text
stringlengths
4
2.78M
meta
dict
--- author: - 'Omrie Ovdat[^1]' - 'Eric Akkermans[^2]' bibliography: - 'refs.bib' title: '[**The breaking of continuous scale invariance to discrete scale invariance: a universal quantum phase transition**]{}' --- Introduction {#sec:1} ============ Continuous scale invariance (CSI) – a common property of physical systems – describes the invariance of a physical quantity $f(x)$ (e.g., the mass) when changing a control parameter $x$ (e.g., the length). This property is expressed by a simple scaling relation, $$f(ax)=b\,f(x),\label{eq: scaling relation}$$ satisfied $\forall a>0$ and corresponding $b(a)$, whose general solution is the power law $$f(x)=C\,x^{\gamma}$$ with $\gamma=\ln b/\ln a$. Other physical systems possess the weaker discrete scale invariance (DSI) expressed by the same scaling relation (\[eq: scaling relation\]) but now satisfied for fixed values $a,b$ and whose solution becomes $$f(x)=x^{\gamma}\,G\left(\ln x/\ln a\right), \label{eq: DSI scaling}$$ where $G(u+1)=G(u)$. Since $G\left(u\right)$ is a periodic function, one can expand it in Fourier series $G\left(u\right)=\sum c_{n}e^{2\pi inu}$, thus, $$f\left(x\right)=\sum_{n=-\infty}^{\infty}c_{n}x^{\gamma+i\frac{2\pi n}{\ln a}}. \label{eq: DSI power law expansion}$$ If $f\left(x\right)$ is required to obey CSI, $G\left(u\right)$ would be constraint to fulfill the relation $G\left(u\right)=G\left(u+a_0\right)\:\forall a_0\in\mathbb{R}$. In this case, $G\left(u\right)$ can only be a constant function, that is, $c_{n}=0$ for all $n\neq0$ eliminating all terms with complex exponents in . Therefore, real exponents are a signature of CSI and complex exponents are a signature of DSI. In this article we describe a variety of distinct quantum systems in which a sharp transition initiates the breaking of CSI into DSI. Essential to all these cases is a DSI phase characterized by a sudden appearance of a low energy spectrum arranged in an infinite geometric series. Accordingly, each transition is associated with exponents that change from real to complex valued at the critical point. We describe the universal properties of this transition. Particularly, in the framework of the renormalization group it is shown that universality in this case is not associated with trajectories terminating at a fixed point but with periodic flow known as a limit cycle. Intrinsic to this phenomena is a special type of scale anomaly in which residual discrete scaling symmetry remains at the quantum level. We discuss the physical realizations of the CSI to DSI transition and present recent experimental observations which provide evidence for the existence of the critical point and for the universal low energy features of the DSI phase. We discuss the basic ingredients that underline these features and the possibility of their occurrence in other, yet to be studied systems. The Schrödinger $1/r^{2}$ potential {#sec:The 1/r^2 -potential} =================================== A well studied example exhibiting the breaking of CSI to DSI is given by the problem of a quantum particle in an attractive inverse square potential [@Case1950b; @Landau1991b] described by the Hamiltonian ($\hbar=1,\,m=1/2$) $$H_{S}=p^{2}-\lambda/r^{2}.\label{eq:H_S Schrodinger equation}$$ This system constitutes an effective description of the “Efimov effect” [@Efimov1970c; @Efimov1971b] and plays a role in various other systems [@Levy-Leblond1967; @Camblong:2001zt; @Kaplan:2009kr; @NisoliBishop2014; @DeMartinoKloepferMatrasulovEtAl2014]. The spectral properties of $H_{S}$ {#subsec: The spectral properties of H_S} ---------------------------------- The Hamiltonian $H_{S}$ has an interesting yet disturbing property – the power law form of the potential matches the order of the kinetic term. As a result, the Schrödinger equation $$H_{S}\psi=E\psi\label{eq: 1/r^2 Schrodinger}$$ depends on the single dimensionless parameter $\lambda$ which raises the question of the existence of a characteristic energy to express the eigenvalues $E_{n}$. This absence of characteristic scale implies the invariance of $H_{S}\psi=E\psi$ under the scale transformation [@Jackiw1995b] $$x^{i}\rightarrow ax^{i},\,E\rightarrow a^{-2}E$$ which indicates that if there is one negative energy bound state $E_{n}$ then there is an unbounded continuum of bound states which render the Hamiltonian nonphysical and mathematically not self-adjoint [@Meetz1964d; @GitmanTyutinVoronov2012c]. The eigenstates of $H_{S}$ can be solved in terms of Bessel functions which confirm these assertions in more detail. For $E<0$ and lowest orbital angular momentum subspace $l=0$, the most general decaying solution is described by the radial function $$\psi\left(r\right)\approx r^{-\frac{d-2}{2}}\left(\left(kr\right)^{-\sqrt{\lambda_{c}-\lambda}}\left(a_{1}+\mathcal{O}\left(kr\right)^{2}\right)+\left(kr\right)^{\sqrt{\lambda_{c}-\lambda}}\left(a_{2}+\mathcal{O}\left(kr\right)^{2}\right)\right)\label{eq: s-wave solution of 1/r^2 potential}$$ where $k\equiv\sqrt{-E}$, $a_{1},\,a_{2}$ are energy independent coefficients, $d$ is the space dimension and $\lambda_{c}\equiv(d-2)^{2}/4$ [^3]. As seen in (\[eq: s-wave solution of 1/r\^2 potential\]), for $\lambda>\lambda_{c}-1$, $\psi_{0}\left(r\right)$ is normalizable $\forall\, \mathrm{Re}\, \left(E\right) < 0$ which constitutes a continuum of complex valued bound states of $H_{S}$. Thus, for $\lambda>\lambda_{c}-1$, $H_{S}$ is no longer self-adjoint, a property that originates from the strong singularity of the potential and is characteristic of a general class of potentials with high order of singularity [@Case1950b]. A simple, physically instructive procedure to deal with the absence of self-adjointness is to remove the singular $r=0$ point by introducing a short distance cutoff $L$ and to apply a boundary condition at $r=L$ [@DeMartinoKloepferMatrasulovEtAl2014; @AlbeverioHoegh-KrohnWu1981a; @BeaneBedaqueChildressEtAl2001a; @MuellerHo2004; @BraatenPhillips2004b; @HammerSwingle2006c; @MorozSchmidt2009]. The most general boundary condition is the mixed condition $$L\frac{\psi'\left(L\right)}{\psi\left(L\right)}=g,\label{eq:mixed condition}$$ $g\in\mathbb{R}$, for which there is an infinite number of choices each describing different short range physics. Equipped with condition (\[eq:mixed condition\]) the operator $H_{S}$ is now a well defined self-adjoint operator on the interval $L<r<\infty$. The spectrum of $H_{S}$ exhibits two distinct features in the low energy $kL\ll1$ regime. For $\lambda<\lambda_{c}\equiv\left(d-2\right)^{2}/4$, the expression of $L\psi'\left(L\right)/\psi\left(L\right)$ as given from (\[eq: s-wave solution of 1/r\^2 potential\]) is independent of $k$ to leading order in $kr$. As a result, equation (\[eq:mixed condition\]) does not hold for a general choice of $g$. For $\lambda>\lambda_{c}$, the insertion of (\[eq: s-wave solution of 1/r\^2 potential\]) into (\[eq:mixed condition\]) leads to $$\left(kL\right)^{2i\sqrt{\lambda-\lambda_{c}}}=e^{i\gamma}\label{eq:k relation}$$ where $\gamma\left(g,\lambda\right)$ is a phase that can be calculated (the explicit expression of $\gamma$ is not important for the purpose of this section). The solution of (\[eq:k relation\]) yields a set of bound states with energies $$k_{n}=k_{0} e^{-\frac{\pi n}{\sqrt{\lambda-\lambda_{c}}}} \label{eq: Efimov spectrum}$$ where $n\in Z$, such that $k_{n}L\ll1$ and $k_{0}\equiv\frac{1}{L}e^{\frac{\gamma}{2\sqrt{\lambda-\lambda_{c}}}}$. Thus, for $\lambda<\lambda_{c}\equiv\left(d-2\right)^{2}/4$, the spectrum contains no bound states close to $E=0$, however, as $\lambda$ goes above $\lambda_{c}$, an infinite series of bound states appears. Moreover, in this ”over-critical” regime, the states arrange in a geometric series such that $$k_{n+1}/k_{n}=e^{-\frac{\pi}{\sqrt{\lambda-\lambda_{c}}}}.$$ The absence of any states for $\lambda<\lambda_{c}$ is a signature of CSI while the geometric structure of (\[eq: Efimov spectrum\]) for $\lambda>\lambda_{c}$ is a signature of DSI since $k_n$ is invariant under $\left\{ k_{n}\right\} \to\left\{ \exp\left(-\pi/\sqrt{\lambda-\lambda_{c}}\right)k_{n}\right\} $. Accordingly, as seen in (\[eq: s-wave solution of 1/r\^2 potential\]), the characteristic behavior of the eigenstates for $kr\ll1$ manifests an abrupt transition from real to complex valued exponents as $\lambda$ exceeds $\lambda_{c}$. Thus, $H_{S}$ exhibits a quantum phase transition (QPT) at $\lambda_{c}$ between a CSI phase and a DSI phase. The characteristics of this transition are independent of the values of $L,g$ which enter only into the overall factor $k_{0}$ in (\[eq: Efimov spectrum\]). The functional dependence of $k_{n}$ on $\sqrt{\lambda-\lambda_{c}}$ is characteristic of Berezinskii-Kosterlitz-Thouless (BKT) transitions as was identified in [@Kaplan:2009kr; @KolomeiskyStraley1992c; @Jensen2011; @JensenKarchSonEtAl2010]. Finally, the breaking of CSI to DSI in the $\lambda>\lambda_{c}$ regime constitutes a special type of scale anomaly since a residual symmetry remains even after regularization (see Table \[tab:QPTtable\]). [ccccc]{} & $\lambda<\lambda_{c}$ & $\lambda>\lambda_{c}-1$ & $\lambda>\lambda_{c}$ & Formal Hamiltonian & CSI & CSI & CSI & Self-adjointness & $H=H^{\dagger}$ & $H\neq H^{\dagger}$ & $H\neq H^{\dagger}$ & Regularization with $L$ & Redundant & Essential & Essential & Symmetry of eigenspace & CSI & CSI & DSI & & Physical realizations of $H_{S}$ {#subsec: Realizations of H_S} -------------------------------- A well known realization of $H_{S}$ for $\lambda>\lambda_{c}$ is the “Efimov effect” [@Efimov1970c; @Efimov1971b; @BraatenHammer2006]. In $1970$, Efimov studied the quantum problem of three identical nucleons of mass $m$ interacting through a short range ($r_{0}$) potential. He pointed out that when the scattering length $a$ of the two-body interaction becomes very large, $a\gg r_{0}$, there exists a scale-free regime for the low-energy spectrum, $\hbar^{2}/ma^{2}\ll E\ll\hbar^{2}/mr_{0}^{2}$, where the corresponding bound-states energies follow the geometric series $E_{n}=-E_{0}e^{-2\pi n/s_{0}}$ where $s_{0}\approx1.00624$ is a dimensionless number and $E_{0}>0$ a problem-dependent energy scale. Efimov deduced these results from an effective Schrödinger equation in $d=3$ with the radial ($l=0$) attractive potential $V\left(r\right)=-\lambda/r^{2}$ with $\lambda=s_{0}+1/4>\lambda_{c}$ ($\lambda_{c}=1/4$ for $d=3$). Despite being initially controversial, Efimov physics has turned into an active field especially in atomic and molecular physics where the universal spectrum has been studied experimentally [@kraemer2006evidence; @TungJimenez-GarciaJohansenEtAl2014a; @PiresUlmanisHaefnerEtAl2014; @pollack2009universality; @GrossShotanKokkelmansEtAl2009; @LompeOttensteinSerwaneEtAl2010; @NakajimaHorikoshiMukaiyamaEtAl2011; @KunitskiZellerVoigtsbergerEtAl2015] and theoretically [@BraatenHammer2006]. The observation of the Efimov geometric spectral ratio $e^{2\pi/s_{0}}\approx515.028$ have been recently determined using an ultra-cold gas of caesium atoms [@huang2014observation]. In addition to the Efimov effect, the inverse square potential also describes the interaction of a point like dipole with an electron in three dimensions. In this case, the dipole potential is considered as an inverse square interaction with non-isotropic coupling [@Camblong:2001zt]. The Klein Gordon equation for a scalar field on an Euclidean AdS $d+1$ space time can be written in the form of (\[eq: 1/r\^2 Schrodinger\]). The over-critical regime $\lambda>\lambda_{c}$ corresponds to the violation of the Breitenlohner-Freedman bound [@Kaplan:2009kr]. Massless Dirac Coulomb system {#sec: Dirac Coulomb system} ============================== The inverse square Hamiltonian (\[eq:H\_S Schrodinger equation\]), a simple system exhibiting a rich set of phenomena, inspires studying the ingredients which lead to the aforementioned DSI and QPT and whether they are found in other systems. One such candidate system is described by a massless Dirac fermion in an attractive Coulomb potential [@MIRANSKY1980421; @PereiraNilssonCastroNeto2007; @ShytovKatsnelsonLevitov2007; @ShytovKatsnelsonLevitov2007d] with the scale invariant Hamiltonian ($\hbar=c=1)$ $$H_{D}=\gamma^{0}\gamma^{j}p_{j}-\beta/r \label{eq: Dirac Hamiltonian}$$ where $\beta$ specifies the strength of the electrostatic potential, $d$ is the space dimension and $\gamma^{\mu}$ are $d+1$ matrices satisfying the anti-commutation relation $$\left\{ \gamma^{\mu},\gamma^{\nu}\right\} =2\eta^{\mu\nu}$$ with $\eta^{00}=\eta^{ii}=-1$, $i=1,\ldots,d$ and $\eta^{\mu\nu}=0$ for $\mu\neq\nu$. Based on the previous example, it may be anticipated that, like $H_{S}$, $H_{D}$ will exhibit a sharp spectral transition at some critical $\beta$ in which the singularity of the potential will ruin self-adjointness. As detailed below, the analog analysis of the Dirac equation $$H_{D}\psi=E\psi\label{eq: Dirac equation}$$ confirm these assertions and details a remarkable resemblance between the low energy features of the two systems. The spectral properties of $H_{D}$ ---------------------------------- Utilizing rotational symmetry, the angular part of equation (\[eq: Dirac equation\]) can be solved and the radial dependence of $\psi$ is given in terms of two functions $\Psi_{1}\left(r\right),\Psi_{2}\left(r\right)$ [@dong2011wave] determined by the following set of equations $$\begin{aligned} \Psi_{2}'\left(r\right)+\frac{\left(d-1+2K\right)}{2r}\Psi_{2}\left(r\right) & =\left(E+\frac{\beta}{r}\right)\Psi_{1}\left(r\right)\nonumber \\ -\Psi_{1}'\left(r\right)-\frac{\left(d-1-2K\right)}{2r}\Psi_{1}\left(r\right) & =\left(E+\frac{\beta}{r}\right)\Psi_{2}\left(r\right)\label{eq:radial Dirac eq}\end{aligned}$$ where $$K\equiv\begin{cases} \pm\left(l+\frac{d-1}{2}\right) & d>2\\ m+1/2 & d=2 \end{cases},\label{eq:def K}$$ $l=0,1,\ldots$ and $m\in\mathbb{Z}$ are orbital angular momentum quantum numbers. In terms of these radial functions, the scalar product of two eigenfunctions $\psi,\tilde{\psi}$ is given by $$\int dV\,\psi^{\dagger}\tilde{\psi}=\int dr\,r^{d-1}\left(\Psi_{1}^{\ast}\left(r\right)\tilde{\Psi}_{1}\left(r\right)+\Psi_{2}^{\ast}\left(r\right)\tilde{\Psi}_{2}\left(r\right)\right).$$ Unlike $H_{S}$ in section \[sec:The 1/r\^2 -potential\], the spectrum of $H_{D}$ does not contain any bound states, a property that reflects the absence of a mass term. As a result, the spectrum is a continuum of scattering states spanning $-\infty<E<\infty$. In the absence of bound states we explore the possible occurrence of “quasi-bound” states. Quasi bound states are pronounced peaks in the density of states $\rho(E)$, embedded within the continuum spectrum. These resonances describe a scattering process in which an almost monochromatic wave packet is significantly delayed by $V\left(r\right)$ as compared to the same wave packet in free propagation [@Friedrich2013b]. An elegant procedure for calculating the quasi-bound spectrum [@Friedrich2013b] is to allow the energy parameter to be complex valued $E\rightarrow\epsilon\equiv E_{R}-i\frac{W}{2}$ and look for solutions of (\[eq:radial Dirac eq\]) with no outgoing $e^{-iEr}$ plane wave component for $r\rightarrow\infty$. The lifetime of the resonance is given by $W^{-1}.$ Consider the lowest angular momentum subspace $K=\pm\left(d-1\right)/2$ and $E<0$, the most general solution with no outgoing component is given by $$\begin{aligned} \begin{pmatrix}\Psi_{1}\left(r\right)\\ \Psi_{2}\left(r\right) \end{pmatrix}\approx r^{-\frac{d-1}{2}} & \left(\left(2iEr\right)^{\sqrt{\beta_{c}^{2}-\beta^{2}}}\left(\begin{pmatrix}a_{11}\\ a_{12} \end{pmatrix}+\mathcal{O}\left(\left|E\right|r\right)\right) \right. \nonumber \\ + & \left. \left(2iEr\right)^{-\sqrt{\beta_{c}^{2}-\beta^{2}}}\left(\begin{pmatrix}a_{21}\\ a_{22} \end{pmatrix}+\mathcal{O}\left(\left|E\right|r\right)\right)\right) \label{eq: ingoing solution}\end{aligned}$$ where $\beta_{c}\equiv (d-1)/2$ [^4] and $a$ is a $2\times2$ energy independent coefficient matrix. As in the case of $H_{S}$ above it is necessary at this point to remove the singularity of the $1/r$ potential by introducing a radial short distance cutoff $L$ and imposing a boundary condition. To identify this explicitly, consider the case where $E=i$ in (\[eq: ingoing solution\]). Since (\[eq: ingoing solution\]) is (asymptotically) an ingoing $e^{iEr}$ plane wave solution if $E\in\mathbb{R}$, it decays exponentially for $E=i$ and $r\to\infty$. If additionally $\beta^2 > \beta_{c}^{2}-1/4$, then (\[eq: ingoing solution\]) is a normalizable eigenfunction with a complex valued eignvalue which renders $H_{D}$ not self-adjoint. The equivalent mixed boundary condition of (\[eq: Dirac equation\]) can be written as follows [@Yang1987b] $$h=\frac{\Psi_{2}\left(L\right)}{\Psi_{1}\left(L\right)}\label{eq:BC Dirac}$$ where $h\in\mathbb{R}$ is determined by the short range physics. Equipped with this condition $H_{D}$ is now a well defined self-adjoint operator on the interval $L<r<\infty$. The spectrum of $H_{D}$ exhibits two distinct pictures in the low energy $\left|E\right|L\ll1$ regime. For $\beta<\beta_{c}\equiv\left(d-1\right)/2$, the expression of $\Psi_{2}\left(L\right)/\Psi_{1}\left(L\right)$ as given from (\[eq: ingoing solution\]) is independent of $E$ to leading order in $\left|E\right|L$. As a result, equation (\[eq:BC Dirac\]) does hold for a general choice of $h$. For $\beta>\beta_{c}$, the insertion of (\[eq: ingoing solution\]) to (\[eq:BC Dirac\]) reduces into $$\left(2iEL\right)^{2i\sqrt{\beta^{2}-\beta_{c}^{2}}}=z_{0}\label{eq:E relation dirac}$$ where $$z_{0}\left(h,\beta\right)\equiv\frac{h\,a_{21}-a_{22}}{a_{12}-h\,a_{11}}\label{eq:z0 explicit}$$ is a complex valued number[^5] (the explicit exporession for $z_0$, which can be found in [@OvdatMaoJiangEtAl2017], is not important for the purpose of this section). The solution of (\[eq:E relation dirac\]) yields a set of quasi-bound energies at $$E_{n}=E_{0}e^{-\frac{\pi n}{\sqrt{\beta^{2}-\beta_{c}^{2}}}}\label{eq: Efimov spectrum-Dirac}$$ where $n\in \mathbb{Z}$, such that $\left|E_{n}\right|L\ll1$ and $E_{0}\equiv\mathrm{Re}\left(\frac{1}{2iL}z_{0}^{\frac{1}{2i\sqrt{\beta^{2}-\beta_{c}^{2}}}}\right)$. It can be directly verified that $E_R = \mathrm{Re} \, E_n < 0 $ and $W = -2 \mathrm{Im} \, E_n >0$ [@OvdatMaoJiangEtAl2017]. Thus, in complete analogy with the $-\lambda/r^2$ inverse squared potential described in section \[sec:The 1/r\^2 -potential\], for $\beta<\beta_{c}\equiv\left(d-1\right)/2$, the spectrum contains a CSI phase with no quasi-bound states close to $E=0$. As $\beta$ exceeds $\beta_{c}$, an infinite series of quasi-bound states appears which arrange in a DSI geometric series such that $$E_{n+1}/E_{n}=e^{-\frac{\pi}{\sqrt{\beta^{2}-\beta_{c}^{2}}}}.$$ As seen explicitly in (\[eq: ingoing solution\]), the characteristic behavior of the eigenstates for $\left|E\right|r\ll1$ manifests an abrupt transition from real to complex valued exponents at $\beta=\beta_{c}$. The characteristics of this transition are independent of the values of $L,h$ which enter only into the overall factor $E_{0}$ in (\[eq:E relation dirac\]). Thus, under a proper trasformation between $\lambda$ and $\beta$, Table \[tab:QPTtable\] represents a valid and consistent description of the massless Dirac Coulomb system as well. Distinct features associated with spin $1/2$ {#subsec: Distinct features associated with spin} -------------------------------------------- On top of the similarities emphasized above, an interesting difference in the quantum phase transition exhibited by $H_{S}$ and $H_{D}$ results from the distinct spin of the associated Schrödinger and Dirac wave functions. Unlike the scalar Schrödinger case, the lowest angular momentum subspace of $H_{D}$ contains two channels corresponding to $K=\pm (d-1)/2$. As a result, not one but two copies of geometric ladders of the form (\[eq: Efimov spectrum-Dirac\]) appear at $\beta=\beta_{c}$ (see Fig. \[fig: Dirac efimov ladder\]). These two ladders may be degenerate or intertwined depending on the choice of boundary condition in (\[eq:BC Dirac\]). The breaking of the degeneracy between the ladders is directly related to the breaking of a symmetry. To understand this point more explicitly consider the case where $d=2$. There, in a basis where $\gamma^0 = \sigma_z$, $\gamma^1 = i\sigma_1$, $\gamma^2 = -i\sigma_2$, $H_{D}$ is given by $$H_D = \sigma_i p_i - \beta/r \label{eq: 2d H_D}.$$ From it is seen that $H_D$ is symmetric under the following parity transformation $$x\rightarrow-x,\,y\to y,H_{D}\to\sigma_{2}H_{D}\sigma_{2},$$ which in terms of $\Psi_{1}\left(r\right),\,\Psi_{2}\left(r\right)$ is equivalent to [@OvdatMaoJiangEtAl2017] $$\Psi_{1}\left(r\right)\to\Psi_{2}\left(r\right),\,\Psi_{2}\left(r\right)\to-\Psi_{1}\left(r\right),\,m\to-m-1\label{eq:parity transformation radial}$$ where $m$ is the orbital angular momentum. Consequently, the Dirac equation (\[eq:radial Dirac eq\]) is invariant under (\[eq:parity transformation radial\]), however, the boundary condition can break (\[eq:parity transformation radial\]). Typical choices of boundary conditions are 1. \[list:BC1\] Continuously connected constant potential $V\left(r<L\right)=-\beta/L$ [@PereiraKotovCastroNeto2008a] corresponding to $h=J_{m+1}(\beta+EL)/J_{m}(\beta+EL),$ where $J_{n}(x)$ is Bessel’s function. 2. \[list:BC2\] Zero wavefunction of one of the spinor components [@ShytovKatsnelsonLevitov2007] corresponding to $h=0$ or $h=\infty$. 3. \[list:BC3\] Infinite mass term on boundary [@PereiraNilssonCastroNeto2007] corresponding to $h=1$. 4. \[list:BC4\] Chiral boundary conditions [@ovdat2018vacancies] $$h=\begin{cases} 0 & m\geq1\\ \infty & m\leq0 \end{cases} \label{eq: chiral BC}$$ inducing a zero mode localized at the boundary. ![\[fig: Dirac efimov ladder\] Density of states $\rho(E)$ of $H_D$ for $d=2$, $\beta = 1.2 > \beta_c$ and different angular momentum eigenstates. The yellow and blue curves correspond to the $m=0,-1$ angular momentum channels respectively. The boundary condition $h$ used here is the chiral boundary condition which breaks parity. The parameter $L$ is the short distance cutoff taken here to be $0.195 \mathrm{nm}$. The numeric values on both axis are in units of $\hbar c = 0.197 \, \mathrm{eV}\, \mu \mathrm{m} $. The set of pronounced peaks in both curves describes the quasi-bound spectrum in the overcritical regime $\beta>\beta_c$ as calculated in . The lower panel displays the detailed structure of the infinite geometric ladders of the quasi-bound states in a logarithmic scale. The $m = 0, -1$ ladders are intertwined, indicative of the breaking of parity by the boundary condition. These results are independent of the specific choice of $L$ or $h$ (provided that it breaks parity).](quasiBoundStatesOverCrit_twoPanels.pdf){width="100.00000%"} Under (\[eq:parity transformation radial\]), a solution of the Dirac equation with angular momentum $m$ obeying boundary condition (\[eq:BC Dirac\]) will transform into a different solution with angular momentum $-m-1$ obeying (\[eq:BC Dirac\]) with $h\to-h^{-1}$. Thus, the boundary condition respects parity if and only if $$h_{m}=-h_{-m-1}^{-1}.\label{eq:parity perserving BC}$$ Thus case \[list:BC1\] above preserves parity while \[list:BC2\], \[list:BC3\], \[list:BC4\] break parity. If (\[eq:parity perserving BC\]) holds, transformation (\[eq:parity transformation radial\]) links between the $m\leftrightarrow-m-1$ eigenspace solutions. The lowest angular momentum subspaces correspond to orbital angular momentum $m=0,-1$. If (\[eq:parity perserving BC\]) holds, then the two geometric ladders (\[eq: Efimov spectrum-Dirac\]) associated with $m=0,-1$ are degenerate. The reason is that, as seen in , under (\[eq:parity transformation radial\]) $$\begin{aligned} a_{11} & \to a_{12},\,a_{12}\to-a_{11},\,a_{21}\to a_{22},\,a_{22}\to-a_{21}\nonumber \\ h & \to-h^{-1}\label{eq: parity a}\end{aligned}$$ which render $z_{0}$ in (\[eq:z0 explicit\]) and consequently $E_{0}$ invariant. Thus $E_{0,m=\pm1/2}$ are identical in this case. If (\[eq:parity perserving BC\]) does not hold, this symmetry is not enforced and the degeneracy between the ladders is broken. The visualization of parity breaking is displayed in Fig. \[fig: Dirac efimov ladder\] where the density of states $\rho(E)$ of $H_D$ is plotted for the $m=0,-1$ channels and $\beta>\beta_c$. The boundary condition that was used in Fig. \[fig: Dirac efimov ladder\] is the chiral boundary condition which breaks parity. Both curves exhibit an identical set of pronounced peaks condensing near $E=0^-$. These peaks describe quasi-bound states and, accordingly, are arranged in a set of two geometric ladders. The separation between the ladder is a distinct signal of parity breaking. Experimental realization ------------------------ The CSI to DSI transition has recently received further validity and interest due to a detailed experimental observation in graphene [@OvdatMaoJiangEtAl2017]. In what follows, we summarize the results of this observation and emphasize its most significant features. Graphene is a particularly interesting condensed matter system where $H_{D}$ is relevant (for $d=2$). The basic reason for this argument is that low energy excitations in graphene behave as a massless Dirac fermion field with a linear dispersion $E=\pm v_{F}\left|p\right|$ where the Fermi velocity $v_{F}\approx10^{6}$ m/s appears instead of $c$ [@Katsnelson2012d]. These characteristics have been extensively exploited to make graphene a useful platform to emulate specific features of quantum field theory, topology and quantum electrodynamics (QED) [@MIRANSKY1980421; @ShytovKatsnelsonLevitov2007; @ShytovKatsnelsonLevitov2007d; @KatsnelsonNovoselovGeim2006; @PhysRevLett.102.026807; @ZhangTanStormerEtAl2005; @WangWongShytovEtAl2013a], since an effective fine structure constant $\alpha_{G}=e^{2}/\hbar v_{F}$ of order unity is obtained by replacing the velocity of light $c$ by $v_{F}$. It has been shown that single-atom vacancies in graphene can host a local and stable charge [@OvdatMaoJiangEtAl2017; @mao2016realization; @LiuWeinertLi2015]. This charge can be modified and measured at the vacancy site by means of scanning tunneling spectroscopy and Landau level spectroscopy [@mao2016realization]. The presence of massless Dirac excitations in the vicinity of the vacancy charge motivates the assumption that these will interact in a way that can be described by a massless Dirac Coulomb system. Particularly, the low energy spectral features of the charged vacancy would be the same as that of a tunable Coulomb source. The experimental results of [@OvdatMaoJiangEtAl2017] provide confirmation of this hypothesis as will be detailed below. The measurements and data analysis presented below were carried out as follows: positive charges are gradually increased into an initially prepared single atom vacancy in graphene. Using a scanning tunneling microscope (STM) the differential conductance $dI/dV \left( V \right)$ through the STM tip is measured at each charge increment at the vacancy site. The conductance $dI/dV \left( V \right)$ is expected to be proportional to the local density of states of the system [@OvdatMaoJiangEtAl2017; @akkermans2007mesoscopic]. Thus, quasi-bound states should also appear as pronounced peaks in the $dI/dV$ curves. For low enough values of the charge, the differential conductance displayed in Fig. \[fig: under critical Theory Vs Experiment\]b, shows the existence of a single quasi-bound state resonance whose distance from the Dirac point increases with charge. The behaviour close to the Dirac point, is very similar to the theoretical prediction of the under-critical regime $\beta<\beta_c$ displayed in Fig. \[fig: under critical Theory Vs Experiment\]a. The $\beta$ value associated with the data of Fig. \[fig: under critical Theory Vs Experiment\]b is obtained from matching the position of the quasi-bound state with the theoretical model where the cutoff $L$ and the boundary condition $h$ are fixed model parameters that will be given later. The theoretical position of the under-critical quasi-bound state as a function of $\beta$ is displayed in Fig. \[fig:CSI to DSI graphene\] along with the positions of the peak extracted from measurements. The existence of a quasi-bound state does not contradict CSI of the undercritical phase since the absence of any states occurs only in the low energy limit. ![\[fig: under critical Theory Vs Experiment\] Experimental and theoretical picture in the undercritical regime. (a) Theoretical behaviour of the density of states $\rho(E)$ of the Dirac Hamiltonian $H_D$ in with $d =2$, $c\to v_F = 0.003 c$ and angular momentum channels $m = -1$ (blue) and $m=0$ (yellow). The cutoff and boundary conditions are assigned here with the optimized values $L = 0.195\,\mathrm{nm}$, $h = -0.85(m + 1)$ as explained in the text. The $m = -1$ (blue) branch contains a single peak and the $m = 0$ (yellow) branch shows no peak. While increasing $\beta$, the resonance shifts to lower energy and becomes broader. (b) The conductance $dI/dV$ measured at a single vacancy in graphene using STM as a function of the applied voltage $V$. The determination of the parameter $\beta$ is obtained from matching the position of the peak in the $dI/dV$ curve with the theoretical model where the cutoff $L$ and the boundary condition $h$ are fixed model parameters.](TheoryVsExpUnderCrit.pdf){width="80.00000%"} At the point where the build up charge exceeds a certain value, three additional resonances emerge out of the Dirac point. These resonances are interpreted as the lowest overcritical $( \beta > 1/2 )$ resonances which we denote $E_1, E'_{1}, E_2$ respectively. The corresponding theoretical and experimental behaviours displayed in Figs. \[fig: Dirac efimov ladder\], \[fig: over critical Theory Vs Experiment\], show a very good qualitative agreement. To achieve a quantitative comparison solely based on the massless Dirac Coulomb Hamiltonian , the theoretical $\beta$ values corresponding to the respective positions of the lowest overcritical experimental resonance $E_1$ (as demonstrated in Fig. \[fig: over critical Theory Vs Experiment\]) are deduced for fixed $L$ and the boundary condition $h$ (as before). This allows to determine the lowest branch $E_1 (\beta)$ represented in Fig. \[fig:CSI to DSI graphene\]. Then, the experimental points $E'_{1}, E_2$ are now free points to be directly compared to their corresponding theoretical branch as seen in Fig. \[fig:CSI to DSI graphene\]. Parameters $L$ and $h$, are determined according to the ansatz $h = a(m + 1)$, and correspond to optimal values of $L = 0.195\,\mathrm{nm}$, $a \simeq -0.85$. The comparison of the experimental $E_2/E_1$ ratio with the universal prediction $E_{n+1} / E_n = e^{- \pi / \sqrt{\beta^2 - 1/4}}$ is given in Fig. \[fig:theory Vs Exp quantitative\]. A trend-line of the form $e^{-b/\sqrt{\beta^2-1/4}}$ is fitted to the ratios $E_2/E_1$ yielding a statistical value of $b = 3.145$ with standard error of $\Delta b = 0.06$ consistent with the predicted value $\pi$. An error of $\pm 1meV$ is assumed for the position of the energy resonances. ![ \[fig: over critical Theory Vs Experiment\] Experimental and theoretical picture in the overcritical regime. Upper plot: Theoretical behaviour of the low energy density of states $\rho(E)$ for overcritical $\beta = 1.33$. The Blue (Yellow) line corresponds to $m = -1$ ($m = 0$) orbital angular momentum. The peaks on the vertical scale describe the first quasi-bound states with two (Blue and Yellow) infinite geometric towers of states. Lower plot: Experimental values of the tunnelling conductance measured at the charged vacancy site in graphene.](theoryVsExperimentOverCrticalbeta128.pdf){width="80.00000%"} A few further comments are appropriate: 1. The points on the $E_2 (\beta)$ curve follow very closely the theoretical prediction $E_{n+1} / E_n = e^{- \pi / \sqrt{\beta^2 - 1/4}}$. This result is relatively insensitive to the choice of $h,L$. 2. In contrast, the correspondence between the $E'_{1}\left(\beta\right)$ points and the theoretical branch is sensitive to the choice of $h,L$. This reflects the fact that while each geometric ladder is of the form , the energy scale $E_0$ is different between the $E_1\left(\beta\right)$ and $E'_{1}\left(\beta\right)$ channels thus leading to a shifted relative position. The ansatz taken for $h$ is phenomenological, however, in order to get reasonable correspondence to theory, the explicit dependence on $m$ is needed. More importantly, it is necessary to use a parity breaking boundary condition (see section \[subsec: Distinct features associated with spin\]) to describe the $E'_{1}(\beta)$ points, otherwise, both angular momentum channels $E_1\left(\beta\right)$, $E'_{1}\left(\beta\right)$ will become degenerate and there would be no theoretical line to describe the $E'_{1}\left(\beta\right)$ points. The existence of the experimental $E'_{1}\left(\beta\right)$ branch is therefore a distinct signal that parity symmetry in the corresponding Dirac description is broken. In graphene, exchanging the triangular sub-lattices is equivalent to a parity transformation. Creating a vacancy breaks the symmetry between the two sub-lattices and is therefore at the origin of broken parity in the Dirac model. 3. The optimal value obtained for the short distance cutoff $L = 0.195 \, \mathrm{nm}$ is fully consistent with the low energy requirement $E_1 L / \hbar v_F \simeq 0.03 \ll 1$ necessary to be in the regime relevant to observe the $\beta$-driven QPT. Furthermore, it is quite close the lattice spacing of graphene ($\approx 0.15 \, \mathrm{nm}$) One of the most interesting features of observed quasi bound states is their similarity with the Efimov spectrum. As discussed in section \[subsec: Realizations of H\_S\], Efimov states are a geometric tower of states with a fixed geometric factor which is derived from an effective Schrödinger equation with a $V = -\lambda/r^2$ potential (as in ) and overcritical potential strength $\lambda = s_0 +1/4$, $s_0 \approx 1.00624$. To emphasize the similarities between the Dirac quasi bound spectrum and the Efimov spectrum or, more generally, between the CSI to DSI transition in the Dirac and Schrödinger Hamiltonians $H_D$, $H_S$, two additional experimental points (pink x’s) are presented in Fig. \[fig:CSI to DSI graphene\]. These points are the values of Efimov energy states measured in Caesium atoms [@kraemer2006evidence; @huang2014observation] and scaled with an appropriate overall factor. The points are placed at the (overcritical) fixed Efimov value $\beta_\mathrm{E} = 1.1236 $ corresponding to the geometric factor of Efimov states. The universality of the transition is thereby emphasized in Fig. \[fig:CSI to DSI graphene\] in which curves calculated from a massless Dirac Hamiltonian, energy positions of tunneling conductance peaks in graphene and resonances of a gas of Caesium atoms are combined in a meaningful context. Relation to universality ======================== In sections \[sec:The 1/r\^2 -potential\], \[sec: Dirac Coulomb system\] we obtained the properties of the CSI to DSI transition from a direct analysis of the corresponding eigenstates of each system. In what follows, we describe the same physics, but this time through the language of the renormalization group (RG). As will be detailed next, the description of this phenomenon in a RG picture provides a notable example of a case in which there is universality even in the absence of any fixed points. To understand this point more clearly, we first recall the physical meaning of the RG formalism and the usual context for which universality is understood with relation to RG. Universality is a central concept of physics. It refers to phenomena for which very different systems exhibit identical behavior when properly coarse-grained to large distance (or low energy) scales. Important representatives of universality are systems that are close to a critical point, e.g., liquid-gas or magnetic systems. Near the critical point, these systems exhibit continuous scale invariance (as in ) where the free energy and correlation length vary as a power of the temperature (or some other control parameter). The exponents of these functions are real valued and are identical for a set of different systems thereby constituting a “universality class”. The contemporary understanding of university in critical phenomena is provided by the tools of RG and effective theory. In the framework of the later, low energy physics is described by a Hamiltonian $H$ with a series of interaction terms $g_{n}\mathcal{O}_{n}$ constrained by symmetries. Intrinsic to this description is an ultraviolet cutoff $\Lambda$ reflecting the conceptual idea that $H$ is obtained from some microscopic Hamiltonian $H_{0}$ by integrating out degrees of freedom with length scale shorter than $1/\Lambda$. The dependence of $\vec{g}\equiv\left(g_{1},g_{2}\ldots\right)$ on $\Lambda$ defines the RG space of parameters $\vec{g}\left(\Lambda\right)$ which represent a large set of Hamiltonians $H\left(\vec{g}\left(\Lambda\right)\right)$. Within this picture, the scale invariant character of critical phenomena is attributed to the case where $H_{0}\left(\vec{g}_{0}\right)$ flows in the infrared limit, $\Lambda\rightarrow0$, to $H\left(\vec{g}_{\ast}\right)$ where $\vec{g}_{\ast}$ is a fixed point. Additionally, universality classes arise since trajectories starting at distinct positions on RG space can flow to the same fixed point for $\Lambda\rightarrow0$. The role of RG fixed points in the description of universality, effective theory and scale invariance is central and extends throughout broad sub-fields in physics. Renomarlization group formalism for the Schrödinger $1/r^2$ potential {#subsec: RG of 1/r^2} --------------------------------------------------------------------- The RG picture which describes the low energy physics of the Schrödinger $-\lambda/r^2$ potential in the $\lambda>\lambda_{c}$ regime cannot be associated with a fixed point because of the absence of CSI. However, even without fixed points, we expect universality to appear in this regime since the geometric series factor $E_{n+1}/E_{n}$ = $\exp\left(-2\pi/\sqrt{\lambda-\lambda_{c}}\right)$ is independent of the short distance parameters associated with the cutoff $L$ and the boundary condition $g$ To see this explicitly [@Kaplan:2009kr; @AlbeverioHoegh-KrohnWu1981a; @BeaneBedaqueChildressEtAl2001a; @MuellerHo2004; @KolomeiskyStraley1992c], consider the radial Schrödinger equation for $H_S$ given by $$-\left(\frac{d^{2}}{dr^{2}}+\frac{d-1}{r}\frac{d}{dr}-\frac{l(l+d-2)}{r^{2}}\right)-\frac{\lambda}{r^{s}} \psi\left(r\right)=E\psi\left(r\right),\hspace{1em}L<r<\infty \label{eq: radial sch equation}$$ where $\psi\left(r\right)$ is the radial wavefunction, $l$ the orbital angular momentum, $d$ the space dimension, $L$ a short distance cutoff and $s=2$ but remains implicit for a reason that will be clear shortly. A well defined eigenstate of (\[eq: radial sch equation\]) is obtained by imposing a boundary condition at $r=L$ $$L\frac{\psi^{\prime}\left(L\right)}{\psi\left(L\right)}=g, \label{eq:BC RG}$$ $g \in \mathbb{R}$, which encodes the short-distance physics. To initiate a RG transformation we transform $$L\rightarrow L+dL\equiv\epsilon L\,;\hspace{1em}0<\epsilon-1\ll1\label{eq: RG tran}$$ and obtain an equivalent effective description with the short distance cut-off $\epsilon L$ and correspondingly, a new boundary condition at $r=\epsilon L$: $$\begin{aligned} \epsilon L\frac{\psi^{\prime}\left(\epsilon L\right)}{\psi\left(\epsilon L\right)} & = & g\left(\epsilon L\right).\label{eq:BC trans}\end{aligned}$$ As a result of , equation is now defined in the range $\epsilon L\leq r<\infty$ with the same functional form. With the help of the rescaling $r'\equiv \epsilon^{-1} r,\, E'\equiv\epsilon^{2}E$, equation is modified to the equivalent form $$-\left(\frac{d^{2}}{dr'^{2}}+\frac{d-1}{r'}\frac{d}{dr'}-\frac{l(l+d-2)}{r'^{2}}\right)-\frac{\lambda\epsilon^{2-s}}{r'^{s}}\psi\left(r'\right)= E'\psi\left(r'\right)\hspace{1em}L<r'<\infty.$$ Thus, transformation (\[eq: RG tran\]) is accounted in (\[eq: radial sch equation\]) by $\lambda \to \lambda \epsilon^{2-s}$ and using (\[eq: RG tran\]) leads to the infinitesimal form $$L\frac{d\lambda}{dL}=\left(2-s\right)\lambda.\label{eq: RG eq lambda}$$ Similarly, $g\left(\epsilon L\right)$ in (\[eq:BC trans\]) can be related to $g\left(L\right)$ as follows. The series expansion of $g\left(\epsilon L\right)$ in $\epsilon-1$ is $$g\left(\epsilon L\right) = L\frac{\psi^{\prime}\left(L\right)}{\psi\left(L\right)}+\left(\epsilon-1\right)\left(L\frac{\psi^{\prime}\left(L\right)}{\psi\left(L\right)}-L^{2}\left(\frac{\psi^{\prime}\left(L\right)}{\psi\left(L\right)}\right)^{2}+L^{2}\frac{\psi^{\prime\prime}\left(L\right)}{\psi\left(L\right)}\right)+\mathcal{O}\left(\epsilon-1\right)^{2}. \label{eq: g expansion 0}$$ Manipulation of by insertion of the radial Schrödinger equation and the definition of $g\left(L\right)$ yield $$g\left(\epsilon L\right) = g\left(L\right)+\left(\epsilon-1\right)\left(\left(2-d\right)g\left(L\right)-g\left(L\right)^{2}-\lambda L^{2-s}+l(l+d-2)-L^{2} E\right) \label{eq: g expansion}$$ where terms of order $\left(\epsilon-1\right)^2$ or higher were eliminated. The equivalent differential form is thus $$L\frac{dg}{dL}=\left(2-d\right)g-g^{2}-\lambda L^{2-s}+l(l+d-2)-L^{2} E.\label{eq: g RG eq}$$ In the low energy regime $$L^2\left|E\right|\ll\left|\lambda-l(l+d-2)\right|\label{eq:low E approx}$$ equation reduces to $$L\frac{dg}{dL} = \left(2-d\right)g-g^{2}-\lambda\label{eq: g RG eq low E}$$ where the orbital angular momentum was taken to be $l=0$ and $s$ set to $s=2$ for brevity. Finally, the combination of , constitutes the RG equations $$\begin{aligned} \beta\left(\lambda\right) \equiv L\frac{d\lambda}{dL} & = & \left(2-s\right)\lambda \nonumber \\ \beta\left(g\right) \equiv L\frac{dg}{dL} & = & -\left(g-g_{+}\right)\left(g-g_{-}\right)\label{eq: RG Schrodinger 1/r^2}\end{aligned}$$ where $$g_{\pm}=\frac{2-d}{2}\pm\sqrt{\lambda_{c}-\lambda}$$ and $\lambda_c = \left(d-2\right)^2/4$. Since $\beta\left(\lambda\right) = 0$ for $s=2$, $\lambda\left(L\right)$ remains unchanged under the RG transformation. In contrary, the function $\beta \left(g\right) $ is not trivial and has two roots $g_{\pm}$ . For $\lambda<\lambda_{c}$, the two roots correspond to two fixed points, $g_{-}$ unstable and $g_{+}$ stable. However, as $\lambda$ increases, the two fixed points get closer and merge for $\lambda=\lambda_{c}$. For $\lambda>\lambda_{c}$, $g_{\pm}$ become complex valued and the two fixed points vanish as can be seen in Fig. \[fig:RGSchrodinger\]a. The solution for $g\left(L\right)$ in this regime is given explicitly by (see Fig. \[fig:RGSchrodinger\]b) $$g\left(L\right)=\frac{2-d}{2}-\sqrt{\lambda-\lambda_{c}}\tan\left[\sqrt{\lambda-\lambda_{c}}\ln\left(L/L_{0}\right) - \phi_{g} \right]\label{eq: limit cycle 1/r^2}$$ where $\phi_{g}\equiv\arctan\left(\frac{g_{0}-\frac{2-d}{2}}{\sqrt{\lambda-\lambda_{c}}}\right)$. Unlike the case of a fixed point, the flow of $g\left(L\right)$ in does not terminate at any specific point but rather oscillate periodically in $\log L$ with period $L\to e^{\pi/\sqrt{\lambda-\lambda_{c}}}L$ independent of the initial condition $g\left(L_{0}\right) = g_0$. The appearance of two fixed points for $\lambda<\lambda_{c}$, which annihilate at $\lambda_{c}$ and give rise to a log-periodic flow for $\lambda>\lambda_{c}$ is the transcription of the CSI to DSI transition in the RG picture. The periodicity $e^{\pi/\sqrt{\lambda-\lambda_{c}}}$, being independent on the initial conditions, $g\left(L_{0}\right) = g_0$, represents a universal content even in the absence of fixed points. ![\[fig:RGSchrodinger\] Visualization of the renormalization group picture associated with the boundary condition $g(L)$ at the short distance cutoff $r=L$ for the case of the Schrödinger $V\left(r\right) = -\lambda/r^2$ potential $H_S$. (a) The $\beta\left(g\right)$ function in the over-critical and under-critical regimes. For $\lambda < \lambda_{c}$, $\beta\left(g\right)$ has two roots correspond to two fixed points, $g_-$ unstable and $g_+$ stable. The point $\lambda = \lambda_{c}$ is a transition point where the roots merge into a single fixed point. For $\lambda > \lambda_{c}$ there are no real fixed points. (b) The behaviour of the boundary condition $g\left(L\right)$ in the overcritical regime $\lambda > \lambda_{c}$ and $d=3$ as a function of $\ln \left(x\right)$ with $x \equiv \sqrt{\lambda-\lambda_{c}}\ln\left(L/L_{0}\right) - \phi_{g}$, $\phi_{g} \equiv \arctan\left(\frac{g_{0}-\frac{2-d}{2}}{\sqrt{\lambda-\lambda_{c}}}\right)$. Independent of the initial condition $g_0\left(L_0\right)$, $g\left(L\right)$ is a log-periodic function of $L$ which, as shown in , is a generic feature of DSI.](RGSchrodinger.pdf){width="\textwidth"} An analogue of the RG equations (\[eq: RG Schrodinger 1/r\^2\]) can be derived for the boundary condition $h\left(L\right)$ in of the massless Dirac Coulomb system described in section \[sec: Dirac Coulomb system\] [@gorsky2014atomic]. Discussion ========== The similarities between the Dirac and Schrödinger system $H_{S},H_{D}$ $$\begin{aligned} H_{S} & = & p^{2}-\lambda/r^{2} \\ H_{D} & = & \gamma^{0}\gamma^{j}p_{j}-\beta/r\end{aligned}$$ presented in sections \[sec:The 1/r\^2 -potential\], \[sec: Dirac Coulomb system\] motivate the study of whether a similar transition from CSI to DSI is possible for a generic class of systems and, if so, what are the common ingredients within this class. Below we briefly survey some other setups which interestingly give rise to a CSI to DSI transition. The relation between all these cases is summarized in table \[tab:QPTtable summary\]. [lcccc]{} System & $O_n$ & $x$ & $x_{c}$ &$b$ $H_{S} = p^2 - \lambda/r^2$ & $E_n$ & $\lambda$ & $\left(d-2\right)^{2}/4$ & $2$$H_{D} = \gamma^{0}\gamma^{j}p_{j}-\beta/r$ & $E_n$ & $\beta^2$ & $\left(d-1\right)^2/4$ & $1$$H_{L} = \left(-\dfrac{d^{2}}{dx^{2}}\right)^{N}-\dfrac{\lambda_{L}}{x^{2N}}$ & $E_n$ & $\lambda_{L}$ & $\left(\dfrac{(2N-1)!!}{2^{N}}\right)^{2}$ & $N \alpha_{N}$QED3 with $N$ massless flavours & $m_n$ & $N^{-1}$ & $\pi^{2}/32$ & $\pi/\sqrt{8}$ Efimov effect in $d$ dimensions & $E_n$ & $d$ & $2.3$ & $1/c_{\pm} \left(d\right)$ BKT & $F$ & $T$ & $T_{c}$ & Lifshitz scaling symmetry {#subsec: Lifshitz} ------------------------- Since $H_{S}$ and $H_{D}$ share the property that the power law form of the corresponding potential matches the order of the kinetic term, it is interesting to examine whether this property is a sufficient ingredient by considering a generalized class of one dimensional Hamiltonians, $$H_{L}=\left(-\frac{d^{2}}{dx^{2}}\right)^{N}-\frac{\lambda_{L}}{x^{2N}},\label{Eq: Lifshitz Hamiltonian}$$ where $N$ is a natural integer and $\lambda_{L}$ a real valued coupling. The Hamiltonian $H_{L}$ describes a quantum system with non-quadratic anisotropic scaling between space and time for $N>1$. This so called “Lifshitz scaling symmetry” [@Alexandre2011], manifest in (\[Eq: Lifshitz Hamiltonian\]), can be seen for example at the finite temperature multicritical points of certain materials [@PhysRevLett.35.1678; @PhysRevB.23.4615] or in strongly correlated electron systems [@FradkinHuseMoessnerEtAl2004; @VishwanathBalentsSenthil2004; @ArdonneFendleyFradkin2004]. Quartic dispersion relations $E\sim p^{4}$ can also be found in graphene bilayers [@0034-4885-76-5-056503] and heavy fermion metals [@2012PhRvL.109q6404R]. It may also have applications in particle physics [@Alexandre2011], cosmology [@Mukohyama2010] and quantum gravity [@KachruLiuMulligan2008; @Horava2009; @Horava2009a]. The detailed solution of the corresponding Schrödinger equation $H_{L}\psi=E\psi$ [@BrattanPhysRevD.97.061701] confirms that a transition from CSI to DSI occurs at $\lambda_{L,c}=\left(2N-1\right)!!^{2}/2^{2N}$, $\forall N\geq$1. The CSI phase contains no low energy, $\left|E\right|^{1/2N}L\ll1$ ($L$ is a short distance cutoff), bound states and the DSI phase is characterized by an infinite set of bound states forming the geometric series $$E_{n}=-E_{0}e^{-\frac{N\alpha_{N}\pi n}{\sqrt{\lambda_{L}-\lambda_{L,c}}}},\,\,0<\lambda_{L}-\lambda_{L,c}\ll1\label{eq:Liphsitz spectrum}$$ where $E_0>0$ and $\alpha_{N}$ is an $N$ dependent real number. For $\lambda_{L}-\lambda_{L,c}\to0^{+}$, the analytic behavior of the spectrum is characteristic of the Berezinskii-Kosterlitz-Thouless (BKT) scaling in analogy with the $N=1$ case. However, unlike the $N=1$ case, the BKT scaling appears only for $\lambda_{L}-\lambda_{L,c}\to0^{+}$. Deeper in the overcritical regime, the dependence on $\left(\lambda_{L}-\lambda_{L,c}\right)^{1/2}$ in (\[eq:Liphsitz spectrum\]) is replaced by a more complicated function of $\lambda_{L}-\lambda_{L,c}$ [@BrattanPhysRevD.97.061701]. The transition as well as the value of $\lambda_{L,c}$ is independent of the short distance physics characterized by the boundary conditions and cutoff $L$. Since $H_L$ is a high order differential operator it requires the specification of several $x=L$ boundary condition parameters (unlike the one parameter $g$ in section \[subsec: The spectral properties of H\_S\]) in order to render it as a well defined self-adjoint operator on the interval $L<x<\infty$. The most general choice of boundary conditions is parameterized by a unitary $N\times N$ matrix. Accordingly, the corresponding $N^{2}$ dimensional RG space is characterized by fixed points in the under-critical regime $\lambda_{L}<\lambda_{L,c}$. Interestingly, the DSI over-critical regime $\lambda_{L}>\lambda_{L,c}$ is not filled with an infinite number of cyclic flows such as represented in Fig. \[fig:RGSchrodinger\]b. Instead, there is a ’limit cycle’ [@strogatz2014nonlinear], i.e., an isolated closed trajectory at which flows terminate [@OvdatJounalOfPhysA] (see Fig. \[fig:isolated limit cycle\]). QED in $2+1$ dimensions and $N$ fermionic flavors {#subsec: QED3} ------------------------------------------------- The study of dynamical fermion mass generation in $2+1$ dimensional quantum electrodynamics (QED) [@AppelquistNashWijewardhana1988; @HerbutPhysRevD.94.025036] provides an interesting many body instance of the CSI to DSI transition. Consider the $2+1$ dimensional QED Lagrangian $$\mathcal{L} = i\bar{\Psi}\gamma^{\mu}\left(\partial_{\mu}-ieA_{\mu}\right)\Psi-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}$$ where $\Psi$ is a vector of $N$ identical types of fermion fields with zero mass. In this theory, $e^2$ or alternatively $\alpha \equiv Ne^2/8$, is a dimension-full coupling. Analogous to the short distance cutoff $L$ of sections \[sec:The 1/r\^2 -potential\], \[sec: Dirac Coulomb system\], \[subsec: Lifshitz\], $\alpha$ constitutes the only energy scale of the theory. Consequently, the low energy regime $E \ll \alpha$ can be shown to exhibit CSI. The understand whether or not fermion mass appears as a result of quantum fluctuations it is required to calculate the fermion propagator, specifically, the self-energy $\Sigma \left(p \right)$. Under a particular (non-perturbative) approximation scheme [@AppelquistNashWijewardhana1988], the expression for $\Sigma \left( p \right)$ can be extracted from the solution of the following differential equation $$-\Sigma '' \left( p \right) - \frac{2}{p}\Sigma ' \left( p \right) - \frac{\lambda_Q}{p^2+\Sigma \left( p \right)^2} \Sigma \left( p \right) = 0,\hspace{0.5em}0<p<\alpha \label{eq: QED3 self energy equation}$$ with boundary condition $$\alpha \frac{\Sigma'\left(\alpha\right)}{\Sigma\left(\alpha\right)}=-1 \label{eq: BC Sigma}$$ where $\lambda_Q \equiv 8/ \left( \pi^2 N \right)$. Close to a transition point the fermion mass and thereby $\Sigma \left( p \right)$ are non-zero but arbitrarily small such that $\Sigma \left( p \right) \ll p < \alpha$. As a result, can be further approximated by assuming $\Sigma \left( p \right)^2$ in the denominator is a constant which we define as $\Sigma \left( p \right)^2 \to m^2/\lambda_Q$. Expanding to order $m^2$ yields $$-\Sigma '' \left( p \right) - \frac{2}{p}\Sigma ' \left( p \right) - \frac{\lambda_Q}{p^2} \Sigma \left( p \right) = -\frac{m^2}{p^4} \Sigma \left( p \right). \label{eq: QED3 self energy equation linearized}$$ ![\[fig:isolated limit cycle\] A two dimensional projection of the (four dimensional) RG picture of the system $H = d_{x}^{4}-2/x^4$. The four boundary conditions at $x = L$ are parametrized by a unitary $2\times 2$ matrix $U$. The initial conditions for the dashed blue flows are specified by choosing $\theta=-\pi,\ldots,-\pi/10,0$ for the $U$ matrix as displayed. All the trajectories flow towards a limit cycle. There exists a non-unitary fixed point, denoted by the blue cross, which is enclosed by the cycle when projected down onto any two dimensional subspace. ](isolated_limit_cycle) A closer look on equations , reveals that they are the same as the radial form of the Schrödinger equation with a $V= -\lambda/r^2$ potential $$\begin{aligned} -\left(\frac{d^{2}}{dr^{2}}+\frac{d-1}{r}\frac{d}{dr}-\frac{l(l+d-2)}{r^{2}}\right)-\frac{\lambda}{r^{2}} \psi\left(r\right) & = & -k^2 \psi\left(r\right),\hspace{0.5em}L<r<\infty \label{eq: Schrodinger 1/r^2 Disscussion} \\ L\frac{\psi^{\prime}\left(L\right)}{\psi\left(L\right)} & = & g \label{eq: BC Schrodinger 1/r^2 Disscussion}\end{aligned}$$ where $k = \sqrt{-E}$ as described in section \[sec:The 1/r\^2 -potential\] and in equations , . To see this explicitly, we rewrite , in terms of $r \equiv 1/p$, $\psi\left(r\right) \equiv \Sigma \left(p \right)$, $L \equiv 1/\alpha$ which then yields $$\begin{aligned} -\psi ''(r) -\frac{\lambda_Q }{r^2} \psi (r) & = & - m^2 \psi (r),\hspace{0.5em}L<r<\infty \\ L\frac{\psi^{\prime}\left(L\right)}{\psi\left(L\right)} & = & 1. \label{eq: QED3 self energy equation linearized transformed }\end{aligned}$$ Thus, the appearance of a non-vanishing fermion self energy constitutes a system of the form , with $d = 1,\,\lambda = \lambda_Q$ and $g = 1$. The resulting implication is that a transition from a CSI to DSI occurs at $\lambda_{Q,c}=1/4$. For $\lambda_{Q,c}<1/4$ there will be no $\Sigma\left(p\right) \neq 0$ solution for the self-energy in the $m / \alpha \ll 1$ regime. However, once $\lambda_Q$ exceeds $\lambda_{Q,c} = 1/4$ an infinite geometric tower of possible non-trivial self-energy solutions appears with eigenvalues $$m_{n+1}/m_n = e^{-\frac{\pi}{\sqrt{\lambda_Q - \lambda_{Q,c}}}}.\label{eq: m_n+1 /m_n}$$ The critical point $\lambda_{Q,c}=1/4$ corresponds to a critical fermion number $N_c = 32/\pi^2$ for which the DSI regime is $N<N_c$. In these term, reduces to $$m_{n+1}/m_n = e^{- \frac{\pi}{\sqrt{\frac{1}{N} -\frac{\pi^2}{32}}}\pi/\sqrt{8}}.$$ Efimov effect in $d$ dimensions {#subsec: Efimov in d dimensions} ------------------------------- As described in \[subsec: Realizations of H\_S\], the Efimov effect [@Efimov1970c; @Efimov1971b; @BraatenHammer2006] is a remarkable phenomenon in which three particles form an infinite geometric ladder of low energy bound states. The effect occurs when at least two of the three pairs interact with a range that is small compared to the scattering length. It can be shown that the Efimov effect is possible only for space dimensions $2.3<d<3.76$ [@NIELSEN2001373] which essentially limits the phenomenon to 3 dimensions. Interestingly, in the case where $d$ is allowed to be tuned continuously, two CSI to DSI transitions are initiated at the critical dimensions $d_- = 2.3$, $d_+ = 3.76$ [@PhysRevA.98.013633]. In what follows we outline the main features of this result. ![\[fig: Efimov EFT diagrams\] Diagrammatic representation of the atom-diatom scattering amplitude and the diatom propagator [@BraatenHammer2006]. (a) Diagrammatic self-consistent equation for the atom-diatom scattering amplitude. (b) Diagrammatic self-consistent equation for the diatom field propagator.](atomDiatomAmp.png){width="\textwidth"} Low energy 3-body observables of locally interacting identical bosons can be described by an effective field theory with Lagrangian $$\mathcal{L} = \psi^{\dagger}\left(i\frac{\partial}{\partial t}+\frac{1}{2}\nabla^{2}\right)\psi+\frac{g_{2}}{4}\left(\Delta^{\dagger}\Delta\right)-\frac{g_{2}}{4}\left(\Delta^{\dagger}\psi^{2}+\psi^{\dagger^{2}}\Delta\right)-\frac{g_{3}}{36}\left(\Delta^{\dagger}\Delta\psi^{\dagger}\psi\right)\label{eq:BHvK Lag}$$ where $\psi$ is a non-relativistic bosonic atom field, $\Delta$ is a non dynamical ’diatom’ field annihilating two atoms at one point and $g_2$, $g_3$ are bare 2-body and 3-body couplings respectively. With the diatom field and these interaction terms, it is possible to reproduce the physics of the Efimov effect [@BedaqueHammerVanKloPhysRevLett.82.463]. The main ingredient of this procedure is the diagramatic calculation of the atom-diatom scattering amplitude as shown in Fig. \[fig: Efimov EFT diagrams\]. The self-consistent equations described in Fig. \[fig: Efimov EFT diagrams\] leads to the following approximate relation for the s-wave atom-diatom amplitude $A_s$ $$A_s \left( p \right) = -\left(\frac{4}{3}\right)^{\frac{d-2}{2}} \frac{4\sin\left(\frac{d}{2} \pi \right)}{\pi} \int_0^\infty\, dq \frac{q}{p^2+q^2} \, {}_2 F_1\left(\begin{matrix}\frac{1}{2}& &1 \\ &\frac{d}{2}\end{matrix};\frac{p^2 q^2}{p^2+q^2} \right) A_s\left( q \right) \label{eq: amplitude CSi relation}$$ Since there are no dimension-full parameters in we are, once again, faced with a CSI equation, in analogy with the characteristics of equations , , and . By inserting the ansatz $A_s \left( p \right) = p^{s-1}$, two possible solutions for are obtained $$A_s \approx a_1 p^{\sqrt{s^2}-1} + a_2 p^{-\sqrt{s^2}-1} \label{eq: A_s solution}$$ where $s^2 \left( d \right)$ is a solution of the $s\to-s$ invariant equation $$2 \sin \left( \frac{d\pi}{2}\right) {}_2 F_1\left(\begin{matrix}\frac{d-1+s}{2}& &\frac{d-1-s}{2} \\ &\frac{d}{2}\end{matrix};\frac{1}{4} \right)+ \cos \left( \frac{s}{2} \pi \right) = 0. \label{eq: s transendal equation}$$ The numerical solution $s^2 \left(d \right)$ of shows that near $d=d_{\pm}$, $s^2 \left( d_\pm \right) = 0$, $\partial_d s^{2} \left(d_-\right) < 0$, $\partial_d s^{2} \left(d_+\right) > 0$ and it is analytic. Consequently, near the critical dimensions $d_\pm$ $$s^2\left( d\right) = \pm c_\pm^2 \left(d-d_\pm\right)+\mathcal{O}\left(d-d_{\pm}\right) \label{eq: near critical dimension s^2}$$ with $c_\pm > 0$. The insertion of into imply a CSI to DSI transition from real to complex valued power law behaviour of $A_s$. The DSI regime $d_-<d<d_+$ is consistent with the strip within which the Efimov states appear. Consequently, close to the critical points $d=d_{\pm}$, $A_s \left(p\right)$ in obeys the following DSI scaling relation (as in ) $$A_s \left(e^{\frac{\pi n}{c_{\pm}\sqrt{|d-d_{\pm}|}}}p\right) = e^{-\frac{\pi n}{c_{\pm}\sqrt{|d-d_{\pm}|}}} A_s \left(p\right).$$ The corresponding RG equation for the couplings is $$\Lambda \frac{d}{d\Lambda} G = \frac{1 - s^2 \left( d \right)}{2} \left(G - G_- \right)\left(G - G_+ \right)$$ where $G\left( \Lambda \right) \equiv \Lambda ^2 g_3 \left(\Lambda\right)/9 g_2 \left(\Lambda\right)^2$, $\Lambda$ is a UV cutoff and $$G_\pm \equiv - \left(1\pm \sqrt{s^2\left( d \right)}\right)/\left(1\mp \sqrt{s^2\left( d \right)}\right).$$ In accordance with the RG picture detailed in section \[subsec: RG of 1/r\^2\], the insertion of shows that the $\beta$-function of $G$ contains two fixed points outside the strip $d_- < d < d_+ $ which annihilate at $d=d_\pm$. Summary ======= The breaking of continuous scale invariance (CSI) into discrete scale invariance (DSI) is a rich phenomenon with roots in multiple fields in physics. Theoretically, this transition plays an important role in various fundamental quantum systems such as the inverse-squared potential (section \[sec:The 1/r\^2 -potential\]), the massless hydrogen atom (section \[sec: Dirac Coulomb system\]), $2+1$ dimensional quantum electrodynamics (section \[subsec: QED3\]) and the Efimov effect (sections \[subsec: Realizations of H\_S\] and \[subsec: Efimov in d dimensions\]). This CSI to DSI transition constitutes a quantum phase transition which appears for single body and strongly coupled many body systems and extends through non-relativistic, relativistic and Lifshitz dispersion relations. In a RG picture the transition describes universal low energy physics without fixed points and constitutes a physical realization of a limit cycle. Remarkably, the features of this transition have been measured recently in various systems such as cold atoms, graphene and Fermi gases [@Deng2016]. In the DSI phase, the dependence of the geometric ladder of states on the control parameter (see Table \[tab:QPTtable summary\]) is in the class of Berezinskii-Kosterlitz-Thouless transitions. This provides an interesting, yet to be studied, bridge between DSI and two dimensional systems associated with BKT physics. The characteristics described above provide the motivation to further study the ingredients associated with CSI to DSI transitions and we expect that these transitions will have an increasingly important role across the physics community in the future. Acknowledgements {#acknowledgements .unnumbered} ================ This work was supported by the Israel Science Foundation Grant No. 924/09 and by the Pazy Foundation. [^1]: [email protected] [^2]: [email protected] [^3]: For higher angular momentum channels $\lambda_{c}$ is larger and given by $(d-2)^{2}/4+l\left(l+d-2\right)$ [^4]: For higher angular momentum channels $\beta_c$ is larger and given by $ \left| K \right| $ where $K$ is defined as in [^5]: Here $z_{0}$ is not a phase like in (\[eq:k relation\]), a reflection of the fact that the solutions for $E$ would have an imaginary component corresponding to a finite lifetime.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Errors in implicative theories coming from binary data are studied. First, two classes of errors that may affect implicative theories are singled out. Two approaches for finding errors of these classes are proposed, both of them based on methods of Formal Concept Analysis. The first approach uses the cardinality minimal (canonical or Duquenne-Guigues) implication base. The construction of such a base is computationally intractable. Using an alternative approach one checks possible errors on the fly in polynomial time via computing closures of subsets of attributes. Both approaches are interactive, based on questions about the validity of certain implications. Results of computer experiments are presented and discussed.' author: - 'Sergei O. Kuznetsov' - Artem Revenko bibliography: - 'mathrefs.bib' - 'my\_refs.bib' title: Interactive Error Correction in Implicative Theories --- Introduction ============ ### Motivation Implicative theories consisting of formulas of the form “if $A$, then $B$” provide a standard way for describing the structure of domain knowledge. They are extensively used in various research areas, e.g., biology [@kknd11], pharmacology [@cla13; @Blinova2003], semantic web [@kltlkl12], etc. It is difficult to overestimate their importance for knowledge discovery [@frawley1992knowledge; @valtchev2004formal], decision making [@rasmussen1985role], classification [@mirkin1996mathematical], ontology engineering [@baader2010description]. In many cases the exactness of rules plays a crucial role, for example in research related to strictly formalized domains like Boolean algebras [@Kwuida2006], algebraic lattices [@Dau2000], or algebraic identities [@revenko2014automatized]. In many applications an exact implicative theory is constructed from a piece of available data. It is well-known that a single mistake in this data can drastically change the resulting implicative theory [@Ganter1999] (the same is true for association rules if there are some exceptions and an error). The implicative theory is not going to recover from this error even if further error-free data is added to the underlying set. Therefore, implicative theories are not error tolerant. However, in the real-world applications, especially if multiple users are expected to work with data, it is hardly imaginable to guarantee the absence of errors. More than that, someone may be willing to spoil the result on purpose. Therefore, a procedure for recovering from errors is essential for the usage of implicative theories. Here we assume that in the beginning there is already some data on hands and new data arrives in the work flow. The goal is to guarantee the correctness of the implicative theory with respect to the initial data which are considered to be reliable. We do not assume that a user, which is going to work with the data and the implicative theory, is always able to explicitly state any knowledge about data domain or has any knowledge about methods in use. That is why it is important to develop a transparent and easy method for error correction. In particular, it is important to find and output possible errors in a human understandable form. To attain this goal a natural framework can be that of Formal Concept Analysis (FCA) [@Ganter1999], where methods and algorithms for finding implicative theories of binary data (formal contexts) are well elaborated and widely used [@Ganter1984; @RyDiB11]. ### Related Work Methods for imputing missing values are well studied. In [@Song2007a] and [@Silva-Ramirez2011a] detailed overviews of existing techniques are presented. Among others there are techniques of ignoring entries with missing values, imputing average values, and more complicated ones such as decision trees, neural networks [@Silva-Ramirez2011a], Nearest Neighbours approach [@Jain2011a]. Having a missing value, there is no need to search for an error, as it is clear from the problem statement which value should be changed (or imputed). An approach proposed in this paper bares some similarity to the Nearest Neighbor method, but aims at solving a different task. Besides that, the imputation techniques (like, e.g. averaging) are mostly not relevant for binary data. Error finding and eliminating are widely discussed in various fields of computer science. The problems of lineage or data provenance, where one needs to explain errors, trace reasons for a query, etc. are well-known in KDD domain [@spg]. These techniques are very useful and efficient, however, they are not appropriate for correcting errors in binary data tables. In [@Fan2010a] an impressive way of using expert knowledge presented in the form of editing rules and certain regions for databases are surveyed. Information in the form of editing rules prevents the errors from getting in to the database. The approach presented in this paper aims at finding and correcting errors without any previously formalized knowledge. The paper [@bs09] presents an interesting approach to dealing with mistakes in answering questions (like the ones we will discuss below) in the process of knowledge base completion within the framework of Description Logics. This approach allows recovering from such mistakes in such an effective manner that the information input is used upon mistake recovery. However, the detection and correction of mistakes is left to pinpointing. Pinpointing is a very successful technique for recovering from inconsistencies. The goal of pinpointing is the following: for a given inconsistent set of rules (not only implicative) find maximal consistent subsets [@baader2007pinpointing; @meyer2006finding]. This technique is successfully applied in different description logics. The complexity of pinpointing is normally beyond polynomial. An approach introduced in this paper (Section \[find\], base approach) is closely related to pinpointing; it proceeds from knowledge base constructed from data. The complexity is also beyond polynomial. However, an alternative approach (Section \[find\], closure approach) takes the advantage of having the data and proposes a polynomial-time solution. In this work we do not modify the knowledge base directly, but we correct the errors in data in such a way that the corresponding implicative theory becomes error-free. As implicative theories is another view of Horn theories [@Ganter1999], the problem of finding explanations in Horn theories turns out to be closely connected to our problem. Namely, an entry in the binary data table can also be considered as a fact to be explained. In [@Kautz1993a] it is shown that such explanations may be found in polynomial time. However, here we aim at explaining existence or absence of all attributes at the same time. Also we state our task and our solutions in a different language and provide algorithms for practical usage. The case of negative attributes is not covered in [@Kautz1993a] as opposed to this work. The present paper is a follow-up work to [@Revenko2012c]. In this paper we assume that working with a data domain we can ask an expert in the domain whose answers are always correct. However, we should ask as few questions as possible. All sets and contexts we consider in this paper are assumed to be finite. ### Contributions We introduce an interactive procedure for making implicative theory of data error-free. This goal is achieved via finding and eliminating errors in rows of data tables. In FCA terms, we propose an approach for finding errors in descriptions of new objects (intents in terms of FCA) that affect the canonical implication base. 1. We introduce two possible classes of errors in data tables (formal contexts, Section \[err\_classes\]). 2. We introduce two approaches to finding and eliminating errors of certain classes (Sections \[find\]). Both aim at restoring dependencies from data domain and eliminating errors in implicative theory. 1. One approach is based on finding those implications from an implication base that are not respected by the new object intent (base approach). However, the base approach leads to an intractable solution, because constructing an implication base is intractable. 2. We introduce another approach (closure approach), where we do not need to compute the set of all implications, and prove its effectiveness. We show that it helps to find all possible errors of certain types (Proposition \[proposition1\]) in polynomial time (Proposition \[proposition2\]). The approaches are experimentally compared in Section \[experiment\]. Main Definitions ================ In what follows we keep to standard definitions of FCA [@Ganter1999]. Let $G$ and $M$ be sets and let $I\subseteq G\times M$ be a binary relation between $G$ and $M$. The triple $\context := \GMI$ is called a *(formal) context*. Set $G$ is called a set of *objects*, set $M$ is called a set of *attributes*, $I$ is called *incidence relation*. Consider mappings $\varphi\colon 2^G\to 2^M$ and $\psi\colon 2^M\to 2^G$: $\varphi(X) := \{m\in M\mid gIm \ \mbox{ for all\ } g\in X\},\ \psi(A) := \{g\in G\mid gIm \ \mbox{ for all\ } m\in A\}.$ Mappings $\varphi$ and $\psi$ define a *Galois connection* between $(2^G,\subseteq)$ and $(2^M,\subseteq)$, i.e. $\varphi(X) \subseteq A \Leftrightarrow \psi(A)\subseteq X$. Hence, for any $X_1, X_2\subseteq G$, $A_1, A_2\subseteq M$ one has 1. $X_1\subseteq X_2 \Rightarrow \varphi(X_2) \subseteq \varphi(X_1)$ \[prop1\] 2. $A_1\subseteq A_2 \Rightarrow \psi(A_2) \subseteq \psi(A_1)$ \[prop2\] 3. $X_1\subseteq \psi\varphi (X_1) \mbox{ and } A_1\subseteq \varphi\psi(A_1)$ \[monoton\] Usually, instead of $\varphi$ and $\psi$ a single notation $(\cdot)^{\prime}$ is used. $(\cdot)'$ is usually called a *derivation operator*. For $X \subseteq G$ the set $X^{\prime}$ is called the *intent* of $X$. Similarly, for $A \subseteq M$ the set $A^{\prime}$ is called the *extent* of $A$. Operator $(\cdot)''$ is idempotent, extensive and monotone, i.e., has the properties of algebraical closure both on $2^G$ and $2^M$. Hence, $(Z)^{\prime\prime}$ is called *closure* of $Z$ in $\context$ for $Z \subseteq M$ or $Z \subseteq G$. If $(Z)^{\prime\prime} = Z$, set $Z$ is called *closed* in $\context$. Applying Properties \[prop1\] and \[prop2\] consequently one gets the *monotonicity* property: for any $Z_1, Z_2 \subseteq G$ or $Z_1, Z_2 \subseteq M$ one has $Z_1 \subseteq Z_2 \Rightarrow Z_1'' \subseteq Z_2''$. In [@rodriguezgeneralized] authors introduce a generalized framework for considering positive and negative attributes. In this paper we also introduce negative attributes, however, we do not need the whole framework for our purpose. Our definitions comply with the definitions from [@rodriguezgeneralized]. The set $\overline{M} := \{\overline{m}\ |\ m \in M\}$ is called the set of *negative* attributes. Consider the following relation $\overline{I} := \{(g, \overline{m})\ |\ (g,m) \in (G \times M) \setminus I\}$ between $G$ and $M$. The context $\context^{\delta} := (G, M \cup \overline{M}, I \cup \overline{I})$ is called the *dichotomized context* to $\context$, the corresponding derivation operator is denoted by $(\cdot)^{\delta}$. Let $X \subseteq G$. Note that $\overline{m} \in X^\delta$ iff $m \notin g'$ for all $g \in X$. If $\overline{m} \in X^\delta$ then, as it does not lead to ambiguity, we informally write $\overline{m} \in X'$. In this paper objects and context are represented without negative attributes, however, in the processing stage they are normally converted to the dichotomized representation in order to be able to work with negative attributes. Consider the context $\context^{\overline{\delta}} = (G, \overline{M} \cup \overline{\overline{M}}, \overline{I} \cup \overline{\overline{I}})$. This context is isomorphic to the context $\context^{\delta} = (G, M \cup \overline{M}, I \cup \overline{I})$ and $\overline{\overline{m}} \in X^{\overline{\delta}}\ \Leftrightarrow\ m \in X^{\delta}$. A *formal concept* of a formal context $\GMI$ is a pair $(X, A)$, where $X \subseteq G,\ A\subseteq M,\ X'=A,$ and $A'=X$. The set $X$ is called the extent, and the set $A$ is called the intent of the concept $(X, A)$. One says that an object $g$ such that $g'\neq \emptyset$ is *reducible* in a context $\GMI$ iff $\exists X \subseteq G\setminus \{g\}:\ g' = \bigcap \limits_{j\in X} j'$. Removing reducible objects does not change the concept lattice up to isomorphism. In this paper implicative theories are formalized in terms of implication bases. An *implication* of $\context := \GMI$ is defined as a pair $(A,B)$, written $A\to B$, where $A, B\subseteq M$. $A$ is called the *premise*, $B$ is called the *conclusion* of implication $A\to B$. Implication $A\to B$ is *respected by a set of attributes* $N$ if $A \nsubseteq N$ or $B \subseteq N$. Implication $A\to B$ holds (is valid) in $\context$ if it is respected by all $g'$, $g\in G$, i.e. every object, that has all the attributes from $A$, also has all the attributes from $B$, or, equivalently, if $A' \subseteq B'$. Implications satisfy *Armstrong rules*: $${{}\over{A\to A}}\quad , \quad {{A\to B}\over{A\cup C\to B}}\quad , \quad {{A\to B, B\cup C\to D}\over{A\cup C\to D}}$$ *Support* of implication $A\to B$ in context $\context$ is $(A\cup B)'$, i.e., the set of all objects of $\context$, whose intents contain the premise and the conclusion of the implication. A *unit implication* is defined as an implication with only one attribute in conclusion, i.e. $A \to b$, where $A \subseteq M,\ b\in M$. Using Armstrong rules, every implication $A \to B$ can be represented as a set of *unit implications* $\{A \to b\ |\ b \in B\}$, so one can always observe only unit implications without loss of generality. Consider implications of the form $A \to \overline{b}$, where $A \subseteq M, \overline{b} \in \overline{M}$ in the dichotomized context $\context^{\delta}$. This implication is said to be respected by $N \subseteq M$ if $A \nsubseteq N$ or $b \in M \setminus N$. This implication holds in $\context^{\delta}$ iff $A^{\delta} \subseteq \overline{b}^{\delta}$. In this paper all the implications with negative attributes are considered as implications of the dichotomized context. An *implication base* of a context $\context$ is defined as a set $\mathfrak{L}$ of implications of $\context$, from which any valid implication for $\context$ can be deduced by Armstrong rules and none of the proper subsets of $\mathfrak{L}$ has this property. A cardinality minimal implication base was characterized in [@Duquenne1986] and is known as the *canonical implication base*, or *Duquenne-Guigues base*, or *stembase*. In [@Ganter1984] the premises of implications of the canonical base were characterized in terms of pseudo-intents. A subset of attributes $P\subseteq M$ is called a *pseudo-intent* if $P\not=P''$ and for every pseudo-intent $Q$ such that $Q\subset P$, one has $Q''\subset P$. The canonical implication base looks as follows: $\{ P \to (P''\setminus P) \mid P$ - pseudo-intent}. Errors in Implicative Theories {#err_classes} ============================== Without loss of generality we consider all observable properties to be expressed in terms of positive attributes from $M$. We aim at restoring valid implications and, therefore, correct errors in implicative theory of data. The goal is achieved if all implications are valid implications of the context. As already mentioned, all implications are reduced to unit ones. Consider the following possible classes of implicative formulas ($A \subseteq M,\ b, c\in M$), which will be called dependencies: 1. If an entity has all attributes from $A$, then it has attribute $b$ ($A \to b$); \[1class\] 2. If an entity has all attributes from $A$, then it does not have attribute $b$ ($A \to \overline b$); \[2class\] Only formulas of Class 1 are standard FCA implications, formulas of Class 2 are FCA implications if the negation of attributes are explicitly introduced in the context. If there are no errors in a context, all the dependencies of Class \[1class\] are deducible from an implication base. However, if not enough data is added to the context yet, we may get false consequences. Therefore, not all valid implications of the context have to necessarily be data domain dependencies. Nevertheless, it is guaranteed that none of valid dependencies is lost, and, as new objects are added, the number of false consequences is reduced (this is essentially the idea behind Attribute Exploration [@Ganter1999]). The situation is different if an erroneous object (data table row) is added. The erroneous object may violate a data domain dependency. In this case, until the error is found and corrected, we are not able to deduce all dependencies valid in the data domain from the implication base, no matter how many error-free objects are added afterwards. Finding Errors {#find} ============== We introduce two different approaches to finding errors. The first one is based on inspecting the canonical base of a context (base approach). When adding a new object to the context one may find all implications from the canonical base of the context such that the implications are not respected by the intent of the new object. These implications are then output as questions to an expert in form of implications. If at least one of these implications is accepted, the object intent is erroneous. Since the canonical base is the most compact (in the number of implications) representation of all valid implications of a context, it is guaranteed that minimal number of questions is asked and no valid dependencies of Class \[1class\] are left out. This approach can be seen as a version of pinpointing in the presence of only implicative rules. Although this approach allows one to reveal all dependencies of Class \[1class\], there are several issues. The problem of producing the canonical basis with known algorithms is intractable. Recent theoretical results  [@Kuznetsov2004On], [@Distel2011a], [@Kuznetsov2008Some], [@Kuznetsov2013Comp] suggest that the canonical base can hardly be computed with better worst-case complexity than that of the existing approaches [@Ganter1984]. One can use other bases (for example, there has been recent progress in computing proper premises [@RyDiB11]), but the algorithms known so far are still too costly and non-minimal bases do not guarantee that the expert is asked minimal sufficient number of questions. However, since we are only interested in implications corresponding to one object at a time, it may be not necessary to compute the whole implication base. The second approach takes this fact into account. Let $A \subseteq M$ be the intent of the object under inspection; we separate it from the context. $m \in A''$ iff $\forall g \in G: A \subseteq g' \Rightarrow m \in g'$, in other words, $A''$ contains the attributes common to all object intents containing $A$. The set of unit implications $\{A \to b\ |\ b \in A''\setminus A\}$ can then be shown to the expert. If all implications are rejected, no attributes are forgotten in the new object intent. Otherwise, there are missing attributes in the object intent. Unfortunately, this simple observation does not allow to correct all the errors in implicative theory. Consider Error4 from Fig. \[err\_cxt\]. Error4 has set of attributes $A$ = {has equal legs, has equal angles, all legs equal, at least 3 different legs}. The closure $A''$ in the context from Fig. \[cxt\] is equal to the set of all attributes $M$. Therefore, closure approach would ask if the user has forgotten to add all the attributes that are still possible to add. The suggestion to add all other attributes is not supported by any example in the context as there are no objects with all attributes. More than that, such solution is not minimal in general. Therefore, such a solution is not satisfactory. A more general description of the situation in the example above is the following. Let $A \subseteq M$ be the intent of the inspected object such that $\nexists g \in G: A \subseteq g^{\prime}$. In this case $A^{\prime\prime} = M$ and the implication $A \to A'' \setminus A$ has an empty support. We could try to solve this problem by allowing to ask only those questions that have a supporting example in the context. Consider again Error4 from Fig. \[err\_cxt\]. As support for every question is required only the following question would be asked: has equal legs, has equal angles, at least 3 different legs $\to$ at least 3 different angles? Support: {Quadrangle with 2 equal legs and 2 equal angles, Rectangular trapezium with 2 equal legs}. However, a smaller and more intuitive correction would be to suggest the user to remove the attribute “at least 3 different legs”. If this is indeed the source of error then even after adding the suggested attribute the error would not be eliminated and would impact the implicative theory. At this point we conclude that it is necessary to be able to suggest corrections for errors of Class \[2class\]. Such errors may be present if the object intent contains subset of attributes that none of the objects in the context has. Crucial Implications -------------------- Suppose we have a new object $g_n$ with intent $A$ and we want to see whether $A$ respects (is consistent with) the previous knowledge given by the context, i.e., does not have errors of Classes 1 or 2. In order to find errors of Class 1 we need to know, whether, according to implications (implicative dependencies) of the context, the new object should have more attributes than just $A$. If this is the case, there should be an implication $B \to c$ not respected by $A$: $B \subseteq A$, but $c \not\in A$. Similarly, in order to find errors of Class 2, we look for implications $B \to \overline{c}$ such that $B \subseteq A$, but $c \in A$. The following proposition shows that we do not need to look for all such implications, but for a much smaller subset of them. \[proposition1\] Let $\context = \GMI,\ g_n' = A, A \subseteq M$. Let $$\mathcal{I}_A^{(\context)} = \{B \to c\ |\ B\in \mathcal{MC}_A, c \in (B'' \setminus A) \cup (\overline{A \setminus B})\},$$ where $\mathcal{MC}_A = \{B \in \mathcal{C}_A\ |\ \nexists C \in \mathcal{C}_A: B \subset C\} \mbox{ and }\mathcal{C}_A = \{A \cap g'\ |\ g \in G\}$. The set $\mathcal{I}_A^{(\context)}$ contains (unit) implications with nonempty support that are valid in $\context$ and not respected by $A$. If an implication $(E \to d),\ E \subseteq A, d \in (M \setminus A) \cup \overline{A}$, with nonempty support is valid in $\context$, then there is an implication $(B \to d) \in {\cal I}_A^{(\context)}$ such that $E \subseteq B \subseteq A$. The last statement says that the set of implications ${\cal I}_A^{(\context)}$ is enough to deduce every attribute that can be deduced from implications of the context with nonempty support. Implications from ${\cal I}_A^{(\context)}$ are called *A-crucial* in $\context$. If ambiguity is excluded we omit the upper index and write simply ${\cal I}_A$. Let $(B\to c) \in {\cal I}_A$, hence $B'\subseteq c'$ by the definition of implication. By definition of ${\cal I}_A$ one has $B = A\cap g'$ for some $g\in G$. Then $B\subseteq g'$ and by the antimonotonicity of $(\cdot)'$ one has $g''\subseteq B'$. Hence, $g''\subseteq B' \subseteq c'$ and $c''\subseteq g''' = g'$. Since $c\in c''$, one has $c\in g'$. Since $c\in g'$ and $B'\subseteq g$, by the properties of $(\cdot)'$, one has $(B\cup c)' = B'\cap c' \subseteq g$. Hence, the support of $B\to c$ contains $g$ and is not empty. Consider the following possible cases: 1. $c \in B''\setminus A$. Since $B''\setminus A \not\subseteq A$ implication $B\to c$ is not respected by $A$; 2. $c \in \overline{A\setminus B}$. Since $A\setminus B\subseteq A$, one has $c\not \in A$ and $B\to c$ is not respected by $A$. Now let $E \to d$ be a valid implication not respected by $A$ with a nonempty support. Then $E \subseteq A, d\notin A$ and there exists $g \in G$ such that $E\subseteq g', d \in g'$. Therefore, there exists $B_{\mathcal{C}_A} \in \mathcal{C}_A$ such that $B_{\mathcal{C}_A} = A \cap g'$ and $E \subseteq B_{\mathcal{C}_A}$. Moreover, there exists $B_{\mathcal{MC}_A} \in \mathcal{MC}_A$ such that $B_{\mathcal{C}_A} \subseteq B_{\mathcal{MC}_A}$. By construction $B_{\mathcal{MC}_A} \subseteq A$, therefore, $E \subseteq B_{\mathcal{MC}_A} \subseteq A$. Consider the following possible cases: 1. $d \in M$. As $E \subseteq B_{\mathcal{MC}_A}$ by the properties of $(\cdot)''$ one has $E'' \subseteq B_{\mathcal{MC}_A}''$. By definition of the validity of an implication one has $d \in E''$, hence, $d \in B_{\mathcal{MC}_A}''$. Therefore, $(B_{\mathcal{MC}_A} \to d) \in {\cal I}_A$; 2. $d \in \overline{M}$. Let $\overline{c} = d$. For any $B \in \mathcal{MC}_A$ there exists $g_* \in G$ such that $B = A \cap g_*'$. If $E \subseteq g_*'$ then by the validity of the implication one has $d \in g_*'$, hence, $c \notin g_*'$. Therefore, $c \notin B$. As $d \notin A$ one has $c \in A$. Hence, $c \in A\setminus B$ and $d \in \overline{A\setminus B}$. Therefore, $(B_{\mathcal{MC}_A} \to d) \in {\cal I}_A$. \[proposition2\] For a new object $g$ with intent $A$ one has ${\cal I}_A \leq |G|\times |M|$. By definition of $\mathcal{MC}_A$ it contains no more than $|G|$ elements. For any $A, B \subseteq M$ one has $|B''| \leq |M|$, hence $|B'' \setminus A| \leq |M| - |A|$; $|\overline{A \setminus B}| \leq |A|$. Hence, $|(B'' \setminus A) \cup (\overline{A \setminus B})|\ \leq\ |(B'' \setminus A)| + |(\overline{A \setminus B})|\ \leq\ |M| - |A| + |A| = |M|$. Therefore, ${\cal I}_A$ contains not more than $|G|\times |M|$ implications. According to Proposition \[proposition2\] to check errors of Classes 1 and 2 one has to consider polynomialy many implications instead of exponentially many implications in the cardinality-minimum canonical base [@Kuznetsov2004On]. Proposition \[proposition1\] allows one to design an algorithm for computing the set of questions (in form of implications) that can help to reveal possible errors of Classes \[1class\] and \[2class\]. \[proposition3\] Let $g$ be a new object with intent $A$. ${\cal I}_A$ can be computed in $O(|G|^2\times |M|)$ time. Consider the following `inspect_closure` algorithm \[pseudocode\] Candidates $= \{$object$'\cap A\ |\ $object $\in G\}$\ MaxCandidates $= \{C \in$ Candidates$\ |\ \nexists B \in $ Candidates$:\ C \subseteq B\}$\ Result $= \emptyset$\ Result Here $A$ is the intent of the new object. In line 3 the algorithm computes the set of all subsets that are candidates for the premises of crucial implications. In line 4 all non-maximal subsets are discarded. In lines 6 and 7 closures of the premises are computed and the corresponding implications are added to the set of crucial implications. To estimate the worst-case complexity of the algorithm, note that executing line 1 and line 3 take at most O$(|G|\times |M|)$ time, line 4 takes O$(|M|)$ time for each of O$(|G|^2)$ containment tests, and lines 6 and 7 take O$(|G|\times |M|)$ time for computing closures of at most O$(|G|)$ premises of crucial implications. Hence, the total worst-case time complexity is O$(|G|^2\times |M|)$. If there are several new objects, the set of crucial implications for each new object is in general dependent on the order of adding objects. The following statement shows which additional questions should be asked in order to compensate for this dependency. \[two\_new\_objects\] Let $g_1$ and $g_2$ be new objects with intents $A_1$ and $A_2$, respectively. Let $\context_1 = (G \cup g_1, M, I \cup \{(g_1, m)\ |\ m \in A_1\})$. If $\nexists g \in G: A_1 \cap A_2 \subseteq g'$ then $I_{A_2}^{(\context_1)} \setminus I_{A_2}^{(\context)} = \{A_2 \cap A_1 \to m\ |\ m \in (A_1 \setminus A_2) \cup (\overline{A_2 \setminus A_1})\}$, otherwise $I_{A_2}^{(\context_1)} = I_{A_2}^{(\context)}$. By definition of $\cal{I}_A^{(\context)}$ only maximal intersections may become premises of implications. Hence, if there exists $g \in G$ such that $A_1 \cap A_2 \subseteq g'$ then no new implications can arise. However, if such $g$ does not exist, the set of new implications, by definition of $\cal{I}_A^{(\context)}$, is $\{(A_1 \cap A_2) \to m\ |\ m \in (A_1 \cap A_2)'' \setminus A_2 \cup \overline{A_2 \setminus (A_1 \cap A_2)}\}$. As $A_2 \setminus (A_1 \cap A_2) = A_2 \setminus A_1$ and, by assumption about maximal intersection, $(A_1 \cap A_2)'' = A_1$, the set of new implications is $\{A_2 \cap A_1 \to m\ |\ m \in (A_1 \setminus A_2) \cup (\overline{A_2 \setminus A_1})\}$. Obviously, no more than $|(A_1 \setminus A_2) \cup (A_2 \setminus A_1)|$ additional questions may arise. It is also easy to see that additional implications only arise in case where both new objects have maximal intents in $\context$. However, as we do not require any information about maximal intents in data domain, we have to be careful when adding an object with maximal intent. The following corollary states clearly which implications should be considered in order to guarantee the absence of errors that may affect the implication base. \[max\_intent\] Let $g_n$ be a new object with intent $A$. The sets of implications $I_1 = \{A \to m\ |\ m \in M\setminus A,\,\nexists g \in G: A \subseteq g'\}$ and $I_2 = \{(A-a) \to \overline{a}\ |\ a \in A,\,\nexists g \in G: (A-a) \subseteq g'\}$ are valid in $\context$, have empty support, and are not respected by $A$. Note that if there exists $g \in G$ such that $A \subseteq g'$ then both sets $I_1$ and $I_2$ are empty. By definition $B \to c$ is respected by $N\subseteq M$ if $B\nsubseteq N$. By definition of $I_1$ and $I_2$ all the premises are not contained in any object intents from the context. Therefore, implications are valid, however, they are not supported by any of the object intents. As $A\subseteq A,\, (M\setminus A) \nsubseteq A$, implications from $I_1$ are not respected by $A$. As $\forall a\in A: (A-a)\subseteq A,\, a \notin (M\setminus A)$, implications from $I_2$ are not respected by $A$. According to Proposition \[max\_intent\] the number of additional questions for new objects that have maximal intents cannot exceed $|M|$. As none of the questions have objects from context in support we suggest that maximal objects should be checked “by hand”. For the sake of compactness in what follows we present implications in non-unit form. The name `inspect_base` is used to denote the function implementing base approach. Example ------- Consider the following example with convex quadrangles. Formal context given by the cross-table in Fig. \[cxt\] contains convex quadrangles and their properties. The context does not cover the domain completely, i.e. not all possible convex quadrangle types are considered. Attributes “has equal legs” and “has equal angles” require at least two angles/legs of a quadrangle to be equal. Some dependencies on attributes are obvious, e.g., it is clear that if all angles are equal in a quadrangle, then this quadrangle definitely has equal angles. Four objects are in the context of tentative errors in Fig \[err\_cxt\]. These objects are added to the context in Fig. \[cxt\] one at a time. Inspecting Case1: [`inspect_base`]{}\ at least 3 different angles $\to$ at least 3 different legs\ all legs equal $\to$ has equal angles, has equal legs [`inspect_closure`]{}\ has equal legs, at least 3 different angles $\to$ at least 3 different legs,\ [$\overline{\mbox{all legs equal}}$]{}\ has equal legs, all legs equal $\to$ has equal angles, [$\overline{\mbox{at least 3 different angles}}$]{} Both algorithms reveal possible errors in a similar manner, although there are obvious differences. In the output of `inspect_base` the premises are smaller than in the output of `inspect_closure`. The latter also reveals dependencies of Class \[2class\]. It is easy to see that all output implications hold in data domain. For example, if all legs are equal in a quadrangle, it should have equal angles and should not have 3 different angles. Hence, this object should be recognized as an error, it should be corrected to rhombus or to quadrangle with two equal legs. Inspecting Case2: [`inspect_base`]{}\ all angles equal $\to$ has equal angles, has equal legs, has right angle\ all legs equal $\to$ has equal angles, has equal legs [`inspect_closure`]{}\ has right angle, has equal legs, all legs equal, all angles equal $\to$ has equal angles In this example we are able to ask even less questions to an expert using `inspect_closure` as with `inspect_base`. This is the result of finding implications generated by maximal subsets of object’s intent. The intent of Case2 occurs in the context (in the intent of Square), that is why we do not get any negative attributes in the output of `inspect_closure`. Again, all implications are valid in data domain, therefore, Case2 is an error, it should be corrected to square. Inspecting Case3: [`inspect_base`]{}\ all angles equal $\to$ has equal angles, has equal legs, has right angle\ all legs equal $\to$ has equal angles, has equal legs [`inspect_closure`]{}\ has equal angles, has right angle, at least 3 different legs, at least 3 different angles $\to$ [$\overline{\mbox{all angles equal}}$]{}, [$\overline{\mbox{all legs equal}}$]{}\ has equal angles, has right angle, all legs equal, all angles equal $\to$ has equal legs, [$\overline{\mbox{at least 3 different angles}}$]{}, [$\overline{\mbox{at least 3 different legs}}$]{} In Case3 we get both implications from the output of `inspect_base` combined in one implication with a bigger premise in the output of `inspect_closure`. In addition we obtain several implications with negative attributes. It is easy to see that all implications hold in the data domain, therefore, Case3 is an error and should be corrected either to rectangular trapezium or to square. Inspecting Case4: [`inspect_base`]{}\ has equal angles, has equal legs, at least 3 different legs, all legs equal $\to$ has right angle, at least 3 different angles, all angles equal [`inspect_closure`]{}\ has equal angles, has equal legs, all legs equal $\to$ [$\overline{\mbox{at least 3 different legs}}$]{}\ has equal angles, has equal legs, at least 3 different legs $\to$ at least 3 different angles, [$\overline{\mbox{all legs equal}}$]{} Case4 is a very special case where the corresponding implication from canonical base has empty support. In the output of `inspect_base` we obtain all questions possible for this intent. As discussed above these questions are not based on any information input so far. The reason for that is that Case4 has maximal intent in the context. So these questions could also be found using Proposition \[max\_intent\]. However, even if we add attributes “at least 3 different angles” and “all angles equal” and reject the last implication we would not be able to recognize this object as an error. On the contrary `inspect_closure` allows us to recognize errors of Class \[2class\] and state that Case4 should be corrected to have the intent of rhombus or quadrangle with two equal legs and two equal angles. Experiment ========== Below the results of experiments on synthetic data are presented. The experiments were conducted as follows: all objects are first taken one by one out from the context and then added as new objects; all the possible errors of Classes 1 and 2 are found and output. This experiment is run for the purpose of testing the efficiency of the algorithm, *not* as an attempt to find errors. An FCA package for Python was used for implementation ([@RevenkoPyFCA]). For computing the canonical base an optimized algorithm based on `Next Closure` was used ([@Obiedkov2007]). All tests described below were run on computer with Intel Core i7 1.6GHz processor and 4 Gb of RAM running Linux Ubuntu 11.10 x64. In Fig. \[graph\] the results of running both algorithms on synthetic contexts are presented. For each context the number of objects is equal to 50. Parameter $d$ represents the density of the context, i.e. the probability of having a cross in the cross-table representing the relation. This result is presented in the semi-logarithmic scale. It is easy to note that with the growth of the number of attributes and the density, the difference between runtime of two algorithms grows as well. coordinates [ (10, 0.167965544595) (12, 0.247017635239) (14, 0.375766846869) (16, 0.543842315674) (18, 0.789092620214) (20, 1.14722710186) (22, 1.62289366457) (24, 2.12516090605) (26, 2.75543173154) (28, 3.8202943537) (30, 5.37453644805) ]{}; coordinates [ (10, 0.0073512395223) (12, 0.00683691766527) (14, 0.00727897220188) (16, 0.00737159781986) (18, 0.00684918297662) (20, 0.008645468288) (22, 0.0088245206409) (24, 0.008004811075) (26, 0.00940683152941) (28, 0.00915070374807) (30, 0.010860840479) ]{}; coordinates [ (10, 0.344418048859) (12, 0.709496061007) (14, 1.25616751777) (16, 1.86424203714) (18, 3.54632928636) (20, 5.22309121821) (22, 9.10261064106) (24, 12.0801133447) (26, 17.8718641599) (28, 24.3004840745) (30, 35.9667735497) ]{}; coordinates [ (10, 0.00849596659342) (12, 0.0115488237805) (14, 0.0140699677997) (16, 0.016548289193) (18, 0.0211648543676) (20, 0.0263027217653) (22, 0.0325202941895) (24, 0.0387960539924) (26, 0.0466770728429) (28, 0.0502827829785) (30, 0.0553248738183) ]{}; coordinates [ (10, 1.11483882533) (12, 3.05852622456) (14, 8.86141722732) (16, 22.6997146871) (18, 61.8963118394) (20, 157.937982559) (22, 295.927075134) (24, 646.7) (26, 1424) ]{}; coordinates [ (10, 0.0178514321645) (12, 0.0246753030353) (14, 0.0366503265169) (16, 0.0489775472217) (18, 0.0676475498411) (20, 0.0750198496713) (22, 0.10060983234) (24, 0.1235) (26, 0.1364) (28, 0.15) (30, 0.164) ]{}; ------------- ------------------ inspect\_base inspect\_closure $\bullet$ $d=0.1$ $\triangle$ $d=0.3$ $\times$ $d=0.6$ ------------- ------------------ Another experiment was conducted to test the quality of finding errors by the introduced method. The information about dependencies between negative attributes is not reflected in the implication base. Therefore, more implications are usually violated by objects having more attributes in their intent. Small intents usually violate only few implications. However, in this experiment we aim at finding not only the errors affecting the implication base; therefore, it is necessary to level out the shift between larger and smaller intents. For this purpose in the following experiments a slight modification of the introduced method is used. The complementary context to a context $\context = \GMI$ is defined as $\context^c := (G, M, (G\times M) \setminus I)$. The method applied to the complementary context will output implications with only negative attributes in the premise. Running the introduced method on both original context and complementary context yields better results. Note that implications with both positive and negative attributes will not be generated. The experiments were conducted in the following settings. An object was picked up from a context, from one to three errors were randomly introduced into the intent of the object. The method was used to find possible errors in the object. If all the erroneous attributes were in the conclusion of the unit implications with *the same* premise then the errors were marked as found. In this case all the erroneous attributes are contained in one question to the user. Afterwards the object already without errors is returned to the context and the next object is picked up. We considered three contexts from the UCI repository ([@Frank2010]): `SPECT`, `house-votes-84` and `kr-vs-kp`. Therefore, nine experiments were conducted. In every experiment 1000 objects (with possible repetitions) were picked up one after another. In Table \[table:quality\] the results of the experiment are presented. Context Name $|G|$ $|G|$ after reducing $|M|$ ------------------ ------- ---------------------- ------- `SPECT` 267 133 23 `house-votes-84` 232 104 17 `kr-vs-kp` 3198 453 36 : Error finding experiment carried out on three contexts from UCI.[]{data-label="table:quality"} \ --- ----- ------- ------ ------ -- 1 548 0.548 2298 2.41 2 242 0.242 2753 2.80 3 131 0.131 2703 2.71 --- ----- ------- ------ ------ -- : Error finding experiment carried out on three contexts from UCI.[]{data-label="table:quality"} --- ----- ------- ------- ------ -- 1 712 0.712 9780 11.4 2 217 0.217 14018 14.7 3 71 0.071 18276 18.4 --- ----- ------- ------- ------ -- : Error finding experiment carried out on three contexts from UCI.[]{data-label="table:quality"} --- ----- ------- ------- ------ -- 1 786 0.786 7520 8.47 2 393 0.393 12863 13.2 3 247 0.247 18322 18.3 --- ----- ------- ------- ------ -- : Error finding experiment carried out on three contexts from UCI.[]{data-label="table:quality"} 0.5cm ${}^1$All objects containing missing values were removed. 0.5cm ${}^2$Attribute 15 was removed due to many-valuedness. As can be seen from the results the more objects there are in the context, the better method works. In `SPECT` intents are very diverse (there are only 5.78 irreducible objects per attribute on average) that is why not more than three implications on average are output and bad ratio of found errors is obtained. In `house-votes-84` intents are more similar, that is why we have more questions per object. The ratio of found errors for one error is relatively high, however, it quickly drops with the increase of errors, as the number of irreducible objects is small. In `kr-vs-kp` there are much more objects per attribute and the results for two and three errors at a time are much better. However, if the user is able to correctly answer all unit implications even better results can be achieved. In this case the user may correct first errors and repeat the procedure having already only one error. In these experiments the error-finding process was considered successful only if there is one implication suggesting all the needed corrections at once. It is worth noting that the chance of random guess in predicting all errors in an object description is only $1/(|M|^n)$ if there are $n$ errors as compared to 50% for the classification task. Conclusion ========== A method for finding errors in implicative theories was introduced. The method uses some techniques based on Formal Concept Analysis. As opposed to finding the canonical (cardinality minimal) base of implications, which can be very time consuming due to intrinsic intractability, the proposed algorithm terminates in polynomial time. Moreover, after checking maximal object descriptions (object intents) “by hand” it is possible to find all errors of two considered types or prove their absence. Computer experiments show that in practice the proposed method works much faster than that based on the generation of the implication base. ### Acknowledgements The first author was supported by German Academic Exchange Service (DAAD). The second author was supported by the Basic Research Program of the National Research University Higher School of Economics, project Mathematical models, algorithms, and software for knowledge discovery in structured and text data. We thank Bernhard Ganter and Sergei Obiedkov for discussion and useful remarks.
{ "pile_set_name": "ArXiv" }
--- abstract: | Motivated by symmetric Cauchy matrices, we define symmetric Cauchy tensors and their generating vectors in this paper. Hilbert tensors are symmetric Cauchy tensors. An even order symmetric Cauchy tensor is positive semi-definite if and only if its generating vector is positive. An even order symmetric Cauchy tensor is positive definite if and only if its generating vector has positive and mutually distinct entries. This extends Fiedler’s result for symmetric Cauchy matrices to symmetric Cauchy tensors. Then, it is proven that the positive semi-definiteness character of an even order symmetric Cauchy tensor can be equivalently checked by the monotone increasing property of a homogeneous polynomial related to the Cauchy tensor. The homogeneous polynomial is strictly monotone increasing in the nonnegative orthant of the Euclidean space when the even order symmetric Cauchy tensor is positive definite. Furthermore, we prove that the Hadamard product of two positive semi-definite (positive definite respectively) symmetric Cauchy tensors is a positive semi-definite (positive definite respectively) tensor, which can be generalized to the Hadamard product of finitely many positive semi-definite (positive definite respectively) symmetric Cauchy tensors. At last, bounds of the largest H-eigenvalue of a positive semi-definite symmetric Cauchy tensor are given and several spectral properties on Z-eigenvalues of odd order symmetric Cauchy tensors are shown. Further questions on Cauchy tensors are raised. [**Keywords:**]{} positive semi-definiteness, positive definiteness, Cauchy tensor, eigenvalue, generating vector. [**AMS Subject Classification(2000):**]{} 90C30, 15A06. author: - 'Haibin Chen[^1], Liqun Qi [^2]' title: '**Positive Definiteness and Semi-Definiteness of Even Order Symmetric Cauchy Tensors**' --- Introduction ============ Let $\mathbb{R}^n$ be the $n$ dimensional real Euclidean space and the set consisting of all natural numbers be denoted by $N$. Suppose $m$ and $n$ are positive natural numbers and denote $[n]=\{1,2,\cdots,n\}$. A Cauchy matrix (maybe not square) is an $m\times n$ structure matrix assigned to $m+n$ parameters $x_1,x_2,\cdots, x_m,y_1,\cdots, y_n$ as follows: [@Polya] $$\label{e11} C=\left[\frac{1}{x_i+y_j}\right],~i\in [m],j\in [n].$$ The Cauchy matrix has been studied and applied in algorithm designing [@T93; @G95; @Go94]. When $x_i=y_i$ in (\[e11\]), it is a real symmetric Cauchy matrix. Stimulated by the notion of symmetric Cauchy matrices, we give the following definition. \[def11\] Let vector $c=(c_1,c_2,\cdots,c_n)\in \mathbb{R}^n$. Suppose that a real tensor $\mathcal{C}=(c_{i_1i_2\cdots i_m})$ is defined by $$c_{i_1i_2\cdots i_m}=\frac{1}{c_{i_1}+c_{i_2}+\cdots+c_{i_m}},\quad j\in [m],~i_j \in [n].$$ Then, we say that $\mathcal{C}$ is an order $m$ dimension $n$ symmetric Cauchy tensor and the vector $c\in \mathbb{R}^n$ is called the generating vector of $\mathcal{C}$. We should point out that, in Definition \[def11\], for any $m$ elements $c_{i_1}, c_{i_2}, \cdots, c_{i_m}$ in generating vector $c$, it satisfies $$c_{i_1}+c_{i_2}+\cdots+c_{i_m}\neq 0,$$ which implies that $c_i\neq 0$, $i\in [n]$. By Definition \[def11\], a dimension $n\times n$ real symmetric Cauchy matrix is an order 2 dimension $n$ real symmetric Cauchy tensor. It is easy to check that every principal subtensors of a symmetric Cauchy tensor is a symmetric Cauchy tensor with a generating vector being a subvector of the generating vector of the original symmetric Cauchy tensor. In this paper, we always consider $m$-th order $n$ dimensional real symmetric Cauchy tensors. Hence, it can be called Cauchy tensors for simplicity. Apparently, Cauchy tensors are a class of structured tensors. In recent years, a lot of research papers on structured tensors are presented [@Chen14; @Ding13; @Ding14; @HH14; @Qi14; @qi14; @QXX; @Song14; @song14; @YY14; @Zhang12], which studied M-tensors, circulant tensors, completely positive tensors, Hankel tensors, Hilbert tensors, P tensors and B tensors. These papers not only established results on spectral theory and positive semi-definiteness property of structured tensors, but also gave some important applications of structured tensors in stochastic process and data fitting [@Chen14; @Ding14]. Actually, Cauchy tensors have close relationships with Hankel tensors and Hilbert tensors. Suppose Cauchy tensor $\mathcal{C}$ and its generating vector $c$ are defined as in Definition \[def11\]. If $$c_{i_1}+c_{i_2}+\cdots+c_{i_m} \equiv c_{j_1}+c_{j_2}+\cdots+c_{j_m}$$ whenever $$i_1 + i_2+\cdots+i_m = j_1+j_2+\cdots+j_m,$$ then Cauchy tensor $\mathcal{C}$ is a Hankel tensor in the sense of [@Qi14]. In general, a symmetric Cauchy tensor is not a Hankel tensor. If entries of $c$ are defined such that $$c_i=i-1+\frac{1}{m},~i\in [n],$$ then Cauchy tensor $\mathcal{C}$ is a Hilbert tensor according to [@Song14]. In this paper, we are interested in the positive semi-definiteness conditions and positive definiteness conditions for even order Cauchy tensors, in addition, several spectral properties of positive semi-definite Cauchy tensors are given. In the next section, we prove that an even order Cauchy tensor is positive semi-definite if and only if its generating vector is positive. An even order Cauchy tensor is positive definite if and only if its generating vector has positive and mutually distinct entries. This extends Fiedler’s result for symmetric Cauchy matrices in 2010 [@Fied10] to symmetric Cauchy tensors. After that, it is shown that an even order Cauchy tensor is positive semi-definite if and only if there is a monotone increasing homogeneous polynomial defined in the nonnegative orthant of $\mathbb{R}^n$. And the homogeneous polynomial is strictly monotone increasing when the even order Cauchy tensor is positive definite. Later in Section 2, we show that the Hadamard product of two positive semi-definite (positive definite respectively) Cauchy tensors is a positive semi-definite (positive definite respectively) tensor, which can be generalized to the Hadamard product of finitely many positive semi-definite (positive definite respectively) Cauchy tensors. In Section 3, several spectral inequalities are presented on the largest H-eigenvalue and the smallest H-eigenvalue of a Cauchy tensor. When a Cauchy tensor is positive semi-definite, bounds of its largest H-eigenvalue are given. Then, for an odd order Cauchy tensor, we prove that the corresponding Z-eigenvector is nonnegative if the Z-eigenvalue is positive and the Z-eigenvector is non-positive if the Z-eigenvalue is negative. Furthermore, if the generating vector of an odd order Cauchy tensor is positive, the Cauchy tensor has nonzero Z-eigenvalues. We conclude this paper with some final remarks in Section 4. Some questions are also presented, which can be considered in the future. By the end of the introduction, we add some comments on the notation that will be used in the sequel. Vectors are denoted by italic lowercase letters i.e. $x,~ y,\cdots$, and tensors are written as calligraphic capitals such as $\mathcal{A}, \mathcal{T}, \cdots.$ Denote $e_i\in \mathbb{R}^n$ as the $i$-th unit coordinate vector for $i\in [n]$, and ${\bf0}$ as the zero vector in $\mathbb{R}^n$. Suppose $x=(x_1,x_2,\cdots,x_n),~y=(y_1,y_2,\cdots,y_n)$. Then $x\geq y$ ($x\leq y$) means $x_i\geq y_i$ ($x_i\leq y_i$) for all $i\in [n]$. If both $\mathcal{A}=(a_{i_1 \cdots i_m})_{1\leq i_j\leq n}$ and $\mathcal{B}=(b_{i_1 \cdots i_m})_{1\leq i_j\leq n}$, $j=1,\cdots,m$, are tensors, then $\mathcal{A}\geq \mathcal{B}$ ($\mathcal{A}\leq \mathcal{B}$) means $a_{i_1 \cdots i_m} \geq b_{i_1 \cdots i_m}$ ($a_{i_1 \cdots i_m} \leq b_{i_1 \cdots i_m}$) for all $i_1,\cdots,i_m \in [n]$. Positive Semi-definite Cauchy Tensors ===================================== An order $m$ dimension $n$ tensor $\mathcal{A}=(a_{i_1i_2\cdots i_m})$ is called positive semi-definite if for any vector $x\in \mathbb{R}^n$, it satisfies $$\mathcal{A}x^m=\sum_{i_1,i_2,\cdots,i_m\in [n]}a_{i_1i_2\cdots i_m}x_{i_1}x_{i_2}\cdots x_{i_n}\geq 0.$$ $\mathcal{A}$ is called positive definite if $\mathcal{A}x^m>0$ for all nonzero vector $x\in \mathbb{R}^n$. Similarly, tensor $\mathcal{A}$ is negative semi-definite (negative definite) if $\mathcal{A}x^m\leq 0$ ($\mathcal{A}x^m<0$ for all nonzero vector $x\in \mathbb{R}^n$). It is obvious that minus positive semi-definite tensors (minus positive definite tensors respectively) are negative semi-definite tensors(negative definite tensors respectively). Clearly, there is no odd order nonzero positive semi-definite tensors. In this section, we will give some sufficient and necessary conditions for even order Cauchy tensors to be positive semi-definite or positive definite. Some conditions are extended naturally from the Cauchy matrix case. \[thm21\] Assume a Cauchy tensor $\mathcal{C}$ is of even order. Let $c\in \mathbb{R}^n$ be the generating vector of $\mathcal{C}$. Then Cauchy tensor $\mathcal{C}$ is positive semi-definite if and only if $c>0$. For necessity, suppose that an even order Cauchy tensor $\mathcal{C}$ is positive semi-definite. It is easy to check that all composites of generating vector $c$ are positive since $$\mathcal{C}e_i^m=\frac{1}{mc_i}\geq0,~i\in [n]$$ where $e_i$ is the $i$-th coordinate vector of $\mathbb{R}^n$. So, $c_i>0$ for all $i\in [n]$, which means $c>0$. On the other hand, assume that $c>0$. For any $x\in \mathbb{R}^n$, it holds that $$\begin{array}{rl} \mathcal{C}x^m=& \sum_{i_1,\cdots,i_m \in [n]}c_{i_1i_2\cdots i_m}x_{i_1}x_{i_2} \cdots x_{i_m}\\ =&\sum_{i_1,\cdots,i_m \in [n]} \frac{x_{i_1}x_{i_2} \cdots x_{i_m}}{c_{i_1}+c_{i_2}+\cdots+c_{i_m}} \\ =&\sum_{i_1,\cdots,i_m \in [n]}\displaystyle \int^1_0 t^{c_{i_1}+c_{i_2}+\cdots+c_{i_m}-1}x_{i_1}x_{i_2} \cdots x_{i_m}dt\\ =&\displaystyle \int^1_0\left(\sum_{i\in [n]}t^{c_i-\frac{1}{m}}x_i\right)^m dt\\ \geq &0. \end{array}$$ Here the last inequality follows that $m$ is even. By the arbitrariness of $x$, we know that Cauchy tensor $\mathcal{C}$ is positive semi-definite and the desired result holds. \[corol21\] Assume that even order Cauchy tensor $\mathcal{C}$ and its generating vector $c\in \mathbb{R}^n$ be defined as in Theorem \[thm21\]. Then Cauchy tensor $\mathcal{C}$ is negative semi-definite if and only if $c<0$. \[corol22\] Assume that even order Cauchy tensor $\mathcal{C}$ and its generating vector $c\in \mathbb{R}^n$ be defined as in Theorem \[thm21\]. Then Cauchy tensor $\mathcal{C}$ is not positive semi-definite if and only if there exist at least one negative element in $c$. From the results about H-eigenvalues and Z-eigenvalues in [@Qi05], we have the following result. \[corol23\] Assume that even order Cauchy tensor $\mathcal{C}$ and its generating vector $c\in \mathbb{R}^n$ be defined as in Theorem \[thm21\]. If $c>0$, then all the H-eigenvalues and Z-eigenvalues of Cauchy tensor $\mathcal{C}$ are nonnegative. \[thm22\] Assume even order Cauchy tensor $\mathcal{C}$ has generating vector $c=(c_1,c_2,\cdots,c_n)\in \mathbb{R}^n$. Suppose $c_1,c_2,\cdots,c_n$ are positive and mutually distinct, then Cauchy tensor $\mathcal{C}$ is positive definite. For the sake of simplicity, without loss of generality, assume that $$0<c_1<c_2<\cdots<c_n.$$ since $c>0$ and $c_1,c_2,\cdots,c_n$ are mutually distinct. From Theorem \[thm21\], we know that Cauchy tensor $\mathcal{C}$ is positive semi-definite. We prove by contradiction that Cauchy tensor $\mathcal{C}$ is positive definite when the conditions of this theorem hold. Assume there exists a nonzero vector $x\in \mathbb{R}^n$ such that $$\mathcal{C}x^m=0.$$ By the proof of Theorem \[thm21\], one has $$\displaystyle \int^1_0\left(\sum_{i\in [n]}t^{c_i-\frac{1}{m}}x_i\right)^m dt=0,$$ which means $$\sum_{i\in [n]}t^{c_i-\frac{1}{m}}x_i\equiv 0,~t\in [0, 1].$$ Thus $$x_1+t^{c_2-c_1}x_2+\cdots+t^{c_n-c_1}x_n\equiv 0,~t\in (0, 1].$$ By continuity and the fact that $c_1,c_2,\cdots,c_n$ are mutually distinct, it holds that $$x_1=0$$ and $$x_2+t^{c_3-c_2}x_3+\cdots+t^{c_n-c_2}x_n\equiv 0,~t\in (0, 1].$$ Repeat the process above. We obtain $$x_1=x_2=\cdots=x_n=0,$$ which is a contradiction with $x\neq0$. So, for all nonzero vectors $x\in \mathbb{R}^n$, it holds $\mathcal{C}x^m>0$ and $\mathcal{C}$ is positive definite. From this theorem, we easily have the following corollary, which was first proved in [@Song14]. \[corol234\] An even order Hilbert tensor is positive definite. From Theorem A of [@Fied10], we know that a symmetric Cauchy matrix $$C=\left[\frac{1}{c_i+c_j}\right]$$ is positive definite if and only if all the $c_i$’s are positive and mutually distinct. In fact, the theorem below shows that conditions in Theorem \[thm22\] is also a sufficient and necessary condition, which is a natural extension of Theorem A of [@Fied10], by Fielder. \[thm23\] Let even order Cauchy tensor $\mathcal{C}$ and its generating vector $c$ are defined as in Theorem \[thm22\]. Then, Cauchy tensor $\mathcal{C}$ is positive definite if and only if the elements of generating vector are positive and mutually distinct. By Theorem \[thm22\], we only need to prove the “only if” part of this theorem. Suppose that Cauchy tensor $\mathcal{C}$ is positive definite. Firstly, by Theorem \[thm21\], we know that $$c_i>0,~i\in [n].$$ We now prove by contradiction that $c_i$’s are mutually distinct. Suppose that two elements of $c$ are equal. Without loss of generality, assume $c_1=c_2=a>0$. Let $x\in \mathbb{R}^n$ be a vector with elements $x_1=1, x_2=-1$ and $x_i=0$ for the others. Then, one has $$\begin{array}{rl} \mathcal{C}x^m=& \sum_{i_1,\cdots,i_m \in [n]}c_{i_1i_2\cdots i_m}x_{i_1}x_{i_2} \cdots x_{i_m}\\ =&\sum_{i_1,\cdots,i_m \in [n]} \frac{x_{i_1}x_{i_2} \cdots x_{i_m}}{c_{i_1}+c_{i_2}+\cdots+c_{i_m}} \\ =&\frac{1}{ma}\sum_{i_1,\cdots,i_m \in [2]}x_{i_1}x_{i_2} \cdots x_{i_m}\\ =&\frac{1}{ma} \left[(-1)^m+m(-1)^{m-1}+\frac{m!}{2!(m-2)!}(-1)^{m-2}+\cdots+(-1)^{m-m}\right]\\ =&\frac{1}{ma}\left[1+(-1)\right]^m\\ =&0, \end{array}$$ where we get a contradiction with the assumption that Cauchy tensor $\mathcal{C}$ is positive definite. Thus, elements of generating vector $c$ are mutually distinct and the desired result follows. We denote the homogeneous polynomial $\mathcal{C}x^m$ as $$f(x)=\mathcal{C}x^m=\sum_{i_1,\cdots,i_m \in [n]}c_{i_1i_2\cdots i_m}x_{i_1}x_{i_2} \cdots x_{i_m}.$$ For all $x,y \in X\subseteq\mathbb{R}^n$, if $f(x)\geq f(y)$ when $x\geq y$($x\leq y$), we say that $f(x)$ is monotone increasing (monotone decreasing respectively) in $X$. If $f(x)>f(y)$ when $x\geq y$, $x\neq y$($x\leq y$, $x\neq y$), we say that $f(x)$ is strictly monotone increasing (strict monotone decreasing respectively) in $X$. The following conclusion means that the positive semi-definite property of a Cauchy tensor is equivalent to the monotonicity of a homogeneous polynomial respected to the Cauchy tensor in $\mathbb{R}^n_+$. \[thm24\] Let $\mathcal{C}$ be an even order Cauchy tensor with generating vector $c$. Then, $\mathcal{C}$ is positive semi-definite if and only if $f(x)$ is monotone increasing in $\mathbb{R}^n_+$. For sufficiency, let $x=e_i, y=0$ and $x\geq y$. Then we have $$\frac{1}{mc_i}=\mathcal{C}x^m=f(x)\geq f(y)=\mathcal{C}y^m=0,$$ which implies that $c_i>0$ for $i\in [n]$. By Theorem \[thm21\], it holds that Cauchy tensor $\mathcal{C}$ is positive semi-definite. For necessary conditions, suppose $x,y\in \mathbb{R}^n_+$ and $x\geq y$. Then, we know that $$\begin{array}{rl} f(x)-f(y)=&\mathcal{C}x^m-\mathcal{C}y^m\\ =&\sum_{i_1,\cdots,i_m \in [n]}c_{i_1i_2\cdots i_m}(x_{i_1}x_{i_2} \cdots x_{i_m}-y_{i_1}y_{i_2} \cdots y_{i_m})\\ =&\sum_{i_1,\cdots,i_m \in [n]}\frac{x_{i_1}x_{i_2} \cdots x_{i_m}-y_{i_1}y_{i_2} \cdots y_{i_m}}{c_{i_1}+c_{i_2}+\cdots+c_{i_m}}\\ \geq &0. \end{array}$$ Here, the last inequality follows that $x\geq y$ and the fact that $c_i>0,$ for $i\in [n]$, which means that $f(x)$ is monotone increasing in $\mathbb{R}^n_+$ and the desired result holds. \[lema21\]Let $\mathcal{C}$ be an even order Cauchy tensor with generating vector $c$. Suppose $\mathcal{C}$ is positive definite, then the homogeneous polynomial $f(x)$ is strictly monotone increasing in $\mathbb{R}^n_+$ From the condition that $\mathcal{C}$ is positive definite, by Theorem \[thm23\], we have $$c_i>0,~i\in [n],$$ where scalars $c_i,~i\in [n]$ are entries of generating vector $c$. For any $x,y\in \mathbb{R}^n_+$ satisfying that $x\geq y$ and $x\neq y$, there exists index $i\in [n]$ such that $$x_i>y_i\geq0.$$ Then, it holds that $$\begin{array}{rl} f(x)-f(y)=&\mathcal{C}x^m-\mathcal{C}y^m\\ =&\sum_{i_1,\cdots,i_m \in [n],(i_1,i_2,\cdots,i_m)\neq(i,i,\cdots,i)}c_{i_1i_2\cdots i_m}(x_{i_1}x_{i_2} \cdots x_{i_m}-y_{i_1}y_{i_2} \cdots y_{i_m})\\ &+c_{ii\cdots i}(x_i^{m}-y_i^m)\\ =&\sum_{i_1,\cdots,i_m \in [n],(i_1,i_2,\cdots,i_m)\neq(i,i,\cdots,i)}\frac{x_{i_1}x_{i_2} \cdots x_{i_m}-y_{i_1}y_{i_2} \cdots y_{i_m}}{c_{i_1}+c_{i_2}+\cdots+c_{i_m}}\\ &+\frac{1}{mc_i}(x_i^{m}-y_i^m)\\ >&0, \end{array}$$ which implies that the homogeneous polynomial $f(x)$ is strictly monotone increasing in $\mathcal{R}^n_+$. Now, we give an example to show that the strictly monotone increasing property for the polynomial $f(x)$ is only a necessary condition for the positive definiteness property of Cauchy tensor $\mathcal{C}$ but not a sufficient condition. \[exam31\] Let Cauchy tensor $\mathcal{C}=(c_{i_1i_2i_3i_4})$ with dimension 3 and it has generating vector $c=(1,1,1)$. Then, $$c_{i_1i_2i_3i_4}=\frac{1}{4},~i_1,i_2,i_3,i_4\in [3]$$ and the homogeneous polynomial $$f(x)=\mathcal{C}x^4=\frac{1}{4}\sum_{i_1,i_2,i_3,i_4\in [3]}x_{i_1}x_{i_2}x_{i_3}x_{i_4}.$$ By direct computation, we know that $f(x)$ is strictly monotone increasing in $\mathbb{R}^3_+$. From Theorem \[thm23\], Cauchy tensor $\mathcal{C}$ is not positive definite. Let $r_i$ denote the sum of the $i$-th row elements of Cauchy tensor $\mathcal{C}$, which can be written such that $$r_i=\sum_{i_2,\cdots,i_m \in [n]}\frac{1}{c_i+c_{i_2}+\cdots+c_{i_m}},~i\in [n],$$ where $c=(c_1,\cdots,c_n)$ is the generating vector of Cauchy tensor $\mathcal{C}$. Suppose $$R=\max_{1\leq i\leq n}r_i,~r=\min_{1\leq i\leq n}r_i.$$ If Cauchy tensor $\mathcal{C}$ is positive semi-definite, by Theorem \[thm21\], it is easy to check that $$R=\sum_{i_2,\cdots,i_m \in [n]}\frac{1}{\underline{a}+c_{i_2}+\cdots+c_{i_m}},~r=\sum_{i_2,\cdots,i_m \in [n]}\frac{1}{\bar{a}+c_{i_2}+\cdots+c_{i_m}},$$ where $\underline{a}=\min_{1\leq i \leq n}c_i,~\bar{a}=\max_{1\leq i\leq n}c_i$. Now, before giving the next conclusion, we give the definition of eigenvalues of tensors and the definition of irreducible tensors, which will be used in the sequel. The definition of eigenvalue-eigenvector pairs of real symmetric tensors comes from [@Qi05]. \[def21\] Let $\mathbb{C}$ be the complex field. A pair $(\lambda, x)\in \mathbb{C}\times \mathbb{C}^n\setminus \{0\}$ is called an eigenvalue-eigenvector pair of a real symmetric tensor $\mathcal{T}$ with order $m$ dimension $n$, if they satisfy $$\label{e21} \mathcal{T}x^{m-1}=\lambda x^{[m-1]},$$ where $\mathcal{T}x^{m-1}=\left(\sum_{i_2,\cdots,i_m=1}^n t_{ii_2\cdots i_m}x_{i_2}\cdots x_{i_m} \right)_{1\leq i\leq n}$ and $x^{[m-1]}=(x_i^{m-1})_{1\leq i\leq n}$ are dimension $n$ vectors. In Definition \[def21\], if $\lambda\in \mathbb{R}$ and the corresponding eigenvector $x\in \mathbb{R}^n$, we call $\lambda,~x$ H-eigenvalue and H-eigenvector respectively. Let $\rho(\mathcal{C})$ denote the spectral radius of Cauchy tensor $\mathcal{C}$. Then $\rho(\mathcal{C})>0$ and $\rho(\mathcal{C})$ is an H-eigenvalue of $\mathcal{C}$ when $\mathcal{C}$ is positive semi-definite The following definition is consistent with [@Chang08] and [@Qi13] respectively. \[def22\] For a tensor $\mathcal{T}$ with order m and dimension n. We say that $\mathcal{T}$ is reducible if there is a nonempty proper index subset $I\subset [n]$ such that $$t_{i_1i_2\cdots i_m}=0,~\forall ~i_1\in I,~\forall~i_2,i_3,\cdots,i_m \notin I.$$ Otherwise we say that $\mathcal{T}$ is irreducible. \[thm25\] Let even order Cauchy tensor $\mathcal{C}$ be positive semi-definite with generating vector $c\in \mathbb{R}^n$. Suppose $x\in \mathbb{R}^n$ is the eigenvector of $\mathcal{C}$ corresponding to $\rho(\mathcal{C})$. Assume $$\label{e22} r_{\bar{i}}=\max_{1\leq i \leq n}r_i,~r_{\underline{i}}=\min_{1\leq i\leq n}r_i.$$ Then, $R=r_{\bar{i}},~r=r_{\underline{i}}.$ Since Cauchy tensor $\mathcal{C}$ is positive semi-definite, from Theorem \[thm21\], all elements of $\mathcal{C}$ and $c$ are positive. By Definition \[def22\], we know that $\mathcal{C}$ is irreducible. Thus $x>0$ from Theorem 1.4 of [@Chang08]. Without loss of generality, suppose $$R=r_l,~r=r_s.$$ By the analysis before this theorem, it holds that $$\bar{a}=\max_{1\leq i\leq n}c_i=c_s,~\underline{a}=\min_{1\leq i \leq n}c_i=c_l.$$ On the other side, by (\[e21\]) of Definition \[def21\], we have $$\begin{array}{rl} \rho(\mathcal{C})x_{\bar{i}}^{m-1}=&(\mathcal{C}x^{m-1})_{\bar{i}}\\ =&\sum_{i_2,\cdots,i_m \in [n]}c_{\bar{i}i_2i_3\cdots i_m}x_{i_2} \cdots x_{i_m} \\ =&\sum_{i_2,\cdots,i_m \in [n]}\frac{x_{i_2} \cdots x_{i_m}}{c_{\bar{i}}+c_{i_2}+\cdots+c_{i_m}} \\ \leq& \sum_{i_2,\cdots,i_m \in [n]}\frac{x_{i_2} \cdots x_{i_m}}{\underline{a}+c_{i_2}+\cdots+c_{i_m}}\\ =&(\mathcal{C}x^{m-1})_l\\ =&\rho(\mathcal{C})x^{m-1}_l,\\ \end{array}$$ which implies that $$x_{\bar{i}}=x_l$$ So we can take $\bar{i}=l$ and $R=r_{\bar{i}}$ holds. Similarly, by (\[e21\]) and (\[e22\]), one has $$\begin{array}{rl} \rho(\mathcal{C})x_{\underline{i}}^{m-1}=&(\mathcal{C}x^{m-1})_{\underline{i}}\\ =&\sum_{i_2,\cdots,i_m \in [n]}c_{\underline{i}i_2i_3\cdots i_m}x_{i_2} \cdots x_{i_m} \\ =&\sum_{i_2,\cdots,i_m \in [n]}\frac{x_{i_2} \cdots x_{i_m}}{c_{\underline{i}}+c_{i_2}+\cdots+c_{i_m}} \\ \geq& \sum_{i_2,\cdots,i_m \in [n]}\frac{x_{i_2} \cdots x_{i_m}}{\bar{a}+c_{i_2}+\cdots+c_{i_m}}\\ =&(\mathcal{C}x^{m-1})_s\\ =&\rho(\mathcal{C})x^{m-1}_s,\\ \end{array}$$ which means that $x_{\underline{i}}=x_s$ and we can take $\underline{i}=s$. Thus $r=r_{\underline{i}}$ and the desired results follows. \[thm26\] Suppose even order Cauchy tensor $\mathcal{C}$ has positive generating vector $c\in \mathbb{R}^n$. Then $\mathcal{C}$ is positive definite if and only if $r_1,r_2,\cdots,r_n$ are mutually distinct. By conditions, all elements of $c$ are positive, so it is obvious that $r_1,r_2,\cdots,r_n$ are mutually distinct if and only if $c_1,c_2,\cdots,c_n$ are mutually distinct. By Theorem \[thm23\], the desired conclusion follows. Suppose $\mathcal{A}=(a_{i_1i_2\cdots i_m})$ and $\mathcal{B}=(b_{i_1i_2\cdots i_m})$ are two order $m$ dimension $n$ tensors, the Hadamard product of $\mathcal{A}$ and $\mathcal{B}$ is defined as $$\mathcal{A}\circ \mathcal{B}=(a_{i_1i_2\cdots i_m}b_{i_1i_2\cdots i_m}),$$ which is still an order $m$ dimension $n$ tensor. \[thm27\] Let $\mathcal{C}$ and $\mathcal{C}'$ be two order $m$ dimension $n$ Cauchy tensors with generating vectors $c=(c_1,c_2,\cdots,c_n)$ and $c'=(c_1',c_2',\cdots,c_n')$ respectively. Suppose $m$ is even. If Cauchy tensors $\mathcal{C}$ and $\mathcal{C}'$ are positive semi-definite, then $\mathcal{C}\circ \mathcal{C}'$ is a positive semi-definite tensor. By conditions, from Theorem \[thm21\], it follows that $$c>0,~c'>0.$$ Let $a'=c_1'+c_2'+\cdots+c_n'>0$. For any vector $x\in \mathbb{R}^n$, by the definition of Hadamard product of tensors, we have $$\begin{array}{rl} (\mathcal{C}\circ \mathcal{C}')x^m=& \sum_{i_1,\cdots,i_m \in [n]}c_{i_1i_2\cdots i_m}c_{i_1i_2\cdots i_m}'x_{i_1}x_{i_2} \cdots x_{i_m}\\ =&\sum_{i_1,\cdots,i_m \in [n]} \frac{x_{i_1}x_{i_2} \cdots x_{i_m}}{(c_{i_1}+c_{i_2}+\cdots+c_{i_m})(c_{i_1}'+c_{i_2}'+\cdots+c_{i_m}')} \\ =&\sum_{i_1,\cdots,i_m \in [n]}\displaystyle \int^1_0 t^{(c_{i_1}+c_{i_2}+\cdots+c_{i_m})(c_{i_1}'+c_{i_2}'+\cdots+c_{i_m}')-1}x_{i_1}x_{i_2} \cdots x_{i_m}dt\\ =&\sum_{i_1,\cdots,i_m \in [n]}\displaystyle \int^1_0 t^{c_{i_1}a'+c_{i_2}a'+\cdots+c_{i_m}a'-1}x_{i_1}x_{i_2} \cdots x_{i_m}dt\\ =&\displaystyle \int^1_0(\sum_{i\in [n]}t^{c_ia'-\frac{1}{m}}x_i)^m dt\\ \geq &0. \end{array}$$ From the arbitrariness of vector $x$, we know that $\mathcal{C}\circ \mathcal{C}'$ is positive semi-definite. \[corol24\] Suppose $\mathcal{C}_1,~\mathcal{C}_2,\cdots,\mathcal{C}_l$ are order $m$ dimension $n$ positive semi-definite Cauchy tensors. Assume $m$ is even. Then, $\mathcal{C}_1\circ\mathcal{C}_2\circ\cdots\circ\mathcal{C}_l$ is a positive semi-definite tensor. By Theorems \[thm22\] and \[thm23\], we have the following conclusion. We omit its proof since it is similar to the proof of Theorem \[thm22\]. \[thm28\] Let $\mathcal{C}$, $\mathcal{C}'$, $c$ and $c'$ be defined as in Theorem \[thm27\]. If Cauchy tensors $\mathcal{C}$ and $\mathcal{C}'$ are positive definite, then $\mathcal{C}\circ \mathcal{C}'$ is a positive definite tensor. Inequalities for Cauchy Tensors =============================== In this section, we give several inequalities about the largest and the smallest H-eigenvalues of Cauchy tensors. The bounds for the largest H-eigenvalues are given for positive semi-definite Cauchy tensors. Moreover, properties of Z-eigenvalues and Z-eigenvectors of odd order Cauchy tensors are also shown. It should be noted that a real symmetric tensor always has Z-eigenvalues and an even order real symmetric tensor always has H-eigenvalues [@Qi05]. We denote the largest and smallest H-eigenvalues of Cauchy tensor $\mathcal{C}$ by $\lambda_{max}$ and $\lambda_{min}$ respectively. When $\mathcal{C}$ is a positive semi-definite Cauchy tensor, then by the Perron-Frobenius theory of nonnegative tensors [@Chang08], we have $$\lambda_{max}=\rho(\mathcal{C}).$$ \[lema31\] Assume $\mathcal{C}$ is a Cauchy tensor with generating vector $c$. If the entries of $c=(c_1,c_2,\cdots,c_n)$ have different signs, then, $$\lambda_{min}\leq \frac{1}{m \max \{c_i|~c_i<0, i\in [n]\}}<0<\frac{1}{m \min \{c_i|~c_i>0, i\in [n]\}}\leq \lambda_{max}.$$ From Theorem 5 of [@Qi05], we have $$\lambda_{max}=\max \left\{\mathcal{C}x^m|~\sum_{i \in [n]}x_i^m=1,~x\in \mathbb{R}^n\right\}$$ and $$\lambda_{min}=\min \left\{\mathcal{C}x^m|~\sum_{i \in [n]}x_i^m=1,~x\in \mathbb{R}^n\right\}.$$ Combining this with the fact that $$\mathcal{C}e_i^m=\frac{1}{mc_i},~i\in [n],$$ we have the conclusion of the lemma. Let $r$, $R$, $\bar{a}$ and $\underline{a}$ be defined as in Section 2. We have the following result. \[thm31\] Assume even order Cauchy tensor $\mathcal{C}$ has generating vector $c=(c_1,c_2,\cdots,c_n)$. Suppose $c>0$ and at least two elements of $c$ are different. Then $$r+\frac{1}{m\bar{a}}\left(\sqrt{\frac{R}{r}}-1\right)< \lambda_{max} < R-\frac{1}{m\bar{a}}\left(1-\sqrt{\frac{r}{R}}\right).$$ Suppose $x\in \mathbb{R}^n$ is the eigenvector of $\mathcal{C}$ corresponding to $\lambda_{max}$. By conditions, Cauchy tensor $\mathcal{C}$ is an irreducible nonnegative tensor and it follows $x>0$ from Theorem 1.4 of [@Chang08]. Without loss of generality, let $x=(x_1,x_2,\cdots,x_n)$ and suppose $0<x_i\leq1,~i\in [n]$ such that $$\label{e31} x_s=\min_{i\in [n]}x_i>0,~x_l=\max_{j\in [n]}x_j=1.$$ By Theorem \[thm25\], we have $$R=r_l,~r=r_s$$ and $R> r$ since at least two entries of $c$ are not equal. On the other side, by the definition of eigenvalues, from (\[e31\]), one has $$\label{e32} \begin{array}{rl} \lambda_{max}x_s^{m-1}=&(\mathcal{C}x^{m-1})_s\\ =&\sum_{i_2,\cdots,i_m \in [n]}c_{si_2\cdots i_m}x_{i_2}x_{i_3}\cdots x_m\\ \leq &\sum_{i_2,\cdots,i_m \in [n]}c_{si_2\cdots i_m} \\ =&r, \end{array}$$ and $$\label{e33} \begin{array}{rl}\lambda_{max}=& \lambda_{max}x_l^{m-1} \\ =&(\mathcal{C}x^{m-1})_l\\ =&\sum_{i_2,\cdots,i_m \in [n]}c_{li_2\cdots i_m}x_{i_2}x_{i_3}\cdots x_m\\ \geq& x_s^{m-1}\sum_{i_2,\cdots,i_m \in [n]}c_{li_2\cdots i_m} \\ =&Rx_s^{m-1}. \end{array}$$ Thus, by (\[e32\]) and (\[e33\]), we have $$0<x_s^{m-1}\leq\frac{\lambda_{max}}{R}\leq\frac{r}{x_s^{m-1}R},$$ which can be written as $$0<x_s^{m-1}\leq \sqrt{\frac{r}{R}}.$$ Combining this with (\[e31\]), we obtain $$\begin{array}{rl}\lambda_{max}=& \lambda_{max}x_l^{m-1} \\ =&(\mathcal{C}x^{m-1})_l\\ =&\sum_{i_2,\cdots,i_m \in [n], (l,i_2,\cdots,i_m)\neq(l,s,\cdots, s)}c_{li_2\cdots i_m}x_{i_2}x_{i_3}\cdots x_m +c_{ls\cdots s}x_s^{m-1}\\ <& \sum_{i_2,\cdots,i_m \in [n]}c_{li_2\cdots i_m}-c_{ls\cdots s}+c_{ls\cdots s}\sqrt{\frac{r}{R}}\\ <&R-\frac{1}{m\bar{a}}(1-\sqrt{\frac{r}{R}}), \end{array}$$ and $$\begin{array}{rl}\lambda_{max}=&\frac{1}{x_s^{m-1}}\sum_{i_2,\cdots,i_m \in [n]}c_{si_2\cdots i_m}x_{i_2}x_{i_3}\cdots x_m\\ =&\frac{c_{sl\cdots l}x_l^{m-1}}{x_s^{m-1}}+\frac{1}{x_s^{m-1}}\sum_{i_2,\cdots,i_m \in [n],(i_2i_3\cdots i_m)\neq (ll\cdots l)}c_{si_2\cdots i_m}x_{i_2}x_{i_3}\cdots x_m\\ =&\frac{c_{sl\cdots l}}{x_s^{m-1}}+\frac{1}{x_s^{m-1}}\sum_{i_2,\cdots,i_m \in [n],(i_2i_3\cdots i_m)\neq (ll\cdots l)}c_{si_2\cdots i_m}x_{i_2}x_{i_3}\cdots x_m\\ >&\sqrt{\frac{R}{r}}c_{sll\cdots l}+r-c_{sll\cdots l}\\ >&r+\frac{1}{m\bar{a}}(\sqrt{\frac{R}{r}}-1), \end{array}$$ from which we get the desired inequalities. In [@Qi05], Qi called a real number $\lambda$ and a real vector $x\in \mathbb{R}^n$ a Z-eigenvalue of tensor $\mathcal{A}$ and a Z-eigenvector of $\mathcal{A}$ corresponding to $\lambda$, it they are solutions of the following system: $$\label{e34} \begin{cases} \mathcal{A}x^{m-1}=\lambda x\\ x^Tx=1 \end{cases}.$$ Next, we will give several spectral properties for odd order Cauchy tensors. \[thm32\] Suppose order $m$ dimension $n$ Cauchy tensor $\mathcal{C}$ has generating vector $c$. Let $m$ be odd and $c>0$. Assume $\lambda\in \mathbb{R}$ is a Z-eigenvalue of $\mathcal{C}$ with Z-eigenvector $x=(x_1,x_2,\cdots,x_n)\in \mathbb{R}^n$. If Z-eigenvalue $\lambda>0$, then $x\geq0$; if Z-eigenvalue $\lambda<0$, then $x\leq0$. By the condition $c>0$, we know that all entries of Cauchy tensor $\mathcal{C}$ are positive. By definitions of Z-eigenvalue and Z-eigenvector, for any $i\in [n]$, we have that $$\label{e35} \begin{array}{rl} \lambda x_i=&(\mathcal{C}x^{m-1})_i\\ =&\sum_{i_2,i_3,\cdots,i_m \in [n]}c_{ii_2\cdots i_m}x_{i_2}x_{i_3}\cdots x_m\\ =&\sum_{i_2,i_3,\cdots,i_m \in [n]}\frac{x_{i_2}x_{i_3}\cdots x_m}{c_i+c_{i_2}+\cdots+c_{i_m}}\\ =&\sum_{i_2,i_3,\cdots,i_m \in [n]}\displaystyle \int^1_0 t^{c_i+c_{i_2}+\cdots+c_{i_m}-1}x_{i_2}x_{i_3}\cdots x_m dt\\ =&\displaystyle \int^1_0 t^{c_i-\frac{1}{m}} (\sum_{j\in [n]}t^{c_j-\frac{1}{m}}x_j)^{m-1}dt. \end{array}$$ Since $m$ is odd, by (\[e35\]), one has $$\lambda x_i\geq 0,~~\mbox{for}~~ i\in [n],$$ which implies that $x\geq0$ when $\lambda>0$ and $x\leq o$ when $\lambda<0$. \[thm33\] Suppose Cauchy tensor $\mathcal{C}$ and its generating vector $c$ are defined as in Theorem \[thm32\]. If all entries of $c$ are mutually distinct, then $\mathcal{C}$ has no zero Z-eigenvalue. By conditions, since entries of generating vector $c$ are mutually distinct, without loss of generality, suppose $$0<c_1<c_2<\cdots<c_n.$$ We prove the result by contradiction. Suppose $\mathcal{C}$ has Z-eigenvalue $\lambda=0$ with Z-eigenvector $x\in \mathbb{R}^n$. Then, by (\[e35\]), for any $i\in [n]$, we have $$\displaystyle \int^1_0 t^{c_i-\frac{1}{m}} \left(\sum_{j\in [n]}t^{c_j-\frac{1}{m}}x_j\right)^{m-1}dt\equiv0.$$ From properties of integration, one has $$t^{c_i-\frac{1}{m}} \left(\sum_{j\in [n]}t^{c_j-\frac{1}{m}}x_j\right)^{m-1}\equiv0,~ t\in [0,1],$$ i.e., $$\label{e36} t^{c_i-\frac{1}{m}}\left(t^{c_1-\frac{1}{m}}x_1+t^{c_2-\frac{1}{m}}x_2+\cdots+t^{c_n-\frac{1}{m}}x_n\right)\equiv0,~t\in [0,1].$$ By (\[e36\]), we obtain $$t^{c_1-\frac{1}{m}}x_1+t^{c_2-\frac{1}{m}}x_2+\cdots+t^{c_n-\frac{1}{m}}x_n\equiv0,~t\in (0,1],$$ which implies that $$\label{e37}x_1+t^{c_2-c_1}x_2+\cdots+t^{c_n-c_1}x_n\equiv0,~t\in (0,1].$$ Since $c_1,c_2,\cdots,c_m$ are mutually distinct, by the continuity property of operators and (\[e37\]), it follows that $$x_1=0.$$ Thus, the equation (\[e37\]) can be written as $$t^{c_2-c_1}x_2+\cdots+t^{c_n-c_1}x_n\equiv0,~t\in (0,1],$$ which is equivalent to $$x_2+t^{c_3-c_2}x_3+\cdots+t^{c_n-c_2}x_n\equiv0,~t\in (0,1].$$ By the continuity property, we have $x_2=0.$ Repeating the process above, we get $$x_1=x_2=\cdots=x_n=0,$$ which is contradicting with the fact that $x$ is a Z-eigenvector corresponding to $\lambda=0$. The desired conclusion follows. Final Remarks ============= In this paper, we give several necessary and sufficient conditions for an even order Cauchy tensor to be positive semi-definite. Some properties of positive semi-definite Cauchy tensors are presented. Furthermore, inequalities about the largest H-eigenvalue and the smallest H-eigenvalue of Cauchy tensors are shown. At last, some spectral properties on Z-eigenvalues of odd order Cauchy tensors are shown. However, there are still some questions that we are not sure now. The Cauchy matrix can be combined with many other structured matrices to form new structured matrices such as Cauchy-Toeplitz matrix and Cauchy-Hankel matrix [@Solak03; @Tyr91; @Tyr92]. Can we get the type of Cauchy-Toeplitz tensors and Cauchy-Hankel tensors? If so, how about their spectral properties? What are the necessary and sufficient conditions for their positive semi-definiteness? [99]{} K.C. Chang, K. Pearson, T. Zhang, [*Perron Frobenius Theorem for nonnegative tensors*]{}, Commu. Math. Sci. 6(2008) 507-520. K.C. Chang, K. Pearson, T. Zhang, [*On eigenvalue problems of real symmetric tensors*]{}, J. Math. Anal. Appl. 350(2009) 416-422. Z. Chen, L. Qi, [*Circulant Tensors with Applications to Spectral Hypergraph Theory and Stochastic Process*]{}, arXiv preprint arXiv:1312.2752v7.2013. W. Ding, L. Qi, Y. Wei, [*M-Tensors and Nonsingular M-Tensors*]{}, Lin. Alg. Appl., 439 (2013) 3264-3278. W. Ding, L. Qi, Y. Wei, [*Fast Hankel Tensor-Vector Products and Application to Exponential Data Fitting*]{}, arXiv preprint arXiv:1401.6238, 2014. M. Fiedler, [*Notes on Hilbert and Cauchy matrices*]{}, Lin. Alg. Appl., 432(1)(2010) 351-356. T. Finck, G. Heinig, K. Rost, [*An inversion formula and fast algorithms for Cauchy-Vandermonde matrices*]{}, Lin. Alg. Appl., 183(1993) 179-191. I. Gohberg, V. Olshevsky, [*Fast algorithms with preprocessing for matrix-vector multiplication problems*]{}, J Complexity, 10(4)(1994) 411-427. J. He, T.Z. Huang, [*Inequalities for M-tensors*]{}, Journal of Inequality and Applications, 2014:114. G. Heinig, [*Inversion of generalized Cauchy matrices and other classes of structured matrices*]{}, Linear Algebra for Signal Processing, Springer New York, (1995) 63-81. G. P$\acute{o}$lya, G. Szeg$\ddot{o}$, [Zweiter Band]{}, Springer, Berlin, 1925. L. Qi, [*Eigenvalue of a real supersymmetric tensor*]{}, J. Symb. Comput. 40(2005) 1302-1324. L. Qi, [*H$^+$-eigenvalues of Laplacian and signless Laplacian tensors*]{}, Communications in Mathematical Sciences, 12 (2014) 1045-1064. L. Qi, [*Hankel tensors: Associated Hankel matrices and Vandermonde decomposition*]{}, Communications in Mathematical Sciences, 12 (2014). L. Qi, Y. Song, [*An even order symmetric B tensor is positive definite*]{}, to appear in: Linear Algebra and Its Applications. L. Qi, C. Xu, Y. Xu, [*Nonnegative tensor factorization, completely positive tensors and an hierarchical elimination algorithm*]{}, arXiv preprint arXiv:1305.5344, 2013. S. Solak, D. Bozkruk, [*On the spectral norms of Cauchy-Toeplitz and Cauchy-Hankel matrices*]{}, Appl. Math. Comput., 140(2)(2003) 231-238. Y. Song, L. Qi, [*Some properties of infinite and finite dimension Hilbert tensors*]{}, Linear Algebra and Its applications, 451 (2014) 1-14. Y. Song, L. Qi, [*An Initial Study on P, P $ _0 $, B and B $ _0 $ Tensors*]{}, arXiv preprint arXiv:1403.1118, 2014. E.E. Tyrtyshnikov, [*Cauchy-Toeplitz matrices and some applications*]{}, Lin. Alg. Appl., 149(1991) 1-18. E.E. Tyrtyshnikov, [*Singular values of Cauchy-Toeplitz matrices*]{}, Lin. alg. appl., 1992, 161: 99-116. P. Yuan, L. You, [*Some remarks on P, P$_0$, B and B$_0$ tensors*]{}, arXiv preprint arXiv:1405.1288, 2014. L. Zhang, L. Qi, G. Zhou, [*M-tensors and some applications*]{}, SIAM Journal on Matrix Analysis and Applications, 35 (2014) 437-452. [^1]: Department of Applied Mathematics, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong. Email: [email protected]. [^2]: Department of Applied Mathematics, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong. Email: [email protected]. This author’s work was supported by the Hong Kong Research Grant Council (Grant No. PolyU 502510, 502111, 501212 and 501913).
{ "pile_set_name": "ArXiv" }
--- abstract: 'We analyze randomized benchmarking for arbitrary gate-dependent noise and prove that the exact impact of gate-dependent noise can be described by a single perturbation term that decays exponentially with the sequence length. That is, the exact behavior of randomized benchmarking under general gate-dependent noise converges exponentially to a true exponential decay of exactly the same form as that predicted by previous analysis for gate-independent noise. Moreover, we show that the operational meaning of the decay parameter for gate-dependent noise is essentially unchanged, that is, we show that it quantifies the average fidelity of the noise between ideal gates. We numerically demonstrate that our analysis is valid for strongly gate-dependent noise models. We also show why alternative analyses do not provide a rigorous justification for the empirical success of randomized benchmarking with gate-dependent noise.' author: - 'Joel J. Wallman' date: 'December 18, 2017' title: 'Randomized benchmarking with gate-dependent noise' --- Introduction ============ The development of practical, large-scale devices that process quantum information relies upon techniques for efficiently characterizing the quality of control operations on the quantum level. Fully characterizing quantum processes [@Chuang1997; @Poyatos1997] requires resources that scale exponentially with the number of qubits despite improvements such as compressed sensing and direct fidelity estimation [@DaSilva2011; @Flammia2011; @Flammia2012; @Reich2013; @Kliesch]. However, some figures of merit can be efficiently estimated without fully characterizing quantum processes. The canonical such figure is the fidelity of an experimental implementation $\tilde{{\ensuremath{\mathcal{G}}}}$ of an ideal unitary channel ${\ensuremath{\mathcal{G}}}(\rho) = G\rho G{\ensuremath{^{\dagger}}}$ averaged over input states distributed according to the Haar measure $d\psi$, $$\begin{aligned} \label{eq:fidelity} f(\tilde{{\ensuremath{\mathcal{G}}}},{\ensuremath{\mathcal{G}}}) = \int d\psi \operatorname{Tr}[{\ensuremath{\mathcal{G}}}(\psi)\tilde{{\ensuremath{\mathcal{G}}}}(\psi)].\end{aligned}$$ At present, the only efficient partial characterizations are randomized benchmarking (RB) [@Emerson2005; @Levi2007; @Knill2008; @Dankert2009; @Magesan2011] and variants thereof [@Emerson2007; @Magesan2012; @Wallman2015; @Wallman2015b; @Wallman2016; @Carignan-Dugas2015; @Cross2016]. The RB protocol of Ref. [@Magesan2011] provides an estimate of the average fidelity over a group of operations from how the fidelity of sequences of operations decays as more operations are applied. RB protocols scale efficiently with the number of qubits and are robust to state-preparation and measurement (SPAM) errors. Consequently, RB protocols have become an important baseline for the validation and verification of quantum operations [@Corcoles2013; @Barends2014] and have recently been used to efficiently optimize over experimental control procedures [@Kelly2014]. However, rigorous analyses of RB protocols make several key assumptions, namely, that the noise is Markovian and time- and gate-independent, so that RB decay curves provide somewhat heuristic estimates of noise parameters (as do all noise characterization protocols). Numerical simulations of RB experiments have shown that the standard single-exponential decay curve holds (approximately) for a variety of physically relevant noise models with gate-dependent and non-Markovian noise, although the ‘true’ and estimated fidelity often differ by a factor of approximately two [@Epstein2014; @Ball2016]. However, it has recently been noted that gate-dependent noise fluctuations can dominate the gate-independent decay curve, resulting in a different decay parameter than expected when compared to a standard definition of the average noise [@Proctor]. In this paper, we prove that, under the assumption of Markovian noise, the full effect of gate-dependent fluctuations can be described by a single perturbation term that decays exponentially with the sequence length $m$. We also prove that the RB decay parameter is the fidelity of a suitably defined average noise process ${\ensuremath{\mathcal{E}}}$. Informally, our main result (\[thm:GD\_exact\]) is that for gate-dependent, but trace-perserving and Markovian noise, the average survival probability over all RB sequences of length $m$ is $$\begin{aligned} Ap({\ensuremath{\mathcal{E}}})^m + B + \epsilon_m\end{aligned}$$ for some constants $A$ and $B$, where $$\begin{aligned} \label{eq:prelation} p({\ensuremath{\mathcal{E}}}) &= \frac{d f({\ensuremath{\mathcal{E}}},{\ensuremath{\mathcal{I}}})-1}{d-1} \notag\\ {\rvert}\epsilon_m{\lvert}&\leq \delta_1 \delta_2^m,\end{aligned}$$ $\delta_1,\delta_2$ quantify the gate-dependence of the noise and $\delta_2$ is small for good implementations of gates by \[thm:smalldelta\]. We prove all theorems without restricting to trace-preserving noise, as we anticipate that near-term implementations of RB on encoded qubits [@Combes2017] will involve significant loss from the encoded space due to imperfect error-correction procedures. We also demonstrate that the alternative analyses of Ref. [@Magesan2011; @Magesan2012a; @Chasseur2015; @Proctor] do not rigorously justify fitting RB decays to a single exponential in \[sec:RBalt\]. This paper is structured as follows. We introduce notation in \[sec:notation\] and review the RB protocol of Ref. [@Magesan2011] in \[sec:RBprotocol\]. We then discuss an ambiguity in the decomposition of noise in \[sec:GD\_noise\] and show how it can be exploited to cancel out the majority of gate-dependent fluctuations in \[sec:GD\_RB\]. We illustrate the accuracy of our new analysis numerically in \[sec:numerics\] and show why alternative analyses of RB for gate-dependent noise do not justify fitting RB experiments to single exponential decays in \[sec:RBalt\]. We conclude by discussing some implications of our results for implementing and interpreting RB. Notation {#sec:notation} ======== We use the following notation throughout this paper. - All operators except density operators are denoted by Roman font (e.g., $A$). The normalized identity matrix is ${\ensuremath{\hat{I}}}_d = I_d/\sqrt{d}$. - Ideal channels are denoted by calligraphic font (e.g., ${\ensuremath{\mathcal{C}}}$), where the ideal channel ${\ensuremath{\mathcal{U}}}$ corresponding to a unitary operator $U$ is ${\ensuremath{\mathcal{U}}}(\rho) = U\rho U{\ensuremath{^{\dagger}}}$. - The unital component of a trace-preserving map is the linear map defined by ${\ensuremath{\mathcal{G}}}_{\mathrm{u}}(A) = {\ensuremath{\mathcal{G}}}(A-\operatorname{Tr}(A)/d)$. - A noisy implementation of an ideal channel is denoted with an overset $\tilde{}$ (e.g., $\tilde{{\ensuremath{\mathcal{C}}}}$ denotes the noisy implementation of ${\ensuremath{\mathcal{C}}}$). - Channel composition is denoted by (noncommutative) multiplication (i.e., ${\ensuremath{\mathcal{C}}}{\ensuremath{\mathcal{B}}}(A) = {\ensuremath{\mathcal{C}}}[{\ensuremath{\mathcal{B}}}(A)]$). - Noncommutative products of subscripted objects are denoted by the shorthand $$\begin{aligned} x_{b:a} &= \begin{cases} x_b x_{b-1}\ldots x_{a+1} x_a & b\geq a, \\ 1 & \mbox{otherwise.} \end{cases}\end{aligned}$$ - Groups and sets will be denoted by blackboard font, e.g., ${\ensuremath{\mathbb{G}}}$. - The uniform average over a set ${\ensuremath{\mathbb{X}}}$ is denoted by ${\ensuremath{\mathbb{E}}}_{x\in{\ensuremath{\mathbb{X}}}} [f(x)] =|X|^{-1}\sum_{x\in {\ensuremath{\mathbb{X}}}}f(x)$. - The trace and operator norms of a matrix are ${\lVert}M {\rVert}_1 = \operatorname{Tr}\sqrt{M{\ensuremath{^{\dagger}}}M}$ and ${\lVert}M {\rVert}= \max_v {\lVert}M v {\rVert}/{\lVert}v{\rVert}$ respectively. - The induced trace norm of a linear map ${\ensuremath{\mathcal{L}}}$ is $$\begin{aligned} {\lVert}{\ensuremath{\mathcal{L}}}{\rVert}_{1\to 1} = \sup_{M:{\lVert}M {\rVert}_1=1} {\lVert}{\ensuremath{\mathcal{L}}}(M) {\rVert}_1,\end{aligned}$$ where the maximization is over positive-semidefinite matrices. We use the following matrix representation of quantum channels (variously known as the Liouville, natural and Pauli transfer matrix representation) wherever a concrete representation is required. Let $\{e_1,\ldots,e_n\}$ be the canonical orthonormal unit basis of ${\ensuremath{\mathbb{C}}}^n$ and $\{B_1,\ldots,B_{d^2}\}$ be a trace-orthonormal basis of ${\ensuremath{\mathbb{C}}}^{d\times d}$, that is, $\operatorname{Tr}(B_j{\ensuremath{^{\dagger}}}B_k) = \delta_{j,k}$. Defining the linear map $|*{\rangle\!\rangle}:{\ensuremath{\mathbb{C}}}^{d\times d}\to{\ensuremath{\mathbb{C}}}^{d^2}$ by $$\begin{aligned} |A{\rangle\!\rangle}= \sum_j \operatorname{Tr}(B_j{\ensuremath{^{\dagger}}}A)e_j \end{aligned}$$ and the adjoint map ${\langle\!\langle}A| = |A{\rangle\!\rangle}{\ensuremath{^{\dagger}}}$, we have ${\langle\!\langle}A|B{\rangle\!\rangle}= \operatorname{Tr}(A{\ensuremath{^{\dagger}}}B)$. A channel ${\ensuremath{\mathcal{C}}}$ maps $\rho$ to ${\ensuremath{\mathcal{C}}}(\rho)$, which can be represented by the matrix $$\begin{aligned} \label{eq:matrix_rep} {\ensuremath{\mathcal{C}}}= \sum_j |{\ensuremath{\mathcal{C}}}(B_j){\rangle\!\rangle}\!{\langle\!\langle}B_j|, \end{aligned}$$ where we abuse notation slightly by using the same notation for the abstract channel and its matrix representation. For the remainder of the paper, we assume the operator basis is Hermitian and that $B_1 = \hat{I}_d$, so that the matrix representation of all Hermiticity-preserving maps, including all completely positive and trace-preserving (CPTP) maps, are real-valued. Randomized benchmarking protocol {#sec:RBprotocol} ================================ The RB protocol for a group of operations ${\ensuremath{\mathbb{G}}}$ that is also a unitary two-design [@Dankert2009] (e.g., the $n$-qubit Clifford group) is as follows. 1. Choose a positive integer $m$. 2. Choose a sequence $\vec{G} = (G_1,\ldots,G_m)\in{\ensuremath{\mathbb{G}}}^m$ of $m$ elements of ${\ensuremath{\mathbb{G}}}$ uniformly at random and set $G_{m+1} = (G_{m:1}){\ensuremath{^{\dagger}}}$. 3. Estimate the expectation value (also known as the survival probability) $Q_{\vec{G}} = \operatorname{Tr}[Q\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:1}(\rho)]$ of an observable $Q$ after preparing a $d$-level system in a state $\rho$ and applying the sequence of $m+1$ operations $G_1, G_2,\ldots,G_{m+1}$. 4. Repeat steps 2–3 to estimate the average over random sequences ${\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(Q_{\vec{G}})$ to a desired precision. 5. Repeat steps 1–4 for different values of $m$ and fit to the decay model $$\begin{aligned} \label{eq:fidelity_decay} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(Q_{\vec{G}}) = Ap^m + B\end{aligned}$$ to estimate $p$ [@Magesan2011]. Despite the ubiquitous usage of RB, best practice choices for the sequence lengths and the number of sequences at each length are still unknown, though see Refs. [@Epstein2014; @Wallman2014; @Granade2014] for some guidance. A unitary two-design is any set of unitary channels ${\ensuremath{\mathbb{G}}}$ such that uniformly sampling ${\ensuremath{\mathbb{G}}}$ reproduces the first and second moments of the Haar measure. The canonical example of a unitary two-design is the $n$-qubit Clifford group. The unitary two-design condition can also be stated in terms of how a general channel is ‘twirled’ by the group, where the twirl of a channel ${\ensuremath{\mathcal{C}}}$ over a group ${\ensuremath{\mathbb{G}}}$ is ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{C}}}{\ensuremath{\mathcal{G}}})$. A unitary two-design is any group ${\ensuremath{\mathbb{G}}}$ such that any CPTP channel is ‘twirled’ into a depolarizing channel [@Dankert2009; @Nielsen2002] $$\begin{aligned} {\ensuremath{\mathcal{D}}}_p(\rho) = p\rho + \frac{1-p}{d}I_d.\end{aligned}$$ Unitary two-designs also twirl a general linear map ${\ensuremath{\mathcal{C}}}$ into a depolarizing map composed with a channel that uniformly decreases the trace, that is, into a channel of the form $$\begin{aligned} {\ensuremath{\mathcal{D}}}_{p,t}(\rho) = \frac{t}{d}I_d + p(\rho-\frac{1}{d}I_d),\end{aligned}$$ where the values of $p$ and $t$ can be computed from the following lemma [@Wallman2014]. Note that $p$ and $t$ are functions from linear maps to the real numbers, however, we will often omit the argument when it is clear from the context. \[lem:twirl\] For any unitary two-design ${\ensuremath{\mathbb{G}}}$ and channel ${\ensuremath{\mathcal{C}}}$, $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{C}}}{\ensuremath{\mathcal{G}}}) = {\ensuremath{\mathcal{D}}}_{p,t},\end{aligned}$$ where $$\begin{aligned} \label{eq:decay_parameter} t({\ensuremath{\mathcal{C}}}) &= \frac{\operatorname{Tr}[{\ensuremath{\mathcal{C}}}(I_d)]}{d} \notag\\ p({\ensuremath{\mathcal{C}}}) &= \frac{\operatorname{Tr}({\ensuremath{\mathcal{C}}}) - t}{d^2-1} = \frac{df({\ensuremath{\mathcal{C}}}) - t}{d-1}.\end{aligned}$$ In \[thm:GD\_exact\], we prove that \[eq:fidelity\_decay\] holds, up to a small and exponentially decaying perturbation, for gate-dependent noise that is time-independent and Markovian, generalizing the original analysis to more realistic noise. Representing gate-dependent noise {#sec:GD_noise} ================================= Let $\tilde{{\ensuremath{\mathcal{G}}}}$ be an experimental implementation of an ideal unitary channel ${\ensuremath{\mathcal{G}}}(\rho) = G\rho G{\ensuremath{^{\dagger}}}$ with Markovian noise. It is convenient to decompose $\tilde{{\ensuremath{\mathcal{G}}}}$ into an ideal gate and a fictitious noise process, for example, $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{E}}}$. However, we can also write $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{L}}}_G {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}_G$ for any suitable ${\ensuremath{\mathcal{L}}}_G$ and ${\ensuremath{\mathcal{R}}}_G$. We now identify an approximate decomposition that will greatly simplify the analysis of RB with gate-dependent noise. \[thm:avnoise\] Let ${\ensuremath{\mathbb{G}}}$ be a unitary two-design, $\{\tilde{{\ensuremath{\mathcal{G}}}}:G\in{\ensuremath{\mathbb{G}}}\}$ be a corresponding set of Hermiticity-preserving maps, and $p$ and $t$ be the largest eigenvalues of ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}_{\mathrm{u}} \otimes \tilde{{\ensuremath{\mathcal{G}}}})$ and ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}})$ in absolute value respectively, where ${\ensuremath{\mathcal{G}}}_{\mathrm{u}}(A) = {\ensuremath{\mathcal{G}}}(A-\operatorname{Tr}(A)/d)$. There exist linear maps ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}$ such that \[eq:conditions\] $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}(\tilde{{\ensuremath{\mathcal{G}}}}{\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}) &= {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{D}}}_{p,t} \label{eq:lcona}\\ {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}\tilde{{\ensuremath{\mathcal{G}}}}) &= {\ensuremath{\mathcal{D}}}_{p,t}{\ensuremath{\mathcal{R}}}\label{eq:rcona}\\ {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}) &= {\ensuremath{\mathcal{D}}}_{p,t} \label{eq:LRscale}.\end{aligned}$$ When the noise is independent of the gate, that is, $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}$ for all $G\in{\ensuremath{\mathbb{G}}}$ for some ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}$, then ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}$ are solutions to \[eq:conditions\], as can be seen by substituting $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}$ into \[eq:conditions\] and using \[lem:twirl\]. For gate-dependent noise such that ${\ensuremath{\mathcal{R}}}$ is invertible, we can set $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{L}}}_G{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}$, so that the noise between two ideal gates ${\ensuremath{\mathcal{G}}}$ and ${\ensuremath{\mathcal{H}}}$ is ${\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}_G$. Substituting $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{L}}}_G{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}$ into \[eq:rcona\] and multiplying by ${\ensuremath{\mathcal{R}}}^{-1}$ gives $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}_G {\ensuremath{\mathcal{G}}}) = {\ensuremath{\mathcal{D}}}_{p,t}.\end{aligned}$$ As the Haar measure is unitarily invariant and $p$ and $t$ are convex functions, $$\begin{aligned} \label{eq:avGDnoise} p = p({\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} [{\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}_G{\ensuremath{\mathcal{G}}}]) = {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (p[{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}_G]) \notag\\ t = t({\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} [{\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}_G{\ensuremath{\mathcal{G}}}]) = {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (t[{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}_G])\end{aligned}$$ give the fidelity and loss of trace of the average noise between ideal gates to the identity via \[lem:twirl\]. If ${\ensuremath{\mathcal{R}}}$ is not invertible, we can introduce an arbitrarily small perturbation into the $\tilde{{\ensuremath{\mathcal{G}}}}$ to make ${\ensuremath{\mathcal{R}}}$ invertible and so \[eq:avGDnoise\] will hold to an arbitrary precision. When $\tilde{{\ensuremath{\mathcal{G}}}}= {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t}$, ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}_{\mathrm{u}} \otimes \tilde{{\ensuremath{\mathcal{G}}}})$ and ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}})$ each have only one nonzero eigenvalue, namely, $p$ and $t$ respectively [@Wallman2015; @Wallman2015b]. By the Bauer-Fike theorem [@Bauer1960], all the eigenvalues $\eta$ of a matrix $M+\delta M$ satisfy $$\begin{aligned} \min_{\lambda\in\operatorname{spec}(M)} {\lvert}\eta -\lambda{\rvert}\leq {\lVert}\delta M{\rVert}, \end{aligned}$$ where $\operatorname{spec}(M)$ is the spectrum of $M$. As the operator norm is submultiplicative, obeys the triangle inequality, and ${\lVert}A\otimes B{\rVert}= {\lVert}A{\rVert}{\lVert}B{\rVert}$, the largest eigenvalues of ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}_{\mathrm{u}} \otimes \tilde{{\ensuremath{\mathcal{G}}}})$ and ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}})$ will differ from $p'$ and $t'$ by at most ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} {\lVert}\tilde{{\ensuremath{\mathcal{G}}}} - {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p',t'}{\rVert}$ for any scalars $p'$ and $t'$. Therefore the largest eigenvalues in absolute value will generally be unique and hence must be real as they are the eigenvalues of real matrices. are essentially a pair of eigenvalue equations, except that ${\ensuremath{\mathbb{G}}}$ acts irreducibly on $|{\ensuremath{\hat{I}}}_d{\rangle\!\rangle}$ and its orthogonal complement. To utilize this structure, let ${\ensuremath{\mathcal{L}}}= |L{\rangle\!\rangle}\!{\langle\!\langle}{\ensuremath{\hat{I}}}_d| + {\ensuremath{\mathcal{L}}}'$ and ${\ensuremath{\mathcal{R}}}= |{\ensuremath{\hat{I}}}_d{\rangle\!\rangle}\!{\langle\!\langle}R| + {\ensuremath{\mathcal{R}}}'$ with ${\ensuremath{\mathcal{L}}}'|{\ensuremath{\hat{I}}}_d{\rangle\!\rangle}=0$ and ${\langle\!\langle}{\ensuremath{\hat{I}}}_d|{\ensuremath{\mathcal{R}}}' = 0$ without loss of generality. With these definitions and as ${\ensuremath{\mathcal{G}}}|{\ensuremath{\hat{I}}}_d{\rangle\!\rangle}= |{\ensuremath{\hat{I}}}_d{\rangle\!\rangle}$ for any unitary (or unital) channel, we can separate each equation in \[eq:lcona,eq:rcona\] into two independent equations \[eq:mncon\] $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}})|L{\rangle\!\rangle}&= t|L{\rangle\!\rangle}\label{eq:m2con}\\ {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}} {\ensuremath{\mathcal{L}}}' {\ensuremath{\mathcal{G}}}_{\mathrm{u}}{\ensuremath{^{\dagger}}}) &= p{\ensuremath{\mathcal{L}}}' \label{eq:n2con}\\ {\langle\!\langle}R|{\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}}) &= t{\langle\!\langle}R| \label{eq:m1con}\\ {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}_{\mathrm{u}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}' \tilde{{\ensuremath{\mathcal{G}}}}) &= p{\ensuremath{\mathcal{R}}}' \label{eq:n1con},\end{aligned}$$ where restricting to ${\ensuremath{\mathcal{G}}}_{\mathrm{u}}$ enforces the orthogonality constraints on ${\ensuremath{\mathcal{L}}}'|{\ensuremath{\hat{I}}}_d{\rangle\!\rangle}=0$ and ${\langle\!\langle}{\ensuremath{\hat{I}}}_d|{\ensuremath{\mathcal{R}}}'=0$. are left- and right-eigenvector problems of a real matrix and so $t$ can be set to any eigenvalue of ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}})$ and $L$ and $R$ to the corresponding non-trivial eigenvectors (recall that the set of left- and right-eigenvalues of a real matrix are identical). We choose the largest eigenvalue as it will result in the smallest $\Delta_G = \tilde{{\ensuremath{\mathcal{G}}}} - {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}$. To turn \[eq:n1con,eq:n2con\] into eigenvalue equations, we can use the vectorization map that maps a matrix to a vector by stacking the columns vertically, that is, $$\begin{aligned} \operatorname{vec}(\sum_j v_j e_j{\ensuremath{^{\dagger}}}) = \sum_j e_j\otimes v_j.\end{aligned}$$ The vectorization map satisfies the identity $$\begin{aligned} \label{eq:vec_product} \operatorname{vec}(ABC) = (C^T\otimes A)\operatorname{vec}(B),\end{aligned}$$ Applying \[eq:vec\_product\] to \[eq:n1con,eq:n2con\] and using ${\ensuremath{\mathcal{G}}}^T = {\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}$ (as the matrix basis is Hermitian) gives $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}_{\mathrm{u}} \otimes \tilde{{\ensuremath{\mathcal{G}}}})\operatorname{vec}({\ensuremath{\mathcal{L}}}') &= p \operatorname{vec}({\ensuremath{\mathcal{L}}}') \notag\\ {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}}\otimes {\ensuremath{\mathcal{G}}}_{\mathrm{u}})^T\operatorname{vec}({\ensuremath{\mathcal{R}}}') &= p \operatorname{vec}({\ensuremath{\mathcal{R}}}'),\label{eq:primeconditions}\end{aligned}$$ and so $p$ can be set to any eigenvalue of ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}_{\mathrm{u}} \otimes \tilde{{\ensuremath{\mathcal{G}}}})$ and ${\ensuremath{\mathcal{L}}}'$ and ${\ensuremath{\mathcal{R}}}'$ to the corresponding non-trivial eigenvectors (or equivalently, to any eigenvalue of ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}}\otimes {\ensuremath{\mathcal{G}}}_{\mathrm{u}})$ as the two are related by a unitary transformation and so have the same spectrum). The maps ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}$ identified are solutions to eigenvector equations and so are only determined up to a normalization constant. As ${\ensuremath{\mathcal{D}}}_{q,u}$ commutes with ${\ensuremath{\mathcal{G}}}$ for all scalars $q$ and $v$ and all ${\ensuremath{\mathcal{G}}}\in{\ensuremath{\mathbb{G}}}$, we can set ${\ensuremath{\mathcal{L}}}\to {\ensuremath{\mathcal{L}}}D_{q,u}$ and ${\ensuremath{\mathcal{R}}}\to {\ensuremath{\mathcal{D}}}_{r,v}{\ensuremath{\mathcal{R}}}$ to satisfy \[eq:LRscale\]. While applying a gauge transformation $\tilde{{\ensuremath{\mathcal{G}}}}\to{\ensuremath{\mathcal{S}}}^{-1}\tilde{{\ensuremath{\mathcal{G}}}}{\ensuremath{\mathcal{S}}}$ changes ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}$ to ${\ensuremath{\mathcal{L}}}\to {\ensuremath{\mathcal{S}}}{\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}\to{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{S}}}^{-1}$ respectively, the noise between ideal gates, namely, ${\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}$, is gauge invariant and so the present analysis is gauge-invariant unlike the original analysis of RB as shown in Ref. [@Proctor]. We now prove that the solutions ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}$ to \[eq:conditions\] can also be chosen to ensure that $\Delta_G = \tilde{{\ensuremath{\mathcal{G}}}} - {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}$ is small when $\tilde{{\ensuremath{\mathcal{G}}}}\approx{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t}$. To prove this, it is convenient to transform to the gauge $\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}}$ such that ${\ensuremath{\mathcal{L}}}={\ensuremath{\mathcal{I}}}$ (i.e., set ${\ensuremath{\mathcal{S}}}= {\ensuremath{\mathcal{L}}}^{-1}$). In this gauge, \[eq:LRscale\] becomes $$\begin{aligned} \label{eq:Rscale} {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}) = {\ensuremath{\mathcal{D}}}_{p,t}.\end{aligned}$$ In particular, $\Delta_G$ is on the order of ${\lVert}\tilde{{\ensuremath{\mathcal{G}}}}-{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t}{\rVert}\in O(\sqrt{1-p})$. \[thm:smalldelta\] Let ${\ensuremath{\mathbb{G}}}$ be a unitary two-design, $\{\tilde{{\ensuremath{\mathcal{G}}}}:G\in{\ensuremath{\mathbb{G}}}\}$ be a corresponding set of Hermiticity-preserving maps, and $p$ and $t$ be the largest eigenvalues of ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}_{\mathrm{u}} \otimes \tilde{{\ensuremath{\mathcal{G}}}})$ and ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}})$ in absolute value respectively, where ${\ensuremath{\mathcal{G}}}_{\mathrm{u}}(A) = {\ensuremath{\mathcal{G}}}(A-\operatorname{Tr}(A)/d)$. There exists a linear map ${\ensuremath{\mathcal{R}}}$ satisfying \[eq:conditions\] with ${\ensuremath{\mathcal{L}}}={\ensuremath{\mathcal{I}}}$ such that $$\begin{aligned} {\lVert}\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}} - {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}{\rVert}\leq {\lVert}\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}} - {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t}{\rVert}+ \frac{{\lVert}{\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}}) - {\ensuremath{\mathcal{D}}}_{p,t}{\rVert}}{1 - {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\lVert}\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}} - {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t}{\rVert}){\lVert}{\ensuremath{\mathcal{D}}}_{1/p,1/t}{\rVert}}.\end{aligned}$$ By the triangle inequality, submultiplicativity, and unitary invariance of the operator norm, $$\begin{aligned} \label{eq:bound_terms} {\lVert}\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}} - {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}{\rVert}&\leq {\lVert}\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}} - {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t}{\rVert}+ {\lVert}{\ensuremath{\mathcal{G}}}({\ensuremath{\mathcal{D}}}_{p,t}-{\ensuremath{\mathcal{R}}}) {\rVert}\notag\\ &\leq {\lVert}\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}} - {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t}{\rVert}+ {\lVert}{\ensuremath{\mathcal{D}}}_{p,t}-{\ensuremath{\mathcal{R}}}{\rVert}\end{aligned}$$ for all $G\in{\ensuremath{\mathbb{G}}}$. holds for any ${\ensuremath{\mathcal{R}}}$. Now let ${\ensuremath{\mathcal{R}}}$ be a solution to \[eq:conditions\] for $\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}}$ and expand ${\ensuremath{\mathcal{R}}}= {\ensuremath{\mathcal{D}}}_{p,t} + {\ensuremath{\mathcal{R}}}_1$ and $\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}} = {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t} + \tilde{{\ensuremath{\mathcal{G}}}}_1$. Substituting these expansions into \[eq:rcona\] and using \[eq:Rscale\] gives $$\begin{aligned} \label{eq:pert_con} {\ensuremath{\mathcal{D}}}_{p,t}^2 + {\ensuremath{\mathcal{D}}}_{p,t}{\ensuremath{\mathcal{R}}}_1 &= {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t}) + {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}\tilde{{\ensuremath{\mathcal{G}}}}_1) \notag\\ &= {\ensuremath{\mathcal{D}}}_{p,t}^2 + {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}\tilde{{\ensuremath{\mathcal{G}}}}_1)\end{aligned}$$ Canceling the common term in \[eq:pert\_con\], multiplying both sides by ${\ensuremath{\mathcal{D}}}_{1/p,1/t}$ from the left, taking the operator norm and using the triangle inequality and submultiplicativity of the operator norm gives $$\begin{aligned} {\lVert}{\ensuremath{\mathcal{R}}}_1 {\rVert}&\leq {\lVert}{\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}\tilde{{\ensuremath{\mathcal{G}}}}_1){\rVert}+ {\lVert}{\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{D}}}_{1/p,1/t} {\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}_1\tilde{{\ensuremath{\mathcal{G}}}}_1) {\rVert}\notag\\ &\leq {\lVert}{\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}\tilde{{\ensuremath{\mathcal{G}}}}_1){\rVert}+ {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\lVert}\tilde{{\ensuremath{\mathcal{G}}}}_1{\rVert}{\lVert}{\ensuremath{\mathcal{R}}}_1{\rVert}{\lVert}{\ensuremath{\mathcal{D}}}_{1/p,1/t}{\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\rVert}).\end{aligned}$$ Rearranging and using the unitary invariance of the operator norm gives $$\begin{aligned} {\lVert}{\ensuremath{\mathcal{R}}}_1{\rVert}&= {\lVert}{\ensuremath{\mathcal{D}}}_{p,t}-{\ensuremath{\mathcal{R}}}{\rVert}\notag\\ &\leq \frac{{\lVert}{\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}\tilde{{\ensuremath{\mathcal{G}}}}_1){\rVert}}{1 - {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\lVert}\tilde{{\ensuremath{\mathcal{G}}}}_1{\rVert}) {\lVert}{\ensuremath{\mathcal{D}}}_{1/p,1/t}{\rVert}} \notag\\ &=\frac{{\lVert}{\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}}) - {\ensuremath{\mathcal{D}}}_{p,t}{\rVert}}{1 - {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\lVert}\tilde{{\ensuremath{\mathcal{G}}}}^{{\ensuremath{\mathcal{I}}}} - {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_{p,t}{\rVert}) {\lVert}{\ensuremath{\mathcal{D}}}_{1/p,1/t}{\rVert}}.\end{aligned}$$ Analyzing RB with arbitrarily gate-dependent noise {#sec:GD_RB} ================================================== Experimental implementations of RB almost always produce relatively clean exponential decays. We now explain this empirical success for general gate-dependent noise by proving that the average survival probability over RB sequences of length $m$ is equal to the model of Ref. [@Magesan2012a] for gate-independent noise with a single perturbation term for a suitably defined average noise model. Moreover, the decay constants correspond to the average loss of trace and fidelity of the noise between ideal gates via \[eq:avGDnoise\]. The perturbation term will generally be negligible for noise that is close to the identity by \[thm:smalldelta\] and for $m\geq 3$ (for numerical values for a single qubit, see \[fig:GDnumerics\]). Even under this condition, more significant (and potentially oscillating) perturbations are possible for $m\approx 1$, potentially explaining experimentally observed fluctuations as seen in, e.g., [@Corcoles2013 Fig. 3b]. While we have not investigated how the perturbation terms scale with dimension, note that we only need $\delta_2\lessapprox 1/2$ in order for the decay to be essentially a single exponential for $m\geq 10$, etc, as $\delta_2$ is exponentially suppressed. \[thm:GD\_exact\] For any unitary two-design ${\ensuremath{\mathbb{G}}}$ and set of Hermiticitiy preserving linear maps $\{\tilde{{\ensuremath{\mathcal{G}}}}:G\in{\ensuremath{\mathbb{G}}}\}$, the average survival probability over all randomized benchmarking sequences of length $m$ is $$\begin{aligned} \label{eq:exact_decay} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m} (Q_{\vec{G}}) &= A p^m + B t^m + \epsilon_m\end{aligned}$$ for some constants $A$ and $B$, where $p$ and $t$ give the fidelity and loss of trace of the average noise between ideal gates via \[eq:avGDnoise\]. Moreover, the perturbation term $\epsilon_m$ satisfies $$\begin{aligned} {\lvert}\epsilon_m {\rvert}\leq \delta_1 \delta_2^m\end{aligned}$$ for some $\delta_1$ and $\delta_2$ that quantify the amount of gate dependence. reduces to the standard single-exponential model plus an exponentially decreasing perturbation when $\tilde{{\ensuremath{\mathcal{G}}}}$ is a CPTP map for all $G$ as then ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}}\otimes {\ensuremath{\mathcal{G}}})$ and ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}\otimes \tilde{{\ensuremath{\mathcal{G}}}})$ are both CPTP maps and so all eigenvalues are in the unit disc and the eigenvalue with eigenvector close to $|{\ensuremath{\hat{I}}}_d{\rangle\!\rangle}$ is 1 [@Perez-Garcia2006]. The constants $A$ and $B$ and the perturbation term are $$\begin{aligned} A &= {\langle\!\langle}Q|{\ensuremath{\mathcal{L}}}|{\ensuremath{\mathcal{R}}}(\rho)-I_d/d{\rangle\!\rangle}\notag\\ B &= {\langle\!\langle}Q|{\ensuremath{\mathcal{L}}}|I_d/d{\rangle\!\rangle}\notag\\ \epsilon_m &= {\langle\!\langle}Q|{\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\Delta_{m+1:1})|\rho{\rangle\!\rangle}.\end{aligned}$$ with $\Delta_j = \Delta_{G_j}= \tilde{{\ensuremath{\mathcal{G}}}}_j - {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}_j {\ensuremath{\mathcal{R}}}$ and where ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}$ are solutions to \[eq:conditions\]. The quantities bounding the perturbation term are $$\begin{aligned} \label{eq:deltas} \delta_1 &= \max_G {\lVert}\Delta_G {\rVert}_{1\to 1}{\lVert}\rho{\rVert}_1 {\lVert}Q{\rVert}_1\notag\\ \delta_2 &= {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\lVert}\Delta_G{\rVert}).\end{aligned}$$ which are in turn bounded in \[thm:smalldelta\] in the gauge where ${\ensuremath{\mathcal{L}}}={\ensuremath{\mathcal{I}}}$. Note that ${\lVert}\rho{\rVert}_1,{\lVert}Q{\rVert}_1\leq 1$ in a standard gauge, however, the gauge transformation required to set ${\ensuremath{\mathcal{L}}}= {\ensuremath{\mathcal{I}}}$ may change these values and introduce negative eigenvalues to $\rho$ so that the maximization for the induced norm in $\delta_1$ is not restricted to positive semi-definite inputs. The average map over all randomized benchmarking sequences $\vec{G}$ of length $m$ is $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:1}).\end{aligned}$$ Let ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}$ satisfy \[eq:mapcon\] $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} (\tilde{{\ensuremath{\mathcal{G}}}}{\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}) &= {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{D}}}_{p,t} \label{eq:lcon}\\ {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}\tilde{{\ensuremath{\mathcal{G}}}}) &= {\ensuremath{\mathcal{D}}}_{p,t}{\ensuremath{\mathcal{R}}}\label{eq:rcon}\\ {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}) &= {\ensuremath{\mathcal{D}}}_{p,t}\end{aligned}$$ and define $$\begin{aligned} \Delta_j = \Delta_{G_j} = \tilde{{\ensuremath{\mathcal{G}}}}_j - {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}_j {\ensuremath{\mathcal{R}}}.\end{aligned}$$ Then for any integer $1\leq j \leq m+1$ we can write $$\begin{aligned} \label{eq:splitting} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:j}\Delta_{j-1:1}) &= {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:j+1} {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}_j {\ensuremath{\mathcal{R}}}\Delta_{j-1:1}) + {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:j+1}\Delta_j\Delta_{j-1:1}) \notag\\ &= {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:j+1} {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}_j {\ensuremath{\mathcal{R}}}\Delta_{j-1:1}) + {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:j+1}\Delta_{j:1}),\end{aligned}$$ where $\Delta_{0:1} = I_d = \tilde{{\ensuremath{\mathcal{G}}}}_{m+1:m+2}$ by convention. Moreover, as $G_{m+1:1} = I_d$ and any set of $m$ elements of $\vec{G}$ are statistically independent and uniformly distributed, we can apply \[eq:lcon\] recursively and use the fact that ${\ensuremath{\mathcal{D}}}_{p,t}$ commutes with unital channels to obtain $$\begin{aligned} \label{eq:j1term} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:2} {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}_1 {\ensuremath{\mathcal{R}}}) &= {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:2} {\ensuremath{\mathcal{L}}}[{\ensuremath{\mathcal{G}}}_{m+1:2}]{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}) \notag\\ &= {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{D}}}_{p,t}^m {\ensuremath{\mathcal{R}}}\end{aligned}$$ for $j=1$. Similarly, we can use \[lem:twirl,eq:mapcon\] for $j\geq 2$ to obtain $$\begin{aligned} \label{eq:otherj} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}[\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:j+1} {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}_j {\ensuremath{\mathcal{R}}}\Delta_{j-1:1}] &= {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}[\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:j+1} {\ensuremath{\mathcal{L}}}({\ensuremath{\mathcal{G}}}_{m+1:j+1}){\ensuremath{^{\dagger}}}({\ensuremath{\mathcal{G}}}_{j-2:1}){\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{G}}}_{j-1}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}\Delta_{j-1}\Delta_{j-2:1}] \notag\\ &= {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{D}}}_{p,t}^{m-j}{\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}[({\ensuremath{\mathcal{G}}}_{j-2:1}){\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{G}}}_{j-1}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}(\tilde{{\ensuremath{\mathcal{G}}}}_{j-1}-{\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}_{j-1}{\ensuremath{\mathcal{R}}}) \Delta_{j-2:1}] \notag\\ &= {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{D}}}_{p,t}^{m-j} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}[({\ensuremath{\mathcal{G}}}_{j-2:1}){\ensuremath{^{\dagger}}}{\ensuremath{\mathbb{E}}}_{G_{j-1}}({\ensuremath{\mathcal{G}}}_{j-1}{\ensuremath{^{\dagger}}}{\ensuremath{\mathcal{R}}}\tilde{{\ensuremath{\mathcal{G}}}}_{j-1}- {\ensuremath{\mathcal{D}}}_{p,t}{\ensuremath{\mathcal{R}}}) \Delta_{j-2:1}] \notag\\ &= 0.\end{aligned}$$ Therefore we can apply \[eq:splitting,eq:j1term\] and then recursively apply \[eq:splitting,eq:otherj\] to obtain $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m} \tilde{{\ensuremath{\mathcal{G}}}}_{m+1:1} &= {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{D}}}_{p,t}^m {\ensuremath{\mathcal{R}}}+ {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}[\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:2}\Delta_{1}] \notag\\ &= {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{D}}}_{p,t}^m {\ensuremath{\mathcal{R}}}+ {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}[\Delta_{m+1:1}].\end{aligned}$$ The average survival probability over all random sequences is then $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m} Q_{\vec{G}} &= {\langle\!\langle}Q|{\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{D}}}_{p,t}^m {\ensuremath{\mathcal{R}}}|\rho{\rangle\!\rangle}+ \epsilon_m \notag\\ &= t^m {\langle\!\langle}Q|{\ensuremath{\mathcal{L}}}|I_d/d{\rangle\!\rangle}+ p^m {\langle\!\langle}Q|{\ensuremath{\mathcal{L}}}|{\ensuremath{\mathcal{R}}}(\rho)-I_d/d{\rangle\!\rangle}+ \epsilon_m\end{aligned}$$ where $$\begin{aligned} \epsilon_m &= {\langle\!\langle}Q|{\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\Delta_{m+1:1})|\rho{\rangle\!\rangle}.\end{aligned}$$ We now bound the perturbation term. By the triangle inequality and submultiplicativity, $$\begin{aligned} {\lVert}{\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m} (\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:1}) - {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{D}}}_{p,t}^m {\ensuremath{\mathcal{R}}}{\rVert}_{1\to1} &= {\lVert}{\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\Delta_{m+1:1}) {\rVert}_{1\to1} \notag\\ &\leq {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}{\lVert}\Delta_{m+1:1} {\rVert}_{1\to1} \notag\\ &\leq {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}{\lVert}\Delta_1{\rVert}_{1\to1} \prod_{j= m+1}^2 {\lVert}\Delta_j{\rVert}\notag\\ &\leq \max_G {\lVert}\Delta_G {\rVert}_{1\to 1} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}\prod_{j= m+1}^2 {\lVert}\Delta_j{\rVert}\notag\\ &\leq \max_G {\lVert}\Delta_G {\rVert}_{1\to 1} [{\ensuremath{\mathbb{E}}}_H {\lVert}\Delta_H{\rVert}]^m .\end{aligned}$$ Numerics {#sec:numerics} ======== ![ Left: upper- and lower- edges of 90% confidence intervals for the decay parameter from the numerical simulations of RB experiments described in \[sec:numerics\] fitted neglecting the perturbation term $\delta_1\delta_2^m$ (red intervals) and the corresponding values of $p$ from \[thm:GD\_exact\] (black dots). Right: numerical values of $\delta_1$ (circles) and $\delta_2$ (crosses) from \[thm:GD\_exact\]. For all simulated experiments, the perturbation term scales as $\delta_1 \delta_2^m \in O(r^{-m/2})$ and so is negligible for $m\geq 3$.[]{data-label="fig:GDnumerics"}](plot_estimates.pdf){width="\linewidth"} In \[fig:GDnumerics\] we illustrate the reliability of our analysis by comparing the estimates obtained by fitting the simulated RB experiments described below under independent random unitary noise on a set of generating gates to the decay parameters predicted by \[thm:GD\_exact\] and by plotting the size of the perturbation terms $\delta_1$ and $\delta_2$ calculated in the gauge where ${\ensuremath{\mathcal{L}}}= {\ensuremath{\mathcal{I}}}$. Mathematica code is available at <https://github.com/jjwallman/numerics>. For simplicity, we simulate RB experiments using the (projective) group $$\begin{aligned} {\ensuremath{\mathbb{G}}}= \{T^t P:t\in{\ensuremath{\mathbb{Z}}}_3,P\in\{I,X,Y,Z\}\}\end{aligned}$$ where $I,X,Y,Z$ are the standard single-qubit Pauli matrices and $$\begin{aligned} T = \frac{1}{\sqrt{2}}\begin{pmatrix} 1 & -i \\ 1 & i \end{pmatrix},\end{aligned}$$ which is a subgroup of the single-qubit Clifford group that is also a unitary two-design. For our simulations, the noisy implementation of $G = T^t P\in{\ensuremath{\mathbb{G}}}$ is $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{G}}}'$ where $$\begin{aligned} G' = U T^t V P\end{aligned}$$ where $U$ and $V$ are independently sampled from the set ${\ensuremath{\mathbb{U}}}_r\subset \mathrm{SU}(d)$ satisfying $1 - f({\ensuremath{\mathcal{U}}},{\ensuremath{\mathcal{I}}}) = r$, that is, the set of unitary channels of fixed infidelity $r$. We then calculate $p({\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}})$ from \[thm:avnoise\] and $\delta_1,\delta_2$ from \[eq:deltas\]. We set the state to $|0{\rangle}$ and the measurement to be a computational basis measurement and do not include SPAM errors for simplicity. The set ${\ensuremath{\mathbb{U}}}_r$ can be sampled by sampling $V\in\mathrm{SU}(d)$ uniformly from the Haar measure and setting $$\begin{aligned} U = V\exp(-i\theta Z) V{\ensuremath{^{\dagger}}}= \cos\theta I - i \sin\theta VZV{\ensuremath{^{\dagger}}},\end{aligned}$$ which gives the unique unitarily invariant distribution over ${\ensuremath{\mathbb{U}}}_r$ for some $r$ as a function of $\theta$. In order for the infidelity of the resulting gates to be $r$, we require [@Nielsen2002] $$\begin{aligned} r = \frac{4-{\lvert}\operatorname{Tr}U{\rvert}^2 }{6} = \frac{2 - \cos^2\theta}{3},\end{aligned}$$ or $\theta = \arcsin(\sqrt{3r/2})$. A single RB experiment for a fixed value of $m$ then consisted of the following. 1. Randomly choose $\vec{a}\in \{0,1,2\}^m$ and $\vec{b}\in\{1,2,3,4\}^m$. 2. Set $G_j = T^{a_j} P_{b_j} P_{b_{j-1}} T^{-a_{j-1}} = T^{a_j - a_{j-1}} P_{\tau_j}$ where $(P_0,P_1,P_2,P_3) = (I,X,Y,Z)$, $a_{m+1} = a_0$, $ b_{m+1} = b_0 = 1$, and $P_{\tau_j} = T^{a_{j-1}} P_{b_j} P_{b_{j-1}} T^{-a_{j-1}}$. 3. Calculate ${\lvert}{\langle}0{\rvert}G'_{m+1:1} {\lvert}0 {\rangle}{\rvert}^2$ We repeat the above steps for 100 random sequences for each value of $m\in \{4,8,16,...,2048\}$ and fit the averages for each sequence to the model ${\ensuremath{\mathbb{E}}}_{\vec{{\ensuremath{\mathcal{C}}}}} = 0.5p_{est}^m + 0.5$ using Mathematica’s NonlinearModelFit function weighted by the variance over the 100 sequences at each $m$, where we have set $A=B=0.5$ because of the choice of SPAM and because the noise is unital. Alternative analysis of RB with gate-dependent noise {#sec:RBalt} ==================================================== We now show how the analyses of gate-dependent noise of Refs. [@Magesan2011; @Magesan2012a; @Chasseur2015; @Proctor] are too loose to justify fitting RB decay curves to a single exponential in practical regimes. We will first show that the higher-order terms in the original analysis of RB give a systematic uncertainty of the infidelity that dominates the infidelity, even when the fidelity is dominated by stochastic errors. We will then discuss subsequent analysis and how it also results in large systematic errors. Initial RB Analysis ------------------- The initial method of analyzing RB under gate-dependent noise was to perform a perturbation expansion to perturb about a perfect twirl [@Magesan2011; @Magesan2012a] of the average noise, $$\begin{aligned} \label{eq:av_noise} {\ensuremath{\mathcal{E}}}= {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{G}}}{\ensuremath{^{\dagger}}}\tilde{{\ensuremath{\mathcal{G}}}}).\end{aligned}$$ Substituting $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{E}}}+ {\ensuremath{\mathcal{G}}}\Delta_G$ into the average map over all sequences, $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:1})\end{aligned}$$ and using the triangle inequality and submultiplicativity, the $k$th order terms in $\Delta_G$ contribute at most $$\begin{aligned} \label{eq:perturbation_bound} \binom{m+1}{k}\gamma^k = \binom{m+1}{k}{\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}} ({\lVert}\Delta_G {\rVert}_{1\to 1})\end{aligned}$$ to the average survival probability over all sequences of length $m$, ${\langle\!\langle}Q{\rvert}{\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(\tilde{{\ensuremath{\mathcal{G}}}}_{m+1:1}) {\lvert}\rho{\rangle\!\rangle}$. Ref. [@Magesan2012a] also obtained similar conditions for time-dependent noise and derived the first-order correction. As stated in Ref. [@Magesan2011], higher-order terms will be negligible provided $m\gamma \ll 1$, so that there is a tension between the amount of gate-dependent noise and the values of $m$ required to obtain a reasonable fit. However, there exist noise models such that $m\gamma$ is not negligible [@Proctor]. The analysis of Ref. [@Magesan2012a] also holds for the more general gate-dependent noise model ${\ensuremath{\mathcal{L}}}_G {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}_G$, where ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}_G$ are gate-independent and gate-dependent noise processes acting from the left and right (that is, after and before a ideal gate) respectively, so that some optimization of $\gamma$ is possible. However, we now show that the higher-order terms cannot be neglected when fitting data in experimental regimes. Suppose the average survival probability over all sequences of some fixed length $m$ is measured to be $f_m$. From the zeroth-order model with $p = 1-2r$, $$\begin{aligned} r = \frac{1}{2} - \frac{1}{2}\sqrt[m]{\frac{f_m - B}{A}},\end{aligned}$$ ignoring any finite measurement statistics. However, by \[eq:perturbation\_bound\], the first-order perturbations add some $\delta\in[-m\gamma,m\gamma]$ to $f_m$, so that the true value of $r$ is $$\begin{aligned} r = \frac{1}{2} - \frac{1}{2}\sqrt[m]{\frac{f_m -B - \delta}{A}} \approx \frac{1}{2} - \frac{1}{2}\sqrt[m]{\frac{f_m - B}{A}} + \frac{\delta}{2m (f_m - B)},\end{aligned}$$ where we have assumed $m\gg 1$ for the approximation, so that the first-order perturbation terms add a systematic uncertainty of $\gamma/2(f_m-B)$ to the estimate of $r$. Therefore the systematic ucertainty will dominate the estimate of the infidelity unless $\gamma \approx r({\ensuremath{\mathcal{E}}})$. To illustrate how impractical this requirement is, and thus the value of the tighter analysis in \[sec:GD\_RB\], consider the following hypothetical implementation of the single-qubit Clifford group (that is, the 24 elements of SU(2) that permute the single-qubit Paulis $X$, $Y$ and $Z$ up to an overall sign). Supoose that half the elements are implemented with only depolarizing noise, that is, $\tilde{{\ensuremath{\mathcal{G}}}} ={\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_\nu$, and the other half are implemented with an additional rotation around the $z$-axis of the Bloch sphere by some angle $\theta\in(0,\tfrac{\pi}{2})$, that is, $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{D}}}_\nu{\ensuremath{\mathcal{Z}}}_\theta$ where $$\begin{aligned} Z_\theta = \left(\begin{array}{cc} 1 & 0 \\ 0 & e^{i\theta} \end{array}\right).\end{aligned}$$ For this noise model, the average noise is ${\ensuremath{\mathcal{E}}}= {\ensuremath{\mathcal{D}}}_\nu(I_4 + {\ensuremath{\mathcal{Z}}}_\theta)/2$ by \[eq:av\_noise\] and so $$\begin{aligned} \gamma &= \frac{1}{2}{\lVert}{\ensuremath{\mathcal{D}}}_\nu - {\ensuremath{\mathcal{D}}}_\nu(I_4 + {\ensuremath{\mathcal{Z}}}_\theta)/2{\rVert}_{1\to 1} + \frac{1}{2}{\lVert}{\ensuremath{\mathcal{D}}}_\nu{\ensuremath{\mathcal{Z}}}_\theta - {\ensuremath{\mathcal{D}}}_\nu(I_4 + {\ensuremath{\mathcal{Z}}}_\theta)/2{\rVert}_{1\to 1} \notag\\ &= \frac{\nu}{2}{\lVert}{\ensuremath{\mathcal{Z}}}_\theta - I_4{\rVert}_{1\to 1} \notag\\ r({\ensuremath{\mathcal{E}}}) &= \frac{1}{6}\Bigl(3 - 2\nu - \nu\cos\theta\Bigr)\end{aligned}$$ using $1-r({\ensuremath{\mathcal{E}}}) = f({\ensuremath{\mathcal{E}}},{\ensuremath{\mathcal{I}}}) = 1/2 + \sum_{\sigma=X,Y,Z} \operatorname{Tr}[\sigma {\ensuremath{\mathcal{E}}}(\sigma)]/12$ [@Bowdrey2002]. The maximization in the induced trace norm will be achieved by any state in the $xy$ plane (e.g., $(|0{\rangle}+ |1{\rangle})/\sqrt{2}$), giving $$\begin{aligned} \gamma = \nu{\lvert}\sin(\theta/2){\rvert}.\end{aligned}$$ For the estimate based on the zeroth-order model to be valid to within a factor of two, we require $\gamma\leq r({\ensuremath{\mathcal{E}}})$, that is, ${\lvert}\theta {\rvert}\lesssim 1-\nu$. For example, the systematic uncertainty in the estimate of $r$ is $9r$ (i.e., under-reporting the error rate $r$ by an order of magnitude) when $\nu = 0.99$ and gate-dependent errors account for 10% of the total infidelity ($\theta = 0.09$ in the above model) or when $\nu=0.999$ and the gate-dependent errors account for just 1% of the total infidelity ($\theta = 0.009$ in the above model). One might expect that the above analysis is pessimistic and that the contributions from gate-dependent perturbations at different sequence lengths would average out. However, Ref. [@Proctor] presented an explicit model wherein the contributions from gate-dependent perturbations are systematic. Subsequent analyses of RB ------------------------- Ref. [@Proctor] demonstrated that the RB decay rate could differ from the average gate infidelity of the average noise defined in \[eq:av\_noise\] by orders of magnitude. Ref. [@Proctor] also presented a significantly improved analysis of the original RB protocol, which can be viewed as an application of the method of [@Chasseur2015] to the original RB protocol. However, this analys has three crucial drawbacks. First, no interpretation of the decay parameter was provided, leading to frequent discussions of “the RB number” $r_{RB}$ at conferences. In the original version, Ref. [@Proctor] states that “it is not clear that $r$ (a function of the RB decay parameter) should be called (the) “fidelity”, except in the special case of gate-independent error maps,” although this statement has since been revised based on the analysis in the present paper. Second, the analysis of Ref. [@Proctor] is vulnerable to the criticism it makes of Ref. [@Magesan2012a], namely, that the infidelity $r({\ensuremath{\mathcal{E}}})$ of a fixed decomposition of the noise (defined in \[eq:av\_noise\]) can differ from $r_{RB}$ by orders of magnitude. This is problematic for the analysis of Ref. [@Proctor] (and also for that of Ref. [@Chasseur2015]) because they bound a perturbation from an exponential of the form $(1-r_{RB})^m$ (ignoring small dimensional factors) by the diamond norm with respect to a fixed decomposition of the noise, giving a perturbation on the order of $\sqrt{r({\ensuremath{\mathcal{E}}})}\gg r({\ensuremath{\mathcal{E}}}) \gg r_{RB}$. While some gauge optimization of their bound is possible (especially using a gauge motivated by the present analysis), it is unclear whether such optimization can close the gap between $r({\ensuremath{\mathcal{E}}})$ and $r_{RB}$. Fortunately, the present paper obviates the potential problem by providing a much sharper bound. Finally, even if gauge optimization allows the perturbation in their method to be bounded by $\sqrt{r_{RB}}$, this still renders their analysis useless in two key regimes, namely, for “large” error rates or in the linear part of the decay. For “large” error rates where $r_{RB} \approx 0.01$, which accounts for most RB experiments to date, the best bound that can be put on the perturbation without knowing the full experimental noise process is approximately $\sqrt{r_{RB}} = 0.1$. With such a large perturbation, any sequence lengths that have an average survival probability over $0.9$ or under $0.6$ are effectively useless assuming SPAM values of $A=B=0.5$ (and the range of values that contain any information decreases with $A$). In the linear part of the decay, such as in the numerics of Ref. [@Proctor], the systematic uncertainty due to the perturbation dominates the decay. For concreteness, consider the linear regime for a single qubit, $$\begin{aligned} {\ensuremath{\mathbb{E}}}_{\vec{G}\in{\ensuremath{\mathbb{G}}}^m}(Q_{\vec{G}})\approx A + B -2A m [1-p({\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}})] .\end{aligned}$$ where $p({\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}})$ is the true RB decay parameter from \[thm:GD\_exact\]. If $p({\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}})$ is estimated via the slope between $m=m_1$ and $m=m_2$, then the unknown perturbation contributes a systematic uncertainty of approximately $\sqrt{r({\ensuremath{\mathcal{E}}})}/(m_1-m_2)$ to the estimate of $p$, so that essentially no information is gained unless $m_1 - m_2 \gg \sqrt{r({\ensuremath{\mathcal{E}}})}$. For the current record value of $1-p \approx 10^{-6}$, this translates to requiring $m_2 - m_1 > 1000$ assuming $1-p({\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}) \approx r({\ensuremath{\mathcal{E}}})$. However, the assumption that $1-p({\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{L}}}) \approx r({\ensuremath{\mathcal{E}}})$ is unjustified as shown by their analysis, so that the required spacing of sequence lengths in order to overcome the systematic uncertainty due to their bound is unknown, rendering their analysis impractical. Consequently, the analysis of Ref. [@Proctor] does not explain the empirical success of RB to date or justify fitting to a single exponential without further analysis. Conclusion ========== We have proven that randomized benchmarking experiments result in a single exponential with a negligible perturbation under gate-dependent noise. Moreover, we proved that the decay parameter is the fidelity of the average noise between idealized gates. More specifically, we can write $\tilde{{\ensuremath{\mathcal{G}}}} = {\ensuremath{\mathcal{L}}}{\ensuremath{\mathcal{G}}}{\ensuremath{\mathcal{R}}}_G$ where ${\ensuremath{\mathcal{L}}}$ is a solution to \[eq:conditions\]. Then for trace-preserving noise the RB decay parameter is ${\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}[p({\ensuremath{\mathcal{R}}}_G{\ensuremath{\mathcal{L}}})]$, that is, the average survival probability between ideal gates. As $\tilde{{\ensuremath{\mathcal{G}}}}\to{\ensuremath{\mathcal{S}}}^{-1}\tilde{{\ensuremath{\mathcal{G}}}}{\ensuremath{\mathcal{S}}}$ maps ${\ensuremath{\mathcal{L}}}\to {\ensuremath{\mathcal{S}}}{\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}\to{\ensuremath{\mathcal{R}}}{\ensuremath{\mathcal{S}}}^{-1}$ in \[thm:avnoise\], the decay parameter $p = {\ensuremath{\mathbb{E}}}_{G\in{\ensuremath{\mathbb{G}}}}({\ensuremath{\mathcal{R}}}_G{\ensuremath{\mathcal{L}}})$ is manifestly gauge invariant, answering the recent criticism of Ref. [@Proctor]. By rigorously reducing the potential impact of gate-dependent errors, the present work enables randomized benchmarking to more reliably diagnose the presence of non-Markovian noise. More precisely, there are two key assumptions that may not hold well enough for the standard exponential decay model to be valid, namely, Markovianity and time-independence. However, the only effect that time-dependent noise can have is to alter the rate at which the survival probability decays [@Emerson2005; @Wallman2014]. Consequently, any “revivals” in RB experiments, that is, statistically significant increases in the average survival probability as $m$ increases [@Epstein2014], can be attributed directly to the noise being non-Markovian. An intriguing observation from the present analysis is that gate-dependent fluctuations may produce a significant deviation from the exponential fit for short sequences. This deviation is possible because the first few values of $m$ are the most sensitive to gate-dependent fluctuations, which may be as large as $0.1$ for $m=1$ and $0.01$ for $m=2$. Consequently, these sequence lengths should be omitted when fitting to \[eq:fidelity\_decay\], but can also indicate the presence of gate-dependent noise. Alternative protocols based on randomized benchmarking have been developed that also assume gate-independent noise [@Emerson2007; @Magesan2012; @Wallman2015; @Wallman2015b; @Wallman2016; @Carignan-Dugas2015; @Cross2016]. We leave applying the current analysis to these protocols to future work. One main open problem of the current work is that the solutions ${\ensuremath{\mathcal{L}}}$ and ${\ensuremath{\mathcal{R}}}$ from \[thm:avnoise\] are not guaranteed to be CPTP maps, even under rescaling. The author acknowledges helpful discussions with Robin Blume-Kohout, Arnaud Carignan-Dugas, Joseph Emerson, Steve Flammia, Chris Granade, Richard Kueng, Robin Harper, Jonas Helsen, Thomas Monz, Jiaan Qi and Philipp Schindler. This research was supported by the U.S. Army Research Office through grant W911NF-14-1-0103. This research was undertaken thanks in part to funding from the Canada First Research Excellence Fund. [37]{} Isaac L. Chuang and Michael A. Nielsen, Prescription for experimental determination of the dynamics of a quantum black box, [Journal of Modern Optics, **44**, 2455 (1997)](https://dx.doi.org/10.1080/09500349708231894). J. F. Poyatos, J. Ignacioi Cirac, and P. Zoller, Complete Characterization of a Quantum Process: The Two-Bit Quantum Gate, [Physical Review Letters **78**, 390 (1997)](https://dx.doi.org/10.1103/PhysRevLett.78.390). Marcus P. da Silva, Olivier Landon-Cardinal, and David Poulin, Practical Characterization of Quantum Devices without Tomography, [Physical Review Letters **107**, 210404 (2011)](https://dx.doi.org/10.1103/PhysRevLett.107.210404). Steven T. Flammia and Yi-Kai Liu, Direct Fidelity Estimation from Few Pauli Measurements, [Physical Review Letters **106**, 230501 (2011)](https://dx.doi.org/10.1103/PhysRevLett.106.230501). Steven T. Flammia, David Gross, Yi-Kai Liu, and Jens Eisert, Quantum tomography via compressed sensing: error bounds, sample complexity and efficient estimators, [New Journal of Physics **14**, 095022 (2012)](https://dx.doi.org/10.1088/1367-2630/14/9/095022). Daniel M. Reich, Giulia Gualdi, and Christiane P. Koch, Optimal Strategies for Estimating the Average Fidelity of Quantum Gates, [Physical Review Letters **111**, 200401 (2013)](https://dx.doi.org/10.1103/PhysRevLett.111.200401). Martin Kliesch, Richard Kueng, Jens Eisert, and David Gross, Guaranteed recovery of quantum processes from few measurements, [arXiv:1701.03135 \[quant-ph\]](https://arxiv.org/abs/1701.03135). Joseph Emerson, Robert Alicki, and Karol Życzkowski, Scalable noise estimation with random unitary operators, [Journal of Optics B **7**, S347 (2005)](https://dx.doi.org/10.1088/1464-4266/7/10/021). Benjamin Lévi, Cecilia C López, Joseph Emerson, and David G. Cory, Efficient error characterization in quantum information processing, [Physical Review A **75**, 022314 (2007)](https://dx.doi.org/10.1103/PhysRevA.75.022314). Emanuel Knill, D. Leibfried, R. Reichle, J. Britton, R. B. Blakestad, J. D. Jost, C. Langer, R. Ozeri, S. Seidelin, and David J. Wineland, Randomized benchmarking of quantum gates, [Physical Review A **77**, 012307 (2008)](https://dx.doi.org/10.1103/PhysRevA.77.012307). Christoph Dankert, Richard Cleve, Joseph Emerson, and Etera Livine, Exact and approximate unitary 2-designs and their application to fidelity estimation, [Physical Review A **80**, 012304 (2009)](https://dx.doi.org/10.1103/PhysRevA.80.012304). Easwar Magesan, Jay M. Gambetta, and Joseph Emerson, Scalable and Robust Randomized Benchmarking of Quantum Processes, [Physical Review Letters **106**, 180504 (2011)](https://dx.doi.org/10.1103/PhysRevLett.106.180504). Joseph Emerson, Marcus P. da Silva, Osama Moussa, Colm A. Ryan, Martin Laforest, Jonathan Baugh, David G. Cory, and Raymond Laflamme, Symmetrized characterization of noisy quantum processes. [Science **317**, 1893 (2007)](https://dx.doi.org/10.1126/science.1145699). Easwar Magesan, Jay M. Gambetta, Blake R. Johnson, Colm A. Ryan, Jerry M. Chow, Seth T. Merkel, Marcus P. da Silva, George A. Keefe, Mary B. Rothwell, Thomas A. Ohki, Mark B. Ketchen, and Matthias Steffen, Efficient Measurement of Quantum Gate Error by Interleaved Randomized Benchmarking, [Physical Review Letters **109**, 080505 (2012)](https://dx.doi.org/10.1103/PhysRevLett.109.080505). Joel J. Wallman, Christopher Granade, Robin Harper, and Steven T. Flammia, Estimating the Coherence of Noise, [New Journal of Physics **17**, 113020 (2015)](https://dx.doi.org/10.1088/1367-2630/17/11/113020). Joel J. Wallman, Marie Barnhill, and Joseph Emerson, Robust Characterization of Loss Rates, [Physical Review Letters **115**, [060501]{} (2015)](https://dx.doi.org/10.1103/PhysRevLett.115.060501). Joel J. Wallman, Marie Barnhill, and Joseph Emerson, Robust characterization of leakage errors, [New Journal of Physics **18**, 043021 (2016)](https://dx.doi.org/10.1088/1367-2630/18/4/043021). Arnaud Carignan-Dugas, Joel J. Wallman, and Joseph Emerson, Characterizing universal gate sets via dihedral benchmarking, [Physical Review A **92**, 060302(R) (2015)](https://dx.doi.org/ 10.1103/PhysRevA.92.060302). Andrew W. Cross, Easwar Magesan, Lev S. Bishop, John A. Smolin, and Jay M. Gambetta, Scalable randomised benchmarking of non-Clifford gates, [npj Quantum Information **2**, 16012 (2016)](https://dx.doi.org/10.1038/npjqi.2016.12). Antonio D. C[ó]{}rcoles, Jay M. Gambetta, Jerry M. Chow, John A. Smolin, Matthew Ware, Joel Strand, B. L. T. Plourde, and Matthias Steffen, Process verification of two-qubit quantum gates by randomized benchmarking, [Physical Review A **87**, 030301(R) (2013)](https://dx.doi.org/10.1103/PhysRevA.87.030301). R. Barends, Julian Kelly, A. Megrant, A. Veitia, D. Sank, E. Jeffrey, T. C. White, J. Mutus, Austin G. Fowler, B. Campbell, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, C. Neill, P. J. J. O‘Malley, P. Roushan, A. Vainsencher, J. Wenner, A. N. Korotkov, A. N. Cleland, and John M. Martinis, Superconducting quantum circuits at the surface code threshold for fault tolerance. [Nature **508**, 500 (2014)](https://dx.doi.org/10.1038/nature13171). Julian Kelly, R. Barends, B. Campbell, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, Austin G. Fowler, I.-C. Hoi, E. Jeffrey, A. Megrant, J. Mutus, C. Neill, P. J. J. O’Malley, C. Quintana, P. Roushan, D. Sank, A. Vainsencher, J. Wenner, T. C. White, A. N. Cleland, and John M. Martinis, Optimal Quantum Control Using Randomized Benchmarking, [Physical Review Letters **112**, 240504 (2014)](https://dx.doi.org/10.1103/PhysRevLett.112.240504). Jeffrey M. Epstein, Andrew W. Cross, Easwar Magesan, and Jay M. Gambetta, Investigating the limits of randomized benchmarking protocols, [Physical Review A **89**, 062321 (2014)](https://dx.doi.org/10.1103/PhysRevA.89.062321). Tobias Chasseur and Frank K. Wilhelm, Complete randomized benchmarking protocol accounting for leakage errors, [Physical Review A **92**, 042333 (2015)](https://doi.org/10.1103/PhysRevA.92.042333). Harrison Ball, Thomas M. Stace, Steven T. Flammia, and Michael J. Biercuk, Effect of noise correlations on randomized benchmarking, [Physical Review A **93**, 022303 (2016)](https://dx.doi.org/10.1103/PhysRevA.93.022303). Timothy Proctor, Kenneth Rudinger, Kevin Young, Mohan Sarovar, and Robin Blume-kohout, What randomized benchmarking actually measures, [Physical Review Letters **119**, 130502 (2017)](https://dx.doi.org/10.1103/PhysRevLett.119.130502). Joshua Combes, Christopher Granade, Christopher Ferrie, and Steven T. Flammia, Logical Randomized Benchmarking, [arXiv:1702.03688 \[quant-ph\]](http://arxiv.org/abs/1702.03688). Easwar Magesan, Jay M. Gambetta, and Joseph Emerson, Characterizing quantum gates via randomized benchmarking, [Physical Review A **85**, 042311 (2012)](https://dx.doi.org/10.1103/PhysRevA.85.042311). Yuval R. Sanders, Joel J. Wallman, and Barry C. Sanders, Bounding quantum gate error rate based on reported average fidelity, [New Journal of Physics **18**, 012002 (2016)](https://dx.doi.org/10.1088/1367-2630/18/1/012002). Joel J. Wallman, and Steven T. Flammia, Randomized benchmarking with confidence, [New Journal of Physics **16**, 103032 (2014)](https://dx.doi.org/10.1088/1367-2630/16/10/103032). Christopher Granade, Christopher Ferrie, and David G. Cory, Accelerated randomized benchmarking, [New Journal of Physics **17**, 013042 (2015)](https://dx.doi.org/10.1088/1367-2630/17/1/013042). Michael A. Nielsen, A simple formula for the average gate fidelity of a quantum dynamical operation, [Physics Letters A **303**, 249 (2002)](https://dx.doi.org/10.1016/S0375-9601(02)01272-0). F. L. Bauer and C. T. Fike, Norms and exclusion theorems, [Numerische Mathematik **2**, 137 (1960)](https://dx.doi.org/10.1007/BF01386217). David Pérez-García, Michael M. Wolf, Denes Petz, and Mary Beth Ruskai, Contractivity of positive and trace-preserving maps under $L_p$ norms, [Journal of Mathematical Physics **47**, 083506 (2006)](https://dx.doi.org/10.1063/1.2218675). Robin Blume-Kohout, John King Gamble, Erik Nielsen, Kenneth Rudinger, Jonathan Mizrahi, Kevin Fortier, and Peter Maunz, Demonstration of qubit operations below a rigorous fault tolerance threshold with gate set tomography, [Nature Communications **8**, 14485 (2017)](https://dx.doi.org/10.1038/ncomms14485). Mark D. Bowdrey, Daniel K. L. Oi, Anthony J. Short, Konrad Banaszek, and Jonathan A. Jones, Fidelity of single qubit maps, [Physics Letters A **294**, 258 (2002)](https://dx.doi.org/10.1016/S0375-9601(02)00069-5).
{ "pile_set_name": "ArXiv" }
--- abstract: 'Naturally ventilated buildings harness freely-available resources such as internal buoyancy gains and wind forcing in achieving comfortable interior conditions. Although these resources are free, they are time-variable and can be difficult to control. As a result of this the nonlinear interplay between sometimes competing resources may lead to sub-optimal ventilation states. This problem has been explored by a number of previous researches e.g. Flynn & Caufield ([*Building and Environment*]{}, [**44**]{}, 216–226, 2009) who, in studying these ventilations states, demonstrated complicated transitions characterized by hysteresis even for the simple case of a one-zone building. The objectives of this research are to extend the previous (theoretical) analysis by conducting complementary numerical experiments using sophisticated algorithms capable of describing turbulent, buoyancy driven flow. A further objective of this study is to specifically “decouple” the source, which was previously assumed to supply both heat and mass to the interior space. Rather, the flow dynamics in a case where heat and mass are supplied independently are examined.' address: | $^{1}$ Section de G[' e]{}nie M[' e]{}canique, Ecole polytechnique f[' e]{}d[' e]{}rale de Lausanne, Lausanne, Switzerland\ $^{2}$Dept. of Mechanical Engineering, University of Alberta, Edmonton, AB, Canada author: - 'Vincent Mayoraz$^{1,2}$ and M.R. Flynn$^2$' bibliography: - 'refs.bib' title: 'Natural vs. blocked ventilation in naturally ventilated buildings – the effect of finite and decoupled sources' --- Fluid mechanics ,Natural ventilation ,Numerical simulations Introduction ============ “Naturally-ventilated” buildings are a special class of net-zero energy buildings that seek to reduce the energy expense by harnessing freely-available resources (solar radiation, wind forcing, internal buoyancy generation) in forcing air into and out of the built environment. Although the previously mentioned resources are free, they are, with the possible exception of patterns of building occupation and the switching on and off of electrical and related equipment, unpredictable. Strictly regulating interior conditions to ensure that they fall within a range deemed comfortable is therefore no trivial task. In this sense, it is important to understand the flow dynamics within the buildings, in order that appropriate control schemes may be devised. Numerous studies have tried to propose practical equations, giving the architects/designers simple rules of thumb that nonetheless capture the essential physics governing architectural flows [@kayeflynn2012].\ In parallel, researchers have used more sophisticated numerical simulations to predict the flow behavior of natural ventilation [@Kaye2009]. Note, however, that numerically studies have generally been limited to certain specific scenarios or flow states and not to the whole range of air flows that can take place in naturally ventilated buildings. Therefore, one of the major objectives of the present research is to examine the details of the unexpected flow behavior when the thermal source is “non-ideal,” i.e. it supplies both heat and mass to the interior space. This particular research interest is one piece of a much larger puzzle aimed at understanding how the principles of natural ventilation, originally developed for application in temperate climates, can be profitably extended to a broader range of locales. In so doing, the overarching objective is to reduce so much as possible the need for expensive and energy-intensive mechanical systems, which have for too long characterized building design. Natural ventilation offers an appealing alternative to traditional practice insofar as the waste heat produced during summer months can be used to remove stale air to be replaced by fresh outside air. By contrast, in winter that heat, no longer unwanted, can be used to generate controlled air flows helping to maintain comfortable temperatures inside.\ Finally this research will try to show that it is possible to perform good quality studies of complex flow phenomenon inside naturally ventilated buildings using commercial CFD softwares on standard workstations. In the long term, this could open the door to a popularization of naturally ventilated buildings since the design of such buildings was until not so far ago too complex to be applied to homes and commercial buildings, this due to time limitations and monetary constraints. The analytical model ==================== Description of the problem -------------------------- Due to the complexity of the problem even for simplified geometries, natural ventilation needs to be modeled. In any case these models can roughly be divided in five different types, taking into account -or not - different parameters influencing the flow. These parameters and the associated models are summarized in the Table \[tab:general\] and are detailed in the following paragraphs. Case $P$ $Q_s$ W Experimental validation ------ ------- ------- ------- ------------------------- 1 $>$ 0 = 0 = 0 @Kaye2004 2 $>$ 0 = 0 $>$ 0 @Flynn2009 3 = 0 $>$ 0 = 0 @Woods2003 4 = 0 $>$ 0 $>$ 0 - 5 $>$ 0 $>$ 0 $>$ 0 - : Different models and their parameters. $P$ is the power of the heat source, $Q_s$ is the source volume flux and W is the speed of the wind[]{data-label="tab:general"} The objective of this section is not only to highlight and show the main relevant studies carried out in the field but also to define the problem and the associated quantities that will be used further together. It will give an idea of the limitations of the mentioned models. The following section emphasizes the main results that will further be compared with the CFD simulations and is by no means a comprehensive list of the existing prediction models.\ The emptying filling box model {#sec:ideal-no-wind-theo} ------------------------------ The emptying filling box model considers an empty box with a floor surface area $S$, height $H$ and bottom and top openings area $A_b$ respectively $A_t$. A plume rises from a source to the top of the box due to thermal differences and spreads out laterally to form a density interface between the plume outflow and the ambient fluid. @Kaye2004 have been studying this flow using the “filling box” model developed by @Baines1969 and @Linden1990. The source $P$ is considered ideal since only heat and no mass is added to the system. The outside temperature and density are fixed to $T_0$ and $\rho_0$ as shown in Fig. \[fig:one\_box\_building\]. ![Illustration of the considered idealized geometry and its relevant parameters.[]{data-label="fig:one_box_building"}](Slide1.JPG){width="\textwidth"} Different properties of the inner flow can be predicted by theory, in function of the initial parameters and the geometry of the box. The “filling time,” i.e. the time for the ventilated box with a single source to be filled with buoyant fluid, can be expressed as $$t_f = \frac{S}{\lambda B^{1/3} H^{2/3} }$$ On the other hand, the “draining time,” i.e. the time taken to flush the box of the aforementioned fluid can also be predicted: $$t_d = \frac{ \lambda^{1/2} H^{4/3} S }{ A^* B^{1/3} }$$ Both of these times depend on the constant $\lambda$ which is widely used in the modeling of plumes and is equal to $$\lambda = \frac{6 \alpha}{5} \left( \frac{9 \alpha}{10} \right)^{1/3} \pi^{2/3}$$ where the coefficient $\alpha$ is the plume entrainment coefficient. It depends on the geometry of the plume and in this case (asymmetric plume with an assumed top-hat distribution for vertical velocity and density) is about $\sim$ 0.114. $B$ represents the plume buoyancy flux created by the temperature difference. It is expressed in function of gravity, the heat source power $P$ and the characteristic properties of the fluid: $$\label{equ:buoyancy1} B = \frac{ Pg }{ c_p \rho_0 T_0 }$$ $A^*$ is a non-dimensional parameter called the effective area defined to facilitate the comparison for different sizes and geometries of openings. It is a function of $A_t$ and $A_b$, and is expressed as $$% A^* = \left[ \frac{1}{(\sqrt{2} c_t A_t)^{2} } + \frac{1}{(\sqrt{2} c_b A_b)^{2}} \right]^{-1/2} A^* = \frac{ \left( 2 c_t c_b A_t A_b \right)^2 }{ \sqrt{2c_t^2 A_t^2 + 2 c_b^2 A_b^2} }$$ where $c_t$ and $c_b$ denote the discharge loss coefficients associated with the flow through the bottom and top openings. @Hunt2000 showed that these coefficients exhibited density dependence but within the framework of this project they will be considered constant and equal to 0.6 for simplicity and since their impact on the final predictions is low.\ Using the filling box model again, @Linden1990 proved that in the case of a single buoyant point source the steady state interface height could be expressed in function of the normalized effective area and the coefficient $\lambda$ only: $$\label{equ:no_wind} \frac{A^*}{H^2} = \lambda^{3/2} \cdot \frac{t_f}{t_d} = \frac{\lambda^{3/2} \xi^{5/2}}{\sqrt{1-\xi}}$$ where the interface height is non-dimensionalized to ease comparison with different geometries ($\xi = h/H$). Fig. \[fig:xi\_vs\_mu\_and\_AH2\] represents the variation of the steady state interface height with respect to the non-dimensionalized area, which corresponds to the equ. (\[equ:no\_wind\]). $A^*$ is “symmetric” in $A_t$ and $A_b$, thus the height of the interface can be modified with a change of $A_t$ or $A_b$. Small effective areas will induce lower interface heights since the size of the openings will not be large enough to allow the escape of the buoyant layer. On the other hand, very large effective areas will induce a density interface close to the roof, corresponding to a case where almost all the heated air flows out of the room. ![Variation of the non-dimensional interface height with respect to the time scale ratio with regard to the effective area.[]{data-label="fig:xi_vs_mu_and_AH2"}](xi_vs_AH2-eps-converted-to.pdf){width="80.00000%"} The transient behavior of this flow has also been studied by @Kaye2004 and @Moradi2017. They showed that the evolution of the interface height could be expressed by the following system of non-dimensional differential equations: $$\frac{d \xi}{d \tau} = \frac{1}{\sqrt{\mu}} \sqrt{\eta (1-\xi)} - \sqrt{\mu} \xi^{5/3}$$ $$\frac{d \eta}{d \tau} = \sqrt{\mu} \ \frac{1 - \xi^{ \frac{5 \delta}{3} }}{1 - \xi}$$ where $\tau$ represents the non-dimensional time, $\mu$ is the ratio of the draining and filling times ($\mu = t_d/t_f$) and $\eta$ represents the dimensionless average reduced gravity of the buoyant layer: $$\eta = g' \frac{\lambda H^{5/3}}{B^{2/3}}$$ Here, the reduced gravity is defined by $$g' = g \left( \frac{T_{up} - T_0}{T_0} \right) = \frac{1}{\lambda} \left[ \frac{B^2}{(\xi H)^5} \right]^{1/3}$$ where the first equality corresponds to the general expression of the reduced gravity and the second one to the equivalent definition for natural ventilation, with regards to the interface height. Fig. \[fig:xi\_vs\_tau\] represents the evolution of the non-dimensional interface height $\xi$ with respect to the non-dimensional time $\tau$ for different values of $\mu$. It can be seen that the interface overshoots before reaching its steady state value. The analytical model was compared with experimental data by @Kaye2004 from a small-scale model using salty water to reproduce the thermal plume. The theoretical model showed good agreement with the small-scale experimental model. ![Evolution of the non-dimensional interface height with respect to the non-dimensional time for different values of $\mu$.[]{data-label="fig:xi_vs_tau"}](xi_vs_tau-eps-converted-to.pdf){width="80.00000%"} Influence of the external atmosphere {#sec:external-atm-theo} ------------------------------------ To improve the realism of the previous model, the influence of the external atmosphere on the inner flows is added. The governing features of the previously mentioned flows have been studied by @Linden1999. It differentiates two main regimes that can be encountered in natural ventilation: mixing ventilation in which the interior is at an approximately uniform temperature, and displacement ventilation where there is strong internal stratification. In this sense, @Flynn2009 have been exploring the role of wind, exerting different static pressures on the leeward and windward facades of the building. To do so, the idealized geometry presented in the Fig. \[fig:wind\_forcing\] is considered. The presence of an adverse wind induces a pressure drop $\Delta p$ between the top and the bottom of the room which needs to be taken into account in the computations of the stratification layer. ![Schematic illustration of observable flow regimes in the presence of an adverse wind assuming a zero volume flux source of infinite temperature. In this circumstance, naturally ventilated (left) and well-mixed (right) states are possible.[]{data-label="fig:wind_forcing"}](Slide3.JPG){width="\textwidth"} Thereby, in the case of a sufficient thermal forcing the influence of an external pressure (produced by wind) can be added to equ. (\[equ:no\_wind\]) which becomes $$\label{equ:wind} \frac{A^*}{H^2} = \frac{ \lambda^{3/2} \xi^{5/3} }{ \sqrt{ \frac{ 1-\xi }{ \xi^{5/3} } - \lambda \frac{ \Delta p}{\rho_0} \left( \frac{H}{B} \right)^{2/3}} }$$ The effect of the pressure difference is taken into account in the equation and weighted by the strength of the buoyancy flux $B$. Intuitively higher adverse wind forces will induce higher pressure differences and will tend to lower the height of the interface. On the other hand a strong buoyancy flux $B$ will compensate the effect of the pressure drop and help to raise the interface height. New variables need to be introduced in order to describe this more complex flow. In the case of a single building the Froude number $Fr$ is equal to $$\label{equ:froude} Fr = \sqrt{ \frac{\Delta p}{\rho_0} \left( \frac{H}{B} \right)^{2/3} }$$ and can be recognized as a term of equ. (\[equ:wind\]). Moreover,a non-dimensional number is defined: $\delta = \lambda Fr^2$ which will be used later to simplify the mathematical relations between the different parameters describing the flow. Fig. \[fig:xi\_vs\_delta\] represents the non-dimensional interface height with respect to that non-dimensional parameter $\delta$ (equ. \[equ:wind\]). It can be seen that the height of the interface is lowered by the action of the wind inducing a higher pressure on the inlet of the building. The influence of the effective area is diminished in the case of high values of $\delta$ which could be expected since the wind becomes the main driving parameter of the flow. ![Variation of the non-dimensional interface height with respect to the normalized Froude number $\delta$ (equ. \[equ:wind\]).[]{data-label="fig:xi_vs_delta"}](xi_vs_delta-eps-converted-to.pdf){width="80.00000%"} Moreover, the wind-driven and buoyancy-driven volume fluxes can be expressed as $$Q_B = A^* (g' H)^{1/2} \qquad and \qquad Q_w = A^* Fr \left( \frac{B}{H} \right)^{1/3}$$ Thereafter, the ventilation rate for the building can be computed in function of the parameters defined above, for the stratified state, respectively mixed state: $$\label{equ:fluxes} Q = A^* \sqrt{g' H (1 - \xi) - \frac{\Delta p}{g' \rho_0}}, Q_w < Q_B -\sqrt{Q_w^2 - Q_B^2}, Q_w > Q_B$$ The flux is considered positive if the buoyancy forces are dominant, i.e. the flow will go out through the upper (top) opening and negative when the wind dominates, where the air will escape through the lower opening. @Hunt2004 proved that the flow state could be predicted with the parameters defined before with a simple polynomial equation: $$\label{equ:states1} \left( \frac{Q}{Q_W} \right)^3 - \frac{Q}{Q_W} - \frac{(H^2/A^*)}{Fr^3} = 0$$ ![Non-dimensionalized total volume flow with respect to the Froude number (equ. \[equ:states1\]).[]{data-label="fig:QQw_vs_Fr"}](QQw_vs_Fr-eps-converted-to.pdf){width="80.00000%"} As shown in Fig.\[fig:QQw\_vs\_Fr\] multiple solutions exist in general. Two of these solutions correspond to the well mixed state (only one of which is stable) and the other corresponds to the naturally ventilated state.\ To explain the transition from mixing to stratified, two cases can be considered. On one hand a case with a constant wind and no heat source ($B$ = 0). This case corresponds to a stable mixing flow with infinite Froude number where the normalized flow rate $Q/Q_W$ = -1. When $B$ is increased from zero, $Fr$ decreases (equ. \[equ:froude\]), and the steady state solution falls on the lower (stable) part of the solution. The buoyancy driven flux $Q_B$ increases in magnitude with $B$ and hence $|Q/Q_w|$ decreases. At the critical Froude number corresponding to the left end of the mixed solution, the wind will be unable the maintain the flow in the room for any further increase in $B$, which will result in a transition from the mixed to the stratified regime.\ On the other hand, if the room contains a non-zero heat source ($B$&gt;0) and the wind speed is initially zero, the system will be located on the upper part of the solution, where $Q/Q_w$ is equal to infinity. Supposing now that the wind is progressively increased, $Fr$ will increase as $Q_w$ so that $Q/Q_w$ decreases and buoyancy accumulates in the upper part of the building. This behavior goes on until the wind is too strong and the flow, unable to maintain the stratified regime, moves to the unstable part of the equation. As explained above, since this state is unstable, the flow will move towards one or the other solution. Physically this corresponds to a hysteresical behavior of the flow at low values of $Q/Q_w$. Blocked ventilation {#sec:non-ideal-no-wind} ------------------- The critical state depicted in Fig.\[fig:QQw\_vs\_Fr\], between the well mixed and stratified state, where the flow will become hysteresical can be modeled more finely if a non-zero mass flux inlet is added at the heat source. This will result in a state where flow escapes from both top and bottom openings, whereas still respecting the mass conservation for the system. This type of regime is called “blocked ventilation.” The buoyancy flux defined in equ. (\[equ:buoyancy1\]) for a heat power source $P$ becomes for a non-zero source mass flux $$B_s = g Q_s \left( \frac{T_{in}-T_0}{T_0} \right)$$ where $Q_s$ is the mass flux in \[m$^3$/s\] and $T_{in}$ is the temperature of the air injected. In order to simplify the further mathematical relations, the source mass flux is non-dimensionalized as follows: $$\zeta_s = \frac{1}{H} \left( \frac{Q_s}{\lambda B_s^{1/3}} \right)^{3/5}$$ Blocked ventilation and the effect of the non-ideal heat source were studied by @Woods2003 among others. They carried out both analytical and experimental campaigns ending up with a validated model for negative plumes initiated by source mass fluxes. They proposed an equation linking the density interface height with the properties of the flow and the inlet parameters. Using the Boussinesq approximation, where positive and negative plumes are equivalent due to the symmetry of the driving forces, it is possible to show that the model studied by @Woods2003 can be applied in the case studied here and re-written with the parameters defined above in a non-dimensional form: $$\label{equ:woods_modified} \frac{A^*}{H^2} = \frac{ \lambda^{3/2} \left[ (\xi+\zeta_s)^{10/3} - \frac{A_T^2 \left( 2 \zeta_s^{5/3} (\xi+\zeta_s)^{5/3} -\zeta_s^{10/3} \right) }{A_T^2 + A_B^2} \right] ^{1/2}} {\sqrt{ \frac{ 1-\xi }{ (\xi+\zeta_s)^{5/3} } } }$$ which would correspond to a “corrected” form of equ. (\[equ:no\_wind\]) for non-ideal plume sources. The variation of the interface height with respect to the source mass flux is plotted in Fig. \[fig:xi\_vs\_qs\_woods\]. ![Variation of the non-dimensional interface height with respect to the non-dimensional source mass flux.[]{data-label="fig:xi_vs_qs_woods"}](xi_vs_zetas-eps-converted-to.pdf){width=".8\textwidth"} The height of the interface is decreasing with the increasing mass flux due to the forcing of the heat mass flux. More air will be introduced inside the building and for a fixed effective area the volume taken by the buoyant layer will be larger. As a consequence the interface will tend to be pushed against the floor. It is interesting to see that for a certain threshold the interface height reaches the bottom of the building. This corresponds to the case where the heat source flux is very large (relative to the given effective area) and discharged plume fluid will completely fill the room. The other extreme case for when the interface height reaches the level of the ceiling corresponds to low heat source flux where the amount of air entering the building is too small to create a buoyant layer, which instead directly flows out through the (top) opening. Full model with non ideal source {#sec:therory-full-non-ideal} -------------------------------- Finally, all the parameters ruling the flow are taken into account. In this scenario, the relation between the non-dimensional interface height, the pressure drop and the source mass flux can be obtained with a combination of equ. (\[equ:wind\]) and equ. (\[equ:woods\_modified\]): $$\label{equ:full_equ} \frac{A^*}{H^2} = \frac{ \lambda^{3/2} \left[ (\xi+\zeta_s)^{10/3} - \frac{A_T^2}{A_T^2 + A_B^2} \left( 2 \zeta_s^{5/3} (\xi+\zeta_s)^{5/3} -\zeta_s^{10/3} \right) \right]^{1/2} }{ \sqrt{ \frac{ 1-\xi }{ (\xi+\zeta_s)^{5/3} } - \lambda \frac{ \Delta p}{\rho_0} \left( \frac{H}{B_s} \right)^{2/3} } }$$ This equation has not yet been validated by experimental data in the literature. By way of validation, the limiting cases for $\Delta p = 0$ (no wind) and $\zeta_s = 0$ (ideal source) have been solved analytically and equ. (\[equ:full\_equ\]) is converging to equ. (\[equ:no\_wind\]), respectively equ. (\[equ:woods\_modified\]) which serves as a validation for those limiting cases. Even though the full validity has not been proven, since no other theoretical model has been found in the literature, this equation will be used for the comparison with the numerical simulations. ![Variation of the non-dimensional interface height with respect to the non-dimensional parameter $\delta$ for different non ideal heat sources $\zeta_s$ and a normalized effective area of $A^*/H^2$ = 0.021.[]{data-label="fig:xi_vs_delta_flynn"}](xi_vs_delta_flynn-eps-converted-to.pdf){width="80.00000%"} Fig. \[fig:xi\_vs\_delta\_flynn\] represents the variation of the non-dimensional interface height with regards to the non-dimensional parameter $\delta$ for different source mass fluxes. As expected the interface height is lower for higher pressure differences over the two sides of the building since the wind will “push” the interface against the bottom. On the other hand, as already seen in Fig. \[fig:xi\_vs\_qs\_woods\] the lowering of the interface will be amplified by the strength of the source mass flux.\ This time, the characteristic polynomials determining the relative importance of wind, the internal buoyancy respectively cannot be determined and expressed with one equation only, similarly to what was done for the case of an ideal source (equ. \[equ:states1\]). Three equations must be used, taking into account the source volume flux. To keep all the parameters non-dimensional the characteristic volume flux is chosen differently and instead of using $Q_w$ (as was done in the derivation of equ.(\[equ:states1\]), but rather $Q_{B_{s}} = A^* B_s^{1/3} H^{1/3}$. Accordingly, the non-dimensional source volume flux and the non-dimensional flux through the top opening are defined as: $$q_s = Q_s/Q_{B_{s}} \quad and \quad q_T = Q_T/Q_{B_{s}}$$ The three equations are corresponding to the three different regimes appearing in that kind of flow are: $$\label{equ:poly1} q_T^{3} - 2(1-a^2)q_s q_T^2 + q_T \left[ (1-a^2) q_s^2 Fr^2 \right] - 1 +\xi = 0$$ $$\label{equ:poly2} (1-2a^2)q_s q_T^2 - 2(1-a^2) q_s^2 q_T + (1-a^2)q_s^3 - q_s Fr^2 + 1 = 0$$ $$\label{equ:poly3} q_T^3 - (3-2a^2) q_s q_T^2 + \left[ 3(1-a^2)q_s^2 -Fr^2 \right] q_T -(1-a^2)q_s^3 + q_s Fr^2 -1 = 0$$ for a naturally-ventilated regime equ. (\[equ:poly1\]), blocked regime equ. (\[equ:poly2\]), and in the case of a well-mixed flow regime equ. (\[equ:poly3\]). The Fig. \[fig:qT\_vs\_Fr\] represents the resulting curves [@Flynn2009], they are plotted for different non-dimensional source mass flux in function of the Froude number at a fixed effective area. The upper part of the solution (continuous line) corresponds to a stratified regime, the lower part (dashed line) to the mixed regime. The two dotted lines show the limits of the blocked regime (i.e. the solution in-between is blocked). The curves look different from the one presented in Fig. \[fig:QQw\_vs\_Fr\] because the normalization flux is different but the general idea is the same and they allow the prediction of the flow in a similar manner. ![Non-dimensional flow rate through the top opening with respect to the Froude number for different source mass flux at a fixed effective area of $A^*/H = 0.021$.[]{data-label="fig:qT_vs_Fr"}](qT_vs_Fr-eps-converted-to.pdf){width="\textwidth"} Decoupled heat sources ---------------------- The heat source which was previously assumed to supply both heat and mass to the interior space is this time considered to supply heat and mass independently. This kind of model has so far never been studied analytically. The main objective of decoupling the heat source is to see if it involves any changes in the prediction of the interface height that might be useful to improve the existing models. Therefore, a numerical investigation only will be carried out to study any relevant influence of that modification on the final prediction. The numerical model =================== The finite volume method ------------------------ Within the framework of this project the commercial software ANSYS Fluent was used to solve the flows. This software is based on the finite volume method, which solves the partial differential equations (cf. equ. \[equ:continuity\]-3) describing the behavior of the flow iteratively, until a solution reasonably (and arbitrarily) close to reality is found. Theses equations are solved numerically on a mesh, constituted of small jointed volumes (or surfaces in 2D). Governing equations ------------------- In fluid mechanics, the Navier-Stokes equations are a set of non-linear partial differential equations describing the Newtonian flows behavior. The resolution of these equations by modeling the fluid as a continuum is extremely complex and analytical solutions only exist for a few simplified cases. Those equations are derived from the conservation laws applied to mass and momentum. In this case, the conservation of energy is added to the Navier-Stokes equations in order to model thermal effects [@Loomans1998].\ *Conservation of mass (continuity equation):* $$\label{equ:continuity} \frac{\partial \rho }{\partial t } + \frac{\partial }{\partial x_i } (\rho u_i) = 0$$ *Conservation of momentum:* $$\label{equ:momentum} \frac{\partial}{\partial t }(\rho u_i) + \frac{\partial}{\partial x_j } ( \rho u_i u_j ) = - \frac{\partial p}{\partial x_i } + \frac{\partial}{\partial x_j } \left[ \nu \left( \frac{\partial u_i }{\partial x_j } + \frac{\partial u_j}{\partial x_i } \right) \right] + \rho g_i$$ *Conservation of energy:* $$\label{equ:energy} \frac{\partial}{\partial t } (\rho H ) + \frac{\partial}{\partial x_i } (\rho u_i H ) = \frac{\partial}{\partial x_i } \left[ \frac{ K }{ c_p } \frac{ \partial H }{ \partial x_i } \right] + S_H$$ where $\rho$ is the density of the fluid, $u_i$ is the velocity component ($u,v,w$), $p$ is the pressure, $\mu$ indicates kinematic viscosity, $H$ the enthalpy and $S_H$ a source term. $K$ and $c_p$ represent the thermal conductivity, respectively the specific heat of the fluid. The time is indicated with $t$, $x_i$ is the coordinate axis (x,y,z) and $g_i$ is the gravitational acceleration. Turbulence modeling ------------------- The Direct Numerical Simulation of these equations (DNS) requires a very fine mesh since the whole range of spatial and temporal scales of turbulence are resolved and the smallest eddies in the flow need to be captured. According to @Nieuwstadt1990 the number of grid points required to describe turbulent motions with this method is at least $Re^{9/4}$ which would induce huge computational times (several years) for the case of naturally-ventilated buildings and are therefore not imaginable nowadays, even with the help of large computing clusters.\ The second type of simulations, called Large Eddy Simulations (LES) is mostly based on the work of @Smagorinsky1963 and @Deardoff1970 proving the turbulent motion can be separated into large eddies and small eddies without having a significant impact on the evolution of the large eddies. Therefore, the large eddies are directly simulated whereas the small eddies are modeled with turbulent transport approximations. Thus, LES is clearly superior to turbulent transport closure wherein the transport terms are treated empirically. This techniques is becoming more and more relevant these days with the rapid increase in computational power on commercial machines.\ Nonetheless, in the case of this study a Reynolds Averaged Navier-Stokes (RANS) model will be used both because of its computational efficiency and because of the broad application of RANS-based models in industry. To this end, the mean parameters are more useful than the instantaneous turbulent-flow parameters in the case of naturally ventilated buildings thus these models will give appreciable results [@Zhiqiang2007] that are of interest to architects, mechanical engineers and others interested in modeling transport processes in the built environment. These average-models for the turbulent flow have calculated the statistical characteristics of the turbulent motions by averaging the flow equations over time. The solution variables in the continuous Navier-Stokes equations are thereby decomposed into the mean and fluctuating components [@theory_guide]. $$\label{equ:vel_average} u = \overline{u}+u'$$ with $\overline{u}$ and $u'$ being the mean and fluctuating components of the velocity field. The other variables (such as pressure and scalar quantities) are also decomposed same way. $$\label{equ:scal_average} \phi = \overline{\phi}+\phi'$$ By introducing equ. (\[equ:vel\_average\]) and equ. (\[equ:scal\_average\]) into the continuous equ. (\[equ:continuity\] to 3.3) and taking a time average (ensemble average), new non-linear terms appear in the set of equations: $$- \rho \overline{u_i' u_j'} \quad and \quad - \rho \overline{u_i' H'}$$ With the apparition of these new terms (Reynolds terms), the set of equations introduces a closure problem and more equations are necessary to solve it. Thus, the effect of turbulence is represented as an increased viscosity/diffusivity by its replacement with linear terms and scalar coefficients based on the hypothesis (@Boussinesq1887) that the turbulent stresses are proportional to the mean velocity gradients. These last non-linear fluctuating terms need to be modeled in order to obtain computationally solvable forms of the RANS equations. Numerous models have been developed over the past years and even though they will all - in general - give a solution, their ability to correctly predict the behavior of a flow may vary significantly. Therefore, choosing the right model for turbulence is a key element to the success of numerical flow simulations.\ Many researchers have been trying to assess different turbulent models for naturally ventilated buildings and the two most popular RANS models that have been successfully used for indoor ventilation flows are SST $k- \omega$ and RNG $k-\epsilon$ [@Zhiqiang2007]. They showed the best overall performance compared to the other models in terms of accuracy, computational efficiency and robustness. Moreover, the RNG $k-\epsilon$ has proven to be superior to the SST $k-\omega$ for low-turbulence flow [@Craven2006; @Zhang2005] such as the ones studied in this thesis. For these reasons, the theoretical development of the RNG $k-\epsilon$ model only will be made in details.\ The standard $k-\epsilon$ model is based on model transport equations for the turbulent kinetic energy $k$ and its dissipation rate $\epsilon$. The derivation of the model is based on the assumption that the flow is fully turbulent and that the effects of molecular viscosity are negligible [@theory_guide]. It is therefore only valid for fully turbulent flows and would fail to predict the majority of low-turbulence flow appearing in natural ventilation. For these reasons, the renormalization group theory (RNG) is used. It has additional terms in the $\epsilon$ equation which provide analytically derived differential formula for the effective viscosity that accounts low-Reynolds number effects. The transport equations (\[equ:continuity\]-3.3) are rewritten in function of the turbulent kinetic energy $k$ and the dissipation rate $\epsilon$: $$\frac{\partial (\rho k)}{\partial t } + \frac{\partial }{\partial x_i } (\rho k u_i) = \frac{\partial}{\partial x_j} \left[ \left( \nu + \frac{\nu_t}{\sigma_k} \right) \frac{\partial k}{\partial x_j} \right] + G_k + G_b - \rho \epsilon - Y_M + S_k$$ and $$\label{equ:RNG_ke_2} \frac{\partial (\rho \epsilon)}{\partial t } + \frac{\partial }{\partial x_i } (\rho \epsilon u_i) = \frac{\partial}{\partial x_j} \left[ \left( \nu + \frac{\nu_t}{\sigma_{\epsilon}} \right) \frac{\partial \epsilon}{\partial x_j} \right] + C_{1 \epsilon} \frac{\epsilon}{k} (G_k + C_{3 \epsilon}G_b) - C_{2 \epsilon} \rho \frac{\epsilon^2}{k} + S_{\epsilon}$$ $G_k$ represents the generation of turbulent kinetic energy due to the mean velocity gradients, $G_b$ is the generation of turbulent kinetic energy due to buoyancy. $Y_M$ represents the contribution of the fluctuating dilatation in compressible turbulence to the overall dissipation rate. $C_{1 \epsilon}$, $C_{2 \epsilon}$ are constant values determined analytically by the RNG theory, $C_{3 \epsilon}$ is also a constant whose determination will be explained below, $\sigma_k$ and $\sigma_{\epsilon}$ are the turbulent Prandtl numbers for $k$ and $\epsilon$. $S_{k}$ and $S_{\epsilon}$ are the source terms that account for the production of kinetic energy, respectively dissipation rate at the simulated sources.\ As it was already mentioned earlier, the effect of buoyancy is one of the major flow-driving aspects of naturally ventilated buildings and it is therefore important to understand the way its modeled. For $k-\epsilon$ the generation of turbulence models is given by $$G_b = \beta g_i \frac{\mu_t}{Pr_t} \frac{\partial T}{\partial x_i}$$ where $Pr_t$ is the turbulent Prandtl number for energy and $g_i$ the component of the gravity in the associated direction. For the standard $k-\epsilon$ model the Prandtl number is considered constant but in the case of the RNG model it is computed using a formula analytically derived by the renormalization group theory to give it more accuracy.\ In ANSYS Fluent the buoyancy effects on the generation of $k$ are relatively well understood [@user_guide] but its effects on $\epsilon$ are less clear. By default, they are neglected by keeping the term $G_b$ in the transport equation (\[equ:RNG\_ke\_2\]) to zero. However, recent versions of ANSYS offer the possibility to include these effects in the solver’s viscous model box. The degree of influence is weighted by the constant $C_{3 \epsilon}$, computed accordingly to $C_{3 \epsilon} = tanh|v/u|$ where $v$ is the component of the flow velocity parallel to gravity and $u$ the component of the flow velocity perpendicular to $v$. In this way, the constant $C_{3 \epsilon}$ will be maximized (equal to 1) if the shear layer of the buoyant flow is aligned with gravity and zero if the buoyant shear layer is perpendicular, which will annihilate the influence of the buoyancy on the dissipation rate. Since the addition of this last parameter is fairly recent to Fluent (version 17.2) no proof of its efficiency has been found in the literature but it will be used accordingly to the recommendations of the user’s guide [@user_guide]. ### Hydrostatic head and Boussinesq approximation Since gravitational acceleration is activated in the solver in order to model the buoyant effects, the pressure field will include the hydrostatic head. Mathematically, this is accomplished by redefining the pressure in terms of a modified pressure that includes the hydrostatic head (denoted p’) as follow : $$p' = p- \rho_0 g_i \cdot r_i$$ where $r_i$ is the position vector component. Noting that $$\frac{\partial}{\partial x_i} (\rho_0 g_i r_i) = \rho_0 g_i$$ it follows that $$\label{equ:nabla_p_prime} \frac{\partial p'}{\partial x_i} = \frac{\partial}{\partial x_i} (p - \rho_0 g_i \cdot r_i) = \frac{\partial p}{\partial x_i} - \rho_0 g_i$$ The substitution of equ. (\[equ:nabla\_p\_prime\]) in the momentum equation (equ. \[equ:momentum\]) gives a pressure gradient and gravitational body force terms of the form $$-\frac{\partial p'}{\partial x_i} + (\rho - \rho_0) g_i$$ where $\rho$ is the actual fluid density. Since the Boussinesq hypothesis is made, the body force term becomes $$(\rho - \rho_0) g_i \approx \rho_0 \beta (T-T_0) g_i$$ The consequence of this treatment of the gravitational body force is that the report of static and total pressures will not show any influence of the hydrostatic pressure. It is important to note this is consistent with equ. (\[equ:wind\]) proposed by @Linden1999, @Flynn2009. Their models do not take into account the hydrostatic head over the building so that the position of the openings on the sides of the building should not influence the prediction of the interface height. The advantage of using the Boussinesq hypothesis over a compressible gas relation is that it simplifies considerably the equations solved since the density is considered constant for all terms except when it is multiplied by gravity. This allows faster convergence thus shorter simulation times. The hypothesis is accurate as long as changes in actual density are small, specifically, it is valid when $\beta(T-T_0) \ll 1$ which is the case in the current setup (thermal expansion coefficient of air: $\beta$ = 0.0034 K$^{-1}$, maximum temperature difference: $\Delta T$ = 10 K). Numerical simulation of the naturally ventilated flow ===================================================== Simulation of the emptying-filling box -------------------------------------- ### Description of the problem The first set of numerical simulations that were carried out aimed at simulating the effect of a buoyancy driven flow inside a single-zone building. To validate the developed model, the numerical simulations performed by @Kaye2009 were reproduced. The domain considered is a 7.5m $\times$ 7.5m $\times$ 3m building as it can be seen in Fig. \[fig:slide6\]. ![Representation of the numerical domain used for the simulations of the building with an ideal heat source only.[]{data-label="fig:slide6"}](Slide6.jpg){width="50.00000%"} Only one-quarter of the domain was simulated due to the symmetry of the situation, allowing to reduce the computational times necessary for the simulations. Different effective areas for openings located on the ceiling and the floor were studied as well as different source powers. These different parameters are presented in the Table \[tab:Kaye2009\]. --- --------- --------- -------- -------------------- -------- $A^*/H$ P \[W\] $\mu$ $\sqrt{t_d / t_f}$ 1 $\quad$ 0.0013 3.20 38.2 27’463 2 0.0013 320.00 38.2 5’917 3 0.0053 0.08 9.5 46’962 4 0.0104 3.20 4.9 9’825 5 0.0319 3.20 1.6 5’612 6 0.0628 3.20 0.8 4’003 7 0.0840 80 0.6 1’834 --- --------- --------- -------- -------------------- -------- : Different parameters used for the simulations with openings on the ceiling/floor and no wind[]{data-label="tab:Kaye2009"} ### Mesh topology {#sec:ideal-no-wind-mesh} Regarding the squareness of the domain, a mesh constituted of hexahedron was generated. It is the simplest and fastest way to mesh such a domain. For information purposes, the mesh used is pictured in Fig. \[fig:building\_mesh\] below even though it is very straightforward to imagine what such mesh would look like (with symmetry). ![Illustration of the mesh used for the simulation of the building with an ideal heat source only.[]{data-label="fig:building_mesh"}](building_mesh.JPG){width="80.00000%"} In order to optimize the number of cells used for the simulations by minimizing the computational times without loosing in accuracy and robustness, a convergence study was carried out. It was realized by comparing the interface height for different meshes for a normalized effective area of $A^*/H^2$ = 0.084 and a heat source of $P$ = 80 W. The mesh was considered robust when the interface height did not change, regardless of the reduced number of cells. A mesh with $\sim$ 90’000 cells proved to give a robust and accurate solution and was therefore kept for the whole set of simulations. ### Operating conditions The operating pressure is set to $P_0$ = 101325 Pa which is the default pressure in Fluent. Gravity is activated and the reference density and temperatures are set to $\rho_0$ = 1.225 kg/m$^3$, respectively $T_0$ = 293 K. ### Boundary conditions {#sec:bc-ideal-no-wind} The boundary conditions were defined as follow: - ***bottom opening:*** pressure inlet with 0 Pa gauge pressure, backflow temperature of $T_0$ and with very low level of turbulence ($k$ = 10$^{-5}$ and $\epsilon$ = 10$^{-6}$) - ***top opening:*** pressure outlet with 0 Pa gauge pressure, backflow temperature $T_0$ and very low level of turbulence ($k$ = 10$^{-5}$ and $\epsilon$ = 10$^{-6}$) - ***heat source:*** the heat source is modeled as a heating plate producing a heat flux corresponding to the heat outputs $P$ shown in Table \[tab:Kaye2009\]. - ***building walls, ceiling and floor:*** the walls and the floor are considered adiabatic with a standard no slip condition. ### Discretization scheme All the discretization schemes (Pressure, momentum, turbulent kinetic energy, turbulent dissipation rate and energy) were chosen of the second order. A SIMPLE scheme was used for the pressure-velocity coupling. ### Wall treatment Two approaches are used for the modeling of boundary layers: near-wall treatment and wall functions. By nature, these two approaches require different types and sizes of cells. The first one consists of having a mesh that is fine enough in the boundary layer region to accurately model the physics happening there with the chosen turbulence model, as it is shown on the left of Fig. \[fig:slide7\]. On the other hand, the wall functions approach “bridges” the near wall region with experimentally validated function that reproduce the boundary layer and the associated logarithmic profile and heat transfer phenomena (right of Fig. \[fig:slide7\]). The advantage of the second method is that it allows to obtain robust and accurate results without requiring a very fine mesh and was therefore chosen in this case. ![Representation of the two main techniques to model boundary layers regions: near-wall treatment (left) and wall functions (right).[]{data-label="fig:slide7"}](Slide7.jpg){width="\textwidth"} ### Time-stepping The management of the time-steps was inspired by the simulations carried out by @Kaye2009. Smaller time-steps with a large number of iterations were used during the early stages of the flow when the flow field is establishing and highly variable gradients exist. Moreover once the flow field is established, larger time-steps were used with larger number of iterations, accordingly to the procedure presented by @Kaye2009: 1 second for the first 10 time steps, 5 seconds for the time steps 11-20, 10 seconds for time steps 21-30, followed by 42 time steps at 20 seconds each. After those initial 1000s fixed time steps of 50 seconds with 50 maximum iterations were used to achieve the final steady state. ### Results **Velocity Field**\ The velocity field inside the building is shown in Fig. \[fig:standard-openings-velocity-field\] on a plane parallel to the wall at the center of the plume. A typical plume profile can be observed with high vertical speeds in the center and low speed regions with recirculation on the edges. The radius of the plume is increased with the height due the entrainment of more air with buoyancy forces. The bottom openings are providing air to the system and it can clearly be seen in the velocity field by the two “bumps” on the sides. This field is similar to what was obtained (and validated) by @Kaye2009 and is therefore assumed correct. ![Velocity field inside the simulated building for a fixed effective area of $A^*/H^2$ = 0.003 and a heat source power of 80W. The side view (left) is plotted on a plane parallel to the wall at the center. The top view (right) is plotted at the height of the density interface. Symmetry is applied for both images.[]{data-label="fig:standard-openings-velocity-field"}](standard-openings-velocity-field.jpg){width="\textwidth"} Looking at the right part of Fig. \[fig:standard-openings-velocity-field\] it is interesting to note that the flow is almost at rest in the part of the building away from the center of the plume. The turbulent flow is confined in the plume region. The main reason for this is that the top openings of the building are located nearby the center of the room and the turbulent flow created by the plume can quickly escape the building without disturbing the rest of the flow.\ **Temperature field**\ The temperature field inside the domain is plotted in Fig. \[fig:standard-openings-temp-field\] on a plane parallel to the wall at the center of the room for two different heat source powers. The plume is once again appearing clearly in the center of the room, with its radius increasing with height. It can be seen that the bottom openings are providing cold air to the interior space. The density interface is observed in the middle of the domain. ![Temperature field inside the simulated building for a fixed effective area of $A^*/H^2$ = 0.003 plotted on a plane parallel to the wall at 0.m off the center. The left part corresponds to a heat source power of 80W whereas the right one corresponds to the temperature field for a heat source of 2000W.[]{data-label="fig:standard-openings-temp-field"}](standard-openings-temp-field.jpg){width="\textwidth"} It is interesting to note that the density interface height is the same, for both cases the height of the interface is only depending on the effective area $A^*$. Those results agree with the model presented in theory and equ. (\[equ:no\_wind\]). The main difference is seen in the transient behavior of the flow where stronger heat sources converge more quickly to a warmer steady state. The second difference is that the interface is more diffuse in the case of strong heat sources because of the stronger temperature gradients in the room and the overall higher degree of turbulence. The detailed impact of different heat source power was not studied in detail due the limited allocated time, but the results presented above show that the thermal effects of the flow should also be modeled correctly.\ **Interface height**\ The interface height simulated with Fluent was compared to the theoretical (validated) value presented in section \[sec:ideal-no-wind-theo\]. The evaluation of the numerical interface height is explained in \[app:measurements\]. The results are displayed in Fig. \[fig:xi\_vs\_AH2\_cfd\]. It shows a good agreement with the theoretical model over the whole range of simulated values. Different heat source power levels were tested but they did not change the final steady state value, only the transient behavior, i.e. higher heat source powers only gave faster convergence to the final steady state. These results agree with the numerical simulations produced by @Kaye2009 and show that the mesh, turbulence model, discretization schemes and different parameters used for the simulations are able to predict the behavior of natural ventilation flows.\ ![Variation of the non-dimensional interface height with respect the normalized effective area for an ideal heat source.[]{data-label="fig:xi_vs_AH2_cfd"}](xi_vs_AH2_cfd-eps-converted-to.pdf){width="80.00000%"} **Transient behavior**\ The transient behavior of the numerical flow was compared with the theoretical model presented in section \[sec:ideal-no-wind-theo\] since these results had also a validation. This was done to verify that the numerical parameters used were performing well enough in reproducing the development of the flow and not only its final state. The transient behavior of the interface height was plotted for case 3 (see Table \[tab:Kaye2009\]) and is represented in Fig. \[fig:xi\_vs\_tau\_cfd\]. The behavior seems to be correctly modeled even though the value is slightly under-predicted before the steady state is reached. ![Evolution of the non-dimensional interface height with respect to the non-dimensional time $\tau$ for a normalized effective area of $A^*/H^2$ = 0.005 and a heat source power of 80mW.[]{data-label="fig:xi_vs_tau_cfd"}](xi_vs_tau_cfd-eps-converted-to.pdf){width="80.00000%"} Simulation of the emptying-filling box with lateral openings ------------------------------------------------------------ ### Description of the problem This first set of simulations was carried out with the setup mentioned above since it was the closest (experimentally validated) study to the final objective of this project found in the literature. Since the further simulations where studying the influence of the external atmosphere and to account for the effect of the wind, it was not possible to put the openings of the building on the ceiling, respectively on the floor, but rather on the sides of the building (see Fig. \[fig:building\_domain\]). Therefore, the simulations where re-run with the same mesh, turbulence model and parameters inputs, but solely changing the locations of the openings to the sides of the building. To asses this effect six cases were studied. The parameters for the different cases are presented in Table \[tab:lat-openings\]. ![Representation of the numerical domain used for the simulation taking in account the influence of the external atmosphere.[]{data-label="fig:building_domain"}](building_domain.JPG){width="65.00000%"} Case $A_b$ \[m$^2$\] $A_t$ \[m$^2$\] $A^*/H^2$ ------ ----------------- ----------------- ----------- 1 0.04 0.04 0.003 2 0.09 0.09 0.006 3 0.16 0.16 0.011 4 0.25 0.49 0.021 5 0.49 0.49 0.033 : Different parameters used for the simulations with lateral openings and no wind[]{data-label="tab:lat-openings"} ### Results **Velocity Field**\ The streamlines starting from the bottom opening of the building as well as the velocity field at the interface height are plotted in Fig. \[fig:lat-openings-velocity-field\]. The flow exhibits much more turbulence than the one presented in Fig. \[fig:standard-openings-velocity-field\]. It has gained one more degree of freedom in the sense that when the openings were located on the floor/ceiling the air exchange with the external atmosphere was in the same direction with the plume whereas now air comes laterally. This induces a high vorticity in the flow that can be observed on the right part of Fig. \[fig:lat-openings-velocity-field\]. ![Streamlines starting from the bottom opening of the building (left) and the velocity field on the density interface (right) for an effective area of $A^*/H^2$ = 0.021 and an ideal heat source of 200W.[]{data-label="fig:lat-openings-velocity-field"}](lat-openings-velocity-field.jpg){width="\textwidth"} Despite these phenomena, stratification still takes place and a buoyant layer is created. The interface is however less well-defined than for openings situated on the floor/ceiling, due to the higher levels of turbulence. Furthermore, the vortex is not steady and even though the plume keeps its conic shape it precesses around the center of the heat source. All these reasons explain why the interface height simulated is under-predicted compared to the theoretical model. A way of correcting this will be discussed in the following paragraphs. **Temperature Field**\ The temperature fields inside the domain are plotted in Fig. \[fig:lat-openings-temp-field\] on a plane parallel to the wall at 0.3 m and 1.5 m off the center of the room for two different heat source powers. Once again the density layer is clearly visible. ![Temperature field inside the simulated building for a fixed effective area of $A^*/H^2$ = 0.021 plotted on a plane parallel to the wall at 0.3m off the center (left), respectively 1.5m (right).[]{data-label="fig:lat-openings-temp-field"}](lat-openings-temp-field.jpg){width="\textwidth"} Similarly to what was said for the velocity field in the previous chapter, the interface is not as defined as for openings situated on the floor/ceiling of the building. Moreover, the more turbulent flow induces a mechanical diffusion of the heat through the interface. The precession of the plume can be seen on the left the part of Fig. \[fig:lat-openings-temp-field\] as the plume is not aligned with the vertical direction but slightly leans to the right.\ **Interface height**\ Fig. \[fig:xi\_vs\_AH2\_cfd\_lat\] shows the results for this case, compared to the ones already presented before. It appears that the interface height is lowered for openings located on the sides rather than on the ceiling and bottom of the building respectively. This phenomenon can be explained by the influence of the openings on the total area of the walls that will represent a non-negligible percentage of the total height of the building.\ ![Variation of the non-dimensional interface height with respect the normalized effective area for an ideal heat source and lateral openings.[]{data-label="fig:xi_vs_AH2_cfd_lat"}](xi_vs_AH2_cfd_lat-eps-converted-to.pdf){width="80.00000%"} The numerical model can be corrected by normalizing the interface height by the corrected height of the building $H'$. The corrected height is taken as the height between the middle of the openings as shown in Fig. \[fig:h\_hprime\]. The corrected results are represented in red in Fig. \[fig:xi\_vs\_AH2\_cfd\_lat\] and seem to work fairly well for the whole range of simulations.\ Regarding the analysis presented above, it would have been better to produce a correction model taking into account the vorticity of the flow and the turbulence level inside the room instead of only applying a correction based on the finite vertical extent of the upper and lower openings. These more complicated models are unfortunately time consuming to build and out of the boundaries of this project. ![Illustration of the lateral view of the building. $H'$ is the corrected building height, taken from the middle of the openings to take into account their effect on the flow.[]{data-label="fig:h_hprime"}](Slide8.jpg){width="50.00000%"} Simulation of the external atmosphere {#sec:ideal-wind} ------------------------------------- ### Description of the problem This set of simulations correspond to the theoretical model presented in section \[sec:external-atm-theo\]. Here wind is taken into account by including the external atmosphere in the numerical domain. The simulated external atmosphere was taken as a rectangle of 20m $\times$ 20m $\times$ 50m to avoid the influence of the wall boundary conditions on the internal flow (see Fig. \[fig:full\_domain\]). The building geometry was taken from the previous (validated) numerical simulations. ![Illustration of the numerical domain used for the simulations with an ideal heat source and the effect of wind.[]{data-label="fig:full_domain"}](full_domain.JPG){width="\textwidth"} The simulations were run for a building with a fixed normalized effective area of $A^*/H^2$ = 0.0210. This particular value was chosen deliberately because it was situated in the middle of the simulation range ($\xi \approx$ 0.5). Different wind speeds and their influence were studied. They are presented in Table \[tab:Flynn2009\]. Case 1 2 3 4 5 6 7 8 9 10 11 -------------------- ----- ----- ------ ----- ------ ----- ----- ----- --- ----- ---- Wind speed \[m/s\] 0.1 0.2 0.25 0.3 0.35 0.4 0.5 0.6 1 1.5 2 : Different cases simulated and their associated wind speeds[]{data-label="tab:Flynn2009"} ### Mesh topology {#sec:ideal-wind-mesh} The mesh used for the simulations taking into account the external atmosphere was based on tetrahedron cells. The reasons for this choice are the simplicity of generating such meshes since they provide easier meshing for complex geometries. To stay consistent with the convergence study performed for the building previously, the previous simulations where re-run with a tetrahedron mesh and proved to give the same results, though requiring a higher number of volumes. In this sense, the same mesh was used for the building inside the atmosphere. The tetrahedron-based mesh allowed the meshing of large cells in the simulated atmosphere whereas fine cells where used on the walls and inside the building, where finer flow phenomena occur. In this regard, if the use of this mesh works and succeeds in modeling the behavior of natural ventilation, it could be used in industry where results are needed quickly. Therefore, the mesh used is showed in Fig. \[fig:mesh-ideal-wind\]. The important parameters assessing the quality of the mesh are regrouped in the table below (Table \[tab:mesh-external-atm\]). -------------- --------------- -------------------- ------------------- -- -- -- Nb. elements Max. skewness Min. orth. quality Max. aspect ratio 1’000’000 0.23 0.76 1.86 -------------- --------------- -------------------- ------------------- -- -- -- : Important properties assessing the quality of the mesh[]{data-label="tab:mesh-external-atm"} ### Operating conditions The operating pressure was set to 101325 Pa. For the full domain, since the pressure will directly enter in the validation of the model, the reference pressure for the numerical domain is probed 10 m upstream of the building, at ground level to avoid any thermal or wind influence. Density and temperatures are set to 1.225 kg/m$^3$, respectively 293 K. ### Boundary conditions {#boundary-conditions} The boundary conditions were defined as follow: - ***inlet:*** velocity inlet with 0 Pa gauge pressure, and the speed specified in Table \[tab:Flynn2009\]. Backflow temperature of $T_0$ and with very low level of turbulence ($k$ = 10$^{-5}$ and $\epsilon$ = 10$^{-6}$). - ***outlet:*** pressure outlet with 0 Pa gauge pressure, backflow temperature $T_0$ and very low level of turbulence ($k$ = 10$^{-5}$ and $\epsilon$ = 10$^{-6}$). - ***heat source:*** the heat source is modeled as a heating plate producing a heat flux corresponding to a power $P$ = 320 W. - ***building walls and floor:*** the walls and the floor are considered adiabatic with a standard no slip condition. - ***atmosphere walls:*** modeled as adiabatic walls with slip conditions to avoid the formation of boundary layers and reproduce the external atmosphere as well as possible. - ***ground:*** adiabatic wall with standard no slip condition. ### Discretization schemes All the terms were discretized with a second order scheme except for pressure where a PRESTO! scheme was chosen. The advantage of the latter is that instead of interpolating the pressure at the center of the cells, the pressure is directly computed on the faces. This scheme is more expensive computationally speaking but allows better precision on pressure, which is a key parameter for naturally ventilated flows. ### Wall treatment In the case studied in this project the precise modeling of the boundary layer only makes sense on the inside of the building where the walls are meshed more finely. The external atmosphere does not need to be discretized extensively since this would increase the required computational power to simulate a flow, but not give exploitable results. Therefore scalable wall functions are used. It allows the solver to choose whether a wall function or a near wall treatment is more appropriate. The use of this method is justified by the fact that velocity will vary along the simulations, along with the $Y^+$ value, it is therefore important that the solver gives solutions as close as possible to reality for all the steps of the simulation. ### Time-stepping The consideration of the external atmosphere in the numerical domain and the addition of wind to the simulations sets aside the possibility of using progressive time steps as it was done for the smaller domain (room only). The reason is that the addition of wind induces higher flow speeds inside the domain and for large time steps numerical instabilities would be created in the simulated atmosphere, thus generating errors in the results. Due to this, the time step size is chosen constant according to the equation given in Fluent’s user guide [@user_guide]: $$\label{equ:time_stepping} \Delta t \leq \frac{L/4}{\sqrt{g \beta \Delta T L}} \approx \frac{L}{4 w}$$ with L the characteristic length scale simulated ($L \sim$ 1m). The chosen time step is set to $\Delta t$ = 1 s in the case studied here, so that equ. (\[equ:time\_stepping\]) is still valid for wind speeds up to 1 m/s. The time step has been lowered to $\Delta t$ = 0.5 s for higher wind speeds in order to avoid divergence and/or errors in the numerical results. ### Results {#sec:ideal-wind-results} **Velocity field**\ The different fields in the external atmosphere are not so important in the framework of this study, since the main interests are the physics happening inside the building especially because the useful value can be taken and calculated independently from the outside. Moreover, it would have been too time-consuming to verify the external parameters quantitatively. For these reasons only a qualitative analysis is done here, and the external velocity profile is plotted in Fig. \[fig:external-atmosphere-streamlines\] below. It appears that the profile is representative of a standard velocity profile that one could expect for the flow around a building. The streamlines show a disturbance of the flow above the building where the air is accelerated. On the other hand, a recirculation zone is created on the lee-side of the building. This is the origin of the pressure drop $\Delta p$ that will be used in the further calculations.\ ![Streamlines of the flow in the middle of the domain for a inlet wind speed of 0.3m/s after 6000s.[]{data-label="fig:external-atmosphere-streamlines"}](external-atmosphere-streamlines.jpg){width="80.00000%"} In the case where a whole neighborhood was simulated, the domain would need to be bigger at least on the upper direction to avoid the over-acceleration on the flow above the building due to the squeezing of the area where the air flows. In any event, in the current case, it will not have an impact on the compared results since the pressure is taken independently from the upstream wind speed and the external velocity field.\ **Temperature field**\ The temperature field at steady state inside the building is presented in Fig. \[fig:external-atmosphere-temp-field\] for a stratified case (left) and a mixed case (right). ![Temperature field inside the building for a wind speed of 0.3m/s (left) and a wind speed of 1m/s (right) after 6000s.[]{data-label="fig:external-atmosphere-temp-field"}](external-atmosphere-temp-field.JPG){width="\textwidth"} For the stratified case, the density interface appears clearly on the figure, with a layer of cold air at $T_0$ and a warmer buoyant layer. It can be seen that the interface has a non-zero thickness, which puts in perspective the fact the data points compared with the theory will allow some incertitude [@Kayeflynn2010]. In the mixed case, the air inside the building is at an almost homogeneous temperature. The colder air in the top right corner corresponds to the cold external air entering the building.\ **Interface height**\ The numerical data obtained were added to the curves presented in the theory chapter (Fig.\[fig:xi\_vs\_delta\]) for a normalized effective area of $A/H^2$ = 0.021 and are displayed in Fig. \[fig:xi\_vs\_delta\_cfd\]. Of the first set of simulations, the pressure discretization was made with a standard second order discretization scheme and under-predicted the height of the density interface. Within Fluent’s user guide [@user_guide], it was recommended to use a PRESTO scheme along with the Boussinesq hypothesis. The simulations were re-run and the predictions were better, as it is shown in the figure below. Furthermore, the interface height was corrected using $H'$ instead of $H$ for the building interface height, accordingly to what was done previously. The final results seemed to agree fairly well with the theory. ![Non-dimensional interface height with respect to the non-dimensional parameter $\delta$ for a fixed normalized effective area of $A/H^2$ = 0.021. The different sets of curves are corresponding to different pressure discretization schemes. The red squares correspond to the data obtained with a corrected building height $H'$.[]{data-label="fig:xi_vs_delta_cfd"}](xi_vs_delta_cfd-eps-converted-to.pdf){width="80.00000%"} It can be observed that the correlation between the numerical results and the theoretical model is diverging with the increasing wind speed (increasing $\delta$). This is likely due to some turbulent phenomenon that is not taken into account in the theoretical model. Moreover, it can be seen that the pressure discretization does not influence the case without wind ($\delta$ = 0). The reason is that this case does not take into account the external pressure drop in the calculations and is therefore not affected. The data points not represented in Fig. \[fig:xi\_vs\_delta\] correspond to the mixed regime. Since no density interface is created, it would not make sense to add them on the graph.\ **Ventilation flux**\ The numerical results were added to the analytical curves presented in the theory (Fig. \[fig:QQw\_vs\_Fr\]) for a normalized effective area of $A/H^2$ = 0.021 and are displayed for both pressure discretization schemes (2$^{nd} $ order, PRESTO) in Fig. \[fig:QQw\_vs\_Fr\_cfd\]. The figure shows good agreement for both schemes. It is interesting to note that since the simulations are initialized with a zero speed everywhere, the solution will never fall to the right of the critical Froude number, defined as that minimum value of Fr where the well-mixed solutions first appear. The numerical model is also performing well at predicting the state of the flow (stratified or mixed). ![Variation of the normalized ventilation flux with respect to the Froude number.[]{data-label="fig:QQw_vs_Fr_cfd"}](QQw_vs_Fr_cfd-eps-converted-to.pdf){width="80.00000%"} This validates the numerical model for the simulation of the building and the external atmosphere. Since the ventilation flux is a parameter that is not directly linked to the effective height of the building, it does not need to be corrected with the same procedure than the interface height. Simulation of the blocked ventilation ------------------------------------- ### Description of the problem For this set of simulations, the effect of a non-ideal heat source are studied but in the absence of an external wind or, for that matter, an explicitly modeled external atmosphere. To this end, the numerical domain of the first set of simulations is used since the only difference is that this time the source provides mass to the system. Therefore, the boundary conditions are the same as before, as well as the discretization schemes and the mesh. The different volume source flux and the associated $\zeta_s$ values are summarized in the table below. ------------------- ------ ------ ------ ------ ------ ------ -- Case 1 2 3 4 5 6 $Q_s$ \[m$^3$/s\] 0.04 0.06 0.08 0.10 0.12 0.14 $\zeta_s$ \[-\] 0.41 0.48 0.53 0.58 0.63 0.67 ------------------- ------ ------ ------ ------ ------ ------ -- : Different volume sources simulated and the associated non-dimensional source flux values for a normalized effective area of $A^*/H^2$ = 0.021.[]{data-label="tab:non-ideal-cases"} ### Boundary conditions {#boundary-conditions-1} The boundary conditions are the same as the ones presented in section \[sec:bc-ideal-no-wind\]. The only change is the heat source, which now provides mass to the system: - ***heat source:*** the heat source is modeled as a mass flow inlet providing air at 300 K and for the values presented in Table \[tab:non-ideal-cases\] ### Results {#sec:non-ideal-no-wind-results} Since the qualitative results were exhibiting the same tendencies as the ones presented in the section before (section \[sec:ideal-wind-results\]), they are not detailed here. Only the quantitative results compared with the theoretical model will be discussed.\ **Interface height**\ Fig. \[fig:xi\_vs\_zetas\_cfd\] represents the non-dimensional interface height with respect to the non-dimensional volume heat source. The black line is the theoretical line presented in section \[sec:non-ideal-no-wind\] and the black dots correspond to the different numerical results. ![Variation of the non-dimensional interface height with respect to the non-dimensional volume heat source for a fixed normalized effective area of $A^*/H^2$ = 0.021.[]{data-label="fig:xi_vs_zetas_cfd"}](xi_vs_zetas_cfd-eps-converted-to.pdf){width="80.00000%"} The results are fitting fairly well to the theory. In general, the interface height is slightly under-predicted by the numerics but the tendency is correct. Since the simulations have been run with the top/bottom openings on the ceiling, respectively on the floor of the room, the results do not need to be corrected with $H'$ instead of $H$. Simulation for a non-ideal heat source {#sec:non-ideal-wind} -------------------------------------- ### Description of the problem The simulations carried out in this section are corresponding to the theory depicted in section \[sec:therory-full-non-ideal\]. The numerical domain is the same that was used for the first set of simulations taking into account the influence of the external atmosphere, as well as the mesh (see section \[sec:ideal-wind-mesh\]). The boundary conditions are kept the same except for the heat source that now provides mass as well as buoyancy The discretization schemes are of the second order, except for pressure for which the influence of a PRESTO discretization scheme was studied. Two different non-ideal heat sources ($\zeta_s = \{0.23, 0.58\}$) were tested, as well as nine different wind speeds. They are summarized in Table \[tab:flynn2009-MF\]. Case 1 2 3 4 5 6 7 8 9 -------------------- ----- ----- ------ ----- ------ ----- ----- ----- --- -- Wind speed \[m/s\] 0.1 0.2 0.25 0.3 0.35 0.4 0.5 0.7 1 : Different wind speeds studied for the simulations of the non-ideal heat source and the influence of the external atmosphere.[]{data-label="tab:flynn2009-MF"} ### Results {#results-2} Since the qualitative results such as the temperature and velocity fields are exhibiting the same behavior than what was presented in section \[sec:ideal-wind-results\] they are not detailed here. The discussion will rather focus on the quantitative aspect of the results.\ **Interface height**\ Fig. \[fig:xi\_vs\_delta\_MF\_cfd\] represents the non-dimensional interface height with respect to the non-dimensional parameter $\delta$. The black line corresponds to the analytical curve, the black circles to the numerical results for a second order pressure discretization scheme and the black squares to the numerical results for a PRESTO pressure discretization scheme. ![Variation of the non-dimensional interface height with respect to the non-dimensional parameter $\delta$ for a fixed normalized effective area of $A^*/H^2$ = 0.021.[]{data-label="fig:xi_vs_delta_MF_cfd"}](xi_vs_delta_MF_cfd-eps-converted-to.pdf){width="80.00000%"} The influence of the discretization scheme for pressure is shown. The PRESTO scheme appears to perform better at predicting the interface height. It means that the pressure is a key parameter that influences the location of the buoyant layer in naturally ventilated flows. The upper graph in Fig. \[fig:xi\_vs\_delta\_MF\_cfd\] represents an average case for the volumetric heat source and theory and numerics agree fairly well with theory. On the other hand, for strong heat source volumes fluxes, such as the lower graph in the figure, the interface height is lowered till the upper boundary of the bottom opening due to the forcing of the source. As a results the interface goes down to its physical limit and not lower but because numerically (and physically) the interface height cannot be lowered under the upper boundary of the bottom opening. As the density interface lowers it will reach the bottom opening at one point and the the air from the buoyant layer will escape through that opening as well. Consequently, the second set of simulations cannot serve as a validation of the theory, even though the results seem to be in general agreement.\ **Flow rate**\ Fig. \[fig:qT\_vs\_Fr\_cfd\] represents the normalized flow rate in function of the Froude number, corresponding to the Fig. \[fig:qT\_vs\_Fr\] presented in the theory. ![Variation of the normalized flow rate with respect to the Froude number for a non-ideal heat source and a fixed normalized effective area of $A^*/H^2$ = 0.021.[]{data-label="fig:qT_vs_Fr_cfd"}](qT_vs_Fr_cfd-eps-converted-to.pdf){width="80.00000%"} The pressure discretization scheme does not have an influence on the behavior of the flow rate, both curves follow the same tendency. The numerics are failing to properly model the flow when the state should be blocked. Instead, the same discontinuous behavior that was observed for an ideal source is seen, with a flow rate that jumps from positive to negative values when close to zero. More than that, the tendencies are not fitted correctly.\ Regarding the fact that the lower part of the curve (mixed state) does not follow the good tendency, it would not be appropriate to draw conclusions from this figure. Further studies should focus on fitting the curve before trying to understand why the behavior predicted by the analytical model does not agree with the numerical simulations. In parallel, a experimental campaign should be carried out since the numerical model has not been validated yet. It is therefore not possible to say at this point whether the theory or the numerics (or both) are failing. Simulation of the decoupled heat source --------------------------------------- ### Description of the problem The geometry presented in Fig.\[fig:one\_box\_building\] and the related atmospheric parameters are kept the same as before. Instead of having a single heat source in the center of the room, the source is decoupled. The first source (represented on the left in Fig. \[fig:slide4\]) will provide only heat to the internal atmosphere, whereas the second source (represented on the right in Fig. \[fig:slide4\]) will provide a source volume flux at $T_0$ (air jet). ![Illustration of the idealized considered geometry and the decoupled heat source.[]{data-label="fig:slide4"}](Slide4.JPG){width="\textwidth"} To asses the effects of the decoupling, the influence of the distance $d$ between the two sources is studied. $d$ is the distance of the ideal source on the left of the jet, so that if the sources are inverted and the ideal source is on the right, $d$ becomes negative. To assess the influence of $d$, different wind speeds, are simulated for $d = \{-H, -\frac{H}{2}, \frac{H}{2}, H \}$. The mesh used is the same that was used for the first set of numerical simulations (see section \[sec:ideal-no-wind-mesh\]) as well as the discretization schemes (PRESTO for pressure) and the boundary conditions. ### Results {#results-3} As mentioned no theoretical model nor experiments were made for decoupled sources. The advantage of studying numerically the decoupling of the heat source is that it provides a fast and relatively reliable way of assessing if decoupling has an influence on the flow inside the building.\ **Velocity field**\ Fig. \[fig:decoupledHs-vel-field\] represents the velocity field inside the building for an ideal source of 80 W, an air jet of 0.01 m$^3$/s and a fixed normalized effective area of $A^*/H^2$ = 0.021. The left part of the figure shows the velocity field on a plane parallel to the front wall (wind-side) of the building, at the center and the right part the velocity field plotted on the density interface. ![Velocity field inside the simulated building for a fixed effective area of $A^*/H^2$ = 0.021 and an ideal source of 80 W and a jet of 0.01 m$^3$/s. The two source are separated by a distance of $d$ = $H/2$.[]{data-label="fig:decoupledHs-vel-field"}](decoupledHS-vel-field.jpg){width="\textwidth"} The velocity field clearly shows that the flow is dominated by buoyancy and not by the air jet. This is mainly seen on the top view where high velocities are composing the field around the heat source and not the jet. The turbulence level created by the heat source is also significantly superior to the one coming from the air jet. As a result, the velocity field inside the building for a decoupled source is very similar to that observed in the case of a single source, for both ideal and non-ideal cases.\ **Temperature field**\ Fig. \[fig:decoupledHs-temp-field\] represents the temperature field inside the building for the same parameters presented before on a plane parallel to the front wall at the center of the sources and at 1.5 m off the center. ![Temperature field inside the simulated building for a fixed effective area of $A^*/H^2$ = 0.021 and an ideal source of 80 W and a jet of 0.01m$^3$/s. The two source are separated by a distance of $d$ = $H/2$.[]{data-label="fig:decoupledHs-temp-field"}](decoupledHS-temp-field.jpg){width="\textwidth"} The two sources can be distinguishably seen on the figure above. The fact that the interface height is well defined above the air jet but more heavily distorted above the heat source confirms the observations made with the velocity field. For the parameters simulated, the buoyancy is dominating the flow. Moreover, the air provided by the air jet seems to stay in the lower density layer and is just creating a bump on the interface by pushing upwards, locally.\ **Interface height**\ Fig. \[fig:xi\_vs\_d\] represents the non-dimensional interface height with respect to the distance $d$ for different wind speeds (and therefore different $\delta$). The data point at $d$ = 0 m are corresponding to a “coupled” source, i.e. a source that provides both heat and mass to the system with a corresponding source strength given by $\zeta_s$ = 0.2343. Concretely, these data points were added from the simulations run for a non-ideal heat source with the influence of the external atmosphere (section \[sec:non-ideal-wind\]). To insure the validity of the comparison, an equivalent buoyancy flux was chosen for the source ($B$ = $B_s$). ![Variation of the non-dimensional interface height with respect to the distance $d$ between the sources for different wind speeds.[]{data-label="fig:xi_vs_d"}](xi_vs_d-eps-converted-to.pdf){width="80.00000%"} It appears that the distance $d$ does not influence the density interface height. The differences that can be seen on the figure above are likely due to the turbulent nature of the flow inside the domain and not a relevant discrepancy. To support this argument, it can be seen that the results are more stable for low wind speed, thus lower turbulence levels. Concretely, it means that it is unnecessary to conduct mathematical investigations where the source is decoupled at least while trying to improve the prediction of the interface height.\ The impact of the flow rate was not studied in the results since the previous results for a non-ideal heat source showed that the tendencies were not correctly modeled by the numerics. It is therefore not possible to extend the results and draw conclusions. Discussion and perspectives =========================== Correction model for the non-dimensional density interface height ----------------------------------------------------------------- The first thing that needs to be pointed out is the way the correction model for the density interface height prediction is built. Since the simulations with a non ideal heat source are fitting well the theory without applying a correction to the non-dimensional interface height emphasizes the fact that using a shorter building height to correct the results might not be the optimal way to do it. As it was already mentioned in the presentation of the results, a corrective model taking into account the inner vorticity of the flow and/or the turbulence level of the domain would certainly have led to a better fit of the model. Moreover, the model should also take into account the type of source since it seems to have an impact on the final prediction.\ To avoid the long process involved, an option would be to modify the type of openings used in the simulations. Even though square openings were easier to model and facilitated the comparisons, they are not optimal since, as explained above it might have an impact on the “real” height of the building. Considering different geometries for the openings for instance as presented in the figure below (Fig. \[fig:different\_effective\_area\]) could help with the problem.\ Moreover, the models described above are trying to represent real buildings where the openings are all the thin cracks and small openings that allow a building to “breathe,” not only the build pipes that allow ventilation to happen. By doing so, one must be aware that changing the geometry of the openings will have an impact on the loss coefficient $c_b$ and $c_t$ that will no longer be approximated by the 0.6 value. Many studies have nevertheless been conducted to study their variations and good papers treating this subject can be found (e.g. [@Karava2004; @Chia2009]). Finally to validate those coefficients, simulations could be run with an emptying box, and the discharge coefficient assessed from the emptying time of the room. ![Different geometries of lateral openings for the same normalized effective area.[]{data-label="fig:different_effective_area"}](Slide5.jpg){width="80.00000%"} In this context, one could think that measuring the interface height is maybe not the optimal parameter to assess the correct modeling of the flow. The fact is that using the height of the density interface as a metric to determine the validity of a model comes from the experimental campaigns. In real-life experiments, all the data such as velocity, temperature, pressure fields, that can be relatively easily known via analysis or numerics are extremely hard to measure precisely. As an example, imagine how hard it is to measure precisely the pressure drop between the top and bottom openings of the building without interfering with the flow. In this sense, the measurement of the density interface height can be made relatively quickly (and not too costly) with optical methods. Simulation of the external atmosphere {#simulation-of-the-external-atmosphere} ------------------------------------- The simulation of the external atmosphere was here done directly in the sense that the whole domain was simulated. The main reason for that was to be as close to reality as possible. For instance, putting a constant pressure drop over the building might have led to results with a better agreement with theory but further from reality. In this sense to improve the efficiency of the simulations without going too far from reality, one option could have been to run simulations of a wind flow around a closed building, for the different studied wind speeds and save the velocity (temperature and pressure) field as a User Defined Function. In a second time, another set of simulations would have been run, using the field computed before as inlets, this under the assumption that the pressure field is comparable between the closed and ventilated building. The problem is that it is hard to assess the numerical behavior of the interaction on the boundaries (e.g. when the air inside the building starts escaping the room through the top opening) if an unsteady boundary condition is fixed. For this reason mainly, the simulations were run as is. Decoupling of the heat source ----------------------------- Specifically concerning the decoupling of the source, the relative importance of the heat source vs. the air jet should be studied. The simulations run above were clearly dominated by the heat source and therefore might have behaved differently if the air jet was stronger. Also, the inlets were specified to have very low levels of turbulence, but as it was mentioned a few times, some discrepancies between the results have certainly turbulence as origin, and it might be interesting to investigate in this direction. Conclusion ========== During this project, the feasibility of simulating naturally ventilated flows in buildings was assessed. The goal was not only to try to model these flows with numerical simulations but to show that all the different regimes that take place can be reproduced with computational fluid dynamics. To do so, existing analytical models were regrouped and standardized, particularly through the reorganization and re-writing of the existing equations incorporating the same parameters. Besides, numerical simulations corresponding to the different analytical models were run and then compared. The simulations showed good agreement with the theory for most cases. Since some of the theoretical work was previously validated, the results can be used to validate the numerical solutions and their ability to predict natural ventilation flows. In addition, the influence of a decoupled heat source was studied. The main objective was to assess if a decoupled heat source could improve the prediction of the flow parameters. The results show that at least for the prediction of the density interface height, this was a priori not the most promising way to investigate. With that being said, one must keep in mind that complementary simulations should be run before drawing a final conclusion.\ Measurement of the numerical data {#app:measurements} ================================= Pressure drop over the building and volumetric fluxes through the openings -------------------------------------------------------------------------- The pressure drop over the building must be known to compute the non-dimensional parameter $\delta$. The pressure values are averaged over the top and bottom opening areas respectively. Moreover, since the values are slightly fluctuating due to turbulence, a time-average is taken from the moment the flow is well established inside the domain (the pressure has reached a steady-state value). A similar method was applied for the computation of the volumetric fluxes flowing in and out of the building. ![Evolution of the pressure drop $\Delta p$ over the wind and lee-sides of the building. The dashed line represents the averaged value when steady state has been reached.[]{data-label="fig:deltap_vs_time"}](deltap_vs_time-eps-converted-to.pdf){width="65.00000%"} Interface height ---------------- Four probes are created inside the building, at different horizontal locations where turbulence coming from the openings or the floor source(s) have low influence. The temperature gradient (Fig. \[fig:deltap\_vs\_time\]) is computed over those lines and the maximum value is taken as the location of the density interface. Generally, the maximums have converged to the same value, if this is not the case, the mean height of the four probes is taken as the interface height. From there, the evolution of the interface height can be determined over time (Fig. \[fig:intheight\_vs\_time\]).\ ![Temperature gradient profile over the height of the building (left) and evolution of the density interface height over time (right).[]{data-label="fig:intheight_vs_time"}](intheight_evolution-eps-converted-to.pdf){width="\textwidth"} The right part of the figure shows that during the establishment of the steady state flow, the turbulence level inside the building is high and stabilizes with time.\ Another promising approach would be to numerically create planes parallel to the ground, spaced by a few centimeters and average the temperature on them instead of using the technique mentioned above. Like that, the temperature (or other variables) field would have been reduced to one direction only and the results might be closer to the theory. The main reason why this technique was not used is the huge amount of computational resources that is required to average one variable on many different planes and over long numerical periods of time. Temperature of the thermal layers --------------------------------- As explained before, the density interface height can be established from the temperature gradient profile inside the room. Since the temperature profile can also be established, it is possible to compute the temperature $T_{int}$ of the density interface, as shown on Fig. \[fig:y\_vs\_temp\]. ![Temperature profile of inside the building for the four different probes. The dashed line represents the density interface and the red dot the corresponding temperature.[]{data-label="fig:y_vs_temp"}](y_vs_temp-eps-converted-to.pdf){width="80.00000%"} Once this temperature is know, an iso-surface at a temperature equal to $T_{int}$ is created inside the building as it is shown in Fig. \[fig:isosurf\_w025\]. The upper (buoyant) layer is defined as the volume contained between the iso-surface and the ceiling of the building (Fig. \[fig:isovol\_w025\]). Finally, the temperature is averaged over the upper layer. In the case of a mixed regime, since the interface height does not exist but the air flows chaotically inside the room, the inner temperature is computed as an average over the entire room. To do so, the cells composing the inside of the building are attributed a “+1” value whether the external atmosphere’s cells are defined at “0.” It is then possible to average the temperature over the cells mentioned previously.
{ "pile_set_name": "ArXiv" }
--- abstract: | Temporal or spatial structures are readily extracted from complex data by modal decompositions like POD or DMD. Subspaces of that decompositions serve as reduced order models and define spatial structures in time or temporal structures in space. Convecting phenomena pose a major problem to those decompositions. A structure travelling with a certain group velocity will be perceived as a plethora of modes in time or space respectively. This manifests itself for example in poorly decaying Singular Values when using a POD. The poor decay is very counter-intuitive, since we expect a single structure to be represented by a few modes. The intuition proves to be correct and we show that in a properly chosen reference frame along the characteristic defined by the group velocity, a POD or DMD reduces moving structures to a few modes, as expected. Beyond serving as a reduced model, the resulting entity can be used to define a constant or minimally changing structure in turbulent flows. This can be interpreted as an empirical counterpart to exact coherent structures. We present the method and its application to a the head vortex of a compressible starting jet. author: - | Jörn Sesterhenn and Amir Shahirpour\ Institut für Fluidmechanik und Technische Akustik\ TU Berlin bibliography: - '/home/jls/Documents/Literatur/Literatur.bib' title: A Lagrangian Dynamic Mode Decomposition --- Introduction ============ Three things come together in this article: so called coherent structures, modal decompositions, and model reduction. The principal aim of this exercise is to extract Lagrangian structures from turbulent flows along characteristics having the slope of the group velocities. Several modes travelling along such a characteristic and interacting amongst them, shall be defined as an empirical coherent structure. The constituent three parts shall be discussed briefly. Turbulent flows appear to be stunningly complex, but even if a measurement of some quantity appears to be random, it is the outcome of a repetitive and recursive application of a few simple rules considered to be coherent structures in the flow [@Townsend1976], [@Cantwell1981]. Relying basically on hot wire measurements and simple flow visualisation measurements, it was extremely difficult to come up with an idea of what structures are behind the observed quantities in turbulent flows, and it was not until the advent of PIV and Numerical Simulations, that the idea of a coherent structure replaced the vague “eddy” in literature. Using those new tools, so called “hairpin vortices” [@Adrian2007] and other coherent structures were described. The availability of the strain rate tensor $S_{ij}=\frac{1}{2}\left(\frac{\partial u_i}{\partial x_j}+\frac{\partial u_j}{\partial x_i}\right)$ from numerical simulations, made the analysis and perception of thoses structures acessible [@ChongPerryCantwell1990] and enabled their visualisation. Fabian Waleffe [@Waleffe1998] computed those structures as fixed point solutions travelling in the flow. This route has been successfully followed such that today exact coherent structures can be computed for relatively high Reynolds numbers [@AvilaMellibovskyRolandHof2013]. On the other hand, data driven methods were adapted from other field of science to extract structures from turbulent flows. Proper orthogonal decompositions were the first to be introduced to fluid dynamics by Lumley [@Lumley67]. These patterns however lack dynamics and have problems with convective flows. The first drawback hampers the construction of reduced models of the flow, and in consequence the success was there, but limited. This problem was addressed by the introduction of Dynamical Modes [@SchmidSesterhenn2008]. The difficulties of these methods to describe convective phenomena remained and is what will be discussed and solved in the present paper. The problem {#sec:problem} =========== From an empirical point of view we may define a structure as something in space which appears somehow recognisable elsewhere at a later time. A most strict example would be a solution $u(x,t)= u(x-\lambda t)$ to the convection equation $$\label{eq:convect} \partial_t u + \lambda \partial_x u=0.$$ Even when nonlinearly distorted, damped and dispersed, e.g. for the Korteweg – de Vries – Burgers equation $$\label{eq:KdVB} \partial_t u + u \partial_x u -\nu \partial_x^2 u + \delta \partial_x^3 u =0,$$ it is possible to find an analytic solution fitting the above definition in form of a soliton. A solution developing from a given initial condition will serve as an introductory example below. But even in in more complex situations, for example a boundary layer, the flow exhibits structures which lack an analytic solution, but clearly fit the above definition. They might be found as fixed point solutions of the Navier-Stoles equations. In what follows, concentrate on the general case, where we do not have descriptive equations — yet. Our main example will be the vortex ring of a starting supersonic jet. Imagine a shorter or longer pulse of high pressure from an orfice. If the pulse is short, a laminar vortex ring forms. If the pulse is long, a jet will follow the vortex ring and both, vortex ring and jet will become turbulent, if the Reynolds number is sufficiently high. POD, DMD or other model reduction techniques should then easily reduce flows containing those structures. The structure, being the principal mode, modified to some degree by higher modes. Unfortunately, this is not so. The failure can be demonstated for already for a solution of the KdVB–equation (\[eq:KdVB\]). We have chossen the parameters $\nu=5\times10^{-4}$, $\delta=4\times10^{-5}$ and initial condition $u(x,0)=1+\alpha e^{-(x-x_o)/\beta^2} $ where $\alpha=0.1$ and $\beta=0.03$. The solution at $t=0.5$ is given in figure \[fig:KdVB-Solution\], ![Solution (black) to equation \[eq:KdVB\] for a Gaussian initial condition (blue)[]{data-label="fig:KdVB-Solution"}](./figs/KdVB-Solution){width="100.00000%"} which is distorted, damped and dispersed, but has primarily experienced a shift in $x$. It would still qualify as a an evolving structure in space and time. A POD of the data $$\label{eq:snaps} X=\left[u(x,t_0),u(x,t_1),...u(x,t_{n-1})\right]$$ using an SVD $$\label{eq:pod} X=U\Sigma V^T$$ yields the poorly reducing singular values, depicted in figure \[fig:svd-t\]. ![Singular Values of snapshots of solutions to equation (\[eq:snaps\])[]{data-label="fig:svd-t"}](./figs/SVD-t){width="100.00000%"} The POD-modes being an average over all shifted solutions and necessary distortions therefor sum up to the desired solution. ![First three POD modes of solution of equation (\[eq:snaps\])[]{data-label="fig:POD-Modes"}](./figs/POD-U1 "fig:"){width="100.00000%"} ![First three POD modes of solution of equation (\[eq:snaps\])[]{data-label="fig:POD-Modes"}](./figs/POD-U2 "fig:"){width="49.00000%"} ![First three POD modes of solution of equation (\[eq:snaps\])[]{data-label="fig:POD-Modes"}](./figs/POD-U3 "fig:"){width="49.00000%"} \[fig:pod-Modes\] This is illustrated in figure \[fig:POD-Modes\]. The dominant mode appears to be the swallowed elephant in the larger subfigure. Any real instance in time is made up by subtracting a large number of modes like the two depicted following subfigures. The resulting linear combination will not yield zero in may spatial locations and give rise to substantial spurious structure, where there should be none. A DMD along $t$ suffers from the same problem. A remedy ======== The main problem above comes neither from the non linearity nor the other factors rather than the mere translation. That the flow has a relatively simple structure is easily guessed from the characteristic diagram \[fig:CharacteristicDiagramm\]. In fact, solutions travel relatively unmolested in $\tau=x-\lambda t$-direction. ![Characteristic diagram of the solution[]{data-label="fig:CharacteristicDiagramm"}](./figs/CharacteristicDiagramm){width="100.00000%"} Generally speaking, if a structure travels in time along the direction given by the wave-vector $\kappa$, the modal decomposition is to be sought in a plane normal to that direction in space-time $$\label{eq:Spacetime} x_i=\left\{t,x_1,x_2,x_3\right\}~~~~~~~i=0...3.$$ Given the example above, the ratio of the first to second singular value of eq.(\[eq:snaps\]) are shown in figure \[fig:drop\]. There is a dramatic drop when the proper direction is chosen for the snapshot direction. *The essence of the menthod presented here, is to perform a modal decomposition along that direction and later transformed back into pysical space.* ![Drop of singular values for varying angles (left) and when aligned with the flow direction in space and time (right)[]{data-label="fig:drop"}](./figs/SingDrop "fig:"){width=".49\textwidth"} ![Drop of singular values for varying angles (left) and when aligned with the flow direction in space and time (right)[]{data-label="fig:drop"}](./figs/SingValuesAligned "fig:"){width=".49\textwidth"} After that operation, the first singular vector, shown in figure \[fig:KdVB-Structure\], looks as expected for the structure of the solution of the KdVB-equation (\[eq:KdVB\]). Higher modes drop of fast and minimally change the overall shape of the mode. ![KdVB-Structure[]{data-label="fig:KdVB-Structure"}](./figs/POD-U1-Aligned){width=".9\textwidth"} It should be noted that figure \[fig:KdVB-Structure\] represents the spatio-temporal structure. A back transform to physical space will be necessary to either see the temporal evolution in a given space interval or conversely the spatial changes in a time interval. The development of the mode while travelling can better be analysed using a DMD (rather than a POD) along that direction. To that end we assume a linear relationship of the rotated snapshots $$\label{eq:Rotated Snapshots} X_{0..n}=\{u(\xi,\tau_o),u(\xi,\tau_1)...u(\xi,\tau_n)\}$$ as $$\label{eq:DMD} X'=AX$$ with $X$ and $X'$ being the first and last $n$ snapshots in $X_{0..n}$ and approximate the transition matrix using the projected matrix $$\tilde{A}= U^TAU = U^TX'V\Sigma^{-1},$$ or another standard method to obtain the DMD. *The important step is to take the columns of $X_{0..n}$ normal to the characteristic direction and the rows along it.* ![DMD-Spectrum of the development along the $\tau$ direction []{data-label="fig:DMD-Spectrum"}](./figs/DMD-Spectrum-Aligned){width="90.00000%"} Note that the resulting structures are defined in planes normal to the group velocity of the structure in space-time. That means they have no imediate temporal or spatial interpretation. If, for example, a structure is running right to left, the values at the top of the snapshots correspond to a later time than those at the bottom. Results for the Starting Jet ============================ Now we turn to the example of a starting jet and apply the method explained above to this three dimensional case. The starting point is a shock tube-like initial condition with two different gases at different pressures on both sides of a membrane. The difference to the shock tube is due to the fact that one side is connected to the ambient instead of another part of the tube. We have chosen a pressure ratio $p_1/p_2=3.4$ such that eventually a supersonic jet will develop. The Reynolds number is about $Re=10000$, based on the fully expanded conditions. One crucial parameter besides the pressure ratio is the non-dimensional mass supply of the jet. It can be expressed as the ratio of length to diameter of the pipe $C=L/D$. If $C$ is close to unity, a vortex ring will form. In order to develop a trailing jet, $C$ needs to be larger than approximately $5$. The temporal evolution is shown in figure \[fig:StartingJetEvolution\] as a pseudo-schlieren image in a two dimensional cut through jet and vortex ring. ![Evolution of the starting Jet[]{data-label="fig:StartingJetEvolution"}](./figs/StartingJet1 "fig:"){width="23.00000%"} ![Evolution of the starting Jet[]{data-label="fig:StartingJetEvolution"}](./figs/StartingJet2 "fig:"){width="23.00000%"} ![Evolution of the starting Jet[]{data-label="fig:StartingJetEvolution"}](./figs/StartingJet3 "fig:"){width="23.00000%"} ![Evolution of the starting Jet[]{data-label="fig:StartingJetEvolution"}](./figs/StartingJet4 "fig:"){width="23.00000%"} The first image shows the inital pressure wave and the developing vortex ring at the wall. If enough vorticity is generated, the self–induced velocity if the vortex ring makes it accelerate and travel in flow direction, slightly expanding it’s diameter. The third picture shows the vortex ring and the trailing jet, which is formed if enough mass is supplied. The last image shows the full jet when the mass supply vanishes and the vortex ring has moved away. We wish roughly to identify the flow in three cases. First for a vortex ring, without trailing jet, next for the vortex ring with trailing jet at an early and a late state. The head vortex in a starting jet --------------------------------- One of the dominant features of the starting jet is the vortex ring. It is initially formed at the tube lip and detaches later to first move with constant speed and finally travels with a velocity decaying as the square root of time. It is now our aim to detect this vortex and describe it with a few DMD modes. The can be seen in the characteristic diagram of the flow given as an $x-t$-diagram of the vorticity along the centre line in figure \[fig:CharacteristicDiagrammJet\]. For a first example, we have chosen a small reservoir with $C=1$ such that the vortex ring is not followed by a jet and rather travels alone like a ring produced by skillful smokers. ![Characteristic diagram of the starting jet. The diagram was extracted at the centre-line of the vortex. On the axis, time– and space indices are shown.[]{data-label="fig:CharacteristicDiagrammJet"}](./figs/LamVortHead_CharDiag_onVR){width=".4\textwidth"} We choose for now to determine the optimal direction manually in order not to exhaust the reader by technicalities. Next, a coordinate transform of the $xyzt$-space into the desired direction has to be performed. One can observe that the group velocity changes with time. Thus also a temporal transformation $t'=\frac{at}{b\sqrt{t}+c}$ with suitable coefficients should be employed. This complication is left for later[^1] and for now we rotate the $xyzt$-cube of data in such a way that we have a new $\xi\eta\zeta\tau$-cube and choose a satisfyingly straight part on the characteristic for our analysis. Any rotation in four dimensional space can be represented by two rotations in two suitably chosen planes. In our example, where the main flow is in $xt$-direction, a single rotation in the $xt$-plane suffices. The procedure is performed as a decomposition of the rotation in three shears $q'=S_1S_2S_1q$ which is fast and accurate [@Paeth1990]. ### Spectrum and Modal decay of the Lagrangian DMD {#sec:SpectrumDecay} A DMD using 100 snapshots in $\tau$-direction yields the spectrum to the left of figure \[fig:DMD-Spectrum-Head\]. The decay of the DMD modes is plotted in the right diagram of the same figure, where the projection coefficients of the initial snapshot onto the DMD-eigenmodes is shown. They are normalised by the norm off the full mode. It can be observed that the first few modes represent the snapshot up to a relative remainder of less then $10^{-2}$. ![DMD-Spectrum of the head vortex and decay of modal amplitudes for $t=t_0$[]{data-label="fig:DMD-Spectrum-Head"}](./figs/SpectrumHead "fig:"){width="45.00000%"} ![DMD-Spectrum of the head vortex and decay of modal amplitudes for $t=t_0$[]{data-label="fig:DMD-Spectrum-Head"}](./figs/ModalDecayRing "fig:"){width="45.00000%"} ### Reduced order representation of a vortex ring The first mode of the DMD captures the vortex head ring. Its eigenvalue is close to zero and thus represents the almost invariable vortex head as it is translated downstream. Figure \[fig:VorticityPhysicalReconstruction\] confronts the physical vorticity field with the first two modes. The Lagrangian DMD ($\cal{L}$DMD) extracts the vortex ring as the major mode and filters out, both the co-flow at the wall as well as the small trailing vortex, as they have a different decay rate and consequently constitute other modes. Already two modes render the differences invisible in the given 3D-plot. ![3D picture of the spatial structure at a given time (right) and the spatial representation of the first mode at the same time (left) both plots show the isosurface of absolute vorticity. []{data-label="fig:VorticityPhysicalReconstruction"}](./figs/vorticityIso_PhysSpace_First2Modes_t30){width="96.00000%"} In in a 2D plot in figure \[fig:LamVortHead\_PhysSpace\_First02Modes\_IsoLine\_t30\] the correspondence is also very good. ![2D contour plot of vorticity for modes 1-2 for the Lagrangian DMD The right figure depicts the original flow-field.[]{data-label="fig:LamVortHead_PhysSpace_First02Modes_IsoLine_t30"}](./figs/LamVortHead_PhysSpace_First02Modes_IsoLine_t30){width="90.00000%"} ### Reduced order representation of the full vortex head at an early time Next a starting jet with $C>1$ is presented for an early and a late time. This corresponds to a laminar and a turbulent vortex head. For sake of simplicity we will show 2D cuts of the flow field only in what follows. The reconstruction of the first few modes in spact is shown in figure \[fig:EarlyVortHead\_PhySpc\_LDMD\_t35\]. The figure shows the first two modes as a contour plot of vorticity magnitude. They are almost indistinguishable from the full field, which is depicted in the left plot. ![2D contour plot of vorticity for modes 1-3 for the Lagrangian DMD for time $t=35$. The right figure depicts the original flowfield.[]{data-label="fig:EarlyVortHead_PhySpc_LDMD_t35"}](./figs/EarlyVortHead_PhySpc_first3Modes_LDMD_t35 "fig:"){width="46.00000%"} ![2D contour plot of vorticity for modes 1-3 for the Lagrangian DMD for time $t=35$. The right figure depicts the original flowfield.[]{data-label="fig:EarlyVortHead_PhySpc_LDMD_t35"}](./figs/EarlyVortHead_PhySpc_fullField_t35 "fig:"){width="46.00000%"} These results have to be confronted with the Eulerian DMD results which are given in figure \[fig:EarlyVortHead\_PhySpc\_first3Modes\_DMD\_t35\]. ![2D contour plot of vorticity for modes 1-10 for the Eulerian DMD The right figure depicts the decay rates.[]{data-label="fig:EarlyVortHead_PhySpc_first3Modes_DMD_t35"}](./figs/EarlyVortHead_PhySpc_first10Modes_DMD_t35 "fig:"){width="56.00000%"} ![2D contour plot of vorticity for modes 1-10 for the Eulerian DMD The right figure depicts the decay rates.[]{data-label="fig:EarlyVortHead_PhySpc_first3Modes_DMD_t35"}](./figs/EarlyVortHead_DecayRate_RotVs_NoRot02 "fig:"){width="40.00000%"} It can be observed that the DMD introduces spurious shadows ahead of the vortex ring. They need to be cancelled out by many modes, but even ten modes are not sufficient. This can be anticipated also from the decay rate given in the right of the same figure. The fully turbulent developed vortex head ----------------------------------------- The situation is more complicated but not essentially different for the later turbulent stage of the developed jet. A cut through the full jet is shown in ![Evolution of the starting Jet[]{data-label="fig:LateVortHead_MeanFlow"}](./figs/LateVortHead_MeanFlow){width="100.00000%"} figure \[fig:LateVortHead\_MeanFlow\]. The shear layer of this jet is entrained in the vortex head and renders it unstable. One can still see the main structure of the vortex ring, but the entrained shear layer modes have lead to internal turbulent structures. The characteristic diagram along the jet axis is shown in figure \[fig:CharacteristicDiagramJetCenter\]. One can observe more than one structure travelling with different velocities. By inspection one can distinguish the acoustic wave, the head wave, shock cells and some other events. The vortex head in the characteristic diagram was identified again by inspection. ![Characteristic diagram along the jet axis \[fig:CharacteristicDiagramJetCenter\] ](./figs/CharDiag_cnter_t200-1990){width="60.00000%"} Also here, the vortex head is extracted nicely in a single mode. The fine scale turbulent structure is not part of the head, moreover, it moves with a slower group velocity as it travels backwards with respect to the main head motion. That means the internal structure of the jet head suffers again from the same deficiencies as did the whole structure before. In principle one should extract those structure using their characteristic velocity. But already six modes give a clear resemblance of the full structure. Some of these modes have a positive real part and correspond to break up of the vortex head which is soon to occur. The modes which travel with the same group velocity can as an entity, be defined as an empirical coherent structure. As indicated in section \[sec:problem\], they resemble a travelling structure which changes along its path but remains very well recognisable. ![Reconstruction of the trailing vortex for $t=65$. On the left is the first spatial mode on the top right the first three. At the bottom is a recostruction of six modes along with the full flow field.](./figs/LateVortHead_PhysSpc_firstMode_LDMD_t65 "fig:"){width="46.00000%"} ![Reconstruction of the trailing vortex for $t=65$. On the left is the first spatial mode on the top right the first three. At the bottom is a recostruction of six modes along with the full flow field.](./figs/LateVortHead_PhysSpc_first03Modes_LDMD_t65 "fig:"){width="46.00000%"} ![Reconstruction of the trailing vortex for $t=65$. On the left is the first spatial mode on the top right the first three. At the bottom is a recostruction of six modes along with the full flow field.](./figs/LateVortHead_PhysSpc_first06Modes_LDMD_t65 "fig:"){width="96.00000%"} \[fig:Jet3Modes\] Conclusion {#sec:conclusion} ========== Performing a modal decomposition along the characteristic directions given by the group velocity of a structure allows for the extraction of moving features. To this end a rotation in four dimensional space is necessary. In general this can be done by two rotations but as in our case of a jet the main direction coincides with one of the axis, one rotation was sufficient. In this rotated frame a modal decomposition is performed. The head vortex of a starting was extracted with a few modes only and also its apparent instability due to the entrainment of the shear layer was reflected by the growth rates revealed by the DMD. The physical structure is recovered after rotation back into the physical frame. In this way Lagrangian structures as well as their development along their path can be described. The method can be used for the definition of empirical coherent structures given as a reduced model described by a few modes, as well as constructing reduced modes for control or other purposes. [^1]: Transformation in time can be achieved by choosing the snapshots equidistantly in t’ as defined above. Since we have abundant time-steps from our numerical simulation, this is easily done by choosing the right snapshots.
{ "pile_set_name": "ArXiv" }
--- abstract: | Background : The isobaric yield ratio difference (IBD) method is found to be sensitive to the density difference of neutron-rich nucleus induced reaction around the Fermi energy. Purpose : An investigation is performed to study the IBD results in the transport model. Methods : The antisymmetric molecular dynamics (AMD) model plus the sequential decay model GEMINI are adopted to simulate the 140$A$ MeV $^{58, 64}$Ni + $^{9}$Be reactions. A relative small coalescence radius R$_c =$ 2.5 fm is used for the phase space at $t =$ 500 fm/c to form the hot fragment. Two limitations on the impact parameter ($b1 = 0 - 2$ fm and $b2 = 0 - 9$ fm) are used to study the effect of central collisions in IBD. Results : The isobaric yield ratios (IYRs) for the large–$A$ fragments are found to be suppressed in the symmetric reaction. The IBD results for fragments with neutron-excess $I = $ 0 and 1 are obtained. A small difference is found in the IBDs with the $b1$ and $b2$ limitations in the AMD simulated reactions. The IBD with $b1$ and $b2$ are quite similar in the AMD + GEMINI simulated reactions. Conclusions : The IBDs for the $I =$ 0 and 1 chains are mainly determined by the central collisions, which reflects the nuclear density in the core region of the reaction system. The increasing part of the IBD distribution is found due to the difference between the densities in the peripheral collisions of the reactions. The sequential decay process influences the IBD results. The AMD + GEMINI simulation can better reproduce the experimental IBDs than the AMD simulation. author: - 'C. Y. Qiao' - 'H. L. Wei' - 'C. W. Ma' - 'Y. L. Zhang' - 'S. S. Wang' date: 'June 29, 2015; Received 11 June 2015; Revised 19 June 2015' title: 'Isobaric yield ratio difference between the 140 $A$ MeV $^{58, 64}$Ni + $^{9}$Be reactions studied by antisymmetric molecular dynamics model' --- [GBK]{} [^1] introduction ============ The isobaric yield ratio difference (IBD) method, which is similar to the isoscaling method [@HShanPRL; @MBTsPRL01iso], has been developed to study chemical potentials of neutrons and protons [@IBD13PRC; @IBD13JPG] or the nuclear density [@IBD14Ca; @IBDCa48EPJA; @NST2015IBD] in heavy-ion collisions. Based on the isobaric yield ratio (IYR), the IBD method provides cancellation of both the system dependence parameters [@Huang10; @Ma11PRC06IYR; @Huang10Powerlaw; @Huang10NPA-Mscaling], the free energies of isobars [@IBD13PRC], and the terms contributing to the free energy of fragments [@Huang10; @WadaNst13; @MA12CPL09bsbv; @MaCW12EPJA; @MaCW12CPL06; @MaCW13CPC] in the formula determining the cross sections of fragment. The target dependence of IBD has been studied by investigating the measured fragments in 140$A$ MeV $^{40, 48}$Ca and $^{58, 64}$Ni projectile fragmentation reactions on the $^9$Be and $^{181}$Ta targets [@IBD15Tgt], which have been performed by Mocko *et al* at the National Superconducting Cyclotron Laboratory (NSCL) in Michigan State University [@Mocko06]. In addition, the Shannon information entropy theory is also adopted to explain the IBD method for a better understanding of it [@info1; @info2]. At the same time, IYRs has been used to extract the temperature for fragments in heavy-ion collisions [@Ma12PRCT; @Ma13PRCT; @Ma2013NST; @XQNst15]. In a typical IBD distribution, there is a plateau part plus a changing (increasing or decreasing) part as the function of the mass numbers $A$ of fragments [@IBD13PRC; @IBD13JPG; @IBD14Ca; @IBDCa48EPJA]. The plateau part is explained as denoting the region where the chemical potential (or the density) difference of neutrons and protons changes very little. The modified statistical abrasion-ablation (SAA) model [@MaCW09PRC] was used to study the IBD and the nuclear density difference of neutrons and protons between the calcium-isotope-induced reactions [@IBD14Ca; @IBDCa48EPJA], which suggests that the IBD is sensitive to the nuclear density extracted from the prefragments, but the sensitivity is weakened in the results for the final fragments. Because the SAA calculation cannot reproduce some of the IYR distributions, the IBD results by SAA cannot well explain the measured data [@IBD14Ca]. The SAA model has a simple collision and deexcitation mechanism [@FangPRC00; @Fang07-iso-JPG], which does not include the system evolution. In this article, the antisymmetric molecular dynamics (AMD) model is adopted to simulate the 140$A$ MeV $^{58, 64}$Ni + $^9$Be reactions, and the simulated fragments will be analyzed using the IBD method. The article is organized as follows. The IBD method and the AMD simulations are described briefly in Sec. \[model\]. The IBD results for the simulated reactions are discussed in Sec. \[RAD\] and a summary is presented in Sec. \[summary\]. methods description {#model} =================== IBD probe {#ibdintro} --------- In canonical ensembles theory within the grand canonical limit, the cross section of a fragment is expressed as [@GrandCan; @Tsang07BET] $$\label{yieldGC} \sigma(I, A) = CA^{\tau}exp\{[-F(I, A) + \mu_{n}N + \mu_{p}Z]/T\},$$ where $C$ is a constant; $I\equiv N - Z$ is the neutron-excess; $T$ is temperature; $\mu_n$ ($\mu_p$) is the chemical potential of neutrons (protons), which depends on the the density and temperature of the system; and $F(I, A)$ is the free energy of fragment, which also depends on the temperature. The IYR differing by 2 units in $I$ is defined as [@Huang10], $$\label{ratiodef} R(I + 2, I, A) = \sigma(I + 2, A)/\sigma(I, A).$$ Considering two reactions where the measurement situations are the same (where the temperature of the reactions can be assumed as the same), the IYR difference between the reactions, i.e., IBD, is defined as [@IBD13PRC; @IBD13JPG; @IBD14Ca; @IBDCa48EPJA], $$\begin{aligned} \label{IBDIS} \Delta\mu/T&=\mbox{ln}[R_{2}(I + 2, I, A)] - \mbox{ln}[R_{1}(I + 2, I, A)], \nonumber\\ &=(\Delta\mu_{n21} - \Delta\mu_{p21})/T, \hspace{2.8cm}\nonumber\\ &=[(\mu_{n2} - \mu_{n1}) - (\mu_{p2} - \mu_{p1})]/T, \hspace{1.5cm}\end{aligned}$$ with the indices 1 and 2 denoting the reaction systems. $\Delta\mu/T$ is related to the density difference between neutrons and protons of the reaction systems [@IBD14Ca; @IBDCa48EPJA]. Although the IBD probe is deduced from the canonical ensemble theory, the modified Fisher model can yield the same form of the IBD probe [@Huang10; @ModelFisher3; @MFM1]. AMD simulations {#amdintro} --------------- As one of the most sophisticated transport models, AMD describes the nuclear reaction at the microscopic level of interactions of individual nucleons [@AMD96; @AMD99; @AMD03; @AMD04]. The extended version of AMD (AMD-V) introduces the wave-packet diffusion effect as a new quantum branching process and calculates the wave-packet diffusion effect with the Vlasov equation, which can predict the excitation energies of fragments better than other microscopic models [@AMD99]. For a complete description of the AMD model and the fragment analysis, the readers are referred to the more original references [@AMD96; @AMD99; @AMD03; @AMD04; @AMD04Wada; @Huang10; @Mocko08-AMD]. In this article, the AMD-V version is used to simulate the fragment produced in the 140$A$ MeV $^{58, 64}$Ni + $^9$Be reactions. The standard Gogny (Gogny-g0) interaction [@Gogny-g0] is used to take into account all reaction processes. More than $10^{5}$ events for the 140$A$ MeV $^{58, 64}$Ni + $^9$Be reactions are treated using the AMD model. In the fragment analysis, a coalescence algorithm is adopted with a relative small coalescence radius $R_c =$ 2.5 fm in the phase space at $t =$ 500 fm/c. The primary fragments recognized in the phase space of the AMD simulation are allowed to decay by the sequential decay code GEMINI [@GEMINI]. To study the effect of central and peripheral collisions, two limitations on the impact parameters are adopted in the fragment analysis, i.e., $b1 =$ 0 – 2 fm, and $b2$ = 0 – 9 fm. In the previous work carried out by M. Mocko *et al.*, the AMD simulations were set for an impact parameter range of 0-10 fm and up to the time of 150 fm/c by adopting the Gogny-AS interaction [@Mocko08-AMD]. results and discussion {#RAD} ====================== The isotopic distributions in the 140$A$ MeV $^{58, 64}$Ni + $^9$Be reactions calculated by the SAA, AMD, AMD + GEMINI and EPAX2/EPAX3 models have been compared in our recent work [@Ma15CPL; @MaCW09PRC]. The AMD and AMD + GEMINI simulations can reproduce the cross sections for the symmetric fragments, but overestimate the cross sections of the neutron-rich fragments. In this work, we first compare the results of fragments with the same neutron-excess $I$. \[width=8.cm\][fig1\_Ni58csr.eps]{} The distributions for fragments with $I$ from 0 to 3 in the 140$A$ MeV $^{58}$Ni + $^9$Be reaction are plotted in Fig. \[IBCSNi58\]. The measured fragment distributions for the $I =$ 0 and 1 chains form plateaus. While for the $I > 1$ chains the measured fragment distribution increases with $A$. The calculated distributions for $I > 1$ fragments by AMD with the $b2$ limitations decrease with the increasing $A$ when $A < \sim 15$ and then form plateaus (or increase) with $A$. Compared to the measured fragments, the calculated fragments by AMD with $b1$ limitation show similar distributions. For fragments from $I = $ 0 to 3 chains, the calculated cross sections by AMD + GEMINI with the $b2$ limitation change from underestimating to overestimating the measured ones, with good reproduction of the measured cross sections for the $I = $ 1 chain. Besides, the distributions are similar for the small–$A$ fragments within the $b1$ and $b2$ limitations, indicating that the small–$A$ fragments are mainly produced in the central collisions. For fragments with larger $A$, the cross section for AMD + GEMINI with the $b2$ limitation decreases suddenly with the increasing $A$ because the hot fragments in AMD cannot survive and decay to smaller ones. The sudden change of the cross sections of large–$A$ fragments in the central collisions has also been observed in previous works, which is explained as one phenomena of skin effects. [@MaCW10PRC; @MACW10JPG; @MaCW13CPC; @Ma13finite]. \[width=8.cm\][fig2\_Ni64csr.eps]{} The fragment distributions in the 140$A$ MeV $^{64}$Ni + $^9$Be reaction are plotted in Fig. \[IBCSNi64\]. The plateau phenomena in the fragment distributions in the $^{58}$Ni reaction is weakened in the $^{64}$Ni reaction. The results by AMD + GEMINI with the $b2$ limitation overestimate the measured one for the $I =$ 0 chain, but well reproduce the measured results for the $I =$ 1, 2, and 3 chains. The cross sections by AMD with the $b1$ and $b2$ limitations are similar for most of the fragments except those with $A$ close to the projectile nucleus, indicating that the cross sections for these fragments are slightly influenced by the impact parameters. The cross sections calculated by AMD + GEMINI with the $b1$ and $b2$ limitations are also similar when $A < \sim 30$, which also indicates that they are less influenced by the impact parameters compared to the $A > \sim 30$ fragments. Besides, in both of the $^{58}$Ni and $^{64}$Ni reactions, an obvious odd–even staggering phenomena is shown in the distribution for the $I =$ 0 chain. \[width=8.cm\][fig3\_IYRP.eps]{} In Fig. \[IYR-pre-I01\], the $I = $ 0 and 1 IYRs for the $^{58}$Ni and $^{64}$Ni reactions simulated by the AMD model are plotted. The IYRs with the $b1$ and $b2$ limitations are very similar both for the $^{58}$Ni and $^{64}$Ni reactions in the AMD and AMD + GEMINI simulations, indicating that the IYRs for the $I = $0 and 1 chains are only slightly influenced by the impact parameters. The experimental IYRs increase with $A$, while in the AMD and AMD + GEMINI simulations, the IYR only increases with $A$ when $A < 30$ and it tends to be constant when $A > 30$. Given the obvious underestimation of the $I = $0 chain and the overestimation of the $I =$ 3 chain in the $^{58}$Ni reaction, the calculated IYRs by AMD do not agree with the measured ones very well. The same results are also shown in the AMD results for the $^{64}$Ni reaction. In the measured and calculated results with $b2$ limitation for the $^{58}$Ni reaction, the increasing IYRs are suppressed in the large–$A$ fragments. The obvious even-odd staggering is found in the experimental IYR distributions of $I =$ 0 chain for both the $^{58}$Ni and $^{64}$Ni reactions. \[width=8.cm\][fig4\_IYRF.eps]{} The IYRs for the $I = $ 0 and 1 chains calculated by AMD + GEMINI are plotted in Fig. \[IYR-final-I01\]. For both the $^{58}$Ni and $^{64}$Ni reactions, after the decayed process through GEMINI, the IYRs with the $b1$ and $b2$ limitations overlap for most of the fragments, which means that after the decay, the IYRs for the $I = $0 and 1 chains are only slightly influenced by the impact parameters. The IYRs increase with $A$ of the fragments, which is different than those of the AMD results. The AMD + GEMINI calculated IYRs overestimate the measured results for both the $^{58}$Ni and $^{64}$Ni reactions, respectively. The suppression of IYRs for the large-$A$ fragments in the $^{58}$Ni reaction can be also found in the measured and calculated results with the $b2$ limitation. \[width=8.cm\][fig5\_IBD.eps]{} The IBD results between the 140$A$ MeV $^{58, 64}$Ni + $^{9}$Be reactions are plotted in Fig. \[IBDcomp\]. When $A < 40$ and $A < 53$, the measured IYRs for the $I = $0 and 1 chains form quite good plateaus around $\Delta\mu/T \sim 1.0$, respectively. In the AMD simulated results with the $b1$ and $b2$ limitations, the IBDs tend to be similar, but a slight difference is shown in the distribution (some of the IBD results are not shown due to the absence of cross sections in calculation). In the AMD results with the $b2$ limitation, for the $I = $ 0 chain, in trend the IBD is smaller than that with the $b1$ limitation. This phenomena is more clearly seen in the $I =$ 1 chain. In the AMD + GEMINI simulations with the $b1$ and $b2$ limitations, the IBDs for the $I = $ 0 and 1 chains almost overlap, indicating that for them the central collisions almost determine the IBD results. Thus the IBDs for the $I = $0 and 1 chains reflect the nuclear density in the central collisions. The measured IBD results for the $I =$ 0 and 1 chains are slightly underestimated by the AMD + GEMINI results, and overestimated by the AMD results. It is indicated that the AMD + GEMINI simulations can better reproduce the experimental IBD distributions. In addition, the calculated IBD results with the $b1$ and $b2$ limitations reveal that the increasing part of the IBD distribution is because of the different trends of IYRs for the peripheral reactions of $^{58}$Ni and $^{64}$Ni. An evident staggering is found in the IBDs for the AMD simulated reactions, but disappears in the IBDs for the measured and the AMD + GEMINI simulated reactions. From Fig. \[IYR-pre-I01\], it can be seen that the staggering in the IYRs for AMD simulated reactions for $^{58}$Ni and $^{64}$Ni occurs in a slightly different manners. In the measured IYRs and the IYRs for the AMD + GEMINI simulated reactions for the $^{58}$Ni and $^{64}$Ni reactions, the staggering occurs in a similar manner. The result of $\Delta\mu/T$ has been related to the density difference between neutrons and protons for the reaction systems [@IBDCa48EPJA; @NST2015IBD], which is $\Delta\mu/T = \mbox{ln}(\rho_{n2}/\rho_{p2})-\mbox{ln}(\rho_{n1}/\rho_{p1})$. For the $^{58}$Ni reaction, if $\rho_{n1}/\rho_{p1} = $ 1 can be assumed [@IBD13PRC], $\Delta\mu/T = \mbox{ln}\rho_{n2}-\mbox{ln}\rho_{p2} \equiv \Delta\mbox{ln}\rho_{np}$ for the asymmetric $^{64}$Ni system can be obtained. According to the IBD plateaus for the $I = $0 and 1 chains, in the $^{64}$Ni + $^9$Be reaction, $\Delta\mu/T \approx 1$ will result in $\Delta\mbox{ln}\rho_{np} = 1$ for the measured reaction, and $\Delta\mu/T \approx 0.5$ will result in $\Delta\mbox{ln}\rho_{np} = 0.65$ for the AMD + GEMINI simulated reaction. The $\Delta\mu/T$ for the $p$ + Kr and $p$ + Xe reactions have been estimated to be around 1.16 [@MFM1]. The IBD plateaus for the measured 140$A$ MeV $^{48}$Ca/$^{48}$Ca + $^{9}$Be reactions [@Mocko06] have been estimated to be around 1.85 $\pm$ 0.25 [@IBD13PRC; @info2; @NST2015IBD]. Assuming that for the $^{40}$Ca reaction $\rho_n/\rho_p = $1, for the $^{48}$Ca reaction, one has $\Delta\mbox{ln}\rho_{np} = 1.85$. A relationship between $\Delta\mu/T$ and $\Delta\rho_{np}$ for the calcium reactions has been roughly shown [@IBDCa48EPJA], in which $\Delta\mu/T = $ 1.85, 1, and 0.5 correspond to $\Delta\rho_{np}$ of 0.015, 0.009, and 0.004 fm$^{-3}$, respectively. summary ======= The AMD (+ GEMINI) models have been used to simulate the measured fragments in the 140$A$ MeV $^{58, 64}$Ni + $^{9}$Be reactions. The cross sections of fragments in the simulated reactions are analyzed by adopting two limitations on the impact parameters, i.e., $b1 = 0 - 2 $ fm and $b2 = 0 - 9$ fm, which reflect the central collisions and the whole reaction system. With a difference between the IBD results for the AMD and AMD + GEMINI simulations, they can reproduce the trend of the experimental IBDs for the $I = $ 0 and 1 chains. It is concluded that for the $I =$ 0 and 1 chains, the IBD plateaus are mainly determined by the central collisions, which reflects the density difference in the core of the reaction system. The increasing part of the IBD distribution is revealed to be the difference of neutron and proton densities between the peripheral reactions of the systems since the IYRs in the peripheral reactions of the symmetric system are suppressed. From the IBD results, it is concluded that the AMD + GEMINI simulation can better reproduce the measured ones. This work is supported by the Program for Science and Technology Innovation Talents in Universities of Henan Province (13HASTIT046), China. We are thankful for the helpful guidance from M.-R. Huang for the AMD simulation and the fragment analysis. The AMD simulation was performed on the High-Performance Computing Center at the College of Physics and Electrical Engineering, HNU. H. S. Xu *et al.*, *Isospin Fractionation in Nuclear Multifragmentation*, Phys. Rev. Lett. **85**, 716 (2000). M. B. Tsang, W. A. Friedman, C. K. Gelbke, *et al.*, *Isotopic Scaling in Nuclear Reactions*, Phys. Rev. Lett. **86**, 5023 (2001). C. W. Ma, S. S. Wang, Y. L. Zhang, and H. L. Wei, *Isobaric yield ratio difference in heavy-ion collisions, and comparison to isoscaling* Phys. Rev. C **87**, 034618 (2013). C. W. Ma, S. S. Wang, Y. L. Zhang, and H. L. Wei, *Chemical properties of colliding sources in $^{124,136}$Xe and $^{112,124}$Sn induced collisions in isobaric yield ratio difference and isoscaling methods*, J. Phys. G: Nucl. Part. Phys. **40**, 125106 (2013). C. W. Ma, J. Yu, X. M. Bai, Y. L. Zhang, H. L. Wei, and S. S. Wang, *Isobaric yield ratio difference and neutron density difference in calcium isotopes*, Phys. Rev. C **89**, 057602 (2014). C. W. Ma, X. M. Bai, J. Yu, and H. L. Wei, *Neutron density distributions of neutron-rich nuclei studied with the isobaric yield ratio difference*, Eur. Phys. J. A **50**, 139 (2014). M. Yu, K. J. Duan, S. S. Wang, Y. L. Zhang, and C. W. Ma, *A nuclear density probe: isobaric yield ratio difference*, Nucl. Sci. Tech. **26**, S20503 (2015). M. Huang, Z. Chen, S. Kowalski, [*et al.*]{}, *Isobaric yield ratios and the symmetry energy in heavy-ion reactions near the Fermi energy*, Phys. Rev. C **81**, 044620 (2010). C. W. Ma, F. Wang, Y. G. Ma, and C. Jin, *Isobaric yield ratios in heavy-ion reactions, and symmetry energy of neutron-rich nuclei at intermediate energies*, Phys. Rev. C **83**, 064620 (2011). M. Huang, R. Wada, Z. Chen, *et al.*, *Power law behavior of the isotope yield distributions in the multifragmentation regime of heavy ion reactions*, Phys. Rev. C **82**, 054602(R) (2010). M. Huang, Z. Chen, S. Kowalski, [*et al.*]{}, *A novel approach to isoscaling: The role of the order parameter*, Nucl. Phys. A **847**, 233 (2010). C. W. Ma, J. B. Yang, M. Yu, [*et al.*]{}, *Surface and Volume Symmetry Energy Coefficients of a Neutron-Rich Nucleus*, Chin. Phys. Lett. **29**, 092101 (2012). C.-W. Ma, J. Pu, H.-L. Wei, [*et al.*]{}, *Symmetry energy extracted from fragments in relativistic energy heavy-ion collisions induced by $^{124,136}$Xe*, Eur. Phys. J. A **48**, 78 (2012). C.-W. Ma, J. Pu, S.-S. Wang, and H.-L. Wei, *The Symmetry Energy from the Neutron-Rich Nucleus Produced in the Intermediate-Energy $^{40,48}$Ca and $^{58,64}$Ni Projectile Fragmentation*, Chin. Phys. Lett. **29**, 062101 (2012). C.-W. Ma, H.-L. Song, J. Pu, [*et al.*]{}, *Symmetry energy from neutron-rich fragments in heavy-ion collisions, and its dependence on incident energy, and impact parameters*, Chin. Phys. C **37**, 024101 (2013). R. Wada, M. Huang, W. Lin, *et al*, *IMF production and symmetry energy in heavy ion collisions near Fermi energy*, Nucl. Sci. Tech. **24**, 050501 (2013). C.-W. Ma, Y.-L. Zhang, C.-Y. Qiao, S.-S. Wang, *Target effects in isobaric yield ratio differences between projectile fragmentation reactions*, *Phys. Rev. C* **91**, 014615 (2015). M. Mocko, M. B. Tsang, L. Andronenko, [*et al.*]{}, *Projectile fragmentation of $^{40}$Ca, $^{48}$Ca, $^{58}$Ni, and $^{64}$Ni at 140 MeV$/$nucleon*, Phys. Rev. C **74**, 054612 (2006). C. W. Ma, and H. L. Wei, *Isotopic Ratio, Isotonic Ratio, Isobaric Ratio and Shannon Information Uncertainty*, Commun. Theor. Phys. **62**, 717 (2014). C. W. Ma, H. L. Wei, S. S. Wang, Y.G. Ma, R. Wada, and Y. L. Zhang, *Isobaric yield ratio difference and Shannon information entropy*, Phys. Lett. B **742**, 19 (2015). C. W. Ma, J. Pu, Y. G. Ma, R. Wada, and S. S. Wang, *Temperature determined by isobaric yield ratios in heavy-ion collisions*, Phys. Rev. C **86**, 054611 (2012). C. W. Ma, X. L. Zhao, J. Pu, [*et al.*]{}, *Temperature determined by isobaric yield ratios in a grand-canonical ensemble theory*, Phys. Rev. C **88**,014609 (2013). C. W. Ma, C. Y. Qiao, S. S. Wang, *et al*., *Temperature and symmetry energy of neutron-rich fragments in the 1A GeV $^{124, 136}$Xe + Pb reactions*, Nucl. Sci. Tech. **24**, 050510 (2013). X. Q. Liu, M. R. Huang, R. Wada, *et al.*, *Symmetry energy extraction from primary fragments in intermediate heavy-ion collisions*, Nucl. Sci. Tech. **25**, S20508 (2015). C. W. Ma, H. L. Wei, J. Y. Wang, [*et al.*]{}, *Isospin dependence of projectile-like fragment production at intermediate energies*, Phys. Rev. C **79**, 034606 (2009). D. Q. Fang, W. Q. Shen, J. Feng, [*et al.*]{}, *Isospin effect of fragmentation reactions induced by intermediate energy heavy ions and its disappearance*, Phys. Rev. C **61**, 044610 (2000). D. Q. Fang, Y. G. Ma, C. Zhong, [*et al.*]{}, *Systematic study of isoscaling behavior in projectile fragmentation by the statistical abrasion-ablation model*, J. Phys. G: Nucl. Part. Phys. **34**, 2173 (2007). C. B. Das, S. Das Gupta, X. D. Liu, M. B. Tsang, *Comparison of canonical and grand canonical models for selected multifragmentation data*, Phys. Rev. C **64**, 044608 (2001). M. B. Tsang, W. G. Lynch, W. A. Friedman, *et al.*, *Fragmentation cross sections and binding energies of neutron-rich nuclei*, Phys. Rev. C **76**, 041302(R) (2007). A. S. Hirsch, A. Bujak, J. E. Finn, [*et al.*]{}, *Experimental results from high energy proton-nucleus interactions, critical phenomena, and the thermal liquid drop model of fragment production*, Phys. Rev. C **29**, 508 (1984). R. W. Minich, S. Agarwal, A. Bujak, [*et al.*]{}, *Critical Phenomena in Hardronic Matter and Experimetenal Isotopic Yields in High Energy Proton-nucleus collisions*, Phys. Lett. B **118**, 458 (1982). A. Ono and H. Horiuchi, *Antisymmetrized molecular dynamics of wave packets with stochastic incorporation of the Vlasov equation*, Phys. Rev. C **53**, 2958 (1996). A. Ono, *Antisymmetrized molecular dynamics with quantum branching processes for collisions of heavy nuclei*, Phys. Rev. C **59**, 853 (1999). A. Ono, P. Danielewicz, W. A. Friedman, W. G. Lynch, and M. B. Tsang, *Isospin fractionation and isoscaling in dynamical simulations of nuclear collisions*, Phys. Rev. C **68**, 051601(R) (2003). A. Ono, P. Danielewicz, W. A. Friedman, W. G. Lynch, and M. B. Tsang, *Symmetry energy for fragmentation in dynamical nuclear collisions*, Phys. Rev. C 70, 041604(R) (2004). R. Wada, T. Keutgen, K. Hagel, *et al.*, *Reaction dynamics and multifragmentation in Fermi energy heavy ion reactions*, Phys. Rev. C 69, 044610 (2004). M. Mocko, M. B. Tsang, D. Lacroix, [*et al.*]{}, *Transport model simulations of projectile fragmentation reactions at 140 MeV$/$nucleon*, Phys. Rev. C **78**, 024612 (2008). J. Dechargé, and D. Gogny, *Hartree-Fock-Bogolyubov calculations with the D1 effective interaction on spherical nuclei*, Phys. Rev. C **21**, 1568 (1980). R. J. Charity, M. A. McMahan, G. J. Wozniak *et al.,* *Systematics of complex fragment emission in niobium-induced reactions*, Nucl. Phys. A **483**, 371 (1988). C. W. Ma, Y. L. Zhang, S. S. Wang, and C. Y. Qiao, *A Model Comparison Study of Fragment Production in 140A MeV $^{58,64}$Ni + $^{9}$Be Reactions*, Chin. Phys. Lett. **32**, 072501 (2015). C.-W. Ma, H.-L. Wei, and M. Yu, *Reexamination of the neutron skin thickness using neutron removal cross sections*, Phys. Rev. C [**82**]{}, 057602 (2010). C. W. Ma, H. L. Wei, G. J. Liu, and J. Y. Wang, *Systematic behavior in the isospin dependence of projectile fragmentation of mirror nuclei with A $= 20-60$*, J. Phys. G: Nucl. Part. Phys. **37**, 015104 (2010). C. W. Ma, S. S. Wang, H. L. Wei, and Y. G. Ma, *Re-examination of Finite-Size Effects in Isobaric Yield Ratios Using a Statistical Abrasion-Ablation Model*, Chin. Phys. Lett. **30**, 052101 (2013); C. W. Ma, H. L. Wei, and Y. G. Ma, *Neutron-skin effects in isobaric yield ratios for mirror nuclei in a statistical abrasion-ablation model*, Phys. Rev. C **88**, 044612 (2013). [^1]: Email: [email protected]
{ "pile_set_name": "ArXiv" }
--- author: - 'J. E. Horvath' - 'G. Lugones' date: Received February 2004 title: 'Self-bound CFL stars in binary systems : are they “hidden” among the black hole candidates ?' --- Introduction ============ The strong gravitational field regime of General Relativity is widely believed to be physically realized in black holes, uniquely possessing event horizons. The positive identification of black holes among existing astrophysical objects has the potential of providing a glimpse into that regime which captures the imagination of scientists and public alike. This is why the study of black hole candidates has consumed so much time of research efforts since the early days of X-ray astronomy, when the first strong sources were identified and some of them tentatively associated with astrophysical black holes. Unfortunately, the task of a positive identification is still plagued with caveats, and even though the advances in the last decade or so have been quite impressive ([@Orosz]), the burden of proof is still with observational astrophysicists. Meanwhile, the discussion continues and new proposals arise as potential unique signatures of event horizons possessed by black holes only. In a recent series of papers ([@Nar1; @Nar2] and references therein), Narayan and coworkers have proposed a class of X-ray binaries (the soft-X transients, or SXT) as attractive candidates to black holes. The argument is that those sources show, in addition to a mass function $f(M)$ larger than the expected $\sim \, 3 \, M_{\odot}$ maximum limit for NS models (the mass function is an absolute lower limit to the mass of the compact object in the SXT), an absence of type I bursts tentatively interpreted as evidence for an event horizon. In fact, a systematic study undertaken ([@Nar2]) to pin down the physical condition for type I thermonuclear bursts has shown that SXT sources should burst if they possess a normal matter crust, therefore the absence of bursts could be interpreted as indicating the presence of event horizons. In a recent paper, Yuan, Narayan and Rees (2004) made a general attempt to show that [*any*]{} neutron star (composed by matter described by a more or less general equation of state) should experience thermonuclear Type I bursts at appropriate mass accretion rates. The question is whether an “abnormal” surface also allows such physical behavior. We shall argue below that general models do not necessarily possess the anomalous high-density surface properties of self-bound quark stars, which is one of the key ingredients that entangle their discrimination within the black hole candidates. Since the proof of the existence of event horizons is so important for modern astrophysics, careful examinations of the possible loopholes and alternatives to the signatures are needed. As an example of the latter, a critical analysis by Abramowicz, Kluźniak and Lasota (2003) concluded, based on a series of persuasive arguments, that a positive proof of the event horizon will be impossible in principle. Specifically, they argue that the case of SXTs as black holes based on the absence of type I bursts would be weakened by the finding of any kind of exotic star without a “normal” crust. We have reexamined this objection with the aim of addressing the recently proposed models of self-bound CFL stars and report our findings below. Stability of color flavor locked (CFL) quark matter and stellar sequences ========================================================================= It is widely accepted that at high temperatures and/or high baryon number densities confinement of quarks will be lost. The phase diagram of QCD seems to be very rich and lattice simulations ([@Fodor; @Kanaya]) are attempting to address this point in full detail. However, it is still not very clear whether the transition points are of actual interest in the low temperature/high density regime reached inside existing compact stars. Different compositions of deconfined matter have been analyzed in the literature, and it seems that at high densities a color-flavor locked state of $u, d$ and $s$ quarks is favored over other forms of paired and unpaired matter on general grounds ([@Alford1999; @ALF]). We have previously addressed the issue of CFL matter absolute stability. The idea is an extension of the celebrated “strange matter” case, and in fact, we found ([@LH1]) that pairing [*enhances*]{} the possibility of absolute stability. The CFL phase at zero temperature has been modelled as an electrically neutral and colorless gas of quark Cooper pairs, in which quarks are paired in such a way that all the flavors have the same Fermi momentum, and hence the same number density ([@rajw]). The model allows CFL strange matter to be the true ground state of strong interactions for a wide range of the parameters $B$, $m_s$ and $\Delta$, and this is why we have called this state “CFL strange matter” ([@LH1]). The exploration of the stellar sequences (see also [@LH2; @AR] for a discussion of the physics and stellar models) constructed with the values of the parameter space of the EOS that give absolutely stable CFL strange matter, yielded very compact configurations (which could help explain the recently claimed compactness of a few neutron stars). In addition, for some parameter choices the models are quite massive and extended (see [@LH2]). With the aim of addressing the SXT systems we have explored the parameter space to check whether self-bound CFL stars sequences end with very massive stars. Actually, the basics of $M_{\rm max}$ for any fluid configuration has been discussed by [@Haen] in a pedagogical form. That work highlighted the value of scaling laws for gaining insight of the problem, particularly for the case of linear equations of state that allow simple solutions of the stellar structure. In the case of strange quark matter, if quark-quark interactions are ignored and the mass of the $s$ quark is small enough, the maximum mass of a stellar sequence would be just (Haensel 2003) $$M_{\rm max} = \frac{1.96 M_{\odot}}{\sqrt{B_{60}}}$$ (being $B_{60} \equiv B / [60 \, {\rm MeV fm^{-3}}]$) as already discussed by [@Witten]. We note that the case of a CFL equation of state in fact closely resembles the well-known theory of strange quark matter, having an equation of state of the form $P = \kappa (\rho - 4 B_{\rm eff})$, with $B_{\rm eff} = B - 3 \mu^2 \Delta^2 / \pi^2$ ([@LH1; @LH2]). This parametrization, however, yields a $B_{\rm eff}$ which is dependent on density through the chemical potential $\mu$, and therefore a simple scaling of the maximum mass of the star is not possible as in the case of strange matter. However, and in order to gain a qualitative insight of the structural properties, we can approximate the chemical potential in $B_{\rm eff}$ by a characteristic mean value $\mu_* \sim 200 MeV$ for CFL strange matter. In this simplified case $B_{\rm eff}$ is a constant that depends just on $B$ and $\Delta$ and allows a very simple scaling $$M_{\rm max} \approx \frac{1.96 M_{\odot}}{\sqrt{B_{60}}} \; (1 + \delta)$$ valid for $B > 3 \mu_*^2 \Delta^2 / (2\pi^2)$, where $$\delta = 0.15 \bigg( \frac{\Delta}{100 \; {\rm MeV}} \bigg)^2 \bigg( \frac{60 \, {\rm MeV fm^{-3}}}{B} \bigg).$$ Therefore, and ignoring for the moment the entanglement between the quark masses and the density dependence of the condensation term, we may state that CFL matter can produce high mass compact stars because the existence of pairing reduces the effective value of the vacuum energy density parametrized by $B$ by adding a term with opposite sign. It is this condensation energy $\Delta$ that causes an increase of the maximum mass from $\sim 2 M_{\odot}$ to $\sim 4 M_{\odot}$ for the range considered in the present paper. Figs. 1-3 show the results of the full calculations, which have taken into account all the important ingredients such as quark masses and the correct density dependence of the condensation term. As discussed above, the value of the pairing gap $\Delta$ is the key ingredient of the equation of state. Lacking of a definitive indication of the latter we conclude that very massive models ($M_{\rm max} \,\sim \, 4 \, M_{\odot}$) are possible if $\Delta$ is high enough ($\geq \, 250$ MeV). Only refined calculations will confirm or rule out this range, which is entirely possible according to the present research. Absence of normal matter crusts in self-bound CFL stars ======================================================= A crucial feature for a wide set of observations (largely explored and debated in the context of strange star (SS) models) is the existence of a normal matter crust (see [@Zdu] and references therein). It has been realized that bare quark surfaces may alter drastically the ability of radiating photons (see e.g. [@AFO; @PU; @Xu]). Conversely, a normal matter crust (held in mechanical equilibrium by the electrostatic potential at the surface) may hide most of the features of exotic matter. It has been recently shown that CFL states force an equal number of $u,d$ and $s$ quarks. As a consequence, this phase is electrically neutral and at the same time it does [*not*]{} allow electrons in beta-equilibrium ([@rajw]). Therefore, it is clear that (in striking contrast with the well-studied SS) no crust could be present in the case of CFL strange stars. CFL strange stars must have bare surfaces within these models. Particularly, photon radiation from a bare CFL surface is vanishingly small and the star is dark in optical frequencies (see [@vogt2003] and references therein). As discussed by Narayan (2003), the Type I bursts observed from many X-ray sources require the presence of a normal matter crust. Abramowicz, Kluźniak and Lasota (2003) correctly (and boldly) stated that “no nuclei, no burst”. It is worth noting that, strictly speaking, the presence of a material surface only implies that energy can be radiated once matter lands on the surface. However, this evidence of the presence of a material surface can be seriously obscured if the energy is emitted in the form of weakly interacting particles. This is the case of the bare surface of a CFL star in which almost all energy is expected to be radiated in neutrinos. Thus, the absence of a crust is an important feature of self-bound CFL models, because normal matter would never accumulate at the surface (as discussed in the strange matter case). Instead, when matter is accreted onto a bare electrically neutral CFL surface, there will be in addition to the usual gravitational energy release $\sim GM{\dot M}/R$, an additional term related to the exothermic fusion of the incoming proton with the quark liquid. The latter is approximately given by $\sim {\dot M} \Delta/m_{p}$, and for typical values it may amount to $\sim \, 30 \%$ of increase over the normal case. As matter falls on the bare surface it is immediately converted to the CFL phase and the gravitational + pairing gap energy is (almost) instantaneously emitted mainly in $\nu$’s (due to the unavoidable $\beta$-decays in the quark phase) and a smaller fraction of (most probably) $\gamma$-rays. Because matter is being continuously converted at the surface, and cannot accumulate, sudden X-ray flashes are not expected. Discussion ========== The photon emission properties of CFL strange stars are expected to resemble those of bare SS ([@ISPE]). Since pairing effects should appear in the plasma frequency $\omega_p$ through the baryon number density as a correction of order $\mu \Delta^2$, the plasma frequency $\omega_p$ will not be very different from (typical) $20$ MeV of unpaired quark matter, and thus the equilibrium photon radiation will be suppressed. A tiny luminosity makes CFL strange stars very difficult to detect directly. The thermal emission of photons from the bare quark surface of an (unpaired) strange star (due mainly by electron-positron pair production) shown to be much higher than the Eddington limit by Page & Usov (2002) may not operate for CFL strange stars since no electrons will be present at the surface and the emissivities due to other processes are negligible at low temperatures. These features (high maximum masses, absence of normal matter crusts and lack of surface emission) show that “exotic” stellar models may be constructed in which Type I thermonuclear bursts can not occur, but which are not black holes either. If physically realized in nature, some of the SXT systems observed to possess a relatively low mass function (e.g. SXT A0620-00, with $f(M) \geq 3 \, M_{\odot}$; [@McR]) may harbor self-bound CFL stars. In these models, the lack of Type I thermonuclear bursts is not interpreted as a signature of an event horizon, but rather as a consequence of the impossibility of a normal crust with $\rho \leq 10^{6} g cm^{-3}$ where accreted matter could accumulate and eventually ignite. Further work is needed to elaborate or rule out this tentative association, but it is already clear that CFL stars provide a definite counterexample of the event horizon proof as a representative of a class of exotic alternatives not easily discarded. Acknowledgements ================ We acknowledge E. M. de Gouveia Dal Pino for helpful discussions. J.E. Horvath wishes to acknowledge the CNPq Agency (Brazil) for partial financial support. G. Lugones acknowledges the Physics Department of the University of Pisa (Italy), and FAPESP (Brazil). [99]{} Abramowicz M., Kluźniak W. & Lasota J-P. 2003, A&A 396, L31 Alford M. & Rajagopal K. 2002, JHEP 06, 031 Alford M., Rajagopal K. & Wilczek F. 1999, Nucl. Phys. B 537, 443 Alford M. & Reddy S. 2003, Phys.Rev. D 67, 074024 Alcock C., Farhi E. & Olinto A. V. 1986, Astrophys. J. 310, 261 Farhi E. & Jaffe R. L. 1984, Phys. Rev. D 30, 2379 Fodor Z. 2003, Nucl. Phys. A715, 319c Haensel, P. 2003, EAS Publications Series, Volume 7, 2003, Final Stages of Stellar Evolution, Proceedings of the conference held 16-21 September, 2001 in Aussois, France. Edited by C. Motch and J.-M. Hameury, pp.249., 7, 249 Kanaya K. 2003, Nucl. Phys. A71, 233c Lugones G. & Horvath J. E. 2002, Phys. Rev. D 66, 074017 Lugones G. & Horvath J. E. 2003, A&A 403, 173 Mc Clintock J. E. & Remillard R. A. 1986, Astrophys. J. 308, 110 Narayan R., García M. R. & Mc Clintock J. E. 2001, In: Proceedings of the The Ninth Marcel Grossmann Meeting, Eds.: Vahe G. Gurzadyan, Robert T. Jantzen, Remo Ruffini (World Scientific Publishing) astro-ph/0107387. Narayan R. 2003, astro-ph/0310692 Orosz J. A. 2002, astro-ph/ 0209041 Page D. & Usov V .V. 2002, Phys. Rev. Lett. 89, 131101 . Rajagopal K. & Wilczek F. 2001, Phys. Rev. Lett. 86, 3492 Shovkovy I. A. & Ellis P. J. 2002, hep-ph/0211049 (2002) Vogt C., Rapp R. & Ouyed R. 2003, hep-ph/0311342 Xu R. X. 2002, Astrophys. J. 570, L65 Yuan Ye-Fei, Narayan R. & Rees M. J. 2004, Astrophys. J. 606, 1112 Witten E. 1984, Phys. Rev. D 30, 272 Zdunik J.L. 2002, A&A 394, 641
{ "pile_set_name": "ArXiv" }
--- abstract: 'We have developed an analytical formulation to calculate the plasmon dispersion relation for a two-dimensional layer which is encapsulated within a narrow spatial gap between two bulk half-space plasmas. This is based on a solution of the inverse dielectric function integral equation within the random-phase approximation (RPA). We take into account the nonlocality of the plasmon dispersion relation for both gapped and gapless graphene as the sandwiched two-dimensional (2D) semiconductor plasma. The associated nonlocal graphene plasmon spectrum coupled to the “sandwich" system is exhibited in density plots, which show a linear mode and a pair of depolarization modes shifted from the bulk plasma frequency.' author: - 'Godfrey Gumbs$^{1,2}$, N. J. M. Horing$^3$, Andrii Iurov$^{ 4}$, and Dipendra Dahal$^{1}$' title: Plasmon Excitations for Encapsulated Graphene --- 0.2in Introduction {#sec1} ============ The properties of high-quality graphene encapsulated between two films such as hexagonal boron-nitride are just beginning to be explored and have now become an active area of research due to recent advances in device fabrication techniques [@encaps1; @encaps2; @encaps3; @encaps4; @encaps5; @encaps6; @encaps7; @encaps8; @encaps9; @encaps10]. Interest in the optical properties of these heterostructures has been focused on their unusual plasmonic behavior including their spatial dispersion and damping. This hybrid system may be employed for tailoring novel metamaterials. Additionally, this fabrication technique provides a clean environment for graphene. This brand new area of nanoscience not only poses challenges for experimentalists, but also for theoreticians seeking to formulate a theory for a model system. Although there already exists a copious literature on graphene on a single substrate [@GG; @ONB; @NJMH; @Pol1; @Pol2; @Pol3; @Pol4; @Pol5], this study shows that encapsulated graphene is vastly different in many ways. The model we use in this paper consists of two identical semi-infinite metallic plasmas with planar boundaries at $z = \pm a/2$. Within the spatial separation between the two bulk conducting plasmas ($|z|<a/2$) is inserted a 2D monolayer graphene sheet at $z=0$, shown schematically in Fig. \[FIG:1\]. The natural first step in our calculations of the plasmon excitation spectrum is to set up and solve the random-phase approximation (RPA) integral equation for the inverse dielectric screening function of this hybrid system. We have solved this equation analytically in position representation for a narrow spatial gap between the bulk half-space plasmas, obtaining a closed-form formula for the inverse dielectric function in terms of the nonlocal polarizability for graphene and the bulk metallic polarizability, for which the latter is well approximated by the hydrodynamical model. Based on this newly derived formula, we have calculated the nonlocal plasmon dispersion relation numerically, considering both gapped and gapless graphene as the two-dimensional (2D) semiconductor plasma. The resulting nonlocal graphene plasmon spectra coupled to the “sandwich" system are exhibited in density plots, which show a linear mode and a pair of depolarization modes shifted from the bulk plasma frequency. Hexagonal boron nitride has been the main substrate material that facilitates graphene based devices to exhibit micrometer-scale ballistic transport. The recent work of Kretinin, et al. [@encaps10] has shown that other atomically flat crystals may also be employed as substrates for making high-quality graphene heterostructures. Alternative substrates for encapsulating graphene include molybdenum and tungsten disulfides which have been found to exhibit consistently high carrier mobilities of about $6\times 10^4$ cm$^2$ V$^{-1}$ s$^{-1}$. On the other hand, when graphene is encapsulated with atomically flat layered oxides such as mica, bismuth strontium calcium copper oxide, or vanadium pentoxide, the result is remarkably low quality graphene with mobilities of about $ 10^3$ cm$^2$ V$^{-1}$ s$^{-1}$. This difference is due mainly to self-cleansing which occurs at interfaces between graphene, hexagonal boron nitride, and transition metal dichalcogenides. In our model calculations, we allow for the possibility that the substrate may affect the energy band structure of graphene by opening a gap in the energy band. We compare the resulting calculated plasmon spectra for encapsulated gapless and gapped graphene. The work of Chuang, et al. [@DD1] is a study of the graphene-like high mobility for p- and n-doped WSe$_2$. Electron transport in the junction between two 2D materials $MoS_2$ and graphene have also been reported recently, showing how spatial confinement can influence physical properties [@DD2]. Mobility, transconductance and carrier inhomogeniety experiments have also been reported in [@DD3; @DD4; @DD5; @DD6] for monolayer and bilayer graphene fabricated on hexagonal BN as well as mica based substrates. The outline of the rest of our paper is as follows. In Sec.  \[sec2\], we give details of our calculation of the inverse dielectric function for a 2D layer sandwiched between two conducting substrates whose separation is very small. We explicitly derive the plasma dispersion equation when the thick substrate layers may be treated in the hydrodynamical model. The 2D RPA ring diagram polarization function for graphene at arbitrary temperature is employed in the dispersion equation. A careful determination of the plasmon spectra for a range of energy gap and carrier doping values is reported in Sec. \[sec3\]. We conclude with a discussion of the highlights of our calculations in Sec. \[sec4\]. ![(Color online) Schematic illustration of a pair of thick conducting plasmas (taken to be semi-infinite in our formulation of the problem) encapsulating a 2D monolayer graphene sheet in a “sandwich" array.[]{data-label="FIG:1"}](fig1){width="35.00000%"} Theoretical Formulation {#sec2} ======================== We consider two identical semi-infinite conductors on either side of a 2D semiconductor layer (Fig. \[FIG:1\]). One of the conductors extends from $z=-a/2$ to $z=-\infty$ while the other conductor has its surface at $z= a/2$ and extends to $z=\infty$. The 2D layer lies in a plane mid-way between the two conductors at $z=0$ in the gap region $|z|<a/2$. The inverse dielectric function satisfies $$\begin{aligned} K(z_1,z_2)&=& K_\infty(z_1,z_2) \nonumber\\ &-& \int_{-\infty}^\infty dz^{\prime} \int_{-\infty}^\infty dz^{\prime\prime} \ K_\infty (z-z^\prime) \left[\alpha_{2D}(z^\prime,z^{\prime\prime})- \alpha_{gap}(z^\prime,z^{\prime\prime}) \right] K(z^{\prime\prime},z_2)\ , \label{eq:1}\end{aligned}$$ where $K_\infty(z_1,z_2)=K_\infty(z_1-z_2)$ is the bulk infinite space symmetric conducting medium inverse dielectric function and $$\alpha_{2D}(z^\prime,z^{\prime\prime})=\frac{2\pi e^2}{\epsilon_s q_\parallel} \Pi^{(0)}_{2D} (q_\parallel,\omega) e^{-q_\parallel |z^\prime|} \delta(z^{\prime\prime}) \equiv \tilde{\alpha}_{2D}(q_\parallel,\omega) \ e^{-q_\parallel |z^\prime|} \delta(z^{\prime\prime}) \ , \label{eq:2}$$ which defines $\tilde{\alpha}_{2D}(q_\parallel,\omega) $ in terms of the 2D ring diagram $\Pi^{(0)}_{2D} (q_\parallel,\omega)$. For a narrow gap between the two identical half-space slabs, i.e., $q_\parallel a \ll 1$, the gap polarizability is given by $$\alpha_{gap}(z^\prime-z^{\prime\prime})= a\delta(z^{\prime\prime}) \alpha_{\infty}(z^\prime-z^{\prime\prime}) = a\delta(z^{\prime\prime}) \int_{-\infty}^\infty \frac{dp_z}{2\pi}\alpha_\infty (p_z,q_\parallel) e^{ip_z ( z^\prime-z^{\prime\prime} ) } \ . \label{eq:3}$$ Employing Eqs.  (\[eq:2\]) and (\[eq:3\]) in Eq.  (\[eq:1\]), we have $K(z_1,z_2)= K_\infty(z_1-z_2)- {\cal F}(z_1;q_\parallel,\omega) K(0,z_2)$, where ${\cal F}(z_1;q_\parallel,\omega)= {\cal F}_{2D}(z_1;q_\parallel,\omega) + {\cal F}_{gap}(z_1;q_\parallel,\omega) $ and $$\begin{aligned} {\cal F}_{2D}(z_1;q_\parallel,\omega)&\equiv & \frac{2\pi e^2}{\epsilon_s q_\parallel} \Pi^{(0)}_{2D} (q_\parallel,\omega) \int_{-\infty}^\infty dz^\prime\ K_\infty (z_1-z^\prime) e^{-q_\parallel |z^\prime|} \ , \nonumber\\ {\cal F}_{gap}(z_1;q_\parallel,\omega)&\equiv & -a\ \int_{-\infty}^\infty dz^\prime\ K_\infty (z_1-z^\prime) \alpha_\infty(z^\prime,0) \ . \label{eq:5-2}\end{aligned}$$ Setting $z_1=0$ in the equation relating $K(0,z_2)$ to $K(z_1,z_2)$ above, we may solve for $K(0,z_2)$ and then obtain $$K(z_1,z_2)= K_\infty(z_1-z_2)- \frac{{\cal F}(z_1;q_\parallel,\omega)} {1+{\cal F}(z_1=0;q_\parallel,\omega)} K_\infty(0,z_2)\ , \label{eq:6}$$ so that the plasma excitation frequencies are determined by solving for the zeros of $$\begin{aligned} {\cal D}(q_\parallel,\omega) &\equiv & 1+{\cal F}(z_1=0;q_\parallel,\omega) \nonumber\\ &=& 1+ \int_{-\infty}^\infty dz^\prime\ K_\infty (0-z^\prime)) \left[ \frac{2\pi e^2}{\epsilon_s q_\parallel} \Pi^{(0)}_{2D} (q_\parallel,\omega) e^{-q_\parallel |z^\prime|} -a\ \alpha_\infty(z^\prime,0)\right] \ . \label{eq:7}\end{aligned}$$ Expressing Eq. (\[eq:5-2\]) in terms of the Fourier transform of $K_\infty (z_1-z^\prime)$, we obtain $$\begin{aligned} {\cal F}_{2D}(z_1;q_\parallel,\omega)&= & \frac{2 e^2}{\epsilon_s} \Pi^{(0)}_{2D} (q_\parallel,\omega) \int_{-\infty}^\infty dp_z\ e^{ip_zz_1} \frac{K_\infty (p_z )}{p_z^2+q_\parallel^2} \ , \nonumber\\ {\cal F}_{gap}(z_1;q_\parallel,\omega)&= & -a\ \int_{-\infty}^\infty \frac{dp_z}{2\pi}\ e^{ip_zz_1} K_\infty (p_z) \alpha_\infty(p_z) \ . \label{eq:5-3}\end{aligned}$$ Using the hydrodynamical model for nonlocality of the conducting substrate [@EGUI; @Kamen], we have $\epsilon(q,\omega)=1-\omega_p^2/(\omega^2-\beta^2q^2)$, where $\omega_p$ is the local bulk metallic plasma frequency, $\beta$ is an adjustable parameter which mabe adjusted to give the correct dispersive shift of the bulk plasma frequency, i.e., $\beta^2=3v_F^{(3D)\ 2}/5$ (where $v_F^{(3D) }$ is the bulk Fermi velocity of the conducting substrate). This approximation leads to the dispersion equation which, after performing the $p_z$ integrations, we obtain $$\begin{aligned} {\cal D}(q_\parallel,\omega)&=&1+ \frac{2 \pi e^2}{\epsilon_s q_\parallel} \Pi^{(0)}_{2D} (q_\parallel,\omega) \frac{1}{|A(q_\parallel,\omega)|} \left( \frac{\omega_p^2 q_\parallel -\omega^2 |A(q_\parallel,\omega)|}{\omega_p^2-\omega^2} \right) \nonumber\\ &-& \frac{ a \omega_p^2}{2\beta^2 |A(q_\parallel,\omega)|}\ , \label{xxx}\end{aligned}$$ where $A(q_\parallel,\omega) \equiv \beta ^{-1} \left( \omega_p^2+\beta^2 q_\parallel^2 -\omega^2\right)^{1/2} $. The 2D RPA ring diagram polarization function for graphene with a gap $\Delta$ may be expressed as $$\begin{aligned} \label{A1} \Pi^{(0)}_{2D}(q,\omega) && = \frac{g}{4 \pi^2} \int d^2 {\bf k} \sum\limits_{s,s' = \pm} \left( 1 + s s' \frac{{\bf k} \cdot ({\bf k}+{\bf q}_{\parallel}) + \Delta^2}{\epsilon_k \,\, \epsilon_{\vert {\bf k}+{ \bf q}_{\parallel} \vert }} \right) \nonumber \\ && \times \frac{f(s \, \epsilon_{{\bf k}}) - f(s' \epsilon_{{\bf k}+{\bf q}_{\parallel}})}{s \, \epsilon_{{\bf k}_{\parallel}} - s' \epsilon_{{\bf k}+{\bf q}_{\parallel}} - \hbar \omega - i \eta^+ } \ ,\end{aligned}$$ where $s, s^\prime$ are band indices, $g=4$ accounts for both spin and valley degeneracies and $\epsilon_{{\bf k}}$ is the band energy. Since we limit our considerations to zero temperature, $T=0$, the Fermi-Dirac distribution function is reduced to the Heaviside step function $f(\epsilon, \mu; T \rightarrow 0) = \eta_{+}(\mu - \epsilon)$, and we use the results of Refs. \[\]. Numerical Results and Discussion {#sec3} ================================ In the figures, we present calculated results for the nonlocal plasmon excitations of encapsulated gapless and gapped graphene. As we demonstrated in previous work [@GG; @ONB], the hybrid plasmon modes and their damping are mainly determined by the doping concentration, i.e., the chemical potential $\mu$, along with the energy bandgap $\Delta$. Consequently, we have paid particular attention in our numerical investigations to the various regimes of the ratio $\Delta/\mu$. In Fig. \[FIG:2\], we exhibit results for gapless graphene. Figures \[FIG:3\] and \[FIG:4\] illustrate the case in which the graphene layer has an intermediate or large energy gap. The regions of strong damping arise from inter-band transitions which are forbidden by Pauli blocking (i.e., when the final transition states are filled, so that further transitions cannot occur). In the case of a [*single*]{} substrate and a 2D layer, the lowest acoustic branch vanishes due to damping in the long wavelength limit for a range of separations between the 2D layer and the surface. However, this is not the present case under consideration involving encapsulation of the graphene sheet within a small gap between the two surfaces. Within the “small gap" framework, $a<1/q_\parallel$, the lower branch is never damped in the long wavelength limit. This result is quite different from that obtained for a [*single*]{} substrate, in which strong damping of the acoustic branch appears in the long wavelength limit for the distance $a$ below a specific critical value. In regard to the two upper branches, attributed to bulk three-dimensional plasmons, their undamped parts appear as upper and lower arcs of a single loop on the left sides of the figures in the cases of low doping or low value of the chemical potential, as clearly demonstrated in Fig.\[FIG:2\]. It is noteworthy that both bulk plasmon modes start from $\omega_p$, [*not*]{} from $\omega_p/\sqrt{2}$ as we previously observed in the case of a single conducting substrate. The presence of a bandgap in the graphene energy spectrum leads to interesting features. First, a finite energy gap modifies the location of each single-particle excitation region. It is especially unusual to observe the extension of the upper plasmon branches, which are understood to arise from a semi-infinite substrate. It is also apparent that the acoustic mode is broken into two separate undamped parts located between the two distinct single-particle excitation regions, as has been previously reported for free-standing gapped graphene [@pavlo] (see Figs. \[FIG:3\] (c) and (d)). The figures show that specifically acoustic plasmon mode behavior persists in our present case of encapsulated graphene. Within the framework of our assumption that the spacing between the conducting half-spaces is narrow, $a\ll 1/q_\parallel$, the results are relatively insensitive to the gap separation, and we consider only small values of $a$. Moreover, we examined modification of the plasmon spectra for various values of the Fermi energy, $\mu$, as depicted in Figs. \[FIG:2\] through \[FIG:4\]. In the matter of experimental realization of our results, it is necessary to achieve a situation in which the surface plasmons are not Landau damped, i.e., the corresponding branches are located outside of the upper inter-band part of the single-particle excitation spectrum. Accordingly, the frequency of the surface plasmon mode should be comparable with the Fermi energy in graphene, as well as with the corresponding $q^{1/2}$-2D-graphene plasmon energy. Only if this condition is satisfied will it be possible to observe the strong plasmon coupling we have reported here. Therefore, one must ensure that the bulk plasmon frequency is in the range of $\sim .1 \, eV$ and below [@NPo1]. Ref \[\] is an experimental paper on $Bi_2Se_3$ which is a heavily doped topological insulator, where the surface plasmon energy was found to be around $0.1 \, eV$. Evidence of mutual interaction between the surface plasmon and the Dirac plasmon of $Bi_2Se_3$ has been provided by using high-resolution electron energy loss spectroscopy. Additionally, at a graphene/$Bi_2Se_3$ interface, which was recently experimentally realized by Kepaptsoglou, et al. [@NPo2], the surface plasmon of $Bi_2Se_3$ is hybridized with acoustic plasmons in graphene. In this vein, the Fermi energy of free-standing graphene corresponding to an electron density $n=10^{16}\, m^{-2}$ is $ E _F = \hbar v_F \sqrt{\pi n } = 0.12\, eV$, so the two quantities are of the same order of magnitude. ![(Color online) Particle-hole modes and plasmon dispersion for doped, gapless graphene sandwiched between two semi-infinite conducting plasmas with separation $a=0.5\ k_F^{-1}$. Panels $(a)$ and $(b)$ present the regions where the plasmons are damped for $\mu=\hbar\omega_p$ and $\mu=1.5\hbar\omega_p$. respectively. Panels $(c)$ and $(d)$ show density plots for the plasmon excitation spectra corresponding to the parameters for zero energy gap and the same bulk plasma separation chosen in $(a)$ and $(b)$. Panels $(e)$ and $(f)$ show the plasmon excitations, both damped (dashed lines) and undamped (solid lines), obtained by solving $ \mbox{Re} \, D (q_\parallel,\omega) = 0 $ for the chosen distance $a=0.5k_F^{-1}$ between the bulk surfaces with the 2D layer at $z=0$, the same as in $(a)$ through $(d)$. The value of the parameter in the hydrodynamical model is $\beta=c_0 v_F$ in terms of the Fermi velocity $v_F$ for graphene and $c_0=(3/5)^{1/2}(v_F^{3D}/v_F)$, where $v_F^{3D}$ is the Fermi velocity of the conducting substrate (but we take $c_0=1$ here).[]{data-label="FIG:2"}](fig2){width="55.00000%"} ![(Color online) Particle-hole modes and plasmon dispersion for doped graphene encapsulated between two semi-infinite conducting plasmas with separation $a=0.5\ k_F^{-1}$. The graphene layer possesses an intermediate value for the energy bandgap $\Delta=0.2\ \mu$ determining the energy dispersion. Panels $(a)$ and $(b)$ present density plots for $ \mbox{Im}\, D^{-1} (q_\parallel,\omega)$, showing the regions where the plasmons are damped when $\mu=\hbar\omega_p$ and $\mu=1.5\hbar\omega_p$, respectively. Panels $(c)$ and $(d)$ exhibit density plots for the plasmon excitation spectra corresponding to the parameters for the energy gap $\Delta=0.2\mu$ and bulk plasma separation chosen in $(a)$ and $(b)$. Panels $(e)$ and $(f)$ show the plasmon excitations, both damped (dashed lines) and undamped (solid lines), obtained by solving $ \mbox{Re} \, D (q_\parallel,\omega) = 0 $ for the chosen distance between the bulk surfaces with the 2D layer at $z=0$, the same as $(a)$ through $(d)$. The value of the parameter used in the hydrodynamical model is $\beta=v_F$ in terms of the Fermi velocity for graphene, as in Fig. \[FIG:2\].[]{data-label="FIG:3"}](fig3){width="55.00000%"} ![(Color online) Particle-hole modes and the plasmons for graphene encapsulated between two semi-infinite conducting substrates. Here, we present a set of graphs, similar to those in Fig. \[FIG:3\], but for the case of a large energy gap $\Delta=0.6\ \mu$.[]{data-label="FIG:4"}](fig4){width="55.00000%"} Concluding Remarks {#sec4} ================== In this paper, we have investigated the properties of the plasmon spectra for a heterostructure consisting of a pair of identical semi-infinite conductors and a 2D graphene layer sandwiched between them. Our formulation is suitable when the separation between the two semi-infinite bulk conducting materials is small compared to the inverse Fermi wave number in graphene, and the whole system is symmetric in the $z-$direction perpendicular to the 2D layer. We have obtained the plasmon dispersion relations for this encapsulated graphene system for both zero and finite energy bandgap and for various values of the chemical potential. In each case, we clearly obtain three hybridized plasmon modes, one of which (the acoustic branch) starts from the origin and is attributed primarily to the graphene layer and the other two modes originating at $\omega=\omega_p$ are considered as optical plasmons. This situation is novel and has not been encountered previously in a system involving a 2D layer with a single semi-infinite conductor. Each of the branches exhibits a specific behavior depending on the chemical potential and consists of various undamped parts which are determined by the energy bandgap and its ratio to the doping parameter, as discussed above. The low-frequency branch has a linear dispersion at long wavelengths and becomes damped by the intra-band and inter-band particle-hole modes as wave vector is increased. There are two bulk plasmon branches which are depolarization shifted by the Coulomb interaction. In this regard, the results demonstrate that there is almost no dependence on the distance $a$ between the two substrates as long as the condition $a \ll k_F^{-1}$ is satisfied. Moreover, there is no evidence of critical damping of the acoustic plasmon branch in the long wavelength limit ($q_\parallel \rightarrow 0$), as was found in the previous comparative study of a monolayer of graphene interacting with a single conducting substrate [@GG]. Furthermore, some crucial properties of the plasmons in free-standing graphene with an energy bandgap, such as extension of the undamped branch and its separation into the two parts for intermediate energy gap $\Delta \backsimeq 0.22 \mu$, are also present in our results. In summary, we have developed a new analytical model and obtained a complete set of numerical results for a two-dimensional layer (graphene) surrounded by two identical thick conducting substrates. While our previous work [@GG] confirmed and offered an adequate theoretical explanation for recent experimental findings [@Pol1; @Pol2; @Pol3; @Pol4], this paper is expected to predict the correct plasmon behavior of the totally realistic and novel situation of encapsulated graphene, which is now being very actively studied experimentally [@Gong; @Kamat]. Hybridized plasmon modes of a graphene-based nanoscale system are at the focus of significant interest in the current fields of technology and applications. The authors of [@encaps5] report evidence of the formation of electron-hole puddles for encapsulated graphene by hexagonal BN. This is an interesting effect which was not taken into account in our model calculations. These localization effects will be investigated in future work by including a concentration of impurities and defects in interaction with the 2D layer and two substrates. The role of impurities is expected to be modest at low concentration. An increase in the concentration of the impurities will lead to a decrease in the electron correlations. The Green’s function and consequently will, in this case, be expressible by vertex corrections by the usual rules of field theory when impurities are present. Principi A., Carrega M., Lundeberg M. B., Woessner A., Koppens F. H. L., Vignale G., and Polini M.,“Plasmon losses due to electron-phonon scattering: The case of graphene encapsulated in hexagonal boron nitride," Phys. Rev. B [**90**]{}, 165408 (2014). Kharche N. and Nayak S.K.,“Quasiparticle Band Gap Engineering of Graphene and Graphone on Hexagonal Boron Nitride Substrate," Nano Lett. [**11**]{}, 5274 (2011). Ristein J., Mammadov S., and Seyller T,“Origin of Doping in Quasi-Free-Standing Graphene on Silicon Carbide," , 246104 (2012). Ryzhii V. and Satou A.,“Plasma waves in two-dimensional electron-hole system in gated graphene heterostructures," Jour. Appl. Phys. [**101**]{}, 024509 (2007). Woessner A., Lundeberg M. B., Gao Y., Principi A., González P.A., Carrega M., Watanabe K., Taniguchi T., Vignale G., Polini M., Hone J., Hillenbrand R., and Koppens F.H.L.,“Highly confined low-loss plasmons in graphene–boron nitride heterostructures," Nature Materials [**14**]{}, 421 (2015). Mayorov A. S., Gorbachev R. V., Morozov S.V., Britnell L., Jalil R., Ponomarenko L.A.,Blake B., Novoselov K.S., Watanabe K., Taniguchi T., and Geim A.K.,“Micrometer-Scale Ballistic Transport in Encapsulated Graphene at Room Temperature," Nano Lett. [**11**]{} (6), 2396 (2011). Guimarães M.H.D., Zomer P.J., Ingla-Aynés J., Brant J.C., Tombros N., and Wees B.J.V.,“Controlling Spin Relaxation in Hexagonal BN-Encapsulated Graphene with a Transverse Electric Field," Phys. Rev. Lett. [ **113**]{} 086602 (2014). Yang S., Feng X., Ivanovici S.,and Müllen K., “Fabrication of Graphene-Encapsulated Oxide Nanoparticles: Towards High-Performance Anode Materials for Lithium Storage," Angewandte Chemie. [**49**]{}, 8408 (2010). Britnell L., Gorbachev R.V., Jalil R., Belle B.D., Schedin F., Mishchenko A., Georgiou T., Katsnelson M.I., Eaves L., Morozov5 S.V., Peres N.M.R., Leist J., Geim A.K, Novoselov K.S., Ponomarenko L.A.,“Field-Effect Tunneling Transistor Based on Vertical Graphene Heterostructures," Science [**335**]{}, 6071 (2012). Kretinin A.V., Cao Y., Tu J.S., Yu G.L., Jalil R. , Novoselov K.S., Haigh S.J., Gholinia A., Mishchenko A., Lozada M., Georgiou T., Woods C.R., Withers F., Blake P., Eda g., Wirsig A., Hucho C., Watanabe K., Taniguchi T., Geim A.K., and Gorbachev R.V.,“Electronic Properties of Graphene Encapsulated with Different Two-Dimensional Atomic Crystals," Nano Lett. [**14** ]{}, 3270 (2014). Gumbs G, Iurov A., and Horing N.J.M.,“Nonlocal plasma spectrum of graphene interacting with a thick conductor," , 235416 (2015). Horing N.J.M., Iurov A., Gumbs G., Politano A., $\&$ Chiarello G.,“Recent Progress on Nonlocal Graphene/Surface Plasmons," (pp. 205-237), Springer International Publishing (2016). Horing N. J. M.,“Coupling of graphene and surface plasmons," , 193401 (2009). Politano A., Marino A.R., Formoso V., Farías D., Miranda R., and Chiarello G.,“Evidence for acoustic-like plasmons on epitaxial graphene on Pt(111)," Phys. Rev. B [**84**]{}, 033401 (2011). Politano A., Marino A. R., and Chiarello G.,“Effects of a humid environment on the sheet plasmon resonance in epitaxial graphene," Phys. Rev. B [**86**]{}, 085420 (2012). Politano A., Formoso V., and Chiarello G.,“Evidence of composite plasmon–phonon modes in the electronic response of epitaxial graphene,", Journal of Physics: Condensed Matter, [**25**]{} (34), 345303 (2013). Politano A. and Chiarello G.,“Quenching of plasmons modes in air-exposed graphene-Ru contacts for plasmonic devices ,", Applied Physics Letters, [**102**]{} , 201608. (2013). Politano A. and Chiarello G., “Unravelling suitable graphene–metal contacts for graphene-based plasmonic devices," Nanoscale, [**5**]{}, 8215(2013). Chuang H. J., Tan X., Ghimire N. J., Perera M. M., Chamlagain B., Cheng M. M. C., Yan J., Mandrus D., Tomanek D., and Zhou Z., “High Mobility WSe$_2$ p- and n-Type Field-Effect Transistors Contacted by Highly Doped Graphene for Low-Resistance Contacts," Nano Lett. [**14**]{} 3594–3601 (2014). Shih C.J., Wang Q. H. , Son Y., Jin Z., Blankschtein D., and Strano M. S., “Tuning On-Off Current Ration and Field-Effect Mobility in a $MoS_2$ Graphene Heterostructure via Schottky Barrier Modulation," ACS Nano, [**8**]{}, 5790 (2014). Low C. G., Zhang Q., Hao Y., and Ruoff R. S.,“Graphene Field Effect Transistors with Mica as Gate Dielectric Layers," Nano Small Micro, [**10**]{}, 4213 ,(2014). Kretinin A. V., Cao Y., Tu J. S., Yu G. L., Jalil R., Novoselov K. S. , Haigh S. J., Gholinia A., Mishchenko A., Lozada M., Georgiou T., Woods C. R., Withers F. , Blake P., Eda G., Wirsig A., Hucho C., Watanabe K., Taniguchi T., Geim A. K. , and Gorbachev R. V., “Electronic Properties of Graphene Encapsulated with Different Two-Dimensional Atomic Crystals," Nano Lett. [**14**]{}, 3270 (2014). Sun T., Wang Z. L., Shi Z. J., Ran G. Z. , Xu W. J., Wang Z. Y. , Li Y. Z., Dai L., and Qin G. G.,“Multilayered graphene used as anode of organic light emitting devices," Applied Phys. Lett. [**96**]{}, 133301 (2010). Dean C.R., Young A.F., Meric I., Lee C., Wang L., Sorgenfrei S. , Watanabe K., Taniguchi T., Kim P., Shepard K. L., and Hone J. “Boron nitride substrates for high quality graphene electron ics," Nature Nanotechnology [**5**]{}, 722 (2010). Eguiluz, A. Ying S. C., and Quinn J. J., “Influence of the electron density profile on surface plasmons in a hydrodynamic model", Phys. Rev. B [**11**]{}, 2118 (1975). Horing N. J. M., Kamen E. and, Gumbs G., “Surface correlation energy and the hydrodynamic model of dynamic, nonlocal bounded plasma response," Phys. Rev. B [**31**]{}, 8269 (1985). Wunsch B., Stauber T., Sols F., and Guinea F.,“Dynamical polarization of graphene at finite doping," New Jour. of Phys. [**8**]{}, 318 (2006). Pyatkovskiy P. K.,“Dynamical polarization, screening, and plasmons in gapped graphene," J. Phys.: Condens. Matter [**21**]{}, 025506 (2009). Politano A., Silkin V.M., Nechaev I.A., Vitiello M.S., Viti L., Aliev Z.S., Babanly M.B., Chiarello G. , Echenique P.M., and Chulkov E.V.,“Interplay of Surface and Dirac Plasmons in Topological Insulators: The Case of $Bi_2Se_3$," Phys. Rev. Lett.[ **115**]{} 216802 (2015) . Kepaptsoglou D. M., Gilks D., Lari L., Ramasse Q. M., Galindo P., Weinert M., Li L., Nicotra G. and Lazarov V. K.,“STEM and EELS study of the Graphene $Bi_2Se_3$ Interface," Microsc. Microanal. [**21**]{} 1151 (2015). Gong C., Hinojos D., Wang W., Nijem N., Shan B., Wallace R. M., Cho K., and Chabal Y. J.,“Metal–Graphene–Metal Sandwich Contacts for Enhanced Interface Bonding and Work Function Control" ACS Nano [**6**]{}, 5381–5387 (2012). Kamat P. V.,“Graphene-Based Nanoarchitectures. Anchoring Semiconductor and Metal Nanoparticles on a Two-Dimensional Carbon Support," J. Phys. Chem. Lett., [**1**]{} (2), 520–527 (2010).
{ "pile_set_name": "ArXiv" }
--- abstract: | We determine necessary conditions on the structure of symbol error rate (SER) optimal quantizers for limited feedback beamforming in wireless networks with one transmitter-receiver pair and $R$ parallel amplify-and-forward relays. We call a quantizer codebook “small” if its cardinality is less than $R$, and “large” otherwise. A “d-codebook” depends on the power constraints and can be optimized accordingly, while an “i-codebook” remains fixed. It was previously shown that any i-codebook that contains the single-relay selection (SRS) codebook achieves the full-diversity order, $R$. We prove the following: Every full-diversity i-codebook contains the SRS codebook, and thus is necessarily large. In general, as the power constraints grow to infinity, the limit of an optimal large d-codebook contains an SRS codebook, provided that it exists. For small codebooks, the maximal diversity is equal to the codebook cardinality. Every diversity-optimal small i-codebook is an orthogonal multiple-relay selection (OMRS) codebook. Moreover, the limit of an optimal small d-codebook is an OMRS codebook. We observe that SRS is nothing but a special case of OMRS for codebooks with cardinality equal to $R$. As a result, we call OMRS as “the universal necessary condition” for codebook optimality. Finally, we confirm our analytical findings through simulations. author: - 'Erdem Koyuncu, and Hamid Jafarkhani, *Fellow*, IEEE[^1]' bibliography: - 'IEEEabrv.bib' - 'letter.bib' title: The Necessity of Relay Selection --- Wireless relay networks, relay selection, diversity, quantizer optimality. Introduction ============ availability of channel state information (CSI) can greatly affect the performance and reliability of amplify-and-forward (AF) cooperative relay networks. With available CSI, each relay can adaptively adjust its transmit power and transmit phase. This network beamforming scheme has been shown to achieve maximal diversity and array gains[@larsson1; @jing1; @koyuncu1]. In contrast, without any CSI at the relays, distributed space-time coding schemes can also achieve maximal diversity, but they also incur an unbounded array gain loss compared to network beamforming[@laneman2; @jing4]. For networks with parallel relays, the optimal beamforming policy requires one or two real numbers to be broadcasted from the receiver to the relays. A more practical assumption is that there is only partial CSI at the relays. For such networks, it has been shown that beamforming with quantized instantaneous CSI can achieve not only the maximal diversity gain but also a very high array gain with only a few feedback bits[@koyuncu1; @zhao5]. A special case of quantized feedback for relay networks is single-relay selection (SRS) [@bletsas1; @riberio1; @anghel1; @hasna1; @zhao3; @zhao6; @jing6], which uses $\lceil \log_2 R \rceil$ feedback bits per channel state for a network with $R$ relays. It allows only one of the relays to cooperate given a constant fading block. This simple quantization scheme has been shown to achieve full-diversity for a very broad class of network topologies[@koyuncu1; @shi1; @koyuncu3], and even under suboptimal selection criteria[@jing6]. In this work, we consider a network with one transmitter-receiver pair and $R$ parallel AF relays. We assume that there is no direct link between the transmitter and the receiver. The transmitter and the relays have their own short-term power constraints. We assume that the receiver has full CSI, while each relay knows only the magnitude of its own receiving channel and has $B$ bits of partial CSI. The feedback bits are conveyed from the receiver to the relays via error-free and delay-free feedback channels, and they represent a quantized beamforming vector. Our performance measure is the symbol error rate (SER). A well-known performance measure that is closely related to SER is diversity. We define the diversity measure for our network model as follows: Let $P_0$ and $P_i,\,i=1,\ldots,R$ represent the transmitter and relay power constraints, respectively. We allow these power constraints to vary linearly with a common power constraint $P$ as $P_i \triangleq p_i P,\,i=0,\ldots,R$, where $p_i$ are fixed positive real numbers that are independent of $P$. Then, as $P\rightarrow\infty$, the SER converges to $\mathtt{a}P^{-\mathtt{d}}$, where $\mathtt{a}$ and $\mathtt{d}$ represent the *array gain*, and the *diversity gain*, respectively. Since there are $R$ independently fading paths between the transmitter and the receiver, the maximal spatial diversity of our network is $R$, which we call the full-diversity order. The set of all $2^B$ quantized beamforming vectors is the quantizer codebook. For clarity of exposition, we classify the codebooks under two criteria, one of which is cardinality: We call a codebook “small” if its cardinality is less than the number of relays, and “large” otherwise. We shall see later on that it is necessary to use a large codebook in order to achieve full-diversity, and correspondingly, the diversity provided by a small codebook is strictly less than $R$. The other criterion that we use is the codebooks’ dependence on the transmitter and relay power constraints, the motivation of which we now explain. In general, we can optimize the codebook with respect to the power constraints, as demonstrated in [@koyuncu1]. We call such power-dependent codebooks as “d-codebooks”. Note that, an optimal codebook given some power constraints will lose its optimality as soon as any of the constraints are changed. Then, in order to achieve the best performance for any choice of constraints, the receiver and the relays need to store a possibly large number of optimal codebooks. A more practical approach might be to consider a power-independent codebook (i-codebook). In this case, a single codebook is used for all possible constraints with the purpose of achieving high diversity and array gains. The main contributions of this paper can be summarized as follows: First, we show that every full-diversity i-codebook necessarily contains the SRS codebook. We obtain an analogous result for power-dependent codebooks: As $P\rightarrow\infty$, the limit of an optimal large d-codebook contains an SRS codebook, provided that it exists. Both results show that full-diversity codebooks should incorporate the SRS codebook structure, and are necessarily large. For small codebooks, we show that the maximal achievable diversity is equal to the cardinality of the codebook. We would like to note that, even though this result is well-known for the case of limited feedback beamforming in multiple-input single-output (MISO) systems[@mukkavilli1; @roh1; @love1; @koyuncu2], its proof requires a completely different approach in our case. Having determined the best achievable diversity of small codebooks, we show that any optimal small i-codebook is an “orthogonal multiple-relay selection” (OMRS) codebook, meaning that it consists of multiple-relay selection vectors that are pairwise orthogonal. We also show that the limit of an optimal small d-codebook is an OMRS codebook. Both results demonstrate the necessity of OMRS for the optimality of small codebooks. We believe that OMRS is also a sufficient condition for optimality, but rather surprisingly, a formal proof seems difficult and will not be considered in this paper. Finally, we observe that SRS is just a special case of OMRS for codebooks with cardinality equal to $R$. As a result, OMRS becomes the universal necessary condition for optimality. Our results in this paper is in contrast to limited feedback beamforming in MISO systems, in which any set of linearly independent beamformers guarantee maximal diversity[@koyuncu2], and the performance of a codebook is invariant under unitary transformations[@mukkavilli1; @roh1; @love1]. In that sense, this paper also shows that the vast literature on limited feedback beamforming for point-to-point systems is not directly applicable to cooperative networks, and we need new methods of analysis. The rest of the paper is organized as follows: In Section \[secNetMod\], we introduce our system model, feedback and data transmission schemes, and problem definition. In Section \[secLowerBounds\], we present a fundamental lemma that we frequently use in our proofs. In Sections \[secSelecLarge\] and \[secSelecSmall\], we state our main results on the necessity of SRS for large codebooks, and the necessity of OMRS for small codebooks, respectively. The numerical results are provided in Section \[secNumerical\]. Finally, in Section \[secConc\], we draw our main conclusions. Some technical proofs are provided in the appendices. *Notation:* $||\cdot||_{\infty}$ indicates the infinite-norm. $\mathbb{C}$, $\mathbb{R}$, and $\mathbb{N}$ represent the sets of complex numbers, real numbers, and natural numbers, respectively. For $z\in\mathbb{C}$, $|z|$ indicates the absolute value. For a random variable $X$, $f_X(\cdot)$ and $F_X(\cdot)$ represent the probability density function (PDF) and the cumulative distribution function (CDF), respectively. $\mathcal{CN}(0, \sigma^2)$ represents a zero-mean complex Gaussian random variable with variance $\tfrac{\sigma^2}{2}$ per complex dimension. $\mathrm{E}[X]$ is the expected value of $X$. For any sets $\mathcal{A}$ and $\mathcal{B}$, $\mathcal{A}-\mathcal{B}$ is the set of elements in $\mathcal{A}$, but not in $\mathcal{B}$. $\mathcal{A}\subset \mathcal{B}$ means $\mathcal{A}$ is a subset of $\mathcal{B}$. $\mathcal{A}\cap\mathcal{B}$ and $\mathcal{A}\cup\mathcal{B}$ are the intersection and the union of $\mathcal{A}$ and $\mathcal{B}$, respectively. $|\mathcal{A}|$ is the cardinality of $\mathcal{A}$, $\mathcal{A}^+ \triangleq \{x:x>0,\,x\in\mathcal{A}\}$, and $\mathcal{A}^r = \left\{(a_1,\ldots,a_r)|a_1,\ldots, a_r\in\mathcal{A}\right\}$, $r\in\mathbb{N}^+$ is the cartesian power. Finally, $\emptyset$ is the empty set, $\mathrm{Q}(\cdot)$ represents the Gaussian tail function, $\Gamma(\cdot)$ is the gamma function, $\log(\cdot)$ is the natural logarithm, and $\sinh(\cdot)$ is the hyperbolic sine. Network Model and Problem Statement {#secNetMod} =================================== System Model {#secSystemModel} ------------ The block diagram of the system is shown in Fig. \[systemblockdiagram\]. We have a relay network with a transmitter-receiver pair and $R$ parallel relays. Let $f_r$ and $g_r$ denote the channel from the transmitter to the $r$th relay, and from the $r$th relay to the receiver, respectively. Also, let $\mathbf{h} = (f_1,g_1,\ldots,f_R,g_R)$ denote the channel state of the entire network. We assume that the entries of $\mathbf{h}$ are independent and $f_r \sim \mathcal{CN}(0, \sigma_{f_r}^2),\,g_r \sim \mathcal{CN}(0, \sigma_{g_r}^2),\,r=1,\ldots,R$. Only the short-term power constraint is considered: For every symbol transmission, the average power levels used at the transmitter and the $r$th relay are no larger than $P_0$ and $P_r$, respectively. Let $P_i = p_i P,\,i=0,\ldots,R$, where $\infty>p_i>0$. In other words, we allow the power constraints of the transmitter and the relays to vary linearly with $P$. In addition, $P$ is the only network parameter that we allow to vary. All the remaining parameters (the channel variances $\sigma_{f_r},\sigma_{g_r},\,r=1,\ldots,R$ and the power constraint scalers $p_r,\,r=0,\ldots,R$) can be arbitrary, but we assume that they are fixed positive constants that do not depend on $P$. We assume a quasi-static channel model, in which the channel realizations vary independently from one channel state to another, while within each state they remain constant. Also, we assume that the receiver knows the channel state of the entire network, $\mathbf{h}$; and the $r$th relay knows $|f_r|$. Each relay has also $B$ bits of partial CSI provided by receiver feedback. Feedback Transmission Scheme ---------------------------- For any $P\in\mathbb{R}^+$ and a finite number of feedback bits $B$, the feedback transmission scheme operates as follows: For each frame, the channel realization $\mathbf{h}$ is quantized by a quantizer $\mathcal{Q}_P(\mathbf{h})\triangleq \mathtt{DEC}_P(\mathtt{ENC}_P(\mathbf{h}))$ defined by the encoder and decoder mappings $\mathtt{ENC}_P:\mathbb{C}^{2R}\rightarrow\mathcal{I}$, and $\mathtt{DEC}_P:\mathcal{I}\rightarrow\mathcal{D}_P$. In this definition, $\mathcal{D}_P$ represents the quantizer codebook for power level $P$, and $\mathcal{I}\triangleq \{1,\ldots,2^B\}$ denotes the index set for the codebook elements. We assume throughout the paper that codebooks for different power levels have the same cardinality, i.e. $|\mathcal{D}_P| = 2^B,\,\forall P$. The encoding operation is performed at the receiver, and the feedback bits represent the encoder output. Each relay uses the decoder to find the corresponding codebook element. Each codebook element corresponds to a quantized beamforming vector. To summarize, we have a collection $\mathscr{Q} \triangleq \{\mathcal{Q}_P:P\in\mathbb{R}^+\}$ of quantizers. For a given $P\in\mathbb{R}^+$, we use the quantizer $\mathcal{Q}_P$ that provides the beamforming vector $\mathcal{Q}_P(\mathbf{h}) = \mathbf{x}$ for some $\mathbf{x}\in\mathcal{D}_P$. Let $\mathcal{D}$ represent the mapping that maps a power level $P$ to its corresponding codebook $\mathcal{D}_P$. With some abuse of language, we call the set-valued map $\mathcal{D}$ a *power-dependent codebook* (d-codebook) in the sense that for a given $P$, the codebook $\mathcal{D}_P$ is employed and can be optimized accordingly. Even though an optimal d-codebook can provide the best possible performance at any $P$, its accommodation requires the receiver and the relays to store a large number of codebooks. A more practical approach might be to use a *power-independent codebook* (i-codebook) $\mathcal{C}$ as a special case of d-codebooks with $\mathcal{D}_P = \mathcal{C},\,\forall P$. We will have more to say on the practicality of i-codebooks later on. Data Transmission Scheme {#secDataTransmisson} ------------------------ We use a two-step AF protocol[@jing1; @koyuncu1]. In the first step, the transmitter selects a symbol $s$ from a constellation $\mathcal{S}$, where $|\mathcal{S}|<\infty$, and sends $\sqrt{P_0}s$. We normalize $s$ as $\mathrm{E}[|s|^2]=1$. Thus, the average power used at the transmitter is $P_0$. During the first step, there is no reception at the receiver, but the $r$th relay receives $$\begin{aligned} t_r = f_r s\sqrt{P_0} + \eta_r,\end{aligned}$$ where $\eta_{r}\sim\mathcal{CN}(0,1)$. Suppose that $\mathcal{Q}_P(\mathbf{h}) = \mathbf{x}$, for some $\mathbf{x}\in\mathcal{D}_P$. Then, the relays use the beamforming vector $\mathbf{x}$ to adjust their transmit power and transmit phase. During the second step, the transmitters remain silent, but the $r$th relay transmits $$\begin{aligned} u_r = x_r \sqrt{\rho_r} t_r,\end{aligned}$$ where $$\begin{aligned} \rho_r \triangleq \frac{P_r}{1 + |f_r|^2P_0}.\end{aligned}$$ The average power used at the $r$th relay can be calculated to be $|x_r|^2P_r$. We require $0\leq|x_r|\leq 1$ as a result of the short term power constraint. The channel state dependent normalization factors $\rho_r$ ensure that the instantaneous transmit power of each relay remains within its power constraint with high probability. Also, note that within the restriction of $0 \leq |x_r| \leq 1$, $\rho_r$ is the maximal normalization factor that we can use. In other words, if a factor $\rho_r''$ satisfies $\rho_r'' > \rho_r$ for some $\mathbf{h}$, then it violates the short term power constraint. Still, one can employ another factor $\rho_r'$ with $\rho_r' \leq \rho_r,\,\forall\mathbf{h}$ (e.g. $\rho_r' = P_r/(1+|f_r|^3P_0)$). We shall see later in Section \[secLowerBounds\] that a different choice of the normalization factor does not change the main results of this paper. After the two steps of transmission, the received signal at the receiver can be expressed as: $$\begin{aligned} y = \sum_{r=1}^R x_r\sqrt{\rho_r}f_{r}g_{r}\sqrt{P_0}s + \sum_{r=1}^R x_r g_{r}\sqrt{\rho_r} \eta_{r} + \eta_{0},\end{aligned}$$ where $\eta_{0}\sim\mathcal{CN}(0,1)$ is the noise at the receiver. We assume that the noises $\eta_i,\,i=0,\ldots,R$ and the channels are all independent. It follows that the received SNR is given by $$\begin{aligned} \label{receivedsnr} \mathtt{SNR}_P(\mathbf{x},\mathbf{h}) \triangleq \frac{P_0\left|\sum_{r=1}^R x_rf_rg_r\sqrt{\rho_r}\right|^2}{1 + \sum_{r=1}^R |x_r|^2|g_r|^2\rho_r}.\end{aligned}$$ Since we have assumed that the power constraint scalers $p_i$ are fixed, the received SNR depends only on $P$ (indicated by a subscript), the beamforming vector $\mathbf{x}$, and the channel state $\mathbf{h}$. In this work, our performance measure is the SER, and we present our results only for the case when $\mathcal{S} = \{+1,-1\}$ is a binary constellation. Then, the SER achieved by the quantizer $\mathcal{Q}_P$ at power level $P$ can be expressed as $$\begin{aligned} \label{serofaquantizer} \mathtt{SER}_P(\mathcal{Q}_P) \triangleq E_{\mathbf{h}} [\mathrm{Q}\sqrt{2 \mathtt{SNR}_P(\mathcal{Q}_P(\mathbf{h}),\mathbf{h})}].\end{aligned}$$ We would like to note that our results can be extended to any finite constellation $\mathcal{S}$. Using (\[serofaquantizer\]), the diversity achieved by the collection $\mathscr{Q}$ of quantizers is given by $$\begin{aligned} \label{divofaquantizer} {\mathtt{d}}(\!\mathscr{Q}\;\!) \triangleq \lim_{P\rightarrow\infty} -\frac{\log \mathtt{SER}_P(\mathcal{Q}_P)}{\log P}.\end{aligned}$$ Since there are $R$ independently fading paths between the transmitter and the receiver, the maximal spatial diversity of our network model is $R$. In other words, for any $\mathscr{Q}$, ${\mathtt{d}}(\mathscr{Q}) \leq R$. A more formal proof of this argument can be found in [@koyuncu3 Theorem 1]. Problem Statement {#secProblemStatement} ----------------- Let $|\mathcal{D}|$ represent the common cardinality of each codebook $\mathcal{D}_P$. In other words, $|\mathcal{D}| = |\mathcal{D}_P| = 2^B,\,\forall P$. We are interested in the structure of the optimal quantizers that minimize the SER subject to $|\mathcal{D}|= 2^B$. The following proposition from [@koyuncu1] determines the optimal quantizers given a fixed codebook. \[optimalquantizerencoderlemma\] Given a fixed d-codebook $\mathcal{D}$ (i.e., for any $P$, the codebook $\mathcal{D}_P$ is fixed), the collection of optimal quantizers that minimize the SER is given by $\mathscr{Q}_{\mathcal{D}}^{\star} \triangleq \{\mathcal{Q}_{P\!,\,\mathcal{D}_P}^{\star}:P\in\mathbb{R}^+ \}$, where $$\begin{aligned} \label{dcodebookoptimalencoder} \mathcal{Q}_{P\!,\,\mathcal{D}_P}^{\star}(\mathbf{h}) \triangleq \arg\max_{\mathbf{x}\in\mathcal{D}_P} \mathtt{SNR}_P(\mathbf{x},\mathbf{h}),\,P\in\mathbb{R}^+.\end{aligned}$$ In particular, given a fixed i-codebook $\mathcal{C}$, the collection of optimal quantizers is given by $\mathscr{Q}_{\mathcal{C}}^{\star} \triangleq \{\mathcal{Q}_{P\!,\,\mathcal{C}}^{\star}:P\in\mathbb{R}^+ \}$, where $$\begin{aligned} \label{icodebookoptimalencoder} \mathcal{Q}_{P\!,\,\mathcal{C}}^{\star}(\mathbf{h}) = \arg\max_{\mathbf{x}\in\mathcal{C}} \mathtt{SNR}_P(\mathbf{x},\mathbf{h}),\,P\in\mathbb{R}^+.\end{aligned}$$ In other words, for d-codebooks, given any power level $P$ and any fixed codebook $\mathcal{D}_P$ for $P$, the optimal quantizer encoder chooses the beamforming vector that maximizes the SNR at $P$. The interpretation of Proposition \[optimalquantizerencoderlemma\] for an i-codebook $\mathcal{C}$ is analogous. We would like to note that in practice, $|\mathcal{D}_P|<\infty$, and thus $\arg\max_{\mathbf{x}\in\mathcal{D}_P} \mathtt{SNR}_P(\mathbf{x},\mathbf{h})$ in (\[dcodebookoptimalencoder\]) will always exist for any $\mathbf{h}$ and $P$. In order to be able to handle codebooks with $|\mathcal{D}_P| = \infty$, we shall further assume throughout the paper that $\mathcal{D}_P$ is compact for all $P$. Similarly, we assume that all i-codebooks are compact without explicit specification. The main motivation for our introduction of i-codebooks was the claim that they are more practical than d-codebooks: One does not need to store different codebooks for different power levels. On the other hand, (\[icodebookoptimalencoder\]) shows us that even if we use an i-codebook, the quantizer encoder will always depend on $P$. In that sense, one might argue that i-codebooks are as impractical as d-codebooks since a large number of quantizer encoders need to be stored anyway. Fortunately, for i-codebooks, we can observe from (\[icodebookoptimalencoder\]) that the optimal encoder is a simple algebraic function of $P$. Therefore, we do not actually need to store the entire set of encoders. In order for a similar situation to hold for d-codebooks though, one needs a simple function that can map any power level to its corresponding optimal codebook. Finding such a function is an open problem. We thus present our results for both d-codebooks and i-codebooks, due to the potential optimality of the former and the practicality of the latter. We shall use the optimal encoder in Proposition \[optimalquantizerencoderlemma\] for the rest of the paper. Then, the codebook uniquely determines the performance of the system, and we set ${\mathtt{d}}(\mathcal{D}) \triangleq {\mathtt{d}}(\mathscr{Q}_{\mathcal{D}}^{\star})$ for a d-codebook $\mathcal{D}$, and similarly, ${\mathtt{d}}(\mathcal{C}) \triangleq {\mathtt{d}}(\mathscr{Q}_{\mathcal{C}}^{\star})$ for an i-codebook $\mathcal{C}$. Any optimal codebook should obey the following proposition from [@koyuncu1]. \[codebookoptimalitylemma\] If $\mathcal{D}$ is an optimal d-codebook, then $\mathcal{D}_P\subset\mathcal{X},\,\forall P$, where $\mathcal{X} = \{\mathbf{x}:\mathbf{x}\in\mathbb{C}^R,\,\|\mathbf{x}\|_{\infty} = 1\}$. In particular, if $\mathcal{C}$ is an optimal i-codebook, then $\mathcal{C}\subset\mathcal{X}$. In other words, at least one component of every beamforming vector in the codebook should have unit norm. Unless otherwise specified, we shall assume that all the codebooks in the rest of this paper are optimal in the sense of Proposition \[codebookoptimalitylemma\]. One simple, yet effective structured i-codebook is the SRS codebook, given by $\mathcal{C}_{\mathtt{SRS}}(\boldsymbol{\theta}) \triangleq \{\mathbf{e}_r(\theta_r):r=1,\ldots,R\}$, where $\boldsymbol{\theta} = {[\begin{array}{ccc} \theta_1 & \cdots & \theta_R \end{array}]}$, and $\mathbf{e}_r(\theta_r) \triangleq {[\begin{array}{ccc} e_{r1}(\theta_r) & \cdots & e_{rR}(\theta_r) \end{array}]}$ with $e_{rq}(\theta_r) = e^{j\theta_r},\,r=q$ and $\mathbf{e}_{rq}(\theta_r) = 0,\,r\neq q$. As an example, both $$\begin{aligned} \label{zerolusrs} \mathcal{C}_{\mathtt{SRS}}\left(\boldsymbol{0}\right) = \bigl\{{[\begin{array}{ccc} 0 & 0 & 1 \end{array}]},\,{[\begin{array}{ccc} 0 & 1 & 0 \end{array}]},\,{[\begin{array}{ccc} 1 & 0 & 0 \end{array}]}\bigr\}\end{aligned}$$ and $$\begin{aligned} \mathcal{C}_{\mathtt{SRS}}\left({[\begin{array}{ccc} \frac{\pi}{4} & \frac{\pi}{2} & \frac{2\pi}{3} \end{array}]}\right) = \bigl\{{[\begin{array}{ccc} e^{j\frac{\pi}{4}} & 0 & 0 \end{array}]},\,{[\begin{array}{ccc} 0 & j & 0 \end{array}]},\,{[\begin{array}{ccc} 0 & 0 & e^{j\frac{2\pi}{3}} \end{array}]}\bigr\}\end{aligned}$$ are SRS codebooks for a network with $3$ relays, where $\boldsymbol{0}$ represents the all-zero vector. Even though there are infinitely many possible SRS codebooks given any $R$, all of them provide the same SER at any given $P$. This follows immediately from \[snrinvariance\] For any beamforming vector $\mathbf{x}$ and channel state $\mathbf{h}$, we have $$\begin{aligned} \mathtt{SNR}_P(\mathbf{x}, \mathbf{h}) = \mathtt{SNR}_P(e^{j\theta}\mathbf{x}, \mathbf{h}),\,\forall\theta\in\mathbb{R}.\end{aligned}$$ The proof is straightforward once we use the definition of $\mathtt{SNR}_P(\mathbf{x}, \mathbf{h})$ in (\[receivedsnr\]). Let $\mathscr{C}_{\mathtt{SRS}} = \{\mathcal{C}_{\mathtt{SRS}}(\boldsymbol{\theta}):\boldsymbol{\theta}\in\mathbb{R}^R\}$ represent the collection of all possible SRS codebooks. It was shown in [@koyuncu1 Theorem 1] that any $\mathcal{C}_{\mathtt{SRS}}\in\mathscr{C}_{\mathtt{SRS}}$ achieves the full-diversity order, $R$. Our first goal is to show that, in order to achieve diversity $R$, it is not only sufficient but also necessary to use SRS. To be more precise, it is necessary to include the SRS vectors to the quantizer codebook to achieve full-diversity. Since there are $R$ SRS vectors, any full-diversity codebook has thus cardinality at least $R$, and is necessarily large. Clearly, we cannot choose the codebook cardinality freely at our will; given $B$ feedback bits, we are restricted to a codebook with cardinality $2^B$. As a result, one needs to use at least $\lceil \log_2 R\rceil$ bits of feedback to accommodate a large codebook and achieve full-diversity. A low-rate application might require the number of feedback bits to be less than $\lceil \log_2 R\rceil$. In this case, we are restricted to using small codebooks and full-diversity is no longer achievable. Optimality conditions for small codebooks are more complicated than the ones for large codebooks and will be discussed later on. All of our results on the necessity of relay selection will be based on a fundamental lemma that provides a lower bound on the SER of a given i-codebook. We introduce this lemma in the next section together with some example applications. We discuss the necessity of SRS for full-diversity immediately afterwards. Lower Bounds on the Performance of I-Codebooks {#secLowerBounds} ============================================== We frequently use the following lemma to prove the main results in this paper. The proof of the lemma can be found in Appendix \[proofofmainlemma\]. \[mainlemma\] For any i-codebook $\mathcal{C}$, not necessarily with $\mathcal{C}\subset\mathcal{X}$, let $$\begin{aligned} \label{indexsets} \mathscr{R}(\mathcal{C}) \triangleq \{\mathcal{R}: \mathcal{R}\subset\{1,\ldots,R\}\mbox{ and }\forall \mathbf{x}\in\mathcal{C},\,\exists r\in\mathcal{R},\,x_r \neq 0\} \end{aligned}$$ be a collection of index sets. Then, there are constants $0 < {\Psi_0},\,{C_{0}},\,{C_{1}}< \infty$ that are independent of $P$ and $\mathcal{C}$ s.t. for all $P \geq {\Psi_0}$ and $\mathcal{R}\in\mathscr{R}(\mathcal{C})$, $$\begin{aligned} \label{mainlemmaeq1} \mathtt{SER}_P(\mathcal{Q}_{P\!,\,\mathcal{C}}^\star) \geq {C_{0}}[\xi(\mathcal{C},\mathcal{R})]^{\frac{3}{2}} \exp\left(-\frac{C_1}{\xi(\mathcal{C},\mathcal{R})}\right) P^{-|\mathcal{R}|},\end{aligned}$$ where $\xi(\mathcal{C},\mathcal{R}) \triangleq \inf_{\mathbf{x}\in\mathcal{C}}\max_{r\in\mathcal{R}} |x_r|^2$. Moreover, (\[mainlemmaeq1\]) holds for any relay normalization factor $\rho_r' \leq \rho_r,\,\forall r$. Since all of our main results will be based on the lower bound in (\[mainlemmaeq1\]), and a different relay normalization factor will not improve this lower bound as stated in the lemma, we fix $\rho_r$ to be our relay normalization factor for the rest of the paper. Before we discuss the consequences of Lemma \[mainlemma\] regarding the necessity of relay selection, let us first present a motivating example application. As an immediate corollary to Lemma \[mainlemma\], the following theorem provides an upper bound on the diversity provided by any finite-cardinality i-codebook. \[sidetheorem\] For any i-codebook $\mathcal{C}$ with $|\mathcal{C}|<\infty$, ${\mathtt{d}}(\mathcal{C}) \leq \min\{|\mathcal{R}|:\mathcal{R}\in\mathscr{R}(\mathcal{C})\}$. Since (\[mainlemmaeq1\]) holds for any $\mathcal{R}\in\mathscr{R}(\mathcal{C})$, we choose the set $\mathcal{R}'$ in $\mathscr{R}(\mathcal{C})$ with the smallest cardinality (if the number of such sets is more than one, we can choose any of them). By definition, $\xi(\mathcal{C}, \mathcal{R}')$ is a positive constant that is independent of $P$. It follows from Lemma \[mainlemma\] that $\mathtt{SER}_P(\mathcal{Q}_{P,\mathcal{C}}^{\star}) \geq {C_{0}}[\xi(\mathcal{C},\mathcal{R}')]^{3/2} \exp(-C_1/\xi(\mathcal{C},\mathcal{R}')) P^{-|\mathcal{R}'|},\,\forall P \geq {\Psi_0}$. Thus, $\mathcal{C}$ provides at most a diversity of $|\mathcal{R}'|$. The rest of this section is devoted to some example applications of this theorem. \[example1\] *For a network with $3$ relays, let $$\begin{aligned} \label{cbc1} \mathcal{C}_1 & = \bigl\{{[\begin{array}{ccc} 0 & 1 & 1 \end{array}]}\bigr\}, \\ \label{cbc2} \mathcal{C}_2 & = \bigl\{{[\begin{array}{ccc} 0 & 1 & 1 \end{array}]},\,{[\begin{array}{ccc} 1 & 0 & 1 \end{array}]}\bigr\},\\ \label{cbc3} \mathcal{C}_3 & = \bigl\{{[\begin{array}{ccc} 0 & 1 & 1 \end{array}]},\,{[\begin{array}{ccc} 1 & 0 & 1 \end{array}]},\,{[\begin{array}{ccc} 1 & 1 & 0 \end{array}]}\bigr\}.\end{aligned}$$ Let us first find an upper bound on the diversity provided by $\mathcal{C}_1$. Using the definition in (\[indexsets\]), we have $\mathscr{R}(\mathcal{C}_1) = \{\{2\},\{3\},\{1,2\},\{2,3\},\{1,3\},\{1,2,3\}\}$. Then, according to Theorem \[sidetheorem\], ${\mathtt{d}}(\mathcal{C}_1) \leq 1$ since $$\begin{aligned} \min\{|\mathcal{R}|:\mathcal{R}\in\mathscr{R}(\mathcal{C}_1)\} & = \min\{|\{2\}|,|\{3\}|,|\{1,2\}|,|\{2,3\}|,|\{1,3\}|,|\{1,2,3\}|\} \\ & = \min\{1,1,2,2,2,3\} \\ & = 1.\end{aligned}$$ Similarly, ${\mathtt{d}}(\mathcal{C}_2) \leq 1$ since $\mathscr{R}(\mathcal{C}_2) = \{\{3\},\{1,2\},\{2,3\},\{1,3\},\{1,2,3\}\}$ and thus $\min\{|\mathcal{R}|:\mathcal{R}\in\mathscr{R}(\mathcal{C}_2)\} = 1$. On the other hand, the “best” that we can say about the diversity of $\mathcal{C}_3$ is that ${\mathtt{d}}(\mathcal{C}_3) \leq 2$ since $\mathscr{R}(\mathcal{C}_3) = \{\{1,2\},\{2,3\},\{1,3\},\{1,2,3\}\}$ with $\min\{|\mathcal{R}|:\mathcal{R}\in\mathscr{R}(\mathcal{C}_3)\} = 2$.* \[example2\] *None of the codebooks $\mathcal{C}_1, \mathcal{C}_2$ and $\mathcal{C}_3$ in Example \[example1\] can achieve the maximal diversity order $3$. Now, suppose that a finite-cardinality i-codebook $\mathcal{C}_4$ achieves diversity $3$. Then, we should have $\{1,2\},\{1,3\},\{2,3\}\notin\mathscr{R}(\mathcal{C}_4)$ (otherwise, if e.g., $\{1,2\}\in\mathscr{R}(\mathcal{C}_4)$, then according to Theorem \[sidetheorem\], ${\mathtt{d}}(\mathcal{C}_4) \leq 2$). Now, since $\{1,2\}\notin\mathscr{R}(\mathcal{C}_4)$, by the definition of $\mathscr{R}(\cdot)$ in (\[indexsets\]), $\exists\mathbf{x}=[x_1\, x_2 \,x_3]\in\mathcal{C}_4$ s.t. $|x_1| = |x_2| = 0$. Also, as a result of Proposition \[codebookoptimalitylemma\], $|x_3| = 1$, and thus $\exists\theta_3\in\mathbb{R}$ s.t. $x_3 = e^{j\theta_3}$. In other words, $\mathbf{x} = \mathbf{e}_3(\theta_3)$ is an SRS vector. Similarly, using the conditions that $\{1,3\}\notin\mathscr{R}(\mathcal{C}_4)$ and $\{2,3\}\notin\mathscr{R}(\mathcal{C}_4)$, we can show that $\exists\theta_2\in\mathbb{R},\,\mathbf{e}_2(\theta_2)\in\mathcal{C}_4$ and $\exists\theta_1\in\mathbb{R},\,\mathbf{e}_1(\theta_1)\in\mathcal{C}_4$, respectively. Therefore, only if $\exists\mathcal{C}_{\mathtt{SRS}}\in\mathscr{C}_{\mathtt{SRS}}$ s.t. $\mathcal{C}_{\mathtt{SRS}}\subset\mathcal{C}_4$, we can have ${\mathtt{d}}(\mathcal{C}_4) = 3$. But, we also know from [@koyuncu1] that $\forall\mathcal{C}_{\mathtt{SRS}}\in\mathscr{C}_{\mathtt{SRS}}$, if $\mathcal{C}_{\mathtt{SRS}}\subset\mathcal{C}_4$, we have ${\mathtt{d}}(\mathcal{C}_4) = 3$. Hence, for a network with $3$ relays, a finite-cardinality i-codebook can achieve diversity $3$ *if and only if* it contains an SRS codebook. In Section \[secSelecLarge\], we shall generalize this result to networks with any number of relays that employ codebooks with possibly infinite cardinality.* \[example3\] *One of the most surprising conclusions that we can draw from Theorem \[sidetheorem\] is that, unlike a MISO system, in a relay network, (i) the performance of a codebook can significantly vary under unitary transformations, and (ii) the existence of linearly independent codebook vectors do not guarantee maximal diversity. We have demonstrated the latter phenomenon by codebooks $\mathcal{C}_2$ and $\mathcal{C}_3$ in Example \[example1\]. Despite the fact that $\mathcal{C}_2$ and $\mathcal{C}_3$ consist of $2$ and $3$ linearly independent codebook vectors, respectively, we have ${\mathtt{d}}(\mathcal{C}_2) \leq 1$ and ${\mathtt{d}}(\mathcal{C}_3) \leq 2$.* *We now demonstrate the former phenomenon. For that purpose, let $\mathcal{C}\cdot\mathbf{U} \triangleq \{\mathbf{x}\mathbf{U} : \mathbf{x}\in\mathcal{C}\}$ denote the transformation of the codebook $\mathcal{C}$ by a unitary matrix $\mathbf{U}$.* *In this example, we consider networks with a sum-power constraint $P$ on relays. For such networks, the $r$th relay transmits with power $|x_r|^2P$ given a beamforming vector $\mathbf{x}$, and we require $\sum_{r=1}^R |x_r|^2 P \leq P \implies \|\mathbf{x}\| \leq 1$. The sum-power constraint on relays makes sure that if $\mathcal{C}$ is a feasible codebook, then for any unitary matrix $\mathbf{U}$, the codebook $\mathcal{C}\cdot\mathbf{U}$ is also feasible.* *Let us now consider the transformations of the codebook $\mathcal{C}_{\mathtt{SRS}}(\boldsymbol{0})$ in (\[zerolusrs\]) by the unitary matrices $$\begin{aligned} \label{someunitaries} \mathbf{U}_1 = \left[\begin{array}{ccc} \frac{1}{\sqrt{2}} & \frac{1}{\sqrt{2}} & 0 \\ -\frac{j}{\sqrt{2}} & \frac{j}{\sqrt{2}} & 0 \\ 0 & 0 & j \end{array}\right],\mbox{ and } \mathbf{U}_2 = \frac{1}{4}\left[\begin{array}{ccc} 1+j & 1-3j & -\sqrt{2}+j\sqrt{2} \\ -3-j & 1-j & \sqrt{2}+j\sqrt{2} \\ \sqrt{2}+j\sqrt{2} & \sqrt{2}+j\sqrt{2} & 2 + 2j \end{array}\right].\end{aligned}$$ Note that the codebooks $\mathcal{C}_{\mathtt{SRS}}(\boldsymbol{0}) \cdot \mathbf{U}_1$ and $\mathcal{C}_{\mathtt{SRS}}(\boldsymbol{0}) \cdot \mathbf{U}_2$ consist of the rows of $\mathbf{U}_1$ and $\mathbf{U}_2$, respectively.* *For limited feedback MISO systems with independent and identically distributed transmitter-to-receiver channels[@roh1; @love1; @mukkavilli1], the performance of a quantizer codebook is invariant under unitary transformations. Moreover, even in the case of arbitrary channel variances, the diversity of a codebook is preserved under unitary transformations. On the other hand, for our example network, the application of Theorem \[sidetheorem\] yields $\mathtt{d}(\mathcal{C}_{\mathtt{SRS}}(\boldsymbol{0}) \cdot \mathbf{U}_1) \leq 2$, and $\mathtt{d}(\mathcal{C}_{\mathtt{SRS}}(\boldsymbol{0}) \cdot \mathbf{U}_2) \leq 1$, whereas ${\mathtt{d}}(\mathcal{C}_{\mathtt{SRS}}(\boldsymbol{0})) = 3$. In general, unless $\mathbf{U}$ is diagonal, it can be shown that ${\mathtt{d}}(\mathcal{C}_{\mathtt{SRS}}(\boldsymbol{0}) \cdot \mathbf{U}) \leq 2$. Therefore, in relay networks, even the diversity performance of a codebook is not preserved under unitary transformations. This unexpected behavior can be attributed to the non-linear nature of the distortion function as well as the noise amplification at the relays.* The Necessity of SRS {#secSelecLarge} ==================== With Lemma \[mainlemma\] at hand, we can now introduce our results on the necessity of relay selection. In this section in particular, we determine the structure of optimal quantizers that achieve the full-diversity order, $R$. First, we consider the power-independent i-codebooks, and show that every i-codebook that achieves full-diversity necessarily contains the SRS codebook. We then focus on d-codebooks that can be optimized with respect to the power level $P$, and show that an optimal large d-codebook contains an SRS codebook asymptotically as $P$ grows to infinity. The Necessity of SRS - I-Codebooks ---------------------------------- In Example \[example2\] in Section \[secLowerBounds\], we showed that an i-codebook for a network with $3$ relays can achieve full-diversity if and only if it contains an SRS codebook. The following theorem generalizes this result to networks with any number of relays that employ codebooks with possibly infinite cardinality. \[litheorem\] For any i-codebook $\mathcal{C}$, ${\mathtt{d}}(\mathcal{C})= R$ if and only if $\exists\mathcal{C}_{\mathtt{SRS}}\in\mathscr{C}_{\mathtt{SRS}}$ s.t. $\mathcal{C}_{\mathtt{SRS}} \subset \mathcal{C}$. The “if” part was proved in [@koyuncu1]. Here, we prove the “only if” part by contradiction. Suppose there is a compact i-codebook $\mathcal{C}$ with ${\mathtt{d}}(\mathcal{C}) = R$ and $\forall\mathcal{C}_{\mathtt{SRS}}\in\mathscr{C}_{\mathtt{SRS}}$, $\mathcal{C}_{\mathtt{SRS}}$ is not a subset of $\mathcal{C}$. The latter condition implies that $\exists r\in\{1,\ldots,R\},\,\forall\theta_r,\,\mathbf{e}_r(\theta_r) \notin \mathcal{C}$ (as otherwise, $\forall r\in\{1,\ldots,R\},\,\exists\vartheta_r,\,\mathbf{e}_r(\vartheta_r) \in \mathcal{C}$ and thus $\mathcal{C}_{\mathtt{SRS}}([\vartheta_1\cdots\vartheta_R])\subset\mathcal{C}$). In other words, $\mathcal{C}$ does not contain the vector(s) that selects the $r$th relay. Let ${\mathcal{T}_2}\triangleq \{\mathbf{e}_r(\theta):\theta\in\mathbb{R}\}$. Note that ${\mathcal{T}_2}$ is the product of the closure of the unit disk by the all-zero vector of dimension $R-1$. Since all the factor sets are compact, ${\mathcal{T}_2}$ is compact. We now show that $$\begin{aligned} \label{firstcondition} \exists \epsilon > 0,\,\forall\mathbf{t}\in{\mathcal{T}_2},\,\forall\mathbf{x}\in\mathcal{C},\,\|\mathbf{x} - \mathbf{t}\| > \epsilon.\end{aligned}$$ Let $(\mathbf{x}',\mathbf{t}') = \arg\min_{(\mathbf{x},\mathbf{t})\in\mathcal{C}\times{\mathcal{T}_2}} \|\mathbf{x} - \mathbf{t}\|$. The minimum will always exist as $\mathcal{C}\times{\mathcal{T}_2}$ is compact and $f(\mathbf{x},\mathbf{t}) = \|\mathbf{x}-\mathbf{t}\|$ is continuous. Moreover, since $\mathcal{C}\cap{\mathcal{T}_2}=\emptyset$, we have $\mathbf{x}'\notin{\mathcal{T}_2}$, and thus $\|\mathbf{x}'-\mathbf{t}'\| > 0$. Therefore, we can pick e.g. $\epsilon = \frac{1}{2} \|\mathbf{x}'-\mathbf{t}'\| > 0$, and (\[firstcondition\]) will hold. According to (\[firstcondition\]), for any $\mathbf{x}\in\mathcal{C}$, we have $$\begin{aligned} \label{ycond} |x_r - e^{j\theta}|^2 + \sum_{\substack{q=1 \\ q\neq r}}^R |x_q|^2 > \epsilon^2,\,\forall \theta\in\mathbb{R}. \end{aligned}$$ Also, since $\mathbf{x}\in\mathcal{X}$, it follows that $\exists r'\in\mathcal{R},\,|x_{r'}| = 1$. If $r' = r$, we choose $\theta' = \angle x_{r'}$. Then, (\[ycond\])$\implies\sum_{q\neq r} |x_q|^2 > \epsilon^2 \implies \max_{q\neq r} |x_q|^2 > (R-1)^{-1}\epsilon^2$. Otherwise, if $r'\neq r$, then $\exists q\neq r,\,|x_q| = 1 \implies \max_{q\neq r} |x_q|^2 = 1$. In either case, $\max_{q\neq r} |x_q|^2 > \epsilon_0$, where $\epsilon_0\triangleq \min\{1, (R-1)^{-1}\epsilon^2\} > 0$. Now, let $\mathcal{U} = \{1,\ldots,R\} - \{r\}$. Clearly, $\mathcal{U}\in\mathscr{R}(\mathcal{C})$. Moreover, $\xi(\mathcal{C},\mathcal{U}) = \inf_{\mathbf{x}\in\mathcal{C}} \max_{q\neq r} |x_q|^2 \geq \epsilon_0$. Using Lemma \[mainlemma\], $\mathtt{SER}_P(\mathcal{Q}^{\star}_{P,\mathcal{C}}) \geq {C_{0}}\epsilon_0^{3/2}\exp(-C_1/\epsilon_0)P^{-|\mathcal{U}|},\,\forall P \geq {\Psi_0}$. Therefore, ${\mathtt{d}}(\mathcal{C}) \leq |\mathcal{U}| = R-1$, which contradicts the assumption that ${\mathtt{d}}(\mathcal{C}) = R$. This concludes the proof. Therefore, an i-codebook can achieve diversity $R$ if and only if it contains an SRS codebook. The Necessity of SRS - D-Codebooks ---------------------------------- Let us now consider the necessity of SRS for power-dependent d-codebooks. In this paper, we are interested in the asymptotic structure of optimal d-codebooks as $P$ grows to infinity. As we have mentioned in Section \[secDataTransmisson\], we can interpret any d-codebook $\mathcal{D}$ as a set-valued map that maps the power level $P\in\mathbb{R}^+$ to the codebook $\mathcal{D}_P\subset\mathcal{X}\subset\mathbb{C}^R$. Therefore, we use the well-established limit definitions for set-valued maps[@sva] to characterize the asymptotic structure (or the limit) of any d-codebook. Let $\mathcal{D}$ be a d-codebook. For any $P$ and $\mathbf{x}\in\mathbb{C}^R$, let $$\begin{aligned} \label{compactminimumlanbnu} d_P(\mathbf{x}) & \triangleq \min_{\mathbf{y}\in\mathcal{D}_P} \|\mathbf{x} - \mathbf{y}\|\end{aligned}$$ as the distance of $\mathbf{x}$ to $\mathcal{D}_P$. The minimum in (\[compactminimumlanbnu\]) always exists since $\mathcal{D}_P$ is compact for all $P$. We now define $$\begin{aligned} \limsup_{P \rightarrow\infty} \mathcal{D}_P \triangleq \left\{\mathbf{x}\in\mathbb{C}^R:\liminf_{P\rightarrow\infty} d_P(\mathbf{x}) = 0\right\}\end{aligned}$$ as the *upper limit* of $\mathcal{D}_P$ as $P \rightarrow \infty$, and $$\begin{aligned} \liminf_{P \rightarrow\infty} \mathcal{D}_P \triangleq \left\{\mathbf{x}\in\mathbb{C}^R:\lim_{P\rightarrow\infty} d_P(\mathbf{x}) = 0\right\}\end{aligned}$$ as the *lower limit* of $\mathcal{D}_P$ as $P \rightarrow \infty$. The upper and lower limits always exist for any given $\mathcal{D}$. If $\liminf_{P \rightarrow\infty} \mathcal{D}_P = \limsup_{P \rightarrow\infty} \mathcal{D}_P = \mathcal{L}$, i.e. if the upper and lower limits agree, we say that the d-codebook converges to $\mathcal{L}$ and write $\lim_{P\rightarrow\infty}\mathcal{D}_P = \mathcal{L}$. We also use the shorthand notation $\liminf \mathcal{D} \triangleq \liminf_{P\rightarrow\infty}\mathcal{D}_P$, $\limsup\mathcal{D} \triangleq \limsup_{P\rightarrow\infty}\mathcal{D}_P$, and similarly, $\lim\mathcal{D} \triangleq \lim_{P\rightarrow\infty}\mathcal{D}_P $. Given Theorem \[litheorem\], we expect intuitively that the limit of any full-diversity d-codebook $\mathcal{D}$ contains an SRS codebook provided that $\lim\mathcal{D}$ exists. The following theorem, whose proof can be found in Appendix \[proofofldtheorem\], verifies this intuition: \[ldtheorem\] The following arguments hold for any d-codebook $\mathcal{D}$ with ${\mathtt{d}}({\mathcal{D}}) = R$. 1. There are $R$ distinct beamforming vectors $\widetilde{\mathbf{e}}_{r,P},\,r=1,\ldots,R$ in $\mathcal{D}_P$ s.t. for all $P>{\Psi_3}$, $$\begin{aligned} \label{rateofconv} |\widetilde{e}_{r,P,q}|^2 \leq \frac{{C_3}}{\log P},\,\forall q\in\{1,\ldots,R\}-\{r\},\,r=1,\ldots,R,\end{aligned}$$ where $\widetilde{e}_{r,P,q}$ represents the $q$th component of $\widetilde{\mathbf{e}}_{r,P}$, and $0<{\Psi_3}, {C_3}<\infty$ are constants that are independent of $P$ and $\mathcal{D}$. 2. If $\lim\mathcal{D}$ exists, 1. If $|\mathcal{D}| = R$, $\exists\mathcal{C}_{\mathtt{SRS}}\in\mathscr{C}_{\mathtt{SRS}}$ s.t. $\mathcal{C}_{\mathtt{SRS}}=\lim \mathcal{D}$. 2. If $|\mathcal{D}| > R$, $\exists\mathcal{C}_{\mathtt{SRS}}\in\mathscr{C}_{\mathtt{SRS}}$ s.t. $\mathcal{C}_{\mathtt{SRS}}\subset \lim \mathcal{D}$. 3. If $\lim \mathcal{D}$ does not exist, $\exists\mathcal{C}_{\mathtt{SRS}}\in\mathscr{C}_{\mathtt{SRS}}$ s.t. $\mathcal{C}_{\mathtt{SRS}}\subset \limsup \mathcal{D}$. Since we can achieve full-diversity using the SRS scheme, any SER-optimal large d-codebook should achieve full-diversity as well. In that sense, the necessary conditions that we have stated in Theorem \[ldtheorem\] hold for optimal large d-codebooks as well. Also note that the rate of convergence indicated in (\[rateofconv\]) is only a necessary condition. In other words, a sequence of codebooks satisfying (\[rateofconv\]) do not necessarily provide maximal diversity. We conjecture that a necessary and sufficient rate of convergence is $\frac{1}{P}$ instead of $\frac{1}{\log P}$ stated in the theorem. In Theorem \[ldtheorem\], we have also taken into account codebooks that may fail to converge. This is not a limitation of the analysis that has been carried out: There exists optimal d-codebooks that do not converge as we demonstrate by the following proposition: For any $R$ and $B <\infty$, there exists an optimal d-codebook $\mathcal{D}$ with $|\mathcal{D}| = 2^B$ that does not converge. We prove the proposition for the trivial case $R=1$ and give a sketch of the proof for $R > 1$. For $R=1$, the received SNR is given by $$\begin{aligned} \mathtt{SNR}_P(\mathbf{x},\mathbf{h}) = \frac{|x_1|^2|f_1|^2|g_1|^2 P_0P_1}{1 + |f_1|^2P_0 + |g_1|^2P_1}. \end{aligned}$$ Hence, at any $P$, it is sufficient to use a single beamforming “vector” $[x]$ with $|x| = 1$ to achieve the best SER performance; it is needless to use a codebook with cardinality greater than $1$. As an example, a d-codebook $\mathcal{D}$ with $\mathcal{D}_P = \{[1]\},\,\forall P$ is SER-optimal. Let us now define another d-codebook $\mathcal{D}'$ as $\mathcal{D}_P' = \{[-1]\}$ if $n \leq P < n+1$ for some nonnegative integer $n$, and $\mathcal{D}_P' = \{[1]\}$, otherwise. Note that $\mathcal{D}'$ provides the same SER performance as $\mathcal{D}$. On the other hand, it is straightforward to show that $\limsup \mathcal{D}' = \{[1],\,[-1]\}$, and $\liminf \mathcal{D}' = \emptyset$, and hence $\lim\mathcal{D}'$ does not exist. In general, for any $R$ and a finite $B$, we can synthesize a non-convergent optimal d-codebook out of a convergent optimal d-codebook $\mathcal{D}$ as follows: If $n \leq P < n+1$ for some nonnegative integer $n$, we replace a beamforming vector $\mathbf{x}\in\mathcal{D}_P$ by $e^{j\theta}\mathbf{x}$ for some $\theta\in\mathbb{R}$, and otherwise, leave it unchanged. As a result of Proposition \[snrinvariance\], the new d-codebook provides the same performance as $\mathcal{D}$, and is thus optimal. However, it fails to have a limit due to the artificial phase oscillations that we have introduced. In this section, we have shown by Theorems \[litheorem\] and \[ldtheorem\] that one needs to include all the SRS vectors to the quantizer codebook to achieve full-diversity. This requires the accommodation of a large codebook, or equivalently, at least $\lceil \log_2 R \rceil$ bits of feedback. On the other hand, the design constraints might require that the number of available feedback bits is less than $\lceil \log_2 R \rceil$, in which case we are restricted to using small codebooks and full-diversity is no longer achievable. Our goal in the next section is to determine the optimal codebook structure for such low feedback rate applications. Small Codebooks and the Necessity of OMRS {#secSelecSmall} ========================================= In this section, we first determine the maximal achievable diversity with small codebooks. Then, we find the optimal small codebook structure that can achieve maximal diversity. We show that a diversity-optimal small i-codebook should contain multiple-relay selection vectors that are pairwise orthogonal, i.e. it should be an OMRS codebook. We also demonstrate the necessity of OMRS for small d-codebooks. Finally, we observe that SRS is actually a special case of OMRS for codebooks with cardinality equal to $R$. Therefore, OMRS becomes the universal necessary condition for codebook optimality. Diversity Limitations of Small Codebooks ---------------------------------------- The following theorem shows that the maximal diversity provided by any small codebook is equal to the cardinality of the codebook. \[smallmaxdiversitytheorem\] For any d-codebook $\mathcal{D}$, ${\mathtt{d}}(\mathcal{D}) \leq \min\{R, |\mathcal{D}|\}$. Let us first prove the theorem for an i-codebook $\mathcal{C}$. For any $\mathbf{x}\in\mathcal{X}$, let $\iota(\mathbf{x})\in\{1,\ldots,R\}$ be any index with $|x_{\iota(\mathbf{x})}| = 1$. Note that as a result of Proposition \[codebookoptimalitylemma\], $\iota(\mathbf{x})$ always exists whenever $\mathbf{x}$ is a member of a quantizer codebook. Now let ${\mathcal{T}_4}= \{\iota(\mathbf{x}):\mathbf{x}\in\mathcal{C}\}$. Note that $|{\mathcal{T}_4}| \leq \min\{R, |\mathcal{C}|\}$, ${\mathcal{T}_4}\in\mathscr{R}(\mathcal{C})$, and $\xi(\mathcal{C}, {\mathcal{T}_4}) =1$. Using Lemma \[mainlemma\], $\forall P \geq {\Psi_0}$, we have $$\begin{aligned} \mathtt{SER}_P(\mathcal{Q}_{P, \mathcal{C}}^{\star}) & \geq {C_4}P^{-|{\mathcal{T}_4}|} \\ \label{smalllowboundd}& \geq {C_4}P^{-\min\{R, |\mathcal{C}|\}},\end{aligned}$$ where ${C_4}\triangleq {C_{0}}\exp(-C_1)$. Thus, ${\mathtt{d}}(\mathcal{C})\leq \min\{R, |\mathcal{C}|\}$, concluding the proof for i-codebooks. One way to deal with the complications that arise from the power-dependency of d-codebooks is to define a lower bound that treats each codebook $\mathcal{D}_{\varrho},\,\varrho\in\mathbb{R}^+$ as an i-codebook. At a given $P$, we can calculate the SERs of all $\mathcal{D}_{\varrho},\,\varrho\in\mathbb{R}^+$. The infimum of these SERs then gives us a lower bound on the performance of $\mathcal{D}$ at $P$. With this observation, $\forall P \geq {\Psi_0}$, we have $$\begin{aligned} \mathtt{SER}_P(\mathcal{Q}_{P\!,\,\mathcal{D}_P}^\star) & \geq \inf_{\mathcal{D}_{\varrho}:\varrho\in\mathbb{R}^+} \mathtt{SER}_P(\mathcal{Q}_{P\!,\,\mathcal{D}_{\varrho}}^\star) \\ \label{infimumfun} & \geq \inf_{\mathcal{D}_{\varrho}:\varrho\in\mathbb{R}^+} {C_4}P^{-\min\{R, |\mathcal{D}|\}} \\ & = {C_4}P^{-\min\{R, |\mathcal{D}|\}},\end{aligned}$$ where (\[infimumfun\]) follows from (\[smalllowboundd\]). This concludes the proof. There are structured small codebooks that can achieve the diversity upper bound in Theorem \[smallmaxdiversitytheorem\]. As an example, for an i-codebook $\mathcal{C}_{\mathtt{SRS}}'(d,\boldsymbol{\theta}) = \{\mathbf{e}_i(\theta_i),\,i=1,\ldots,d\}$ that contains $d < R$ SRS vectors, ${\mathtt{d}}(\mathcal{C}_{\mathtt{SRS}}'(d,\boldsymbol{\theta})) = d,\,\forall\boldsymbol{\theta}$, as shown in [@koyuncu1]. In other words, an “incomplete” SRS scheme, in which the selection of only a subset of the relays is considered, can achieve maximal diversity. What is left is thus to determine the structure of a general diversity-optimal small codebook. Unlike large codebooks where SRS is the only way to achieve maximal diversity, we show in the following that for small codebooks, a more general OMRS structure can potentially provide maximal diversity. OMRS ---- The necessity of SRS for large codebooks “generalizes” to the necessity of OMRS for small codebooks. Let us first describe what we mean by OMRS in a more formal manner. An i-codebook $\mathcal{C}$ is an OMRS codebook if and only if either $|\mathcal{C}| = 1$, or $\forall\mathbf{x},\mathbf{y}\in\mathcal{C},\,\mathbf{y}\neq\mathbf{x},\,\sum_{r=1}^R |x_r||y_r| = 0$. OMRS is the scheme induced by an OMRS codebook. In other words, an OMRS codebook contains multiple-relay selection vectors that are pairwise orthogonal.[^2] As an example, $\mathcal{C}_5 = \{[0\;1\;0\;0.8], [0\;0\;1\;0], [1\;0\;0\;0]\}$ is an OMRS codebook. By definition, the cardinality of an OMRS codebook cannot be more than $R$. An OMRS codebook that has cardinality equal to $R$ should be familiar: it is an SRS codebook. The Necessity of OMRS - I-Codebooks ----------------------------------- Now let us demonstrate the necessity of OMRS for i-codebooks by the following theorem: \[sitheorem\] A diversity-optimal i-codebook $\mathcal{C}$ with $|\mathcal{C}| \leq R$ is an OMRS codebook. The case $|\mathcal{C}|=1$ is trivial. We prove the other cases by contradiction. Suppose that there exists a non-OMRS i-codebook $\mathcal{C}$ with $1 < |\mathcal{C}| \leq R$ and ${\mathtt{d}}(\mathcal{C}) = |\mathcal{C}|$. Since $\mathcal{C}$ is not an OMRS, $\exists\mathbf{x},\mathbf{y}\in\mathcal{C},\,\mathbf{y}\neq\mathbf{x},\,\exists r\in\{1,\ldots,R\},\,|x_r| \neq 0 ,\,|y_r|\neq 0$. Now, let ${\mathcal{T}_5}= \{r\}\cup\{\iota(\mathbf{z}): \mathbf{z}\in\mathcal{C} - \{\mathbf{x},\mathbf{y}\}\}$, where $\iota(\mathbf{z})$ is any index that satisfies $|z_{\iota(\mathbf{z})}| = 1$. Note that $|{\mathcal{T}_5}| \leq |\mathcal{C}| - 1$, ${\mathcal{T}_5}\in\mathscr{R}(\mathcal{C})$, and $\xi(\mathcal{C}, {\mathcal{T}_5}) = \min\{|x_r|, |y_r|\}$. Applying Lemma \[mainlemma\], we have ${\mathtt{d}}(\mathcal{C}) \leq |\mathcal{C}| - 1$. This contradicts the assumption that ${\mathtt{d}}(\mathcal{C}) = |\mathcal{C}|$. In other words, an i-codebook $\mathcal{C}$ with $|\mathcal{C}| \leq R$ achieves diversity $|\mathcal{C}|$ only if it is an OMRS codebook. In particular, if $|\mathcal{C}| = R$, ${\mathtt{d}}(\mathcal{C}) = R$ if and only if $\mathcal{C}$ is an OMRS codebook, in which case it is also an SRS codebook. Unlike the necessity and sufficiency of SRS for large codebooks, we can only show the necessity of OMRS for small codebooks. We leave the sufficiency as a conjecture: If a small i-codebook $\mathcal{C}$ is an OMRS codebook, ${\mathtt{d}}(\mathcal{C}) = |\mathcal{C}|$. The Necessity of OMRS - D-Codebooks ----------------------------------- Let us now generalize our result on the necessity of OMRS for i-codebooks to d-codebooks by the following theorem. Its proof can be found in Appendix \[proofofsdtheorem\]. \[sdtheorem\] Let $\mathscr{O}(c)$ denote the collection of all possible OMRS codebooks with cardinality $c$. The following arguments hold for any optimal d-codebook $\mathcal{D}$ with $1 \leq |\mathcal{D}| \leq R$. 1. There are constants $0<{C_6},\,{\Psi_{6}}< \infty$ that are independent of $P$ and $\mathcal{D}$ s.t. for all $P > {\Psi_{6}}$, $$\begin{aligned} \label{sdtheoremeq} \max_{\substack{\mathbf{x},\mathbf{y}\in\mathcal{D}_P \\ \mathbf{x}\neq\mathbf{y}}} \sum_{r=1}^R |x_r||y_r| \leq \frac{{C_6}}{\log P}.\end{aligned}$$ 2. If $\lim \mathcal{D}$ exists, $\exists\mathcal{O}\in\mathscr{O}(|\mathcal{D}|)$ s.t. $\mathcal{O}= \lim \mathcal{D}$. 3. If $\lim \mathcal{D}$ does not exist, $\exists\mathcal{O}\in\mathscr{O}(|\mathcal{D}|)$ s.t. $\mathcal{O}\subset \limsup \mathcal{D}$. Therefore, any two distinct beamforming vectors in an optimal d-codebook $\mathcal{D}$ with $|\mathcal{D}| \leq R$ are asymptotically orthogonal, and thus $\mathcal{D}$ converges asymptotically to an OMRS codebook. In particular, for codebooks with cardinality equal to $R$, Theorem \[sdtheorem\] provides the same arguments as Theorem \[ldtheorem\]. This follows from our previous observation that an OMRS codebook with cardinality equal to $R$ is also an SRS codebook. From all the results that we have obtained up to now, we can conclude that OMRS is a universal necessary condition in the sense that for any SER-optimal d-codebook $\mathcal{D}$, there exists $\mathcal{O}\in\mathscr{O}(\min\{R,|\mathcal{D}|\})$ s.t. $\mathcal{O}\subset\limsup \mathcal{D}$. In other words, as $P$ grows to infinity, the upper limit of every optimal codebook should contain an OMRS codebook with the largest possible cardinality. Numerical Results {#secNumerical} ================= In this section, we provide numerical evidence regarding the validity of our analytical results. For all the figures, the horizontal and the vertical axes represent $P$, and the SER, respectively. Diversity Bounds for Finite-Cardinality I-Codebooks --------------------------------------------------- In Fig. \[letsimfig1\], we show the simulation results with i-codebooks for a $3$-relay network with power constraints $p_0 = 1$, $p_1 = 0.5$, $p_2 = p_3 = 2$, and channel variances $\sigma_{f_1}^2 = 1.2$, $\sigma_{f_2}^2 = 0.8$, $\sigma_{f_3}^2 = 1$, $\sigma_{g_1}^2 = 1.5$, $\sigma_{g_2}^2 = 1.7$, $\sigma_{g_3}^2 = 0.7$. The codebooks $\mathcal{C}_1$, $\mathcal{C}_2$, and $\mathcal{C}_3$ are as defined in (\[cbc1\]), (\[cbc2\]), and (\[cbc3\]), respectively. $\mathcal{O}_1 = \{\,[1\,\,0\,\,0],\,[0\,-\!\!0.8\,\,1]\}$ is an OMRS codebook, and $\mathcal{C}_{\mathtt{SRS}}$ represents an arbitrary SRS codebook. $\mathcal{C}_{\mathtt{SRS}}\cdot\mathbf{U}_1$ and $\mathcal{C}_{\mathtt{SRS}}\cdot\mathbf{U}_2$ represent the transformations of an arbitrary SRS codebook by the unitary matrices $\mathbf{U}_1$ and $\mathbf{U}_2$ in (\[someunitaries\]), respectively. Note that all SRS codebooks provide the same SER at any given $P$, as we have discussed in Section \[secProblemStatement\] and as shown by Proposition \[snrinvariance\]. Similarly, given any unitary matrix $\mathbf{U}$, all the codebooks $\mathcal{C}_{\mathtt{SRS}}\cdot\mathbf{U},\,\mathcal{C}_{\mathtt{SRS}} \in \mathscr{C}_{\mathtt{SRS}}$ provide the same SER at any given $P$. We can observe from Fig. \[letsimfig1\] that ${\mathtt{d}}(\mathcal{C}_1) \leq 1$, ${\mathtt{d}}(\mathcal{C}_2) \leq 1$, and ${\mathtt{d}}(\mathcal{C}_3) \leq 2$, verifying Theorem \[sidetheorem\]. Moreover, both codebooks seem to actually achieve their diversity bounds dictated by the theorem. This suggests that Theorem \[sidetheorem\] also provides an accurate estimate on the diversity of any finite cardinality codebook. Also, $\mathcal{O}_1$ yields second order diversity as we have conjectured, and $\mathcal{C}_{\mathtt{SRS}}$ provides full-diversity. We have analytically shown earlier in Example \[example3\] that unlike a MISO system, in a relay network, (i) the performance of a codebook can significantly vary under unitary transformations, and (ii) the existence of linearly independent codebook vectors do not guarantee maximal diversity. Regarding the latter phenomenon, Fig. \[letsimfig1\] demonstrates that even though $\mathcal{C}_2$ and $\mathcal{C}_3$ consist of $2$ and $3$ linearly independent codebook vectors, respectively, we have ${\mathtt{d}}(\mathcal{C}_2) \leq 1$ and ${\mathtt{d}}(\mathcal{C}_3) \leq 2$. For the former phenomenon, despite the fact that ${\mathtt{d}}(\mathcal{C}_{\mathtt{SRS}}) = 3$, we have ${\mathtt{d}}(\mathcal{C}_{\mathtt{SRS}} \cdot \mathbf{U}_2) \leq 1$ and ${\mathtt{d}}(\mathcal{C}_{\mathtt{SRS}} \cdot \mathbf{U}_1) \leq 2$, as we can infer from Fig. \[letsimfig1\]. Hence, in relay networks, even the diversity provided by a codebook is not preserved under unitary transformations. As a final remark for this set of simulations, we would like to note that we have chosen the power constraint scalers and channel variances in a random manner so as to demonstrate the validity of our results in “asymmetric” scenarios. We have obtained similar results for other (including identical) choices of these parameters. The Necessity of SRS {#simthenecessofrs} -------------------- Let us now demonstrate the validity of Theorems \[litheorem\] and \[ldtheorem\] for a network with $R=2$. We assume that the power constraint scalers and the channel variances of the network are equal to unity. In this set of simulations, we use a special type of codebook that we define in what follows: For any $0 \leq \epsilon \leq 1$, let $$\begin{aligned} \label{specialcodebookdefi} \widetilde{\mathcal{C}}(\epsilon, r) & \triangleq \{\mathbf{x}:\mathbf{x}\in\mathcal{X},\,|x_r|^2 \geq \epsilon\}.\end{aligned}$$ In Fig. \[letsimfig2\], we show the SERs for our $2$-relay network with i-codebooks $\widetilde{\mathcal{C}}(\epsilon, r),\,\epsilon = 1,\,\frac{1}{4},\,\frac{1}{16},\,r=1,2$, $\mathcal{C}_{\mathtt{SRS}}$, $\mathcal{X}$, and the d-codebooks $\smash{\widetilde{\mathcal{D}}}_r \triangleq \widetilde{\mathcal{C}}(\frac{1}{\log P}, r),\,r=1,2$. Note that, as a result of our choice of the network parameters, the SER with $\smash{\widetilde{\mathcal{C}}}(\epsilon,1)$ is the same as the SER with $\smash{\widetilde{\mathcal{C}}}(\epsilon,2)$ at any given $P$. Similarly, the SER with $\smash{\widetilde{\mathcal{D}}}_1$ is the same as the SER with $\smash{\widetilde{\mathcal{D}}}_2$ at any given $P$. We first demonstrate the validity of Theorem \[litheorem\]. Since $\{r\}\in\mathscr{R}(\smash{\widetilde{\mathcal{C}}}({\epsilon,r}))$, and $\xi(\smash{\widetilde{\mathcal{C}}}(\epsilon,r), \{r\}) = \epsilon$, by Lemma \[mainlemma\], $\smash{\widetilde{\mathcal{C}}}(\epsilon,r)$ provides at most a diversity of $1$ for any fixed $\epsilon>0$. This is precisely what we observe in Fig. \[letsimfig2\]. In general, we expect a similar behavior for any given $\epsilon > 0$. Thus, if we use an i-codebook $\mathcal{C}$ with either $\mathcal{C}\subset\smash{\widetilde{\mathcal{C}}}(\epsilon, 1)$ or $\mathcal{C}\subset\smash{\widetilde{\mathcal{C}}}(\epsilon, 2)$ for some $\epsilon > 0$, $\mathcal{C}$ will not be able to provide diversity more than $1$. In other words, if $\mathcal{C}\subset\bigcup_{\epsilon > 0} \widetilde{\mathcal{C}}(\epsilon, 1)$ or $\mathcal{C}\subset\bigcup_{\epsilon > 0} \widetilde{\mathcal{C}}(\epsilon, 2)$, then ${\mathtt{d}}(\mathcal{C}) \leq 1$. Hence, if $\mathcal{C}^{\star}$ is an i-codebook that achieves diversity $2$, then $\exists\mathbf{e}_1^{\star},\mathbf{e}_2^{\star}\in\mathcal{C}^{\star}$ s.t. $$\begin{aligned} \mathbf{e}_1^{\star} \in \left[ \bigcup_{\epsilon > 0} \widetilde{\mathcal{C}}(\epsilon, 1) \right]^c = \bigcap_{\epsilon > 0} \left[ \widetilde{\mathcal{C}}(\epsilon, 1) \right]^c = \bigcap_{\epsilon > 0}\{\mathbf{x}\in\mathcal{X}:|x_1|^2 < \epsilon\} = \{\mathbf{x}\in\mathcal{X}:|x_1|=0\},\end{aligned}$$ and $\mathbf{e}_2^{\star} \in \{\mathbf{x}\in\mathcal{X}:|x_2|=0\}$, where $\mathcal{C}^c \triangleq \mathcal{X} - \mathcal{C}$. Note that $\mathbf{e}_1^{\star}$ and $\mathbf{e}_2^{\star}$ are SRS vectors. Therefore, if $\mathcal{C}^{\star}$ achieves full diversity, it should contain an SRS codebook. This verifies Theorem \[litheorem\]. The verification of Theorem \[ldtheorem\] is analogous: Let $\mathcal{D}^\star$ denote an optimal d-codebook, and $\epsilon_P \triangleq \sup \{\epsilon: \mathcal{D}_P^{\star} \subset \smash{\widetilde{\mathcal{C}}}(\epsilon, 1)\mbox{ or } \mathcal{D}_P^{\star} \subset \smash{\widetilde{\mathcal{C}}}(\epsilon, 2)\}$. Since $\mathcal{D}^\star$ is an optimal d-codebook, it achieves full-diversity. Thus, using the same arguments in the previous paragraph, $\epsilon_P\rightarrow 0$ as $P\rightarrow\infty$. On the other hand, by the definition of $\epsilon_P$, we have $\exists\mathbf{x}_{r,P}^{\star} \in \mathcal{D}_P^{\star}$ s.t. $|x_{r,P,r}^{\star}|^2 \leq \epsilon_P + \epsilon_{r,P}',\,r=1,2$, where $\epsilon_{r,P}' > 0$ can be chosen arbitrarily. Let us choose $\epsilon_{r,P}' = \epsilon_P,\,r=1,2$. Then, we have $|x_{r,P,r}^{\star}|^2 \leq 2\epsilon_P,\,r=1,2$. This shows the existence of two beamforming vectors in $\mathcal{D}_P^{\star}$, namely $\mathbf{x}_{1,P}^{\star}$ and $\mathbf{x}_{2,P}^{\star}$, that converges to two distinct SRS vectors as $P\rightarrow\infty$. This verifies the limit arguments in Theorem \[ldtheorem\]. Theorem \[ldtheorem\] also provides an estimate on how fast $\epsilon_P$ should decay. The performance of the d-codebooks $\widetilde{\mathcal{D}}_1$ and $\widetilde{\mathcal{D}}_2$ in Fig. \[letsimfig2\] demonstrate that the decay should be no slower than $\tfrac{1}{\log P}$, and thus verifies the theorem. On the other hand, since both codebooks do not provide maximal diversity, the estimate of Theorem \[ldtheorem\] might be rather loose. The Necessity of OMRS --------------------- We now demonstrate the validity of Theorems \[sitheorem\] and \[sdtheorem\] for a network with $R=3$. We assume that the power constraint scalers and the channel variances of the network are equal to unity. Our goal is to determine the structure of optimal codebooks that have cardinality equal to $2$ and thus provide a diversity of $2$. For that purpose, similar to what we have done in Section \[simthenecessofrs\], we use the special i-codebook $\widetilde{\mathcal{C}}(\epsilon, r)$ as defined in (\[specialcodebookdefi\]). In Fig. \[letsimfig3\], we show the SERs for our $3$-relay network with i-codebooks $\widetilde{\mathcal{C}}(\epsilon, r),\,\epsilon = 1,\,\frac{1}{4},\,\frac{1}{16},\,r=1,2,3$, $\mathcal{C}_{\mathtt{SRS}}$, $\mathcal{X}$, and the d-codebooks $\smash{\widetilde{\mathcal{D}}}_r \triangleq \widetilde{\mathcal{C}}(\frac{1}{\log P}, r),\,r=1,2,3$. As a result of our choice of the network parameters, for a given $\epsilon$, the SERs with $\smash{\widetilde{\mathcal{C}}}(\epsilon,1),\,\smash{\widetilde{\mathcal{C}}}(\epsilon,2)$ and $\smash{\widetilde{\mathcal{C}}}(\epsilon,3)$ are the same at any given $P$. Similarly, the SERs with $\smash{\widetilde{\mathcal{D}}}_1,\,\smash{\widetilde{\mathcal{D}}}_1$ and $\smash{\widetilde{\mathcal{D}}}_3$ are the same at any given $P$. We first demonstrate the validity of Theorem \[litheorem\]. We can observe from Fig. \[letsimfig2\] that $\smash{\widetilde{\mathcal{C}}}(\epsilon,r)$ provides at most a diversity of $1$ for any fixed $\epsilon>0$. In general, we expect a similar behavior for any given $\epsilon > 0$. Thus, if we use an i-codebook $\mathcal{C}$ with $\mathcal{C}\subset\smash{\widetilde{\mathcal{C}}}(\epsilon, r)$ for some $r\in\{1,2,3\}$ and $\epsilon > 0$, then ${\mathtt{d}}(\mathcal{C}) \leq 1$. As a result, using the same arguments in Section \[simthenecessofrs\], if ${\mathtt{d}}(\mathcal{C}^*) = 2$, then $\exists\mathbf{y}_r\in\mathcal{C}^*\mbox{ s.t. }\mathbf{y}_r^* \in \{\mathbf{y}\in\mathcal{X}:|y_r|=0\},\,r=1,2,3$. In other words, for any $r\in\{1,2,3\}$, there exists a beamforming vector in $\mathcal{C}^* = \{\mathbf{x}_1^*, \mathbf{x}_2^*\}$ with a vanishing $r$th component. Therefore, $\sum_{r=1}^3 |x_{1r}^*| |x_{2r}^*| = 0$, which means that $\mathcal{C}^*$ is an OMRS codebook. This verifies Theorem \[sitheorem\]. In order to verify Theorem \[sdtheorem\], let $\mathcal{D}^\star$ with $\mathcal{D}_P^* = \{\mathbf{x}_{1,P}^*,\mathbf{x}_{2,P}^*\},\,P\in\mathbb{R}$ and $|\mathcal{D}^{\star}| = 2$ denote an optimal small d-codebook, and $\epsilon_P \triangleq \sup \{\epsilon: \exists r\in\{1,2,3\}\mbox{ s.t. }\mathcal{D}_P^{\star} \subset \smash{\widetilde{\mathcal{C}}}(\epsilon, r)\}$. Since $\mathcal{D}^\star$ is optimal, it achieves second order diversity. Using the same arguments in the previous paragraph, $\epsilon_P\rightarrow 0$ as $P\rightarrow\infty$. On the other hand, $\exists\mathbf{y}_{r,P}^{\star} \in \mathcal{D}_P^{\star}$ s.t. $|y_{r,P,r}^{\star}| \leq \sqrt{2\epsilon_P},\,r=1,2,3,$ by the definition of $\epsilon_P$. As a result, $\sum_{r=1}^3 |x_{1,r,P}^*| |x_{2,r,P}^*| \leq 6\epsilon_P,\,\forall P\in\mathbb{R}$, and $\sum_{r=1}^3 |x_{1,r,P}^*| |x_{2,r,P}^*| \rightarrow 0$ as $P\rightarrow\infty$. In other words, the two beamforming vectors in $\mathcal{D}_P^*$ should become asymptotically orthogonal. Finally, the performance of the codebooks $\widetilde{\mathcal{D}}_r,\,r=1,2,3$ in Fig. \[letsimfig3\] demonstrate that $\epsilon_P$ should decay no slower than $\frac{1}{\log P}$. These verify Theorem \[sdtheorem\]. Conclusions {#secConc} =========== We have determined some necessary structural properties of symbol error rate optimal quantizers for limited feedback beamforming in wireless networks with a single transmitter-receiver pair and $R$ parallel amplify-and-forward relays. We have shown that any power-independent codebook (i-codebook) necessarily contains an orthogonal multiple-relay selection (OMRS) codebook with the largest possible cardinality. In particular, if the cardinality of the codebook is no less than $R$, an i-codebook achieves maximal diversity if and only if it contains the single-relay selection (SRS) codebook. We have obtained similar results for the general case of power-dependent codebooks (d-codebooks): An optimal d-codebook should contain an OMRS codebook with the largest possible cardinality, asymptotically as the transmitter powers grow to infinity. Proof of Lemma \[mainlemma\] {#proofofmainlemma} ============================ Note that $\mathcal{R}\neq\emptyset$, since if $\mathcal{R}=\emptyset$ then $\mathbf{x} = \mathbf{0},\,\forall\mathbf{x}\in\mathcal{C}$, contradicting Proposition \[codebookoptimalitylemma\]. Using (\[receivedsnr\]), for any set of indices $\mathcal{R}\neq\emptyset$ and relay normalization factors $\rho_r'\leq\rho_r,\,\forall r$, the SNR with any beamforming vector $\mathbf{x}\in\mathcal{C}$, can be upper bounded by $$\begin{aligned} \mathtt{SNR}_P(\mathbf{x}, \mathbf{h}) & \leq \frac{RP_0 \sum_{r=1}^R |x_rf_rg_r|^2 \rho_r'}{1 + \sum_{r=1}^R |x_rg_r|^2\rho_r'} \\ & \leq \frac{RP_0 \sum_{r=1}^R |x_rf_rg_r|^2 \rho_r}{1 + \sum_{r=1}^R |x_rg_r|^2\rho_r} \\ \label{longinek1} & = \frac{R\sum_{r=1}^R \frac{|x_r|^2|f_{r}|^2P_0 |g_{r}|^2 P_r}{1+|f_r|^2 P_0}}{1 + \sum_{r=1}^R \frac{|x_r|^2|g_{r}|^2 P_r}{1+|f_r|^2P_0}}, \end{aligned}$$ where the first and the second inequalities follow from Hölder’s inequality, and the fact that $\rho_r' \leq \rho_r,\,\forall r$, respectively. The proof of the lemma for $R=1$ is now straightforward. If $R=1$, we have $|x_1| = 1,\,\forall \mathbf{x}\in\mathcal{C}$, and thus $\xi(\mathcal{C},\mathcal{R}) = 1$ for any $\mathcal{C}$ and $\mathcal{R}$ (indeed the only available $\mathcal{R}$ will be $\mathcal{R} = \{1\}$). Using (\[longinek1\]), we have $\mathtt{SNR}_P(\mathbf{x}, \mathbf{h}) \leq \frac{|f_1|^2P_0 |g_1|^2P_1}{1 + |f_1|^2P_0 + |g_1|^2P_1} \leq |f_1|^2P_0$. Since this final upper bound is the SNR of a fading channel with single transmitter and receiver antennas, we have $\mathtt{SER}_P(\mathcal{Q}^{\star}_{P, \mathcal{C}}) \geq {C_{7}}P^{-1}$ for some constant $0<{C_{7}}<\infty$ independent of $P$. For $R \geq 2$, we shall further bound $\mathtt{SNR}_P(\mathbf{x}, \mathbf{h})$. For the numerator of (\[longinek1\]), we have $$\begin{gathered} \label{numbound} R \sum_{r=1}^R \frac{|x_r|^2|f_{r}|^2P_0|g_{r}|^2 P_r}{1+|f_r|^2 P_0} \leq R \sum_{r=1}^R \frac{|f_{r}|^2P_0|g_{r}|^2 P_r}{1+|f_r|^2 P_0} \leq R \sum_{r=1}^R |g_{r}|^2 P_r \leq R^2 \max_r |g_r|^2 P_r \\ = R^2 P \max_r \{\sigma_{g_r}^{2}\sigma_{g_r}^{-2}|g_r|^2 p_r\} \leq R^2 \max_r \{p_r\sigma_{g_r}^2\} P \max_r Z_r. \end{gathered}$$ where $Z_r \triangleq \sigma_{g_r}^{-2}|g_r|^2$. Note that $Z_r \sim \Gamma(1,1)$. Now, for the denominator of (\[longinek1\]), $$\begin{gathered} \label{denombound} 1 + \sum_{r=1}^R \frac{|x_r|^2|g_{r}|^2 P_r}{1+|f_r|^2P_0} \geq \sum_{r=1}^R \frac{|x_r|^2|g_{r}|^2 P_r}{1+|f_r|^2P_0} \geq \max_r|x_r|^2 \min_r \frac{|g_{r}|^2 P_r}{1+|f_r|^2P_0} \\ \geq \max_r|x_r|^2 \frac{\min_{r\in\mathcal{R}} |g_r|^2 P_r}{\max_{r\in\mathcal{R}} (1+|f_r|^2 P_0)} \geq \frac{\max_r|x_r|^2 \min_r \{p_r\sigma_{g_r}^2\} P \min_r Z_r}{\max\{1, p_0\max_r\sigma_{f_r}^2\}\left(1 + P \sum_{r\in\mathcal{R}} \sigma_{f_r}^{-2}|f_r|^2\right)}. \end{gathered}$$ Now let $V \triangleq YZ$ with $Y \triangleq \frac{1}{P} + \sum_{r\in\mathcal{R}}\sigma_{f_r}^{-2}|f_{r}|^2 $ and $Z \triangleq \frac{\max_r Z_r }{\min_r Z_r}$. Using (\[denombound\]) and (\[numbound\]) in the final upper bound in (\[longinek1\]), and then taking the supremum over all possible $\mathbf{x}\in\mathcal{C}$, we have $$\begin{aligned} \label{snruboundofaquantizer} \mathtt{SNR}_P(\mathcal{Q}^{\star}_{P, \mathcal{C}}(\mathbf{h}), \mathbf{h}) & \leq {C_{8}}PV,\end{aligned}$$ where ${C_{8}}\triangleq [\xi(\mathcal{C},\mathcal{R})]^{-1}{C_{9}}$ is a finite constant with $\xi(\mathcal{C}, \mathcal{R})$ is as defined in the statement of the lemma, and ${C_{9}}\triangleq R^2 \max\{1,p_0\max_r\sigma_{f_r}^2\}\max_r \{p_r\sigma_{g_r}^2\}/ \min_r \{p_r\sigma_{g_r}^2\}$. Note that since ${C_{9}}\geq 1$ and $\xi(\mathcal{C}, \mathcal{R}) \leq 1$, we have ${C_{8}}\geq 1$. Now, substituting (\[snruboundofaquantizer\]) to (\[serofaquantizer\]), we have $$\begin{aligned} \label{mainlbound} \mathtt{SER}_P(\mathcal{Q}^{\star}_{P, \mathcal{C}}) & \geq \int_{0}^{\infty} \mathrm{Q}(\sqrt{2{C_{8}}vP})f(v)\mathrm{d}v.\end{aligned}$$ In the following, we find a lower bound on the PDF of $V=YZ$. Since $Y$ is the sum of $R$ independent $\Gamma(1,1)$ random variables and a constant $\frac{1}{P}$, it follows a “shifted” gamma distribution: $$\begin{aligned} \label{ybond} f_Y(y) = \frac{1}{\Gamma(|\mathcal{R}|)}e^{-\left(y-\frac{1}{P}\right)} \left(y-\frac{1}{P}\right)^{|\mathcal{R}|-1},\,y \geq \frac{1}{P}.\end{aligned}$$ Now, let us evaluate $f_Z(z)$. Note that for $z<1$, $F_Z(z) = 0$, and thus $f_Z(z) = 0,\,z<1$. For $z\geq 1$, the CDF of $Z$ can be expressed as $F_Z(z) = P(E)$ where $E$ is the event that $\max_r Z_r \leq z \min_r Z_r$, with $Z_r \triangleq |g_r|^2$. Moreover, $E$ is the union of $R(R-1)$ disjoint events $E_{ij},\,i\neq j,\,i,j\in\{1,\ldots, R\}$, where $E_{ij}$ is the event that $Z_i = \min Z_r,\,Z_j = \max Z_r,\, Z_i\in[0,\infty),\,Z_j\in(Z_i,zZ_i],\,Z_k\in(Z_i,Z_j),\,k\neq i,\,k\neq j$.[^3] Since $Z_r$ are identically distributed, and each $E_{ij}$ has the same probability, for $z\geq 1$, we have $$\begin{aligned} \label{aciklaeq1} F_Z (z) & = R(R\!-\!1) \int_0^{\infty} \int_x^{xz} \underbrace{\int_x^y \cdots \int_x^y}_{R-2\,\,\mathrm{integrals}} e^{-x-y-\sum_{i}\!\!w_i} \prod_{i} \mathrm{d}w_i \mathrm{d}y \mathrm{d}x \\ \label{sonradan1} & = R(R\!-\!1) \int_0^{\infty} \int_x^{xz} e^{-x-y} (e^{-x} - e^{-y})^{R-2}\mathrm{d}y \mathrm{d}x \\ \label{sonradan2} & = R(R\!-\!1) \int_0^{\infty} \int_x^{xz} e^{-x-y} \sum_{r=0}^{R-2} {R-2 \choose r} (-1)^r e^{-yr} e^{-x(R-2-r)}\mathrm{d}y \mathrm{d}x\\ \label{aciklaeq2} & = (R-1)\sum_{r=0}^{R-2}{R-2 \choose r} \frac{(-1)^r(z-1)}{R+(z-1)(1+r)}\\ \label{aciklaeq3} & = (z\!-\!1)(R\!-\!1)\int_0^{\infty}e^{-x(R+(z-1))} \sum_{r=0}^{R-2}{R-2 \choose r}(-1)^r e^{-rx(z-1)} \mathrm{d}x \\ \label{aciklaeq4} & = (z\!-\!1)(R\!-\!1)\int_0^{\infty}\!\!e^{-x(R+(z-1))} (1\! - \! e^{-x(z-1)})^{R-2} \mathrm{d}x \\ \label{aciklaeq5} & = 2^{R-2}(z\!-\!1)(R\!-\!1)\int_0^{\infty}\!\!e^{-x\frac{R(z+1)}{2}} \sinh^{R-2}\left[\frac{x(z-1)}{2}\right] \mathrm{d}x \\ \label{aciklaeq6} & = \frac{\Gamma(R)\Gamma(1+\frac{R}{z-1})}{\Gamma(R+\frac{R}{z-1})} \\ \label{aciklaeq7} & = \frac{\Gamma(R)}{ \prod_{r=1}^{R-1} (r + \frac{R}{z-1})},\end{aligned}$$ where (\[sonradan2\]) follows from the binomial expansion of the term $(e^{-x} - e^{-y})^{R-2}$ in (\[sonradan1\]). In order to obtain (\[aciklaeq3\]), we have rewritten the denominator of the fraction in (\[aciklaeq2\]) in integral form by using the identity $\int_0^{\infty}e^{-\alpha x}\mathrm{d}x = \frac{1}{\alpha},\,\alpha > 0$. Also, (\[aciklaeq4\]) is a result of the fact that $\sum_{r=0}^{R-2} (-1)^r \beta^r = (1-\beta)^{R-1}$ for $0\leq\beta\leq 1$, and (\[aciklaeq6\]) follows from [@toi Eq. 3.541.1]. In order to derive (\[aciklaeq7\]) from (\[aciklaeq6\]), we have used the identity $\Gamma(1+x) = x\Gamma(x),\,x\in\mathbb{R}$, which implies $\Gamma(R+\frac{R}{z-1}) = (R-1+\frac{R}{z-1})\Gamma(R-1+\frac{R}{z-1}) = \prod_{r=1}^{R-1} (r+\frac{R}{z-1}) \Gamma(1 + \frac{R}{z-1})$. We can now find the PDF of $Z$ using (\[aciklaeq7\]). We have $$\begin{aligned} f_Z(z) & = \frac{\partial}{\partial z}F_Z(z) \\ & = \frac{-\Gamma(R)}{ \prod_{r=1}^{R-1} (r + \frac{R}{z-1})^2}\frac{\partial}{\partial z}\left\{\prod_{r=1}^{R-1} \left(r + \frac{R}{z-1}\right) \right\} \\ & = \frac{-\Gamma(R)}{ \prod_{r=1}^{R-1} (r + \frac{R}{z-1})^2}\sum_{r=1}^R\frac{\partial}{\partial z}\left\{r + \frac{R}{z-1}\right\}\prod_{\substack{q=1 \\ q\neq r}}^{R-1}\left(q + \frac{R}{z-1}\right) \\ & = \frac{-\Gamma(R)}{ \prod_{r=1}^{R-1} (r + \frac{R}{z-1})^2}\sum_{r=1}^{R-1}\frac{-R}{(z-1)^2}\frac{\prod_{q=1}^{R-1}(q + \frac{R}{z-1})}{r + \frac{R}{z-1}} \\ & = \frac{\Gamma(R+1)\sum_{r=1}^{R-1}(r+\frac{R}{z-1})^{-1}}{(z-1)^2\prod_{r=1}^{R-1}(r+\frac{R}{z-1})} \\ & \geq \frac{\Gamma(R+1)\sum_{r=1}^{R-1}(R+\frac{R}{z-1})^{-1}}{(z-1)^2\prod_{r=1}^{R-1}(R+\frac{R}{z-1})} \\ \label{zbond} & = \frac{(R-1)\Gamma(R+1)}{R^R} \frac{(z-1)^{R-2}}{z^R}.\end{aligned}$$ Now, we find a lower bound on the PDF of $V = YZ$. It can be shown[@aofrvs] that the PDF of $V$ is given by $f_V(v) = \int_{-\infty}^{\infty} f_{Z}(x) f_{Y}\left(\frac{v}{x}\right)\frac{1}{|x|}\mathrm{d}x$. Substituting the PDF of $Y$ in (\[ybond\]), and the lower bound in (\[zbond\]) on the PDF of $Z$, we have $$\begin{aligned} f_V(v) & \geq \frac{(R-1)\Gamma(R+1) }{R^R\Gamma(|\mathcal{R}|)} \! \int_1^{vP}\!\!\left(\frac{v}{x} - \frac{1}{P}\right)^{|\mathcal{R}|-1} \!\! e^{-\left(\frac{v}{x} - \frac{1}{P}\right)} \frac{(x-1)^{R-2}}{x^{R+1}} \mathrm{d}x \\ & \label{coff} = \frac{R(R-1) }{\Gamma(|\mathcal{R}|)} \int_0^{\infty}\left(\frac{vP-1}{(1+w)P}\right)^{|\mathcal{R}|-1} e^{-\frac{vP-1}{(1+w)P}} \left(\frac{w(vP-1)}{vP(1+w)}\right)^{R+1}\left( \frac{w+vP}{w(vP-1)}\right)^3 \mathrm{d}w \\ & \label{cofftansonra} \geq \frac{(R-1)\Gamma(R+1) }{R^R\Gamma(|\mathcal{R}|)} \left(\frac{vP-1}{vP}\right)^{R+|\mathcal{R}|-3}e^{-v}v^{|\mathcal{R}|-1}\int_0^{\infty} \frac{w^{R-2}}{(1+w)^{R+|\mathcal{R}|}}\mathrm{d}w,\end{aligned}$$ where we have applied a change of variables $w = \tfrac{vP(x-1)}{vP-x}$ to obtain (\[coff\]), and (\[cofftansonra\]) follows from the facts that $\exp(-\frac{vP-1}{(1+w)P}) = \exp(-v\frac{vP-1}{(1+w)vP}) \geq \exp(-v)$, and $(\frac{w+vP}{w(vP-1)})^3 \geq \frac{1}{w^3}(\frac{vP}{vP-1})^3$. The integral in (\[cofftansonra\]) is non-zero and finite for $R\geq 2$. Thus, $$\begin{aligned} \label{fvlbound} f_V(v) \geq {C_{10}}\Bigl(\frac{vP-1}{vP}\Bigr)^{R+|\mathcal{R}|-3}e^{-v}v^{|\mathcal{R}|-1},\end{aligned}$$ for some constant $0<{C_{10}}<\infty$. Combining (\[mainlbound\]), (\[fvlbound\]), and using the fact that $\mathrm{Q}(x) \geq \frac{1}{\sqrt{2\pi}}\frac{x}{1+x^2}\exp(-\frac{x^2}{2})$, we have $$\begin{aligned} \mathtt{SER}_P(\mathcal{Q}^{\star}_{P, \mathcal{C}}) & \geq \frac{{C_{10}}}{\sqrt{2\pi}} \int_{\frac{1}{P}}^{\infty} \! v^{|\mathcal{R}|-1}e^{-v} \left(\frac{vP-1}{vP}\right)^{R+|\mathcal{R}|-3} \frac{\sqrt{2{C_{8}}vP}}{1+2{C_{8}}vP}e^{-v{C_{8}}P}\mathrm{d}v \\ & = \frac{{C_{10}}e^{-{C_{8}}-\frac{1}{P}}}{\sqrt{\pi}P^{|\mathcal{R}|}} \int_{0}^{\infty} e^{-w({C_{8}}+\frac{1}{P})}\frac{ w^{R+|\mathcal{R}|-3}}{(1+w)^{R-2}}\frac{\sqrt{{C_{8}}(1+w)}}{2{C_{8}}w+1+2{C_{8}}} \mathrm{d}w \\ & \label{buintegraliburadakullan} \geq \frac{{C_{10}}\sqrt{{C_{8}}}e^{-1-{C_{8}}}}{(1+2{C_{8}})\sqrt{\pi}P^{|\mathcal{R}|}} \int_{0}^{\infty} e^{-2{C_{8}}w}w^{R+|\mathcal{R}|-3}(1+w)^{\frac{1}{2}-R} \mathrm{d}w, $$ where the equality follows from a change of variables $w = vP-1$. The second inequality follows from ${C_{8}}\geq 1$, and the assumption that $P>1$. Now let us find a lower bound for the integral in (\[buintegraliburadakullan\]), i.e. $I \triangleq \int_0^{\infty} e^{-\alpha w} w^{\beta} (1+w)^{-\gamma} \mathrm{d}w$, where $\alpha = 2{C_{8}}$, $\beta = R+|\mathcal{R}|-3$, and $\gamma = R - \frac{1}{2}$. Note that $\alpha,\,\gamma > 0$ and $\beta \geq 0$. We have $$\begin{aligned} I & = \int_0^{1} e^{-\alpha w} w^{\beta} (1+w)^{-\gamma} \mathrm{d}w + \int_1^{\infty} e^{-\alpha w} w^{\beta} (1+w)^{-\gamma} \mathrm{d}w \\ & \geq e^{-\alpha} 2^{-\gamma} \int_0^{1} w^{\beta} \mathrm{d}w + 2^{-\gamma} \int_1^{\infty} e^{-\alpha w} w^{\beta-\gamma} \mathrm{d}w \\ & \geq \frac{e^{-\alpha} 2^{-\gamma}}{1+\beta}+ 2^{-\gamma} \int_1^{\infty} e^{-\alpha w} w^{-\beta-\gamma} \mathrm{d}w \\ & \geq \frac{e^{-\alpha} 2^{-\gamma}}{1+\beta}+ 2^{-\gamma} \int_1^{\infty} e^{-\alpha w} e^{-w(\beta+\gamma)} \mathrm{d}w \\ & = \frac{e^{-\alpha} 2^{-\gamma}}{1+\beta}+ \frac{2^{-\gamma} e^{-(\alpha+\gamma+\beta)}}{\alpha + \gamma + \beta}\\ \label{buintegralbudur} & \geq \frac{2^{1-\gamma} e^{-(\alpha+\gamma+\beta)}}{1 + \alpha + \gamma + \beta}\end{aligned}$$ Substituting the values of $\alpha,\,\beta$ and $\gamma$ to (\[buintegralbudur\]), and combining with (\[buintegraliburadakullan\]), we have $$\begin{aligned} \mathtt{SER}_P(\mathcal{Q}^{\star}_{P, \mathcal{C}}) & \geq \frac{{C_{10}}\sqrt{{C_{8}}}e^{-1-{C_{8}}}}{(1+2{C_{8}})\sqrt{\pi}P^{|\mathcal{R}|}} \frac{2^{\frac{3}{2}-R} e^{-(2{C_{8}}+2R+|\mathcal{R}|-\frac{7}{2})}}{2{C_{8}}+2R+|\mathcal{R}|-\frac{5}{2}}\end{aligned}$$ Since $C_5 \geq 1$, and $2R + |\mathcal{R}| - \frac{5}{2} > 0$ for all $R \geq 2$ and $1 \leq |\mathcal{R}| \leq R$, we have $$\begin{aligned} \label{yettigayri} \mathtt{SER}_P(\mathcal{Q}^{\star}_{P, \mathcal{C}}) & \geq \frac{{C_{10}}\sqrt{{C_{8}}}e^{-1-{C_{8}}}}{3{C_{8}}\sqrt{\pi}P^{|\mathcal{R}|}} \frac{2^{\frac{3}{2}-R} e^{-(2{C_{8}}+2R+R-\frac{7}{2})}}{{C_{8}}(2R+|\mathcal{R}|-\frac{1}{2})}. $$ Finally, using the fact that $2R+|\mathcal{R}|-\frac{1}{2} \leq 3R$ on the denominator of the second fraction in (\[yettigayri\]), we can show that (\[mainlemmaeq1\]) holds for any $R$ with constants ${C_{0}}= \min\{{C_{7}}, {C_{10}}\smash{C_{9}}^{-3/2}\,2^{3/2-R}\,\exp(-3R+\frac{5}{2})/(9R\sqrt{\pi})\}$ and ${C_{1}}= 3{C_{9}}$ that are independent of $\mathcal{C}$ and $P$. This concludes the proof. Proof of Theorem \[ldtheorem\] {#proofofldtheorem} ============================== We carry out the proof in two parts: First we prove (\[rateofconv\]), and then the limit arguments in the statement of the theorem. Proof of (\[rateofconv\]) ------------------------- For notational convenience, let $\vartheta_P \triangleq \mathtt{SER}_P(\mathcal{Q}_{P,\mathcal{D}_P}^{\star})$. Also, let $\mathscr{T} = \{\{1,\ldots,R\}-\{r\},r=1,\ldots,R\}$. Using Lemma \[mainlemma\], we have $$\begin{aligned} \label{logic1} \forall P\geq{\Psi_0},\,\forall\mathcal{T}\in\mathscr{T},\,\biggl\{\xi(\mathcal{D}_P, \mathcal{T}) > \frac{2 C_1}{ \log P} \implies \vartheta_P > \frac{{C_{0}}(2C_1)^{\frac{3}{2}}}{P^{|\mathcal{T}|+\frac{1}{2}} \log^{\frac{3}{2}}P }\biggr\}.\end{aligned}$$ This, using logical transposition, is equivalent to $$\begin{aligned} \label{logic2} \forall P\geq{\Psi_0},\forall \mathcal{T}\in\mathscr{T},\,\biggl\{\vartheta_P \leq \frac{{C_{0}}(2C_1)^{\frac{3}{2}}}{P^{|\mathcal{T}|+\frac{1}{2}} \log^{\frac{3}{2}}P } \implies \xi(\mathcal{D}_P, \mathcal{T}) \leq \frac{2C_1}{ \log P} \biggr\}.\end{aligned}$$ Since $\vartheta_P \leq \frac{{C_{0}}(2C_1)^{\frac{3}{2}}}{P^{R-\frac{1}{2}} \log^{\frac{3}{2}}P } \implies \vartheta_P \leq \frac{{C_{0}}(2C_1)^{\frac{3}{2}}}{P^{|\mathcal{T}|+\frac{1}{2}} \log^{\frac{3}{2}}P },\,\forall P \geq 1$, it follows from (\[logic2\]) that $$\begin{aligned} \label{logic3} \forall P\geq{\Psi_{7}},\forall \mathcal{T}\in\mathscr{T},\,\biggl\{\vartheta_P \leq \frac{{C_{0}}(2C_1)^{\frac{3}{2}}}{P^{R-\frac{1}{2}} \log^{\frac{3}{2}}P } \implies \xi(\mathcal{D}_P, \mathcal{T}) \leq \frac{2C_1}{ \log P} \biggr\},\end{aligned}$$ where ${\Psi_{7}}\triangleq \max\{1,{\Psi_0}\}$. It was shown in [@koyuncu1] that $$\begin{aligned} \forall\mathcal{C}_{\mathtt{SRS}} \in\mathscr{C}_{\mathtt{SRS}},\,\mathtt{SER}_P(\mathcal{Q}_{P\!,\,\mathcal{C}_{\mathtt{SRS}}}^{\star}) \leq {C_{11}}P^{-R},\,\forall P \geq {\Psi_{8}}, \end{aligned}$$ where $0<{C_{11}},{\Psi_{8}}<\infty$ are constants that are independent of $P$. This upper bound on the SER with $\mathcal{C}_{\mathtt{SRS}}$ holds for any optimal codebook of cardinality at least $R$. Thus, $$\begin{aligned} \label{combobmmbob1} \vartheta_P \leq {C_{11}}P^{-R},\,\forall P \geq {\Psi_{8}}. \end{aligned}$$ Moreover, there exists a constant $0 < {\Psi_{9}}< \infty$ that is independent of $P$ s.t. $$\begin{aligned} \label{combobmmbob2} {C_{11}}P^{-R} \leq \frac{{C_{0}}(2C_1)^{\frac{3}{2}}}{P^{R-\frac{1}{2}} \log^{\frac{3}{2}}P },\,\forall P\geq{\Psi_{9}}. \end{aligned}$$ Combining (\[combobmmbob1\]) and (\[combobmmbob2\]), we have $$\begin{aligned} \label{combobmmbob3} \vartheta_P \leq \frac{{C_{0}}(2C_1)^{\frac{3}{2}}}{P^{R-\frac{1}{2}} \log^{\frac{3}{2}}P },\,\forall P\geq\max\{{\Psi_{8}}, {\Psi_{9}}\}.\end{aligned}$$ Letting ${\Psi_{10}}\triangleq \max\{{\Psi_{7}}, {\Psi_{8}}, {\Psi_{9}}\}$, and noting that the left hand side of the implication in (\[logic3\]) does not depend on $\mathcal{R}$, we have $$\begin{aligned} \label{logic4} \forall P\geq{\Psi_{10}},\,\vartheta_P \leq \frac{{C_{0}}(2C_1)^{\frac{3}{2}}}{P^{R-\frac{1}{2}} \log^{\frac{3}{2}}P } \!\implies\! \forall P\geq{\Psi_{10}},\,\forall \mathcal{T}\in\mathscr{T},\, \xi(\mathcal{D}_P, \mathcal{T}) \leq \frac{2 C_1}{\log P}.\end{aligned}$$ According to (\[combobmmbob3\]), the left hand side of (\[logic4\]) is true. But (\[logic4\]) itself is true. We thus have $$\begin{aligned} \label{logic5} \forall P\geq{\Psi_{11}},\,\forall \mathcal{T}\in\mathscr{T},\, \xi(\mathcal{D}_P, \mathcal{T}) \leq \frac{2C_1}{\log P},\end{aligned}$$ where ${\Psi_{11}}\triangleq \max\{{\Psi_{10}}, \exp(\frac{2C_1}{0.6})\}$. Note that we have also further restricted the power levels that we consider by choosing $P > \exp(\frac{2C_1}{0.6})$ so that $\frac{2C_1}{\log P} < 0.6$. Now, consider a fixed ${\Psi_{12}}\geq {\Psi_{11}}$. According to (\[logic5\]), $$\begin{aligned} \label{logic6} \forall r\in\{1,\ldots,R\},\,\exists \widetilde{\mathbf{e}}_r\in\mathcal{D}_{{\Psi_{12}}},\,\forall q\in\{1,\ldots,R\}-\{r\},\,|\widetilde{e}_{rq}|^2 \leq \frac{2C_1}{\log P} + \epsilon,\end{aligned}$$ where $\epsilon > 0$ can be arbitrary and $\widetilde{e}_{rq}$ represents the $q$th component of $\widetilde{\mathbf{e}}_r$. Let us choose $\epsilon = \frac{C_1}{\log P}$. Note that, with this choice of $\epsilon$, any $|\widetilde{e}_{rq}|^2$ in (\[logic6\]) satisfies $|\widetilde{e}_{rq}|^2 \leq 0.9 < 1$. We now show by contradiction that $\widetilde{\mathbf{e}}_i \neq \widetilde{\mathbf{e}}_j$ whenever $i \neq j$. Suppose that $\widetilde{\mathbf{e}}_i = \widetilde{\mathbf{e}}_j$ with $i\neq j$. Then, $|\widetilde{e}_{ir}|^2 < 1,\,\forall r$, which contradicts the optimality of $\mathcal{D}_{{\Psi_{12}}}$ due to Proposition \[codebookoptimalitylemma\]. Therefore, for any ${\Psi_{12}}\geq {\Psi_{11}}$, there should be $R$ distinct vectors $\widetilde{\mathbf{e}}_r,\,r=1,\ldots,R$ in $\mathcal{D}_{{\Psi_{12}}}$ with the $r$th satisfying $|\widetilde{e}_{rq}|^2 \leq \frac{3C_1}{\log P},\,\forall q\in\{1,\ldots,R\}-\{r\}$. This concludes the proof of (\[rateofconv\]). Proof of the Limit Arguments ---------------------------- We can now prove the limit arguments in the statement of the theorem using (\[rateofconv\]). Let $\mathcal{E} = \{\mathbf{x}\in\mathcal{X} : |x_1|=1\}$ represent the set of all SRS vectors that selects the first relay. First, we show that $\exists\mathbf{e}\in\mathcal{E}$ s.t. $\mathbf{e}\in\limsup_{P\rightarrow\infty}\mathcal{D}_P$. Using (\[rateofconv\]), we have $$\begin{aligned} \forall P > {\Psi_3},\,\exists\mathbf{e}\in\mathcal{E}\mbox{ s.t. }d_P(\mathbf{e}) & = \min_{\mathbf{y}\in\mathcal{D}_P} \|\mathbf{e}-\mathbf{y}\| \leq \|\mathbf{e}-\widetilde{\mathbf{e}}_1\| \leq \sqrt{\frac{3C_1(R-1)}{\log P}},\end{aligned}$$ and therefore, $\min_{\mathbf{e}\in\mathcal{E}} d_P(\mathbf{e}) \leq \sqrt{\frac{3C_1(R-1)}{\log P}}$. It follows that $$\begin{aligned} \label{limminsifir} \lim_{P\rightarrow\infty} \min_{\mathbf{e}\in\mathcal{E}} d_P(\mathbf{e}) = 0.\end{aligned}$$ Now, let $\mathbf{e}_P^{\star} \triangleq \min_{\mathbf{e}\in\mathcal{E}} d_P(\mathbf{e})$, and $\mathbf{e}_n^{\star},\,n\in\mathbb{N}$ be a sequence of beamforming vectors. Since $\mathbf{e}_n^{\star}\in\mathcal{E},\,\forall n$ and $\mathcal{E}$ is compact, by the Bolzano-Weierstrass theorem, the sequence $\mathbf{e}_n^{\star},\,n\in\mathbb{N}$ contains a subsequence $\mathbf{e}_{n_i}^{\star},\,i\in\mathbb{N}$ with $\lim_{i\rightarrow\infty} \mathbf{e}_{n_i}^{\star} = \mathbf{e}^{\star}$ for some $\mathbf{e}^{\star}\in\mathcal{E}$. Note that for any $\mathbf{y}\in\mathbb{C}^R$, $ \|\mathbf{y} - \mathbf{e}^{\star}\| \leq \|\mathbf{y} - \mathbf{e}_{n_i}\| + \|\mathbf{e}^{\star} - \mathbf{e}_{n_i}\|$ by triangle inequality. It follows that $\min_{\mathbf{y}\in\mathcal{D}_{n_i}}\|\mathbf{y} - \mathbf{e}^{\star}\| \leq \min_{\mathbf{y}\in\mathcal{D}_{n_i}}\left\{\|\mathbf{y} - \mathbf{e}_{n_i}\| + \|\mathbf{e}^{\star} - \mathbf{e}_{n_i}\|\right\} = \min_{\mathbf{y}\in\mathcal{D}_{n_i}}\|\mathbf{y} - \mathbf{e}_{n_i}\| + \|\mathbf{e}^{\star} - \mathbf{e}_{n_i}\|$. Rearranging the terms, we have $d_{n_i}(\mathbf{e}^{\star}) - d_{n_i}(\mathbf{e}_{n_i}^{\star}) \leq \| \mathbf{e}^{\star} - \mathbf{e}_{n_i}^{\star}\|,\,\forall i\in\mathbb{N}$, and thus $$\begin{aligned} \liminf_{i\rightarrow\infty} \left(d_{n_i}(\mathbf{e}^{\star}) - d_{n_i}(\mathbf{e}_{n_i}^{\star}) \right) & \leq \liminf_{i\rightarrow\infty}\|\mathbf{e}^{\star} - \mathbf{e}_{n_i}^{\star}\| \\ \label{budabaskbirliminf} & = 0.\end{aligned}$$ The equality follows from the fact that $\lim_{i\rightarrow\infty} \mathbf{e}_{n_i}^{\star} = \mathbf{e}^{\star}$. We now have $$\begin{aligned} \liminf_{P\rightarrow\infty} d_{P}(\mathbf{e}^{\star}) - \limsup_{P\rightarrow\infty} d_{P}(\mathbf{e}_{P}^{\star}) & = \liminf_{P\rightarrow\infty} d_{P}(\mathbf{e}^{\star}) + \liminf_{P\rightarrow\infty} \left(-d_{P}(\mathbf{e}_{P}^{\star})\right) \\ & \label{liminffirsteq} \leq \liminf_{P\rightarrow\infty}\left(d_{P}(\mathbf{e}^{\star}) - d_{P}(\mathbf{e}_{P}^{\star}) \right) \\ & \label{liminfsecondeq} \leq \liminf_{i\rightarrow\infty} \left(d_{n_i}(\mathbf{e}^{\star}) - d_{n_i}(\mathbf{e}_{n_i}^{\star})\right), \\ \label{liminfinfinalinequalitisbu} & \leq 0,\end{aligned}$$ where (\[liminffirsteq\]) follows since $$\begin{aligned} \label{thatliminfbound} \liminf_{x\rightarrow\infty}f(x)+\liminf_{x\rightarrow\infty}g(x) \leq \liminf_{x\rightarrow\infty}\left(f(x) + g(x)\right),\end{aligned}$$ for any functions $f$ and $g$. For (\[liminfsecondeq\]), we have used the fact that the lower limit of a sequence is less than the lower limit of any of its subsequences. For (\[liminfinfinalinequalitisbu\]), we have used (\[budabaskbirliminf\]). Now, since $\liminf_{P\rightarrow\infty} d_{P}(\mathbf{e}^{\star}) - \limsup_{P\rightarrow\infty} d_{P}(\mathbf{e}_{P}^{\star}) \leq 0$ as shown in the derivation above, we have $$\begin{aligned} \liminf_{P\rightarrow\infty} d_{P}(\mathbf{e}^{\star}) \leq \limsup_{P\rightarrow\infty} d_{P}(\mathbf{e}_{P}^{\star}) = \limsup_{P\rightarrow\infty} \min_{\mathbf{e}\in\mathcal{E}} d_P(\mathbf{e}) = 0,\end{aligned}$$ where the last equality follows from (\[limminsifir\]). Therefore, $\liminf_{P\rightarrow\infty} d_{P}(\mathbf{e}^{\star}) \leq 0$. On the other hand, obviously we have $\liminf_{P\rightarrow\infty} d_{P}(\mathbf{e}^{\star}) \geq 0$. Combining the two inequalities yields $\liminf_{P\rightarrow\infty} d_{P}(\mathbf{e}^{\star}) = 0$. This shows the existence of an SRS vector $\mathbf{e}\in\mathcal{E}$ (namely $\mathbf{e}^{\star}$) that selects the first relay and satisfies $\liminf_{P\rightarrow\infty} d_P(\mathbf{e}) = 0$, or equivalently, $\mathbf{e}\in\limsup_{P\rightarrow\infty}\mathcal{D}_P$. We can similarly show the existence of the remaining $R-1$ SRS vectors in the upper limit of $\mathcal{D}$. Therefore, $\exists\mathcal{C}_{\mathtt{SRS}}\in\mathscr{C}_{\mathtt{SRS}}$ s.t. $\mathcal{C}_{\mathtt{SRS}}\subset\limsup_{P\rightarrow\infty}\mathcal{D}_P$. What is left is to show that if $|\mathcal{D}| = R$, and $\lim_{P\rightarrow\infty}\mathcal{D}_P$ exists, we have $\exists\mathcal{C}_{\mathtt{SRS}}'\in\mathscr{C}_{\mathtt{SRS}}$ s.t. $\mathcal{C}_{\mathtt{SRS}}'=\lim_{P\rightarrow\infty}\mathcal{D}_P$. We have shown that $\mathcal{C}_{\mathtt{SRS}}' \subset \limsup_{P\rightarrow\infty}\mathcal{D}_P$. If $\lim_{P\rightarrow\infty}\mathcal{D}_P$ exists, then $\limsup_{P\rightarrow\infty}\mathcal{D}_P = \lim_{P\rightarrow\infty}\mathcal{D}_P$, and thus $\mathcal{C}_{\mathtt{SRS}}' \subset \lim_{P\rightarrow\infty}\mathcal{D}_P$ with $|\lim_{P\rightarrow\infty}\mathcal{D}_P| \geq R$. To complete the proof, it is therefore sufficient to show that $|\lim_{P\rightarrow\infty}\mathcal{D}_P| \leq R$. The following lemma shows that when $\lim_{P\rightarrow\infty}\mathcal{D}_P$ exists, its cardinality cannot be more than $|\mathcal{D}|$, and thus concludes the proof of the theorem. \[limitexistssebokolur\] For any d-codebook $\mathcal{D}$ with $|\mathcal{D}| < \infty$, if $\lim_{P\rightarrow\infty}\mathcal{D}_P$ exists, then $|\lim_{P\rightarrow\infty}\mathcal{D}_P| \leq |\mathcal{D}|$. Let $\mathcal{L} = \lim_{P\rightarrow\infty}\mathcal{D}_P$. Suppose that $|\mathcal{L}| \geq |\mathcal{D}|+1$. Then, $\exists\mathbf{x}_i,\ldots,\mathbf{x}_{|\mathcal{D}|+1}\in\mathcal{L}$, with $\forall i,j\in\{1,\ldots,|\mathcal{D}|+1\},\,\mathbf{x}_i \neq \mathbf{x}_j \Longleftrightarrow i \neq j$. Since $\mathcal{L} = \liminf_{P\rightarrow\infty}\mathcal{D}_P$ as well, we have $\forall i\in\{1,\ldots, |\mathcal{D}| + 1\}$, $\lim_{P\rightarrow\infty}\min_{\mathbf{y}\in\mathcal{D}_P} \|\mathbf{y} - \mathbf{x}_i\| = 0$ by the definition of $\liminf_{P\rightarrow\infty}\mathcal{D}_P$. This implies that $\forall i\in\{1,\ldots,|\mathcal{D}| + 1\},\,\forall \epsilon > 0,\,\exists P_{i,\epsilon} > 0\mbox{ s.t. }\forall P > P_{i,\epsilon},\,\min_{\mathbf{y}\in\mathcal{D}_P}\|\mathbf{y} - \mathbf{x}_i\| \leq \epsilon$. Letting $P_{\epsilon} = \max_i P_{i,\epsilon}$, we have $$\begin{aligned} \label{limitargument} \forall \epsilon > 0,\,\exists P_{\epsilon} > 0\mbox{ s.t. }\forall P > P_{\epsilon},\,\forall i\in\{1,\ldots,|\mathcal{D}| + 1\},\,\min_{\mathbf{y}\in\mathcal{D}_P}\|\mathbf{y} - \mathbf{x}_i\| \leq \epsilon. \end{aligned}$$ Now, let $$\begin{aligned} \label{deltaargument} \delta = \min_{\substack{i,j\in\{1,\ldots,|\mathcal{D}|+1\}\\i \neq j}} \|\mathbf{x}_i - \mathbf{x}_j\|,\end{aligned}$$ set $\epsilon = \delta/4$, and consider a fixed $P_0 > P_{\delta/4}$. Also, let $\mathbf{y}_i = \arg\min_{\mathbf{y}\in\mathcal{D}_{P_0}} \|\mathbf{y} - \mathbf{x}_i\|,\,i=1,\ldots,|\mathcal{D}|+1$. Since $|\mathcal{D}_{P_0}| = |\mathcal{D}|$, $\mathbf{y}_k = \mathbf{y}_\ell \triangleq \widetilde{\mathbf{y}}$ for some $k \neq \ell$. Note that, as a result of (\[limitargument\]) and the definition of $\widetilde{\mathbf{y}}$, we have $\|\widetilde{\mathbf{y}} - \mathbf{x}_k\| \leq \delta/4$ and $\|\widetilde{\mathbf{y}} - \mathbf{x}_\ell\| \leq \delta/4$. However, by triangle inequality, $\|\mathbf{x}_k - \mathbf{x}_\ell\| \leq \|\mathbf{x}_k - \widetilde{\mathbf{y}}\| + \|\mathbf{x}_\ell - \widetilde{\mathbf{y}}\| \leq \delta/4+\delta/4 = \delta/2$, and this contradicts (\[deltaargument\]). Therefore, the cardinality of $\mathcal{L}$ cannot be more than $|\mathcal{D}|$, concluding the proof. Proof of Theorem \[sdtheorem\] {#proofofsdtheorem} ============================== Similar to what has been done in the proof of Theorem \[ldtheorem\], we carry out the proof in two parts: First we prove (\[sdtheoremeq\]), and then the limit arguments in the statement of the theorem. Proof of (\[sdtheoremeq\]) -------------------------- Let $\mathscr{T}$ represent the collection of all subsets of $\{1,\ldots,R\}$ with cardinality no greater than $|\mathcal{D}|-1$. Then, using the same ideas as in the proof of Theorem \[ldtheorem\], we have $$\begin{aligned} \label{condgaribus} \forall P>{\Psi_{13}},\,\forall \mathcal{T}\in\mathscr{T},\, \xi(\mathcal{D}_P, \mathcal{T}) \leq \frac{{C_{12}}}{\log P},\end{aligned}$$ for constants $0 < {C_{12}},\, {\Psi_{13}}<\infty$ independent of $P$. Let us set the constants in the statement of the theorem as ${C_6}= R{C_{12}}$ and ${\Psi_{6}}= {\Psi_{13}}$. We now prove (\[sdtheoremeq\]) by contradiction. Suppose that (\[sdtheoremeq\]) is false. Then, $\exists {\Psi_{14}}> {\Psi_{6}}= {\Psi_{13}},\,\exists\mathbf{x},\mathbf{y}\in\mathcal{C},\,\mathbf{x}\neq\mathbf{y},\,\sum_{r=1}^R |x_r||y_r| > \frac{ R{C_{12}}}{\log P}$. This implies that $\exists r\in\{1,\ldots,R\},\,|x_r||y_r| > \frac{ {C_{12}}}{\log P}$. Hence, either we have $|x_r|> (\frac{{C_{12}}}{\log P})^{1/2}$ or $|y_r| > (\frac{{C_{12}}}{\log P})^{1/2}$. Now, let ${\mathcal{T}_6}= \{r\}\cup\{\iota(\mathbf{z}): \mathbf{z}\in\mathcal{D}_{{\Psi_{14}}} - \{\mathbf{x},\mathbf{y}\}\}$, where $\iota(\mathbf{z})$ is any index that satisfies $|z_{\iota(\mathbf{z})}| = 1$. Since $|{\mathcal{T}_6}| \leq |\mathcal{D}| - 1$, ${\mathcal{T}_6}\in\mathscr{T}$. Also, either $\max_{t\in{\mathcal{T}_6}}|x_t|^2> \frac{{C_{12}}}{\log P}$ or $\max_{t\in{\mathcal{T}_6}}|y_t|^2 > \frac{{C_{12}}}{\log P}$. Moreover, $\forall \mathbf{z}\in\mathcal{D}_{{\Psi_{14}}} - \{\mathbf{x},\mathbf{y}\},\,\max_{t\in{\mathcal{T}_6}} |z_t|^2 = 1$ by the construction of ${\mathcal{T}_6}$. Therefore, $\xi(\mathcal{D}_{{\Psi_{14}}}, {\mathcal{T}_6}) = \inf_{\mathbf{x}\in\mathcal{D}_{{\Psi_{14}}}} \max_{t\in{\mathcal{T}_6}} |x_t|^2 > \frac{{C_{12}}}{\log P}$. But, this contradicts (\[condgaribus\]), and thus concludes the proof of (\[sdtheoremeq\]). Proof of the Limit Arguments ---------------------------- We can now prove the limit arguments in the statement of the theorem using (\[sdtheoremeq\]). Let $\mathcal{O} = \{\mathbf{o}_1, \ldots,\mathbf{o}_{|\mathcal{D}|}\}$ represent an OMRS codebook given a beamforming codebook cardinality $1 \leq |\mathcal{D}| \leq R$. For any OMRS codebook $\mathcal{O}$, we define its “vectorized” version $\mathtt{vec}(\mathcal{O}) \triangleq [\mathbf{o}_1\cdots\mathbf{o}_{|\mathcal{D}|}]$ as an alternative representation for $\mathcal{O}$. Also, let $\mathfrak{O}\triangleq\bigcup_{\mathcal{O}\in\mathscr{O}}\mathtt{vec}(\mathcal{O})$ represent the collection of all vectorized OMRS codebooks. We now need the following lemma to proceed: For any $\mathcal{D}_P$ that satisfies (\[sdtheoremeq\]), $\exists\mathbf{o}\in\mathfrak{O}$ s.t. $\min_{\mathbf{y}\in\mathcal{D}_P^{|\mathcal{D}|}} \|\mathbf{o} - \mathbf{y}\| \leq \frac{{C_6}\sqrt{R|\mathcal{D}|}}{\log P}$. Let $\mathcal{D}_P = \{\mathbf{x}_1, \ldots, \mathbf{x}_{|\mathcal{D}|}\}$. For convenience, we rewrite the condition in (\[sdtheoremeq\]) as $$\begin{aligned} \label{orthocond} \max_{\substack{i,j\in\{1,\ldots,|\mathcal{D}|\} \\ i \neq j}} \sum_{r=1}^R |x_{ir}||x_{jr}| \leq \delta,\end{aligned}$$ where $\delta = \frac{{C_6}}{\log P}$. Now, (\[orthocond\]) implies that for all $i,j\in\{1,\ldots,|\mathcal{D}|\}$ with $i \neq j$ we have $|x_{ir}||x_{jr}| \leq \delta,\,\forall r\in\{1,\ldots,R\}$. Then, given any $r\in\{1,\ldots,R\}$, either $|x_{ir}| \leq \sqrt{\delta},\,\forall i\in\{1,\ldots,|\mathcal{D}|\}$, or there exists only one index $i_r^{\star}\in\{1,\ldots,|\mathcal{D}|\}$ s.t. $|x_{i_r^{\star}, r}|> \sqrt{\delta}$, and $|x_{ir}| \leq \sqrt{\delta},\,\forall i\in\{1,\ldots,|\mathcal{D}|\}-\{i_r^{\star}\}$. Let us now define the function $f:\mathbb{C}\rightarrow\mathbb{C}$ with $f(x) = x$ if $|x| > \sqrt{\delta}$, and $f(x) = 0$, otherwise. Now, let $o_{ir} = f(x_{ir}),\,i=1,\ldots,|\mathcal{D}|,\,r=1,\ldots,R$. As a result of the properties of $f$ and $x_{ir}$, not only $\mathbf{o}\in\mathfrak{O}$ is an OMRS codebook, but also, $$\begin{aligned} \min_{\mathbf{y}\in\mathcal{D}_P^{|\mathcal{D}|}} \|\mathbf{o} - \mathbf{y}\| & = \min_{\mathbf{y}\in\mathcal{D}_P^{|\mathcal{D}|}} \sqrt{\sum_{r=1}^R \sum_{i=1}^{|\mathcal{D}|} |o_{ir} - y_{ir}|^2} \\ & \leq \sqrt{\sum_{r=1}^R\sum_{i=1}^{|\mathcal{D}|} |o_{ir} - x_{ir}|^2} \\ & \leq \sqrt{\sum_{r=1}^R |\mathcal{D}|\delta^2 } \\ & = \sqrt{R|\mathcal{D}|} \delta,\end{aligned}$$ and this concludes the proof. In other words, if $\mathcal{D}$ is an optimal d-codebook, for all $P$ sufficiently large, we can find an OMRS codebook $\mathbf{o}\in\mathfrak{O}$ s.t. $\mathbf{o}$ is as close as $\frac{{C_6}\sqrt{R|\mathcal{D}|}}{\log P}$ to $\mathcal{D}_P$. We thus have $$\begin{aligned} \label{infimumluyaklasim} \lim_{P\rightarrow\infty} \inf_{\mathbf{o}\in\mathcal{O}} \min_{\mathbf{y}\in\mathcal{D}_P^{|\mathcal{D}|}} \|\mathbf{y} - \mathbf{o}\|= 0.\end{aligned}$$ We now prove that $\mathfrak{O}$ is a compact set so that we can replace the infimum in (\[infimumluyaklasim\]) by a minimum. $\mathfrak{O}$ is compact. It is sufficient to show that $\mathfrak{O}$ is bounded and closed. Since $\mathfrak{O} \subset \mathcal{X}^{|\mathcal{D}|}$ and $\mathcal{X}^{|\mathcal{D}|}$ is bounded, $\mathfrak{O}$ is bounded. We prove that $\mathfrak{O}$ is also closed by showing that it can be expressed as the union of a finite number of closed sets. First, we need the following definitions: - Let $\mathcal{V}$ represent the set of all vectors $[\begin{array}{cccccc} \alpha_1 & \cdots & \alpha_{|\mathcal{D}|} & \beta_1 & \cdots & \beta_{R - |\mathcal{D}|}\end{array}]$ that satisfy the following: 1. $\alpha_1,\,\ldots,\,\alpha_{|\mathcal{D}|},\,\beta_1,\,\ldots,\,\beta_{R - |\mathcal{D}|}$ are positive integers. 2. $1 \leq \alpha_1,\ldots,\alpha_{|\mathcal{D}|} \leq R$. 3. $\forall i,j\in\{1,\ldots,|\mathcal{D}|\},\,\alpha_i \neq \alpha_j \Leftrightarrow i\neq j$. 4. $1 \leq \beta_1, \ldots, \beta_{R - |\mathcal{D}|} \leq |\mathcal{D}|$. Note that $|\mathcal{V}| = R(R-1)\cdots(R-|\mathcal{D}|+1)|\mathcal{D}|^{R - |\mathcal{D}|}$. - Let $\mathtt{DISK} = \{x\in\mathbb{C}:\|x\|\leq 1\}$ and $\mathtt{CIRC} = \{x\in\mathbb{C}:\|x\|=1\}$ represent the unit disk and the unit circle, respectively. - Given $\mathbf{v} = [\begin{array}{cccccc} \alpha_1 & \cdots & \alpha_{|\mathcal{D}|} & \beta_1 & \cdots & \beta_{R - |\mathcal{D}|}\end{array}]\in\mathcal{V}$, let $\mathfrak{X}_{\mathbf{v}}$ represent the collection of all vectorized codebooks $[\begin{array}{ccc}\mathbf{x}_1 & \cdots & \mathbf{x}_{|\mathcal{D}|}\end{array}] = [\begin{array}{ccccccc} x_{11} & \cdots & x_{1R} & \cdots & x_{|\mathcal{D}|,1} & \cdots & x_{|\mathcal{D}|, R}\end{array}]$ with the following properties: 1. $\forall i\in\{1,\ldots,|\mathcal{D}|\},\,x_{i\alpha_i} \in \mathtt{CIRC},\,\forall j\in\{1,\ldots,|\mathcal{D}|\}-\{i\},\,x_{j\alpha_i} = 0$. 2. $\forall i\in\{1,\ldots,R - |\mathcal{D}|\},\,x_{\beta_i\gamma_i} \in \mathtt{DISC},\,\forall j\in\{1,\ldots,R - |\mathcal{D}|\}- \{\beta_i\},\,x_{j\gamma_i} = 0$, where $\gamma_1 < \cdots < \gamma_{R - |\mathcal{D}|}$ satisfy $\{\gamma_1, \ldots, \gamma_{R - |\mathcal{D}|}\} = \{1,\ldots,R\}-\{\alpha_1, \ldots, \alpha_{|\mathcal{D}|}\}$. According to these properties, for any given $\mathbf{v}\in\mathcal{V}$, $\mathfrak{X}_{\mathbf{v}}$ can be expressed as a finite cartesian product of the closed sets $\mathtt{DISC}$, $\mathtt{CIRC}$ and $\{0\}$. Hence, $\mathfrak{X}_{\mathbf{v}}$ is closed for any $\mathbf{v}$. It is straightforward to show that $\mathfrak{X}_{\mathbf{v}}$ is a set of OMRS codebooks for any given $\mathbf{v}\in\mathcal{V}$, and thus $\mathbf{o}\in\bigcup_{\mathbf{v}\in\mathcal{V}} \mathfrak{X}_{\mathbf{v}} \implies \mathbf{o}\in \mathfrak{O}$. Each $\mathbf{v}\in\mathcal{V}$ actually corresponds to a particular OMRS structure. As an example, for $R=2$ and $|\mathcal{D}| = 1$, let $\mathbf{w} = [\begin{array}{cccc} 1 & 3 & 1 & 1 \end{array}]$. Then, $\mathfrak{X}_{\mathbf{w}}$ is the union of all OMRS codebooks of structure $[\begin{array}{cccccccc} x_{11} & x_{12} & 0 & x_{14} & 0 & 0 & x_{23} & 0 \end{array}]$, where $x_{11},x_{23}\in\mathtt{CIRC}$ and $x_{12},x_{14}\in\mathtt{DISC}$. We now show the converse, i.e. $\mathbf{o}\in \mathfrak{O} \implies \mathbf{o}\in\bigcup_{\mathbf{v}\in\mathcal{V}} \mathfrak{X}_{\mathbf{v}}$. Consider some $\mathbf{o}\in\mathfrak{O}$. We shall construct a $\mathbf{v} = [\begin{array}{cccccc} \alpha_1 & \cdots & \alpha_{|\mathcal{D}|} & \beta_1 & \cdots & \beta_{R - |\mathcal{D}|}\end{array}]\in\mathcal{V}$ s.t. $\mathbf{o}\in\mathfrak{X}_{\mathbf{v}}$. Since $\mathbf{o}$ is an optimal codebook, by Proposition \[codebookoptimalitylemma\], at least one component of every beamforming vector in $\mathbf{o}$ has unit norm, and thus we choose the $\alpha_i$ in such a way that $|o_{i\alpha_i}| = 1$, or equivalently, $o_{i\alpha_i}\in\mathtt{CIRC}$. Also, since $\mathbf{o}$ is an OMRS codebook, we have $o_{j\alpha_i} = 0,\,\forall j \neq i$ by definition. This satisfies the first property in the definition of $\mathfrak{X}_{\mathbf{v}}$. Now, let $\gamma_1 < \cdots < \gamma_{R - |\mathcal{D}|}$ satisfy $\{\gamma_1, \ldots, \gamma_{R - |\mathcal{D}|}\} = \{1,\ldots,R\}-\{\alpha_1, \ldots, \alpha_{|\mathcal{D}|}\}$. For any given $\gamma_i$, there are two possibilities: 1. $x_{j\gamma_i}=0,\,\forall j\in\{1,\ldots,|\mathcal{D}|\}$. In this case we can pick any $1 \leq \beta_i \leq |\mathcal{D}|$. 2. Since $\mathbf{o}$ is an OMRS codebook, there is at most one non-zero $x_{j\gamma_i},\,j=1,\ldots,|\mathcal{D}|$. Suppose that $x_{j'\gamma_i}\neq 0$. Then, we set $\beta_i = j'$. This satisfies the second property in the definition of $\mathfrak{X}_{\mathbf{v}}$. Therefore, for the particular $\alpha$s and $\beta$s we have chosen, $\mathbf{o}\in\mathfrak{X}_v$, and thus in general $\mathbf{o}\in \mathfrak{O} \implies \mathbf{o}\in\bigcup_{\mathbf{v}\in\mathcal{V}} \mathfrak{X}_{\mathbf{v}}$. Combining this with the fact that $\mathbf{o}\in\bigcup_{\mathbf{v}\in\mathcal{V}} \mathfrak{X}_{\mathbf{v}} \implies \mathbf{o}\in \mathfrak{O}$, we have $\mathfrak{O} = \bigcup_{\mathbf{v}\in\mathcal{V}} \mathfrak{X}_{\mathbf{v}}$. Hence $\mathfrak{O}$ is the union of a finite number of closed sets. Therefore, it is closed, and this concludes the proof. Hence, we can rewrite (\[infimumluyaklasim\]) as $$\begin{aligned} \label{minimumluyaklasim} \lim_{P\rightarrow\infty} \min_{\mathbf{o}\in\mathcal{O}} \min_{\mathbf{y}\in\mathcal{D}_P^{|\mathcal{D}|}} \|\mathbf{y} - \mathbf{o}\|= 0.\end{aligned}$$ Note that this equality has the same form as (\[limminsifir\]). Using the exact same steps as in Appendix \[proofofldtheorem\], we can show that $$\begin{aligned} \label{minimumluyaklasimmmm} \exists\mathbf{o}^{\star}\in\mathfrak{O}\mbox{ s.t. } \liminf_{P\rightarrow\infty} \min_{\mathbf{y}\in\mathcal{D}_P^{|\mathcal{D}|}} \|\mathbf{y} - \mathbf{o}^{\star}\|= 0.\end{aligned}$$ Now, we have $$\begin{aligned} 0 & = \liminf_{P\rightarrow\infty} \min_{\mathbf{y}\in\mathcal{D}_P^{|\mathcal{D}|}} \|\mathbf{y} - \mathbf{o}^{\star}\|\\ \label{hhhhhooooollllddddeeerrrr}& \geq \liminf_{P\rightarrow\infty} \min_{\mathbf{y}\in\mathcal{D}_P^{|\mathcal{D}|}} \frac{1}{\sqrt{|\mathcal{D}|}}\sum_{i=1}^{|\mathcal{D}|}\|\mathbf{y}_i - \mathbf{o}_i^{\star}\| \\ & = \frac{1}{\sqrt{|\mathcal{D}|}} \liminf_{P\rightarrow\infty} \min_{\mathbf{y}_1\in\mathcal{D}_P} \cdots \min_{\mathbf{y}_{|\mathcal{D}|}\in\mathcal{D}_P} \sum_{i=1}^{|\mathcal{D}|}\|\mathbf{y}_i - \mathbf{o}_i^{\star}\| \\ & = \frac{1}{\sqrt{|\mathcal{D}|}} \liminf_{P\rightarrow\infty} \sum_{i=1}^{|\mathcal{D}|} \min_{\mathbf{y}_i\in\mathcal{D}_P} \|\mathbf{y}_i - \mathbf{o}_i^{\star}\| \\ \label{againliminfineq} & \geq \frac{1}{\sqrt{|\mathcal{D}|}} \sum_{i=1}^{|\mathcal{D}|} \liminf_{P\rightarrow\infty} \min_{\mathbf{y}_i\in\mathcal{D}_P} \|\mathbf{y}_i - \mathbf{o}_i^{\star}\|,\end{aligned}$$ where (\[hhhhhooooollllddddeeerrrr\]) follows from Hölder’s inequality. For (\[againliminfineq\]), we have used (\[thatliminfbound\]). Finally, using (\[againliminfineq\]), $\forall i\in\{1,\ldots,|\mathcal{D}|\}$, we have $\liminf_{P\rightarrow\infty} \min_{\mathbf{y}\in\mathcal{D}_P} \|\mathbf{y} - \mathbf{o}_i^{\star}\| = 0$, or equivalently $\mathbf{o}_i^{\star} \in \limsup_{P\rightarrow\infty} \mathcal{D}_P$. This shows the existence of an OMRS codebook $\mathcal{O}^{\star}\in\mathscr{O}(|\mathcal{D}|)$ s.t. $\mathcal{O}^{\star}\subset\limsup_{P\rightarrow\infty} \mathcal{D}_P$. If the limit exists, according to Lemma \[limitexistssebokolur\], $\mathcal{O}^{\star}=\lim_{P\rightarrow\infty} \mathcal{D}_P$, concluding the proof. [^1]: The authors are with the Center for Pervasive Communications and Computing, University of California, Irvine, Irvine CA 92697-2625 USA. Email: {ekoyuncu, hamidj}@uci.edu. [^2]: Note that this orthogonality condition is not the same as the “usual” orthogonality condition for complex vectors with respect to the Hermitian inner product. [^3]: We ignore the events that have zero probability, e.g. the event that $\max Z_r = \min Z_r$
{ "pile_set_name": "ArXiv" }
--- abstract: | We study an alternative to the prevailing approach to modelling qualitative spatial reasoning (QSR) problems as constraint satisfaction problems. In the standard approach, a relation between objects is a constraint whereas in the alternative approach it is a variable. By being declarative, the relation-variable approach greatly simplifies integration and implementation of QSR. To substantiate this point, we discuss several specific QSR algorithms from the literature which in the relation-variable approach reduce to the customary constraint propagation algorithm enforcing generalised arc-consistency. author: - Sebastian Brand title: | Relation Variables in\ Qualitative Spatial Reasoning --- Introduction ============ Qualitative spatial representation and reasoning (QSR) [@cohn:2001:qualitative] lends itself well to modelling by constraints. In the standard approach, a spatial object, such as a region, is described by a variable, and the qualitative relation between spatial objects, such as a topological relation between two regions, contributes a constraint. For many QSR calculi, it is known that if all the constraints represent definite (base) relations and path-consistency (PC) holds, then this description of a spatial scene is consistent. If the relation is not fully specified, the corresponding constraint is a disjunction of basic constraints. By establishing PC, such a disjunctive constraint is refined in view of the constraints with which it shares a variable. A combination of PC with search over the disjunctive constraints decides the consistency of indefinite scene descriptions. We examine here an alternative constraint-based formulation of QSR. In this approach, a spatial object is a constant, and the relation between spatial objects is a variable. We call this the *relation-variable* approach, in contrast to the conventional *relation-constraint* approach above. Although modelling QSR with relation variables is not original, see [@tsang:1987:consistent], it is mentioned very rarely. This fact surprises in view of the advantages of this approach. In particular, the following two important issues are tackled successfully: Integration. : Space has several aspects that can be characterised qualitatively, such as size, shape, orientation. These aspects are interdependent, but no convenient canonical representation exists that provides a link (the role of time points in temporal reasoning). Spatial reasoning problems in practice are also not likely to occur in pure form. They may be embedded into a non-spatial context, or contain application-specific side constraints. The relation-variable approach to QSR is declarative in a strict sense and is thus well-suited for these integration problems. Systems. : Typical current constraint solving platforms focus on domain reduction, and accordingly provide convenient access to variable domains. Modifying the constraint network, on the other hand, is usually difficult. This task is, however, required for enforcing PC. A formulation of QSR according to the relation-variable approach means that generic domain-reducing propagation algorithms and conventional constraint solving platforms can be used instead of dedicated spatial reasoning systems. #### Plan of the paper. We begin by introducing briefly the necessary constraint solving concepts and methods, and qualitative spatial reasoning, using the example of the RCC-8 calculus. The next section presents in-depth the two modelling approaches for constraint-based QSR. In the following sections, we discuss several aspects of space and contrast the relation-variable and relation-domain approach. We finally mention some new modelling options, and end with a summary. Constraint Satisfaction ----------------------- Recent coverage of the field can be found in [@apt:2003:principles; @dechter:2003:constraint; @fruehwirth:2003:essentials]. Consider a sequence $X = {{x_1, \ldots ,x_m}}$ of pairwise different variables with respective domains ${{D_1, \ldots ,D_m}}$. By a *constraint* $C$ on $X$, written $C(X)$, we mean a subset of ${{D_1\times \cdots \timesD_m}}$. The arity of $C$ is $m$. A *constraint satisfaction problem (CSP)* consists of a finite sequence of variables $X= {{x_1, \ldots ,x_n}}$ with respective domains ${\mathcal{D}}= {{D_1, \ldots ,D_n}}$, and a finite set ${\mathcal{C}}$ of constraints, each on a subsequence of $X$. We write it as ${\left\langle{{\mathcal{C}}};{x_1 \in D_1, \ldots, x_n \in D_n}\right\rangle}$, or shorter as ${\left\langle{{\mathcal{C}}};{X \in {\mathcal{D}}}\right\rangle}$. Given an element $d = {{d_1, \ldots ,d_n}}$ of ${{D_1\times \ldots \timesD_n}}$ and a subsequence of $X$ we denote by $d[Y]$ the sequence ${{d_{i_1}, \ldots ,d_{i_\ell}}}$; in particular, we have $d[x_{k}] = d_{k}$. A *solution* to ${\left\langle{{\mathcal{C}}};{X \in {\mathcal{D}}}\right\rangle}$ is an element $d \in {\mathcal{D}}$ such that for each constraint $C \in {\mathcal{C}}$ on the variables $Y$ we have $d[Y] \in C$. ### Constraint propagation. One method to establish satisfiability of CSPs when the search space is finite is systematic search for a solution. For reducing the search space and overall search effort, constraint propagation is often very useful; the principle is to replace a given CSP by another one that is equivalent with respect to the solutions but that is easier to solve. Constraint propagation is typically characterised by the resulting *local consistency*. The two notions most relevant for this paper are: *Path Consistency (PC)*: : A CSP of binary constraints is path-consistent [@montanari:1974:networks] if for every triple of variables $x,y,z$ $$C(x,z) = {\left\{\,{ (a,c) }\mid{ b \ \text{exists s.t.}\ (a,b) \in C(x,y) \ \text{and}\ (b,c) \in C(y,z) }\,\right\}}.$$ It is assumed here that a unique constraint $C(u,w)$ for each pair of variables $u,w$ exists, and that $C(u,w) = C^{-1}(w,u)$. *Generalised Arc-Consistency (GAC)*: : A constraint $C(X)$ is generalised arc-consistent [@mohr:1988:good] if for all $x_k \in X$ and all $a \in D_k$ $$d \in C(X) \quad\text{exists such that}\quad d[x_k]=a.$$ In short, every domain value must participate in a local solution. A CSP is generalised arc-consistent if each of its constraints is. For example, the CSP ${\left\langle{x + y = z};{\; x,y,z \in \{1,2,3\}}\right\rangle}$ can be reduced to ${\left\langle{x + y = z};{\; x,y \in \{1,2\}, z \in \{2,3\}}\right\rangle}$ which is GAC. Enforcing PC means reducing constraints but not domains, whereas enforcing GAC means reducing domains but not constraints. A number of generic methods to establish GAC for a constraint are known, and many constraint solving systems have implementations. One example is the *GAC-schema* [@bessiere:1997:arc-consistency] available in ILOG Solver [@ilog:2001:solver]. Qualitative Spatial Reasoning ----------------------------- The topological calculus RCC-8 [@randell:1992:spatial] is one of the best-known formalisations in spatial reasoning. We use it to illustrate a number of concepts. In one distinguishes 8 topological relations between two regions, see Fig. \[figure:rcc8\]: *disconnected, externally connected, partially overlapping, equal, tangential proper part, non-tangential proper part*, and inverses of the latter two. These are denoted ${\mathsf{DC}}, {\mathsf{EC}}, {\mathsf{PO}}, {\mathsf{EQ}}, {\mathsf{TPP}}, {\mathsf{NTPP}}, {\mathsf{TPPi}}, {\mathsf{NTPPi}}$, respectively; together they form a set that we call ${{\mathsf{RCC8}}}$. (0.5,1)(11.6,5) (1,3.7)[0.5]{} (1,3.7)[a]{} (1,2.3)[0.5]{} (1,2.3)[b]{} (1.3,3)[disjoint]{} (1,1.4)[${\mathsf{DC}}$]{} (2.6,3)[meet1]{} (3,3.5)[0.5]{} (3,3.5)[a]{} (3,2.5)[0.5]{} (3,2.5)[b]{} (3.4,3)[meet2]{} (3,1.6)[${\mathsf{EC}}$]{} (4.5,3)[overlap1]{} (5,3.3)[0.5]{} (5,3.5)[a]{} (5,2.7)[0.5]{} (5,2.5)[b]{} (5.5,3)[overlap2]{} (5,1.8)[${\mathsf{PO}}$]{} (7,3)[0.5]{}[equal]{} (6.8,3.1)[a]{} (6.6,2)[${\mathsf{EQ}}$]{} (7.2,2.9)[b]{} (9,4)[0.5]{}[coveredby]{} (8.8,4)[a]{} (8.8,4)[0.3]{} (9.3,4)[b]{} (9,4.8)[${\mathsf{TPP}}$]{} (9,2)[0.5]{}[covers]{} (8.7,2)[a]{} (9.2,2)[0.3]{} (9.2,2)[b]{} (9,1.2)[${\mathsf{TPPi}}$]{} (11,4)[0.5]{}[inside]{} (10.9,4)[a]{} (10.9,4)[0.3]{} (11.35,4)[b]{} (11,4.8)[${\mathsf{NTPP}}$]{} (11,2)[0.5]{}[contains]{} (10.65,2)[a]{} (11.1,2)[0.3]{} (11.1,2)[b]{} (11,1.2)[${\mathsf{NTPPi}}$]{} #### Jointly exhaustive and pairwise disjoint. Any two spatial regions are in one and exactly one of the RCC-8 relations to each other. #### Composition table. Considering the triple $R_{a,b}, R_{b,c}, R_{a,c}$ of relations between regions $a,b,c$, one finds that not all triples of RCC-8 relations are semantically feasible. The consistent triples are collected in the RCC-8 composition table. It contains 193 relation triples, such as $({\mathsf{NTPP}}, {\mathsf{EC}}, {\mathsf{DC}})$. Bennett [@bennett:1998:determining] proved that compositional consistency entails global consistency: if for all triples of regions the relations between them respect the composition table then this topological scenario is consistent. #### Converse relation table. In analogy to the composition table, it is helpful to think of a converse relation table consisting of the 8 pairs $(R,Ri)$ of RCC-8 relations such that $Ri$ is the converse of $R$. It contains for example $({\mathsf{EQ}}, {\mathsf{EQ}})$ and $({\mathsf{TPP}}, {\mathsf{TPPi}})$. If we agree on $({\mathsf{EQ}})$ for the relation of a region with itself then the converse relation table follows from the composition table. Approaches to Constraint-based QSR ================================== A spatial topological scenario consists of a set of region names denoted by ${{\mathit{Regions}}}$, and possibly some restrictions on the topological relation for regions pairs. A scenario is fully specified if for each region pair exactly one RCC-8 relation is given. We examine now how scenarios can be modelled as constraint satisfaction problems. We continue using topology with RCC-8 as an example, but most of the concepts below are immediately transferable to other spatial aspects. Relations as Constraints ------------------------ In this conventional approach, ${{\mathit{Regions}}}$ is considered to be a set of region variables. Their infinite domain is the set of all spatial regions in the underlying topological space; for example, if we model 2D space then a region variable represents a set of points in the plane. Information about the topological relation between two regions is expressed as a binary constraint ${{\mathit{Rel}}}$ that corresponds to a subset of ${{\mathsf{RCC8}}}$. One usually writes this in infix notation as $$\text{constraint}\quad x \mathbin{{{\mathit{Rel}}}} y \quad {\quad\text{where}\quad} {{\mathit{Rel}}}\subseteq {{\mathsf{RCC8}}}{\quad\text{and}\quad} x,y \in {{\mathit{Regions}}}.$$ Such a CSP describes a possibly partially specified scenario. Whether a corresponding fully specified and satisfiable scenario exists is checked by path-consistency and search over the relations. A PC-enforcing algorithm revises the constraints between regions according to the converse relation and composition tables of RCC-8, and search branches over disjunctive constraints. Establishing satisfiability of a scenario processes only the constraints, for compositional consistency. The variables remain unassigned. Relations as Variables {#section:relvar-approach} ---------------------- Here we interpret every element of ${{\mathit{Regions}}}$ as a constant. The topological relation between two regions is a variable with a subset of ${{\mathsf{RCC8}}}$ as its domain. Such a relation variable exists for each ordered pair of regions, and we collect all these variables in an array ${{\mathit{Rel}}}$. We write an individual relation as $$\text{variable}\quad {{\mathit{Rel}}}[a,b] \quad {\quad\text{where}\quad} {{\mathit{Rel}}}[a,b] \subseteq {{\mathsf{RCC8}}}{\quad\text{and}\quad} a,b \in {{\mathit{Regions}}}.$$ ### Integrity constraints. Relation converse and composition in this setting are captured at the constraint level. The binary constraint ${\mathsf{conv}}$ represents the converse relation table: $${\mathsf{conv}}(\, {{\mathit{Rel}}}[a,b],\; {{\mathit{Rel}}}[b,a] \,) \quad{\quad\text{for all}\quad} \{a,b\} \subseteq {{\mathit{Regions}}}.$$ The composition table is represented by the ternary constraint ${\mathsf{comp}}$, with $${\mathsf{comp}}(\,{{\mathit{Rel}}}[a,b],\; {{\mathit{Rel}}}[b,c], \; {{\mathit{Rel}}}[a,c]\,) \qquad\text{for all}\quad \{a,b,c\} \subseteq {{\mathit{Regions}}}.$$ In presence of $${{\mathit{Rel}}}[a,a] = {\mathsf{EQ}} \quad{\quad\text{for all}\quad} a \in {{\mathit{Regions}}}$$ and a ${\mathsf{conv}}$ constraint for all pairs of different regions, one ${\mathsf{comp}}$ constraint per three different regions suffices. Comments -------- By modelling the items of interest as variables and static information as constraints, the relation-variable approach yields plain finite-domain CSPs in which the solutions ([i.e.]{}, assignments) are relevant. There is a straightforward correspondence between a solution and a fully specified, consistent scenario. Obtaining the latter from a partially specified scenario amounts to the standard task of solving a finite-domain CSP. Constructing a relation-variable model means finding integrity constraints that embody the intended semantics. Once that has been established, the origin or meaning of the constraints is irrelevant. For example, a constraint solver can ignore whether ${\mathsf{comp}}$ represents the composition operation in a relation algebra; we also discuss examples below in which other restrictions on the relations must be satisfied. There is thus a clear distinction between specification and execution. The relation-variable approach is declarative in a strict sense. ### Constraint propagation. The relation-variable approach is independent of the particular constraint solving method. We could, however, choose a solver based on search and propagation, and furthermore we could choose a GAC-enforcing propagation algorithm. Path consistency in the relation-constraint approach and generalised arc-consistency in the relation-variable approach simulate each other. This can be seen by analysing, in both approaches, the removal of one topological relation from the disjunctive constraint $a \mathbin{{{\mathit{Rel}}}} b$, or from the domain of the variable ${{\mathit{Rel}}}[a,b]$, respectively. The reason in both cases must be the lack of supporting relations between $a,c$ and $b,c$, for some third region $c$; that is, compositional consistency. ### Complexity. It is perhaps not surprising but useful to mention that establishing the respective local consistency in either approach ([i.e.]{}, PC and GAC) requires the same computational effort. Let $n$ denote the number of regions. Enforcing PC by an algorithm as the one given in [@mackworth:1977:consistency] requires time in $O(n^3)$ [@mackworth:1985:complexity]. For this, one assumes that one PC step, restricting $a \mathbin{{{\mathit{Rel}}}} c$ by $a \mathbin{{{\mathit{Rel}}}} b$ and $b \mathbin{{{\mathit{Rel}}}} c$, takes constant time. Analogue reasoning entails that GAC can be enforced in constant time on a single ${\mathsf{comp}}( {{\mathit{Rel}}}[a,b], {{\mathit{Rel}}}[b,c], {{\mathit{Rel}}}[a,c] )$ constraint — observe that the three variables have domains of size at most eight. In this way, the overall time complexity depends only on the number of such constraints, and is thus in $O(n^3)$. ### Previous work. Tsang [@tsang:1987:consistent] describes the relation-variable approach in qualitative temporal reasoning, a field similar to QSR. The idea appears not to have caught on, however. One reason is probably that integration in temporal reasoning is simpler because the canonical representation of time points on the real line exists. By referring to its end points, a time interval can directly be related to its duration or another time interval. Space, in contrast, has no such convenient canonical representation — but many aspects to be integrated. In QSR, the possibility of the relation-variable approach is mentioned occasionally in passing, but without examining its potential. For actually modelling and solving QSR problems using relation variables I am only aware of [@apt:2003:principles pages 30-33], which deals with a single aspect (topology) only. Relation Variables in Use {#section:integration} ========================= An essential advantage of the relation-variable approach is that the relevant information is available in variables. This means that linking pieces of information reduces to merely stating additional constraints on the variables. In that way, embedding a QSR problem into an application context or adding side restrictions, for example, can be dealt with easily and declaratively. We illustrate the issue of composite models with the case of aspect integration. Combining Topology and Size {#section:topology-size} --------------------------- Following Gerevini and Renz [@gerevini:2002:combining], we study scenarios combining topological and size information. We collect information about both these aspects and their link in one CSP. Let $n$ be the number of regions. ### Topological aspect. As in Section \[section:relvar-approach\], the $$n \times n \text{ array } {\mathit{TopoRel}}$$ of RCC-8 relation variables stores the topological relation between two regions. The integrity constraints ${\mathsf{conv}}_{{\mathrm{RCC8}}}$, ${\mathsf{comp}}_{{\mathrm{RCC8}}}$ need to hold. ### Size aspect. Relative size of regions is captured by one of $\{<, =, >\}$, as in [@gerevini:2002:combining]. The $$n \times n \text{ array } {\mathit{SizeRel}}$$ of variables stores the relative sizes of region pairs. The converse relation and composition tables are straightforward; the integrity constraints are $$\begin{aligned} {\mathsf{conv}}_{{\mathrm{Size}}} &= \{\, (<, >),\; (=, =),\; (>, <) \,\}, \qquad\text{and}\\ {\mathsf{comp}}_{{\mathrm{Size}}} &= \{\, (<, <, <),\; (<, =, <),\; \ldots \} \qquad\text{(13 triples).}\end{aligned}$$ ### Linking the aspects. The topological relation between two regions is dependent on their relative size. A table with this information is given in [@gerevini:2002:combining], it contains rules such as the following: $$\begin{array}{lcl} {\mathit{TopoRel}}[x,y] = {\mathsf{TPP}} &{\qquad\text{implies}\qquad}& {\mathit{SizeRel}}[x,y] = (<), \\ {\mathit{SizeRel}}[x,y] = (=) &{\qquad\text{implies}\qquad}& {\mathit{TopoRel}}[x,y] \in \{{\mathsf{DC}}, {\mathsf{EC}}, {\mathsf{PO}}, {\mathsf{EQ}}\}. \end{array}$$ In [@gerevini:2002:combining], these rules represent a meta constraint. Here, we infer the linking constraint $${\mathsf{link}}_{{\mathrm{Topo\&Size}}} = \{\, ({\mathsf{TPP}}, <),\; ({\mathsf{DC}}, =),\; \ldots \} \qquad\text{(14 pairs)}$$ which is to be stated as $${\mathsf{link}}_{{\mathrm{Topo\&Size}}}(\, {\mathit{TopoRel}}[a,b],\; {\mathit{SizeRel}}[a,b] \,)$$ for all regions $a,b$. #### Example. Let us pick up the combined scenario from [@gerevini:2002:combining p. 14]. Five regions, denoted by $\{0,\ldots,4\}$, are constrained by $$\begin{aligned} &{\mathit{TopoRel}}[0,2] \in \{{\mathsf{TPP}}, {\mathsf{EQ}}\} &&{\mathit{SizeRel}}[0,2] \in \{<\} \\ &{\mathit{TopoRel}}[1,0] \in \{{\mathsf{TPP}}, {\mathsf{EQ}}, {\mathsf{PO}}\} &&{\mathit{SizeRel}}[3,1] \in \{<,=\} \\ &{\mathit{TopoRel}}[1,2] \in \{{\mathsf{TPP}}, {\mathsf{EQ}}\} &&{\mathit{SizeRel}}[2,4] \in \{<,=\} \\ &{\mathit{TopoRel}}[4,3] \in \{{\mathsf{TPP}}, {\mathsf{EQ}}\}\end{aligned}$$ Independently, the topological and the size scenarios are consistent while the combined scenario is not. It is pointed out in [@gerevini:2002:combining] that naive propagation scheduling schemes do not suffice to detect inconsistency. A formulation of this scenario as a combined topological & size CSP in the relation-variable approach is straightforward. The resulting CSP can be entered into a constraint programming platform such as [@wallace:1997:eclipse]. is focused on search and domain-reducing propagation; in particular, it offers a GAC-enforcing propagation algorithm for user-defined constraints. Given our CSP in , solely executing GAC-propagation for all constraints yields failure, which proves that this CSP is inconsistent. For the same purpose but within the relation-constraint approach, Gerevini and Renz proposed a new algorithm called <span style="font-variant:small-caps;">Bipath-consistency</span> [@gerevini:2002:combining]. Its principle is the computation of path-consistency for both types of relations in an interleaved fashion while taking into account the interdependency. The ${\mathsf{link}}_{{\mathrm{Topo\&Size}}}$ constraint is in essence treated as a *meta constraint* on the algorithm level. Moreover, the <span style="font-variant:small-caps;">Bipath-consistency</span> algorithm fixes in part the order of propagation. The relation-variable method, on the other hand, is declarative; all information is in the five types of constraints. They are handled by repeated, interleaved calls to the same GAC-enforcing algorithm. The actual propagation order is irrelevant for the result. <span style="font-variant:small-caps;">Bipath-consistency</span> is restricted to combining two types of relations ([e.g.]{}, two aspects of space). In contrast, the relation-variable approach is compositional in the sense that adding a third aspect, such as morphology [@cristani:1999:complexity] or orientation, is straightforward. It amounts to formulating integrity constraints ([e.g.]{}, ${\mathsf{conv, comp}}$), linking constraints to each of the already present aspects, and a constraint linking all three aspects. Some of these constraints may be logically redundant. Combining Cardinal Directions and Topology {#section:cardinal-directions} ------------------------------------------ In orientation, another important aspect of space, one studies the relation of two objects, the primary and the reference object, with respect to a frame of reference. It is thus inherently a ternary relation, but by agreeing on the frame of reference, a binary relation is obtained. The binary relation approach is realised in the cardinal direction model [@frank:1992:qualitative], based on the geographic (compass) directions. Points as well as regions have been studied as the objects to be oriented. The point-based models can be cast in the relation-variable approach analogously to topology, Section \[section:relvar-approach\]. For instance, Frank [@frank:1992:qualitative] distinguishes the jointly exhaustive and pairwise disjoint relations ${\mathsf{N}}, {\mathsf{NW}}, {\mathsf{W}}, \ldots$ for points; denoting North, Northwest, West, and so on. Ligozat [@ligozat:1998:reasoning] gives a composition table. ### Orienting regions. Goyal and Egenhofer [@goyal:1997:direction] and Skiadopoulos and Koubarakis [@skiadopoulos:2001:composing] study a more expressive model, in which the oriented objects are regions. The exact shape of the primary region is taken into account, and a ninth atomic relation ${\mathsf{B}}$ exists, describing overlap of the primary region and the axes-parallel minimum bounding box of the reference region. *Sets* of the atomic relations are then used to describe directional information. In this way, for example, the position of South America for an observer located in Ecuador can be fully described by the set $\{{\mathsf{B}}, {\mathsf{N}}, {\mathsf{NE}}, {\mathsf{E}}, {\mathsf{SE}}, {\mathsf{S}}\}$. In contrast, the position of Ecuador with respect to South America is just $\{{\mathsf{B}}\}$. Relation variables for directional information are thus naturally *set variables*: they take their value from a set of sets of constants, unlike relation variables for topology and size whose domain is a set of atomic constants. For each pair $a,b$ of regions, the direction is a relation variable $${\mathit{DirRel}}[a,b] \in \mathcal{P}({\mathsf{Dir}}) \qquad \text{where}\quad {\mathsf{Dir}} = \{{\mathsf{B}}, {\mathsf{N}}, {\mathsf{NW}}, \ldots, {\mathsf{NE}}\}.$$ $\mathcal{P}$ denotes the power set function. ### Integrity constraints. A restriction on the set values that ${\mathit{DirRel}}[a,b]$ can take arises if $a,b$ are internally connected regions, which is often assumed. Only 218 of the 512 subsets of ${\mathsf{Dir}}$ are then semantically possible. This knowledge can be represented in a unary integrity constraint, which for example allows $\{{\mathsf{N}}, {\mathsf{NE}}, {\mathsf{E}}\}$ but excludes $\{{\mathsf{N}}, {\mathsf{S}}\}$. The usual integrity constraints ${\mathsf{comp}}$ and ${\mathsf{conv}}$ can be derived from studies of composition [@skiadopoulos:2001:composing] and converse [@cicerone:2004:cardinal] (but it is outside of our focus whether these are the only integrity constraints needed). ### Integration with topology. Let us briefly consider linking directional information to topology. The relevant knowledge could be expressed by rules as $$\begin{array}{lcl} {\mathit{TopoRel}}[x,y] \in \{{\mathsf{EQ}},{\mathsf{NTPP}},{\mathsf{TPP}}\} &{\qquad\text{implies}\qquad}& {\mathit{DirRel}}[x,y] = \{{\mathsf{B}}\}, \\ {\mathit{TopoRel}}[x,y] \in \{{\mathsf{NTPPi}},{\mathsf{TPPi}}\} &{\quad\text{implies}\quad}& {\mathit{DirRel}}[x,y] \supseteq \{{\mathsf{B}}\},\\ \end{array}$$ from which a constraint ${\mathsf{link}}_{{\mathrm{Topo\&Dir}}}$ can be defined. It is to be stated as $${\mathsf{link}}_{{\mathrm{Topo\&Dir}}}(\, {\mathit{TopoRel}}[a,b],\; {\mathit{DirRel}}[a,b] \,)$$ for all regions $a,b$. We now have some components of a combined cardinal directions & topology model. It can be given to any sufficiently expressive constraint solver, which in particular would provide constraints on set variables. Constraint solving with set variables is discussed in [@gervet:1997:interval]. Many contemporary constraint programming systems support set variables. Cyclic Ordering of Orientations with Relation Variables ------------------------------------------------------- From the several formalisations of orientation information with an explicit frame of reference, let us examine the approach of Isli and Cohn to cyclic ordering of 2D orientations [@isli:2000:new]. Here, the spatial objects are orientations, [i.e.]{} directed lines. At the root of the framework is the qualitative classification of the angle $\alpha = \sphericalangle(a, b)$ between the two orientations $a$ and $b$ by $${\mathsf{Or}}(\alpha) = \begin{cases} {\mathsf{e}} {\quad\text{(equal)}\quad} &\text{if}\quad \alpha = 0,\\ {\mathsf{l}} {\quad\text{(left)}\quad} &\text{if}\quad 0 < \alpha < \pi,\\ {\mathsf{o}} {\quad\text{(opposite)}\quad} &\text{if}\quad \alpha = \pi,\\ {\mathsf{r}} {\quad\text{(right)}\quad} &\text{if}\quad \pi < \alpha < 2\pi \end{cases}$$ into the jointly exhaustive and pairwise disjoint relations ${\mathsf{e}},{\mathsf{l}},{\mathsf{o}},{\mathsf{r}}$. See Fig. \[fig:relvars:orientations\] for an illustration. (0,0)(25,7) (1,6)(1,0) (0.5,5.2)[a]{} (1.5,4.5)[b]{} [ (3;90)(3;270) (3;50)(3;230) ]{}(7.3,5.3)[a]{} (9.8,4)[b]{} (15,6)(15,0) (14.5,5.2)[a]{} (15.5,0.8)[b]{} [ (3;90)(3;270) (3;200)(3;20) ]{}(21.5,5.2)[a]{} (20,3)[b]{} For three orientations $a,b,c$, we now consider the triple $$\langle\ {\mathsf{Or}}(\sphericalangle(b,a)),\; {\mathsf{Or}}(\sphericalangle(c,b)),\; {\mathsf{Or}}(\sphericalangle(c,a)) \ \rangle.$$ Of all $4^3$ triples over $\{{\mathsf{e}},{\mathsf{l}},{\mathsf{o}},{\mathsf{r}}\}$, only 24 combinations are geometrically possible. We denote this set by ${\mathsf{Cyc}}$. Fig. \[fig:relvars:cyc\] shows three of its elements. (0,0)(22,7) [ (3;90)(3;270) (3;135)(3;-45) (3;200)(3;20) ]{}(3.8,5.5)[a]{} (1,1.5)[b]{} (1,4)[c]{} [ (3;90)(3;270) (3;135)(3;-45) ]{}(11.8,5.5)[a]{} (10,4.7)[b]{} (9,4)[c]{} [ (3;90)(3;270) (3;230)(3;50) ]{}(19.7,5.5)[a]{} (17.2,2)[b]{} (21,4.5)[c]{} Such cyclic ordering information can be expressed within the relation variable approach in an array ${\mathit{CycRel}}$, which in particular is ternary. We have thus a relation variable $${\mathit{CycRel}}[a,b,c] \in {\mathsf{Cyc}} \qquad \text{with}\quad {\mathsf{Cyc}} = \{ {\mathsf{lrl}}, {\mathsf{orl}}, \ldots, {\mathsf{rle}} \}$$ for every three orientations $a,b,c$. The integrity constraints here are $$\begin{gathered} {\mathsf{conv}}(\,{\mathit{CycRel}}[a,b,c],\; {\mathit{CycRel}}[a,c,b]\,),\\ {\mathsf{comp}}(\,{\mathit{CycRel}}[a,b,c],\; {\mathit{CycRel}}[a,c,d],\; {\mathit{CycRel}}[a,b,d]\,),\end{gathered}$$ and a new constraint $$\begin{gathered} {\mathsf{rotate}}(\,{\mathit{CycRel}}[a,b,c],\; {\mathit{CycRel}}[c,a,b]\,).\end{gathered}$$ Details and definitions can be found in [@isli:2000:new]. Working within the relation constraint approach, Isli and Cohn construct a new algorithm called *s4c* that enforces $4$-consistency [@freuder:1978:synthesizing] on the ternary relation constraints that correspond to ${\mathit{CycRel}}$. They are able to prove that this algorithm decides consistency, [i.e.]{}, 2D geometric feasibility, of fully specified scenarios. The *s4c* algorithm uses exactly the information that we represent in the ${\mathsf{conv}}$, ${\mathsf{comp}}$ and ${\mathsf{rotation}}$ constraints. Consequently, we can conclude that in our relation variable model these constraints guarantee geometric consistency. We hypothesise further that *s4c* in the relation constraint model propagates at most as much information as a GAC-enforcing algorithm does in our relation variable model. Intuitively, this should be clear: every possible reduction of a disjunctive constraint in the relation constraint model corresponds to a domain reduction of a relation variable in our model. Combining Cardinal Direction with Relative Orientation ------------------------------------------------------ Isli [@isli:2003:combining; @isli:2004:combining] studies the problem of exchanging information between a cardinal direction model for pairs of points as in Section \[section:cardinal-directions\], and a relative orientation model for triples of points, derived from Freksa and Zimmermann’s formalisation [@freksa:1992:utilization]. This problem is again similar to combining topology and size, Section \[section:topology-size\]. Isli works with the relation-constraints and proposes a new algorithm for this integration issue. We formulate a relation-variable model. The cardinal direction subproblem can straightforwardly be expressed in this approach; we omit the obvious details here. The relative orientation subproblem leads to a model similar to that of orientations in the preceding section; in particular, it is based on a ternary array. The arrays in the combined model are: $$\begin{aligned} &n \times n \times n \text{ array } {\mathit{ROrientRel}}, \text{ and}\quad\\ &n \times n \text{ array } {\mathit{CDirRel}},\end{aligned}$$ if we assume $n$ points. For linking the two models, Isli [@isli:2004:combining] devises functions for both directions of the information transfer. They can be transformed into the two constraints $$\begin{array}{l} {\mathsf{link}}_{{\mathrm{CD}}\rightarrow{\mathrm{RO}}} (\, {\mathit{CDirRel}}[a,b],\; {\mathit{CDirRel}}[b,c],\; {\mathit{ROrientRel}}[a,b,c] \,), \\ {\mathsf{link}}_{{\mathrm{CD}}\leftarrow{\mathrm{RO}}} (\, {\mathit{ROrientRel}}[a,b,c],\; {\mathit{CDirRel}}[a,b],\; {\mathit{CDirRel}}[b,c],\; {\mathit{CDirRel}}[a,c] \,). \end{array}$$ For the relation-constraint model it is necessary to treat the information in ${\mathsf{link}}_{{\mathrm{CD}}\rightarrow{\mathrm{RO}}}, {\mathsf{link}}_{{\mathrm{CD}}\leftarrow{\mathrm{RO}}}$ as meta-constraints, embedded inside an algorithm that moreover integrates *s4c* of [@isli:2000:new] and a path-consistency algorithm. Using relation variables, it suffices to state the constraints and provide a generic GAC-enforcing algorithm. Also, for a given triple of points, the first constraint ${\mathsf{link}}_{{\mathrm{CD}}\rightarrow{\mathrm{RO}}}$ should just be the restriction of the second constraint ${\mathsf{link}}_{{\mathrm{CD}}\leftarrow{\mathrm{RO}}}$ in which the variable ${\mathit{CDirRel}}[a,c]$ is projected away. The former constraint is then redundant, and we just need one constraint $${\mathsf{link}}_{{\mathrm{CD}}\&{\mathrm{RO}}} (\, {\mathit{ROrientRel}}[a,b,c],\; {\mathit{CDirRel}}[a,b],\; {\mathit{CDirRel}}[b,c],\; {\mathit{CDirRel}}[a,c] \,).$$ On the grounds that both the relation-variable and the relation-constraint approach are based on the same semantic information, for one embedded in an algorithm, for the other in constraints, we conclude that both accept exactly the same point configuration scenarios. Extensions ========== ### Variables ranging over spatial objects. In the relation-variable model, spatial objects are denoted by constants. An *object variable*, whose domain is the set of object constants, has thus a different meaning than in the relation-constraint approach. This issue is best demonstrated by an example. Suppose we wish to identify two regions among all given regions such that - the first is smaller than the second, and - they are disconnected or externally connected. We use topological and size information, formalised as in Section \[section:topology-size\], so we have arrays ${\mathit{SizeRel}}$ and ${\mathit{TopoRel}}$ recording the qualitative relations. Let ${{\mathit{Regions}}}$ be the set of the $n$ region constants. We define the $$\text{region variables } x_1, x_2$$ whose domain is the set ${{\mathit{Regions}}}$, and constrain them by $$\begin{aligned} &{\mathit{SizeRel}}[x_1, x_2] = (<), \label{eq:regionarrays-example-1}\tag{$C_1$} \\ &{\mathit{TopoRel}}[x_1, x_2] \in \{{\mathsf{DC}},{\mathsf{EC}}\}. \label{eq:regionarrays-example-2}\tag{$C_2$}\end{aligned}$$ $C_1$ is a constraint on the variables $x_1,x_2$ and on all size relation variables in the array ${\mathit{SizeRel}}$. Namely, region constants $r_1,r_2 \in {{\mathit{Regions}}}$ must be assigned to $x_1,x_2$ such that the size relation variable ${\mathit{SizeRel}}[r_1,r_2]$ is assigned a ‘$<$’. We call such constraints, in which arrays are indexed by variables instead of constants, *array constraints*. They are a generalisation of the better-known constraint, which corresponds to a one-dimensional array indexed by a variable. Constraint propagation to establish GAC for array constraints is studied in [@brand:2001:constraint2]. The constraint programming system ILOG Solver [@ilog:2001:solver] accepts and propagates array constraints. ### Reasoning about spatial change. It is not difficult to augment a relation-variable model with temporal information. It suffices to add a new time index to each array of qualitative relations, and to link the new time-annotated scenarios appropriately. We extend ${{\mathit{Rel}}}$ from a binary to a ternary array such that $${{\mathit{Rel}}}[a, b, t]$$ is a variable specifying the relation between the spatial objects $a$ and $b$ at time $t$. Suppose we view time as linear and discrete, such that only atomic relational changes can occur between subsequent time points. We can specify these atomic changes (the so-called conceptual neighbourhood) by pairs of qualitative relations and define accordingly a new binary constraint ${\mathsf{neighbour}}$. For example, the pair $({\mathsf{DC}}, {\mathsf{EC}})$ in the constraint ${\mathsf{neighbour}}_{{\mathrm{Topo}}}$ indicates that the topological relation *disconnected* between two regions may change in one time step to *externally connected*. The ${\mathsf{neighbour}}$ constraint is then stated on all variable pairs $({{\mathit{Rel}}}[a, b, t], {{\mathit{Rel}}}[a, b, t'])$ where $t$ directly precedes $t'$ temporally. Summary ======= We have presented an alternative formulation of qualitative spatial reasoning problems as constraint satisfaction problems. Contrary to the conventional approach, we model qualitative relations as variables. Uncertain relational information is naturally expressed by variables with domains; consistency of this information is naturally expressed by static constraints. The propagation of these constraints is a well-understood issue in research on constraint programming, and corresponding generic algorithms are provided by many constraint solving systems. While the principle of the relation-variable approach is not new, the advantages of applying it to QSR, especially for integration tasks, have so far very rarely been realised. We have argued that several algorithms that are custom-designed for integrating spatial aspects become unnecessary if a relation-variable model and a generic GAC-establishing constraint propagation algorithm is used: the <span style="font-variant:small-caps;">Bipath-consistency</span> algorithm of [@gerevini:2002:combining], the *s4c* algorithm of [@isli:2000:new], the algorithm combining *s4c* and a path-consistency algorithm of [@isli:2004:combining]. We have shown how the relation-variable approach can accommodate composite qualitative relations as investigated in [@cicerone:2004:cardinal; @skiadopoulos:2001:composing] with the help of set variables and constraints. We have indicated that extending or combining a relation-variable model often consists mainly in defining appropriate constraints, contrary to what is the case in the relation-constraint approach where new algorithms must be designed. Finally, we remark that the strictly declarative model that results from using relation-variables can be solved by any sufficiently expressive solver of CSPs. This includes typical CP systems based on search and propagation, but also for example solvers based on local search. [10]{} K. R. Apt. . Cambridge University Press, 2003. B. Bennett. Determining consistency of topological relations. , 2:213–225, 1998. C. Bessi[è]{}re and J.-C. R[é]{}gin. Arc consistency for general constraint networks: preliminary results. In [*Proc. of 15th International Joint Conference on Artificial Intelligence ([IJCAI]{}’97)*]{}, pages 398–404, 1997. S. Brand. Constraint propagation in presence of arrays. In K. R. Apt, R. Bart[á]{}k, E. Monfroy, and F. Rossi, editors, [ *Proc. of 6th Workshop of the ERCIM Working Group on Constraints*]{}, 2001. S. Cicerone and P. Di Felice. Cardinal directions between spatial objects: the pairwise-consistency problem. , 164:165–188, 2004. A. G. Cohn and S. M. Hazarika. Qualitative spatial representation and reasoning: An overview. , 46(1-2):1–29, 2001. M. Cristani. The complexity of reasoning about spatial congruence. , 11:361–390, 1999. R. Dechter. . Morgan Kaufmann, 2003. A. U. Frank. Qualitative spatial reasoning about distance and directions in geographic space. , 3:343–373, 1992. C. Freksa and K. Zimmermann. On the utilization of spatial structures for cognitively plausible and efficient reasoning. In [*Proc. of IEEE International Conference on Systems, Man, and Cybernetics*]{}, pages 18–21. IEEE, 1992. E. C. Freuder. Synthesizing constraint expressions. , 21(11):958–966, 1978. T. Fr[ü]{}hwirth and S. Abdennadher. . Springer, 2003. A. Gerevini and J. Renz. Combining topological and size constraints for spatial reasoning. , 137(1-2):1–42, 2002. C. Gervet. Interval propagation to reason about sets: [D]{}efinition and implementation of a practical language. , 1(3):191–244, 1997. R. K. Goyal and M. J. Egenhofer. The direction-relation matrix: A representation of direction relations for extended spatial objects. In [*Proc. of UCGIS Annual Assembly and Summer Retreat*]{}, 1997. ILOG S.A. , 2001. A. Isli. Combining cardinal direction relations and relative orientation relations in qualitative spatial reasoning. Technical report, University of Hamburg, Dept. of Informatics, 2003. A. Isli. Combining cardinal direction relations and other orientation relations in [QSR]{}. In [*Proc. of 8th International Symposium on Artificial Intelligence and Mathematics ([AI&M]{}’04)*]{}, 2004. A. Isli and A. G. Cohn. A new approach to cyclic ordering of [2D]{} orientations using ternary relation algebras. , 122(1-2):137–187, 2000. G. Ligozat. Reasoning about cardinal directions. , 9(1):23–44, 1998. A. K. Mackworth. Consistency in networks of relations. , 8(1):118–126, 1977. A. K. Mackworth and E. C. Freuder. The complexity of some polynomial network algorithms for constraint satisfaction problems. , 25:65–74, 1985. R. Mohr and G. Masini. Good old discrete relaxation. In Y. Kodratoff, editor, [*Proc. of European Conference on Artificial Intelligence ([ECAI]{}’88)*]{}, pages 651–656. Pitman publishers, 1988. U. Montanari. Networks of constraints: Fundamental properties and applications to picture processing. , 7:95–132, 1974. D. A. Randell, Z. Cui, and A. G. Cohn. A spatial logic based on regions and connection. In B. Nebel, C. Rich, and W. R. Swartout, editors, [*Proc. of 2nd International Conference on Principles of Knowledge Representation and Reasoning ([KR]{}’92)*]{}, pages 165–176. Morgan Kaufmann, 1992. S. Skiadopoulos and M. Koubarakis. Composing cardinal direction relations. In C.S. Jensen, M. Schneider, B. Seeger, and V.J. Tsotras, editors, [*Proc. of 7th International Symposium on Advances in Spatial and Temporal Databases ([SSTD]{}’01)*]{}, volume 2121 of [*LNCS*]{}, pages 371–386. Springer, 2001. E. P. K. Tsang. The consistent labeling problem in temporal reasoning. In K. S. H. Forbus, editor, [*Proc. of 6th National Conference on Artificial Intelligence ([AAAI]{}’87)*]{}, pages 251–255. AAAI Press, 1987. M. G. Wallace, S. Novello, and J. Schimpf. : A platform for constraint logic programming. , 12(1):159–200, 1997.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Equations of state for strange quark matter in beta equilibrium at high densities are used to investigate the structure (mass and radius) of compact objects. The chromodielectric model is used as a general framework for the quark interactions, which are mediated by chiral mesons, $\sigma$ and $\vec \pi$, and by a confining chiral singlet dynamical field, $\chi$. Using a quartic potential for $\chi$, two equations of state for the same set of model parameters are obtained, one with a minimum at around the nuclear matter density $\rho_0$ and the other at $\rho \sim 5 \, \rho_0$. Using the latter equation of state in the Tolman-Oppenheimer-Volkoff equations we found solutions corresponding to compact objects with $R\sim 5 - 8$ km and $M\sim M_\odot$. The phenomenology of recently discovered X-ray sources is compatible with the type of quark stars that we have obtained.' author: - 'M. Malheiro' - 'E.O. Azevedo' - 'L.G. Nuss' - 'M. Fiolhais' - 'A. Taurines' title: Small quark stars in the chromodielectric model --- [ address=[Instituto de Física, Universidade Federal Fluminense, Av. Litorânea, 24210-310 Niterói, Brazil]{} ]{} [ address=[Instituto de Física, Universidade Federal Fluminense, Av. Litorânea, 24210-310 Niterói, Brazil]{} ]{} [ address=[Instituto de Física, Universidade Federal Fluminense, Av. Litorânea, 24210-310 Niterói, Brazil]{} ]{} [ address=[Departamento de Física and Centro de Física Computacional, Universidade de Coimbra, P-3004-516 Coimbra, Portugal]{} ]{} [ address=[Instituto de Física, Universidade Federal do Rio Grande do Sul CP 15051, 91501-970 Porto Alegre, Brazil]{} ]{} Introduction ============ Various effective models using quarks as fundamental dynamical fields, originally designed for the nucleon, have also been successfully used to describe infinite quark matter, and the resulting equations of state (EOS’s) have been applied to investigate the structure of compact stars \[1–6\]. The chromodielectric model (CDM) \[7–9\], for example, provides a reasonable phenomenology for the nucleon [@neuber; @drago1] and also allows us to obtain EOS’s for dense quark matter. The model yields soliton solutions representing single baryons with three quarks dynamically confined by a scalar field, $\chi$, whose quanta can be assigned to $0^{++}$ glueballs. When it is applied to quark matter in two or three flavors \[12–14\] the resulting EOS turns out to be relatively soft at large densities. Using a quadratic potential for the $\chi$ field, Drago et al. [@drago; @drago3] applied the CDM to describe the inner part of neutron stars, obtaining masses in the range $1-2 M_\odot$ and radii of the order 10 km or higher, with a small hadron crust of 2 km. In this work we consider an extension of the model used in Ref. [@drago], taking quartic instead of quadratic potentials. In addition to the structures found by Drago et al., the quartic model predicts another type of compact objects made out of quarks only, smaller and denser than neutron stars. From the observational point of view, the recent discovery of X-ray sources, by the Hubble and Chandra telescopes, increased the plausibility that these sources might be strange quark stars \[17–19\]. In particular, the isolated compact object RX J1856.5-3574 with a small radius does not show evidence of spectral lines or edge features [@pons; @drake], reinforcing the conjecture for the existence of stars made out of strange matter. The phenomenology of these objects seems to be compatible with the small and dense quark stars reported in this work. The model ========= CDM Lagrangian -------------- We write the CDM Lagrangian in the form \[8–10\] $$\mathcal{L} = \mathcal{L}_q + \mathcal{L}_{\sigma,\pi} + \mathcal{L}_{q-\mathrm{meson}} + \mathcal{L}_\chi\;, \label{langrangian}$$ where $$\mathcal{L}_q = \mathrm{i}\bar{\psi}\gamma^\mu \partial_\mu\psi \;, \qquad \mathcal{L}_{\sigma,\pi} = \half\partial_\mu\hat{\sigma}\partial^\mu\hat{\sigma} + \half\partial_\mu\hat{\vec{\pi}}\cdot\partial^\mu\hat{\vec{\pi}} - {W}(\hat{\vec{\pi}},\hat{\sigma})\;, \label{langrangian1}$$ and ${W}(\hat{\vec{\pi}},\hat{\sigma})$ is the Mexican hat potential. In the $u,\, d$ sector the quark-meson interaction is described by $$\mathcal{L}_{q-{\mathrm{meson}}} = {g\over\chi}\, \bar{\psi} (\hat{\sigma}+\mathrm{i}\vec{\tau}\cdot\hat{\vec{\pi}}\gamma_5) \psi\;. \label{langrangian2}$$ The last term in (\[langrangian\]) contains the kinetic and the potential piece for the $\chi$-field: $$\mathcal{L}_\chi = \half\partial_\mu\hat{\chi}\,\partial^\mu\hat{\chi} - U(\hat{\chi})\, . \label{langrangian3}$$ The potential term is $$U(\chi)= {1 \over 2} M^2 \chi^2 \left[ 1+\left( {8 \eta^4 \over \gamma^2}-2 \right) {\chi \over \gamma M} + \left( 1- {6 \eta^4 \over \gamma^2} \right) {\chi^2 \over (\gamma M)^2} \right] \, , \label{uchi}$$ where $M$ is the $\chi$ mass. The parameterization used in (\[uchi\]) allows for a physically meaningful interpretation of the parameters $\gamma$ and $\eta$: $U(\chi)$ has a global minimum at $\chi=0$ and a local one at $\chi=\gamma M$, and $U(\gamma M)=\eta^4 M^4$ (see Fig. \[figstar1\]). ![Quartic potential, Eq. (\[uchi\]), for fixed $M$ and $\gamma$, in dependence of $\chi$. For comparison, the quadratic potential $(\gamma \rightarrow \infty)$ is also shown. The value of the potential at the second minimum can be interpreted as a bag pressure, $B$. []{data-label="figstar1"}](fstar1){width="13cm"} The height of the local minimum, $B=(\eta M)^4$, is interpreted as a “bag pressure" and this is used to fix the parameters in $U(\chi)$. Assuming the wide range $0.150 \le B^{1/4} \leq 0.250$ GeV, one has $0.08\le \eta \le 0.15$, using $M=1.7$ GeV. We note that $\gamma$ is not a free parameter since the quartic term of $U(\chi)$ must be positive and the cubic term negative, which implies $\gamma^2\ge 6 \eta^4$. In the soliton sector of the model, best nucleon properties are obtained for $G=\sqrt{g M}\sim 0.2$ GeV (only $G$ matters for the nucleon properties) and we keep that $G$ in our quark matter calculations. In order to study strange quark matter, we add to the interaction Lagrangian (\[langrangian2\]) the term [@birsegovern] $$\mathcal{L}_{s-{\mathrm{meson}}} = {g_s \over\chi}\, \bar{\psi}_s \psi_s\;. \label{langrangian4}$$ accounting for the coupling between the strange quark and the $\chi$ field. Strange quark matter -------------------- In order to study strange quark matter in beta equilibrium, an electron gas must also be considered. The mean-field energy per unit volume for strange quark matter in the CDM (plus electrons) is given by $$\begin{aligned} \varepsilon &=& \alpha \sum_{f=u,d}\int_{0}^{k_{f}}\frac{d^3k}{(2\pi)^3}\sqrt{k^2+m_f(\sigma,\chi)^2} + \alpha \int_{0}^{k_{s}}\frac{d^3k}{(2\pi)^3}\sqrt{k^2+m_s(\chi)^2} \nonumber \\ &+& 2 \int_{0}^{k_{e}}\frac{d^3k}{(2\pi)^3}\sqrt{k^2+m_e^2} +{ U}(\chi)+{ m _\sigma^2 \over 8 \, f_\pi^2}(\sigma^2-f_\pi^2)^2, \label{eov2}\end{aligned}$$ where the first two terms refer to quarks and the third one to the electrons, all described by plane waves. The degeneracy factor is $\alpha=6$ (for spin and color). The last term is the Mexican hat potential (with $\vec\pi=0$ and $f_\pi=93$ MeV). The $k_i$ in (\[eov2\]) are the Fermi momenta of quarks and electrons. The quark masses in (\[eov2\]) are [@birsegovern]: $m_{u,d} = {g_{u,d}\, \sigma}/{(\chi \, f_\pi)} $ and $m_{s} = {g_{s}}/{\chi}$ with the coupling constants given by $g_u= g \, (f_\pi+\xi_3)$, $g_d= g \, (f_\pi-\xi_3)$ and $g_s=g\, (2 f_k-f_\pi)$ \[$\xi_3=-0.75$ MeV, $f_K=113$ MeV\]. A variational principle applied to the energy density, Eq. (\[eov2\]), leads to two gap equations for $\sigma$ and $\chi$. In the interior of a compact star the matter should satisfy both the electrical charge neutrality and chemical equilibrium. These conditions should supplement the gap equations, and altogether we have a system of six algebraic equations to solve at each baryon density $\rho= \left( \rho_u+\rho_d+\rho_s \right)/3$ \[here, $\rho_i=\alpha k_i^3 / (6\pi)^2$ stand for each flavor density\]. The solution of the system of equations are the meson fields, $\sigma$ and $\chi$, and the Fermi momenta, $k_u$, $k_d$, $k_s$ and $k_e$. For the same set of model parameters we found two stable solutions, hereafter denoted by I and II, shown in Fig. \[figstar2\]. ![Fermi momenta (solid lines) and quark masses (dashed lines) for solution I and II in dependence of the baryon density. For solution II the quark masses and the electron Fermi momentum almost vanish. The model parameters are $g=0.023$ GeV, $M=1.7$ GeV, $\gamma=0.2$ and $\eta=0.12$. The vertical scales are in fm$^{-1}$.[]{data-label="figstar2"}](fstar2){width="9.cm"} For both solutions $\sigma$ is always close to $f_\pi$. In solution I, the $\chi$ field is a slowly increasing function of the density, remaining always smaller than $\sim 0.05$ GeV. For such a small $\chi$, the quartic potential and the quadratic potential are indistinguishable (see Fig. \[figstar1\]), thus, in practice, solution I corresponds to the one obtained and used by Drago et al. [@drago] in the framework of the quadratic potential. Due to the smallness of the $\chi$ field, quark masses are large and the system is in a chiral broken phase. The solution II exhibits a large confining field, $\chi\sim \gamma M$ (local minimum of $U$), independent of the density. The resulting quark masses are similar for the three flavors and very close to zero (chiral restored phase). Therefore, the chemical potentials in solution II are dominated by the Fermi momentum contribution, $\mu_u\simeq\mu_d\simeq\mu_s$ and $\mu_e\simeq0$, i.e. in solution II there are almost no electrons. Besides solutions I and II, there is an additional [*unstable*]{} solution corresponding to $\chi\sim \gamma M /2$ \[local maximum of $U(\chi$)\]. Equations of state ------------------ The energy per baryon number as a function of the baryon density (EOS) is readily evaluated for each solution. We stress that EOS-I is not sensitive to $\gamma$ and $\eta$ (since $\chi$ is small), just depends on $G$. The saturation density occurs at a relatively low density ($\rho=1.2-1.6 \, \rho_0$) and the behavior of EOS-I at intermediate densities, is similar to the hadronic EOS’s (see Ref. [@drago2] for the two flavors sector). The EOS-II is also insensitive to $\gamma$, but does depend on $\eta$ \[in fact, the dependence is on $(\eta M)^4$, as we have already pointed out\]: the energy per baryon number increases with $\eta$ and so does the saturation density. For $\eta\sim 0.12$ the saturation density is $\rho\sim 5 \rho_0$ and the energy per baryon number is some 20% higher than for solution I at its saturation density. ![Pressure versus energy density for the two types of solution for various parameters $\eta$. \] The other model parameters are $M=1.7$ GeV, $g=0.023$ GeV and $\gamma=0.2$.[]{data-label="figstar3"}](fstar3){width="16cm"} In Fig. \[figstar3\] we present the EOS’s in the form of pressure $P=\rho^2 {\partial \ \over \partial \rho}\left( {\epsilon \over \rho} \right)$ versus energy density plots for solutions I and II and various $\eta$. It is worth noticing that our results with two distinct EOS predicted by the CDM, are consistent with the results from perturbative QCD: it is remarkable indeed that, for $\eta\sim0.12$ the CDM reproduces accurately the EOS’s recently obtained in a perturbative QCD calculation (compare our third panel of Fig. \[figstar3\] with figure 1 of Ref. [@fraga]). Regarding energetics, both phases are almost degenerated at high densities and have similar shapes in the $P \ttimes \epsilon$ plane even at intermediate densities (or energy densities) in the narrow range $0.1 \le \eta \le 0.12$. In that region of $\eta$, one solution is not clearly lower in energy than the other. However, we should point out that they correspond to two different $\chi$ values and for the system to undergo a transition from the chiral restored to the chiral broken phase it has to go through a high potential energy barrier. In a 3D plot of the energy per baryon number versus $(\rho,\chi)$ the stable solutions correspond to two distinct “valleys", and the unstable solution mentioned at the end of the previous section corresponds to the top of the barrier between the two valleys. Quark stars =========== In order to investigate the structure of stars we solved the Tolman-Oppenheimer-Volkoff (TOV) equation using the two EOS. Since EOS-I is identical to the one using a quadratic potential, it leads to stars that have the same phenomenology as the hybrid stars obtained by Drago et al. [@drago]: $R\sim 10 - 12$ km, a hadron crust and a mass $M\sim 1- 2 M_\odot$. At low densities, hadronization occurs and an hadronic equation of state should be used, replacing EOS-I. Since EOS-II saturates at a high density and, in addition, the system is not likely to undergo a transition to solution I, one should not perform any connection to the hadronic sector: the EOS-II alone generates a new family of strange quark stars. In Fig. \[figstar4\] it is shown the mass-radius relation for different values of $\eta$. These quark stars are smaller and denser in comparison with those resulting from EOS-I. For $\eta\sim0.115$ (and $M=1.7$ GeV, yielding $B^{1/4}\sim 0.195$ GeV) one obtains a maximum radius $R\sim 6$ km and a corresponding mass $M\sim 0.9 M_\odot$, which are the fitted radius and mass for the nearby compact object RX J 1856.5-3754 [@pons; @drake]. According to our calculation, such star has a central density of $10\rho_0$ ($\rho_0$ is the nuclear matter density) and a central energy density $\epsilon\sim 3\ttimes 10^{15}$ g/cm$^3$. At the edge, the density drops to $5\rho_0$ and $\epsilon\sim 1.35\ttimes 10^{15}$ g/cm$^3$. The ratio $\epsilon/\rho$ remains approximately constant inside the star. The maximum period of the star, computed using the expression given in Ref. [@haensel], is $\sim0.4$ ms. From Fig. \[figstar4\] one concludes that the mass-radius relation for these strange small stars mainly depends on the height of the local minimum of the $\chi$ potential. ![Mass versus radius for the pure quark stars (solution II) in the CDM model. The dot indicates the maximum radius star for $\eta=0.115$. For the other model parameters see caption of Fig. \[figstar3\].[]{data-label="figstar4"}](fstar4){width="12cm"} Conclusions =========== Using a mean-field variational method we obtained two solutions for homogeneous strange matter in beta equilibrium using the CDM with a quartic potential, with its parameters fixed in the nucleon sector and to yield a reasonable bag constant. One solution is similar to the already known solution for quadratic potentials in the CDM, with massive quarks. In the other solution, quarks are almost massless. The EOS for both solutions are similar to those found recently in the framework of perturbative QCD. The pure quark stars emerging from the chiral symmetric solution are small and dense compact objects, showing a phenomenology compatible with the nearby X-ray source whose mass and radius was recently fitted [@pons; @drake]. This work was supported by FCT (POCTI/FEDER program), Portugal and by CNPq/ICCTI through the Brazilian-Portuguese scientific exchange program. We thank G. Marranghello and B. Garcia for some useful discussions. E.O.A. and L.G.N. acknowledge the support of the PIBIC/CNPq program for young researchers. [99]{} N. K. Glendenning, Compact Stars – Nuclear Physics, Particle Physics, and General Relativity (Springer, New York, 1997) F. Weber, Pulsars as Astrophysical Laboratories for Nuclear and Particle Physics, (IOP, UK, 1999) H. Heiselberg, M. Hjorth-Jensen, [*Phys. Rept.*]{} [**328**]{}, 237 (2000) M. Hanauske, L. M. Satarov, I. N. Mishustin, H. Stocker, W. Greiner, [*Phys. Rev. D*]{} [**64**]{}, 043005 (2001) E. Witten, [*Phys. Rev. D*]{} [**30**]{}, 272 (1984); C. Alcock, E. Farhi, A. Olinto, [*Astrophys. J.*]{} [**310**]{}, 261 (1986); P. Haensel, J. L. Zdunik, R. Schaeffer, [*Astron. Astrophys.*]{} [**160**]{}, 121 (1986) A. R. Taurines, C. A. Z. Vasconcellos, M. Malheiro, M. Chiapparini, [*Phys. Rev. C*]{} [**63**]{}, 065801 (2001) H. B. Nielsen, A. Patkós, [*Nucl. Phys. B*]{} [**195**]{}, 137 (1982) H. J. Pirner, [*Prog. Part. Nucl. Phys.*]{}, [**29**]{}, 33 (1992) M. K. Banerjee, [*Prog. Part. Nucl. Phys.*]{}, [**31**]{}, 77 (1993) T. Neuber, M. Fiolhais, K. Goeke, J. N. Urbano, [*Nucl. Phys. A*]{} [**560**]{}, 909 (1993) A. Drago, M. Fiolhais, U. Tambini, [*Nucl. Phys. A*]{} [**609**]{}, 488 (1996) S. K. Ghosh, S. C. Phatak, [*Phys. Rev. C*]{} [**52**]{}, 2195 (1995) A. Drago, M. Fiolhais, U. Tambini, [*Nucl. Phys. A*]{} [**588**]{}, 801 (1995) W. Broniowski, M. Čibej, M. Kutschera, M. Rosina, [*Phys. Rev. D*]{} [**41**]{}, 285 (1990) A. Drago, U. Tambini, M. Hjorth-Jensen, [*Phys. Lett. B*]{} [**380**]{}, 13 (1996); [*Prog. Part. Nucl. Phys.*]{} [**36**]{}, 407 (1996) A. Drago, A. Lavagno [*Phys. Lett. B*]{} [**511**]{}, 229 (2001) I. Bombaci, A. V. Thampan, B. Datta, [*Astrophys. J.*]{} [**541**]{}, L71 (2000) X. D. Li, I. Bombaci, M. Dey, J. Dey, E. P. J. van den Heuvel, [*Phys. Rev. Lett.*]{} [**83**]{}, 3776 (1999) M. Dey, I. Bombaci, J. Dey, S. Ray, B. C. Samanta, [*Phys. Lett. B*]{} [**438**]{}, 123 (1998) J. A. Pons, F. M. Walter, J. M. Lattimer, M. Prakash, R. Neuhäuser, P. An, Astrophys. J. [**564**]{}, 981 (2002) J. J. Drake et al., astro-ph/0204159 J. A. McGovern, M. Birse, [*Nucl. Phys. A*]{} [**506**]{}, 367 (1990); [*Nucl. Phys.*]{} [**506**]{}, 392 (1990) E. S. Fraga, R. D. Pisarski, J. Schaffner-Bielich, [*Phys. Rev. D*]{} [**63**]{}, 121702 (2001) P. Haensel, J. L. Zdunik, [*Nature*]{} [**340**]{} 617 (1989)
{ "pile_set_name": "ArXiv" }
--- abstract: 'This paper proposes an unsupervised representation model, based on rank-fusion graphs, for general applicability in multimodal tasks, either unsupervised or supervised prediction. Rank-fusion graphs encode information from multiple descriptors and retrieval models, thus being able to capture underlying relationships between modalities, samples, and the collection itself. By doing so, our method is able to promote a fusion model better than either early-fusion and late-fusion alternatives. The solution is based on the encoding of multiple ranks of a query, defined according to different criteria, into a graph. Later, we embed the generated graph into a feature space, creating fusion vectors. Those embeddings are employed to build an estimator that infers whether an input (even multimodal) object refers to a class (or event) or not. Performed experiments in the context of multiple multimodal and visual datasets, evaluated on several descriptors and retrieval models, demonstrate that our representation model is highly effective for different detection scenarios involving visual, textual, and multimodal features, yielding better detection results than state-of-the-art methods.' address: - 'Institute of Computing, University of Campinas (UNICAMP), Campinas, Brazil' - 'Université de Lorraine-LORIA UMR 7503, Vandoeuvre-lès-Nancy, France' - 'Department of ICT and Natural Sciences, Norwegian University of Science and Technology (NTNU), [Å]{}lesund, Norway' author: - Icaro Cavalcante Dourado - Salvatore Tabbone - Ricardo da Silva Torres bibliography: - 'bib.bib' title: 'Multimodal Representation Model based on Graph-Based Rank Fusion' --- multimodal fusion ,representation model ,graph-based rank fusion ,graph embedding ,rank aggregation ,prediction tasks Introduction ============ Nowadays, data analysis involving multimedia and heterogeneous content is a hot topic that attracts a lot of attention from not only public and private sectors, but also from academia. The proliferation of digital content and social media is expanding substantially the volume and diversity of available digital content. For the most part, these data is unlabeled, heterogeneous, unstructured, and derived from multiple modalities. Despite such challenges, such content is of great relevance to support the development of prediction and retrieval models. In particular, multimodal data analysis is required in several scenarios, such as , and of natural disasters such as flooding. Massive digital content has demanded the development of textual, visual, and multimedia descriptors for content-based data analysis. Despite the continuous advance on feature extractors and machine learning techniques, a single descriptor or a single modality is often insufficient to achieve effective prediction results in real case scenarios. Descriptors have specific pros and cons, because each one often focuses on a specific point of view of a single modality. For example, dedicated descriptors may be created to characterize scenes, textual descriptions, movement, symbols, signals, etc. For this reason, descriptors and retrieval models often provide complementary views, when adopted in combination. Many research works have been proposed to combine heterogeneous data sources (remotely sensed information and social media) to promote multimodal analysis. In [@zhou:2017:CBIR], authors point out the benefit of exploring and fusing multimodal features with different models. Moreover, combining different kinds of features (local vs holistic) improves substantially the retrieval effectiveness [@zhang:2015:queryRankFusion], and most search fusion approaches are based on rank fusion [@dao:2018; @Dourado:2019:FG; @zhang:2015:queryRankFusion]. Scenarios involving heterogeneous data impose a challenge of selecting features or models to combine, which is performed by either supervised or unsupervised approaches. Unsupervised approaches are necessary in the absence of labeled data, which is prominent nowadays, or in scenarios involving lower computation capabilities or large amount of data. Most previous initiatives for multimodal prediction are solely based on either CNN-based descriptors in isolation, feature concatenation [@dirsm17:BBischkeEtAl; @dirsm17:MSDaoEtAl; @dirsm17:XFuEtAl; @dirsm17:MHanifEtAl; @dirsm17:LLopezFuentesEtAl; @dirsm17:NTkachenkoEtAl], or graph-based feature-fusion [@dirsm17:KAvgerinakisEtAl; @werneck:2018:BocgBokg]. These approaches still ignore the correlation between modalities, as well as object correlation, and they are not consistently better than ranking models that do not rely on fusion. This paper presents an unsupervised representation model, based on rank-fusion graphs, for general applicability in multimodal prediction tasks, such as classification and regression. We explore and extend the concept of a rank-fusion graph, that was originally proposed as part of a rank aggregation approach for retrieval tasks [@Dourado:2019:FG]. We present a fusion method based on the representation of multiple ranks, defined according to different criteria, into a graph. Graphs provide an efficient representation of arbitrary structures and inter-relationships among different elements of a model. We embed the generated graph into a feature space, creating fusion vectors. Next, an estimator is trained to predict if an input multimodal object refers to a target label (or event) or not, following their fusion vectors. In [@Dourado:2019:FV], a methodology to apply rank-fusion graphs for efficient retrieval was presented, in the context of retrieval tasks. Here we explore those graph embedding approaches of rank-fusion graphs, now targeting a representation model for prediction tasks, either supervised or unsupervised. For this purpose, we propose specific components for the training and prediction phases, as well as a new application for the fusion vectors. Our method has the advantage to be unsupervised, it also explores and captures relationships from the collection into the representation model, and it works on top of any descriptors for multimodal data such as visual or textual. By promoting a representation model solely based on base descriptors and unsupervised data analysis over the collection, we conjecture that our approach leads to a competitive multimodal representation model that explores and encodes information from multiple descriptors and underlying sample relationships automatically, while not requiring labeled data. Experimental results over multiple multimodal and visual datasets demonstrate that the proposal is robust for different detection scenarios involving textual, visual, and multimodal features, yielding better detection results than state-of-the-art methods from both early-fusion and late-fusion approaches. This paper extends the work presented in [@Dourado:2019:fusionGraphPrediction] where we introduced the notion of a representation model from rank-fusion graphs, and demonstrated its application for multimedia flood detection. Here we extend this formulation, proposing and discussing alternative approaches for the representation model. We also evaluate the method extensively over multiple multimodal and image scenarios, to analyze its applicability for prediction tasks in general. We also evaluate the method against additional baselines, either early-fusion and late-fusion approaches. Related work is presented and discussed in a more depth way as well. Finally, we evaluate the influence of hyper-parameters and provide guidelines for their selection. Related Work ============ Representation learning models have been developed and advanced for the data modalities individually, such as image, text, video, and audio. However, their combined exploration for multimodal tasks is still an open issue. Multimodality may also impose even more challenges, depending on the scenario, such as translation between modalities, exploration of complementarity and redundancy, co-learning, and semantic alignment [@baltruvsaitis:2018:multimodal]. Some works have focused on multimodal events, which require modeling spatio-temporal characteristics of data [@tzelepis2016event]. @faria2016fusion proposed a time-series descriptor that generates recurrence plots for the series, coupled with a bio-inspired optimization of how to combine classifiers. In this paper, we focus on multimodal tasks that do not depend on temporal modeling as well as unsupervised models. In order to achieve fusion capabilities, *early-fusion* approaches emphasize the generation of composite descriptions for samples, thus working at feature level. Conversely, *late-fusion* approaches perform a combination of techniques focused on a target problem, fusing at score or decision level. On a smaller scale, some papers propose hybrid solutions based on both approaches [@zadeh2017tensor; @lan2014multimedia]. As we propose a representation model, our solution can be seen as an early-fusion approach. Nevertheless, it is based on retrieval models, without the need to work directly on feature level. In this sense, we categorize the method as hybrid. While early-fusion approaches are theoretically able to capture correlations between modalities, often a certain modality produces unsatisfactory performance and leads to biased or over-fitted models [@zadeh2017tensor]. Most early-fusion methods work in a two-step procedure, first extracting features from different modalities, then fusing them by strategies such as concatenation [@kiela2014learning], singular values decomposition [@bruni2014multimodal] or autoencoders [@silberer2014learning]. A few others focus on multimodal features jointly [@kottur2016visual; @hill2014learning], although generally restricted to a pre-defined textual attribute set. Concatenation is a straightforward yet widely used early-fusion approach, which merges vectors obtained by different descriptors. As a drawback, concatenation does not explore inherent correlations between modalities. Supervised early-fusion optimizes a weighted feature combination, either during or after feature extraction. A common strategy is to build a neural architecture with multiple separate input layers, then including a final supervised layer such as a regressor [@nogueira2018exploiting]. Another approach is the design of a composite loss function, suited for the particular desired task [@park2016image]. Composite loss functions work well in practice, but, as they need both multimodal composition and supervision, they are tied to the domain of interest. Supervised early-fusion usually suffers from high memory and time consumption costs. Besides, they usually have difficulty in preserving feature-based similarities and semantic correlations [@lan2014multimedia]. Late-fusion approaches are particularly useful when the raw data from the objects are not available. Besides, they are less prone to over-fit. Mixture of experts (MoE) approaches focus on performing decision fusion, combining predictors to address a supervised learning problem [@yuksel2012twenty]. Majority voting of classifiers [@nogueira2018exploiting], rank aggregation functions [@Dourado:2019:FG] and matrix factorization [@dong2018late] are examples of late-fusion methods. Both fusions based on rank aggregation functions and matrix factorization are based on manifold learning, i.e. the exploration of dataset geometry. Majority voting is a well-known approach to combine multiple estimators, being effective due to bias reduction. It is applied to scenarios involving an odd number of estimators, so that each predicted output is taken as the one most frequently predicted by the base estimators. Rank aggregation functions combine results from different retrieval models. They target a good permutation of retrieved objects obtained from different input ranks, in order to promote more effective retrieval results [@Dourado:2019:FG]. In general, unsupervised learning needs more investigation, specially for multimodal representation models. We explore how unsupervised rank aggregation capabilities can be applied to prediction tasks. Despite their original intent regarding better retrieval accuracy, we claim that unsupervised rank aggregation functions can provide an effective dataset exploitation. Considering previous works regarding fusion approaches for prediction tasks, a considerable amount of them are still based on classic visual descriptors [@dirsm17:MSDaoEtAl; @dirsm17:MHanifEtAl; @dirsm17:NTkachenkoEtAl; @werneck:2018:BocgBokg]. Most of them resorted to pre-trained CNN-based models for visual feature extraction [@dirsm17:KAhmadEtAl; @dirsm17:SAhmadEtAl; @dirsm17:KAvgerinakisEtAl; @dirsm17:BBischkeEtAl; @dirsm17:XFuEtAl; @dirsm17:LLopezFuentesEtAl; @dirsm17:KNogueiraEtAl], from which just a few fine-tuned their models [@dirsm17:LLopezFuentesEtAl; @dirsm17:KNogueiraEtAl]. When dealing with specific tasks, specially for competitions, some works use preprocessing steps, such as image cropping and filtering [@dirsm17:ZZhao], but this is beyond our general intent in this paper, regarding fusion methods and representation models. In order to explore the textual modality, most initiatives used BoW, using either TF or TF-IDF weighting, while others presented more complex formulations, such as word embeddings [@dirsm17:BBischkeEtAl; @dirsm17:NTkachenkoEtAl], Long Short-Term Memory (LSTM) networks [@dirsm17:LLopezFuentesEtAl], or relation networks [@dirsm17:KNogueiraEtAl]. Regarding multimodal scenarios, most works rely on early-fusion approaches, such as a concatenation of visual and textual feature vectors [@dirsm17:BBischkeEtAl; @dirsm17:MSDaoEtAl; @dirsm17:XFuEtAl; @dirsm17:MHanifEtAl; @dirsm17:LLopezFuentesEtAl; @dirsm17:NTkachenkoEtAl] or graph-based attribute fusion [@dirsm17:KAvgerinakisEtAl; @werneck:2018:BocgBokg], while only a few others adopted late-fusion approaches [@dirsm17:KAhmadEtAl; @dirsm17:SAhmadEtAl]. Preliminaries {#secPrelim} ============= Let a *sample* $s$ be any digital object, such as a document, an image, a video, or even a hybrid (multimodal) object. A sample is characterized by a *descriptor* ${\mathcal{D}}$, which relies on a particular point of view to describe $s$ as a vector, graph or another data structure $\epsilon(s)$. Descriptions allow samples to be compared to each other. Therefore, descriptors are the basis for retrieval and learning models. A *comparator* ${\mathcal{C}}$, applied over a tuple $(\epsilon(s_i),\epsilon(s_j))$, produces a *score* ${\varsigma}\in \mathbb{R^+}$ (e.g., the Euclidean distance or the cosine similarity). Either similarity or dissimilarity functions can be used to implement ${\mathcal{C}}$. A query sample, or just *query* $q$, refers to a particular sample taken as an input object in the context of a search, whose purpose is to retrieve *response items* from a *response set* ($S$) according to relevance criteria. A response set ${S = \{ s_i\}}_{i=1}^n$ is a collection of $n$ *samples*, where $n$ is the collection size. In the context of prediction tasks, a query is called *test sample*, and a response set is called *train set*. These terms can be used interchangeably, but a train set is demanded to contain labeled samples. A *ranker* is a tuple ${R}= ({\mathcal{D}},{\mathcal{C}})$, which is employed to compute a rank ${\tau}$ for $q$, denoted by ${{\tau}_{q}}$ to distinguish its corresponding query. A rank is a permutation of $S_L \subseteq S$, where $L \ll n$ in general, such that ${{\tau}_{q}}$ provides the most similar – or equivalently the least dissimilar – samples to $q$ from $S$, in order. $L$ is a cut-off parameter. A ranker establishes a ranking system, but different descriptors and comparators can compose rankers. Besides, descriptors are commonly complementary, as well as comparators. For $m$ rankers, $\{R_j\}_{j=1}^m$, used for query retrieval over a collection $S$, for every query $q$ we can obtain ${\mathcal{T}_{q}}= \{{{\tau}_{j}}\}_{j=1}^m$, from which a *rank aggregation* function $f$ produces a combined rank ${{\tau}_{q,f}}=f({\mathcal{T}_{q}})$, presumably more effective than the individual ranks ${{\tau}_{1}}$, ${{\tau}_{2}}$, $\dots$, ${{\tau}_{m}}$. Representation and Prediction Based on Graph-Based Rank Fusion {#secMethod} ============================================================== Figure \[fig:fusionVectorEstimation\] presents an overview of our method – a multimodal representation and estimator based on rank-fusion graphs. The solution is composed of three main generic components, briefly described here and detailed in the following sections. The multimodal representation is completely unsupervised, thus able to be adopted in any tasks in the absence of labeled data. Two phases are defined. The [*training*]{} phase comprehends the modeling of the train set in terms of multiple rankers, as fusion graphs and then as fusion vectors. This step performs a graph embedding of fusion graphs and the training of a multimodal prediction model. The [*inference*]{} phase refers to the multimodal prediction preceded by a rank-based fusion approach for multimodal representation. The training phase is performed only once, while the inference phase is performed per prediction. The first two components – [*fusion graph extraction*]{} and [*graph embedding*]{} – are used in both phases. ![Proposed graph-based rank fusion for multimodal prediction.[]{data-label="fig:fusionVectorEstimation"}](fusionVectorEstimation){width="0.95\linewidth"} The [*fusion graph extraction*]{} (component 1 in the figure) generates a fusion graph ${\mathcal{G}}$ for a given query sample $q$. ${\mathcal{G}}$ consists of an aggregated representation of multiple ranks for $q$, thus capturing and correlating information of multiple ranks. This formulation is presented in Section \[sec:fusionGraphExtraction\]. [*Graph Embedding*]{} (2) projects fusion graphs into a vector space model, producing a corresponding fusion vector ${\mathcal{V}}$ for ${\mathcal{G}}$. We propose an embedding formulation in Section \[sec:fusionGraphEmbedding\]. At the end, an [*event predictor*]{} (3) is built based on the response fusion vectors, in order to predict for queries (also modeled as fusion vectors). This component is detailed in Section \[sec:fusionVectorSearch\]. Rank-Fusion Graphs {#sec:fusionGraphExtraction} ------------------ This component produces a fusion graph ${\mathcal{G}}$ for a given query sample $q$, also in terms of $m$ rankers and $n$ response items. A fusion graph is a graph-based encoding of multiple ranks for $q$, that encapsulates and correlates ranks. We follow the fusion graph formulation from [@Dourado:2019:FG], referred to as *FG*, that defines a mapping function $q \mapsto {\mathcal{G}}$, based on its ranks ${\tau}\in {\mathcal{T}_{q}}$ and ranks’ inter-relationships. The proposed formulation also includes a dissimilarity function for *FG*, and a retrieval model based on fusion graphs. Here, however, we focus on the definition of *FG*, extending its use as part of a rank-based late-fusion approach for representation model in prediction tasks, without these components. The process is illustrated in Figure \[fig:graphFusionFromSample\]. Given a query $q$, $m$ rankers, and a training set of size $n$, $m$ ranks are generated. These ranks are then normalized to allow for producing the fusion graph ${\mathcal{G}}$ for $q$. In rank normalization, the scores in the ranks generated by dissimilarity-based comparators are converted to similarity-based scores. Besides, all ranks have their scores rescaled to the same interval. ${\mathcal{G}}$, for $q$, includes all response items from each ${{\tau}_{q}} \in {\mathcal{T}_{q}}$, as vertices. Vertices are connected by taking into account the degree of relationship between their corresponding response items, and the degree of their relationships to $q$. ![Extraction of a fusion graph (adapted from [@Dourado:2019:FG]).[]{data-label="fig:graphFusionFromSample"}](graphFusionFromSample2){width=".8\linewidth"} Embedding of Rank-Fusion Graphs {#sec:fusionGraphEmbedding} ------------------------------- Let ${\mathbb{G}}=\{{\mathcal{G}}_i\}_{i=1}^n$ be the fusion graph set related to the response set of a given collection. Based on ${\mathbb{G}}$, an embedding function ${\mathcal{E}}$ defines a vector space model in order to project a fusion graph ${\mathcal{G}}$ into that space as a fusion vector ${\mathcal{V}}$, i.e. ${\mathcal{V}}={\mathcal{E}}({\mathcal{G}})$ for any ${\mathcal{G}}$. We investigate how ${\mathcal{V}}$ can be adopted as a representation model of multi-ranked objects. It encodes the use of multiple rankers and allows the fusion of multiple modalities, being therefore suitable for prediction tasks. ${\mathcal{E}}$ can be defined by unsupervised or supervised approaches. We explore three approaches in this paper, preliminarly presented in [@Dourado:2019:FV] in the context of retrieval tasks. Different from that paper, here we focus its use on prediction tasks involving supervised learning. This first one, ${\mathcal{E}}_V$, derives the vector space based on vertex analysis. Let $w(v)$ be the weight of the vertex $v$, if $v \in {\mathcal{G}}$, otherwise 0. Similarly, let $w(e)$ be the weight of the edge $e$, if $e \in {\mathcal{G}}$, otherwise 0. Also, let $N$ be the dimensionality of the vector space model defined by ${\mathcal{E}}$, such that ${\mathcal{V}}\in {\rm I\!R}^N$. ${\mathcal{E}}_V$ derives ${\mathcal{V}}$ from the vertices of ${\mathcal{G}}$. There is one vector attribute relative to each response object, therefore $N=|{\mathbb{G}}|$. ${\mathcal{V}}$ is derived from ${\mathcal{G}}$ such that $|{\mathcal{V}}|=N$, ${\mathcal{V}}[i]$ is the importance value of the $i$-th attribute, and ${\mathcal{V}}[i]=w(v_i)$. Despite the vector space increases linearly to the collection size, the resulting fusion vectors are mainly sparse, i.e., composed of few non-zero entries, which makes this embedding formulation simple and efficient in practice. ${\mathcal{E}}_H$ is a hybrid embedding approach based on both vertex and edge analysis. ${\mathcal{E}}_H$ encodes more information into the vector space, at a cost of a higher dimensionality. The third approach, ${\mathcal{E}}_K$, extends the archetype to embed graphs as a histogram of kernels, where the vectorial attributes are selected by unsupervised selection of common subgraph patterns. The kernels are obtained from the centroids of a graph clustering process. Then, a vector quantization process, consisting of assignment and pooling procedures, is adopted to embed an input graph to the vector space. We adopt SOFT assignment and AVG pooling, as suggested by the authors [@Dourado:2019:FV]. has been successfully applied in scenarios involving graph classification, textual representation and information retrieval [@Silva:2018:BoG; @Dourado:2019:BoTG; @Dourado:2019:FV]. To the best of our knowledge, this is the first extension of in the context of multimodal representation. We refer to FV-V as the fusion vector generated by ${\mathcal{E}}_V$, while FV-H is generated by ${\mathcal{E}}_H$, and FV-K by ${\mathcal{E}}_K$. Prediction based on Fusion Vectors {#sec:fusionVectorSearch} ---------------------------------- Fusion vectors allow the creation of predictors, such as classifiers or regressors, and also ad-hoc retrieval systems, depending on the underlying demanded task. In this work, we adopt them to build predictors, where training objects – associated with ground-truth information – are used to train an estimator for a certain input object be considered a label (or event) or not. Let $S$ be a training corpus of size $n$. A predictor can be modeled as $f (X, \beta) \approx Y$, where $f$ is an approximation function, $X$ are the independent variables, $Y$ is the dependent variable (target), and $\beta$ are unknown parameters. A learning model explores $S$ to find a $f$ that minimizes a certain error metric. The training samples are generally labeled, so $Y$ may be categorical. Still, a regressor can be built, as $E(Y|X)=f(X, \beta)$, so that posterior probabilities are inferred in order to estimate a confidence of a sample to refer to a class of not. $X$, in our case, refers to the fusion vectors, acting as variables that describes the samples in terms of their multiple multimodal ranks. Experimental Evaluation {#secExpEval} ======================= We present, in this section, the experimental protocol used to evaluate the method, and the results achieved comparatively to state-of-the-art baselines. We evaluate the effectiveness of our method as a representation model in prediction tasks. The focus is on validating our fusion method comparatively to the individual use of descriptors with no fusion, as well as to compare it to early-fusion and late-fusion approaches. Evaluation Scenarios -------------------- We evaluate the proposed method on multiple datasets, comprised of heterogeneous multimodal data, in order to assess its general applicability. Our experimental evaluation comprises the following scenarios: - **ME17-DIRSM** dataset, the acronym for MediaEval 2017 Disaster Image Retrieval from Social Media [@bischke:2017:meSatellite], is a multimodal dataset of a competition whose goal is to infer whether images and/or texts refer to flood events or not. The samples contain images along with textual metadata, such as title, description, and tags, and they are labeled as either flood (1) or non-flood (0). The task predefines a development set (devset) of samples, and a test set of samples, as well as its own evaluation protocol. - **Brodatz** [@brodatz:1966] is a dataset of texture images, labeled across $111$ classes. There are $16$ samples per class, composing a total of samples. - **Soccer** [@van:2006:soccer] is an image dataset, labeled across 7 categories (soccer teams), containing $40$ images each. - **UW** [@Deselaers:2008:UW], also called University of Washington dataset, is a multimodal collection of images annotated by textual keywords. The images are pictures labeled across $22$ classes (locations). Pictures per class vary from 22 to 255. The number of keywords per picture vary from 1 to 22. Evaluation Protocol ------------------- For any dataset that does not explicitly define train and test sets, we initially split it in train and test sets, at a proportion of 80% and 20% respectively, in a stratified way so that the proportions per class remain equal. The same train and test sets per dataset are adopted to evaluate all methods under the same circumstances, as well as the evaluation metrics. For each representation model, we fit a multiclass SVM classifier, with one-vs-all approach and linear kernel, as it is a good fit for general applicability. Hyper-parameters are selected by grid search on the train set, using an internal 5-fold cross validation. We evaluate the effectiveness of each method by the balanced accuracy score, which is suitable to evaluate on either balanced and imbalanced datasets. The methods are compared by their balanced accuracy. The descriptors compose rankers, which are employed to generate ranks in our late-fusion representation. Our method varies with respect to which rankers are used, whether visual or textual rankers, or even their combinations for the multimodal scenario are applied. Besides that, it also varies with respect to which embedding approach is adopted. We evaluate these aspects experimentally. We model our solution as a rank-fusion approach, followed by an estimator based on rank-fusion vectors. This approach intends to validate our hypothesis that unsupervised graph-based rank-fusion approaches can lead to effective representation models for prediction tasks in general. We adopt the same experimental evaluation for all datasets but ME17-DIRSM, that defines its own procedure. In this case, the task imposes three evaluation scenarios, as follows. In the first one, called “visual”, only visual data can be used. In the second scenario, called “textual”, only textual data are used. In the third scenario, called “multimodal”, both visual and textual data are expected to be used. The correctness is evaluated, over the test set, by the metric Average Precision at $K$ (AP@K) at various cutoffs (50, 100, 250, 480), and by their mean value (mAP). Although the ME17-DIRSM task may be seen as a multimodal binary classification problem, the evaluation metrics require ranking-based solutions, or equivalently confidence-level regressors, so that the first positions are the most likely to refer to a flood event. For the estimator component in ME17-DIRSM, we adopt SVR, an L2-regularized logistic regression based on linear SVM in its dual form, with probabilistic output scores, and trained over the fusion vectors from devset. Probabilistic scores are used so that we can sort the test samples by confidence expectancy of being flood. In ME17-DIRSM, our results are compared to those from state-of-the-art baselines. In Soccer, Brodatz, and UW, our results are compared to those from two major fusion approaches: concatenation, and majority vote. They cover baselines from both early-fusion and late-fusion families. For the concatenation procedure, besides the concatenation itself, we normalize the vectors to the $[0,1]$ interval for each attribute, in order to avoid disparities due to different descriptor attribute ranges. We apply majority voting in the scenarios involving an odd number of descriptors, so that each predicted class is taken as the one most frequently predicted by the estimators constructed for each descriptor. Descriptors and Rankers ----------------------- For ME17-DIRSM, we selected three visual descriptors and three textual descriptors, for individual analysis in the designed evaluation scenarios, and to evaluate different possibilities of rank-fusion aggregations. We adopt the following state-of-the-art visual descriptors: - *ResNet50IN*: 2048-dimensional average pooling of the last convolutional layer of ResNet50 [@he:2016:ResNet50], pre-trained on ImageNet [@Russakovsky:2015:ImageNet], a dataset of about 14M images labeled for object recognition; - *VGG16P365*: 512-dimensional average pooling of the last convolutional layer of VGG16 [@simonyan:2014:VGG16], pre-trained on Places365-Standard [@zhou:2017:Places], a dataset of about 10M images of labeled scenes; - *NASNetIN*: 2048-dimensional average pooling of the last convolutional layer of NASNet [@barret:2017:NASNet], pre-trained on ImageNet dataset. Based on the textual metadata in ME17-DIRSM, we adopt the following descriptors: - *BoW*: Bag of Words (BoW) with Term Frequency (TF) weighting; - *2grams*: 2grams with TF weighting; - *doc2vecWiki*: 300-dimensional doc2vec [@le:2014:doc2vec] pre-trained on English Wikipedia dataset, of about 35M documents and dumped at 2015-12-01. For the other datasets, we elected a number of heterogeneous descriptors: - Soccer: , , and . - Brodatz: , , and . - UW: , , , and , as visual descriptors, and *word2vecSum*, *word2vecAvg*, *doc2vecWiki*, and *doc2vecApnews*, as textual descriptors. For the deep networks used for CNN-based visual feature extraction, as well as in the textual feature extraction with doc2vec, we take advantage of pre-trained models. This practice, known as transfer learning, has been effective in many scenarios [@kornblith:2018:imagenetModelTransfer], and it is also particularly beneficial for datasets that are not large enough to generalize the training of such large architectures, as in our case. Because the problem requires prediction of flood images, we prioritize, in the selection of visual descriptors, datasets for pre-training that focus on images of scenes, aiming at better generality to the target problem. We perform the same preprocessing steps for every textual descriptor: lower case conversion, digit and punctuation removal, and English stop word removal. For *BoW* and *2grams*, we also apply Porter stemming. The *word2vecSum* descriptor produces, for any input document, a vector corresponding to the sum of the word embedding vectors [@mikolov:2013:word2vec] related to each term within that document, while *word2vecAvg* computes the mean vector of them. The doc2vec model promotes document-level embeddings for texts, and it is based on word embeddings [@mikolov:2013:word2vec], a preliminary work that assigns vector representations for words in order to capture their semantic relationships. *doc2vecApnews* stands for a 300-dimensional *doc2vec* [@le:2014:doc2vec] model, pre-trained over the Associated Press News textual dataset, of about 25M news articles from 2009 to 2015. From the descriptors, rankers are defined as tuples of (descriptor, comparator), where the comparator corresponds to a dissimilarity function. We compose a ranker for each descriptor by choosing an appropriate comparator. Given that our method works on top of rankers, we have to define dissimilarity functions to be used along with those descriptors that is not explicitly associated with one. This is the case for the four textual descriptors adopted in UW, as well as the descriptors adopted in ME17-DIRSM. All remaining descriptors define their own comparators. For the textual descriptors BoW and 2grams, we adopt the Weighted Jaccard distance, defined as $1 - J(\mathbf{u}, \mathbf{v})$, where $J$ is the Ruzicka similarity metric (Equation \[eq:Jaccard\]). Jaccard is a well-known and widely-used comparison metric for classic textual descriptors, specially for short texts, as in our case. For the remaining descriptors, we choose the Pearson correlation distance, defined as $1 - \rho(\mathbf{u}, \mathbf{v})$ (Equation \[eq:correlation\]), which is a general-purpose metric due to its suitability for highly dimensional data and scale invariance. $$\label{eq:Jaccard} J(\mathbf{u},\mathbf{v}) = \frac{\sum_i \min(u_i,v_i)}{\sum_i \max(u_i,v_i)}$$ $$\label{eq:correlation} \rho(\mathbf{u}, \mathbf{v}) = \frac{(\mathbf{u} - \bar{u}) \cdot (\mathbf{v} - \bar{v})} {{\|(\mathbf{u} - \bar{u})\|}_2 {\|(\mathbf{v} - \bar{v})\|}_2}$$ The datasets, descriptors, and evaluation criteria, are summarized in Table \[tab:datasets\]. Fusion Setups ------------- For both visual and textual scenarios in ME17-DIRSM, we analyze three variants of our method with respect to the input rankers for late-fusion. For the visual scenario, the combinations are ResNet50IN + NASNetIN, ResNet50IN + VGG16P365, and ResNet50IN + NASNetIN + VGG16P365. For the textual scenario, the combinations are BoW + 2grams, BoW + doc2vecWiki, and BoW + 2grams + doc2vecWiki. As for the multimodal scenario, we investigate some combinations taking one ranker of each type, two of each, and three of each. Six multimodal combinations are evaluated: ResNet50IN + BoW, ResNet50IN + NASNetIN + BoW + 2grams, ResNet50IN + NASNetIN + BoW + doc2vecWiki, ResNet50IN + VGG16P365 + BoW + 2grams, ResNet50IN + VGG16P365 + BoW + doc2vecWiki, and ResNet50IN + NASNetIN + VGG16P365 + BoW + 2grams + doc2vecWiki. We report three results for the adoption of *FV*, in its different embedding approaches, as a representation model for prediction tasks, in Soccer, Brodatz, and UW. We report the results for multiple descriptor combinations, in order to analyze: (i) the method against baselines, (ii) the embedding approaches, and (iii) the comparative effectiveness between the descriptor combinations. The descriptor combinations selected, although not exhaustive, are targeted for a large number of scenarios. In Soccer and Brodatz, all possible combinations were selected for evaluation. In UW, several visual combinations and multimodal combinations were selected. The descriptor combinations are: - In Soccer: ACC + BIC, BIC + GCH, ACC + GCH, and ACC + BIC + GCH. - In Brodatz: CCOM + FCTH, CCOM + JCD, FCTH + JCD, and CCOM + FCTH + JCD. - In UW: ACC + CEDD, ACC + JCD, CEDD + JAC, CEDD + JCD, ACC + CEDD + JCD, and CEDD + JAC + JCD, for visual fusion, and ACC + doc2vecApnews, JCD + doc2vecApnews, ACC + JCD + doc2vecApnews, ACC + JCD + doc2vecWiki, ACC + JCD + word2vecAvg, and ACC + JCD+word2vecSum, for multimodal fusion. Results and Discussion ---------------------- ### Base Results Here we report results by the use of individual descriptors. They constitute an initial baseline for our method as well as for other fusion approaches. We report, in Tables \[tab:resultsVisual\] and \[tab:resultsTextual\], the results for the visual and textual scenarios in ME17-DIRSM, achieved by the three visual and textual selected descriptors, along with a SVR regressor. As the task only mentioned AP@480 and mAP in their leaderboard, we focus our discussions on these two metrics. The correctness for the visual scenario is already high within these baselines, around $85\%$ in AP@480. In the textual scenario, AP@480 is around $65\%$, which suggests more room for improvement. We report, in Tables \[tab:descriptorResultsSoccer\], \[tab:descriptorResultsBrodatz\], \[tab:descriptorResultsUWvisual\], and \[tab:descriptorResultsUWtextual\], the results obtained in Soccer, Brodatz, UW (visual), and UW (textual), respectively, by the use of the descriptors along with a SVM classifier. [.49]{} \[tab:resultsVisual\] [.49]{} \[tab:resultsTextual\] [0.49]{} Descriptor Balanced acc. ------------ --------------- BIC 60.84 GCH 56.38 ACC 51.79 : UW, textual[]{data-label="tab:descriptorResultsUWtextual"} Descriptor Balanced acc. ------------ --------------- JCD 78.65 FCTH 75.10 CCOM 74.65 : UW, textual[]{data-label="tab:descriptorResultsUWtextual"} [0.49]{} Descriptor Balanced acc. ------------ --------------- JAC 83.74 ACC 77.06 JCD 73.99 CEDD 72.46 : UW, textual[]{data-label="tab:descriptorResultsUWtextual"} ### Parameter Analysis The resulting size of ${\mathcal{G}}$ is affected by the input rank sizes, defined by the hyper-parameter *L*. For the same reason, larger *FG*’s either increase the vocabulary sizes of *FV* or the complexity to generate them. @Dourado:2019:FG showed that an increase in *L* leads to more discriminate graphs up to a saturation point. A practical upper bound for the choice of *L* tends to be the maximum rank size of users’ interest, indirectly expressed here by the evaluation metrics. As ME17-DIRSM defines evaluation metrics for ranks up to 480, we start by empirically evaluating the influence of *L* in the mAP score, for values to up 480. Figure \[fig:parameterEvaluationL\] reports the influence of L for some of the elected fusion scenarios. The results were as expected: the effectiveness usually increased as *L* was larger. For the next evaluation scenarios in ME17-DIRSM, we adopt $L=480$. For the other datasets, we adopt $L=10$. ![Effect of the rank size limit (*L*) for the fusion graph extraction, in the mAP score, for different fusion scenarios in ME17-DIRSM.[]{data-label="fig:parameterEvaluationL"}](parameterAnalysisL){width="\linewidth"} ### Fusion Results in Event Detection We present our results achieved for the three scenarios in ME17-DIRSM, using the combinations proposed, along with the results of the $11$ teams that participated in the competition. We also show the results achieved in [@werneck:2018:BocgBokg] in the visual and multimodal scenarios, which relied on early-fusion techniques. These results are presented in Tables \[tab:resultsRunVisual\], \[tab:resultsRunTextual\], and \[tab:resultsRunMultimodal\], respectively for the visual, textual, and multimodal scenarios. In ME17-DIRSM, we focused on the ${\mathcal{E}}_V$ embedding approach. For the other datasets, we evaluate all of them. In the visual scenario, only [@dirsm17:SAhmadEtAl; @dirsm17:BBischkeEtAl] performed better, in terms of AP@480 and mAP, than our preliminary base setup, based on individual descriptors along with the SVR regressor. As for the textual scenario, only [@dirsm17:NTkachenkoEtAl] in 12 initiatives surpassed BoW + SVR in AP@480, and [@dirsm17:BBischkeEtAl; @dirsm17:KNogueiraEtAl; @dirsm17:ZZhao] in mAP. This indicates that descriptors properly selected to the target problem can overcome more complex models, also requiring less effort. Our method was superior in the visual scenario to all baselines, for two of three proposed variants of ranker combinations. Compared to the strongest baselines considering this scenario, our method presents gains from around 1 to 2% in AP@480, and 1% in mAP. Compared to the visual base results, from the individual descriptors, 3 to 4% in AP@480, and 1 to 2% in mAP. In the textual scenario, our gains were even more expressive. It was superior in the textual scenario to all related works, for all three proposed variants of ranker combinations. Compared to the strongest baselines, our method presents gains from 5 to 7% in AP@480, and 6 to 11% in mAP. Compared to the textual base results, from the individual descriptors, 6 to 8% in AP@480, and 14 to 16% in mAP. In the multimodal scenario, considered baselines were more competitive. Again, however, our method presents gains over them, of 0.5% in AP@480 and 0.23% in mAP. ### Fusion Results in Classification Tasks We report in Tables \[tab:fusionResultsSoccer\], \[tab:fusionResultsBrodatz\], \[tab:fusionResultsUWvisual\], and \[tab:fusionResultsUWmultimodal\], the results obtained by our method variants, respectively in Soccer, Brodatz, UW for image data, and UW for multimodal data, besides the results obtained by the baselines. Our method led to significant gains when compared to the best base result from the descriptors in each dataset: around 8 p.p. in Soccer, 13.2 p.p. in Brodatz, 2.2 p.p. in UW for visual fusion, and 5.9 p.p. in UW for multimodal fusion. The gains in UW were comparatively lower than others, yet consistent, because the base results were already higher, so that there were less room for improvement. Our method, when compared to the best baseline in each descriptor combination in each dataset, had gains in all cases: up to 12.4 p.p. in Soccer, up to 2.9 p.p. in Brodatz, up to 7.1 p.p. in UW for visual fusion, and up to 3.9 p.p. in UW for multimodal fusion. Overall, all the FV approaches performed better than all baselines in all datasets. In only 6 of 20 descriptor combinations evaluated, any of the baselines surpassed any of the FV approaches. The accuracy disparities, obtained by FV or any other fusion method, across the multiple descriptor combinations in each dataset, show that there is no obvious choice when dealing with which descriptors to be used together. As our representation model is meant to be unsupervised, this choice could only be guided by general heuristics, such as a selection of effective and low correlated descriptors, as discussed in [@Dourado:2019:FG]. We leave this exploration for future work. FV-K usually performed better than FV-H and FV-V, and FV-H usually performed better than FV-V, although in both cases the gains were at most 3 p.p one over the other. On the opposite side, FV-V is the simplest among the three, and FV-K requires more computational steps. These two aspects combined impose that the practical choice among the three must take into account the trade-off between accuracy vs computational cost. In any case, our method is unsupervised and feasible for general applicability. Conclusions {#secConc} =========== In this paper, we presented an unsupervised graph-based rank-fusion approach as a representation model for multimodal prediction tasks. Our solution is based on encoding multiple ranks into a graph representation, which is later embedded into a vectorial representation. Next, an estimator is built to predict if an input multimodal object refers to a target event or not, given their graph-based fusion representations. The proposed method extends the fusion graphs – first introduced in [@Dourado:2019:FG] – for supervised learning tasks. It also applies a graph embedding mechanism in order to obtain the fusions vectors, a late-fusion vector representation that encodes multiple ranks and their inter-relationships automatically. Performed experiments in multiple prediction tasks, such as flood detection and multimodal classification, demonstrate that our solution leads to highly effective results overcoming state-of-the-art solutions from both early-fusion and late-fusion underlying approaches. Future work will focus on investigating the impact of semi-supervised and supervised approaches for the fusion graph and fusion vector constructions. We also plan to investigate the use of our solution in other multimodal problems, such as recommendation and hierarchical clustering. Finally, we plan to evaluate the proposed approach for other multimedia data, such as audio and video. References {#references .unnumbered} ==========
{ "pile_set_name": "ArXiv" }
--- bibliography: - 'main.bib' - 'LHCb-PAPER.bib' - 'LHCb-CONF.bib' - 'LHCb-DP.bib' --- =1
{ "pile_set_name": "ArXiv" }
--- abstract: 'Fermi acceleration has not been considered to be viable to explain non-thermal electrons (20$\sim$100 KeV) produced in solar flares, because of its high injection energy. Here, we propose that non-thermal electrons are efficiently accelerated by first-order Fermi process at the fast shock, as a natural consequence of the new magnetohydrodynamic picture of the flaring region revealed with [*Yohkoh*]{}. An oblique fast shock is naturally formed below the reconnection site, and boosts the acceleration to significantly decrease the injection energy. The slow shocks attached to the reconnection X-point heat the plasma up to 10$\sim$20 MK, exceeding the injection energy. The combination of the oblique shock configuration and the pre-heating by the slow shock allows bulk electron acceleration from the thermal pool. The accelerated electrons are trapped between the two slow shocks due to the magnetic mirror downstream of the fast shock, thus explaining the impulsive loop-top hard X-ray source discovered with [*Yohkoh*]{}. Acceleration time scale is $\sim$ 0.3–0.6 s, which is consistent with the time scale of impulsive bursts. When these electrons stream away from the region enclosed by the fast shock and the slow shocks, they are released toward the footpoints and may form the simultaneous double-source hard X-ray structure at the footpoints of the reconnected field lines.' author: - Saku Tsuneta - Tsuguya Naito title: 'FERMI ACCELERATION AT FAST SHOCK IN A SOLAR FLARE AND IMPULSIVE LOOP-TOP HARD X-RAY SOURCE' --- INTRODUCTION ============ Hard X-ray spectral observations of solar flares made in past 20 years have established that efficient electron acceleration ($\sim10^{34-35}$ electrons s$^{-1}$) occurs during impulsive solar flares (Lin & Hudson 1971). The [*Yohkoh*]{} Hard X-ray Telescope clearly showed that almost all flares have double footpoint sources, the signature of bombardment by the accelerated electrons into the dense chromosphere (Hoyng et al. 1981, Sakao et al. 1997a, 1997b). These spectral and imaging observations showed that rich amount of non-thermal electrons were accelerated somewhere in the system, and were injected in the soft X-ray loop, which turned out to be the reconnected field lines filled with evaporated plasma (Tsuneta 1996). Furthermore, a surprising discovery from the HXT is the detection of an impulsive hard X-ray source located above the soft X-ray loop (Masuda et al. 1994). The overall time profile of the loop-top source is similar to those from the footpoints, implying common origin of acceleration. See Miller et al. (1997) for a review of these recent observations and the theoretical implications. [*Yohkoh*]{} soft X-ray observations, on the other hand, strongly suggest that magnetic reconnection serves as an efficient engine to convert magnetic energy to plasma kinetic and thermal energies (Tsuneta 1996). In this process, MHD slow-mode shocks attached to the reconnection X-point appear to convert magnetic energy, brought with the inflow to the X-point from the vast active region corona, to plasma heating and an outflow jet with the Alfvén speed on the upstream side. The outflow jet can form a fast-mode shock (quasi-perpendicular shock) to further heat the plasma (Masuda et al. 1994, Shibata et al. 1995, Tsuneta et al. 1997). The observed geometry indicated that the loop-top hard X-ray source was located at the downstream side of reconnection site (below the X-point). This implied that the formation of the loop-top source was related to the fast shock, regardless of whether it was of thermal or non-thermal origin (Tsuneta et al. 1997). Furthermore, the electron spectra of the loop-top hard X-ray source turned out to be more consistent with the power-law with differential power index of 2–4 (Alexander & Metcalf 1997). Aschwanden et al. (1996) concluded from their innovative time-of-flight method that the acceleration site of this flare must have been located around the loop-top hard X-ray source or above, if the acceleration site was spatially concentrated. All these three independent approaches point to the loop-top hard X-ray source as the acceleration site. In this Letter, we suggest that the electron acceleration and the resultant formation of the loop-top hard X-ray source are natural consequences of the new MHD framework being established by the [*Yohkoh*]{} soft X-ray observations: The highly oblique fast shock sandwiched by the slow shocks provides an ideal site for electron acceleration with the first-order shock Fermi acceleration (eg. Blandford & Eichler 1987; Jones & Ellison, 1991). SHOCK FERMI ACCELERATION ======================== Requirements from Observations ------------------------------ There are four stringent requirements from the observations to quantitatively examine whether first-order Fermi acceleration is viable in the 10–100 keV electron acceleration. \(1) Since the background plasma density is high in the solar corona, the acceleration rate has to exceed the collisional loss at the thermal energy of the background plasma. The net acceleration rate is given by $$\frac{dE}{dt} = \frac{dE}{dt}_{\rm a} - \left|\frac{dE}{dt}_{\rm c}\right|,$$ where the first term on the right hand side is the acceleration rate and the second term the collisional energy loss. The net acceleration rate has to be positive at the thermal energy of the background plasma. (2) The energy gain has to be high enough to explain the loop-top hard X-ray source, that is observed at $\sim$ 50 keV. (If the loop-top source provides electrons responsible for the footpoint hard X-rays, the energy has to reach $>$100 keV.) Electrons, therefore, have to be accelerated within a relatively short time, over which an oblique (quasi-perpendicular) field line crosses the fast shock structure. (3) Acceleration takes place on a time scale of impulsive bursts, which is about 1s or faster. (4) Number of accelerated electrons is consistent with number of accelerated electrons deduced from observations (10$^{34-35}$ electrons s$^{-1}$.) Injection Energy and Energy Gain -------------------------------- In this paper, we call the energy, at which the net energy gain rate \[Eq. (1)\] equals zero, the injection energy. Although the injection energy is defined with regard both to energy loss and particle scattering, we assume that there exists sufficient magnetic field turbulence such as whistler waves to be able to scatter the electrons with energies exceeding 1 keV (eg. Melrose 1986, Melrose 1994). \[These waves can energize electrons (eg. Miller et al. 1997). We, however, do not include these alternative acceleration processes such as resonance acceleration and DC electric acceleration in this analysis.\] The collisional energy loss in Eq.(1) is estimated to be $$\left|\frac{dE}{dt}_{\rm c}\right| \sim 47 \frac{n_{\rm 10}}{\surd E},$$ where $n_{\rm 10}$ is the background electron density in unit of 10$^{10}$ cm$^{-3}$, and $E$ the electron energy in keV (eg. Jackson 1975). In the scheme of shock acceleration, the acceleration rate is derived from energy gain $\Delta E$ and required time $\Delta t$ per each shock crossing, and represented by $$\frac{dE}{dt}_{\rm a} = \frac{\Delta E}{\Delta t} = \frac{2}{3} E \frac{u}{l \cos\theta},$$ where $l$ is the (energy-independent) diffusion length measured along the field line, $u$ the upstream speed with respect to the downstream speed, and $\theta$ the angle between the fast shock normal and the field line crossing the shock(Fig. 1). Here, we apply $\Delta E = 8uE/3v\cos\theta$, where $v$ is an electron velocity ($E=mv^2/2$), and $1/\cos\theta$ is the enhancement factor originating from the obliqueness of the shock. Assuming an isotropic electron distribution, we apply $\Delta t = l/(v/4)$. In the standard shock acceleration theory, $l = \kappa_{\rm 1}/u_{\rm 1}+\kappa_{\rm 2}/u_{\rm 2}$, where $\kappa_{\rm 1 (2)}$ and $u_{\rm 1 (2)}$ are the diffusion coefficient and flow speed in the upstream (downstream), respectively. The time during which a particular field line crosses the fast shock is $L/2u\tan\theta$. The crossing time decreases with increasing obliqueness. From Eq. (3), the energy gain is given by $$\frac{E}{E_{\rm 0}} = \exp ( \frac{L}{3 l \sin\theta}),$$ for particles with energy above the injection energy. We simply represent the diffusion length as a parameter $l$. We adopt $\theta$ as another parameter expressing obliqueness of the shock. It is proposed that the efficiency of shock acceleration is boosted due to the obliqueness (Jokipii 1987, Naito & Takahara 1995a, 1995b and references therein). Fig. 2 shows that the injection energy rapidly decreases with increasing obliqueness. Here, we adopt the following values from the observations; the density of the background plasma $n = 10^{10}$ cm$^{-3}$, the temperature $T = 2 \times 10^{7}$ K (Tsuneta et al. 1997), the length of the fast shock $L = 7000$ km, which is the width of the loop-top hard X-ray source (Masuda et al. 1984). We assume the diffusion length $l = 500 {\rm km}$, which will be estimated later to satisfy the observational and theoretical requirements. The speed of the outflow from the reconnection region is equal to the Alfvén velocity of the upstream side (Petschek 1964), and is taken to be $u = 1000$ km s$^{-1}$ (Tsuneta 1996). Diffusion Length and Shock Angle -------------------------------- The number of accelerated electrons critically depends on the injection energy with respect to the thermal energy of the background plasma. Here, we conservatively assume that the injection energy is equal to the thermal energy of the background plasma heated by the slow shocks ($\sim$ 2 keV), implying bulk electron acceleration, though higher injection energy resulting in acceleration of tail electrons in the thermal distribution function may be acceptable. The initial energy of electrons $E_{\rm 0}$ is, therefore, 2 keV, which is accepted as the injection energy for $\theta = 85$ deg as shown in Fig. 2. Combining with the electron energy at the loop-top hard X-ray source of about 100 keV, we assume that the energy gain of 50 or more is necessary. Since the diffusion (scattering) length $l$ and the shock angle $\theta$ are parameters that cannot be directly determined from the observations, we obtain the two parameters from the following two conditions to satisfy the first two requirements in section 2.1; $$\frac{dE}{dt} > 0 \;\;\;{\rm at}\;\;\; E_{\rm 0} = 2 {\rm keV},$$ $$\frac{E}{E_{\rm 0}} > 50.$$ Fig. 3 shows $dE/dt=0$ at $E_{\rm 0}=2$ and $1$KeV, and $E/E_{\rm 0}=25,50$ and $100$ in the $\theta - l$ space. The following two additional conditions have to be satisfied to choose the proper diffusion length $l$: (1) The length must be comparable to or be larger than the diffusion length derived from the assumption of the Bohm diffusion (eg. Kirk 1994). It is customarily expressed by using Bohm diffusion coefficient $ \kappa_{\rm B} $ as $l \sim \eta \kappa_{\rm B}/u \sim 2 \eta E/(3 m \omega_{\rm ce} u) \sim 0.06 \eta $ km for $E=100$keV and $B=10$G where the factor $\eta $ is anticipated to be in the range of $\eta \sim 1-100 $ from recent observations in interplanetary shocks (Kang & Jones 1997; Baring et al. 1997). Since scattering by whistler waves in transient phenomena such as solar flares is poorly understood, we conservatively assume $\eta \sim 10^{4} $, and the diffusion length $l = 600 {\rm km}$. (2) The length must be smaller than the distance between the slow shocks ($\sim L$), which contain the electrons, thus $l \leq L/2\sin\theta \sim 3500/\sin\theta$ km. These conditions given above are met when the shock is highly oblique ($\theta \sim 80-85$ degree). The diffusion length in this case is about $l \sim 600$ km. The diffusion length $l$ needs to be shorter for higher energy gain. For instance, $l\sim 500$ km and $\theta\sim 85$ degree are needed to accelerated electrons to 1 MeV from 2 keV. Since the required diffusion length logarithmically depends on the energy gain $E/E_{\rm 0}$, we need the diffusion length of 270 km to accelerate electrons to 10 MeV from 2 keV, 230 km to 50 MeV from 2 keV for the same shock angle $\theta\sim 85$ degree. The acceleration to ultra-high energy may be feasible, if we assume an energy-independent diffusion length. The entire acceleration time for $\theta\sim 80-85$ degree is $L/2u\tan\theta \sim$ 0.6-0.3 s, which is consistent with the time scale of the intensity fluctuation of the impulsive hard X-ray bursts, satisfying the third requirement in section 2.1. The forth requirement will be discussed in section 3.3. DISCUSSION AND SUMMARY ====================== Fermi acceleration has often been proposed to explain electron acceleration in solar flares (eg. Bai 1983, Ellison & Ramaty 1985, Somov & Kosugi 1997; first-order Fermi acceleration, Ramaty 1979, LaRosa, Moore, & Shore 1994; second-order Fermi acceleration). However, one of the principal problems associated with the first and second-order Fermi acceleration is the high injection energy of about 20–100 keV (Bai 1983, Ramaty 1979). The oblique shock configuration can drastically decrease the injection energy. Nevertheless, the injection energy is still higher than the energy corresponding to the coronal temperatures (1–3MK), and it may not be possible to accelerate electrons from the coronal plasma, unless we assume a very small diffusion length. \[There is no injection problem with some versions of the stochastic acceleration (LaRosa, Moore, & Shore, 1994, and Miller, LaRosa, & Moore, 1996).\] New Observations from [***Yohkoh***]{} -------------------------------------- Here, we point out that recent [*Yohkoh*]{} observations open up the new possibility to accelerate electrons with first-order Fermi acceleration: (1) Reconnection may produce a fast outflow with speed exceeding the local sound speed. (The sound speed is much faster than the Alfvén speed in the outflow.) Thus, it can produce a fast shock, providing the site for first-order Fermi acceleration. (2) Magnetic field-line structure of the reconnection site naturally creates a highly oblique shock. (The magnetic field lines are almost parallel to the fast shock.) (3) The outflow from the reconnection site is heated upto 10–20MK by the slow shocks (Tsuneta 1996). Therefore, the plasma temperature can reach the injection energy, and the bulk thermal plasma can be accelerated. (4) The acceleration site is bounded by the two facing slow shocks, so that electrons that escape in the upstream side during the course of acceleration most probably return to the fast shock due to the magnetic mirror by the slow shocks, subject to further acceleration. The electrons escaping in the downstream side also returns the shock front, and continue to be accelerated (Fig. 1). Thus, the fast shock bounded by the two slow shocks may provide an escape-free accelerator. Indeed, we have shown that when the shock is oblique enough ($\theta\sim80-85$ degree), the electrons can be accelerated with energy gain $E/E_{\rm 0} > 50$ in 0.3–0.6 s. The injection energy can be low enough so that bulk acceleration from the thermal pool appears to be feasible. The diffusion length $l$ is safely as large as 600 km. Although we have assumed an energy-independent diffusion, the conclusion here is essentially the same as the energy-dependent case. If the shock angle deviates to below about 80 degrees, then the injection energy exceeds the thermal energy of the background plasma, resulting in the significant reduction of the number of accelerated electrons (unless we assume smaller diffusion length). The shock angle is not likely to be uniform due to non-uniform high-$\beta$ outflow in the upstream side of the fast shock (Tsuneta 1996). The variety of flares from purely thermal flares to flares with highly non-thermal nature may be related to this. Location of Hard X-ray Emitting Region -------------------------------------- Since there is no plasma density enhancement in the loop-top hard X-ray source, the loop-top hard X-rays are essentially emitted from the thick target process (Brown 1972). The column density $N$ \[cm$^{-2}$\] to stop electrons with energy $E$ \[keV\] is given by $ N = 8.3 \times 10^{17} E^2$. The column density required to stop $E$ = 50 keV electrons is $N \sim 2 \times 10^{21}$ cm$^{-2}$, and the electrons have to bounce 100–200 times in the loop-top hard X-ray source (Tsuneta et al. 1997, Wheatland & Melrose 1995). It takes about 5–10s to completely thermalize the accelerated electrons. This dissipation time scale is much longer than the acceleration time scale at the fast shock (0.3–0.6 s). The loss cone angle for the slow shocks is about 14 degrees for $B_{\rm u} = 52$G (field strength of the upstream of the slow shock) and $B_{\rm d} = 3$ G (downstream of the slow shock) (Tsuneta et al. 1997). Thus, the slow shocks can provide an efficient confinement mechanism. These considerations indicate that the fast shock is located above the loop-top hard X-ray source, and that the slow shock extends downward below the fast shock to confine the non-thermal electrons. The accelerated electrons are carried downward with slow outflow downstream of the fast shock, and emit hard X-rays as long as they are confined with the slow shock. They are no longer trapped outside the slow shock region. The electrons are then immediately released toward the footpoints, and may form the simultaneous double-source hard X-ray structure at the footpoints of the reconnected field lines. Number of Accelerated Electrons ------------------------------- Masuda (1994) reported that $2\times10^{35}$ electrons s$^{-1}$ are accelerated during the peak phase of the 1992 January 13 flare (Masuda 1994). The reconnection outflow to the fast shock supplies thermal electrons with rate $nuL^2 \sim 5\times10^{35}$ electrons s$^{-1}$, so that the reconnection outflow can provide enough seed electrons to be accelerated. Although acceleration efficiency is expected to be very high as discussed above, we need more detailed modeling with Monte Carlo simulation to obtain the spectra of the accelerated electrons, and to answer whether the forth requirement described in section 2.1 is met. This is left to our future work. The authors thank H. Hudson for comments on the paper. One of the authors (TN) is indebted to the Research Fellowships of the Japan Society for the Promotion of Science for Young Scientists. Alexander, D., & Metcalf, T. R. 1997, , in press Aschwanden, M. J., Kosugi, T., Hudson, H. S., & Wills, M. J. 1996, , 470, 1198 Bai, T. 1983, , 267, 433 Baring M. G., Ogilvie, K. W., Ellison, D. C., & Forsyth, R. J. 1997, , 476, 889 Blandford, R.D., & Eichler, D. 1987, [*Phys. Rep.*]{}, 154, 1 Brown, J. C. 1971, , 18, 489 Ellison, D. C., & Ramaty, R. 1985, , 298, 400 Hoyng, P., Duijveman, A., Machado, M. E., Rust, D. M., Svestka, Z., Boelee, A., de Jager, C., et al. 1981, , 246, L155 Jackson, J. D. 1975, in Classical Electrodynamics (John Wiley & Sons: New York), chap. 13 Jokipii, J. R. 1987, , 313, 842 Jones, F. C. & Ellison, D. C. 1991, [*Space Sci. Rev.*]{}, 58, 259 Kang, H. & Jones, T. W 1997, , 476, 875 Kirk, J. G. 1994, in Plasma Astrophysics, Saas-Fee Advanced Course 24, eds: Benz, A. O. & Courvoisier, T.J.-L (Springer-Verlag: Berlin) p225 LaRosa T. N., Moore R. L., & Shore S. N. 1994, , 425, 856 Lin R. P., & Hudson H. S. 1971, , 17, 412 Masuda, S. 1994, Ph. D. thesis, Univ. Tokyo Masuda, S., Kosugi, T., Hara, H., Tsuneta, S., & Ogawara, Y. 1994, , 371, 495 Melrose, D. B. 1986 in Instabilities in Space and Laboratory Plasmas (Cambridge University Press: Cambridge), chap. 13.6 Melrose, D. B. 1994, in Plasma Astrophysics, Saas-Fee Advanced Course 24, eds: Benz, A. O. & Courvoisier, T.J.-L (Springer-Verlag: Berlin) p113 Miller J. A., LaRosa T. N., & Moore R. L. 1996, ApJ, 461, 445 Miller, J. A. et al. 1997, , 102, A7, 14631 Naito, T. & Takahara, F. 1995, [*Prog. Theor. Phys.*]{}, 93, 287 Naito, T. & Takahara, F. 1995, , 275, 1077 Petschek, H. E. 1964, AAS – NASA Symp. on Solar Flares, ed. W. N. Hess (NASA SP-50), 425 Ramaty, R., 1979, Particle Acceleration Mechanism in Astrophysics, eds. J. Arons, C. McKee, & C. Max, (AIP: New York) p135 Sakao, T., Kosugi, K., Masuda, S., Makishima, K., Inda-Koide, M., & Murakami, T. 1997a, , in press Sakao, T., Kosugi, K., & Makishima, K. 1997b, , in press Shibata, K, Masuda, S., Shimojo, M., Hara, H., Yokoyama, T, Tsuneta, S., Kosugi, T., & Ogawara, Y. 1995, , 451, L83 Somov, B. V. & Kosugi, T. 1997, , 485, in press Tsuneta, S. 1996, , 456, 840 Tsuneta, S., Masuda, S., Kosugi, T., & Sato, J. 1997, , 478,787 Wheatland, M. S., & Melrose, D. B. 1995, , 158, 283
{ "pile_set_name": "ArXiv" }
--- abstract: | Recent inapproximability results of Sly (2010), together with an approximation algorithm presented by Weitz (2006) establish a beautiful picture for the computational complexity of approximating the partition function of the hard-core model. Let $\lambda_c({\mathbb{T}}_\Delta)$ denote the critical activity for the hard-model on the infinite $\Delta$-regular tree. Weitz presented an ${\mathsf{FPTAS}}$ for the partition function when $\lambda<\lambda_c({\mathbb{T}}_\Delta)$ for graphs with constant maximum degree $\Delta$. In contrast, Sly showed that for all $\Delta\geq 3$, there exists $\eps_\Delta>0$ such that (unless $RP=NP$) there is no ${\mathsf{FPRAS}}$ for approximating the partition function on graphs of maximum degree $\Delta$ for activities $\lambda$ satisfying $\lambda_c({\mathbb{T}}_\Delta)<\lambda<\lambda_c({\mathbb{T}}_\Delta)+\eps_\Delta$. We prove that a similar phenomenon holds for the antiferromagnetic Ising model. Sinclair et al. recently extended Weitz’s approach to the antiferromagnetic Ising model, yielding an ${\mathsf{FPTAS}}$ for the partition function for all graphs of constant maximum degree $\Delta$ when the parameters of the model lie in the uniqueness region of the infinite $\Delta$-regular tree. We prove the complementary result that for the antiferrogmanetic Ising model without external field that, unless $RP=NP$, for all $\Delta\geq 3$, there is no ${\mathsf{FPRAS}}$ for approximating the partition function on graphs of maximum degree $\Delta$ when the inverse temperature lies in the non-uniqueness region of the infinite tree ${\mathbb{T}}_\Delta$. Our proof works by relating certain second moment calculations for random $\Delta$-regular bipartite graphs to the tree recursions used to establish the critical points on the infinite tree.\ [**Keywords:**]{} Antiferromagnetic Ising model, Hard-core model, Phase transition, Computational Hardness, Approximate Counting\ [**AMS subject classifications:**]{} 68Q87, 82B20, 60J10, 05C80. author: - 'Andreas Galanis[^1]' - 'Daniel Štefankovič[^2]' - 'Eric Vigoda$^\star$' title: - 'Inapproximability of the Partition Function for the Antiferromagnetic Ising and Hard-Core Models' - 'Inapproximability of the Partition Function for the Antiferromagnetic Ising and Hard-Core Models' --- Introduction ============ A remarkable computational transition has recently been established for the complexity of approximating the partition function of the hard-core model. Amazingly, this computational transition coincides exactly with the statistical physics phase transition on infinite $\Delta$-regular trees. In this work we prove that a similar phenomenon holds for a much more general class of 2-spin models. Our general setup is 2-spin systems on an input graph $G=(V,E)$ with maximum degree $\Delta$. We follow the setup of several related previous works [@GJP; @SST; @LLY]. Configurations of the system are assignments $\sigma: V\rightarrow\{-1,+1\}$. For a given configuration $\sigma\in \{-1,+1\}^V$, denote by $n_{-}(\sigma)$ the number of vertices assigned $-1$, $m^{-}(\sigma)$ the number of edges with both endpoints assigned $-1$ and $m^{+}(\sigma)$ the number of edges with both endpoints assigned $+1$. There are three non-negative parameters $B_1,B_2$ and $\lambda$ where $B_1$ is the edge activity for $(-,-)$ edges, $B_2$ is the edge activity for $(+,+)$ edges, and $\lambda$ is the vertex activity (or external field). A configuration $\sigma\in \{-1,+1\}^V $ has weight $$w_G(\sigma)=\lambda^{n_{-}(\sigma)} {B}_1^{m^{-}(\sigma)}{B}_2^{m^{+}(\sigma)}.$$ The Gibbs distribution $\mu_G(\cdot)$ is over the set $\{-1,+1\}^V$ where $\mu_G(\sigma)=w_G(\sigma)/Z$, where $Z=Z_G(B_1,B_2,\lambda)$ is a normalizing factor, known as the partition function, defined as: $$Z := \sum_{\sigma\in \{-1,+1\}^V} w_G(\sigma)=\sum_{\sigma\in \{-1,+1\}^V}\lambda^{n_{-} (\sigma)} {B}_1^{m^{-}(\sigma)}{B}_2^{m^{+}(\sigma)}.$$ Let us point out a few examples of 2-spin models that are of particular interest. Setting $B_1=0$ and $B_2=1$, the only configurations with positive weight in the Gibbs distribution induce independent sets of $G$. This case is known as the [*hard-core model*]{} with fugacity $\lambda$. Setting $B_1=B_2=B$, one recovers the Ising model. The case $\lambda=1$ is the Ising model without external field. When $B<1$ it is the antiferromagnetic Ising model, and when $B\geq1$ it is the ferromagnetic Ising model. In general, when $B_1B_2<1$ the model is called [*antiferromagnetic*]{}, and when $B_1B_2\geq1$ it is ferromagnetic. The Ising and hard-core models (and, more generally, spin models) have been studied thoroughly in various contexts, especially in statistical physics. Their interpretation as idealized models of microscopic interaction within a body have nurtured research on understanding their macroscopic behavior on appropriate underlying graph structures, such as the infinite grid or the infinite triangular lattice. The intriguing questions arising have made an impact on other fields as well, most notably probability and computer science. For illuminating accounts of such directions, we refer the reader to [@Mezard] and [@Talagrand]. In this paper, we focus on the computational complexity of approximating the partition function in 2-spin models. A long series of works, which we only touch upon later in this introduction, have identified the connection of the problem with decay of correlation properties in Gibbs measures of the infinite $\Delta$-regular tree. In statistical physics terms, this is known as the phase transition on the Bethe lattice. To briefly sketch this important concept, let ${\mathbb{T}_{\Delta}}$ denote the infinite $\Delta$-regular tree. When the parameters of the 2-spin model lie in the so-called uniqueness region of ${\mathbb{T}}_\Delta$, for finite complete trees of height $h$, the influence on the root from fixing a configuration on the leaves dies off in the limit as the height goes to infinity. In sharp contrast, in the non-uniqueness region there exist configurations of the leaves for which their influence on the root persists in the limit. For convenience, we often refer to the uniqueness region omitting the reference to ${\mathbb{T}}_\Delta$. The uniqueness/non-uniqueness regions can be determined rather easily on ${\mathbb{T}}_\Delta$ (for other infinite graphs this is far from easy, see [@Beffara] and [@RSTVY] for recent advances in the two dimensional grid). For the hard-core model, Kelly [@Kelly] showed non-uniqueness holds iff $\lambda>\lambda_c({\mathbb{T}}_\Delta):= \frac{(\Delta-1)^{\Delta-1}}{(\Delta-2)^\Delta}$. For the antiferromagnetic Ising model without external field, non-uniqueness holds iff $B<B_c({\mathbb{T}}_\Delta):=\frac{\Delta-2}{\Delta}$ (see, e.g., [@SST]). For general antiferromagnetic 2-spin models with soft constraints (i.e., $B_1B_2>0$), non-uniqueness holds iff $\sqrt{B_1B_2}<(\Delta-2)/\Delta$ and $\lambda\in (\lambda_1,\lambda_2)$ for some critical values $\lambda_1(\Delta,B_1,B_2),$ $\lambda_2(\Delta,B_1,B_2)$ (see, e.g., [@LLY]). Before stating our results, it will be instructive to review some known results from the literature. When $\lambda=1$, the partition function is the number of independent sets in $G$, and this quantity is \#P-complete to compute exactly [@Valiant], even when $\Delta=3$ [@Greenhill]. For the Ising model, once again, exact computation of the partition function for graphs of maximum degree $\Delta$ is \#P-complete, even for the ferromagnetic model [@JS]. Hence, the focus is on the computational complexity of approximating the partition function. This is where the phase transition on ${\mathbb{T}}_\Delta$ comes into play. Specifically, for the hard-core model, recently, a deep connection between uniqueness/non-uniqueness and the complexity of approximating the partition function was established. Namely, for graphs of constant maximum degree $\Delta$, Weitz [@Weitz] presented a beautiful ${\mathsf{FPTAS}}$ for approximating the partition function when $\lambda<\lambda_c({\mathbb{T}}_\Delta)$. On the other hand, Sly [@Sly10] presented a clever reduction which proves that for every $\Delta\geq 3$, there exists $\eps_\Delta>0$, such that it is NP-hard (unless RP=NP) to approximate the partition function for graphs of maximum degree $\Delta$ for any $\lambda$ where $\lambda_c({\mathbb{T}}_\Delta)<\lambda<\lambda_c({\mathbb{T}}_\Delta)+\eps_\Delta$. It was believed that Sly’s inapproximability result should hold for all $\lambda>\lambda_c({\mathbb{T}}_\Delta)$. However, Sly’s work utilized results of Mossel, Weitz and Wormald [@MWW] which showed that for a random $\Delta$-regular bipartite graph, an independent set chosen from the Gibbs distribution is “unbalanced” with high probability for $\lambda$ under the same condition as in Sly’s result. Mossel et al.’s work [@MWW] involves a technically complicated second moment argument that is characterised in [@Sly10] as a “technical tour-de-force”. In [@Galanis] we extended Mossel et al.’s work to all $\lambda>\lambda_c({\mathbb{T}}_\Delta)$ for $\Delta=3$ and $\Delta\geq 6$. However, that work built upon Mossel et al.’s work and had an even more complicated analysis. In this work, as a byproduct of the approach we devise for coping with the Ising model, we can prove inapproximability for the hard-core model for all $\lambda>\lambda_c({\mathbb{T}}_\Delta)$ for $\Delta=3,4,5$, and our proof is simpler than in [@Galanis]. This settles the picture for the hard-core model, and also proves Conjecture 1.2 of Mossel et al. [@MWW]. \[thm:hardcore\] Unless $NP=RP$, for the hard-core model, for all $\Delta\geq 3$, for all $\lambda$ in the non-uniqueness region of the infinite tree ${\mathbb{T}}_\Delta$, there does not exist an ${\mathsf{FPRAS}}$ for the partition function at activity $\lambda$ for graphs of maximum degree at most $\Delta$. Our main focus in this paper is to address whether the above phenomenon for the hard-core model occurs for other models. In particular, our goal is to address for 2-spin antiferromagnetic models whether the computational complexity of approximating the partition function for graphs with maximum degree $\Delta$ undergoes a transition that coincides exactly with the uniqueness/non-uniqueness phase transition on the infinite tree ${\mathbb{T}}_\Delta$. Let us first review existing results in the literature. For the ferromagnetic Ising model, Jerrum and Sinclair [@JS] presented an ${\mathsf{FPRAS}}$ for all graphs, for all $B\geq 1$ where $B=B_1=B_2$. For the antiferromagnetic Ising model, for general graphs there does not exist an ${\mathsf{FPRAS}}$ for $B<1$ [@JS], unless $NP=RP$. For constant $\Delta$, Sinclair et al. [@SST] extend Weitz’s approach to the antiferromagnetic Ising model (see also, [@ZLB; @RSTVY; @LLY] for other results on the Ising model), yielding in the uniqueness region an ${\mathsf{FPTAS}}$ for the partition function of graphs with maximum degree $\Delta$. We prove for the antiferromagnetic Ising model without external field that for all $B$ in the non-uniqueness region of the infinite tree there does not exist an ${\mathsf{FPRAS}}$ for the antiferromagnetic Ising model. \[thm:Ising\] Unless $NP=RP$, for the antiferromagnetic Ising model without external field, for all $\Delta\geq 3$, for all $B$ in the non-uniqueness region of the infinite tree ${\mathbb{T}}_\Delta$, there does not exist an ${\mathsf{FPRAS}}$ for the partition function at inverse temperature $B$ for graphs of maximum degree at most $\Delta$. The hard-core model and Ising model are the two most well-studied examples of 2-spin systems. For ferromagnetic 2-spin systems with no external field, i.e., $B_1B_2>1$ and $\lambda=1$, Goldberg et al. [@GJP] presented an ${\mathsf{FPRAS}}$ for the partition function for any graph. For any antiferromagnetic 2-spin model, for constant $\Delta$, an ${\mathsf{FPTAS}}$ for the partition function was obtained by Sinclair et al. [@SST] for $\Delta$-regular graphs in the uniqueness region of the infinite tree ${\mathbb{T}}_\Delta$, and by Li et al. [@LLY] for graphs of maximum degree $\Delta$ in the intersection of the uniqueness regions for infinite trees ${\mathbb{T}}_d$ for $d\leq\Delta$. For antiferromagnetic 2-spin models we obtain a complementary inapproximability result in the non-uniqueness region. Our results do not reach the uniqueness/non-uniqueness threshold in general. \[thm:2spin\] Unless $NP=RP$, for all $d\geq 2$, for all $B_1,B_2,\lambda$ that lie in the non-uniqueness region of the infinite tree ${\mathbb{T}}_{d+1}$ and $\sqrt{B_1B_2} \geq \frac{\sqrt{d}-1}{\sqrt{d}+1}$, there does not exist an ${\mathsf{FPRAS}}$ for the partition function at parameters $B_1,B_2,\lambda$ for graphs of maximum degree at most $\Delta=d+1$. Independent Results of Sly and Sun ---------------------------------- In an independent and simultaneous work, Sly and Sun [@SlySun] obtained closely related results. They prove inapproximability of the partition function in the case of the hard-core model and the Ising model. Their result for the Ising model also covers the case of an external field, which then extends to general 2-spin antiferromagnetic models [@SST; @SlySun]. The main technical result in our proof is a second moment argument to analyze the partition function for the Ising model on random $\Delta$-regular bipartite graphs, as outlined in the following Section \[sec:proof-approach\]. As a consequence we can estimate the partition function within any arbitrarily small polynomial factor, which allows us to use the reduction in [@Sly10]. In contrast, Sly and Sun’s approach build upon the recent interpolation scheme of Dembo et al. [@DMS] to analyze the logarithm of the partition function, which yields estimates of the partition function within an arbitrarily small exponential factor. To get their NP-hardness results, they use a modification of the approach appearing in [@Sly10]. Proof Outline {#sec:proof-approach} ============= The main element of the proof of Theorem \[thm:2spin\] (and similarly, the proofs of Theorems \[thm:hardcore\] and \[thm:Ising\]) is to analyze random $\Delta$-regular bipartite graphs to show a certain bimodality as in Mossel et al.’s work [@MWW] for the hard-core model. From there the same reduction of Sly [@Sly10] applies with some non-trivial modifications in the analysis. Hence, in some sense our main technical result is to show that in random $\Delta$-regular bipartite graphs, under the conditions of Theorem \[thm:2spin\], a configuration selected from the Gibbs distribution is “unbalanced” with high probability. To formally state our results, let us introduce some notation. Let ${\mathcal{G}}(n,\Delta)$ denote the probability distribution over bipartite graphs with $n+n$ vertices formed by taking the union of $\Delta$ random perfect matchings. We will denote the two sides of the bipartition of the graphs as $V_1,V_2$. Strictly speaking, this distribution is over bipartite multi-graphs. However, since our results hold asymptotically almost surely (a.a.s.) over ${\mathcal{G}}(n,\Delta)$, as noted in [@MWW], by contiguity arguments they also hold a.a.s. for the uniform distribution over bipartite $\Delta$-regular graphs. For a complete account of contiguity, we refer the reader to [@JLR Chapter 9]. For a graph $G\sim{\mathcal{G}}(n,\Delta)$ and a given configuration $\sigma\in \{-1,+1\}^V$, denote by $N_{-}(\sigma)$ the set of vertices assigned $-1$ and by $N_{+}(\sigma)$ the set of vertices assigned $+1$. For $\alpha,\beta\geq0$, let $$\Sigma^{\alpha,\beta} = \{\sigma\in\{-1,+1\}^V \big|\ |N_{-}(\sigma) \cap V_1| = \alpha n,\ |N_{-}(\sigma) \cap V_2| = \beta n\},$$ that is, $\alpha$ (resp. $\beta$) is the fraction of the vertices in $V_1$ (resp. $V_2$) whose spin is $-1$. Denote by $\Sigma^{Bal}$ the set of all balanced configurations, and for $\rho>0$, denote by $\Sigma^\rho$ the set of $\rho n$-unbalanced configurations. Formally, $$\Sigma^{Bal}=\mathop{\bigcup}_\alpha\Sigma^{\alpha,\alpha}, \quad \Sigma^\rho=\mathop{\bigcup}_{|\alpha-\beta|\geq \rho}\Sigma^{\alpha,\beta}.$$ \[thm:bimodality\] Under the hypotheses of Theorem \[thm:2spin\], there exist $a>1$ and $\rho>0$, such that asymptotically almost surely, for a graph $G$ sampled from ${\mathcal{G}}(n,\Delta)$, the Gibbs distribution $\mu$ satisfies: $$\mu_G\left(\Sigma^{Bal}\right)\leq a^{-n}\cdot \mu_G(\Sigma^\rho).$$ Therefore, the Glauber dynamics is torpidly mixing. Theorem \[thm:bimodality\], while interesting in its own right, provides us also with the necessary ingredients to carry out Sly’s reduction. To prove Theorem \[thm:bimodality\], we use a second moment approach as in [@MWW]. Specifically, for $G\sim {\mathcal{G}}(n,\Delta)$ we analyze the random variable $Z^{\alpha,\beta}_G$, where $Z^{\alpha,\beta}_G = \sum_{\sigma \in \Sigma^{\alpha,\beta}} w_G(\sigma)$, for some well-suited values of $\alpha,\beta$. Intuitively, the $\alpha,\beta$ should be selected so as to maximize the first moment of $Z^{\alpha,\beta}_G$. For Theorem \[thm:bimodality\] to hold we need that the maximum occurs for $\alpha\neq\beta$. This is indeed true in the non-uniqueness region and moreover there is an explicit connection with three specific Gibbs measures in the infinite tree ${\mathbb{T}_{\Delta}}$. These Gibbs measures are extremal and are invariant under parity preserving transformations. We next introduce the relevant notation here, deferring a slightly more thorough discussion to Section \[sec:tree-recursions\]. For a more complete account of extremality, we refer the reader to [@Georgii]. For every $B_1,B_2,\lambda>0$ there exists a unique translation invariant Gibbs measure $\mu^*$ on ${\mathbb{T}_{\Delta}}$, known as the free measure (free boundary condition). We will denote the marginal probability for the root $\rho$ being assigned spin $-1$ by $p^* := \mu^*(\sigma_\rho=-1)$. In the non-uniqueness region, there also exist two semi-translation invariant measures $\mu^+$ and $\mu^-$, corresponding to the even and odd boundary conditions. More precisely, the measure $\mu^+$ (resp. $\mu^{-}$) is obtained by conditioning on the spins of the vertices at level $2\ell$ (resp. $2\ell+1$) being $-1$ and taking the weak limit as $\ell\rightarrow\infty$. Analogously to $p^{*}$, let $p^+ := \mu^+(\sigma_\rho=-1)$, $p^-:=\mu^-(\sigma_\rho=-1)$ denote the marginals for the root being assigned spin $-1$ under the even and odd boundaries (note that $\pm$ refers to the parity of the level where the boundary condition is imposed, while $*$ to the free boundary). Using (anti)monotonicity arguments, it can easily be proved that $p^-<p^*<p^+$. \[lem:firstmomentmax\] For the distribution ${\mathcal{G}}(n,\Delta)$, $\displaystyle\lim_{n\rightarrow\infty}\frac{1}{n}\log{\mathbb{E}}_{\mathcal{G}}[Z^{\alpha,\beta}_G]$ is maximized for: 1. $(\alpha,\beta)=(p^*,p^*)$, whenever $B_1,B_2,\lambda$ are in the uniqueness region of the infinite tree ${\mathbb{T}_{\Delta}}$. 2. $(\alpha,\beta)=(p^{\pm},p^{\mp})$, whenever $B_1,B_2,\lambda$ are in the non-uniqueness region of the infinite tree ${\mathbb{T}_{\Delta}}$. Note that Lemma \[lem:firstmomentmax\] already yields a weak form of a bimodality in the Gibbs distribution on random $\Delta$-regular graphs: in the non-uniqueness region, the configurations with the largest contribution are unbalanced in expectation. Since $Z^{\alpha,\beta}_G$ is typically exponential in $n$, Markov’s inequality implies an upper bound on the number of balanced configurations, which holds with high probability over the choice $G\sim {\mathcal{G}}(n,\Delta)$. However, to get tight results, we need a strong form of this bimodality. To do this, we look more carefully at the random variable $Z^{\alpha,\beta}_G$ for $(\alpha,\beta)=(p^\pm,p^\mp)$. Note that if this variable is concentrated around its expectation, Lemma \[lem:firstmomentmax\] establishes that it dominates exponentially variables $Z^{\alpha',\beta'}_G$ with ${\lVert(\alpha',\beta')-(p^\pm,p^\mp)\rVert}\geq \epsilon$ where $\epsilon$ can be made arbitrarily small for sufficiently large $n$. Hence, one obtains the required bottleneck that is needed for the proof of Theorem \[thm:bimodality\]. The torpid mixing of Glauber dynamics follows then by a conductance argument given in [@DFJ]. In fact, the same idea yields something stronger: a typical configuration of the system will be from the set $\Sigma^{p^{\pm},p^{\mp}}$ (roughly). Informally, this allows us to say that with high probability $G$ behaves as a Boolean gadget depending on the phase of a configuration, where the phase is the bipartition which has the largest number of vertices assigned -1. Pasting $G$ on the vertices of a graph $H$, it turns out that the partition function of $H$ is dominated by configurations where the phases of the gadgets correspond to a maximum cut of $H$. Of course, the quantification of this nice idea requires serious effort, but this has already be done in [@Sly10]. To hint why such a scheme succeeds, note for example that in the antiferromagnetic Ising model, a ground state of the Gibbs distribution corresponds to a maximum cut. Thus, the reduction ensures that the phase configurations induced by maximum cut partitions of the original graph $H$ dominate exponentially the partition function, and hence one can recover a maximum cut from an approximate sample from the Gibbs distribution. Let us now return to the technical core of the method, which requires analyzing the second moment of $Z^{\alpha,\beta}_G$ for $(\alpha,\beta)=(p^\pm,p^\mp)$. By symmetry, we may clearly focus on $(\alpha,\beta)=(p^+,p^-)$. Pick two configurations $\sigma_1,\sigma_2$ from the set $\Sigma^{p^+,p^-}$ (not necessarily distinct). The two most important variables here are the overlap of the configurations $\sigma_1,\sigma_2$. Formally, let $$\Sigma^{\alpha,\beta}_{\gamma,\delta}=\big\{(\sigma_1,\sigma_2)\,\big|\,\sigma_1,\sigma_2\in \Sigma^{\alpha,\beta}, |N_{-}(\sigma_1)\cap N_{-}(\sigma_2) \cap V_1|=\gamma n,\ |N_{-}(\sigma_1)\cap N_{-}(\sigma_2) \cap V_2|=\delta n\big\},$$ and $Y^{\gamma,\delta}_G=\sum_{(\sigma_1,\sigma_2)\in\Sigma^{\alpha,\beta}_{\gamma,\delta}}w_G(\sigma_1)w_G(\sigma_2).$ We will study $Y^{\gamma,\delta}_G$ for $(\alpha,\beta)=(p^+,p^-)$, so this is why we dropped the dependence on $\alpha,\beta$. Clearly, ${\mathbb{E}}_{\mathcal{G}}[Y^{\gamma,\delta}_G]$ is the contribution of pairs of configurations in $\Sigma^{\alpha,\beta}$ with overlap $\gamma n,\delta n$ to the second moment ${\mathbb{E}}_{\mathcal{G}}[(Z^{\alpha,\beta}_G)^2]$. In our case, the second moment approach has any chance of succeeding when the largest contribution to the second moment comes from uncorrelated pairs of configurations. This is captured by the following condition. \[cond:maxima\] For $(\alpha,\beta)=(p^+,p^-)$, $\displaystyle\lim_{n\rightarrow\infty}\frac{1}{n}\log{\mathbb{E}}_{\mathcal{G}}[Y^{\gamma,\delta}_G]$ is maximized at $\gamma=\alpha^2, \delta=\beta^2$. \[lem:secondmax2spin\] Under the hypotheses of Theorem \[thm:2spin\], Condition \[cond:maxima\] holds. \[lem:secondmaxising\] For the antiferromagnetic Ising model with $\lambda=1$, $\Delta=3$ and $B$ in the non-uniqueness region, Condition \[cond:maxima\] holds. \[lem:secondmaxhardcore\] For the hard-core model, $\Delta=3,4,5$ and $\lambda$ in the non-uniqueness region, Condition \[cond:maxima\] holds. The innocuous statements of Lemmas \[lem:secondmax2spin\], \[lem:secondmaxising\], and \[lem:secondmaxhardcore\] are rather misleading. Namely proving that uncorrelated pairs of random variables “dominate” the second moment is a standard complication of the second moment approach in a variety of settings, see for example [@AN; @AP; @MWW]. To highlight the technical difficulties, the second moment typically has a number of extra other variables that need to be considered and those are usually treated by sophisticated analysis arguments. We escape this paradigm in the following sense: our arguments try to explain the success of the second moment by relating it to the intrinsic nature of the problem at hand (in our context, uniqueness on the infinite tree). To stress the gains of such a scheme, it underlies perhaps a more generic method to translate uniqueness proofs to similar second moment arguments. In the hard-core model, due to the hard constraints, only one extra variable (besides $\gamma,\delta$) was needed and it could be expressed explicitly in terms of $\gamma,\delta$, so that, in fact, one had a 2-variable optimization. Even with this seemingly simple setup and despite the variety and complexity of approaches appearing in [@MWW], [@Sly10] and [@Galanis], certain regions for $\Delta=4,5$ remained open. In the general 2-spin model the second moment is more complex than in the hard-core model in the sense that the soft constraints cause the number of extra variables to increase to nine. To make matters even worse, one cannot get rid of those extra variables in some easy algebraic way, since such a scheme reduces to a sixth order polynomial equation in terms of $\gamma,\delta$. These obstacles are surmounted in the proof of Lemma \[lem:secondmax2spin\] by a very clean argument: the analysis of the second moment is reduced to a certain stable form of tree recursions on ${\mathbb{T}_{\Delta}}$, allowing for a short inequality argument. After establishing Condition \[cond:maxima\], one obtains that for $(\alpha,\beta)=(p^\pm,p^\mp)$ it holds that ${\mathbb{E}}_{\mathcal{G}}[(Z^{\alpha,\beta}_G)^2]\approx {\mathbb{E}}_{\mathcal{G}}[Y^{\alpha^2,\beta^2}_G]$. At this point, we are in position to study the ratio ${\mathbb{E}}_{\mathcal{G}}[(Z^{\alpha,\beta}_G)^2]/({\mathbb{E}}_{\mathcal{G}}[Z^{\alpha,\beta}_G])^2$. This is a technical computation and requires an asymptotic tight approximation of sums by appropriate Gaussian integrals. While the asymptotic value of the ratio is a constant, as in [@MWW], the constant is greater than 1. \[lem:ratiofirstsecond\] When Condition \[cond:maxima\] is true, for $(\alpha,\beta)=(p^\pm,p^\mp)$, it holds that $$\lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{{\mathcal{G}}}\left[(Z^{\alpha,\beta}_G)^2\right]}{\left({\mathbb{E}}_{{\mathcal{G}}}[Z^{\alpha,\beta}_G]\right)^2}=\big(1-\omega^2\big)^{-(\Delta-1)/2}\big(1-(\Delta-1)^2\omega^2\big)^{-1/2},$$ where $\omega$ is given by . The formula given in the lemma is valid for the hard-core model as well ($B_1=0, B_2=1$), agreeing with the statement of [@MWW Theorem 3.3]. This complication does not allow one to get high probability statements using the second moment method. This is by now a standard scenario in the theory of random regular graphs and it can be dealt with by the small graph conditioning method [@Wormald; @JLR]. The application of the method utilizes the moments’ analysis. In our setting, the method explains the variance of the variables $Z^{\alpha,\beta}_G$ by conditioning on the occurrence of a cycle in a random $\Delta$-regular graph. The consequence of applying the small graph conditioning method is the following result. \[lem:smallgraph\] When Condition \[cond:maxima\] is true, for $(\alpha,\beta)=(p^\pm,p^\mp)$, it holds asymptotically almost surely over the graph $G\sim{\mathcal{G}}(n,\Delta)$ that $$Z^{\alpha,\beta}_G\geq \frac{1}{n}{\mathbb{E}}_{\mathcal{G}}[Z^{\alpha,\beta}_G].$$ Using Lemmas \[lem:firstmomentmax\] and \[lem:smallgraph\], one can carry out the proof of Theorem \[thm:bimodality\] as in [@MWW Proof of Theorem 2.2] (see also [@Galanis Proofs of Theorems 2 and 3]). This remaining argument was sketched earlier in this section. The outline of the rest of the paper is as follows. In Section \[sec:tree-recursions\] we derive certain properties for the infinite tree which are used throughout the remainder of the paper. We then look at the first and second moments in Section \[sec:moments\]. In Section \[sec:maxima\] we look at the optimization problems for the first and second moments and prove Lemmas \[lem:firstmomentmax\] and  \[lem:secondmax2spin\]. In Section \[sec:NP-outline\] we outline the proof of our inapproximability results based on Sly’s reduction for the hard-core model. In Section \[sec:calculating-moments\] we analyze the ratio of the second moment to the first moment squared for random $\Delta$-regular graphs and for the modified graph used in Sly’s reduction. We then apply the small graph conditioning method in Section \[sec:smallgraph\] to get asymptotically almost sure results. Finally, in Section \[sec:remainingproofs\] we give the proofs of some remaining lemmas, including Lemmas \[lem:secondmaxising\] and \[lem:secondmaxhardcore\]. Extremal Measures on the Infinite Tree for 2-Spin Models {#sec:tree-recursions} ======================================================== In Section \[sec:proof-approach\], we defined the measures $\mu^{*},\mu^{\pm}$ on ${\mathbb{T}_{\Delta}}$ which will be of interest to us. In our context, the phase transition on the infinite $(\Delta-1)$-ary tree ${\hat{\mathbb{T}}_{\Delta}}$ rooted at $\rho$ will also be of interest. The uniqueness/non-uniqueness regions for this slightly modified tree coincide with those of ${\mathbb{T}_{\Delta}}$, albeit with different occupation probabilities for the root in the respective free and fixed boundary measures. We denote by $\hat{\mu}^{\pm}, \hat{\mu}^*$ the measures on ${\hat{\mathbb{T}}_{\Delta}}$ which are the analogues of the measures $\mu^{\pm}, \mu^*$ on ${\mathbb{T}_{\Delta}}$, respectively. Recall that we use $+,-$ to denote the even and odd boundary conditions, and $*$ for the free boundary condition. We will denote by $q^{\pm},q^*$ the marginal probabilities of the root being assigned $-1$ in each of these measures, i.e., $q^{\pm}=\hat{\mu}^{\pm}(\sigma_\rho=-1)$ and $q^{*}=\hat{\mu}^{*}(\sigma_\rho=-1)$. Once again, it can easily be proved that $q^-<q^*<q^+$ using (anti)monotonicity arguments. A well known and interesting property that the measures $\mu^{\pm}$ (and $\hat{\mu}^{\pm}$) possess is that of extremality. Namely, in the non-uniqueness region they cannot be written as nontrivial convex combinations of other Gibbs measures. In this particular setting, this follows from (and can be interpreted as) the fact that the even and odd boundaries force the maximal bias on the probability that the root $\rho$ is assigned the spin $-1$ as the height of the tree goes to infinity; in other words, for every Gibbs measure $\mu\neq \mu^{\pm}$ on ${\mathbb{T}_{\Delta}}$, it holds that $p^{-}<\mu(\sigma_\rho=-1)<p^{+}$. In this section, we state some further properties for the densities $q^\pm, q^*$ and $p^\pm,p^*$, which we are going to utilize throughout the text. Standard tree recursions for Gibbs measures (e.g., see [@SST]) establish that $$\frac{q^\pm}{1-q^\pm}=\lambda\left(\frac{B_1\cdot \frac{q^\mp}{1-q^\mp}+1}{\frac{q^\mp}{1-q^\mp}+B_2}\right)^{\Delta-1}, \quad \frac{q^*}{1-q^*}=\lambda\left(\frac{B_1\cdot \frac{q^*}{1-q^*}+1}{\frac{q^*}{1-q^*}+B_2}\right)^{\Delta-1}. \label{eq:recurone}$$ We define separate expressions for $\frac{q^\pm}{1-q^\pm}, \frac{q^*}{1-q^*}$, i.e., $$Q^{+}=\frac{q^+}{1-q^+},\qquad Q^{-}=\frac{q^-}{1-q^-}, \qquad Q^{*}=\frac{q^*}{1-q^*}.\label{eq:qpqm}$$ The following lemma summarizes the properties of the occupation densities $q^+, q^-, q^*$ which we are going to utilize. Its rather folklore proof is a straightforward analysis of appropriate tree recursions and is omitted. \[lem:densitiesone\] For $\Delta\geq 3$ and ${B}_1,{B}_2,\lambda>0$, consider the system of equations $$x=\lambda\left(\frac{B_1 y+1}{y+B_2}\right)^{\Delta-1},\ y=\lambda\left(\frac{B_1 x+1}{x+B_2}\right)^{\Delta-1}, \label{eq:densitiesone}$$ with $x,y\geq 0$. 1. In the uniqueness region, the only solution to is $(Q^*,Q^*)$. 2. In the non-uniqueness region, has exactly three solutions which are given by $(Q^\pm,Q^\mp)$ and $(Q^*,Q^*)$. It holds that $Q^-<Q^*<Q^+$. We also define the following quantities $\omega,\ \omega^{*}$ which will emerge throughout the paper (and will be motivated shortly afterwards): $$\omega:=\frac{(1-B_1B_2)^2Q^+Q^-}{(B_2+Q^-)(B_2+Q^+)(1+B_1Q^-)(1+B_2Q^+)},\qquad \omega^*:=\frac{(1-B_1B_2)^2(Q^*)^2}{(B_2+Q^*)^2(1+B_1Q^*)^2}.\label{eq:definitionofomega}$$ Note that $\omega^*$ is obtained from $\omega$ after identifying $Q^\pm$ with $Q^*$. The following lemma for $\omega,\omega^*$ will be crucial to prove some optimality conditions required to establish Lemma \[lem:firstmomentmax\] and Condition \[cond:maxima\]. Its proof is given in Section \[sec:technicalinequality\]. \[lem:technicalinequality\] For $(B_1,B_2,\lambda)$ in the non-uniqueness region, it holds that $$(\Delta-1)^2\omega<1<(\Delta-1)^2\omega^*.$$ Let us explain the mysterious looking inequality in Lemma \[lem:technicalinequality\]. While not obvious, it is related with the two-step recurrence on the infinite tree ${\hat{\mathbb{T}}_{\Delta}}$. The two step-recurrence is obtained by the composition of two successive one step recurrences. In the non-uniqueness region, the fixed points of the two step recurrence are exactly $q^\pm,q^*$ (this is basically Lemma \[lem:densitiesone\]). The inequality in Lemma \[lem:technicalinequality\] establishes, after some calculations, that the fixed point $q^+$ (resp. $q^-$) is an attractor of the two step recurrence in $\hat{\mu}^+$ (resp. $\hat{\mu}^-$), while the fixed point $q^*$ is repulsive. More specifically, the derivative of the two step function at $q^\pm$ is equal to $(\Delta-1)^2\omega$, and similarly at $q^*$ is equal to $(\Delta-1)^2\omega^*$. This property will be used to show that the measures $\hat{\mu}^{\pm}$ satisfy a strong form of non-reconstruction with exponential decay of correlation. We should also note that the analysis in [@SST] (see also [@LLY]) establishes the following criterion for uniqueness, which we reformulate for our purposes: uniqueness holds on ${\hat{\mathbb{T}}_{\Delta}}$ iff $(\Delta-1)^2\omega^*\leq 1$. Finally, let us translate the above to obtain a handle on $p^\pm, p^*$. Using a slight modification of the tree recursions in ${\hat{\mathbb{T}}_{\Delta}}$, one obtains that $$\frac{p^\pm}{1-p^\pm}=\frac{q^\pm}{1-q^\pm}\left(\frac{B_1\cdot \frac{q^\mp}{1-q^\mp}+1}{\frac{q^\mp}{1-q^\mp}+B_2}\right), \qquad \frac{p^*}{1-p^*}=\frac{q^*}{1-q^*}\left(\frac{B_1\cdot \frac{q^*}{1-q^*}+1}{\frac{q^*}{1-q^*}+B_2}\right). \label{eq:pppmtemp}$$ Using and , an easy calculation gives $$p^{\pm}=\frac{Q^{\pm}(1+B_1 Q^{\mp})}{B_2+Q^{+}+Q^{-}+B_1 Q^{+} Q^{-}}.\label{eq:pppm}$$ Equation will prove to be very useful to simplify seemingly intricate calculations. For the case of the Ising model with zero external field ($B_1=B_2=B,\lambda=1$), it holds that $p^++p^-=q^++q^-=1$, $p^*=q^*=1/2$. These properties may be proved easily using the symmetries of the model. Namely, observe that if $(x,y)$ satisfy then so do $(1/y,1/x)$. By items 1 and 2 of Lemma \[lem:densitiesone\], this implies that $q^++q^-=1$ and $q^*=1/2$. By , it also follows that $p^++p^-=1$, $p^*=1/2$. Moment Analysis {#sec:moments} =============== In this section, we give expressions for ${\mathbb{E}}_{{\mathcal{G}}}\big[Z^{\alpha,\beta}_G\big]$, ${\mathbb{E}}_{\mathcal{G}}\big[Y^{\gamma,\delta}_G\big]$ and then study the asymptotics of $\frac{1}{n}\log{\mathbb{E}}_{{\mathcal{G}}}\big[Z^{\alpha,\beta}_G\big]$, $\frac{1}{n}\log{\mathbb{E}}_{{\mathcal{G}}}\big[Y^{\gamma,\delta}_G\big]$. First Moment {#sec:firstmomentform} ------------ The first moment of $Z^{\alpha,\beta}_G$ is given by $${\mathbb{E}}_{{\mathcal{G}}}\big[Z^{\alpha,\beta}_G\big] = \lambda^{n(\alpha+\beta)}\binom{n}{\alpha n}\binom{n}{\beta n}\left(\sum_{\delta\leq \alpha, \delta\leq \beta} \frac{\binom{\alpha n}{\delta n}\binom{(1-\alpha)n}{(\beta-\delta) n}}{\binom{n}{\beta n}}B_1^{\delta n}B_2^{(1-\alpha-\beta+\delta)n}\right)^{\Delta}.\label{eq:firstmoment1}$$ Ignoring the contribution of the external field $\lambda$, the terms outside the sum count $|\Sigma^{\alpha,\beta}_G|$ . The sum gives the expected contribution to $w_G(\sigma)$ (for $\sigma\in \Sigma^{\alpha,\beta}_G$) from a *single* matching, where $\delta n$ is the number of edges between $N_-(\sigma)\cap V_1$ and $N_-(\sigma)\cap V_2$ in a single matching. The weight of $\sigma$ for a single matching depends only on $\delta n$ and is equal to $B_1^{\delta n}B_2^{(1-\alpha-\beta+\delta)n}$. The fraction in the sum is the probability that there are exactly $\delta n$ edges from $N_-(\sigma)\cap V_2$ to $N_+(\sigma)\cap V_2$. Since the matchings are chosen independently, the exponentiation to the power $\Delta$ gives the expected contribution of $\sigma\in \Sigma^{\alpha,\beta}_G$ to the partition function for $\Delta$ matchings. Finally, the $\lambda$ term in the beginning accounts for the contribution of the external field for $\sigma\in \Sigma^{\alpha,\beta}_G$. A different way to arrive at a slightly different formulation for the first moment follows. Here, we introduce some redundant variables at the cost of introducing some complexity of the presentation, but underlying a general approach to treat such expressions. Observe that the sets $N^{-}(\sigma)\cap V_1,\ N^{+}(\sigma)\cap V_1$ form a partition of $V_1$ and similarly for $V_2$. Thus for a perfect matching let $x_{ij}$, ($1\leq i,j\leq 2$) denote the number of edges having their endpoints in partition $i$ of $V_1$ and in partition $j$ of $V_2$. Note that the $x_{ij}$ must satisfy $$\begin{array}{ll} x_{11}+x_{12}=\alpha,& x_{11}+x_{21}=\beta,\\ x_{21}+x_{22}=1-\alpha,& x_{12}+x_{22}=1-\beta,\\ x_{ij}\geq 0.&\\ \end{array}\label{eq:xregion}$$ The first moment may be written as $${\mathbb{E}}_{{\mathcal{G}}}\big[Z^{\alpha,\beta}_G\big] = \lambda^{(\alpha+\beta)n}\binom{n}{\alpha n}\binom{n}{\beta n}\left(\sum_{\mathbf{X}} \frac{\binom{\alpha n}{x_{11}n}\binom{(1-\alpha)n}{x_{21}n}\binom{\beta n}{x_{11}n}\binom{(1-\beta)n}{x_{21}n}}{\binom{n}{x_{11}n,x_{12}n,x_{21}n,x_{22}n}}B_1^{x_{11}n}B_2^{x_{22}n}\right)^{\Delta}. \label{eq:firstmoment2}$$ In the above sum, we denote by $\mathbf{X}$ the $x_{ij}$ which satisfy . Note that the expressions and are completely equivalent after identifying $\delta$ with $x_{11}$ and using to express the variables $x_{12},x_{21},x_{22}$ in terms of $x_{11}$. Second Moment -------------  \[sec:secondmomentform\] We continue with the ideas presented in Section \[sec:firstmomentform\], to get an expression for ${\mathbb{E}}_{\mathcal{G}}\big[Y^{\gamma,\delta}_G\big]$. Recall that for the formulation of the second moment $\alpha$ and $\beta$ are fixed, and hence suppressed from the notation. Now there are four induced parts by $\sigma_1,\sigma_2$ on each $V_1$ and $V_2$. Let $\gamma n,\delta n$ stand for the number of vertices in $V_1$ and $V_2$ whose spin is $-1$ in both $\sigma_1,\sigma_2$ and $y_{ij}n$ the number of edges incident with endpoints in partition $i$ of $V_1$ and partition $j$ of $V_2$ in a single matching. For ease of exposition, define $$\begin{array}{lll} L_1=\gamma,& L_2=L_3=\alpha-\gamma,& L_4=1-2\alpha+\gamma,\\ R_1=\delta,& R_2=R_3=\beta-\delta,& R_4=1-2\beta+\delta. \end{array}$$ In the same spirit with expression for the first moment, we have that $$\begin{aligned} {\mathbb{E}}_{\mathcal{G}}\big[Y^{\gamma,\delta}_G\big]&=\lambda^{2(\alpha+\beta)n}\binom{n}{\alpha n}\binom{n}{\beta n}\binom{\alpha n}{\gamma n}\binom{(1-\alpha)n}{(\alpha-\gamma)n}\binom{\beta n}{\delta n}\binom{(1-\beta)n}{(\beta-\delta)n}\label{eq:secondmoment}\\ &\times \bigg(\sum_{\mathbf{Y}}\prod^4_{i=1} \binom{L_i n}{y_{i1}n,y_{i2}n,y_{i3}n,y_{i4}n}\prod^4_{j=1} \binom{R_jn}{y_{1j}n,y_{2j}n,y_{3j}n,y_{4j}n} /\binom{n}{y_{11}n, y_{12}n, \dots ,y_{44}n}\notag\\ &\qquad \times B_1^{(2y_{11} + y_{12} + y_{13}+ y_{21} + y_{22}+y_{31}+y_{33})n}B_2^{(y_{22}+y_{24}+y_{33}+y_{34}+y_{42}+y_{43}+2y_{44})n}\bigg)^{\Delta}.\notag\end{aligned}$$ In the above sum, we denote by $\mathbf{Y}$ the $y_{ij}$ which satisfy $$\begin{array}{ll} \sum_{j}y_{ij}=L_i,&\mbox{ for } i=1,\hdots,4\\ \sum_{i}y_{ij}=R_j,&\mbox{ for } j=1,\hdots,4\\ y_{ij}\geq 0.&\\ \end{array}\label{eq:yregion}$$ The interpretation of formula should be clear: ignoring the external field, the first line accounts for $|\Sigma^{\alpha,\beta}_{\gamma,\delta}|$, a term in the sum accounts for the weight of a matching specified by the cardinalities $Y$ (last line) as well as the probability of such a matching occurring (middle line). Asymptotics of the Logarithms of the Moments {#sec:generallog} -------------------------------------------- We next turn to finding the asymptotics of $\frac{1}{n}\log{\mathbb{E}}_{{\mathcal{G}}}\big[Z^{\alpha,\beta}_G\big]$, $\frac{1}{n}\log{\mathbb{E}}_{{\mathcal{G}}}\big[Y^{\gamma,\delta}_G\big]$. The asymptotic order of the sums in and are dominated by their maximum terms. For the sole purpose of having a unifying treatment for soft and hard constrained 2-spin models, here and in the rest of the paper, we adopt the usual conventions that $\ln 0\equiv -\infty$ and $0\ln 0\equiv 0$. Standard application of Stirling’s approximation yields for the first moment: $$\begin{aligned} \lim_{n\rightarrow \infty}&\ \frac{1}{n}\log{\mathbb{E}}_{{\mathcal{G}}}\big[Z^{\alpha,\beta}_G\big] = \max_{\mathbf{X}}\Phi_1(\alpha,\beta,\mathbf{X}),\label{eq:limitfirst}\\ \mbox{ where } \ \ \ \Phi_1(\alpha,\beta,\mathbf{X})&:=(\alpha+\beta)\ln \lambda+(\Delta-1)f_1(\alpha,\beta)+ \Delta g_1(\mathbf{X})\notag\\ f_1(\alpha,\beta)&:=\alpha\ln\alpha+(1-\alpha)\ln(1-\alpha)+\beta\ln\beta+(1-\beta)\ln(1-\beta)\notag\\ g_1(\mathbf{X})&:=(x_{11}\ln B_1+x_{22}\ln B_2)-\mbox{$\sum_{i,j}$}x_{ij}\ln x_{ij}.\notag\end{aligned}$$ And for the second moment: $$\begin{aligned} &\lim_{n\rightarrow \infty} \frac{1}{n}\log{\mathbb{E}}_{{\mathcal{G}}}\big[Y^{\gamma,\delta}_G\big]= \max_\mathbf{Y}\Phi_2(\gamma,\delta,\mathbf{Y}), \label{eq:limitsecond}\\ \mbox{ where } \ \ \ \Phi_2(\gamma,\delta,\mathbf{Y})&:=2(\alpha+\beta)\ln \lambda+ (\Delta-1)f_2(\gamma,\delta)+\Delta g_2(\mathbf{Y})\notag\\[0.15cm] f_2(\gamma,\delta)&:=2(\alpha-\gamma)\ln(\alpha-\gamma)+\gamma \ln\gamma+(1-2\alpha+\gamma)\ln(1-2\alpha+\gamma)+\notag\\ &\qquad \qquad 2(\beta-\delta)\ln(\beta-\delta)+\delta \ln\delta+(1-2\beta+\delta)\ln(1-2\beta+\delta)\notag\\[0.15cm] g_2(\mathbf{Y})&:=(2y_{11} + y_{12} + y_{13}+ y_{21} + y_{22}+y_{31}+y_{33})\ln B_1\notag\\ &\qquad +(y_{22}+y_{24}+y_{33}+y_{34}+y_{42}+y_{43}+2y_{44})\ln B_2-\mbox{$\sum_{i,j}$}y_{ij}\ln y_{ij}.\notag\end{aligned}$$ Recall that we are going to study the second moment for specific values of $\alpha,\beta$, namely $(\alpha,\beta)=(p^\pm,p^\mp)$, so this is why we dropped the dependence of the functions $\Phi_2,f_2$ on $\alpha,\beta$. The functions $g_1, g_2$ are defined on the regions , respectively. For fixed $\alpha,\beta,\gamma,\delta$, they are strictly concave over the convex regions on which they are defined and hence they have a unique maximum, so this also the case for the functions $\Phi_1, \Phi_2$. The limits and can thus be justified using the standard Laplace method (see for example [@deBru Chapter 4]). Finding the Maxima ================== In this section, we use the information obtained in Section \[sec:generallog\] to establish Lemmas \[lem:firstmomentmax\] and \[lem:secondmax2spin\]. The proofs for Lemmas \[lem:secondmaxising\] and \[lem:secondmaxhardcore\] are given in Sections \[sec:secondmaxisingproof\] and \[sec:hardhard\] respectively, see also Section \[sec:remarks\] for an overview. These proofs are based on the approach given in this section. \[sec:maxima\] Optimizing Entropy Distributions {#sec:maxentropy} -------------------------------- By the limits and , to establish Lemmas \[lem:firstmomentmax\] and \[lem:secondmax2spin\] (and \[lem:secondmaxising\], \[lem:secondmaxhardcore\] as well), it suffices to study the maxima of the functions $$\phi_1(\alpha,\beta):=\max_{\mathbf{X}} \Phi_1(\alpha,\beta, \mathbf{X})\mbox{ and }\phi_2(\gamma,\delta):=\max_{\mathbf{Y}} \Phi_2(\gamma,\delta, \mathbf{Y}).$$ A small abstraction allows us to treat the two functions in a unifying way. Namely, observe that $\phi_1,\phi_2$ have the form $\phi(u,v)=\max_{\mathbf{Z}}\Phi(u,v,\mathbf{Z})$, where $\Phi$ can be decomposed as $\Phi(u,v,\mathbf{Z})=f(u,v)+g(\mathbf{Z})$, for some functions $f,g$ and $\mathbf{Z}$ stands for $mn$ non-negative variables $Z_{ij}$, $i=1,\hdots,m$, $j=1,\hdots,n$. Thus in the context of $\phi_1$ (resp. $\phi_2$), $u,v,\mathbf{Z}$ coorespond to $\alpha,\beta,\mathbf{X}$ (resp. $\gamma,\delta,\mathbf{Y}$). Moreover, $\mathbf{Z}$ satisfies some prescribed marginals, i.e., $$\label{e1} \mbox{$\sum_{j}$} Z_{ij} = \alpha_i(u),\quad \mbox{$\sum_{i}$} Z_{ij} = \beta_j(v), \quad Z_{ij}\geq 0,$$ for some functions $\alpha_i(u)$ and $\beta_j(v)$, which satisfy $\sum_{i} \alpha_i(u) =\sum_{j} \beta_j(v) = 1$. In this section, we will only consider $u,v$ such that $\alpha_i(u)\neq 0\neq \beta_j(v)$ for all $i,j$, since the remaining cases correspond to boundary conditions in our setup and these are treated more easily in the specific instance of the maximization (see the relevant Lemmas \[lem:firboundary\] and \[lem:secboundary\]). Finally, the function $g(\mathbf{Z})$ is an entropy-like function and has the form $$g(\mathbf{Z})=\mbox{$\sum_{i}\sum_j$} \left(Z_{ij} \ln M_{ij} - Z_{ij}\ln Z_{ij}\right),$$ where $M\in\mathbb{R}^{m\times n}$ is a matrix with non-negative entries. We will use another interpretation of the function $g(\mathbf{Z})$, which will be handy when we perform integrations: let $\mathbf{Z^f}$ denote the variables $Z_{ij}$ for $i=1,\hdots, m-1,$ $j=1,\hdots,n-1$. Using the equations in , we may write $g(\mathbf{Z})$ as a function of the variables $\mathbf{Z^f}$. Note that the region must be adjusted to account for the non-negativity of the substituted variables. Other than this, the “new” function has exactly the same behaviour as $g(\mathbf{Z})$ but in a full-dimensional space. Thus we will also think of $g(\mathbf{Z})$ as a function of the variables $\mathbf{Z^f}$ where the $Z_{ij}$ with $i=m$ or $j=n$ are just shorthands for the expressions of these variables when substituted by the equations in . We will refer to this reformulation as the *full-dimensional representation* of $g(\mathbf{Z})$. We define similarly the full-dimensional representation of $\Phi(u,v,\mathbf{Z})$. It is immediate to see that $g(\mathbf{Z})$ is strictly concave in the convex region , hence it attains its maximum over the region at a unique point. A tedious but straightforward optimization of $g(\mathbf{Z})$ using Lagrange multipliers, yields: \[lem:gmax\] Fix $u,v$ such that $\alpha_i(u)\neq 0 \neq \beta_j(v)$ for all $i,j$. Let $\mathbf{Z}^*=\arg\max_{\mathbf{Z}}g(\mathbf{Z})$ (the maximum taken over the region ). Then $$\label{e2} Z^*_{ij} = M_{ij} R_i C_j,$$ where $R_i$ and $C_j$ are positive real numbers, which satisfy $$\label{e3} R_i \sum_{j} M_{ij} C_j = \alpha_i\quad\mbox{and}\quad C_j \sum_{i} M_{ij} R_i = \beta_j.$$ Moreover, the full-dimensional representation of $g(\mathbf{Z})$ decays quadratically in a sufficiently small ball around $\mathbf{Z}^*$. The proof of Lemma \[lem:gmax\] is given in Section \[sec:maxentropyproof\]. It is relevant to point out that while the values of $Z^*_{ij}$ are unique in the region , the values of $R_i,C_j$ which satisfy are not unique since a positive solution $\{\{R_i\}_i,\{C_j\}_j\}$ of gives rise to a positive solution $\{\{tR_i\}_i,\{C_j/t\}_j\}$ for arbitrary $t>0$. Modulo this mapping, the uniqueness of $Z^*_{ij}$ easily implies that the positive $R_i,C_j$ which satisfy are otherwise unique. Plugging in $g(\mathbf{Z})$, we obtain $$\label{eq:QQ} \max_{\mathbf{Z}}g(\mathbf{Z}) = -\mbox{$\sum_{i} \sum_{j}$} M_{ij} R_i C_j \ln (R_i C_j):=g^{*}(u,v).$$ In our setup, it will be hard in general to solve for the $R_i,C_j$, so that we need to study $\frac{\partial g^{*}}{\partial u},\frac{\partial g^{*}}{\partial v}$ using implicit differentiation. The following lemma will be very useful. \[lem:maxentropy\] For $g^{*}=\max_{\mathbf{Z}} g(\mathbf{Z})$, we have $$\frac{\partial g^{*}}{\partial u}= -\sum_{i} \ln(R_i) \frac{\partial \alpha_i}{\partial u} \mbox{\ and \ } \frac{\partial g^{*}}{\partial v}= - \sum_{j} \ln(C_j) \frac{\partial \beta_j}{\partial v}.$$ In particular, $$\frac{\partial (f+g^{*})}{\partial u}= \frac{\partial f}{\partial u}-\sum_{i} \ln(R_i) \frac{\partial \alpha_i}{\partial u} \mbox{\ and \ } \frac{\partial (f+g^{*})}{\partial v}= \frac{\partial f}{\partial v}- \sum_{j} \ln(C_j) \frac{\partial \beta_j}{\partial v}.$$ where the $R_i,C_j$ are positive solutions of . The proof of Lemma \[lem:maxentropy\] is given in Section \[sec:maxentropyproof\]. The Logarithm of the First Moment - Maximum {#sec:logfirstmax} ------------------------------------------- In this section, we prove Lemma \[lem:firstmomentmax\]. As observed in Sections \[sec:generallog\] and  \[sec:maxentropy\], it suffices to study the maxima of $\phi_1(\alpha,\beta)$. Lemma \[lem:firstmomentmax\] is immediately implied by the following lemmas. \[lem:fircritical\] The only critical points of $\phi_1(\alpha,\beta)$ are: 1. $(\alpha,\beta)=(p^\pm,p^\mp)$ and $(\alpha,\beta)=(p^*,p^*)$, when $B_1,B_2,\lambda$ lie in the non-uniqueness region. 2. $(\alpha,\beta)=(p^*,p^*)$, when $B_1,B_2,\lambda$ lie in the uniqueness region. \[lem:firboundary\] The function $\phi_1(\alpha,\beta)$ does not have any local maximum on the boundary of the region $$\big\{(\alpha,\beta)\big|0\leq \alpha\leq 1,\ 0\leq\beta\leq 1\big\}.$$ \[lem:firhessian\] Let $\mathbf{X}^{*}=\arg\max_{\mathbf{X}}\Phi_1(p^+,p^-,\mathbf{X})$ and $\mathbf{X}^{o}=\arg\max_{\mathbf{X}}\Phi_1(p^*,p^*,\mathbf{X})$. 1. When $B_1,B_2,\lambda$ lie in the non-uniqueness region, the function $\Phi_1(\alpha,\beta,\mathbf{X})$ has a local maximum at the point $(p^+,p^-, \mathbf{X}^*)$ and a saddle point at $(p^*,p^*, \mathbf{X}^o)$. 2. When $B_1,B_2,\lambda$ lie in the uniqueness region, the function $\Phi_1(\alpha,\beta,\mathbf{X})$ has a local maximum at $(p^*,p^*, \mathbf{X}^o)$. We defer the proofs of Lemmas \[lem:firboundary\] and \[lem:firhessian\] to Section \[sec:logfirstmaxproof\]. We give the more interesting proof of Lemma \[lem:fircritical\]. We apply Lemma \[lem:maxentropy\] with $$\begin{gathered} \mathbf{Z}\leftarrow \mathbf{X},\ u\leftarrow\alpha,\ v\leftarrow\beta,\ \alpha_1\leftarrow\alpha,\ \alpha_2\leftarrow1-\alpha \ \beta_1\leftarrow\beta,\ \beta_2\leftarrow1-\beta,\\ f\leftarrow(\alpha+\beta)\ln\lambda+(\Delta-1)f_1,\ g\leftarrow \Delta g_1.\end{gathered}$$ The matrix $M$ is given by $$M= \left(\begin{array}{cc} B_1 & 1 \\ 1 & B_2 \\ \end{array}\right).$$ Lemma \[lem:maxentropy\] gives that the critical points of $\max_\mathbf{X}\Phi_1 (\alpha,\beta,\mathbf{X})$ must satisfy $$\begin{aligned} \frac{\partial \phi_1 }{\partial\alpha}&=\ln\left(\lambda \left(\frac{\alpha}{1-\alpha}\right)^{\Delta-1} \left(\frac{R_2}{R_1}\right)^\Delta \right)=0,\label{ea1}\\ \frac{\partial \phi_1}{\partial\beta}&=\ln\left(\lambda \left(\frac{\beta}{1-\beta}\right)^{\Delta-1} \left(\frac{C_2}{C_1}\right)^\Delta \right)=0,\label{ea2}\end{aligned}$$ where $$\begin{array}{lll} R_1 (B_1 C_1 + C_2)= \alpha,& & C_1 (B_1 R_1 + R_2)= \beta,\\ R_2 (C_1 + B_2 C_2)= 1-\alpha,& & C_2 (R_1 + B_2 R_2)=1-\beta. \end{array}\label{eb}$$ For $r=R_1/R_2$ and $c=C_1/C_2$ we have $$r \frac{B_1 c + 1}{c + B_2} = \frac{\alpha}{1-\alpha} = r^{1+1/(\Delta-1)} (1/\lambda)^{1/(\Delta-1)}\Longrightarrow r=\lambda \left(\frac{B_1 c + 1}{c + B_2}\right)^{\Delta-1},$$ $$c \frac{B_1 r + 1}{r + B_2} = \frac{\beta}{1-\beta} = c^{1+1/(\Delta-1)} (1/\lambda)^{1/(\Delta-1)}\Longrightarrow c=\lambda \left( \frac{B_1 r + 1}{r + B_2}\right)^{\Delta-1}.$$ The right hand side equations are exactly the equations and in light of Lemma \[lem:densitiesone\] it must be the case that in the non-uniqueness region either that $r=\frac{q^\pm}{1-q^{\pm}}, c=\frac{q^\mp}{1-q^{\mp}}$ or $r=c=\frac{q^{*}}{1-q^{*}}$. By , we hence obtain that the critical points for the first moment in the non-uniqueness region are given by $(\alpha,\beta)=(p^\pm,p^\mp)$ or $(\alpha,\beta)=(p^{*},p^{*})$. Similarly for the uniqueness region. The Logarithm of the Second Moment - Maximum {#sec:logsecmax} -------------------------------------------- In this section, we prove Lemma \[lem:secondmax2spin\]. As established in Sections \[sec:generallog\] and  \[sec:maxentropy\], it suffices to study the maxima of $\phi_2(\gamma,\delta)$. Lemma \[lem:secondmax2spin\] is immediately implied by the following lemmas. \[lem:seccritical\] Under the hypotheses of Theorem \[thm:2spin\], for $(\alpha,\beta)=(p^\pm,p^\mp)$, the only critical points of $\phi_2(\gamma,\delta)$ satisfy $\gamma=\alpha^2,\ \delta=\beta^2$. \[lem:secboundary\] For $(\alpha,\beta)=(p^\pm,p^\mp)$, the function $\phi_2(\gamma,\delta)$ does not have any local maximum on the boundary of the region $$\big\{(\gamma,\delta)\big|\ \gamma\geq0,\ \alpha-\gamma\geq0,\ 1-2\alpha+\gamma\geq 0,\ \delta\geq 0,\ \beta-\delta\geq0,\ 1-2\beta+\delta\geq0\big\}.$$ \[lem:sechessian\] For $(\alpha,\beta)=(p^+,p^-)$, let $\mathbf{Y}^{*}=\arg\max_{\mathbf{Y}}\Phi_2(\alpha^2,\beta^2,\mathbf{Y})$. The function $\Phi_2(\gamma,\delta,\mathbf{Y})$ has a local maximum at $(\alpha^2,\beta^2,\mathbf{Y}^{*})$. The proofs of Lemmas \[lem:secboundary\] and \[lem:sechessian\] are valid for all settings of $B_1,B_2$, so that to prove Lemmas \[lem:secondmaxising\] and \[lem:secondmaxhardcore\], one only needs to obtain the analogues of Lemma \[lem:seccritical\] in the respective settings of $B_1,B_2$. The proofs of these analogs are given in Sections \[sec:secondmaxisingproof\] and \[sec:hardhard\], see also Section \[sec:remarks\]. We defer the proofs of Lemmas \[lem:secboundary\] and \[lem:sechessian\] to Section \[sec:logsecmaxproof\] and give here the more interesting proof of Lemma \[lem:seccritical\]. We apply once more Lemma \[lem:maxentropy\] with $$\begin{gathered} \mathbf{Z}\leftarrow \mathbf{Y},\ u\leftarrow \gamma,\ v\leftarrow \delta,\ \alpha_1\leftarrow\gamma,\ \alpha_2=\alpha_3\leftarrow\alpha-\gamma,\ \alpha_4\leftarrow1-2\alpha+\gamma\\ \beta_1\leftarrow\gamma,\ \beta_2=\beta_3\leftarrow\beta-\delta,\ \beta_4\leftarrow1-2\beta+\delta,\ f\leftarrow (\Delta-1)f_2,\ g\leftarrow \Delta g_2.\end{gathered}$$ The matrix $M$ is given by $$M=\left(\begin{array}{cccc} B_1^2 & B_1 & B_1 & 1 \\ B_1 & B_1 B_2 & 1 & B_2 \\ B_1 & 1 & B_1 B_2 & B_2 \\ 1 & B_2 & B_2 & B_2^2 \\ \end{array}\right).$$ We obtain that the critical points of $\max_{\mathbf{Y}}\Phi_2(\gamma,\delta,\mathbf{Y})$ must satisfy $$\label{es1} \frac{\partial \phi_2}{\partial \gamma} = \ln\left( \left(\frac{\gamma(1-2\alpha+\gamma)}{(\alpha-\gamma)^2}\right)^{\Delta-1}\left(\frac{R_2 R_3}{R_1 R_4}\right)^\Delta \right)=0.$$ $$\label{es2} \frac{\partial \phi_2}{\partial \delta} = \ln\left( \left(\frac{\delta(1-2\beta+\delta)}{(\beta-\delta)^2}\right)^{\Delta-1}\left(\frac{C_2 C_3}{C_1 C_4}\right)^\Delta \right)=0.$$ where the $R_i,C_j$ satisfy the following equations (symbol $\circ$ denotes the Hadamard product). $$\label{et} \left(\begin{array}{c}R_1\\R_2\\R_3\\R_4\end{array}\right)\circ M\left(\begin{array}{c}C_1\\C_2\\C_3\\C_4\end{array}\right) =\left(\begin{array}{c}\gamma \\ \alpha-\gamma \\\alpha-\gamma \\1-2\alpha+\gamma\end{array}\right) \quad\quad \left(\begin{array}{c} C_1\\C_2\\C_3\\C_4\end{array}\right)\circ M \left(\begin{array}{c}R_1\\R_2\\R_3\\R_4\end{array}\right) =\left(\begin{array}{c}\delta \\\beta-\delta \\\beta-\delta \\1-2\beta+\delta\end{array}\right).$$ We will write out an explicit form of the equations after establishing the following claim. \[claim:equalityrc\] It holds that $R_2=R_3$ and $C_2=C_3$. Observe that the matrix $M$ remains invariant upon interchanging its 2nd,3rd rows and 2nd,3rd columns, i.e., $$(R_1,R_2,R_3,R_4,C_1,C_2,C_3,C_4) \mbox{ satisfy } \eqref{et}\Longleftrightarrow (R_1,R_3,R_2,R_4,C_1,C_3,C_2,C_4) \mbox{ satisfy } \eqref{et}.$$ Since $g_2(\mathbf{Y})$ is strictly concave, this yields the claim. The rest of the proof will work as follows: as in the first moment, we first do some manipulations with the equations so that we arrive to a nice form. This form coincides with an inequality version of the tree equations for the ferromagnetic Ising model without external field, which can be analyzed easily. Let us begin by using Claim \[claim:equalityrc\], to write the equalities as $$\begin{aligned} &R_1 (B^2_1 C_1+2B_1 C_2+C_4)=\gamma, &\ \ &C_1 (B^2_1R_1+2B_1R_2+R_4)= \delta,\notag\\ &R_2 (B_1 C_1+(B_1B_2+1)C_2+B_2C_4)=\alpha-\gamma, &\ \ &C_2 (B_1R_1+(B_1B_2+1)R_2+B_2R_4)= \beta-\delta,\label{eq:frty}\\ &R_4 (C_1+2B_2 C_2+B^2_2 C_4)=1-2\alpha+\gamma, &\ \ &C_4(R_1+2B_2R_2+B^2_2R_4)= 1-2\beta+\delta.\notag \end{aligned}$$ $$\begin{gathered} \frac{\partial \phi_2}{\partial \gamma} = \ln\left( \left(\frac{\gamma(1-2\alpha+\gamma)}{(\alpha-\gamma)^2}\right)^{\Delta-1}\left(\frac{R_2^2}{R_1 R_4}\right)^\Delta \right)=0,\label{eq:dergamma}\\ \frac{\partial \phi_2}{\partial \delta} = \ln\left( \left(\frac{\delta(1-2\beta+\delta)}{(\beta-\delta)^2}\right)^{\Delta-1}\left(\frac{C_2^2}{C_1 C_4}\right)^\Delta \right)=0.\label{eq:derdelta}\end{gathered}$$ We now set $r_1=R_1/R_2, r_4=R_4/R_2, c_1=C_1/C_2, c_4=C_4/C_2$. After dividing the appropriate pairs of equations in , we obtain $$\begin{aligned} r_1=\frac{\gamma}{\alpha-\gamma}\cdot\frac{B_1c_1+(B_1B_2+1)+B_2c_4}{B^2_1c_1+2B_1+c_4},&\quad r_4 =\frac{1-2\alpha+\gamma}{\alpha-\gamma}\cdot\frac{B_1c_1+(B_1B_2+1)+B_2c_4}{c_1+2B_2+B^2_2c_4},\label{r1r4}\\ c_1 = \frac{\delta}{\beta-\delta}\cdot\frac{B_1r_1+(B_1B_2+1)+B_2r_4}{B^2_1r_1+2B_1+r_4},& \quad c_4 = \frac{1-2\beta+\delta}{\beta-\delta}\cdot\frac{B_1r_1+(B_1B_2+1)+B_2r_4}{r_1+2B_2+B^2_2r_4}.\label{c1c4}\end{aligned}$$ Equations and become $$\begin{aligned} \left(r_1r_4\right)^\Delta=\left(\frac{\gamma(1-2\alpha+\gamma)}{(\alpha-\gamma)^2}\right)^{\Delta-1},\quad \left(c_1c_4\right)^\Delta=\left(\frac{\delta(1-2\beta+\delta)}{(\beta-\delta)^2}\right)^{\Delta-1}.\label{secab}\end{aligned}$$ Using the identity $$(B^2_1r_1+2B_1+r_4)(r_1+2B_2+B^2_2r_4)=(B_1r_1+(B_1B_2+1)+B_2r_4)^2+(1-B_1B_2)^2(r_1r_4-1).$$ For convenience, we now work with $d=\Delta-1$. Multiplying the equations in , we obtain $$\begin{aligned} r_1r_4&=\frac{\gamma(1-2\alpha+\gamma)}{(\alpha-\gamma)^2}\cdot\frac{(B_1c_1+(B_1B_2+1)+B_2c_4)^2}{(B_1c_1+(B_1B_2+1)+B_2c_4)^2+(1-B_1B_2)^2(c_1c_4-1)}.\label{eq:prodr}\end{aligned}$$ It will be convenient at this point to work with $d=\Delta-1$. We plug the first equation in into , yielding $$(r_1r_4)^{1/d}-1=\frac{(1-B_1B_2)^2(c_1c_4-1)}{(B_1c_1+(B_1B_2+1)+B_2c_4)^2},\ (c_1c_4)^{1/d}-1=\frac{(1-B_1B_2)^2(r_1r_4-1)}{(B_1r_1+(B_1B_2+1)+B_2r_4)^2}.\label{recurone}$$ The second equality in follows by a completely symmetric argument for $c_1,c_4$. Observe that implies that one of the following three cases can hold. $$\mbox{\sc{Case I}}: r_1 r_4=c_1c_4=1, \quad \mbox{\sc{Case II}}: r_1 r_4>1,c_1c_4>1, \quad \mbox{\sc{Case III}}: r_1 r_4<1,c_1c_4<1.$$ reduces to $\gamma=\alpha^2, \ \delta=\beta^2$ by . Thus we may focus on and . We further restrict our attention to , with being completely analogous. Use the AM-GM inequalities $B_1c_1+B_2c_4\geq 2\sqrt{B_1B_2c_1 c_4}, \ B_1r_1+B_2r_4\geq 2\sqrt{B_1B_2r_1 r_4}$ to obtain $$(r_1r_4)^{1/d}-1\leq\frac{(1-B_1B_2)^2(c_1c_4-1)}{\big(B_1B_2+1+2\sqrt{B_1B_2c_1c_4}\big)^2}, \quad (c_1c_4)^{1/d}-1\leq\frac{(1-B_1B_2)^2(r_1r_4-1)}{\big(B_1B_2+1+2\sqrt{B_1B_2r_1r_4}\big)^2}.$$ It is straightforward to verify the identity $$\Big(B_1B_2+1+2z\sqrt{B_1B_2}\Big)^2+(1-B_1B_2)^2(z^2-1)=\Big((1+B_1B_2)z+2\sqrt{B_1 B_2}\Big)^2.$$ Set $x=\sqrt{r_1 r_4}$, $y=\sqrt{c_1c_4}$. Using the identity with $z=x,y$ and taking square roots, we obtain $$x^{1/d}\leq\frac{(1+B_1B_2)y+2\sqrt{B_1 B_2}}{2\sqrt{B_1B_2}y+B_1B_2+1}, \quad y^{1/d}\leq\frac{(1+B_1B_2)x+2\sqrt{B_1 B_2}}{2\sqrt{B_1B_2}x+B_1B_2+1}.\label{ineq:recura}$$ Using the substitution $B'=(1+B_1 B_2)/2\sqrt{B_1B_2}$, this can be rewritten in the form $$x^{1/d}\leq\frac{B'y+1}{y+B'}, \quad y^{1/d}\leq\frac{B'x+1}{x+B'}.\label{ineq:recurb}$$ The astute reader will immediately realize the analogy of with the tree equations for the Ising model without external field. Moreover, since $B'\geq 1$ we are in the case of the [*ferromagnetic*]{} Ising model. For the ferromagnetic Ising model on the infinite tree ${\mathbb{T}}_{d+1}$, we have uniqueness when $1<B\leq \frac{d+1}{d-1}$. Hence intuitively the lemma holds when $B'\leq \frac{d+1}{d-1}$. However, in we have an inequality version of the tree recursions and therefore a bit more work is required. In our setting, we obtain that cannot hold when $x>1,y>1$. To see this, multiply the inequalities in to obtain $$\label{eq:stablea} x^{1/d}y^{1/d}\leq \frac{B'x+1}{x+B'}\cdot \frac{B'y+1}{y+B'}.$$ Note that is reversed in . The following lemma implies that cannot hold in , and similarly for the reverse inequality in . \[lem:maininequality\] Let $d\geq 2$. When $B'\leq 1/B_c({\mathbb{T}}_{d+1})=\frac{d+1}{d-1}$, for $z>1$ it holds that $z^{1/d}> \displaystyle \frac{B'z+1}{z+B'}$. The inequality is reversed for $z<1$. We give the proof of Lemma \[lem:maininequality\] after observing that $B'\leq \frac{d+1}{d-1}$ reduces to $\sqrt{B_1 B_2}\geq \frac{\sqrt{d}-1}{\sqrt{d}+1}$. This concludes the proof of Lemma \[lem:seccritical\]. The case $z<1$ reduces to the case $z>1$ using the inversion $z\leftarrow 1/z$. Thus we focus on proving the Lemma for $z>1$. Let $f(w):=w^{d+1}-B' w^d+B' w-1$. The lemma reduces to proving $f(w)>0$ when $w>1$ (under the substitution $w=z^{1/d}$). We have $$\begin{aligned} f'(w)&=(d+1)w^d-dB'w^{d-1}+B'.\\ f''(w)&=dw^{d-2}\big((d+1)w-(d-1)B'\big).\end{aligned}$$ The assumptions imply that $(d+1)-(d-1)B'\geq 0$. Hence, when $w>1$, we have $f''(w)>f''(1)$. Note that $f''(1)\geq 0$ only if $(d+1)-(d-1)B'\geq 0$. Thus, $f''(w)$ is strictly positive when $w>1$ and hence $f'(w)$ is strictly increasing. It follows that for $w>1$, we have $f'(w)> f'(1)$ and $$f'(1)=(d+1)-(d-1)B'\geq0.$$ Thus, $f$ is strictly increasing for $w>1$, yielding $f(w)>f(1)=0$. Some Remarks {#sec:remarks} ------------ Here we give the idea behind the arguments establishing Lemmas \[lem:secondmaxising\] and \[lem:secondmaxhardcore\]. As observed in Section \[sec:logsecmax\], these lemmas boil down to proving that, for $(\alpha,\beta)=(p^+,p^-)$, the only critical point of $\phi_2(\gamma,\delta)$ is at $(\gamma,\delta)=(\alpha^2,\beta^2)$ under the respective hypotheses. Our point of departure is once again the equations in , albeit we proceed with a more thorough analysis. Looking at the argument for the case $\frac{\sqrt{d}-1}{\sqrt{d}+1}\leq \sqrt{B_1B_2}$, the point of the proof which is subject to tighter analysis is the use of the AM-GM inequalities. These are weak if the ratios $r_1/r_4,c_1/c_4$ are much bigger than $B_2/B_1$. In Section \[sec:secondmaxisingproof\], we turn this weakness into our favor. Namely, for the antiferromagnetic Ising model without external field, in the region $0<B<\frac{\sqrt{d}-1}{\sqrt{d}+1}$, the densities $p^\pm$ are heavily biased towards 1 and 0 respectively, and this reflects at the values of $r_1,r_4,c_1,c_4$. This observation can be turned into an elementary proof of Lemma \[lem:secondmaxising\]. For the proof of Lemma \[lem:secondmaxhardcore\], it is easier to get an explicit handle on the values $p^\pm$ (and thus on the bias of the $r_i,c_j$), which can be used to analyze the equations (see Section \[sec:hardhard\]). Further, let us make some small minor observations. The reader may have noticed that the proof of Lemma \[lem:seccritical\] applies in a more general setup than the one stated in Condition \[cond:maxima\]. Namely, in the region $\sqrt{B_1B_2}\geq \frac{\sqrt{d}-1}{\sqrt{d}+1}$, it verifies Condition \[cond:maxima\] for all values of $\alpha,\beta$ and not just for $(\alpha,\beta)=(p^\pm,p^\mp)$. The proof also extends in the uniqueness region of the antiferromagnetic Ising model (even with external field). While we do not draw this path here, since we are only interested in the values of $(\alpha,\beta)$ which maximize ${\mathbb{E}}_{\mathcal{G}}[Z^{\alpha,\beta}_G]$, one can readily determine explicitly a broader range of $\alpha,\beta$ for which Lemma \[lem:smallgraph\] continues to hold. NP-Hardness Results {#sec:NP-outline} =================== In this section, we give the proofs for our NP-hardness results. We will prove the following theorem, which allows us to prove Theorems \[thm:hardcore\], \[thm:Ising\], \[thm:2spin\]. \[thm:mainNP\] For a given $\Delta$, assume that $(B_1,B_2,\lambda)$ lie in the non-uniqueness region. Moreover, assume that Condition \[cond:maxima\] holds. Then, unless NP$=$ RP, there is no FPRAS for the partition function of the 2-spin model with parameters $B_1,B_2,\lambda$ in graphs with maximum degree $\Delta$. With Theorem \[thm:mainNP\] at hand, it is straightforward to prove Theorems \[thm:hardcore\], \[thm:Ising\], \[thm:2spin\]. Theorem \[thm:2spin\] follows immediately from Lemma \[lem:secondmax2spin\] and Theorem \[thm:mainNP\]. Theorem  \[thm:hardcore\] was known to hold for $\Delta=3$ and $\Delta\geq 6$ ([@Sly10], [@Galanis]). Lemma \[lem:secondmaxhardcore\] in combination with Theorem \[thm:mainNP\] prove the cases $\Delta=4,5$ as well. We now prove Theorem \[thm:Ising\]. In the case of Ising model with no external field, note that Lemma \[lem:secondmax2spin\] and Theorem \[thm:mainNP\] give hardness for $\Delta\geq 3$ when $$\frac{\sqrt{d}-1}{\sqrt{d}+1}\leq B<\frac{d-1}{d+1} \mbox{ where } d=\Delta-1.\label{eq:intervals}$$ The theorem will be proved once we show that this can be extended to the region $0<B<\frac{\sqrt{d}-1}{\sqrt{d}+1}$. Observe that hardness for $B,\Delta$ gives hardness for $B,\Delta+1$. It is easy to check that for $d\geq 2$, it holds that $$\frac{\sqrt{d+1}-1}{\sqrt{d+1}+1}<\frac{d-1}{d+1}<\frac{d}{d+2},$$ so that the intervals corresponding to $\Delta$ and $\Delta+1$ are overlapping for $\Delta\geq 3$. Thus it suffices to check NP hardness for $d=2$ and $0<B<\frac{\sqrt{d}-1}{\sqrt{d}+1}$. This follows from Lemma \[lem:secondmaxising\] and Theorem \[thm:mainNP\], thus completing the proof. We thus focus on proving Theorem \[thm:mainNP\]. Sly’s reduction [@Sly10], while presented for the particular case of the hard-core model, is general enough to cover general 2-spin models. The only part in Sly’s reduction which needs modifications is in establishing the properties of the gadget he uses. Let us first describe the gadget used in [@Sly10]. The construction of the gadget has two parameters $0<\theta,\psi<1/8$. Let $k:=(\Delta-1)^{\lfloor \theta \log_{\Delta-1}n\rfloor}$ and $\ell:=2\lfloor \frac{\psi}{2}\log_{\Delta-1}n\rfloor$ and let $m':=k(\Delta-1)^{\ell}$. Note that $m'=o(n^{1/4})$. The gadget is constructed in two steps: first, a random bipartite graph $\overline{G}$ with $n+m'$ vertices on each side is constructed. The two sides of the graph will be labelled with $+,-$. For $s\in\{+,-\}$, let the vertices on the $s$-side be $W_s\cup U_s$, where $|W_s|=n$ and $|U_s|=m'$. The edges of the graph are the union of $\Delta-1$ uniformly random perfect matchings between $W_+\cup U_+$ and $W_-\cup U_-$, together with a uniformly random perfect matching between $W_+$ and $W_-$. Thus, in the resulting random graph $\overline{G}$ , the vertices in $W:=W_+\cup W_-$ have degree $\Delta$ and the vertices in $U:=U_+\cup U_-$ have degree $\Delta-1$. We denote the graph distribution defined by this construction as $\overline{\mathcal{G}}$. The second part of the construction appends $k$ disjoint $(\Delta-1)$-ary trees of (even) depth $\ell$ to $\overline{G}$. Specifically, the leaves of each tree form a subset of one of $U_+$, $U_-$ and the leaves of the trees form a partition of the set $U_+\cup U_-$ (note that this is possible since $|U_+|=|U_-|=k (\Delta-1)^\ell$). The trees have otherwise no other common vertices with the graph $\overline{G}$. Denote by $R_+$ (resp. $R_-$) the roots of those trees whose leaves are a subset of $U_+$ (resp. $U_-$). Let $R:=R_+\cup R_-$. Analogously to [@Sly10], for a configuration $\sigma$ on $H$, the phase $Y(\sigma)$ of $\sigma$ is defined to be $+$ if the number of vertices assigned $-1$ in $W_+$ is greater than the number of vertices assigned $-1$ in $W_-$, otherwise $Y(\sigma)=-$. In other words, the phase $Y(\sigma)$ of a configuration $\sigma$ simply points to the set between $W_+,W_-$ with the greatest number of vertices assigned $-1$ in $\sigma$. The properties of Sly’s gadget are stated in the following lemma. We use the following notation: for a configuration $\sigma$ on $H$, $\sigma_R$ denotes the restriction of $\sigma$ to vertices in $R$. Moreover, for a spin $s\in\{-1,+1\}$, $\sigma^{-1}(s)$ denotes the vertices assigned the spin $s$. \[lem:gadget2\] For a given $\Delta$, assume that $(B_1,B_2,\lambda)$ lie in the non-uniqueness region of the infinite tree ${\mathbb{T}_{\Delta}}$. Moreover, assume that Condition \[cond:maxima\] holds. There exist constants $\theta(\Delta,B_1,B_2,\lambda),\psi(\Delta,B_1,B_2,\lambda)>0$ such that for all sufficiently large $n$ the graph $H$ satisfies with probability $1-o_n(1)$ the following: 1. The two phases occur with roughly equal probability, i.e., for $i\in\{+,-\}$, we have \[it:phaseprob2\] $$\mu_H\big(Y(\sigma)=i\big)\geq \frac{1}{n}.$$ 2. Conditioned on the phase $i\in\{+,-\}$, the spins of vertices in $R$ are approximately independent, that is, \[it:approxind2\] $$\max_{\eta\in \{-1,+1\}^{R}}\Big|\frac{\mu_H\big(\sigma_R=\eta\, |\, Y(\sigma)=i\big)}{Q_R^{i}(\eta)}-1\Big|\leq n^{-2\theta},$$ where $Q^{i}_R$ is the following product distribution on configurations $\eta: R_+\cup R_-\rightarrow \{\pm 1\}$: $$Q^{i}_R(\eta)= (q^i)^{ |\eta^{-1}(-1)\cap R_+| } (1-q^i)^{ |\eta^{-1}(+1)\cap R_+| } (q^{-i})^{ |\eta^{-1}(-1)\cap R_-| } (1-q^{-i})^{ |\eta^{-1}(+1)\cap R_-| }.$$ Recall that for $i\in\{+,-\}$, $q^i$ is the probability that the root of the infinite $(\Delta-1)$-ary tree ${\hat{\mathbb{T}}_{\Delta}}$ is assigned the spin $-1$ in the extremal measure $\hat{\mu}^{j}$ (see Section \[sec:tree-recursions\]). We now describe briefly how the required property of $H$ is established in [@Sly10]. This is done in two steps: 1. First prove that in the graph $\overline{G}$, conditioned on the phase $i\in \{+,-\}$, the spins of the vertices in $U$ behave in expectation as a product distribution $Q^i_U$ on $\{-1,+1\}^{U}$. The latter distribution is obtained by replacing $R$ with $U$ in the definition of $Q^i_R$ in Lemma \[lem:gadget2\]. Roughly, this can be obtained as follows from what we have seen so far: in Section \[sec:proof-approach\] we showed that for a random $\Delta$-regular bipartite graph, with high probability over the choice of the graph, in the Gibbs distribution the total weight of configurations from the set $\Sigma^{p^+,p^-}$ dominate exponentially over the rest. An analogous phenomenon occurs for a graph $\overline{G}\sim \overline{{\mathcal{G}}}$, which allows to quantify the aforementioned behavior, see Section \[sec:modifiedaaa\] for more details. 2. Translate the results to $R$. This is accomplished in a crafty way in [@Sly10]: the product measures on $\{-1,+1\}^{U}$ are analogous to the even and odd boundaries in the infinite $(\Delta-1)$-ary tree, providing for an elegant translation of the results. Moreover, the extremality of the measures implies that the roots of the trees, conditioned on the phase, are strongly concentrated around their expected value. The latter part follows after studying non-reconstruction in the extremal measures and proving that the correlation of the spins in $R$ to the spins in $U$ decays doubly exponentially in $\ell$, see Section \[sec:reco\]. In Section \[sec:modifiedaaa\], we give the lemmas in [@Sly10] which require adjustment in the 2-spin model. Their proofs follow the same way as in [@Sly10], by comparing the quantities in the graph distribution $\overline{\mathcal{G}}$ with those in the distribution ${\mathcal{G}}$. In Section \[sec:reco\], we describe the non-reconstruction results that Sly uses. Finally, in Section \[sec:sketchproof\], we give the proof of Lemma \[lem:gadget2\] and conclude Theorem \[thm:mainNP\], checking that Sly’s arguments go through in our setting as well. The Partition Function for Graphs Sampled from $\overline{\mathcal{G}}$ {#sec:modifiedaaa} ----------------------------------------------------------------------- In this section, we give an outline of the lemmas which are needed to extend Sly’s reduction for general 2-spin models. Let $\eta$ be an assignment of spins to $U$, i.e., $\eta\in\{-1,+1\}^U$. For a graph $\overline{G}\sim\overline{{\mathcal{G}}}$, let $Z^{\alpha,\beta}_{\overline{G}}(\eta)$ be the partition function over configurations which agree with $\eta$ on $U$, i.e., $Z^{\alpha,\beta}_{\overline{G}}(\eta)$ is the total weight of configurations in $\sigma\in\{-1,+1\}^{V(\overline{G})}$ subject to $|N_-(\sigma)\cap W_+|=\alpha n$, $|N_-(\sigma)\cap W_-|=\beta n$ and $\sigma_U=\eta$. Denote by $\eta_1^+$ the number of vertices in $U_+$ which are assigned spin $+1$ and let $\eta_1^- = m'-\eta_i^+$ (the number of vertices in $U_+$ that are assigned spin $-1$). Define similarly $\eta_2^+,\eta_2^-$ for $U_-$. We have the following analogue of [@Sly10 Lemma 3.1]. \[lem:modfirst\] For any $\alpha,\beta$ and all $\eta\in\{-1,+1\}^U$ we have $$\frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}[Z^{\alpha,\beta}_{\overline{G}}(\eta)]}{{\mathbb{E}}_{\mathcal{G}}[Z^{\alpha,\beta}_{G}]}= (1+o(1)) \left(\frac{\alpha^{\eta_1^-}(1-\alpha)^{\eta_1^+}\beta^{\eta_2^-}(1-\beta)^{\eta_2^+}}{(\alpha-x^*)^{\eta_1^-} (1-\alpha-\beta+x^*)^{\eta_1^+-\eta_2^-} (\beta-x^*)^{\eta_2^-}}B_2^{\eta_1^+ - \eta_2^-} \right)^{\Delta-1}\lambda^{\eta_1^-+\eta_2^-},$$ where $x^*$ is the (unique) non-negative solution of the equation $B_1B_2(\alpha-x)(\beta-x)=x(1-\alpha-\beta+x)$. In particular, when $(\alpha,\beta)=(p^+,p^-)$, it holds that $$\label{eq:productmeasure} \frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}[Z^{\alpha,\beta}_{\overline{G}}(\eta)]}{{\mathbb{E}}_{\mathcal{G}}[Z^{\alpha,\beta}_{G}]}= (1+o(1))\frac{C^*}{(1-q^+)^{m'}(1-q^-)^{m'}}(q^+)^{\eta_1^-}(1-q^+)^{\eta_1^+}(q^-)^{\eta_2^-}(1-q^-)^{\eta_2^+},$$ where $$\label{eq:cstar} C^* = \left(\frac{(B_2+Q^+)(B_2+Q^-)}{B_2+Q^++Q^-+B_1Q^+Q^-}\right)^{(\Delta-1)m'}.$$ The proof of Lemma \[lem:modfirst\] is given in Section \[sec:modfirstproof\]. Seeking high probability lower bounds on $Z^{\alpha,\beta}_{\overline{G}}(\eta)$, we study the second moment of $Z^{\alpha,\beta}_{\overline{G}}(\eta)$. \[lem:modsecond\] When Condition \[cond:maxima\] is true, for $(\alpha,\beta)=(p^+,p^-)$ and all $\eta\in\{-1,+1\}^U$ we have $$\frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[\big(Z^{\alpha,\beta}_{\overline{G}}(\eta)\big)^2\big]}{{\mathbb{E}}_{{\mathcal{G}}}\big[\big(Z^{\alpha,\beta}_{G}\big)^2\big]}= (1+o(1)) (C^*)^2 \left(\lambda\left(\frac{1+B_1Q^-}{B_2+Q^-}\right)^{\Delta-1}\right)^{2\eta_1^-} \left(\lambda\left(\frac{1+B_1Q^+}{B_2+Q^+}\right)^{\Delta-1}\right)^{2\eta_2^-},\label{eq:secondmodratio}$$ where $Q^-,Q^+$ are given by  and $C^*$ is given by . The proof of Lemma \[lem:modsecond\] is given in Section \[lem:modsecondproof\]. Once again, it turns out that the second moment method fails to give high probability bounds, as is captured by the following lemma.  \[lem:rfsgadget\] When Condition \[cond:maxima\] is true, for $(\alpha,\beta)=(p^+,p^-)$ it holds that $$\lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[\big(Z^{\alpha,\beta}_{\overline{G}}(\eta)\big)^2\big]}{\left({\mathbb{E}}_{\overline{{\mathcal{G}}}}[Z^{\alpha,\beta}_{\overline{G}}(\eta)]\right)^2}=\big(1-\omega^2\big)^{-(\Delta-1)/2}\big(1-(\Delta-1)^2\omega^2\big)^{-1/2},$$ where $\omega$ is given by . The proof of Lemma \[lem:rfsgadget\] is given in Section \[sec:rfsgadgetproof\]. We then apply the small graph conditioning method to obtain the following analogue of [@Sly10 Lemma 3.9].  \[lem:smallgraphgadget\] For $\Delta\geq 3$, suppose that for $B_1,B_2, \lambda$ in the non-uniqueness region, Condition \[cond:maxima\] is true. Then, for $(\alpha,\beta)=(p^+,p^-)$, asymptotically almost surely over the choice of the graph $\overline{G}\sim\overline{{\mathcal{G}}}$, for $\eta\in \{-1,+1\}^U$, it holds that $$Z^{\alpha,\beta}_{\overline{G}}(\eta)>\frac{1}{\sqrt{n}}{\mathbb{E}}_{\overline{{\mathcal{G}}}}[Z^{\alpha,\beta}_{\overline{G}}(\eta)].$$ The proof of Lemma \[lem:smallgraphgadget\] is given in Section \[sec:smallgraphregular\]. The Reconstruction Problem in the Extremal Measures {#sec:reco} --------------------------------------------------- The final ingredient we need is proving that a strong form of non-reconstruction holds in the extremal measures. Let us first introduce some notation. Denote by $\hat{T}_{\rho,\ell,\Delta}$ the subtree of ${\hat{\mathbb{T}}_{\Delta}}$ rooted at $\rho$ up to depth $\ell$. Let $S_{\rho,\ell}$ be the leaves of $\hat{T}_{\rho,\ell,\Delta}$. Denote by $X_{\rho,\ell,+}$ the marginal probability that the root $\rho$ is occupied in a configuration $\sigma$ generated by the following process: 1. Sample a configuration $\hat{\sigma}$ from the projection of the measure $\hat{\mu}^+$ on $\hat{T}_{\rho,\ell,\Delta}$, 2. Conditioning on the configuration $\hat{\sigma}_S$ on $S_{\rho,\ell}$, sample a configuration $\sigma$ from the Gibbs distribution on $\hat{T}_{\rho,\ell,\Delta}$. Note that the configuration of the vertices in $S_{\rho,\ell}$ is a random vector, so that $X_{\rho,\ell,+}$ is a random variable. Define similarly $X_{\rho,\ell,-}$. We need the following analogue of [@Sly10 Lemma 4.2]. \[lem:Sly42\] Assume that $\Delta\geq 3$ and $B_1,B_2,\lambda$ lie in the non-uniqueness region. There exist constants $\zeta_1(\Delta,B_1,B_2,\lambda), \zeta_2(\Delta,B_1,B_2,\lambda) > 0$, such that for all sufficiently large $\ell$, $$\mathbb{P}\big[\big\vert X_{\rho,\ell,\pm} - q^\pm\big\vert \geq \exp(-\zeta_1\ell)\big]\leq \exp(-\exp(\zeta_2\ell)).$$ We omit the proof of Lemma \[lem:Sly42\]: Martinelli et al. [@Martinelli Proof of Theorem 5.1] prove the required strong concentration for the variables $X_{\rho,\ell,\pm}$ for the Ising model with arbitrary external field (in a different form than the one stated here but completely equivalent). Their results can be translated to the general 2-spin model, after carrying out the reduction described in [@SST]. Finally, in the case of hard constraints Lemma \[lem:Sly42\] follows by the respective result in [@Sly10 Lemma 4.2] (slightly extended in [@Galanis Lemma 30]). Of course, one could also apply the methods in [@Martinelli] or [@Sly10] to directly get the result in general. Proof of Theorem \[thm:mainNP\] {#sec:sketchproof} ------------------------------- The reduction of [@Sly10 Section 2.2] goes through as is, once we establish the properties of the gadget stated in Lemma \[lem:gadget2\]. The proof of Lemma \[lem:gadget2\] is a straightforward extension of the proof of [@Sly10 Theorem 2.1], so we give a desciption of the main elements and point to the relevant lemmas which verify that Sly’s proof goes through in our setting as well. We first introduce some notation. Define $\Sigma^{\pm}$ to be the configurations $\sigma\in \{-1,+1\}^{V(\Go)}$ such that $|N_-(\sigma)\cap W_{\pm}|\geq |N_-(\sigma)\cap W_\mp|$, that is, a configuration in $\Sigma^{+}$ assigns the spin $-1$ to more vertices in $W_{+}$ than in $W_{-}$ (and conversely for $\Sigma^-$). Similarly, for $\eta\in\{-1,+1\}^U$, define $\Sigma^{\pm}(\eta)$ to be the configurations $\sigma$ in $\sigma\in \Sigma^{\pm}$ such that $\sigma_U=\eta$. Further, define $$Z^{\pm}_{\Go}(\eta)=\sum_{\sigma\in \Sigma^{\pm}(\eta)}w_G(\sigma), \quad Z^{\pm}_{\Go}:=\sum_{\sigma\in \Sigma^{\pm}}w_G(\sigma).$$ Note that $Z^{\pm}_{\Go}=\sum_{\eta\in\{-1,+1\}^U}Z^{\pm}_{\Go}(\eta)$. Using that $(\alpha,\beta)=(p^\pm,p^\mp)$ maximizes the logarithm of ${\mathbb{E}}_{\Gc}[Z^{\alpha,\beta}_G]$ (see Lemma \[lem:firstmomentmax\]), the first item of Lemma \[lem:modfirst\] together with Markov’s inequality, shows that the largest contribution to ${\mathbb{E}}_{\Gco}[Z^{\pm}_{\Go}(\eta)]$ is made by ${\mathbb{E}}_{\Gco}[Z^{p^{\pm},p^{\mp}}_{\Go}(\eta)]$, i.e., $$\label{eq:aazzss} {\mathbb{E}}_{\Gco}[Z^{\pm}_{\Go}(\eta)]=(1+o(1)){\mathbb{E}}_{\Gco}[Z^{p^{\pm},p^{\mp}}_{\Go}(\eta)]=(1+o(1))(C')^{m'}Q^{\pm}_U(\eta){\mathbb{E}}_{\Gc}[Z^{p^{\pm},p^{\mp}}_G],$$ where in the last equality we used the conclusion of Lemma \[lem:modfirst\] (the constant $C'$ can be inferred from , though it will not be important for what follows). The derivation of is identical to the proof of [@Sly10 Lemma 3.3]. Summing over $\eta\in \{-1,+1\}^{U}$ yields $$\label{eq:aazzssb} {\mathbb{E}}_{\Gco}[Z^{\pm}_{\Go}]=(1+o(1))(C')^{m'}Q^{\pm}_U(\eta){\mathbb{E}}_{\Gc}[Z^{p^{\pm},p^{\mp}}_G].$$ Our results from the small graph conditioning method (see Section \[sec:oversmall\]), combined with the above observations, yield that $Z^{\pm}_{\Go}(\eta)$ is close to its expectation, in particular $$\label{eq:wertyb} \lim_{n\rightarrow\infty}\max_{\eta\in\{-1,+1\}^U} \PrGco\left(Z^{\pm}_{\Go}(\eta)<\frac{1}{\sqrt{n}}{\mathbb{E}}_{\Gco}[Z^{\pm}_{\Go}(\eta)]\right)=0.$$ see the proof of [@Sly10 Theorem 3.10] for details. At this point, one needs to transfer these results to the graph $H$. Define $Z_H^{\pm}$ to be the contribution to the partition function of $H$ from configurations with phase $\pm$ (note that the restriction of such configurations on $\Go$ is a configuration in $\Sigma^{\pm}$). To connect the partition function of $H$ with the partition function of $\Go$, let $T$ be the graph induced by the edges $E(H)\backslash E(\Go)$. In particular, $T$ is a union of disjoint trees and the union of the leaves of the trees is precisely the set $U$. For $\eta\in\{-1,+1\}^U$, define $Z_T(\eta)$ to be the contribution to the partition function of $T$ from configurations $\sigma\in \{-1,+1\}^T$ such that $\sigma_U=\eta$. These definitions allow to write $$Z^{\pm}_H=\sum_{\eta\in\{-1,+1\}^U}Z^{\pm}_{\Go}(\eta)Z_T(\eta).$$ It follows by and Markov’s inequality that $$\begin{aligned} \lim_{n\rightarrow\infty}\PrGco\left(\sum_{\eta\in\{-1,+1\}^U}Z^{\pm}_{\Go}(\eta)Z_T(\eta)<\frac{1}{2\sqrt{n}}\sum_{\eta\in\{-1,+1\}^U}Z_T(\eta){\mathbb{E}}_{\Gco}[Z^{\pm}_{\Go}(\eta)]\right)=0.\\ \lim_{n\rightarrow\infty}\PrGco\left(\sum_{\eta\in\{-1,+1\}^U}Z^{\pm}_{\Go}(\eta)Z_T(\eta)>\frac{\sqrt{n}}{3}\sum_{\eta\in\{-1,+1\}^U}Z_T(\eta){\mathbb{E}}_{\Gco}[Z^{\pm}_{\Go}(\eta)]\right)=0, \end{aligned}$$ see also [@Sly10 (4.22) & (4.23)]. It follows that $\mu_H(Y(\sigma)=\pm)=Z_H^{\pm}/Z_H\geq 1/n$ a.a.s. over the choice of the graph $\Gc$. This proves the first item in Lemma \[lem:gadget2\]. The proof of the second item in Lemma \[lem:gadget2\] utilizes to a greater extent the connection to the tree and relies crucially on the non-reconstruction results we overviewed in Section \[sec:reco\]. First observe that $$\label{eq:Hproductmeasure} \mu_H(\sigma_U=\eta\, |\, Y(\sigma)=\pm)=\frac{Z^{\pm}_{\Go}(\eta)Z_T(\eta)}{Z_H^{\pm}}=\frac{Z^{\pm}_{\Go}(\eta)Z_T(\eta)}{\sum_{\eta'\in\{-1,+1\}^U}Z^{\pm}_{\Go}(\eta')Z_T(\eta')}.$$ Now, provided that $Z^{\pm}_{\Go}(\eta)<\frac{1}{\sqrt{n}}{\mathbb{E}}_{\Gco}[Z^{\pm}_{\Go}(\eta)]$ (see ), the right hand side of is (with large probability) within a multiplicative factor $n$ from $$\nu^{\pm}(\eta):=\frac{Z_T(\eta){\mathbb{E}}_{\Gco}[Z^{\pm}_{\Go}(\eta)]}{\sum_{\eta'\in\{-1,+1\}^U}Z_T(\eta'){\mathbb{E}}_{\Gco}[Z^{\pm}_{\Go}(\eta')]}=\frac{Z_{T}(\eta)Q^{\pm}_U(\eta)}{\sum_{\eta'\in\{-1,+1\}^U}Z_T(\eta')Q^{\pm}_U(\eta')},$$ where in the first equality we used . The crucial idea, captured in the proof of [@Sly10 Lemma 4.3], is that the measures $\nu^{\pm}$ on configurations $\{-1,+1\}^U$ are the same as the distributions induced by the random processes we described in Section \[sec:reco\], see the proof of [@Sly10 Lemma 4.3] for more details. This allows to use Lemma \[lem:Sly42\] together with to bound the probability mass of “bad" configurations $\eta$ on $U$, i.e., configurations on $U$ which exert large influence on the roots $R$ of the trees, see [@Sly10 Proof of Theorem 2.1]. This yields Item 2 in Lemma \[lem:gadget2\], concluding the argument. Calculating Moments {#sec:calculating-moments} =================== Random Bipartite $\Delta$-Regular Graphs {#sec:ratiofirstsecondproof} ---------------------------------------- We use the information we have obtained for the first and second moments to prove Lemma \[lem:ratiofirstsecond\]. Its proof is a very technical calculation. Recall that $Q^\pm=\frac{q^\pm}{1-q^\pm}$. To aid the presentation, it will be convenient to also have the following expressions defined beforehand. $$\mbox{\begin{tabular}{ll} \multicolumn{2}{c}{$E_1:=B_1 Q^{-} Q^{+}+Q^{-}+Q^{+}+B_2, \qquad E_2:=B_1 Q^{-} Q^{+}+B_1 B_2(Q^{+}+Q^{-})+B_2$,}\\ \multicolumn{2}{l}{$E_3:=(1-B_1 B_2)^2 Q^{-}Q^{+}+\Big(1+B_1(Q^{+}+Q^{-}) + B_1^2Q^{-}Q^{+}\Big)\Big(B^2_2+B_2(Q^{-}+Q^{+})+Q^{-}Q^{+}\Big).$} \end{tabular}}\label{eq:helpfulAs}$$ These expressions will emerge in the analysis of the asymptotics of the first and second moments, as is captured by the following lemmas. We are also going to need the explicit values of $\mathbf{X}^*=\arg\max_{\mathbf{X}}g_1(\mathbf{X})$ and $\mathbf{Y}^*=\arg\max_{\mathbf{X}}g_2(\mathbf{Y})$ when $\alpha=p^+,\beta=p^-,\gamma=\alpha^2,\delta=\beta^2$. These are given by: $$\begin{gathered} \left(\begin{array}{cc}x^{*}_{11}&x^{*}_{12}\\x^{*}_{21}&x^{*}_{22}\end{array}\right)=\frac{1}{E_1}\left(\begin{array}{cc}B_1 Q^{-}Q^{+}& Q^{+}\\Q^{-}& B_2\end{array}\right),\label{optimalxfirst}\\[0.2cm] \left(\begin{array}{cccc} y^*_{11}&y^*_{12}&y^*_{13}&y^*_{14}\\ y^*_{21}&y^*_{22}&y^*_{23}&y^*_{24}\\ y^*_{31}&y^*_{32}&y^*_{33}&y^*_{34}\\ y^*_{41}&y^*_{42}&y^*_{43}&y^*_{44} \end{array}\right)=\frac{1}{E_1^2}\left(\begin{array}{cccc} B_1^2 (Q^{-})^2 (Q^{+})^2 & B_1 Q^{-} (Q^{+})^2 & B_1 Q^{-} (Q^{+})^2 & (Q^{+})^2 \\ B_1 (Q^{-})^2 Q^{+} & B_1 B_2 Q^{-} Q^{+} & Q^{-} Q^{+} & B_2 Q^{+} \\ B_1 (Q^{-})^2 Q^{+} & Q^{-} Q^{+} & B_1 B_2 Q^{-} Q^{+} & B_2 Q^{+} \\ (Q^{-})^2 & B_2 Q^{-} & B_2 Q^{-} & B_2^2 \end{array}\right).\label{eq:optimalxsecond}\end{gathered}$$ The easiest way to argue about the validity of these values is to use the strict concavity of the functions $g_1(\mathbf{X}), g_2(\mathbf{Y})$ and just check that the required equalities for the critical points are satisfied. \[lem:asympfirst\] For $(\alpha,\beta)=(p^+,p^-)$, we have that $$\lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{{\mathcal{G}}}[Z^{\alpha,\beta}_G]}{\frac{1}{n}e^{n\Phi_1(\alpha,\beta,\mathbf{X}^{*})}}=\frac{1}{2\pi} \big(\alpha\beta(1-\alpha)(1-\beta)\big)^{(\Delta-1)/2}\left(\frac{Q^{+}Q^{-}E_2}{E^3_1}\right)^{-\Delta/2}.$$ \[lem:asympsec\] For $(\alpha,\beta)=(p^+,p^-)$, we have that $$\begin{aligned} \lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{{\mathcal{G}}}[(Z^{\alpha,\beta}_G)^2]}{\frac{1}{n^2}e^{n\Phi_2(\gamma^*,\delta^*,\mathbf{Y}^{*})}}&=\frac{1}{4\pi^2}\big(\alpha\beta(1-\alpha)(1-\beta)\big)^{2(\Delta-1)}\\ &\qquad\qquad\left(\frac{(Q^{+}Q^{-})^4E^3_2E_3}{E^{13}_1}\right)^{-\Delta/2}\left(\frac{(Q^+ Q^-)^2 E_2 E_3}{E_1^{7}}\right)^{1/2}\left(1-(\Delta-1)^2\omega^2\right)^{-1/2},\end{aligned}$$ where $\gamma^*=\alpha^2,\delta^*=\beta^2$, $\mathbf{Y}^{*}$ is given by and $\omega$ by . The proofs of Lemmas \[lem:asympfirst\] and \[lem:asympsec\] are given in Sections \[sec:asympfirst\] and \[sec:asympsec\] respectively. We need to state one more lemma before we can give the proof of Lemma \[lem:ratiofirstsecond\]. \[lem:firstsecvalues\] It holds that $2\Phi_1(\alpha,\beta,\mathbf{X^*})=\Phi_2(\alpha^2,\beta^2,\mathbf{Y}^*)$ It clearly suffices to prove $$\begin{aligned} f_2(\alpha^2,\beta^2)=2f_1(\alpha,\beta)\mbox{ and } g_2(\mathbf{Y}^*)=2g_1(\mathbf{X^*}).\end{aligned}$$ Both of these can be verified by straightforward brute force algebra. A neater way to do this, using more directly the uncorrelated property that Condition \[cond:maxima\] represents, is to use entropy arguments. We illustrate this for convenience in the case of $f_1$ and $f_2$, with the same technique extending to $g_1,g_2$ as well. Define $A,B$ to be independent Bernoulli variables with success probabilities $\alpha,\beta$ respectively. It is straightforward to see that $f_1(\alpha,\beta)=-\big(H(A)+H(B)\big)$ where $H(X)$ is the entropy of the random variable $X$. Let $A'=(A_1,A_2)$, where $A_1,A_2$ are independent copies of the random variable $A$. Define similarly $B'$. Using the definition of entropy, it is easy to check that $f_2(\alpha^2,\beta^2)=-\big(H(A')+H(B')\big)$. But clearly $H(A')=H(A_1)+H(A_2)=2H(A)$ and similarly for $B,B'$, thus verifying that $f_2(\alpha^2,\beta^2)=2f_1(\alpha,\beta)$. The observation which extends this idea for $g_2(\mathbf{Y}^*)=2g_1(\mathbf{X^*})$ is that the solution $\mathbf{Y}^*$, as given in , is the tensor product of $\mathbf{X}^*$, as given in , with itself. We omit the tedious details. We should emphasize that the proof of Lemma \[lem:firstsecvalues\] does not rely on the fact that $Q^+,Q^-$ are a function of the critical probabilities on the infinite tree, i.e., we do not need Lemma \[lem:densitiesone\], but rather the representation given in and the uncorrelated property that Condition \[cond:maxima\] represents. Employ Lemmas \[lem:asympfirst\], \[lem:asympsec\], and \[lem:firstsecvalues\] to obtain $$L:=\lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{{\mathcal{G}}}\left[(Z^{\alpha,\beta}_G)^2\right]}{\left({\mathbb{E}}_{{\mathcal{G}}}[Z^{\alpha,\beta}_G]\right)^2}=\left(1-(\Delta-1)^2\omega^2\right)^{-1/2}\big(\alpha\beta(1-\alpha)(1-\beta)\big)^{\Delta-1} \left(\frac{(Q^+ Q^-)^2 E_2 E_3}{E_1^{7}}\right)^{1/2}R^{-\Delta/2},$$ where $$R=\left(\frac{(Q^{+}Q^{-})^4E^3_2E_3}{E^{13}_1}\right)\cdot \left(\frac{Q^{+}Q^{-}E_2}{E^3_1}\right)^{-2}=\frac{(Q^{+}Q^{-})^2E_2E_3}{E^7_1}.$$ Thus $$\begin{aligned} L&=\big(1-(\Delta-1)^2\omega^2\big)^{-1/2}\big(\alpha\beta(1-\alpha)(1-\beta))^{\Delta-1}\left(\frac{(Q^{+}Q^{-})^2E_2E_3}{E^7_1}\right)^{-(\Delta-1)/2}\\ &=\big(1-(\Delta-1)^2\omega^2\big)^{-1/2}\left(\frac{1}{\alpha^2\beta^2(1-\alpha)^2(1-\beta)^2}\cdot\frac{(Q^{+}Q^{-})^2E_2E_3}{E^7_1}\right)^{-(\Delta-1)/2}.\end{aligned}$$ By , we have $\alpha=\frac{Q^+(1 +B_1 Q^-)}{E_1}$ and $\beta=\frac{Q^-(1 +B_1 Q^+)}{E_1}$, so that $$\begin{aligned} L&=\big(1-(\Delta-1)^2\omega^4\big)^{-1/2}\left(\frac{E_1 E_2 E_3}{(B_2 + Q^-)^2(B_2+Q^+)^2(1 +B_1 Q^-)^2(1 +B_1 Q^+)^2}\right)^{-(\Delta-1)/2}.\end{aligned}$$ The lemma follows after observing the easy to prove identity $$\frac{E_1 E_2 E_3}{(B_2 + Q^-)^2(B_2+Q^+)^2(1 +B_1 Q^-)^2(1 +B_1 Q^+)^2}=1-\omega^2.$$ ### The asymptotics of the first moment {#sec:asympfirst} In this section we prove Lemma \[lem:asympfirst\], i.e. compute the asymptotics of ${\mathbb{E}}_{{\mathcal{G}}}[Z^{\alpha,\beta}_G]$ for $(\alpha,\beta)=(p^+,p^-)$, see the relevant Sections \[sec:firstmomentform\] and \[sec:generallog\]. We will use the full-dimensional representation (see Section \[sec:maxentropy\]) of $g_1(\mathbf{X})$: while we are still going to refer to the values $x^{*}_{ij}$, in what follows the only free variable is going to be $x_{11}$. By the approximation $$\binom{bn}{an}=(1+o(1))\frac{1}{\sqrt{2\pi n}}\sqrt{\frac{b}{a(b-a)}}e^{n\big(b\ln b-a\ln a-(a-b)\ln(a-b)\big)},$$ we obtain $$\begin{aligned} \frac{{\mathbb{E}}_{{\mathcal{G}}}[Z^{\alpha,\beta}_G]}{\frac{1}{n}e^{n\Phi_1(\alpha,\beta,\mathbf{X}^{*})}}&=(1+o(1))\frac{1}{2\pi}\cdot \Big(\alpha\beta(1-\alpha)(1-\beta)\Big)^{\frac{\Delta-1}{2}}\cdot\\ &\qquad\left(\sum_{x_{11}}\Big(\prod_{i,j}x_{ij}\Big)^{-1/2}\frac{1}{\sqrt{2\pi n}}\cdot e^{n\big(g_1(x_{11})-g_1(x_{11}^*)\big)}\right)^{\Delta}.\end{aligned}$$ Note that in the above sum terms with $|x_{11}-x_{11}^*|\leq O(1)$ dominate exponentially over the rest and in a small enough ball around $x^*_{11}$ the function decays quadratically, as established by Lemma \[lem:gmax\]. Thus, standard integration arguments (for precise details see [@JLR Chapter 9]) yield $$\begin{aligned} \lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{{\mathcal{G}}}[Z^{\alpha,\beta}_G]}{\frac{1}{n}e^{n\Phi_1(\alpha,\beta,\mathbf{X}^{*})}}&=\frac{1}{2\pi} \Big(\alpha\beta(1-\alpha)(1-\beta)\Big)^{(\Delta-1)/2}\Big(\prod_{i,j} x^{*}_{ij}\Big)^{-\Delta/2}\\ &\qquad\left(\frac{1}{\sqrt{2\pi}}\int^\infty_{-\infty}\exp\left(\frac{\partial^2 g_1}{\partial x_{11}^2}(x_{11}^*)\cdot\frac{(x_{11}-x_{11}^*)^2}{2}\right)d x_{11}\right)^\Delta.\end{aligned}$$ Let $H:=\frac{\partial^2 g_1}{\partial x_{11}^2}(x_{11}^*)$. Since $g_1$ is strictly concave and it achieves its maximum at $x_{11}^*$ we have $H<0$. It follows that $$\frac{1}{\sqrt{2\pi}}\int^\infty_{-\infty}\exp\left(\frac{\partial^2 g_1}{\partial x_{11}^2}(x_{11}^*)\cdot\frac{(x_{11}-x_{11}^*)^2}{2}\right)d x_{11}=(-H)^{-1/2},$$ and consequently $$\begin{aligned} \lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{{\mathcal{G}}}[Z^{\alpha,\beta}_G]}{\frac{1}{n}e^{n\Phi_1(\alpha,\beta,\mathbf{X}^{*})}}&= \frac{1}{2\pi} \big(\alpha\beta(1-\alpha)(1-\beta)\big)^{(\Delta-1)/2}\Big((-H)\prod_{i,j} x^{*}_{ij}\Big)^{-\Delta/2}.\end{aligned}$$ We have $$\frac{\partial^2 g_1}{\partial x_{11}^2}=-\left(\frac{1}{x_{11}}+\frac{1}{\alpha-x_{11}}+\frac{1}{\beta-x_{11}}+\frac{1}{1-\alpha-\beta+x_{11}}\right).$$ Substituting the values of $\alpha,\beta$ by and the solution for $x^*_{11}$ by , we obtain $$-H=\frac{(B_2+Q^-+Q^++B_1 Q^- Q^+) (B_2+B_1 B_2 Q^-+B_1 B_2 Q^++B_1 Q^- Q^+)}{B_1 B_2 Q^- Q^+},$$ $$\prod_{i,j} x^{*}_{ij}=\frac{B_1B_2(Q^{+}Q^{-})^2}{(B_2+Q^-+Q^++B_1 Q^- Q^+)^4},$$ so that $$(-H)\prod_{i,j} x^{*}_{ij}=\frac{Q^{+}Q^{-}E_2}{E^3_1}.$$ ### The Asymptotics of the Second Moment {#sec:asympsec} In this section we prove Lemma \[lem:asympsec\], i.e., compute the asymptotics of ${\mathbb{E}}_{{\mathcal{G}}}[(Z^{\alpha,\beta}_G)^2]$ for $(\alpha,\beta)=(p^+,p^-)$. We have $${\mathbb{E}}_{{\mathcal{G}}}[(Z^{\alpha,\beta}_G)^2]=\sum_{\gamma,\delta}{\mathbb{E}}_{{\mathcal{G}}}[Y^{\gamma,\delta}_G],$$ where ${\mathbb{E}}_{{\mathcal{G}}}[Y^{\gamma,\delta}_G]$ was determined in Section \[sec:secondmomentform\] (see also the relevant Section \[sec:generallog\]). As in the proof of Lemma \[lem:asympfirst\], in the following we are going to refer to the full-dimensional representations (see Section \[sec:maxentropy\]) of $g_2,\Phi_2$, i.e. our free variables are going to be $\{y_{ij}\}$ with $1\leq i,j\leq3$, so that the rest of the $y_{ij}$ are just shorthand for their respective values (in terms of the free variables). We have that $$\begin{aligned} \frac{{\mathbb{E}}_{{\mathcal{G}}}[(Z^{\alpha,\beta}_G)^2]}{\frac{1}{n^2}e^{n\Phi_2(\gamma^*,\delta^*,\mathbf{Y}^{*})}}&= (1+o(1))\frac{1}{4\pi^2}\sum_{\gamma,\delta}\left(\frac{1}{\sqrt{2\pi n}}\right)^2\Big(\gamma\delta(1-2\alpha+\gamma)(1-2\beta+\delta)(\alpha-\gamma)^2(\beta-\delta)^2\Big)^{\frac{\Delta-1}{2}}\\ &\qquad\qquad\qquad\left(\sum_\mathbf{Y}\left(\frac{1}{\sqrt{2\pi n}}\right)^9\left(\prod^4_{i=1}\prod^4_{j=1}\frac{1}{\sqrt{y_{ij}}}\right)e^{n(\Phi_2(\gamma,\delta,\mathbf{Y})-\Phi_2(\gamma^*,\delta^*,\mathbf{Y}^*))/\Delta}\right)^{\Delta}.\end{aligned}$$ The function $\Phi_2(\gamma,\delta,\mathbf{Y})$ has a unique maximum at $(\gamma^*,\delta^*,\mathbf{Y}^*)$. Thus, in the above sum, terms with ${\lVert(\gamma,\delta,\mathbf{Y})-(\gamma^*,\delta^*,\mathbf{Y}^*)\rVert}\leq O(1)$ dominate exponentially over the rest. Moreover, in the region ${\lVert(\gamma,\delta,\mathbf{Y})-(\gamma^*,\delta^*,\mathbf{Y}^*)\rVert}\leq O(1)$ the function $\Phi_2$ decays quadratically. As in the proof of Lemma \[lem:asympfirst\], we thus obtain $$\begin{aligned} \lim_{n\rightarrow \infty}\frac{{\mathbb{E}}_{{\mathcal{G}}}[(Z^{\alpha,\beta}_G)^2]}{\frac{1}{n^2}e^{n\Phi_2(\gamma^*,\delta^*,\mathbf{Y}^{*})}}&=\left(\frac{1}{\sqrt{2\pi}}\right)^{15}\Big(\gamma^*\delta^*(1-2\alpha+\gamma^*)(1-2\beta+\delta^*)(\alpha-\gamma^*)^2(\beta-\delta^*)^2\Big)^{\frac{\Delta-1}{2}}\notag\\ &\left(\prod^4_{i,j=1}y^*_{ij}\right)^{-\frac{\Delta}{2}}\int^{\infty}_{-\infty}\int^{\infty}_{-\infty}\left(\frac{1}{\left(\sqrt{2\pi}\right)^9}\int^{\infty}_{-\infty}e^{\frac{1}{2}(\gamma,\delta,\mathbf{Y})\cdot\mathbf{H}\cdot(\gamma,\delta,\mathbf{Y})^{T}}d\mathbf{Y}\right)^\Delta d\gamma d\delta\notag\\ &=\frac{1}{4\pi^2}\Big(\alpha \beta(1-\alpha)(1-\beta)\Big)^{2(\Delta-1)}\left(\prod^4_{i,j=1}y^*_{ij}\right)^{-\frac{\Delta}{2}}\notag\\ &\ \ \ \left(\frac{1}{\sqrt{2\pi}}\right)^{2}\int^{\infty}_{-\infty}\int^{\infty}_{-\infty}\left(\frac{1}{\left(\sqrt{2\pi}\right)^9}\int^{\infty}_{-\infty}e^{\frac{1}{2}(\gamma,\delta,\mathbf{Y})\cdot\mathbf{H}\cdot(\gamma,\delta,\mathbf{Y})^{T}}d\mathbf{Y}\right)^\Delta d\gamma d\delta,\label{eq:secmain}\end{aligned}$$ where $\mathbf{H}$ denotes the Hessian matrix of $\Phi_2$ evaluated at $(\gamma^*,\delta^*,\mathbf{Y}^*)$ scaled by $1/\Delta$ and the operator $\cdot$ stands for matrix multiplication. We thus focus on computing the integral in . This will be achieved by two successive Gaussian integrations. First, we split the exponent of the quantity inside the integral as follows. $$\frac{1}{2}(\gamma,\delta,\mathbf{Y})\cdot\mathbf{H}\cdot(\gamma,\delta,\mathbf{Y})^{T}=\frac{1}{2}(\gamma,\delta)\cdot\mathbf{H}_{\gamma,\delta}\cdot(\gamma,\delta)^{T}-\frac{1}{2}\mathbf{Y}\cdot(-\mathbf{H}_{\mathbf{Y}})\cdot\mathbf{Y}^T+ \sum^3_{i=1}\sum^3_{j=1}T_{ij} y_{ij},$$ where $\mathbf{H}_{\gamma,\delta}$ denotes the principal minor of $\mathbf{H}$ corresponding to $\gamma,\delta$, $\mathbf{H}_{\mathbf{Y}}$ denotes the principal minor of $\mathbf{H}$ corresponding to $\mathbf{Y}$ and $$T_{ij}=\frac{1}{\Delta}\left(\gamma \cdot \frac{\partial^2 \Phi_2}{\partial \gamma \partial y_{ij}}+\delta \cdot \frac{\partial^2 \Phi_2}{\partial \delta \partial y_{ij}}\right).$$ It follows that $$\begin{aligned} \left(\int^{\infty}_{-\infty}e^{\frac{1}{2}(\gamma,\delta,\mathbf{Y})\cdot\mathbf{H}\cdot(\gamma,\delta,\mathbf{Y})^{T}}d\mathbf{Y}\right)^\Delta &=e^{\frac{\Delta}{2}(\gamma,\delta)\cdot\mathbf{H}_{\gamma,\delta}\cdot(\gamma,\delta)^{T}}\left(\int^{\infty}_{-\infty}e^{-\frac{1}{2}\mathbf{Y}\cdot(-\mathbf{H}_{\mathbf{Y}})\cdot\mathbf{Y}^T+ \sum^3_{i=1}\sum^3_{j=1}T_{ij} y_{ij}}d\mathbf{Y}\right)^\Delta.\end{aligned}$$ Let $\mathbf{T}$ denote the row vector with entries $T_{ij}$. Note that $\mathbf{H}_\mathbf{Y}$ is the Hessian of $g_2(\mathbf{Y})$ evaluated at $\mathbf{Y}^*$. Since $g_2(\mathbf{Y})$ is concave and $g_2$ achieves its maximum at $\mathbf{Y}^*$, we have that $\mathbf{H}_\mathbf{Y}$ is negative definite. Hence, $$\begin{aligned} \left(\frac{1}{\left(\sqrt{2\pi}\right)^9}\int^{\infty}_{-\infty}e^{\frac{1}{2}(\gamma,\delta,\mathbf{Y})\cdot\mathbf{H}\cdot(\gamma,\delta,\mathbf{Y})^{T}}d\mathbf{Y}\right)^\Delta &=\frac{1}{\left(\mathrm{Det}(-\mathbf{H}_{\mathbf{Y}})\right)^{\Delta/2}}e^{\frac{\Delta}{2}\left(\mathbf{T}\cdot (-\mathbf{H}_\mathbf{Y})^{-1}\cdot\mathbf{T}^T+(\gamma,\delta)\cdot\mathbf{H}_{\gamma,\delta}\cdot(\gamma,\delta)^{T}\right)}.\end{aligned}$$ We are thus left with the task of computing the integral $$\left(\frac{1}{\sqrt{2\pi}}\right)^2\int^{\infty}_{-\infty}\int^{\infty}_{-\infty}e^{\frac{\Delta}{2}\left(\mathbf{T}\cdot (-\mathbf{H}_\mathbf{Y})^{-1}\cdot\mathbf{T}^T+(\gamma,\delta)\cdot\mathbf{H}_{\gamma,\delta}\cdot(\gamma,\delta)^{T}\right)}.\label{eq:firstgaussian}$$ It can easily be seen that the expression in the exponent is a quadratic polynomial in $\gamma,\delta$, so that we may again perform Gaussian integration. We have $$\frac{\Delta}{2}\left(\mathbf{T}\cdot (-\mathbf{H}_\mathbf{Y})^{-1}\cdot\mathbf{T}^T+(\gamma,\delta)\cdot\mathbf{H}_{\gamma,\delta}\cdot(\gamma,\delta)^{T}\right)= D\gamma^2+E\gamma\delta+F\delta^2,\label{eq:defDEF}$$ for some complicated $D,E,F$, which can be determined explicitly by , once we perform the substitutions. Hence, $$\left(\frac{1}{\sqrt{2\pi}}\right)^2\int^{\infty}_{-\infty}\int^{\infty}_{-\infty}e^{\frac{\Delta}{2}\left(\mathbf{T}\cdot (-\mathbf{H}_\mathbf{Y})^{-1}\cdot\mathbf{T}^T+(\gamma,\delta)\cdot\mathbf{H}_{\gamma,\delta}\cdot(\gamma,\delta)^{T}\right)}d\gamma d\delta=\frac{1}{(4DF-E^2)^{1/2}},\label{eq:secondgaussian}$$ provided that $4DF-E^2>0$, so that the integration is meaningful. We will check this condition later. Combining equations , , , we obtain $$\begin{aligned} \lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{{\mathcal{G}}}[(Z^{\alpha,\beta}_G)^2]}{\frac{1}{n^2}e^{n\Phi_2(\gamma^*,\delta^*,\mathbf{Y}^{*})}}&=\frac{1}{4\pi^2}\Big(\alpha \beta(1-\alpha)(1-\beta)\Big)^{2(\Delta-1)}\left(\mathrm{Det}(-\mathbf{H}_{\mathbf{Y}})\prod^4_{i,j=1}y^*_{ij}\right)^{-\frac{\Delta}{2}}(4DF-E^2)^{-1/2}\notag\\ &=\frac{1}{4\pi^2}\Big(\alpha \beta(1-\alpha)(1-\beta)\Big)^{2(\Delta-1)}\cdot R,\label{eq:secmomentalmost}\end{aligned}$$ where $$R:=\left(\mathrm{Det}(-\mathbf{H}_{\mathbf{Y}})\prod^4_{i,j=1}y^*_{ij}\right)^{-\frac{\Delta}{2}}(4DF-E^2)^{-1/2}.$$ At this point, we resort to Maple to perform the necessary computations. The solution in the form of is particularly handy, since it rationalizes the expressions in the computations. We have: $$\mathrm{Det}(-\mathbf{H}_{\mathbf{Y}})=\frac{E_1^{19} E_2^3 E_3}{(B_1B_2)^8(Q^+ Q^-)^{12}},\ \prod^4_{i,j=1}y^*_{ij}=\frac{(B_1B_2)^8(Q^+ Q^-)^{16}}{E_1^{32}},$$ $$4DF-E^2=\frac{E_1^{7}}{(Q^+Q^-)^2 E_2 E_3}\cdot\left(1-(\Delta-1)^2\omega^2\right).$$ Note that by Lemma \[lem:technicalinequality\], we have $4DF-E^2>0$, so that the integration in is indeed meaningful. After plugging the above values into $R$ and substituting back in , the lemma follows. Gadget {#sec:gadget} ------ In this subsection we prove Lemmas \[lem:modfirst\] and \[lem:modsecond\]. We will need the following bound. \[lllo\] Let $b_1,\dots,b_\ell$ and $y_1,\dots,y_\ell$ be non-negative integers. Let $a=b_1+\dots+b_\ell$ and $x=y_1+\dots+y_\ell$. Assume $y_i^2\leq b_i$ for each $i\in\{1,\dots,\ell\}$. Then $$\frac{\binom{a+x}{ b_1+y_1, b_2+y_2,\dots,b_\ell+y_\ell}}{\binom{a}{b_1,b_2,\dots,b_\ell}} = \left(1+O\left(\sum_{i=1}^\ell y_i^2/b_i \right)\right)\frac{a^x}{b_1^{y_1}\dots b_\ell^{y_\ell}}.$$ We have $$\begin{aligned} \frac{\binom{a+x}{ b_1+y_1, b_2+y_2,\dots,b_\ell+y_\ell}}{\binom{a}{b_1,b_2,\dots,b_\ell}}& = \frac{(a+x)!}{a!}\prod_{i=1}^\ell\frac{b_i!}{(b_i+y_i)!} \\ &= \frac{a^x}{b_1^{y_1}\dots b_\ell^{y_\ell}}\left(\prod_{j=1}^x(1+j/a)\right)/ \prod_{i=1}^\ell \left(\prod_{j=1}^{y_i} (1+j/b_i)\right) \\ & = \frac{a^x}{b_1^{y_1}\dots b_\ell^{y_\ell}}\left(1+O\left(\frac{x^2}{a}+\frac{y_1^2}{b_1}+\dots+\frac{y_\ell^2}{b_\ell}\right)\right).\end{aligned}$$ The lemma now follows from the fact that $x^2/a\leq y_1^2/b_1+\dots+y_\ell^2/b_\ell$ (which follows from the Cauchy-Schwarz inequality). ### First Moment {#sec:modfirstproof} By , recall that $$\begin{aligned} {\mathbb{E}}_{\mathcal{G}}\big[ Z^{\alpha,\beta}_{G}\big] =\lambda^{(\alpha+\beta) n} \binom{n}{\alpha n}\binom{n}{\beta n} \left( \sum_{x\leq\min\{\alpha,\beta\}} \kappa^{\alpha,\beta,x}_{G} \right)^{\Delta},\end{aligned}$$ where $$\begin{aligned} \kappa^{\alpha,\beta,x}_{G} = & \frac{\binom{\alpha n}{x n}\binom{(1-\alpha)n}{(\beta-x)n}}{\binom{n}{\beta n}} B_1^{xn} B_2^{(1-\alpha-\beta+x)n}.\end{aligned}$$ Similarly, we have $$\begin{gathered} {\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[ Z^{\alpha,\beta}_{\overline{G}}(\eta)\big] \\ = \lambda^{(\alpha+\beta)n+\eta^-_1+\eta^-_2} \binom{n}{\alpha n}\binom{n}{\beta n} \left( \sum_{x\leq\min\{\alpha+\eta_1^-/n,\beta+\eta_2^-/n\}} \kappa^{\alpha,\beta,x}_{\overline{G}}(\eta) \right)^{\Delta-1} \left( \sum_{x\leq\min\{\alpha,\beta\}} \kappa^{\alpha,\beta,x}_{G} \right),\end{gathered}$$ where $$\begin{aligned} \kappa^{\alpha,\beta,x}_{\overline{G}}(\eta) = & \frac{\binom{\alpha n+\eta_1^-}{x n}\binom{(1-\alpha)n+\eta_1^+}{(\beta-x)n+\eta_2^-}}{\binom{n+m'}{\beta n+\eta^-_2}} B_1^{xn} B_2^{(1-\alpha-\beta+x)n+\eta_1^+-\eta_2^-}.\end{aligned}$$ From Lemma \[lllo\] and the fact that $\eta_1^+,\eta_1^-,\eta_2^+,\eta_2^-\leq n^{1/4}$, we have the following estimate on the ratio of $\kappa^{\alpha,\beta,x}_{\overline{G}}(\eta)$ and $\kappa^{\alpha,\beta,x}_{G}$: $$\label{wwzzz} \frac{\kappa^{\alpha,\beta,x}_{\overline{G}}(\eta)}{\kappa^{\alpha,\beta,x}_{G}} = (1+o(1)) \frac{\alpha^{\eta_1^-}(1-\alpha)^{\eta_1^+}\beta^{\eta_2^-}(1-\beta)^{\eta_2^+}}{(\alpha-x)^{\eta_1^-} (1-\alpha-\beta+x)^{\eta_1^+-\eta_2^-} (\beta-x)^{\eta_2^-}}B_2^{\eta_1^+ - \eta_2^-}. $$ It is easy to see that for any $\alpha,\beta$, the logarithm of the function $\kappa^{\alpha,\beta,x}_{G}$ scaled by $n$ is exactly the full-dimensional representation of the function $g_1(\mathbf{X})$ (see Section \[sec:generallog\]). Since $g_1(\mathbf{X})$ is strictly concave with respect to $\mathbf{X}$, it has a unique maximum. With these observations, it can easily be calculated that the value of $x=x_{11}$ for which $g_1(\mathbf{X})$ achieves its maximum is given by the unique positive solution of $$B_1B_2(\alpha-x^*)(\beta-x^*)=x^*(1-\alpha-\beta+x^*).$$ By Lemma \[lem:gmax\], the logarithm of $\kappa^{\alpha,\beta,x}_{G}$ decays quadratically in the distance from $x^*$. Let $${\cal A} = \{x : |x-x^*|\leq n^{-1/4}\}.$$ The relative contribution of terms outside ${\cal A}$ to ${\mathbb{E}}_{\mathcal{G}}\big[Z^{\alpha,\beta}_{G}\big]$ is $\exp(-\Omega(n^{1/2}))$ and hence can be omitted. Similarly, using , the relative contribution of terms outside ${\cal A}$ to ${\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[ Z^{\alpha,\beta}_{\overline{G}}(\eta)\big]$ is $\exp(-\Omega(n^{1/2}))$ and hence can also be omitted. For $x\in{\cal A}$ we have $$\begin{aligned} \frac{\kappa^{\alpha,\beta,x}_{\overline{G}}(\eta)}{\kappa^{\alpha,\beta,x}_{G}} = & (1+o(1)) \frac{\alpha^{\eta_1^-}(1-\alpha)^{\eta_1^+}\beta^{\eta_2^-}(1-\beta)^{\eta_2^+}}{(\alpha-x^*)^{\eta_1^-} (1-\alpha-\beta+x^*)^{\eta_1^+-\eta_2^-} (\beta-x^*)^{\eta_2^-}}B_2^{\eta_1^+ - \eta_2^-}.\end{aligned}$$ Thus, $$\frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[ Z^{\alpha,\beta}_{\overline{G}}(\eta)\big]}{{\mathbb{E}}_{\mathcal{G}}\big[ Z^{\alpha,\beta}_{G}\big]}=(1+o(1))\left( \frac{\alpha^{\eta_1^-}(1-\alpha)^{\eta_1^+}\beta^{\eta_2^-}(1-\beta)^{\eta_2^+}}{(\alpha-x^*)^{\eta_1^-} (1-\alpha-\beta+x^*)^{\eta_1^+-\eta_2^-} (\beta-x^*)^{\eta_2^-}}B_2^{\eta_1^+ - \eta_2^-} \right)^{\Delta-1}\lambda^{\eta_1^-+\eta_2^-},$$ which proves the first part of the lemma. When $(\alpha,\beta)=(p^+,p^-)$, we express $\alpha,\beta$ and $x^*=x^*_{11}$ in terms of $Q^+,Q^-$ using equations  and . We obtain $$\begin{aligned} \lefteqn{ \frac{\alpha^{\eta_1^-}(1-\alpha)^{\eta_1^+}\beta^{\eta_2^-}(1-\beta)^{\eta_2^+}}{(\alpha-x^*)^{\eta_1^-} (1-\alpha-\beta+x^*)^{\eta_1^+-\eta_2^-} (\beta-x^*)^{\eta_2^-}}B_2^{\eta_1^+ - \eta_2^+} } \hspace{1in} \\ & = \left(\frac{B_2(1-\alpha)(1-\beta)}{1-\alpha-\beta+x^*}\right)^{m'} \left(\frac{\alpha(1-\alpha-\beta+x^*)}{B_2 (\alpha-x^*)(1-\alpha)}\right)^{\eta_1^-} \left(\frac{\beta(1-\alpha-\beta+x^*)}{B_2 (\beta-x^*)(1-\beta)}\right)^{\eta_2^-}\\ & = \left(\frac{(B_2+Q^+)(B_2+Q^-)}{B_2+Q^++Q^-+B_1Q^+Q^-}\right)^{m'} \left(\frac{1+B_1Q^-}{B_2+Q^-}\right)^{\eta_1^-} \left(\frac{1+B_1Q^+}{B_2+Q^+}\right)^{\eta_2^-}.\end{aligned}$$ and hence $$\frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[ Z^{\alpha,\beta}_{\overline{G}}(\eta)\big]}{{\mathbb{E}}_{\mathcal{G}}\big[ Z^{\alpha,\beta}_{G}\big]}= (1+o(1)) C^* \left(\lambda\left(\frac{1+B_1Q^-}{B_2+Q^-}\right)^{\Delta-1}\right)^{\eta_1^-} \left(\lambda\left(\frac{1+B_1Q^+}{B_2+Q^+}\right)^{\Delta-1}\right)^{\eta_2^-},\label{eq:qwertya}$$ where $C^*$ is given by . Using the fact that $Q^+,Q^-$ satisfy , we have $$Q^+=\lambda\left(\frac{1+B_1Q^-}{B_2+Q^-}\right)^{\Delta-1}\quad\mbox{and}\quad Q^-=\lambda\left( \frac{1+B_1Q^+}{B_2+Q^+}\right)^{\Delta-1}.$$ Hence, we obtain $$\begin{aligned} \frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[ Z^{\alpha,\beta}_{\overline{G}}(\eta)\big]}{{\mathbb{E}}_{\mathcal{G}}\big[ Z^{\alpha,\beta}_{G}\big]} &= & (1+o(1))C^* (1+Q^+)^{m'}(1+Q^-)^{m'} \\ & & \times\left(\frac{Q^+}{1+Q^+}\right)^{\eta_1^-} \left(\frac{1}{1+Q^+}\right)^{\eta_1^+} \left(\frac{Q^-}{1+Q^-}\right)^{\eta_2^-} \left(\frac{1}{1+Q^-}\right)^{\eta_2^+}.\end{aligned}$$ The second part of the lemma follows after observing that $q^{\pm}=\frac{Q^\pm}{1+Q^\pm},\ 1-q^\pm=\frac{1}{1+Q^\pm}$. ### Second Moment {#lem:modsecondproof} By , we have $$\begin{aligned} {\mathbb{E}}_{{\mathcal{G}}}\big[\big(Z^{\alpha,\beta}_{G}\big)^2\big]&=\lambda^{2(\alpha+\beta) n}\binom{n}{\alpha n}\binom{n}{\beta n}\sum_{\gamma,\delta}\binom{\alpha n}{\gamma n}\binom{(1-\alpha)n}{(\alpha-\gamma)n}\binom{\beta n}{\delta n}\binom{(1-\beta)n}{(\beta-\gamma)n}\bigg(\sum_{\mathbf{Y}}\kappa^{\gamma,\delta,\mathbf{Y}}_G\bigg)^\Delta\label{OOUU1}\end{aligned}$$ where $$\begin{aligned} \kappa^{\gamma,\delta,\mathbf{Y}}_G = & \prod_{i=1}^4\binom{L_i n}{y_{i1}n,y_{i2}n,y_{i3}n,y_{i4}n}\prod^4_{j=1} \binom{R_jn}{y_{1j}n,y_{2j}n,y_{3j}n,y_{4j}n}/\binom{n}{y_{11}n, y_{12}n, \dots ,y_{44}n}\\ & \times B_1^{(2y_{11} + y_{12} + y_{13}+ y_{21} + y_{22} +y_{31}+y_{33})n} B_2^{(y_{22} + y_{24} + y_{33} + y_{34}+y_{42}+y_{43}+2y_{44})n},\end{aligned}$$ $$\begin{array}{lll} L_1=\gamma,& L_2=L_3=\alpha-\gamma,& L_4=1-2\alpha+\gamma,\\ R_1=\delta,& R_2=R_3=\beta-\delta,& R_4=1-2\beta+\delta, \end{array}$$ and $\mathbf{Y}$ stands for the nonnegative variables $y_{ij}$, $1\leq i,j\leq 4$ such that $$\mbox{$\sum_j$}y_{ij}=L_i\mbox{ and }\mbox{$\sum_i$}y_{ij}=R_j.$$ Similarly, we have that $$\begin{aligned} {\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[\big(Z^{\alpha,\beta}_{\overline{G}}(\eta)\big)^2\big]&=\lambda^{2(\alpha+\beta) n + 2(\eta_1^- + \eta_2^-)}\binom{n}{\alpha n}\binom{n}{\beta n}\sum_{\gamma,\delta}\binom{\alpha n}{\gamma n}\binom{(1-\alpha)n}{(\alpha-\gamma)n}\binom{\beta n}{\delta n}\binom{(1-\beta)n}{(\beta-\gamma)n}\label{OOUU2}\\ &\hskip 6.5cm \times \bigg(\sum_{\hat{\mathbf{Y}}}\kappa^{\gamma,\delta,\hat{\mathbf{Y}}}_{\overline{G}}(\eta)\bigg)^{\Delta-1}\bigg(\sum_{\mathbf{Y}}\kappa^{\gamma,\delta,\mathbf{Y}}_G\bigg),\nonumber\end{aligned}$$ where $$\begin{aligned} \kappa^{\gamma,\delta,\hat{\mathbf{Y}}}_{\overline{G}}(\eta)= & \prod_{i=1}^4\binom{\hat{L}_i n}{\hat{y}_{i1}n,\hat{y}_{i2}n,\hat{y}_{i3}n,\hat{y}_{i4}n} \prod^4_{j=1} \binom{\hat{R}_jn}{\hat{y}_{1j}n,\hat{y}_{2j}n,\hat{y}_{3j}n,\hat{y}_{4j}n}/\binom{n+m'}{\hat{y}_{11}n, \hat{y}_{12}n, \dots ,\hat{y}_{44}n}\\ & \times {B}_1^{(2\hat{y}_{11} + \hat{y}_{12} + \hat{y}_{13}+ \hat{y}_{21} + \hat{y}_{22} +\hat{y}_{31}+\hat{y}_{33})n} {B}_2^{(\hat{y}_{22} + \hat{y}_{24} + \hat{y}_{33} + \hat{y}_{34}+\hat{y}_{42}+\hat{y}_{43}+2\hat{y}_{44})n},\end{aligned}$$ $$\begin{array}{lll} \hat{L}_1=\gamma+\eta_1^-/n,& \hat{L}_2=\hat{L}_3=\alpha-\gamma,& \hat{L}_4=1-2\alpha+\gamma+\eta_1^+/n,\\ \hat{R}_1=\delta+\eta_2^-/n,& \hat{R}_2=\hat{R}_3=\beta-\delta,& \hat{R}_4=1-2\beta+\delta+\eta_2^+/n, \end{array}$$ and $\hat{\mathbf{Y}}$ stands for the nonnegative variables $\hat{y}_{ij}$, $1\leq i,j\leq 4$ such that $$\mbox{$\sum_j$}\hat{y}_{ij}=\hat{L}_i\mbox{ and }\mbox{$\sum_j$}\hat{y}_{ij}=\hat{R}_j.$$ For a given $\mathbf{Y}$, consider $\hat{\mathbf{Y}}$ such that $y_{ij}=\hat{y}_{ij}$ for all $i,j\in\{1,2,3,4\}$, except $\hat{y}_{14}=y_{14}+\eta_1^-$, $\hat{y}_{41}=y_{41}+\eta_2^-$, and $\hat{y}_{44}=y_{44}+m'-\eta_1^- - \eta_2^-$. This is always possible, except at the boundary, but as we will see shortly, these cases can be safely ignored. By Lemma \[lllo\], we have $$\label{eyy} \frac{\kappa^{\gamma,\delta,\hat{\mathbf{Y}}}_{\overline{G}}(\eta)}{\kappa^{\gamma,\delta,\mathbf{Y}}_G}= (1+o(1)) \frac{\gamma^{\eta_1^-}(1-2\alpha+\gamma)^{\eta_1^+}\delta^{\eta_2^-}(1-2\beta+\delta)^{\eta_2^+}}{y_{14}^{\eta_1^-} y_{41}^{\eta_2^-} y_{44}^{m'-\eta_1^--\eta_2^-}} B_2^{2m'-2\eta_1^--2\eta_2^-}.$$ Let $${\cal A} = \{(\gamma,\delta,\mathbf{Y}) : \|(\gamma,\delta,\mathbf{Y})-(\gamma^*,\delta^*,\mathbf{Y}^*)\|_2\leq n^{-1/4}\},$$ where $\gamma^*=\alpha^2$, $\delta^*=\beta^2$ and $\mathbf{Y}^*$ is given by . By Condition \[cond:maxima\] and Lemma \[lem:gmax\], the relative contribution of terms outside ${\cal A}$ to ${\mathbb{E}}_{{\mathcal{G}}}\big[\big(Z^{\alpha,\beta}_{G}\big)^2\big]$ is $\exp(-\Omega(n^{1/2}))$ and hence can be omitted. Similarly, using , the relative contribution of terms outside ${\cal A}$ to ${\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[\big(Z^{\alpha,\beta}_{\overline{G}}(\eta)\big)^2\big]$ is $\exp(-\Omega(n^{1/2}))$ and hence can also be omitted. For $(\gamma,\delta,{\mathbf Y})\in{\cal A}$ we have $$\frac{\kappa^{\gamma,\delta,\hat{\mathbf{Y}}}_{\overline{G}}(\eta)}{\kappa^{\gamma,\delta,\mathbf{Y}}_G}= (1+o(1)) \frac{(\gamma^*)^{\eta_1^-}(1-2\alpha+\gamma^*)^{\eta_1^+}(\delta^*)^{\eta_2^-}(1-2\beta+\delta^*)^{\eta_2^+}}{(y_{14}^*)^{\eta_1^-} (y_{41}^*)^{\eta_2^-} (y_{44}^*)^{m'-\eta_1^--\eta_2^-}} B_2^{2m'-2\eta_1^--2\eta_2^-}.$$ Using equations  and , we express $\alpha,\beta$ and $\mathbf{Y}$ in terms of $Q^+,Q^-$. For $(\gamma,\delta,{\mathbf Y})\in{\cal A}$, we obtain that $$\begin{aligned} \frac{\kappa^{\gamma,\delta,\hat{\mathbf{Y}}}_{\overline{G}}(\eta)}{\kappa^{\gamma,\delta,\mathbf{Y}}_G} & = & (1+o(1)) \left(\frac{B_2^2 (1-2\alpha+\gamma^*)(1-2\beta+\delta^*)}{y^*_{44}}\right)^{m'} \\& & \times \left(\frac{\gamma^* y^*_{44}}{B_2^2 y^*_{14} (1-2\alpha+\gamma^*)}\right)^{\eta_1^-} \left(\frac{\delta^* y^*_{44}}{B_2^2 y^*_{41} (1-2\beta+\delta^*)}\right)^{\eta_2^-}\\ & = &(1+o(1)) \left(\frac{(B_2+Q^+)(B_2+Q^-)}{B_2+Q^++Q^-+B_1Q^+Q^-}\right)^{2m'} \left(\frac{1+B_1Q^-}{B_2+Q^-}\right)^{2\eta_1^-} \left(\frac{1+B_1Q^+}{B_2+Q^+}\right)^{2\eta_2^-}.\end{aligned}$$ Plugging the last equation into  and  we obtain the lemma. ### Proof of Lemma \[lem:rfsgadget\] {#sec:rfsgadgetproof} By equations and , we have $$\lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[\big(Z^{\alpha,\beta}_{\overline{G}}(\eta)\big)^2\big]}{\big({\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[Z^{\alpha,\beta}_{\overline{G}}(\eta)\big]\big)^2}=\lim_{n\rightarrow\infty}\frac{{\mathbb{E}}_{{\mathcal{G}}}\big[\big(Z^{\alpha,\beta}_{G}\big)^2\big]}{\big({\mathbb{E}}_{{\mathcal{G}}}\big[Z^{\alpha,\beta}_{G}\big]\big)^2}.$$ The lemma follows, after using Lemma \[lem:ratiofirstsecond\]. Asymptotically Almost Surely results {#sec:smallgraph} ==================================== Overview {#sec:oversmall} -------- In this section, we apply the small graph conditioning method to obtain Lemmas \[lem:smallgraph\] and \[lem:smallgraphgadget\], see Section \[sec:proof-approach\] for a brief discussion. Closely related results have appeared in [@MWW; @Sly10] for the hard-core model, but here we need to account for the extra parameters $B_1,B_2$. The proofs are similar, however the technical calculations are a bit more complicated due to the extra parameters. The following theorem is taken from [@MWW]. The notation $[X]_{m}$ refers to the $m$-th order falling factorial of the variable $X$. \[thm:smallgraphmethod\] Let $\lambda_i>0$ and $\delta_i>-1$ be real numbers for $i=1,2,\hdots$. Let $r(n)\rightarrow 0$ and suppose that for each $n$ there are random variables $X_i=X_i(n)$, $i=1,2,\hdots$ and $Y=Y(n)$, all defined on the same probability space ${\mathcal{G}}={\mathcal{G}}_n$ such that $X_i$ is nonnegative integer valued, $Y$ is nonnegative and ${\mathbb{E}}[ Y]>0$ (for $n$ sufficiently large). Suppose furthermore that 1. For each $k\geq 1$, the variables $X_1,\hdots,X_k$ are asymptotically independent Poisson random variables with ${\mathbb{E}}X_i\rightarrow \lambda_i$, 2. for every finite sequence $m_1,\hdots,m_k$ of nonnegative integers, $$\frac{{\mathbb{E}}[Y[X_1]_{m_1}\cdots [X_k]_{m_k}]}{{\mathbb{E}}[Y]}\rightarrow \prod^k_{i=1}\big(\lambda_i(1+\delta_i)\big)^{m_i},\label{eq:prodsmall}$$ 3. $\sum_i \lambda_i \delta^2_i<\infty$. 4. ${\mathbb{E}}[Y^2]/({\mathbb{E}}[Y])^2\leq \exp\big(\sum_i \lambda_i \delta^2_i)+o(1)$ as $n\rightarrow \infty$. Then $Y>r(n){\mathbb{E}}[Y]$ asymptotically almost surely. Recall that $Z^{\alpha,\beta}_G$ is the measure of configurations in $\Sigma^{\alpha,\beta}_G$ for a random $\Delta$-regular bipartite graph $G$. Let $X_i$ be the number of circles of length $i$ in $G$. Clearly $X_i=0$ iff $i$ is odd. We have the following lemmas. \[lem:cycle\] Condition 1 of Theorem \[thm:smallgraphmethod\] holds for even $i$ with $$\lambda_i=\frac{r(\Delta,i)}{i},$$ where $r(\Delta,i)$ is the number of ways one can properly edge color a cycle of length $i$ with $\Delta$ colors. The proof of Lemma \[lem:cycle\] is given in [@MWW]. \[lem:ratiosmallgraph\] For $(\alpha,\beta)=(p^\pm,p^\mp)$, we have that $$\frac{{\mathbb{E}}_{\mathcal{G}}[Z^{\alpha,\beta}_G X_i]}{{\mathbb{E}}_{\mathcal{G}}[Z^{\alpha,\beta}_G]}\rightarrow \lambda_i(1+\delta_i)\quad\mbox{ and }\quad \frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}[Z^{\alpha,\beta}_{\overline{G}}(\eta) X_i]}{{\mathbb{E}}_{\overline{{\mathcal{G}}}}[Z^{\alpha,\beta}_{\overline{G}}(\eta)]}\rightarrow \lambda_i(1+\delta_i),$$ where $$\delta_i=\frac{(1-B_1B_2)^{i/2}(Q^+Q^-)^{i/2}}{\Big((1+B_1 Q^+)(1+B_1 Q^-)(B_2+Q^+)(B_2+Q^-)\Big)^{i/2}}.$$ The proof of Lemma \[lem:ratiosmallgraph\] is given in Section \[sec:smallgraphregular\]. \[lem:sumasymptotics\] It holds that $$\exp\left(\sum_{\mbox{even }i\geq 2}\lambda_i\delta^2_i\right)=\big(1-\omega^2\big)^{-(\Delta-1)/2}\big(1-(\Delta-1)^2\omega^2\big)^{-1/2},$$ where $\omega$ is given by . The proof of Lemma \[lem:sumasymptotics\] is given in Section \[sec:smallgraphregular\]. \[lem:finitesequence\] For $(\alpha,\beta)=(p^\pm,p^\mp)$ and for every finite sequence $m_1,\hdots,m_k$ of nonnegative integers, it holds that $$\frac{{\mathbb{E}}_{\mathcal{G}}\big[Z^{\alpha,\beta}_G(\eta) [X_2]_{m_1}\cdots [X_{2k}]_{m_k}\big]}{{\mathbb{E}}_{\mathcal{G}}[Z^{\alpha,\beta}_G]}\rightarrow \prod^{k}_{i=1}\big(\lambda_i(1+\delta_i)\big)^{m_i},$$ and $$\frac{{\mathbb{E}}_{\overline{{\mathcal{G}}}}\big[Z^{\alpha,\beta}_{\overline{G}}(\eta)[X_2]_{m_1}\cdots [X_{2k}]_{m_k}\big]}{{\mathbb{E}}_{\overline{{\mathcal{G}}}}[Z^{\alpha,\beta}_{\overline{G}}(\eta)]}\rightarrow \prod^{k}_{i=1}\big(\lambda_i(1+\delta_i)\big)^{m_i}.$$ The proof of Lemma \[lem:finitesequence\] is identical to [@MWW Proof of Lemma 7.5]. We are now in position to prove Lemmas \[lem:smallgraph\] and  \[lem:smallgraphgadget\]. We verify the conditions of Theorem \[thm:smallgraphmethod\] in each case: Condition 1 holds (Lemma \[lem:cycle\]), Condition 2 holds (Lemma \[lem:finitesequence\]), Conditions 3 and 4 hold (Lemma \[lem:sumasymptotics\] and Lemmas \[lem:ratiofirstsecond\] and \[lem:rfsgadget\]). Lemmas \[lem:smallgraph\] and \[lem:smallgraphgadget\] follow by picking $r(n)=1/n$ and $r(n)=1/\sqrt{n}$ respectively. Random Bipartite $\Delta$-regular Graphs {#sec:smallgraphregular} ---------------------------------------- In this section, we prove Lemma \[lem:ratiosmallgraph\]. We give the proof for the variables $Z^{\alpha,\beta}_G$ and omit the proof for the variables $Z^{\alpha,\beta}_{\overline{G}}(\eta)$, which can be carried out the same way as in [@Sly10 Lemma 3.8] and along the lines of Section \[sec:modfirstproof\]. Recall that $r(\Delta,i)$ denotes the number of proper $\Delta$-edge colorings of a cycle of length $i$. \[lem:properedgecolorings\] $r(\Delta,i)=(\Delta-1)^i+(-1)^i(\Delta-1)$. The (simple) proof of Lemma \[lem:properedgecolorings\] is given in [@MWW Lemma 7.6]. For sets $S,T$ such that $S\subset V_1, T\subset V_2,|S|=\alpha n, |T|=\beta n$, denote by $Y_{S,T}$ the measure of the configuration $\sigma$ that $S,T$ induce, i.e. for a vertex $v\in V(G)$, $\sigma(v)=1$ iff $v\in S\cup T$. Fix a specific pair of $S,T$. By symmetry, $$\frac{{\mathbb{E}}[Z^{\alpha,\beta}_GX_i]}{{\mathbb{E}}[Z^{\alpha,\beta}_G]}=\frac{{\mathbb{E}}[Y_{S,T} X_i]}{{\mathbb{E}}[Y_{S,T}]}. \label{eq:smallratio}$$ We now decompose $X_i$ as follows: - $\xi$ will denote a proper $\Delta$-edge colored, rooted and oriented $i$-cycle ($r(\Delta,i)$ possibilities), in which the vertices are colored with red($R$), blue($B$) and white($W$) such that no two adjacent vertices in the cycle have both $R$ or $B$ color, i.e. the only assignments which are prohibited for an edge are $(R,R)$, $(B,B)$. The color of the edges will prescribe which of the $\Delta$ perfect matchings an edge of a (potential) cycle will belong to. $R$ denotes vertices belonging to $S$, $B$ denotes vertices belonging to $T$ and $W$ the remaining vertices. - Given $\xi$, $\zeta$ denotes a position that an $i$-cycle can be in (i.e. the exact vertices it traverses, in order) such that prescription of the vertex colors of $\xi$ is satisfied. - $\mathbf{1}_{\xi,\zeta}$ is the indicator function whether a cycle specified by $\xi,\zeta$ is present in the graph $G$. Note that each possible cycle corresponds to exactly $2i$ different configurations $\xi$ (the number of ways to root and orient the cycle). For each of those $\xi$, the respective sets of configurations $\zeta$ are the same. Hence, we may write $$X_i=\frac{1}{2i}\sum_\xi\sum_\zeta \mathbf{1}_{\xi,\zeta}.$$ Let $P_1:= \Pr[\mathbf{1}_{\xi,\zeta}=1]$. It follows that $$\begin{aligned} {\mathbb{E}}[Y_{S,T}X_i]&=\frac{1}{2i}\sum_\xi\sum_\zeta P_1\cdot{\mathbb{E}}[Y_{S,T}|\mathbf{1}_{\xi,\zeta}=1].\end{aligned}$$ In light of , we need to study the ratio ${\mathbb{E}}[Y_{S,T}|\mathbf{1}_{\xi,\zeta}=1]/{\mathbb{E}}[Y_{S,T}]$. At this point, to simplify notation, we may assume that $\xi,\zeta$ are fixed. We have shown in Section \[sec:firstmomentform\] that $${\mathbb{E}}[Y_{S,T}]=\lambda^{(\alpha+\beta)n}\left(\sum_{\delta} \frac{\binom{\alpha n}{\delta n}\binom{(1-\alpha)n}{(\beta-\delta)n}}{\binom{n}{\beta n}}B_1^{\delta n}B_2^{(1-\alpha-\beta+\delta)n}\right)^\Delta,$$ where $\delta$ denotes the number of edges between $S,T$ in one matching. To calculate ${\mathbb{E}}[Y_{S,T}|\mathbf{1}_{\xi,\zeta}=1]$, we need some notation. For colors $c_1,c_2\in\{R,B,W\}$, we say that an edge is of type $\{c_1,c_2\}$ if its endpoints have colors $c_1,c_2$. Let $j_1,j_2$ denote the number of red and blue vertices in the coloring prescribed by $\xi$. For $k = 1,\hdots,\Delta$, let $a_1(k)$ denote the number of edges of color $k$ of type $\{R,B\}$, $a_2(k)$ denote the number of edges of color $k$ of type $\{W,W\}$, $d_1(k)$ denote the number of edges of color $k$ of type $\{R,W\}$, $d_2(k)$ denote the number of edges of color $k$ of type $\{B,W\}$. Finally, for $j=1,2$ let $a_j=\sum_k a_j(k)$ and $d_j=\sum_k d_j(k)$. By considering the sum of the degrees of $R$ vertices, the sum of the degrees of $B$ vertices and the total number of edges of the cycle, we obtain the following equalities. $$a_1+d_1=2j_1,\ a_1+d_2=2j_2,\ a_1+a_2+d_1+d_2=i.\label{eq:graphsimple}$$ A straightforward modification of the arguments in Section \[sec:firstmomentform\] yields $$\begin{aligned} {\mathbb{E}}[Y_{S,T}|\mathbf{1}_{\xi,\zeta}=1]&=\lambda^{(\alpha+\beta)n}\\ &\times\prod^\Delta_{k=1}\left(\sum_{\delta_k} \frac{\binom{\alpha n-a_1(k)-d_1(k)}{\delta_k n}\binom{(1-\alpha)n-a_2(k)-d_2(k)}{(\beta-\delta_k)n-a_1(k)-d_2(k)}}{\binom{n-a_1(k)-a_2(k)-d_1(k)-d_2(k)}{\beta n-a_1(k)-d_2(k)}}B_1^{\delta_k n+a_1(k)}B_2^{(1-\alpha-\beta+\delta_k) n+a_1(k)}\right).\end{aligned}$$ By Lemma \[lllo\], we have $$\begin{aligned} \frac{\binom{\alpha n-a_1(k)-d_1(k)}{\delta_k n}}{\binom{\alpha n}{\delta_k n}}&\sim \left(\frac{\alpha-\delta_k}{\alpha}\right)^{a_1(k)+d_1(k)},\\ \frac{\binom{(1-\alpha)n-a_2(k)-d_2(k)}{(\beta-\delta_k)n-a_1(k)-d_2(k)}}{\binom{(1-\alpha)n}{(\beta-\delta_k)n}}&\sim \left(\frac{1-\alpha-\beta+\delta_k}{1-\alpha}\right)^{a_2(k)+d_2(k)}\left(\frac{\beta-\delta_k}{1-\alpha-\beta+\delta_k}\right)^{a_1(k)+d_2(k)},\\ \frac{\binom{n-a_1(k)-a_2(k)-d_1(k)-d_2(k)}{\beta n-a_1(k)-d_2(k)}}{\binom{n}{\beta n}}&\sim\beta^{a_1(k)+d_2(k)}\left(1-\beta\right)^{a_2(k)+d_1(k)}.\end{aligned}$$ By the same line of arguments as in the proofs of Lemmas \[lem:modfirst\] and \[lem:modsecond\], we obtain that $$\begin{aligned} \frac{{\mathbb{E}}[Y_{S,T}|\mathbf{1}_{\xi,\zeta}=1]}{{\mathbb{E}}[Y_{S,T}]}&\sim\frac{(\delta^{*})^{a_2-a_1}(\alpha-\delta^{*})^{a_1+d_1}(\beta-\delta^{*})^{a_1+d_2}(B_1B_2)^{a_1}}{\alpha^{a_1+d_1}\beta^{a_1+d_2}(1-\alpha)^{a_2+d_2}(1-\beta)^{a_2+d_1}}.\end{aligned}$$ Clearly $P_1\sim n^{-i}$ and for given $\xi$, the number of possible $\zeta$ is asymptotic to $$\alpha^{j_1}\beta^{j_2}(1-\alpha)^{i/2-j_1}(1-\beta)^{i/2-j_2}n^i.$$ Thus, for the given $\xi$, we have $$\begin{aligned} \frac{\sum_\zeta P_1{\mathbb{E}}[Y_{S,T}|\mathbf{1}_{\xi,\zeta}=1]}{{\mathbb{E}}[Y_{S,T}]}\sim\frac{(1-\alpha-\beta+\delta^{*})^{a_2-a_1}(\alpha-\delta^{*})^{a_1+d_1}(\beta-\delta^{*})^{a_1+d_2}(B_1B_2)^{a_1}}{\alpha^{a_1+d_1-j_1}\beta^{a_1+d_2-j_2}(1-\alpha)^{a_2+d_2+j_1-i/2}(1-\beta)^{a_2+d_1+j_2-i/2}}:=R.\end{aligned}$$ Utilize to express everything in terms of $i,j_1,j_2,a_2$ to get $$R=\frac{(1-\alpha-\beta+\delta^*)^{i-2 j_1-2j_2}(\alpha-\delta^*)^{2 j_1} (\beta-\delta^*)^{2 j_2}(B_1B_2)^{a_2 - i + 2 j_1 + 2 j_2}}{\alpha^{ j_1}\beta^{ j_2}(1-\alpha)^{\frac{i}{2}- j_1} (1-\beta)^{\frac{i}{2}- j_2} }.$$ Substituting $$\alpha= \frac{Q^{+}(1+B_1 Q^{-})}{B_2+Q^{-}+Q^{+}+B_1 Q^{-} Q^{+}},\ \beta=\frac{Q^{-}(1+B_1Q^{+})}{B_2+Q^{-}+Q^{+}+B_1 Q^{-} Q^{+}},\ \delta^*=\frac{B_1 Q^{-}Q^{+}}{B_2+Q^{-}+Q^{+}+B_1 Q^{-} Q^{+}},$$ it follows that $$R=\frac{B_1^{a_2 - i + 2 j_1 + 2 j_2}B_2^{a_2}(Q^+)^{j_1}(Q^-)^{j_2}}{(1+B_1 Q^{-})^{j_1}(1+B_1 Q^{+})^{j_2}(B_2+Q^-)^{i/2-j_1}(B_2+Q^+)^{i/2-j_2}}.$$ We thus obtain $$\frac{{\mathbb{E}}[Y_{S,T}|\mathbf{1}_{\xi,\zeta}=1]}{{\mathbb{E}}[Y_{S,T}]}\sim x^i y^{j_1} z^{j_2} w^{a_2},$$ where $$x=\frac{1}{B_1 \sqrt{(B_2+Q^-)(B_2+Q^+)}},\ y=\frac{B_1^2 Q^+(B_2+Q^-)}{1+B_1 Q^{-}},\ z=\frac{B^2_1 Q^-(B_2+Q^+)}{1+B_1 Q^{+}},\ w=B_1B_2.$$ Thus, by , we obtain $$\frac{{\mathbb{E}}[YX_i]}{{\mathbb{E}}[Y]}=\frac{r(\Delta,i)}{2i}\cdot x^i\sum_{j_1,j_2,a_2}a_{j_1,j_2,a_2}y^{j_1} z^{j_2} w^{a_2},\label{eq:poissonsum}$$ where $a_{i,j_1,j_2,a_2}$ is the number of possible $\xi$ with $j_1$ red vertices, $j_2$ blue vertices and $a_2$ edges of type $\{W,W\}$. To analyze such sums, an elegant technique (given in [@Janson]) is to define an appropriate transition matrix. The powers of the matrix count the (multiplicative) weight of walks in the implicit graph. In our setting, the transition matrix is given by $$\mathbf{A}=\left(\begin{array}{ccccccc} 0& & y& & 0& & y\\ z& & 0& & z& & 0\\ 0& & 1& & 0& & w\\ 1& & 0& & w& & 0 \end{array}\right).$$ Note that a rooted oriented cycle with specification $\xi,\zeta$ corresponds to a unique closed walk in the implicit graph defined by $\mathbf{A}$. We have that $\mathrm{Tr}(\mathbf{A}^i)=\sum^4_{j=1}e^i_j$, where $e_j$ are the eigenvalues of $\mathbf{A}$ ($j=1,\hdots,4$). We thus obtain $$\frac{{\mathbb{E}}[YX_i]}{{\mathbb{E}}[Y]}=\frac{r(\Delta,i)}{2i}\sum^4_{j=1} (x e_j)^i.$$ A computation gives $$e_1=-e_2=B_1 \left(1-B_1 B_2\right) \sqrt{\frac{Q^+ Q^-}{(1+B_1 Q^+) (1+B_1 Q^-)}},\ e_3=-e_4=B_1 \sqrt{(B_2+Q^+) (B_2+Q^-)}.$$ It follows that for even $i$, we have $$\frac{{\mathbb{E}}[YX_i]}{{\mathbb{E}}[Y]}=\frac{r(\Delta,i)}{i}\left(1+\left(\frac{(1-B_1B_2)\sqrt{Q^+Q^-}}{\sqrt{(1+B_1 Q^+)(1+B_1 Q^-)(B_2+Q^+)(B_2+Q^-)}}\right)^i\right),$$ thus proving the lemma. Using Lemma \[lem:properedgecolorings\], we have $$\sum_{\mbox{\small{even} }i\geq 2}\lambda_i\delta^2_i=\sum_{\mbox{\small{even} }i\geq 2}\frac{r(\Delta,i)}{i}\cdot \omega^i=\sum_{\mbox{\small{even} }i\geq 2}\frac{(\Delta-1)^i+(\Delta-1)}{i}\cdot\omega^i.$$ Observe that $\sum_{j\geq 1}\frac{x^{2j}}{2j}=-\frac{1}{2}\ln(1-x^2)$ for all $|x|<1$. Clearly this holds for $\omega$ and by Lemma \[lem:technicalinequality\], it also holds for $(\Delta-1)\omega$. It follows that $$\sum_{\mbox{\small{even} }i\geq 2}\lambda_i\delta^2_i=-\frac{1}{2}\Big(\ln\big(1-(\Delta-1)^2\omega^2\big)+(\Delta-1)\ln\big(1-\omega^2\big)\Big),$$ thus proving the lemma. Remaining Proofs {#sec:remainingproofs} ================ Proof of Lemma \[lem:technicalinequality\] {#sec:technicalinequality} ------------------------------------------ In this section, we give the proof of Lemma \[lem:technicalinequality\]. We prove only the inequality for $\omega$, since the inequality for $\omega^*$ is the condition for non-uniqueness, see the discussion in Section \[sec:tree-recursions\]. We will in fact establish a slightly stronger inequality, as is captured by the following lemma. To aid the presentation, let us work with $x=Q^+,y=Q^-, d=\Delta-1$. \[lem:aidforproof\] For any $d\geq 2$, it holds that $B_1 xy+B_1B_2(x+y)+B_2> (d-1)(1-B_1B_2)\sqrt{xy}.$ Let $W:=B_1 xy+B_1B_2(x+y)+B_2$. Observe $$W=(B_1x+1)(y+B_2)-(1-B_1B_2)y=(B_1y+1)(x+B_2)-(1-B_1B_2)x.$$ Using the two expressions of $W$ and the AM-GM inequality, we obtain $$\begin{aligned} W^2&=(B_1x+1)(y+B_2)(B_1y+1)(x+B_2)+(1-B_1B_2)^2xy \notag\\ &\qquad -(1-B_1B_2)\big(y(B_1x+1)(B_2+y)+x(B_1y+1)(B_2+x)\big)\notag\\ &\leq\big(\sqrt{(B_1x+1)(B_1y+1)(B_2+x)(B_2+y)}-(1-B_1B_2)\sqrt{xy}\big)^2.\label{eq:mediuminequality}\end{aligned}$$ Trivially $$(B_1x+1)(B_1y+1)(B_2+x)(B_2+y)>xy>(1-B_1B_2)^2 xy,$$ so and Lemma \[lem:aidforproof\] give $$\sqrt{(B_1x+1)(y+B_2)(B_1y+1)(x+B_2)}-(1-B_1B_2)\sqrt{xy}>(d-1)(1-B_1B_2).$$ which after trivial manipulations reduces to the desired inequality. The $x,y$ satisfy $$x=\lambda \left(\frac{B_1 y+1}{y+B_2}\right)^d\mbox{ and } y=\lambda \left(\frac{B_1 x+1}{x+B_2}\right)^d.$$ It follows that $$x\left(\frac{B_1 x+1}{x+B_2}\right)^d=y\left(\frac{B_1 y+1}{y+B_2}\right)^d \Rightarrow x\Big((B_1 x+1)(y+B_2)\Big)^d=y\Big((B_1 y+1)(x+B_2)\Big)^d.\label{eq:startingpoint}$$ Let $W:=B_1 xy+B_1B_2x+B_1B_2y+B_2$. Observe that $$\begin{aligned} (B_1 x+1)(y+B_2)&=B_1 xy+B_1B_2x+y+B_2=W+(1-B_1B_2)y,\\ (B_1 y+1)(x+B_2)&=B_1 xy+B_1B_2y+x+B_2=W+(1-B_1B_2)x.\end{aligned}$$ Hence, gives $$x\left(W+(1-B_1B_2)y\right)^d=y\left(W+(1-B_1B_2)x\right)^d.$$ Expanding using Newton’s formula gives $$x\sum^d_{k=0}\binom{d}{k}W^{d-k}\big((1-B_1B_2)y\big)^{k}=y\sum^d_{k=0}\binom{d}{k}W^{d-k}\big((1-B_1B_2)x\big)^{k},$$ which is equivalent to $$W^d(x-y)=xy\sum^d_{k=2}\binom{d}{k}W^{d-k} (1-B_1B_2)^k\left(x^{k-1}-y^{k-1}\right).$$ Since $x\neq y$, this can be rewritten as $$W^d=xy\sum^d_{k=2}\binom{d}{k}W^{d-k} (1-B_1B_2)^k\left(\frac{x^{k-1}-y^{k-1}}{x-y}\right).\label{eq:startingpointa}$$ \[cl:simplebound\] For $k\geq 2$ and $x,y>0$ with $x\neq y$, it holds that $\displaystyle\frac{x^{k-1}-y^{k-1}}{x-y}> (k-1)(xy)^{(k-2)/2}$. The simple proof of Claim \[cl:simplebound\] is given at the end. Using Claim \[cl:simplebound\], gives $$W^d> \sum^d_{k=2}\binom{d}{k}(k-1)W^{d-k} (1-B_1B_2)^k (xy)^{k/2}, $$ or equivalently $$\underbrace{W^d+ \sum^d_{k=2}\binom{d}{k}W^{d-k}(1-B_1B_2)^k (xy)^{k/2}}_{C}> \underbrace{\sum^d_{k=2}\binom{d}{k}kW^{d-k} (1-B_1B_2)^k (xy)^{k/2}}_{D}.\label{eq:endingpoint}$$ Using again Newton’s formula and the identity $\binom{d}{k}=\frac{d}{k}\binom{d-1}{k-1}$, we have $$\begin{aligned} C&=\big(W+(1-B_1B_2)\sqrt{xy}\big)^d-dW^{d-1}(1-B_1B_2)\sqrt{xy},\\ D&=d(1-B_1B_2)\sqrt{xy}\sum^d_{k=2}\binom{d-1}{k-1}W^{d-k} (1-B_1B_2)^{k-1} (\sqrt{xy})^{k-1}\\ &=d(1-B_1B_2)\sqrt{xy}\Big(\big(W+(1-B_1B_2)\sqrt{xy}\big)^{d-1}-W^{d-1}\Big).\end{aligned}$$ Thus, gives $$W> (d-1)(1-B_1B_2)\sqrt{xy},$$ which is exactly the inequality we wanted. Finally, we give the proof of Claim \[cl:simplebound\]. Since $k\geq 2$, observe that $$\frac{x^{k-1}-y^{k-1}}{x-y}=x^{k-2}+x^{k-3}y+\hdots+y^{k-2}\geq (k-1)\Big((xy)^{(k-1)(k-2)/2}\Big)^{1/(k-1)}=(k-1)(xy)^{(k-2)/2}.$$ The inequality is an application of the AM-GM inequality to $x^{k-2},\hdots,y^{k-2}$. Equality holds iff $x~=~y$. This completes the proof of Lemma \[lem:aidforproof\]. Proofs for Section \[sec:maxentropy\] {#sec:maxentropyproof} ------------------------------------- We first give the proof when $M_{ij}>0$ for all $i,j$. Then it is easy to see that the function $g(\mathbf{Z})$ attains its maximum in the interior of the region , since $$\frac{\partial g}{\partial Z_{ij}}=\ln M_{ij}-\ln Z_{ij}-1\rightarrow +\infty \mbox{ as } Z_{ij}\rightarrow 0^+.$$ Since $g(\mathbf{Z})$ is differentiable in the interior of the region (by $\alpha_i\neq 0\neq \beta_j$), we may formulate the maximization of $g(\mathbf{Z})$ using Lagrange multipliers. Let $$\Lambda(\mathbf{Z},\boldsymbol{\lambda}, \boldsymbol{\lambda'})=g(\mathbf{Z})+\lambda_i(\sum_j Z_{ij}-\alpha_i)+\lambda_j'(\sum_i Z_{ij}-\beta_j).$$ The optimal $\mathbf{Z}$ must satisfy $$\frac{\partial \Lambda}{\partial Z_{ij}}=0 \Longleftrightarrow \ln M_{ij}-\ln Z_{ij}-1+\lambda_i+\lambda_j'\Longleftrightarrow Z_{ij}=\exp\left(\lambda_i+\lambda'_j-1\right)M_{ij}.$$ Setting $R_i=\exp(\lambda_i), C_j=\exp(\lambda'_j-1)$ one obtains that the stationary points of $g(\mathbf{Z})$ have the form stated in Lemma \[lem:gmax\]. The equalities come from the conditions $\frac{\partial \Lambda}{\partial \lambda_i}=\frac{\partial \Lambda}{\partial \lambda_j'}=0$. We next treat the case where some entries of $M_{ij}$ are zero. Recall the convention (Section \[sec:generallog\]) that $\ln 0\equiv -\infty$ and $0\ln 0\equiv 0$. To see why these conventions are relevant, notice that whenever $M_{ij}=0$, $g(\mathbf{Z})$ can have a maximum at $\mathbf{Z}^*$ only if $Z^*_{ij}=0$. Going back to the interpretations of $g(\mathbf{Z})$ in the first and second moments, the reader will immediately notice that these are in complete accordance with the combinatorial meaning of the variables $Z_{ij}$. Let $P=\{(i,j):\ M_{ij}=0\}$. We may restrict our attention for the maximum of $g(\mathbf{Z})$ in the region $$\label{rege1} \begin{array}{ll} \sum_{j} Z_{ij} = \alpha_i(u),& \mbox{for } i\in 1,\hdots, m,\\ \sum_{i} Z_{ij} = \beta_j(v),& \mbox{for } j\in 1,\hdots,n,\\ Z_{ij}\geq 0,& \mbox{for } (i,j)\notin P,\\ Z_{ij}=0& \mbox{for }(i,j)\in P. \end{array}$$ The proof given above (under the assumptions $M_{ij}>0$ for all $i,j$) now goes through for the region , as it can readily be checked. Since $M_{ij}=0$ for $(i,j)\in P$, the form of $\mathbf{Z}^*$ stated in Lemma \[lem:gmax\] is consistent with the region . Finally, the desired quadratic decay is an immediate consequence of the stationarity of $\mathbf{Z}^*$, the strict concavity of $g$ and Taylor’s expansion. The lemma is essentially an application of the envelope theorem, but we give the short proof for the sake of completeness. From  we have $$\begin{aligned} \left(\frac{\partial}{\partial u} R_i\right) \sum_{j=1}^m M_{ij} C_j + R_i \sum_{j=1}^m M_{ij} \frac{\partial}{\partial u} C_j & = & \frac{\partial}{\partial u} \alpha_i,\label{e4A} \\ \left(\frac{\partial}{\partial u} C_j\right) \sum_{i=1}^n M_{ij} R_i + C_j \sum_{i=1}^n M_{ij} \frac{\partial}{\partial u} R_i & = & 0 \label{e4B}.\end{aligned}$$ Then $$\begin{aligned} \frac{\partial g^{*}}{\partial u} = -\sum_{i=1}^n \sum_{j=1}^m M_{ij} \Bigg( & \left(\frac{\partial}{\partial u} R_i\right)(C_j\ln (R_i) + C_j\ln (C_j)) + \left(\frac{\partial}{\partial u} C_j\right)(R_i\ln (R_i) + R_i\ln (C_j)) \\ & + \left(\frac{\partial}{\partial u} R_i\right) C_j + R_i \left(\frac{\partial}{\partial u} C_j\right) \Bigg),\end{aligned}$$ which, using  and , simplifies to $$\begin{aligned} \frac{\partial g^{*}}{\partial u} = - & \sum_{i=1}^n \ln (R_i)\left( \left(\frac{\partial}{\partial u} R_i\right)\sum_{j=1}^m M_{ij} C_j + R_i\left(\sum_{j=1}^m M_{ij}\frac{\partial}{\partial u} C_j \right)\right)\\ & - \sum_{j=1}^m \ln (C_j)\left( \left(\frac{\partial}{\partial u} C_j\right)\sum_{i=1}^n M_{ij} R_i + C_j\left(\sum_{i=1}^n M_{ij}\frac{\partial}{\partial u} R_i \right)\right)\\ & - \sum_{j=1}^m \left(\left(\frac{\partial}{\partial u} C_j\right) \sum_{i=1}^n M_{ij} R_i + C_j \sum_{i=1}^n M_{ij} \frac{\partial}{\partial u} R_i\right)\\ =-& \sum_{i=1}^n \ln(R_i) \frac{\partial}{\partial u}\alpha_i.\end{aligned}$$ Proofs for Section \[sec:logfirstmax\] {#sec:logfirstmaxproof} -------------------------------------- To avoid unnecessary complications, we give the proof for $B_1 B_2>0$. The case with $B_1=0$ (essentially hard-core model) is covered in [@DFJ Claim 2.2]. By and we have $$\frac{\partial \phi_1 }{\partial\alpha}=\ln\left(\lambda \left(\frac{\alpha}{1-\alpha}\right)^{\Delta-1} \left(\frac{R_2}{R_1}\right)^\Delta \right)=\ln\left(\lambda \frac{1-\alpha}{\alpha}\left(\frac{\alpha}{1-\alpha}\frac{R_2}{R_1}\right)^\Delta \right),$$ $$\frac{\partial \phi_1}{\partial\beta}=\ln\left(\lambda \left(\frac{\beta}{1-\beta}\right)^{\Delta-1} \left(\frac{C_2}{C_1}\right)^\Delta \right)=\ln\left(\lambda \frac{1-\beta}{\beta}\left(\frac{\beta}{1-\beta}\frac{C_2}{C_1}\right)^\Delta \right),$$ where the $R_i,C_j$ satisfy . By , we have $$\frac{\alpha}{1-\alpha}=\frac{R_1}{R_2}\cdot\frac{B_1C_1+C_2}{C_1+B_2C_2},\quad \frac{\beta}{1-\beta}=\frac{C_1}{C_2}\cdot\frac{B_1R_1+R_2}{R_1+B_2R_2}.$$ Hence, the derivatives $\displaystyle\frac{\partial \phi_1}{\partial \alpha},\displaystyle \frac{\partial \phi_1}{\partial \beta}$ become $$\frac{\partial \phi_1 }{\partial\alpha}=\ln\left(\lambda \frac{1-\alpha}{\alpha}\left(\frac{B_1C_1+C_2}{C_1+B_2C_2}\right)^\Delta \right),\label{ea1copy}$$ $$\frac{\partial \phi_1}{\partial\beta}=\ln\left(\lambda \frac{1-\beta}{\beta}\left(\frac{B_1R_1+R_2}{R_1+B_2R_2}\right)^\Delta \right).\label{ea2copy}$$ The following claim, which we prove at the end, will be helpful. \[cl:helpfulfir\] When $B_1>0, B_2>0$, the function $$f(x,y):=\frac{B_1x+y}{x+B_2y},$$ defined for $x,y\geq 0$ with $xy\neq 0$ is upper and lower bounded by strictly positive constants depending only on $B_1$, $B_2$. Observe that Claim \[cl:helpfulfir\] is applicable to our case since not all of $R_1,R_2$ or $C_1,C_2$ can be equal to 0 (by equations ). With Claim \[cl:helpfulfir\], we can easily analyze the boundary cases. We consider the boundaries with $\alpha(1-\alpha)\rightarrow 0^+$ and prove that the derivatives with respect to $\alpha$ go the right way. The remaining cases (and their combinations) are completely analogous. We consider cases. 1. $\alpha\rightarrow 0^+$. Then $1-\alpha\rightarrow1$. By Claim \[cl:helpfulfir\] and , we obtain $\frac{\partial \phi_1}{\partial \alpha}\rightarrow +\infty$. 2. $1-\alpha\rightarrow 0^+$. Then $\alpha\rightarrow 1$. By Claim \[cl:helpfulfir\] and , we obtain $\frac{\partial \phi_1}{\partial \alpha}\rightarrow -\infty$. We now give the proof of Claim \[cl:helpfulfir\]. Observe that for $t>0$ it holds that $f(x,y)=f(tx,ty)$, so that we may assume $x+y=1$. Thus, $B_1x+y$ and $x+B_2 y$ are convex combinations of $\{1,B_1\}$ and $\{1,B_2\}$ respectively. Since we assumed that $B_1,B_2>0$, $f(x,y)$ is bounded by strictly positive constants. This concludes the proof of Lemma \[lem:firboundary\]. Denote by $\mathbf{H_1}$ the Hessian of $\Phi_1$ evaluated at $(p^+,p^-,X^*)$. We keep only $\alpha,\beta,x_{11}$ as free variables in $\Phi_1$ (the rest of the $x_{ij}$ are substituted by ). We check that the principal minors along the diagonal are positive, the first principal minor being the lower right corner of the matrix. This is equivalent to applying Sylvester’s criterion in a rearranged version of $-\mathbf{H_1}$. We use Maple to perform the necessary computations and plug into the resulting formulas the values of $p^\pm,\mathbf{X}^*$ as given in and respectively. Denote by $P_i$ the determinant of the $i$-th leading principal minor. We obtain $$\begin{aligned} P_1=\frac{\Delta E_1 E_2}{B_1 B_2 Q^- Q^+},\ P_2=\frac{\Delta E_1^2 (B_2+Q^-) (1+B_1 Q^-)\big(1+(\Delta-1)\omega\big)}{B_1 B_2 (Q^-)^2 Q^+},\ P_3=\frac{\Delta E^4_1\big(1-(\Delta-1)^2\omega\big)}{B_1B_2\big(Q^+Q^-\big)^2},\end{aligned}$$ where $E_1,E_2$ and $\omega$ are given by and . Trivially $P_1, P_2>0$. The fact that $P_3>0$ is just Lemma \[lem:technicalinequality\]. To prove that $(p^*,p^*,X^o)$ is a saddle point, observe that the respective expressions for $P_1,P_2,P_3$ can be obtained by substituting $Q^-\leftarrow Q^*, Q^+\leftarrow Q^*$, where $Q^*=\frac{q^*}{1-q^*}$. We have that $P_3<0$ by Lemma \[lem:technicalinequality\]. Proofs for Section \[sec:logsecmax\] {#sec:logsecmaxproof} ------------------------------------ To avoid unnecessary complications, we give the proof for $B_1 B_2>0$. The case with $B_1=0$ (essentially hard-core model) is covered in [@Galanis Lemma 12]. By equations and , we have $$\frac{\partial \phi_2}{\partial \gamma} = \ln\left( \left(\frac{\gamma(1-2\alpha+\gamma)}{(\alpha-\gamma)^2}\right)^{\Delta-1}\left(\frac{R_2^2}{R_1 R_4}\right)^\Delta \right)=\ln\left(\frac{(\alpha-\gamma)^2}{\gamma(1-2\alpha+\gamma)}\left(\frac{\gamma(1-2\alpha+\gamma)}{(\alpha-\gamma)^2}\frac{R_2^2}{R_1 R_4}\right)^\Delta \right),$$ $$\frac{\partial \phi_2}{\partial \delta} = \ln\left( \left(\frac{\delta(1-2\beta+\delta)}{(\beta-\delta)^2}\right)^{\Delta-1}\left(\frac{C_2^2}{C_1 C_4}\right)^\Delta \right)=\ln\left(\frac{(\beta-\delta)^2}{\delta(1-2\beta+\delta)}\left(\frac{\delta(1-2\beta+\delta)}{(\beta-\delta)^2}\frac{C_2^2}{C_1 C_4}\right)^\Delta \right),$$ where the $R_i,C_j$ satisfy . By , we have $$\frac{\gamma(1-2\alpha+\gamma)}{(\alpha-\gamma)^2}=\frac{R_1 R_4}{R^2_2}\cdot \frac{(B^2_1 C_1+2B_1 C_2+C_4)(C_1+2B_2 C_2+B^2_2 C_4)}{(B_1 C_1+(B_1B_2+1)C_2+B_2C_4)^2},$$ $$\frac{\delta(1-2\beta+\delta)}{(\beta-\delta)^2}=\frac{C_1 C_4}{C^2_2}\cdot \frac{(B^2_1 R_1+2B_1 R_2+R_4)(R_1+2B_2 R_2+B^2_2 R_4)}{(B_1 R_1+(B_1B_2+1)R_2+B_2 R_4)^2},$$ Hence, the derivatives $\displaystyle\frac{\partial \phi_2}{\partial \gamma},\displaystyle \frac{\partial \phi_2}{\partial \delta}$ become $$\label{eq:dergammacopy} \frac{\partial \phi_2}{\partial \gamma} = \ln\left(\frac{(\alpha-\gamma)^2}{\gamma(1-2\alpha+\gamma)}\left( \frac{(B^2_1 C_1+2B_1 C_2+C_4)(C_1+2B_2 C_2+B^2_2 C_4)}{(B_1 C_1+(B_1B_2+1)C_2+B_2C_4)^2} \right)^\Delta\right),$$ $$\label{eq:derdeltacopy} \frac{\partial \phi_2}{\partial \delta} = \ln\left(\frac{(\beta-\delta)^2}{\delta(1-2\beta+\delta)}\left( \frac{(B^2_1 R_1+2B_1 R_2+R_4)(R_1+2B_2 R_2+B^2_2 R_4)}{(B_1 R_1+(B_1B_2+1)R_2+B_2R_4)^2} \right)^\Delta\right).$$ Before proceeding we need the following claim, which we prove at the end. \[cl:veryhelpful\] When $B_1>0, B_2>0$, the function $$f(x,y,z):=\frac{(B^2_1 x+2B_1 y+z)(x+2B_2 y+B^2_2 z)}{(B_1 x+(B_1B_2+1)y+B_2z)^2},$$ defined for $x,y,z\geq 0$ with $xyz\neq 0$ is upper and lower bounded by strictly positive constants depending only on $B_1$, $B_2$. Observe that Claim \[cl:veryhelpful\] is applicable to our case since not all of $R_1,R_2,R_4$ or $C_1,C_2, C_4$ can be equal to 0 (by equations ). With Claim \[cl:veryhelpful\] at hand, it is immediate to treat the boundary cases. We consider the boundaries with $\gamma(\alpha-\gamma)(1-2\alpha+\gamma)\rightarrow 0^+$ and prove that the derivatives with respect to $\gamma$ go the right way. The other cases (and their combinations) are completely analogous. We consider cases. 1. $\gamma\rightarrow 0^+$. This case may only occur if $1-2\alpha\geq 0$. Then $\alpha-\gamma\rightarrow \alpha>0$ and $1-2\alpha+\gamma\rightarrow 1-2\alpha\geq 0$. By Claim \[cl:veryhelpful\] and , we obtain $\frac{\partial \phi_2}{\partial \gamma}\rightarrow +\infty$. 2. $1-2\alpha+\gamma\rightarrow 0^+$. This case may only occur if $2\alpha-1\geq 0$. Then $\gamma\rightarrow 2\alpha-1\geq 0$ and $\alpha-\gamma\rightarrow 1-\alpha>0$. By Claim \[cl:veryhelpful\] and , we obtain $\frac{\partial \phi_2}{\partial \gamma}\rightarrow +\infty$. 3. $\alpha-\gamma\rightarrow 0^+$. Then $\gamma\rightarrow\alpha>0$ and $1-2\alpha+\gamma\rightarrow 1-\alpha>0$. By Claim \[cl:veryhelpful\] and , we obtain $\frac{\partial \phi_2}{\partial \gamma}\rightarrow -\infty$. We now give the proof of Claim \[cl:veryhelpful\]. Observe that for $t>0$ it holds that $f(x,y,z)=f(tx,ty,tz)$, so we may assume that $x+y+z=1$. Thus, $B^2_1 x+2B_1 y+z$, $x+2B_2 y+B^2_2 z$ and $B_1x+(B_1B_2+1)y+B_2z$ are convex combinations of $\{B^2_1,2B_1,1\},\{1,2B_2,B^2_2\}$ and $\{B_1,B_1B_2+1,B_2\}$ respectively. Since we assumed that $B_1,B_2>0$, $f(x,y,z)$ is bounded by strictly positive constants. This concludes the proof of Lemma \[lem:secboundary\]. For $(\alpha,\beta)=(p^+,p^-)$, denote by $\mathbf{H_2}$ the Hessian of $\Phi_2$ evaluated at $(\alpha^2,\beta^2,\mathbf{Y}^*)$. We keep only $\gamma,\delta,y_{ij}$ with $i,j=1,2,3$ as free variables in $\Phi_2$ (the rest of the $y_{ij}$ are substituted by ). We use Sylvester’s criterion to check that $-\mathbf{H_2}$ is positive definite. We check that the principal minors are positive, the first principal minor being the lower right corner of the matrix. This is equivalent to applying Sylvester’s criterion in a rearranged version of $-\mathbf{H_2}$. Note that the first nine principal minors are guaranteed to be positive since they correspond to the Hessian of the convex function $-g_2(\mathbf{Y})$ evaluated at its minimum point. Thus we need to check only the remaining two principal minors. We use Maple to perform the necessary computations and we plug into the resulting formulas the values of $p^\pm$ from and the $\mathbf{Y}^*$ as given in . Denote by $P_i$ the determinant of the $i$-th leading principal minor. We have: $$\begin{aligned} P_{10}&=\frac{\Delta^9 E^{22}_1E^2_2(B_2+Q^-)^2(1+B_1 Q^-)^2\big(1+(\Delta-1)\omega^2\big)}{(B_1 B2)^8\big(Q^-\big)^{14}\big(Q^{+}\big)^{12}},\\ P_{11}&=\mathrm{det}(-\mathbf{H_2})=\frac{\Delta^9 E_1^{26}E^2_2\big(1-(\Delta-1)^2\omega^2\big)}{(B_1 B2)^8\big(Q^-\big)^{14}\big(Q^{+}\big)^{14}},\end{aligned}$$ where $E_1,E_2$ and $\omega$ are given by and . Trivially $P_{10}>0$. And we also have $P_{11}>0$ by Lemma \[lem:technicalinequality\]. Proof of Lemma \[lem:secondmaxising\] {#sec:secondmaxisingproof} ------------------------------------- We now prove Condition \[cond:maxima\] for the Ising model without external field and for $d=2$, that is $B_1=B_2=B, \lambda=1$ and $\Delta=3$. See Section \[sec:remarks\] for the proof overview. Our point of departure is . As in the proof of Lemma \[lem:secondmax2spin\], we may assume that $r_1r_4,c_1c_4$ are either both greater than 1 or both are less than 1. We are going to argue that cannot hold in these cases. For $B_1=B_2=B$, multiplying the equations in gives $$\label{eq:stable} (Bc_1+(B^2+1)+Bc_4)^2(Br_1+(B^2+1)+Br_4)^2=\frac{(1-B^2)^2(c_1c_4-1)}{(c_1c_4)^{1/2}-1}\frac{(1-B^2)^2(r_1r_4-1)}{(r_1r_4)^{1/2}-1}.$$ We will prove that for $0<B<\frac{\sqrt{2}-1}{\sqrt{2}+1}$, when $r_1 r_4\neq 1$ (and similarly when $c_1 c_4\neq 1$) $$\big(Br_1+(B^2+1)+Br_4\big)^2>\frac{(1-B^2)^2(r_1r_4-1)}{\sqrt{r_1r_4}-1}.\label{ineq:onea}$$ In the region $0<B<\frac{\sqrt{2}-1}{\sqrt{2}+1}$ the values of $\alpha$ and $\beta$ are heavily biased towards 1 and 0 respectively, and this reflects at the values of $r_1,r_4,c_1,c_4$. The required bias of the values of $\alpha,\beta$ is captured by the following lemma. \[lem:biasdtwo\] Assume $0<B<\displaystyle\frac{\sqrt{2}-1}{\sqrt{2}+1}$. For $(\alpha,\beta)=(p^+,p^-)$ it holds that $\displaystyle\frac{\alpha}{1-\alpha}=\displaystyle\frac{1-\beta}{\beta}>\displaystyle\frac{4}{9}\cdot \frac{1}{B^3}$. The proof of Lemma \[lem:biasdtwo\] is given in Section  \[sec:biasdtwo\]. We utilize Lemma \[lem:biasdtwo\] to prove: \[lem:biasedvaluesdtwo\] Assume $0<B<\frac{\sqrt{2}-1}{\sqrt{2}+1}$. For $(\alpha,\beta)=(p^+,p^-)$ and $r_1,r_4,c_1,c_4$ satisfying the equations in , , it holds that - $r_1>\displaystyle\frac{1}{3}\cdot\frac{1}{B^{2}}$ and $c_4>\displaystyle\frac{1}{3}\cdot\displaystyle\frac{1}{B^{2}}$. - If in addition $r_1 r_4,c_1 c_4>1$, then $r_1>\displaystyle\frac{4}{9}\cdot\frac{1}{B^{2}}$ and $c_4>\displaystyle\frac{4}{9}\cdot\displaystyle\frac{1}{B^{2}}$. The proof of Lemma \[lem:biasedvaluesdtwo\] is given in Section \[sec:biasedvaluesdtwo\]. We now prove inequality . Rewrite as $$B^2(r_1+r_4)^2+2B(B^2+1)(r_1+r_4)+(B^2+1)^2>(1-B^2)^2\left(1+\sqrt{r_1r_4}\right),$$ or $$B^2(r_1+r_4)^2+2B(B^2+1)(r_1+r_4)+4B^2>(1-B^2)^2\sqrt{r_1r_4}.\label{ineq:equivdtwo}$$ At this point, we split the analysis into the cases $r_1r_4>1$ and $r_1r_4<1$, the first being considerably harder. \[claim:proofdtwoa\] Inequality holds when $r_1r_4<1$. We prove the stricter inequality (since $1>B>0$ and $r_1,r_4>0$) $$2B r_1>\sqrt{r_1r_4}.$$ By item 1 of Lemma \[lem:biasedvaluesdtwo\] and the assumptions, we have $$2Br_1>\frac{2}{3}\cdot\frac{1}{B}>1>\sqrt{r_1 r_4}.$$ \[claim:proofdtwo\] Inequality holds when $r_1r_4>1$. We prove the stricter inequality (since $1>B>0$ and $r_1,r_4>0$) $$B^2(r_1+r_4)^2>\sqrt{r_1r_4}.\label{ineq:maindtwo}$$ We first identify the regions where is hard to prove. Note that $B^2(r^2_1+2r_1r_4+r^2_4)\geq4B^2r_1r_4$, so that is true if $4B^2r_1r_4>\sqrt{r_1r_4}$. Thus we may assume that $$\frac{1}{16B^4}\geq r_1 r_4.\label{rprodupper}$$ Since $r_1r_4>1$, by item 2 of Lemma \[lem:biasedvaluesdtwo\] $$B^2 r_1^2>\frac{4^2}{9^2B^{2}},\label{ineq:ronelow}$$ so that is true if $\frac{4^2}{9^2B^{2}}\geq\sqrt{r_1r_4}$. Thus we may assume that $$r_1r_4>\frac{4^4}{9^4B^4}.\label{rprodlower}$$ We are now ready to prove . Using , and we obtain $$\begin{aligned} B^2(r^2_1+2r_1r_4+r^2_4)\geq B^2(r^2_1+2r_1r_4)>\frac{1}{B^2}\left(\frac{4^2}{9^2}+2\cdot \frac{4^4}{9^4}\right)>\frac{1}{B^2}\cdot \frac{1}{4}\geq \sqrt{r_1r_4}.\end{aligned}$$ ### Proof of Lemma \[lem:biasdtwo\] {#sec:biasdtwo} By the remarks in the end of Section \[sec:tree-recursions\], for the Ising model with no external field it holds that $\frac{\alpha}{1-\alpha}=\frac{1-\beta}{\beta}$ and $Q^+Q^-=1$. For ease of presentation, let $x=Q^+, y=Q^-$. By and , we have $$\frac{\alpha}{1-\alpha}=\frac{x(By+1)}{y+B},\mbox{ where } x=\left(\frac{By+1}{y+B}\right)^2 \mbox{ and } x>1>y>0.$$ Since $xy=1$, we have that $$\label{eq:algebraicbound} \frac{\alpha}{1-\alpha}=\frac{By+1}{y(y+B)},\mbox{ where } \frac{1}{y^3}=\left(\frac{By+1}{y(y+B)}\right)^2\mbox{ and } 0<y<1.$$ For $0<B<\frac{\sqrt{2}-1}{\sqrt{2}+1}=3-2\sqrt{2}$, we want to prove that $\displaystyle\frac{\alpha}{1-\alpha}>\displaystyle\frac{4}{9}\cdot\frac{1}{B^3}$, which in light of , is equivalent to $y<(3/2)\sqrt[3]{3/2} B^2$. For convenience, let $c:= (3/2)\sqrt[3]{3/2}$. The equation for $y$ in may be rewritten as $$\begin{array}{lcl} (y+B)^2=y(By+1)^2&\Rightarrow &B^2 y^3+(-1+2 B) y^2+(1-2 B) y -B^2=0\Rightarrow\\[0.2cm] (y-1)\big(B^2 y^2+(B^2+2B-1) y+B^2\big)=0& \Rightarrow & B^2 y^2+(B^2+2B-1) y+B^2=0, \end{array}$$ since by we are interested in $0<y<1$. Let $f(z):= B^2 z^2+(B^2+2B-1) z+B^2$. By Lemma \[eq:densitiesone\], we know that $y$ is the unique root of $f(z)$ for $0<z<1$. Observe that $f(0)=B^2>0$, so the desired inequality will follow if we prove that $f(B^2c)<0$. Now, $f$ may be rewritten as $$f(z)=B^2z\left(z+\frac{1}{z}-\frac{1-2 B - B^2}{B^2}\right).$$ For $z=B^2c$ and $0<B<3-2\sqrt{2}$, we have $$\begin{gathered} z+\frac{1}{z}=\frac{B^4 c^2+1}{B^2c}\leq \frac{(3-2\sqrt{2})^4c^2+1}{c}\cdot \frac{1}{B^2},\\ \frac{1-2 B - B^2}{B^2}\geq \frac{1-2(3-2\sqrt{2})-(3-2\sqrt{2})^2}{B^2}=\frac{16\sqrt{2}-22}{B^2}.\end{gathered}$$ A straightforward calculation gives $\frac{(3-2\sqrt{2})^4c^2+1}{c}<\frac{3}{5}<16\sqrt{2}-22$, thus proving $f(B^2c)<0$. The lemma follows. ### Proof of Lemma \[lem:biasedvaluesdtwo\] {#sec:biasedvaluesdtwo} We need the following lemma. \[lem:boundsalphabetaRC\] When $B_1B_2>0$ and $r_1,r_4,c_1,c_4$ satisfy and , it holds that $$\frac{\alpha}{1-\alpha}<\frac{1+r_1/B_2}{1+B_2r_4}, \qquad \frac{1+B_1c_1}{1+c_4/B_1}<\frac{\beta}{1-\beta}$$ We first give the proof of Lemma \[lem:biasedvaluesdtwo\] and then give the proof of Lemma \[lem:boundsalphabetaRC\]. When $B_1=B_2=B$, Lemma \[lem:boundsalphabetaRC\] gives $$\frac{\alpha}{1-\alpha}<\frac{1+r_1/B}{1+Br_4},\qquad \frac{1+Bc_1}{1+c_4/B}<\frac{\beta}{1-\beta}.$$ Since $r_4,c_1>0$, it follows that $$r_1>B\left(\frac{\alpha}{1-\alpha}-1\right)>B\left(\frac{4}{9}\cdot \frac{1}{B^3}-1\right) \mbox{ and } c_4>B\left(\frac{1-\beta}{\beta}-1\right)>B\left(\frac{4}{9}\cdot\frac{1}{B^3}-1\right),$$ where we have used Lemma \[lem:biasdtwo\] to bound $\frac{\alpha}{1-\alpha}$ and $\frac{1-\beta}{\beta}$. This proves the first item of the lemma, after observing that $\frac{4}{9}\cdot \frac{1}{B^3}-1>\frac{1}{3}\cdot \frac{1}{B^3}$ for any $B<1/3$. To prove the second item of the Lemma, note that when $r_1r_4,c_1c_4>1$ it holds $$\frac{1+r_1/B}{1+Br_4}<\frac{1+r_1/B}{1+B/r_1}=\frac{r_1}{B}\mbox{ and } \frac{1+Bc_1}{1+c_4/B}>\frac{1+B/c_4}{1+c_4/B}=\frac{B}{ c_4}.$$ It follows that $$r_1>B\cdot \frac{\alpha}{1-\alpha} \mbox{ and } c_4>B\cdot \frac{1-\beta}{\beta}.$$ The desired bounds now follow after using Lemma \[lem:biasdtwo\] to bound $\frac{\alpha}{1-\alpha}$ and $\frac{1-\beta}{\beta}$. It will be convenient to have the following variables at hand. $$\begin{array}{lll} C_1=\displaystyle\frac{B^2_1c_1+2B_1+c_4}{B_1c_1+(B_1B_2+1)+B_2c_4}, & & C_4=\displaystyle\frac{c_1+2B_2+B^2_2c_4}{B_1c_1+(B_1B_2+1)+B_2c_4},\\ & &\\ R_1=\displaystyle\frac{B^2_1r_1+2B_1+r_4}{B_1r_1+(B_1B_2+1)+B_2r_4}, & & R_4=\displaystyle\frac{r_1+2B_2+B^2_2r_4}{B_1c_1+(B_1B_2+1)+B_2c_4}. \end{array}$$ Note that the $C_1,C_4,R_1,R_4$ do not have anything to do with the $R_i,C_j$ in the proof of Lemma \[lem:seccritical\], but this notation is convenient since the equations in , may be written as $$\begin{gathered} r_1C_1=\frac{\gamma}{\alpha-\gamma},\qquad r_4C_4=\frac{1-2\alpha+\gamma}{\alpha-\gamma},\label{eq:rs}\\ c_1R_1=\frac{\delta}{\beta-\delta},\qquad c_4R_4=\frac{1-2\beta+\delta}{\beta-\delta}.\label{eq:cs}\end{gathered}$$ Eliminating $\gamma,\delta$ from and respectively, we obtain $$\frac{\alpha}{1-\alpha}=\frac{1+r_1C_1}{1+r_4C_4},\qquad \frac{\beta}{1-\beta}=\frac{1+c_1R_1}{1+c_4R_4}.\label{eq:mainrcRC}$$ We have the following loose bounds on $R_1,R_4,C_1,C_4$. $$\label{cl:boundRC} B_1<R_1, \qquad R_4<\displaystyle\frac{1}{B_1},\qquad C_1<\displaystyle\frac{1}{B_2}, \qquad B_2<C_4.$$ These bounds may be proved by picking any of the inequalities, multiplying out and using $B_1B_2<1$. It is immediate now to check that the bounds in Lemma \[lem:boundsalphabetaRC\] follow after combining and . Proof of Lemma \[lem:secondmaxhardcore\] {#sec:hardhard} ---------------------------------------- For the hard-core model (that is, $B_{1}=0$, $B_{2}=1$) for $\Delta=3,4,5$ and $(\alpha,\beta)=(p^+,p^-)$ the solution of , and satisfies $\gamma=\alpha^2$ and $\delta=\beta^2$. Let $d:=\Delta-1$ and let $c_1:=C_1/C_2$, $c_4:=C_4/C_2$, $r_1:=R_1/R_2$, $r_4:=R_4/R_2$ (the same notation as we used in the proof of Lemma \[lem:seccritical\]). Equations in , for $B_1=0$ and $B_2=1$, become $$\label{ettt} (c_1c_4)^{1/d}-1=\frac{r_1r_4-1}{(1+r_4)^2}\quad\mbox{and}\quad (r_1r_4)^{1/d}-1=\frac{c_1c_4-1}{(1+c_4)^2}.$$ Let $x=(r_1r_4)^{1/d}$, $y=(c_1c_4)^{1/d}$, $a=1/r_4$, and $b=1/c_4$. We can rewrite equation  as follows $$\label{xxxyy} y=\frac{1+2a+a^2 x^d}{1+2a+a^2}\quad\mbox{and}\quad x=\frac{1+2b+b^2 y^d}{1+2b+b^2}.$$ Solving for $a$ and $b$ we obtain $$\label{barn} a=\frac{y-1 + q_a}{x^d-y}\quad\mbox{and}\quad b=\frac{x-1 + q_b}{y^d-x},$$ where $$\label{barn2} q_a^2=(y-1)(x^d-1)\quad\mbox{and}\quad q_b^2 =(x-1)(y^d-1).$$ Note that $a>0$ and $b>0$. Assume $y>1$. Then $x>1$ (otherwise the sides of  would have different signs) and also $y\leq x^d$ (the right hand side of is a convex combination of $1$ and $x^d$) and hence $|q_a|>y-1$. Thus, using $a>0$ in Equation , we obtain $q_a>0$ and similarly $q_b>0$. Now assume $y<1$. Then $x<1$ and also $y\geq x^d$ (again, the right hand side of is a convex combination of $1$ and $x^d$) and hence $|q_a|>1-y$. Thus, using $a>0$ in Equation , we obtain $q_a<0$ and similarly $q_b<0$. Because of symmetry between $x$ and $y$ we only need to consider two cases (in the case $x=y=1$ we have $\gamma=\alpha^2$ and $\delta=\beta^2$): - CASE 1: $1<y\leq x\leq y^d$, $q_a:=\sqrt{(y-1)(x^d-1)}$, $q_b:=\sqrt{(x-1)(y^d-1)}$, and - CASE 2: $0<y^d\leq x\leq y<1$, $q_a:=-\sqrt{(y-1)(x^d-1)}$, $q_b:=-\sqrt{(x-1)(y^d-1)}$. From Equation  one obtains the following expressions for $\alpha$ and $\beta$ (see equation ) $$\frac{\alpha}{1-\alpha}=\frac{\frac{c_4}{1+c_4}r_1+1}{\frac{c_1+2+c_4}{1+c_4}r_4+1}\quad\mbox{and}\quad \frac{\beta}{1-\beta}=\frac{\frac{r_4}{1+r_4}c_1+1}{\frac{r_1+2+r_4}{1+r_4}c_4+1}.$$ Solving for $\alpha,\beta$ and using the parametrization with $x,y,a,b$ we have $$\label{alphabeta2} \begin{array}{l} \alpha=\displaystyle\frac{y(1+a)^2-1-a+ab}{x(1+b)^2+y(1+a)^2+2ab-1},\vspace{0.2cm}\\ \beta=\displaystyle\frac{x(1+b)^2-1-b+ab}{x(1+b)^2+y(1+a)^2+2ab-1}. \end{array}$$ Equations , for $B_1=0,\,B_2=1,\, (\alpha,\beta)=(p^+,p^-)$ easily imply $$\label{hardcore3} \alpha (1-\alpha)^d = \beta (1-\beta)^d.$$ Equation  is equivalent to (assuming $\alpha\neq\beta$) $$\label{hardcore2} \frac{\alpha (1-\alpha)^d - \beta (1-\beta)^d}{\alpha-\beta}=0,$$ which using the expressions for $\alpha$ and $\beta$ from  becomes $$\label{eeer} \textstyle\frac{(y(1+a)^2-1-a+ab)(x(1+b)^2+ab+a)^d - (x(1+b)^2-1-b+ab)(y(1+a)^2+ab+b)^d)}{x(b+1)^2-y(a+1)^2+a-b}=0.$$ The final part of the proof will require some computational assistance (just to manipulate polynomials). First we are going to plug-in the expression for $a$ and $b$ from  into . Then we are going to use the expressions in  to reduce the powers of $q_a$ and $q_b$ occurring in the expression obtaining an expression of the form $$c_{00} + c_{10} q_a + c_{01} q_b + c_{11} q_a q_b,$$ where $c_{00},c_{01},c_{10},c_{11}$ are polynomials in $x$ and $y$. Then we are going to show that in CASE 1 we have $c_{00}>0$, $c_{01}>0$, $c_{10}>0$, $c_{11}>0$ and in CASE 2 we have $c_{00}>0$, $c_{01}<0$, $c_{10}<0$, $c_{11}>0$ (and hence  cannot be zero) . This will be accomplished by reparameterizing $x=(ty + y^d)/(t+1)$, factoring the expressions, and observing that: - factor $y-1$ occurs with even power in $c_{00}$ and $c_{11}$, - factor $y-1$ occurs with odd power in $c_{01}$ and $c_{10}$, and - all the other factors have all coefficients positive. The details of the argument appear in Appendix \[sec:appendixhardcore\]. [99]{} D. Achlioptas and A. Naor. The two possible values of the chromatic number of a random graph. [*Annals of Mathematics*]{}, 162(3):1333–1349, 2005. D. Achlioptas and Y. Peres. The threshold for random k-SAT is $2^k\log{2} - O(k)$. [*Journal of the AMS*]{}, 17(4):947-973, 2004. V. Beffara and H. Duminil-Copin. The self-dual point of the two-dimensional random cluster model is critical for $q\geq 1$. To appear in [*Probability Theory and Related Fields*]{}. Preprint available from the arXiv at: <http://arxiv.org/abs/1006.5073>. N. G. de Bruijn. [*Asymptotic Methods in Analysis*]{}, Dover, New York, 1981. A. Dembo, A. Montanari, and N. Sun. Factor models on locally tree-like graphs. Preprint, 2011. Available from the arXiv at: <http://arxiv.org/abs/1110.4821> M. Dyer, A. M. Frieze, and M. Jerrum. On counting independent sets in sparse graphs. [*SIAM Journal on Computing*]{}, 31(5):1527-1541, 2002. A. Galanis, Q. Ge, D. Štefankovič, E. Vigoda, and L. Yang. Improved Inapproximability Results for Counting Independent Sets in the Hard-Core Model. In [*Proceedings of the 15th International Workshop, RANDOM*]{}, 567-578, 2011. A. Galanis, D. Štefankovič, and E. Vigoda. Inapproximability of the Partition Function for the Antiferromagnetic Ising and Hard-Core Models. Preprint, 2012. Available from the arXiv at: <http://arxiv.org/abs/1203.2226> H.-O. Georgii. [*Gibbs Measures and Phase Transitions*]{}, volume 9 of de Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, 1988. L.A. Goldberg, M. Jerrum and M. Paterson. The computational complexity of two-state spin systems. [*Random Structures and Algorithms*]{}, 23(2):133-154, 2003. C. Greenhill. The complexity of counting colorings and independent sets in sparse graphs and hypergraphs. [*Computational Complexity*]{}, 9(1):52–72, 2000. S. Janson. Random Regular Graphs: Asymptotic Distributions and Contiguity. [*Combinatorics, Probability [&]{} Computing*]{}, 4:369-405, 1995. S. Janson, T. [Ł]{}uczak, and A. Rucinski. [*Random Graphs*]{}. Wiley-Interscience, New York, 2000. M. Jerrum and A. Sinclair. Polynomial-time Approximation Algorithms for the Ising Model. [*SIAM Journal on Computing*]{}, 22(5):1087-1116, 1993. F. P. Kelly. Loss Networks. [*Annals of Applied Probability*]{}, 1(3):319-378, 1991. L. Li, P. Lu, Y. Yin. Correlation Decay up to Uniqueness in Spin Systems. In [*Proceedings of the 24th Annual ACM-SIAM Symposium on Discrete Algorithms*]{} (SODA), 47-66, 2013. F. Martinelli, A. Sinclair, and D. Weitz. The Ising Model on Trees: Boundary Conditions and Mixing Time. In [*Proceedings of the 43rd Annual IEEE Symposium on Foundations of Computer Science*]{} (FOCS), 628-639, 2003. M. Mezard and A. Montanari. [*Information, Physics, and Computation*]{}, Oxford University Press, USA, 2009. M. Molloy, H. Robalewska, R. W. Robinson, and N. C. Wormald. 1-Factorizations of random regular graphs. [*Random Structures and Algorithms*]{}, 10(3):305-321, 1997. E. Mossel, Survey: Information flow on trees. In [*Graphs, Morphisms, and Statistical Physics*]{}. [*DIMACS Series in Discrete Mathematics and Theoretical Computer Science*]{}, 63:155-170, 2004. E. Mossel, D. Weitz, and N. Wormald. On the hardness of sampling independent sets beyond the tree threshold. [*Probability Theory and Related Fields*]{}, 143(3-4):401-439, 2009. R. Restrepo, J. Shin, P. Tetali, E. Vigoda, and L. Yang. Improved mixing condition on the grid for counting and sampling independent sets. In [*Proceedings of the 52nd Annual IEEE Symposium on Foundations of Computer Science*]{} (FOCS), 140-149, 2011. A. Sinclair, P. Srivastava, and M. Thurley. Approximation algorithms for two-state anti-ferromagnetic spin systems on bounded degree graphs. In [*Proceedings of the 23rd Annual ACM-SIAM Symposium on Discrete Algorithms*]{} (SODA), 941-953, 2012. A. Sly. Computational Transition at the Uniqueness Threshold. In [*Proceedings of the 51st Annual IEEE Symposium on Foundations of Computer Science*]{} (FOCS), 287-296, 2010. A. Sly and N. Sun. The Computational Hardness of Counting in Two-Spin Models on d-Regular Graphs. In [*Proceedings of the 53rd Annual IEEE Symposium on Foundations of Computer Science*]{} (FOCS), 361-369, 2012.. Full version available from the arXiv at: <http://arxiv.org/abs/1203.2602> M. Talagrand. [*Spin Glasses: A Challenge for Mathematicians*]{}, Springer, Berlin, 2003. L. G. Valiant. The Complexity of Enumeration and Reliability Problems. [*SIAM Journal on Computing*]{}, 8(3):410-421, 1979. D. Weitz. . In [*Proceedings of the 38th Annual ACM Symposium on Theory of Computing*]{} (STOC), 140–149, 2006. N.C. Wormald. Models of random regular graphs. In [*Surveys in Combinatorics, 1999 (Canterbury)*]{}, J.D. Lamb and D.A. Preece (eds), pp. 239-298. Cambridge University Press, Cambridge, 1999. J. Zhang, H. Liang, and F. Bai. Approximating Partition Functions of Two-State Spin Systems. [*Information Processing Letters*]{}, 111(14):702-710, 2011. Computer Assisted Proofs for the Hardcore Model {#sec:appendixhardcore} =============================================== Case $\Delta=3$ --------------- d = 2; F = Factor[((y*(1 + a)^2 - 1 - a + a*b)*(x*(1 + b)^2 + a*b + a)^d - (x*(1 + b)^2 - 1 - b + a*b)*(y*(1 + a)^2 + a*b + b)^d)][[3]]; [**Comment:**]{} now $F$ contains the left-hand side of (we took the third factor; the other two factors are $-1$ and the denominator of ). H = Factor[Expand[F /. {a -> (-1 + y + Qa)/(x^d - y), b -> (-1 + x + Qb)/(y^d - x)}]][[3]]; [**Comment:**]{} now $H$ contains the left-hand side of after substituting the values of $a,b$ given by  (the result has three factors: $1/(x-y^d)^{2d-1}$, $1/(x^d-y)^{2d-1}$, and the factor we assigned to $H$; note that the first two factors cannot have value zero). da = Exponent[H, Qa]; For [i = da, i >= 2, i--, H = Expand[H /. {Qa^i -> Qa^(i - 2)*(1 - x^d - y + x^d*y)}]]; db = Exponent[H, Qb]; For [i = db, i >= 2, i--, H = Expand[H /. {Qb^i -> Qb^(i - 2)*(1 - y^d - x + y^d*x)}]]; [**Comment:**]{} now $H$ contains the left-hand side of (multiplied by $(x-y^d)^{2d-1}(x^d-y)^{2d-1}$) after reducing the powers of $q_a$ and $q_b$ using . c00 = Factor[Coefficient[Coefficient[H, Qa, 0], Qb, 0]]; c01 = Factor[Coefficient[Coefficient[H, Qa, 0], Qb, 1]]; c10 = Factor[Coefficient[Coefficient[H, Qa, 1], Qb, 0]]; c11 = Factor[Coefficient[Coefficient[H, Qa, 1], Qb, 1]]; [**Comment:**]{} the following reparameterization will reveal the signs of $c_{00}, c_{01}, c_{10}, c_{11}$. We show the expressions for $d=2$; for larger $d$ we will print the first $6$ factors and check the positivity of the last factor by Mathematica. u00 = Factor[Expand[c00 /. {x -> (t*y + y^d )/(t + 1)}]] u01 = Factor[Expand[c01 /. {x -> (t*y + y^d )/(t + 1)}]] u10 = Factor[Expand[c10 /. {x -> (t*y + y^d )/(t + 1)}]] u11 = Factor[Expand[c11 /. {x -> (t*y + y^d )/(t + 1)}]] [**OUTPUT:**]{} $$\begin{aligned} &-\frac{1}{(1+t)^9}(-1+y)^6 y^2 (1+t+y) \Big(4 t+28 t^2+84 t^3+140 t^4+140 t^5+84 t^6+28 t^7+4 t^8+12 t y\\ &+84 t^2 y+248 t^3 y+400 t^4 y+380 t^5 y+212 t^6 y+64 t^7 y+8 t^8 y+4 y^2+50 t y^2+228 t^2 y^2+532 t^3 y^2\\ &+714 t^4 y^2+570 t^5 y^2+264 t^6 y^2+64 t^7 y^2+6 t^8 y^2+16 y^3+141 t y^3+502t^2 y^3+954 t^3 y^3+1054 t^4 y^3\\ &+683 t^5 y^3+248 t^6 y^3+46 t^7 y^3+4 t^8 y^3+32 y^4+233 t y^4+694 t^2 y^4+1089 t^3 y^4+976 t^4 y^4+521 t^5 y^4\\ &+179t^6 y^4+43 t^7 y^4+5 t^8 y^4+40 y^5+243 t y^5+592 t^2 y^5+759 t^3 y^5+600 t^4 y^5+352 t^5 y^5+161 t^6 y^5\\ &+43 t^7 y^5+4 t^8 y^5+32 y^6+159 t y^6+318 t^2 y^6+387 t^3 y^6+385 t^4 y^6+302 t^5 y^6+139 t^6 y^6+26 t^7 y^6\\ &+t^8 y^6+16 y^7+63 t y^7+132 t^2 y^7+241 t^3 y^7+322 t^4 y^7+236 t^5 y^7+72 t^6 y^7+6 t^7 y^7+4 y^8+19 t y^8\\ &+82 t^2 y^8+201 t^3 y^8+234 t^4 y^8+110 t^5 y^8+15 t^6 y^8+12 t y^9+69 t^2 y^9+137 t^3 y^9+100 t^4 y^9 \\ &+20 t^5 y^9+10 t y^{10}+44 t^2 y^{10}+54 t^3 y^{10}+15 t^4 y^{10}+6 t y^{11}+16 t^2 y^{11}+6 t^3 y^{11}+2 t y^{12}+t^2 y^{12}\Big)\end{aligned}$$ $$\begin{aligned} &-\frac{1}{(1+t)^8}(-1+y)^5 y^2 (1+t+y) \Big(4 t+24 t^2+60 t^3+80 t^4+60 t^5+24 t^6+4 t^7+8 t y+50 t^2 y\\ &+126 t^3 y+164 t^4 y+116 t^5 y+42 t^6 y+6 t^7 y+4 y^2+38 t y^2+136 t^2 y^2+248 t^3 y^2+254 t^4 y^2+148 t^5 y^2\\ &+46 t^6 y^2+6 t^7 y^2+12 y^3+89 t y^3+258 t^2 y^3+383 t^3 y^3+311 t^4 y^3+138 t^5 y^3+35 t^6 y^3+6 t^7 y^3\\ &+20 y^4+118 t y^4+274 t^2 y^4+312 t^3 y^4+197 t^4 y^4+92 t^5 y^4+36 t^6 y^4+5 t^7 y^4+20 y^5+95 t y^5+165 t^2 y^5+\\ &156 t^3 y^5+127 t^4 y^5+87 t^5 y^5+29 t^6 y^5+2 t^7 y^5+12 y^6+42 t y^6+65 t^2 y^6+94 t^3 y^6+113 t^4 y^6\\ &+70 t^5 y^6+12 t^6 y^6+4 y^7+11 t y^7+36 t^2 y^7+85 t^3 y^7+90 t^4 y^7+30 t^5 y^7+6 t y^8+35 t^2 y^8\\ &+65 t^3 y^8+40 t^4 y^8+6 t y^9+25 t^2 y^9+30 t^3 y^9+4 t y^{10}+12 t^2 y^{10}+2 t y^{11}\Big)\end{aligned}$$ $$\begin{aligned} &-\frac{1}{(1+t)^7}(-1+y)^5 y^2 (1+t+y) \Big(4 t+20 t^2+40 t^3+40 t^4+20 t^5+4 t^6+10 t y+50 t^2 y+96 t^3 y\\ &+88 t^4 y+38 t^5 y+6 t^6 y+4 y^2+38 t y^2+120 t^2 y^2+174 t^3 y^2+126 t^4 y^2+44 t^5 y^2+6 t^6 y^2+16 y^3 \\ &+107 t y^3+267 t^2 y^3+324 t^3 y^3+197 t^4 y^3+55 t^5 y^3+6 t^6 y^3+32 y^4+177 t y^4+371 t^2 y^4+366 t^3 y^4\\ &+179 t^4 y^4+47 t^5 y^4+5 t^6 y^4+40 y^5+183 t y^5+310 t^2 y^5+250 t^3 y^5+118 t^4 y^5+28 t^5 y^5+2 t^6 y^5\\ &+ 32y^6+119 t y^6+164 t^2 y^6+132 t^3 y^6+56 t^4 y^6+8 t^5 y^6+16 y^7+47 t y^7+68 t^2 y^7+52 t^3 y^7+12 t^4 y^7\\ &+4 y^8+13 t y^8+23 t^2 y^8+8 t^3 y^8+4 t y^9+2 t^2 y^9\Big)\end{aligned}$$ $$\begin{aligned} &-\frac{1}{(1+t)^6}(-1+y)^4 y^2 \Big(4 t+20 t^2+40 t^3+40 t^4+20 t^5+4 t^6+10 t y+48 t^2 y+88 t^3 y+76 t^4 y\\ & +30 t^5 y+4 t^6 y+4 y^2+38 t y^2+117 t^2 y^2+164 t^3 y^2+116 t^4 y^2+42 t^5 y^2+7 t^6 y^2+16 y^3+105 t y^3\\ &+253 t^2 y^3+294 t^3 y^3+173 t^4 y^3+49 t^5 y^3+6 t^6 y^3+32 y^4+171 t y^4+339 t^2 y^4+310 t^3 y^4+139 t^4 y^4\\ &+39 t^5 y^4+4 t^6 y^4+40 y^5+173 t y^5+268 t^2 y^5+192 t^3 y^5+92 t^4 y^5+20 t^5 y^5+32 y^6+109 t y^6+132 t^2 y^6\\ &+102 t^3 y^6+40 t^4 y^6+16 y^7+41 t y^7+54 t^2 y^7+40 t^3 y^7+4 y^8+11 t y^8+20 t^2 y^8+4 t y^9\Big)\end{aligned}$$ Case $\Delta=4$ --------------- d = 3; F = Factor[((y*(1 + a)^2 - 1 - a + a*b)*(x*(1 + b)^2 + a*b + a)^d - (x*(1 + b)^2 - 1 - b + a*b)*(y*(1 + a)^2 + a*b + b)^d)][[3]]; H = Factor[Expand[F /. {a -> (-1 + y + Qa)/(x^d - y), b -> (-1 + x + Qb)/(y^d - x)}]][[3]]; da = Exponent[H, Qa]; For [i = da, i >= 2, i--, H = Expand[H /. {Qa^i -> Qa^(i - 2)*(1 - x^d - y + x^d*y)}]]; db = Exponent[H, Qb]; For [i = db, i >= 2, i--, H = Expand[H /. {Qb^i -> Qb^(i - 2)*(1 - y^d - x + y^d*x)}]]; c00 = Factor[Coefficient[Coefficient[H, Qa, 0], Qb, 0]]; c01 = Factor[Coefficient[Coefficient[H, Qa, 0], Qb, 1]]; c10 = Factor[Coefficient[Coefficient[H, Qa, 1], Qb, 0]]; c11 = Factor[Coefficient[Coefficient[H, Qa, 1], Qb, 1]]; u00 = Factor[Expand[c00 /. {x -> (t*y + y^d )/(t + 1)}]]; Length[u00]; u01 = Factor[Expand[c01 /. {x -> (t*y + y^d )/(t + 1)}]]; Length[u01]; u10 = Factor[Expand[c10 /. {x -> (t*y + y^d )/(t + 1)}]]; Length[u10]; u11 = Factor[Expand[c11 /. {x -> (t*y + y^d )/(t + 1)}]]; Length[u11]; [**Comment:**]{} the following code checks the positivity of the coefficients of the last factor of $c_{00}, c_{01}, c_{10}, c_{11}$. BAD = False; u00[[1]]*u00[[2]]*u00[[3]]*u00[[4]]*u00[[5]]*u00[[6]] For[i = 1, i <= Length[u00[[7]]], i++, If[ (u00[[7]][[i]] /. {T -> 1, y -> 1}) < 0, BAD = True]]; u01[[1]]*u01[[2]]*u01[[3]]*u01[[4]]*u01[[5]]*u01[[6]] For[i = 1, i <= Length[u01[[7]]], i++, If[ (u01[[7]][[i]] /. {T -> 1, y -> 1}) < 0, BAD = True]]; u10[[1]]*u10[[2]]*u10[[3]]*u10[[4]]*u10[[5]]*u10[[6]] For[i = 1, i <= Length[u10[[7]]], i++, If[ (u10[[7]][[i]] /. {T -> 1, y -> 1}) < 0, BAD = True]]; u11[[1]]*u11[[2]]*u11[[3]]*u11[[4]]*u11[[5]]*u11[[6]] For[i = 1, i <= Length[u11[[7]]], i++, If[ (u11[[7]][[i]] /. {T -> 1, y -> 1}) < 0, BAD = True]]; Print[BAD]; [**OUTPUT:**]{} $$-\frac{(-1+y)^{10} y^4 (1+y)^4 \left(1+t+y+y^2\right)^2}{(1+t)^{20}}$$ $$-\frac{(-1+y)^9 y^4 (1+y)^4 \left(1+t+y+y^2\right)}{(1+t)^{19}}$$ $$-\frac{(-1+y)^9 y^4 (1+y)^4 \left(1+t+y+y^2\right)}{(1+t)^{17}}$$ $$-\frac{(-1+y)^8 y^4 (1+y)^4 \left(1+t+y+y^2\right)}{(1+t)^{16}}$$ $\text{False}$ Case $\Delta=5$ --------------- d = 4; F = Factor[((y*(1 + a)^2 - 1 - a + a*b)*(x*(1 + b)^2 + a*b + a)^d - (x*(1 + b)^2 - 1 - b + a*b)*(y*(1 + a)^2 + a*b + b)^d)][[3]]; H = Factor[Expand[F /. {a -> (-1 + y + Qa)/(x^d - y),b -> (-1 + x + Qb)/(y^d - x)}]][[3]]; da = Exponent[H, Qa]; For [i = da, i >= 2, i--, H = Expand[H /. {Qa^i -> Qa^(i - 2)*(1 - x^d - y + x^d*y)}]]; db = Exponent[H, Qb]; For [i = db, i >= 2, i--, H = Expand[H /. {Qb^i -> Qb^(i - 2)*(1 - y^d - x + y^d*x)}]]; c00 = Factor[Coefficient[Coefficient[H, Qa, 0], Qb, 0]]; c01 = Factor[Coefficient[Coefficient[H, Qa, 0], Qb, 1]]; c10 = Factor[Coefficient[Coefficient[H, Qa, 1], Qb, 0]]; c11 = Factor[Coefficient[Coefficient[H, Qa, 1], Qb, 1]]; u00 = Factor[Expand[c00 /. {x -> (t*y + y^d )/(t + 1)}]]; Length[u00]; u01 = Factor[Expand[c01 /. {x -> (t*y + y^d )/(t + 1)}]]; Length[u01]; u10 = Factor[Expand[c10 /. {x -> (t*y + y^d )/(t + 1)}]]; Length[u10]; u11 = Factor[Expand[c11 /. {x -> (t*y + y^d )/(t + 1)}]]; Length[u11]; (* proof for Delta=5, d=4 *) BAD = False; u00[[1]]*u00[[2]]*u00[[3]]*u00[[4]]*u00[[5]]*u00[[6]] For[i = 1, i <= Length[u00[[7]]], i++, If[ (u00[[7]][[i]] /. {T -> 1, y -> 1}) < 0, BAD = True]]; u01[[1]]*u01[[2]]*u01[[3]]*u01[[4]]*u01[[5]]*u01[[6]] For[i = 1, i <= Length[u01[[7]]], i++, If[ (u01[[7]][[i]] /. {T -> 1, y -> 1}) < 0, BAD = True]]; u10[[1]]*u10[[2]]*u10[[3]]*u10[[4]]*u10[[5]]*u10[[6]] For[i = 1, i <= Length[u10[[7]]], i++, If[ (u10[[7]][[i]] /. {T -> 1, y -> 1}) < 0, BAD = True]]; u11[[1]]*u11[[2]]*u11[[3]]*u11[[4]]*u11[[5]]*u11[[6]] For[i = 1, i <= Length[u11[[7]]], i++, If[ (u11[[7]][[i]] /. {T -> 1, y -> 1}) < 0, BAD = True]]; Print[BAD]; [**OUTPUT:**]{} $$-\frac{(-1+y)^{14} y^6 \left(1+y+y^2\right)^6 \left(1+t+y+y^2+y^3\right)^2}{(1+t)^{35}}$$ $$-\frac{(-1+y)^{13} y^6 \left(1+y+y^2\right)^6 \left(1+t+y+y^2+y^3\right)^2}{(1+t)^{34}}$$ $$-\frac{(-1+y)^{13} y^6 \left(1+y+y^2\right)^6 \left(1+t+y+y^2+y^3\right)^2}{(1+t)^{31}}$$ $$-\frac{(-1+y)^{12} y^6 \left(1+y+y^2\right)^6 \left(1+t+y+y^2+y^3\right)}{(1+t)^{30}}$$ $\text{False}$ [^1]: School of Computer Science, Georgia Institute of Technology, Atlanta GA 30332. Email: {agalanis,vigoda}@cc.gatech.edu. Research supported in part by NSF grant CCF-1217458. [^2]: Department of Computer Science, University of Rochester, Rochester, NY 14627. Email: [email protected]. Research supported in part by NSF grant CCF-0910415.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We report transport measurements on a semiconductor quantum dot with a small number of confined electrons. In the Coulomb blockade regime, conduction is dominated by cotunneling processes. These can be either elastic or inelastic, depending on whether they leave the dot in its ground state or drive it into an excited state, respectively. We are able to discriminate between these two contributions and show that inelastic events can occur only if the applied bias exceeds the lowest excitation energy. Implications to energy-level spectroscopy are discussed.' address: '[$^1$Department of Applied Physics, DIMES, and ERATO Mesoscopic Correlation Project, Delft University of Technology, PO Box 5046, 2600 GA Delft, The Netherlands]{} ' author: - 'S. De Franceschi$^1$, S. Sasaki$^2$, J. M. Elzerman$^1$, W. G. van der Wiel$^1$, S. Tarucha$^{2,3}$, and L. P. Kouwenhoven$^1$' title: Electron Cotunneling in a Semiconductor Quantum Dot --- Quantum-dot devices consist of a small electronic island connected by tunnel barriers to source and drain electrodes [@NATO]. Due to on-site Coulomb repulsion, the addition of an electron to the island implies an energy change $U=e^2/C$, where $C$ is the total capacitance of the island. Hence the number of confined electrons is a well-defined integer, $N$, that can be controlled by varying the voltage on a nearby gate electrode. Transport of electrons through the dot is allowed only at the transition points where the $N$- and ($N+1$)-states are both energetically accessible. Otherwise, $N$ is constant and current is strongly suppressed. This is known as Coulomb blockade [@NATO]. At low temperature, however, higher-order tunneling events can become dominant. These are commonly known as $cotunneling$ events since they involve the simultaneous tunneling of two or more electrons . Electron cotunneling has received considerable attention over the last decade. Initially it was recognized as a limitation to the accuracy of single-electron devices. More recently, it has acquired a broader relevance, especially since an increasing activity has been focused on quantum dots with a deliberately enhanced tunnel coupling to their leads. These systems allow the investigation of high-order transport processes and many-body phenomena, such as the Kondo effect [@Kondo98; @Sasaki]. The latter can be regarded as the result of a coherent superposition of higher-order cotunneling events. Here, we will focus on the lowest order contribution to cotunneling. Previous experiments were performed with metallic islands [@Geerligs90; @Eiles92; @Hanna92] or large semiconductor dots [@Glattli91; @Pasquier93; @Cronenwett97], where the energy spectrum is essentially continuous and many levels contribute to cotunneling. Here, we study cotunneling through a small quantum dot where the energy levels are well separated, and where the absolute value of $N$ is precisely known. A cotunneling event is called inelastic when it leaves the dot in an excited state. Otherwise it is classified as elastic. We identify two regimes: one consisting of elastic processes only, and one including both elastic and inelastic contributions. We note that the transition between these regimes can be sharper than the characteristic life-time broadening of the dot states. In such a case, the onset of inelastic cotunneling can be exploited to measure the energy spectrum of a quantum dot with improved resolution. The stability diagram of a generic quantum dot can be obtained by plotting the differential conductance (d$I$/d$V_{sd}$) as a function of bias, $V_{sd}$, and gate voltage, $V_g$. Coulomb blockade occurs within the diamond-shaped regions in Fig. 1a. The diamond size is proportional to the addition energy, defined as $E_{add}(N) \equiv \mu_{dot}(N+1)- \mu_{dot}(N)$, where $\mu_{dot}(N)$ is the electrochemical potential of an $N$-electron dot. Inside the $N$-electron diamond, $\mu_{dot}(N) < \mu_L, \mu_R < \mu_{dot}(N+1)$, with $\mu_L$, $\mu_R$ the Fermi energies of the leads. The diamond edges correspond to level alignment: $\mu_{dot}(N)$ = $\mu_L$ or $\mu_R$ (see angled solid lines). This alignment determines the onset for first-order tunneling via the ground state of the dot, leading to a peak in d$I$/d$V_{sd}(V_{sd})$. The onset for first-order tunneling via the first excited state occurs at a somewhat higher bias (see dot-dashed lines in Fig. 1a, and the corresponding diagrams in Fig. 1b and 1e). These first-order processes have been exploited as a spectroscopic tool on the discrete energy spectrum of dots [@NATO]. Here, we are interested in second-order tunneling of charge, which becomes more apparent when the tunnel coupling between the dot and the leads is enhanced. We neglect contributions from spin that could give rise to the Kondo effect. Elastic cotunneling is the dominant off-resonance process at low bias. It gives rise to current inside the Coulomb diamond (light-grey region in Fig. 1a). The corresponding two-electron process (Fig. 1c) transfers one electron from the left to the right lead, thereby leaving the dot in the ground state. For $e|V_{sd}| \ge \Delta(N)$, where $\Delta(N)$ is the lowest on-site excitation energy for a constant $N$ [@note1], similar two-electron processes can occur which drive the dot into an excited state. For instance, an electron can leave the dot from the ground state to the lowest Fermi sea, while another electron from the highest Fermi sea tunnels into the excited state (see Fig. 1d). Although this type of process is called inelastic , the total electron energy is conserved. The on-site excitation is created at the expense of the energy drop $eV_{sd}$. To first approximation, the onset of inelastic cotunneling yields a step in d$I$/d$V_{sd} (V_{sd})$ [@Funabashi99]. This step occurs when $e|V_{sd}| = \Delta(N)$, which is not or only weakly affected by $V_g$ (see also Fig. 1c in Ref. [@Schmid00]). As a result, inelastic cotunneling turns on along the vertical (dotted) lines in Fig. 1a. At the edge of the Coulomb diamond the condition for the onset of inelastic cotunneling connects to that for the onset of first-order tunneling via an excited state (dot-dashed lines). Our device has the external shape of a 0.5-$\mu$m-high pillar with a $0.6 \times 0.45$ $\mu$m$^2$ rectangular base (inset to Fig. 2). It is fabricated from an undoped AlGaAs(7 nm)/InGaAs(12 nm)/AlGaAs(7 nm) double barrier heterostructure, sandwiched between n-doped GaAs source and drain electrodes [@Sasaki]. The quantum dot is formed within the InGaAs layer. The lateral confinement potential is close to that of an ellipse [@Austing99]. Its strength is tuned by a negative voltage, $V_g$, applied to a metal gate surrounding the pillar. A dc bias voltage, $V_{sd}$, applied between source and drain, drives current vertically through the pillar. In addition, we apply a small bias modulation with rms amplitude $V_{ac} = 3$ $\mu$V at 17.7 Hz for lock-in detection. Measurements are carried out in a dilution refrigerator with a base temperature of 15 mK. We find an effective electron temperature $T_e= 25 \pm 5$ mK, due to residual electrical noise. Figure 2 shows d$I$/d$V_{sd}$ in grey-scale versus ($V_{sd}$,$V_g$) at 15 mK. Diamond-shaped regions of low conductivity (light grey) identify the Coulomb blockade regimes for $N$ = 1 to 4. The diamonds are delimited by dark narrow lines (d$I$/d$V_{sd} \sim e^2/h$) corresponding to the onset of first-order tunneling. For $N=1$, as well as for $N=3$, sub-gap transport is dominated by elastic cotunneling with no evidence for inelastic cotunneling. The differential conductance is uniformly low inside the Coulomb diamond. (Slight modulations are seen due to a weak charging effect in the GaAs pillar above the dot [@foot1].) This is different for $N=2$, where the onset of inelastic cotunneling is clearly observed. As argued before, this onset follows (dotted) lines, nearly parallel to the $V_{g}$ axis [@foot3]. At the diamond edges they connect to (dot-dashed) lines where first-order tunneling via an excited state sets in. Similar considerations apply to $N=4$. The different behavior observed for $N=even$ and $N=odd$ stems from the fact that inelastic cotunneling occurs only if $E_{add}(N) > \Delta(N)$, as apparent from Fig. 1a. In the case of non-interacting electrons, $E_{add}(N) = U(N)$ for $N=odd$ and $E_{add}(N) = U(N) + \Delta(N)$ for $N=even$, where $U(N)$ is the charging energy for $N$ electrons; energy levels are spin degenerate and consecutively filled with pairs of electrons. This is a reasonable picture if the level spacing exceeds the exchange interaction energy [@Tarucha00]. In our small quantum dot the first three levels are indeed widely spaced as already discussed in Ref. [@Sasaki]. For low $N$, $\Delta(N)$ exceeds not only the exchange energy but also $U(N)$. This implies that for $N=odd$, $\Delta(N)$ lies outside the Coulomb diamond (i.e. $\Delta(N) > E_{add}(N)$) and thus inelastic cotunneling is not observed. (Note that cotunneling between spin-degenerate states is an elastic process as initial and final state have the same energy.) For $N=even$, $\Delta(N)$ is always smaller than $E_{add}(N)$ and inelastic cotunneling can be observed. We now discuss the difference in life-time broadening between first- and higher-order tunneling. At the onset of first-order tunneling a certain level is aligned to one of the Fermi energies. In this case, an electron can escape from the dot, which leads to a finite life-time broadening of the observed resonance by an amount $\hbar \Gamma$. Here, $\Gamma = \Gamma_L + \Gamma_R$, where $\Gamma_L$ and $\Gamma_R$ are the tunneling rates through the left and the right barrier, respectively. (Note that these rates are independent of $V_{sd}$, since our bias window ($\sim$meV) is much smaller than the height of the AlGaAs tunnel barriers ($\approx$50 meV). The onset of inelastic cotunneling is also characterized by a width. In the zero-temperature limit, this is determined by the life-time broadening of the excited state. Two types of situations can occur. First, the excited state can be between $\mu_L$ and $\mu_R$ (see right inset to Fig. 3) so that inelastic cotunneling can be followed by first-order tunneling. Such a decay event leads to a life-time broadening of at least $\hbar \Gamma_R \approx \hbar \Gamma/2$. Second, the ground and excited state are both well below $\mu_L$ and $\mu_R$, implying that only higher-order tunneling is allowed (see right inset to Fig. 4). Now, decay from the excited state can only rely on cotunneling. For this higher-order perturbation, the corresponding rate, $\Gamma_{co}$, is much smaller than $\Gamma$, leading to a reduced life-time broadening. To illustrate these arguments, we select different d$I$/d$V_{sd}-vs-V_{sd}$ traces and analyse their shape in detail. Figure 3 shows two traces for $N=2$, taken at 15 mK for gate voltages at the horizontal lines in the left inset. The dashed trace has several peaks. The two inner ones, at $|V_{sd}| \approx 1.1$ mV, correspond to first-order tunneling of the 3rd electron via the 3-electron ground state; i.e. $\mu_{dot}(3)$ = $\mu_L$ or $\mu_R$. The right (left) peak has a full width at half maximum (FWHM) of $\approx$200 ($\approx$400) $\mu$V. This is somewhat larger than the width, $\hbar \Gamma/e \simeq 150$ $\mu$V, measured in the zero-bias limit. Indeed a finite bias allows non-energy-conserving tunneling events leading to additional broadening. The most likely source for energy relaxation is acoustic-phonon emission [@Fujisawa]. The following pair of peaks, at $|V_{sd}| \approx 2$ mV, corresponds to the onset of first-order tunneling via the first excited state for $N=2$ (see Fig. 1b). Because of the larger bias voltage, these peaks are visibly broader than the inner ones. Additional peak structures occur near the edges of the bias window. The origin of these peaks can not be precisely identified. The solid trace contains structure from both first- and second-order tunneling. The peaks labeled by open squares arise from first-order tunneling at the edges of the Coulomb diamond (see left inset). Steps, labeled by open circles, identify the onset of inelastic cotunneling and correspond to the open circles in the left inset. Their different heights are probably due to a left-right asymmetry in the tunnel coupling to the leads. Their $V_{sd}$-position, which is symmetric around zero bias, provides a direct measure of $\Delta(2)$. The width of these steps is $\approx$150 $\mu$V [@foot2]. Since $\Delta(2) \approx U(2)$, the first excited state lies unavoidably within the bias window when $|V_{sd}| = \Delta(2)/e$ and hence is allowed to decay into the lowest-energy lead (see the right inset to Fig. 3). As argued above, this situation leads to a step-width exceeding $\hbar \Gamma_R/e$, consistent with our finding. Another structure occurs at $V_{sd} \approx 2.6$ mV and is probably due to the onset of inelastic cotunneling via the second excited state for $N=2$. The corresponding line in the stability diagram is hardly visible due to its vicinity to the diamond edge. To study inelastic cotunneling when both ground and excited state lie well below the Fermi energies of the leads (Fig. 1d) we need $\Delta(N) \ll E_{add}(N)$. To this aim we move to $N=6$, since $\Delta(6)$ can be effectively tuned by a magnetic field applied along the vertical axis. We tune the field to 0.35 T, such that $\Delta(6) \approx 0.1$ meV, i.e. several times smaller than $E_{add}(6)$. From a previous study we know that the ground state is a spin singlet, and the first excited state is a spin triplet [@Sasaki]. The d$I$/d$V_{sd}-vs-V_{sd}$ traces shown in Fig. 4 are taken at two different temperatures, but for the same $V_g$, at the horizontal line in the left inset. The solid trace (15 mK) shows a broad minimum around $V_{sd}=0$, where transport is dominated by elastic cotunneling via the ground state (see also the light-grey region in the left inset). The differential conductance increases rapidly at the onset of inelastic cotunneling with a step-width of $\approx$20 $\mu$V, i.e. much smaller $\hbar \Gamma/e$. This reduced width stems from the fact that the excited state can not decay directly into the lower energy lead (see right inset). The corresponding life-time broadening, $\hbar \Gamma_{co}$, can be estimated from the cotunneling current, $I_{co}$, at $V_{sd}=\Delta(6)/e$. We find $\hbar \Gamma_{co} = \hbar I_{co}/e \approx (\hbar/e) \int_0^{\Delta(6)/e}$ d$I$/d$V_{sd} (V_{sd})$ d$V_{sd} \simeq 10$ $\mu$eV, consistent with the observed step-width. At $T_e = 25$ mK the thermal broadening of the Fermi distribution functions leads to a step-width of $5.44 k_B T_e/e \simeq 12 $ $\mu$eV [@Wolf]. Hence life-time broadening has been reduced here to the thermal limit. The cotunneling onset in Fig. 4 shows a peak structure at low temperature (solid trace) in addition to the expected step structure. This is likely due to Kondo correlations, as discussed in Ref. [@Sasaki]. On increasing temperature to 200 mK these Kondo correlations are suppressed such that only lowest order cotunneling contributes. This recovers the step structure (dashed trace). We thank Yu. V. Nazarov, M. R. Wegewijs, M. Eto, K. Maijala, and J. E. Mooij for discussions. We acknowledge financial support from the Specially Promoted Research, Grant-in-Aid for Scientific Research, from the Ministry of Education, Science and Culture in Japan, from the Dutch Organisation for Fundamental Research on Matter (FOM), from the NEDO joint research program (NTDP-98), and from the EU via a TMR network. L. P. Kouwenhoven, C. M. Marcus, P. L. McEuen, S. Tarucha, R. M. Westervelt, and N. S. Wingreen, in [*Mesoscopic Electron Transport*]{}, edited by L.L. Sohn, L. P. Kouwenhoven, and G. Schön, (Kluwer, Series E 345, 1997), p. 105-214. D. V. Averin and Yu. V. Nazarov, in [*Single Charge Tunneling - Coulomb Blockade Phenomena in Nanostructures*]{}, edited by H. Grabert and M. H. Devoret (Plenum Press and NATO Scientific Affairs Division, New York, 1992), p. 217. D. Goldhaber-Gordon [*et al.*]{}, Nature [**391**]{}, 156, (1998); Cronenwett [*et al.*]{}, Science [**281**]{}, 540 (1998); Schmid [*et al.*]{}, Physica B [**256-258**]{}, 182 (1998). S. Sasaki [*et al.*]{}, Nature [**405**]{}, 764 (2000). L. J. Geerligs, D. V. Averin, and J. E. Mooij, Phys. Rev. Lett. [**65**]{}, 3037 (1990). T. M. Eiles [*et al.*]{}, Phys. Rev. Lett. [**69**]{}, 148 (1992). A. E. Hanna, M. T. Tuominen, and M. Tinkham, Phys. Rev. Lett. [ **68**]{}, 3228 (1992). D. C. Glattli [*et al.*]{}, Z. Phys. B [**85**]{}, 375 (1991). C. Pasquier [*et al.*]{}, Phys. Rev. Lett. [**70**]{}, 69 (1993). S. M. Cronenwett [*et al.*]{}, Phys. Rev. Lett. [**79**]{}, 2312 (1997). In our definition $\Delta(N)$ is strictly positive. It reduces to the single-particle level spacing for non-interacting electrons (for instance, in the case of $N=1$ and $N=2$, $\Delta(N)$ is the spacing between the first two single-particle levels). Y. Funabashi [*et al.*]{}, Jpn. J. Appl. Phys. [**38**]{}, 388 (1999). J. Schmid [*et al.*]{}, Phys. Rev. Lett. [**84**]{}, 5824 (2000). D. G. Austing, [*et al.*]{}, Phys. Rev. B [**60**]{}, 11514 (1999). The top contact is obtained by deposition of Au/Ge and annealing at 400 $^\circ$C for 30 s. This thermal treatment is gentle enough to prevent the formation of defects near the dot, but does not allow the complete suppression of the native Schottky barrier. The residual barrier leads to electronic confinement and corresponding charging effects in the GaAs pillar. $V_g$ affects not only the bottom but also the shape of the confining potential. As a result, the level spacing (and hence $\Delta(N)$) depends weakly on $V_g$, leading to a non-zero angle between the dotted lines and the $V_g$ axis. S. Tarucha [*et al.*]{}, Phys. Rev. Lett. [**84**]{}, 2485 (2000). T. Fujisawa [*et al.*]{}, Science [**282**]{}, 932 (1998). The step-width is estimated by taking the full width at half maximum of the corresponding peak (or dip) in d$^2 I$/d$V_{sd}^2(V_{sd})$. E. L. Wolf, [*Principles of Electron Tunneling Spectroscopy*]{}, (Oxford, New York, 1985) p. 438.
{ "pile_set_name": "ArXiv" }
--- address: | Osservatorio Astronomico di Bologna, via Ranzani 1, I-40127 Bologna, Italy\ [In collaboration with:\ T. VENTURI, R. MORGANTI, R.W. HUNSTEAD (radio)\ S. ETTORI, S. DE GRANDI, S. MOLENDI, G. ZAMORANI (X-rays)\ A. BALDI (optical) ]{} author: - | SANDRO BARDELLI & ELENA ZUCCA\ ([email protected], [email protected]) title: | The VIOLENT ENVIRONMENT of the SHAPLEY CONCENTRATION:\ a MULTIWAVELENGTH VIEW --- Introduction ============ Merging clusters are the most energetic and common phenomena in the Universe, but it is not clear in which way the kinetic energy of this phenomenon is dissipated and which is the influence on the galaxy population. There are some observational features that have been associated with the cluster merging, like shocks in the hot gas, radio halos, relics and wide angle tail radiosources, and the presence of starburst galaxies, even if the precise theorical description is not yet completely assesed.\ Rich superclusters are the ideal environment for the detection of cluster mergings, because the peculiar velocities induced by the enhanced local density of the large-scale structure favour the cluster-cluster and cluster-group collisions, in the same way as the caustics seen in the simulations.\ The most remarkable examples of cluster merging seen at an early stage are found in the central region of the Shapley Concentration, the richest supercluster of clusters within 300 . This supercluster represents a unique laboratory where it is possible to follow cluster mergings and to test related astrophysical consequences.\ We present the results of an extensive, multiwavelength survey of the central part of the Shapley Concentration, with the use of optical spectra, radio and X-ray data (see also the poster by Zucca et al., these proceedings).\ Updates to this project can be found at the WWW page:\ http://boas5.bo.astro.it/$\sim$bardelli/shapley/shapley\_new.html OPTICAL DATA ============ We performed extensive redshift surveys in the central part of the Shapley Concentration, covering both the cluster and the intercluster galaxies. We focussed in particular on the A3558 and the A3528 complexes, two chains of interacting clusters. We derived the dynamical parameters of each cluster (Bardelli et al. 1994, 1998a, 2000b) and we studied their substructures (Bardelli et al. 1998b, 2000b).\ The analysis of substructures is important in order to fully describe the galaxy dynamics in major mergings: in this sense the DEDICA algorithm (Pisani 1996) is particularly powerful, because it is possible to perform both a bi-dimensional and a three-dimensional (i.e. using the galaxy velocity as the third coordinate) analysis.\ In Figures 1 and 2 we show the results of this analysis on the A3528 and the A3558 complexes, respectively.\ In the A3558 complex, the members of the chain appear to be fragmented in a large number of components: in fact we found 21 significant three-dimensional subclumps. On the contrary, the A3528 complex is less fragmented: in fact, even if these two complexes have similar sizes, in the A3528 complex we detect only 8 significant three-dimensional subclumps. RADIO DATA ========== We performed a radio survey at 22cm with the ATCA telescope to 1 mJy in the region of the A3558 and A3528 cluster complexes, obtaining a total of 450 sources (Venturi et al. 2000a, 2000b). Among these sources, $\sim 100$ have optical counterpart and 43 have redshifts in the Shapley Concentration.\ In Figure 3 we show the detected radiosources overlaid on the optical isodensity contours: filled circles and squares are sources with optical counterparts. The position of the radiosources at the distance of the Shapley Concentration is reported in Figure 4, where the wedge diagrams of the optical sample are presented.\ The bivariate radio-optical lunimosity function has been calculated for the A3528 and A3558 cluster complexes and compared with the luminosity function of normal clusters (Ledlow & Owen 1996) in Figure 5: it is clear that in the A3558 complex a significant lack of radiogalaxies is present.\ In the region of interacting clusters, a number of extended radiosources has been detected (Figure 6 and Figure 7). The study of these sources is particularly important for deriving physical constraints on the interaction of electrons with the ICM. In particular, Venturi et al. (1999), studying a relic radio source at the periphery of the A3558 complex, individuated a possible shock front between the A3556 cluster and a smaller group.\ Following this strategy we will integrate the optical, X-ray and radio data in order to have a general description of the ongoing phenomena (Figure 8). X-RAY DATA ========== Fundamental data on the physics of the ICM during a merging come from the X-ray band. Figure 11 (taken from Ettori et al. 1997) shows a mosaic of ROSAT-PSPC images of clusters in the A3558 complex.\ Bardelli et al. (1996) and Kull & Böhringer (1999) showed that the three clusters (A3562, A3558 and A3556) and the two poor groups (SC1329-131 and SC1327-312) of this complex are enbedded in a hot gas filament. This fact points toward a major merging event. A careful analysis of the distribution and the temperature of the ICM is currently in progress with the use of ROSAT and BeppoSAX data (Ettori et al. 2000, see Figures 12 and 13).\ For the A3528 cluster complex, pointed XMM observations have been allocated. CONCLUSIONS =========== In this contribution we described a multiwavelength study of two cluster complexes at the center of the Shapley Concentration: both these structures are formed by interacting clusters and have sizes of the order of $\sim 7$ . In the A3558 complex, the members of the chain appear to be fragmented in a large number of components, detected with high significance both in the optical and in the X-ray band. All these components are embedded in a hot gas filament. Moreover, Venturi et al. (2000) found a deficiency of radiogalaxies in the A3558 complex with respect to the radio-optical luminosity function of normal clusters and Bardelli et al. (1998b) found an excess of blue galaxies in the expected position of the shock.\ These facts could suggest that the dynamical processes acting on this complex are in a rather advanced stage and that the merging events were already able to induce modifications in the galaxy population. Our conclusion is that the A3558 complex is a cluster-cluster collision (otherwise called “major merging") seen just after the first core-core encounter, where an intervening cluster impacted onto the richer object A3558. Indeed the clumpiness found eastward of A3558 could be due to the galaxies of this intervening cluster, which is now emerging from the main component. This scenario is also confirmed by numerical simulations of merging clusters (see f.i. lower panels in Fig.2 of Burns et al. 1994, which show the structure after the core-core encounter). In the A3528 cluster complex, the effect of interactions is less spectacular. This complex appear less fragmented than the A3558 complex and the bivariate radio-optical luminosity function is not different with respect to the other clusters.\ All these characteristics of the A3528 complex seem to indicate that this is a “young" structure (Bardelli et al. 2000b): the two interacting cluster pairs (A3528N-A3528S and A3530-A3532) resemble the simulations of Burns et al. (1994) for the pre-merger case (see upper panels of their Fig.2). This scenario is consistent with the suggestions of Reid et al. (1998), based on the analysis of extended radiosources in this region and with the fact that the merging effects on the galaxy population and on the cluster dynamics are not yet evident.\ Given the overall overdensity of this region, at the end these two main components will merge together in a major merging event, forming a structure similar to the A3558 complex. The “age" difference of the two complexes is mainly due to their overdensities, which lead to different collapse times (see Table 5 in Bardelli et al. 2000a). This fact confirms that the Shapley Concentration is a “laboratory" where it is possible to study the formation of clusters at different stages. =8.5cm Figure 1: Substructures in the A3528 region found in the three–dimensional sample using the DEDICA algorithm. Solid lines connect the position of each galaxy to the common limiting position $\vec{x}_{lim}$ of the group.\ Upper panel: projection on the plane of the sky. Lower panel: projection on the velocity–Y plane. =0.49 =0.49 =0.49 =0.49 Figure 2: The positions of subcluster galaxy found by DEDICA in the three-dimensional sample covering the A3558 complex are overplotted on the 2D isodensity contours of galaxy distribution. The high number of substructures reveals that this structure is far from the equilibrium state.\ Upper panels: Substructures found in the A3558 cluster; for clarity, the group members have been split in two panels.\ Lower panels: The same as above for SC1329-313 & SC1327-312 (left) and A3556 (right). = 0.8 =0.75 Figure 3: Radio sources detected in the 22cm ATCA survey in the A3558 and A3528 complex regions overlaid on the optical isodensities. Filled circles represent optical counterparts with measured redshift, filled squares represent those without velocity information and crosses are radiosources without optical counterparts. = = 0.9 Figure 4: Wedge diagrams of galaxies in the A3558 and A3528 regions. Crosses are radiosources with optical counterpart. In the A3558 complex, radio galaxies with spectral index $\alpha^{22}_{13}>1$ have been marked with a square and the extended sources have been highlighted. In the A3528 complex, extended radio galaxies are indicated by triangles. =0.8 =0.65 Figure 5: Bivariate radio-optical luminosity function for the A3558 (upper panel) and the A3528 (lower panel) complexes. Given the uncertainty in the fraction of ellipticals, the real luminosity function of the A3528 complex should lie in the shaded region. Triangles represent the normal cluster luminosity function of Ledlow & Owen (1996): note the lack of radio sources in the A3558 complex with respect to normal clusters. = Figure 6: Radio(20cm VLA)-optical superposition at the center of the cluster A3560. This kind of information, added to X-ray maps and spectra, will constrain the dynamics of merging clusters. =0.75 =0.75 Figure 7: The A3528 cluster is actually a double cluster. In this figure we superimposed the radio maps of the two components (dubbed A3528N and A3528S, Schindler 1996) on the Digital Sky Survey. Note the central galaxy of A3528N, which is formed by two radiosources. =0.9 Figure 8: The radio isophotes from the NVSS survey are superimposed to a part of the ROSAT PSPC image centered on A3562. At the center of the cluster A3562 (on the left) the emission of an Head Tail radio galaxy is clearly visible. Additional extended emission is present in the cluster center. At the center of the figure, an extended radio emission (possibly a mini-halo) is located at the border of the poor cluster SC 1329-313. This emission is associated to a bright cluster galaxy. = Figure 9: The soft X-ray emission (ROSAT mosaic from Ettori et al. 1997) in the A3558 cluster complex, with the main clusters labelled. We propose that this structure is the remnant of a cluster-cluster collision seen after the first core-core encounter. =0.7 =0.7 Figure 10: Upper panel: ROSAT PSPC image of the A3562 cluster. The image has been adaptively smoothed in order to enhance the features with a significance higher than 5 $\sigma$ with respectc to the background. Lower panel: Residuals in the emission map after subtraction of a best fit $\beta$ model. The darker regions represent excess in emission, the lighter ones indicate a deficit. =0.7 =0.7 Figure 11: Upper panel: The LECS (open squares) and PSPC (filled circles) counts of A3562 are here divided by the best-fit absorbed MEKAL model with fixed galactic absorption obtained from the MECS data (open triangles). Note that the LECS data are shifted by a factor $\sim 0.77$ in normalization with respect to MECS and that the PSPC counts present a large disagreement below 0.5 keV. Lower panel: Projected radial temperature and abundance profiles for A3562, as determined from the BeppoSAX map. Filled circles and open triangles indicate the measurements obtained including and excluding point sources. = 0.7 =0.7 Figure 12: BeppoSAX MECS maps of the two poor groups SC1327-312 and SC1329-313, which are in between A3558 and A3562. References {#references .unnumbered} ========== [99]{} Baldi A., Bardelli S., Zucca E., 2000, MNRAS submitted Bardelli S., Zucca E., Vettolani G., Zamorani G., Scaramella R., Collins C.A., MacGillivray H.T., 1994, MNRAS 267, 665 Bardelli S., Zucca E., Malizia A., Zamorani G., Scaramella R., Vettolani G., 1996, A&A 305, 435 Bardelli S., Zucca E., Zamorani G., Vettolani G., Scaramella R., 1998a, MNRAS 296, 599 Bardelli S., Pisani A., Ramella M., Zucca E., Zamorani G., 1998b, MNRAS 300, 589 Bardelli S., Zucca E., Zamorani G., Moscardini L., Scaramella R., 2000a, MNRAS 312, 540 Bardelli S., Zucca E., Baldi A., 2000b, MNRAS in press (astro-ph/0009008) Burns J.O., Roettiger K., Ledlow M., Klypin A., 1994, ApJL 427, L87 Ettori S., Fabian A.C., White D.A., 1997, MNRAS 289, 787 Ettori S., Bardelli S., DeGrandi S., Molendi S., Zamorani G., Zucca E., 2000, MNRAS in press Kull A., Böhringer H., 1999, A&A 341, 23 Ledlow M.J., Owen F.N., 1996, AJ 112, 9 Pisani A., 1996, MNRAS 278, 697 Reid A.D., Hunstead R.W., Pierre M.M., 1998, MNRAS 296, 949 Schindler S., 1996, MNRAS 280, 309 Venturi T., Bardelli S., Morganti R., Hunstead R.W., 1997, MNRAS 285, 898 Venturi T., Bardelli S., Morganti R., Hunstead R.W., 1998, MNRAS 298, 1113 Venturi T., Bardelli S., Zambelli G., Morganti R., Hunstead R.W., 1999, in [*Diffuse thermal and relativistic plasma in galaxy clusters*]{}, H.Böhringer et al. eds., MPE report 271 p.27 Venturi T., Bardelli S., Morganti R., Hunstead R.W., 2000a, MNRAS 314, 594 Venturi T., Bardelli S., Zambelli G., Morganti R., Hunstead R.W., 2000b, MNRAS submitted
{ "pile_set_name": "ArXiv" }
--- abstract: 'We show that the quantum phase transition of the Tavis-Cummings model can be realised in a linear ion trap of the kind proposed for quantum computation. The Tavis-Cummings model describes the interaction between a bosonic degree of freedom and a collective spin. In an ion trap, the collective spin system is a symmetrised state of the internal electronic states of N ions, while the bosonic system is the vibrational degree of freedom of the centre of mass mode for the ions.' address: | $^*$ Centre for Quantum Computer Technology,\ and Department of Physics, The University of Queensland,QLD 4072 Australia.\ $^\dagger$ The University of New Mexico\ Albuquerque High Performance Computing Center\ Albuquerque, New Mexico 87131 author: - 'G.J.Milburn $^*$ and Paul Alsing$^\dagger$' title: Quantum phase transitions in a linear ion trap --- Introduction ============ More than two decades ago, when quantum optics was young, the quantum dynamics of collective spin systems interacting with a single bosonic degree of freedom was a major research problem. The model arose as an attempt to describe the interaction between a collection of two level atoms and a single mode of the radiation field. Walls and co workers[@Collective] were among the first to realise that such models provided ideal examples of the role of quantum fluctuations in the nonlinear interaction between matter and light. Quantum fluctuations were shown to drastically change the predictions of semiclassical theory in such systems. This phenomenon has appeared more recently in the discovery of quantum phase transitions in quantum spin glasses[@Reiger] and other many body quantum systems. While the collective spin models did not directly apply to achievable experiments at the time, they did provide insight that subsequently proved important for many other quantum optical experiments including anti-bunching, squeezing[@WallsMilb], and cavity QED[@Kimble]. In this paper we show that the models of a collective spin interacting with one or more bosonic modes can now be experimentally realised in modern ion trap systems of the kind proposed for quantum computation[@NIST; @LANL]. An enormous effort has gone into making such systems work at the quantum level, with little interference form classical sources of noise, and a number of such experiments exist today. It would thus appear worthwhile to reconsider the collective spin models, and the associated quantum many-body effects exhibited by such systems, with a view to direct experimental realisation. In particular we consider the Tavis-Cummings (TC) model[@tavis-cum], which can be realised in a linear ion trap of $N$ ions with the bosonic degree of freedom appearing as the quantised collective centre-of-mass motion. If each ion is coupled to the vibrational motion using an identical external (classical) laser detuned to the first red-sideband transition, the symmetry is such that the electronic degree of freedom for the ions can be described as a collective spin ($N$) and the reversible dynamics is well described by the TC model. The TC model is known to exhibit important nonlinear quantum effects including a quantum phase transition[@Reiger] in which the (zero temperature) ground state undergoes a morphological change as a parameter is varied and averages of intensive quantities undergo a bifurcation. The Tavis-Cummings model ======================== The interaction Hamiltonian for N ions interacting with the centre of mass vibrational mode can be controlled by using different kinds of Raman laser pulses. A considerable variety of interactions has already been achieved or proposed [@NIST; @LANL; @James]. In this paper we consider the first red-sideband transition. The ion is assumed to be in a three dimensional anisotropic harmonic potential. Two dimensions are very tightly bound and are neglected. In the remaining dimension, an external laser couples the electronic state to the vibrational motion. If the vibrational frequency is large enough and the Lamb-Dicke limit[@NIST] applies the motional sidebands of the absorption of the electronic transition can be resolved and a laser detuned below the electronic resonance by one unit of the trap frequency can excite the electronic transition by absorbing one vibrational phonon, the additional energy required being made up by the laser. We will assume that the laser ( or lasers if a Raman process is used) is sufficeintly strong that it can be treated classically. Under these assumptions the Hamiltonian, in the interaction picture, is $$H_I=\hbar\Omega\sum_{i=1}^N(a\sigma_+^{(i)}+a^\dagger \sigma_-^{(i)})$$ where the coupling constant is $ \Omega=\eta\Omega_0 $ where $\eta^2=E_r/(\hbar M\omega_0)$ is the Lambe-Dicke parameter with $E_r$ the recoil kinetic energy of the atom, $\omega_0$ is the trap vibrational frequency, and $M$ is the effective mass for the centre-of-mass mode. The Lamb-Dicke limit assumes $\eta\ <<\ 1$, which is easily achieved in practice. The frequency, $\Omega_0$ is the effective Rabi frequency for the electronic transition involved. The raising and lowering operators for each ion are defined by $\sigma_-=|g\rangle \langle e|$ and $\sigma_+=|e\rangle\langle g|$. This sideband transition can be used to efficiently cool the ions to the collective centre-of-mass ground state, thus preparing the system in the vibrational ground state[@NIST]. If the external laser field on each ion is identical (in amplitude and phase) the interaction Hamiltonian is $$H_I=\hbar\Omega(a\hat{J}_+ +a^\dagger \hat{J}_-) \label{TC}$$ where we have introduced the bosonic annihilation operator $a$ for the centre-of-mass vibrational mode and where we have used the definition of the collective spin operators, $$\hat{J}_\alpha=\sum_{i=1}^N\sigma^{(i)}_\alpha$$ where $\alpha=x,y,z$ . Identical laser fields could easily be obtained by splitting a single, stabilised laser into multiple beams. The interaction Hamiltonian in Eq (\[TC\]) specifies the Tavis-Cummings model[@tavis-cum]. This model first appeared in quantum optics where the bosonic mode is the quantised field in a cavity. However this realisation is difficult to achieve experimentally. In contrast the vibrational mode realisation should be readily achieved. The dynamics resulting from this Hamiltonian is quite rich. Collective spin models of this kind were considered many decades ago in quantum optics[@Drummond; @Dicke]. In much of that work however the collective spin underwent an irreversible decay. In the case of an ion trap model however we can neglect such decays due to the long lifetimes of the excited states. On the other hand heating of the vibrational centre-of-mass mode can induce irreversible dynamics in the system in a manner that has not been previously considered, and that is reminiscent of thermal effects in condensed matter physics. We are interested in the driven Tavis-Cummings model in which the vibrational mode is subject to a linear forcing term which can easily be achieved by a suitable combination of Raman laser pulses, or by appropriate AC voltages applied to the trap electrodes[@NIST]. In this case the Hamiltonian, in the interaction picture, is given by $$H_I=\hbar\Omega(a\hat{J}_+ +a^\dagger \hat{J}_-)+\hbar E(a+a^\dagger) \label{drivenTC}$$ This may be written in terms of the hermitian canonical oscillator variables $\hat{X}=(a+a^\dagger)/\sqrt{2}$, $\hat{Y}=-i(a-a^\dagger)/\sqrt{2}$, and the canonical angular momentum variables $\hat{J}_x=(\hat{J}_++\hat{J}_-)/2$, $\hat{J}_y=-i(\hat{J}_+-\hat{J}_-)/2$, $\hat{J}_z=[\hat{J}_+,\hat{J}_-]/2$. It takes the form $$H=\hat{X}\hat{J}_x-\hat{Y}\hat{J}_y+\chi\hat{X}$$ with $\chi=E/\Omega$ and we have scaled the Hamiltonian by $H\rightarrow H/\sqrt{2}\Omega$. This indicates that time is measured in units of $\frac{1}{\sqrt{2}\Omega}$. Alsing[@Alsing] has shown that the ground state of this system, for weak driving, is a product state in which the bosonic mode is squeezed and the electronic states are rotated in the angular momentum space. We provide a direct proof of this statement below. However it is first useful to consider the dynamics of the equivalent semiclassical model as many of the results in the quantum case can be interpreted in terms of the features of the semiclassical model. Semiclassical Tavis-Cummings model ---------------------------------- The Tavis-Cummings model represents an interaction between a simple harmonic oscillator and a linear top for which there is a classical model which we now define. We choose the classical model so that the equations of motion are of the same form as the Heisenberg equations of motion for the quantum model. The classical Hamiltonian is defined as $${\cal H}=X{\cal J}_x-Y{\cal J}_y+\chi EX$$ where $X,Y$ are respectively the canonical oscillator position and momentum variables with the canonical Poisson bracket $\{X,Y\}=1$, while ${\cal J}_k$ are the three components of angular momentum for a classical top with the canonical Poisson brackets $\{{\cal J}_i,{\cal J}_k\}=\sum_{k}\epsilon_{ijk}{\cal J}_k$. The equations of motion for a canonical coordinate $w$ is given as usual by Poisson bracket with the Hamiltonian $\dot{w}=\{w,H\}$. The equations of motion are, $$\begin{aligned} \dot{X} & = & -{\cal J}_Y\\ \dot{Y} & = & -{\cal J}_X-\chi\\ \dot{{\cal J}_x} & = & -Y{\cal J}_z\\ \dot{{\cal J}_y} & = & -X{\cal J}_z\\ \dot{{\cal J}_z} & = & X{\cal J}_y+Y{\cal J}_x\end{aligned}$$ Note that these equations have a conservation law ${\cal J}_x^2+{\cal J}_y^2+{\cal J}_z^2=\mbox{constant}$. We now justify this choice of classical Hamiltonian by noting that the Heisenberg equations of motion for the Hamiltonian Eq(\[drivenTC\]) have the same form as the semiclassical equations of motion with all variables replaced by the corresponding operators. We thus see that the semiclassical equations result form taking moments of the Heisenberg equations and factorising all product moments. The factorisation assumptions ignores correlations which scale as $1/N$ for the scaled operators $\hat{J}_\sigma/N$. The conservation law ${\cal J}_x^2+{\cal J}_y^2+{\cal J}_z=constant$ is a reflection of the operator relation $$\hat{J}^2=\frac{N}{2}\left (\frac{N}{2}+1\right )$$ which in the semiclassical limit indicates that ${\cal J}_x^2+{\cal J}_y^2+{\cal J}_z=\frac{N^2}{4}$. The classical equations have one nontrivial fixed point at $X^*=Y^*={\cal J}_y^*=0$ and ${\cal J}_x^*=-\chi$, $ {\cal J}_z^*=\sqrt{N^2/4-\chi^2}$. However as the conservation law requires that $|{\cal J}_x|\leq N/2$ we see that we must have $$\frac{2E}{N\Omega}\leq 1\ \ \ \ \mbox{(below threshold)}$$ which corresponds to an energy of ${\cal H}=0$. We will refer to this as the [*below threshold*]{} case. As $E$ is increased from zero, the fixed point for the angular momentum system rotates about the ${\cal J}_y$ direction eventually reaching the equatorial plane at ${\cal J}_x=-N/2$ at the threshold condition. The oscillator system always has zero amplitude below threshold. If we linearise around this fixed point we discover that it is an unstable hyperbolic point with time constant proportional to $\frac{1}{\sqrt{{\cal J}_z^*}}$. Note that this time constant goes to infinity as the fixed point is approached as is typical for a hyperbolic fixed point. We now consider the [*above threshold*]{} case $$\frac{2E}{N\Omega}\geq 1\ \ \ \ \mbox{(above threshold)}$$ Clearly the value of $|{\cal J}_x|$ cannot increase above $N/2$. Indeed there is no fixed point above threshold. However there is a special solution curve that continuously joins to the below threshold case for phase curves with ${\cal H}=0$. To see this we consider making a canonical transformation by a rotation in both the $X-Y$ plane and in the ${\cal J}_x,{\cal J}_y$ plane (see figure \[fig\_one\]). The canonical transformations are $$\begin{aligned} X & = & \bar{X}\cos\theta+\bar{Y}\sin\theta\\ Y & = & \bar{Y}\cos\theta-\bar{X}\sin\theta\\ {\cal J}_x & = & \bar{{\cal J}}_x\cos\theta-\bar{{\cal J}}_y\sin\theta\\ {\cal J}_y & = & \bar{{\cal J}}_x\cos\theta+\bar{{\cal J}}_y\sin\theta\end{aligned}$$ The Hamiltonian then takes the form $${\cal H}=\bar{X}(\bar{{\cal J}}_x+\chi\cos\theta)-\bar{Y}(\bar{{\cal J}}_y-\chi\sin\theta) \label{sc_ham}$$ The phase curves with ${\cal H}=0$ now correspond to either $$\begin{aligned} \bar{X} & = & 0\ \ \ \ ;\ \ \ \bar{{\cal J}}_y = \chi\sin\theta\end{aligned}$$ or $$\begin{aligned} \bar{Y} & = & 0\ \ \ \ ;\ \ \ \bar{{\cal J}}_x = -\chi\cos\theta\end{aligned}$$ These phase curves smoothly join the fixed point at threshold if ${\cal J}_x=-N/2$ which implies $$\cos\theta=\frac{N\Omega}{2E}$$ These solutions are illustrated in figure \[fig\_one\]. Note that as $E\rightarrow \infty$ we have that $\bar{{\cal J}}_x$ eventually points in the direction of $-{\cal J}_y$ while phase curve in the oscillator phase space points along the $Y$ axis, indicating that for large driving the system is essentially a particle in a linear potential which accelerates at constant rate. These results were first obtained by Alsing and Carmichael[@AlsingCar]. Quantum states ============== First note that the ground state when there is no driving is $|j,-j\rangle_z\otimes|0\rangle_v$ with a zero eigenvalue. This ground state corresponds to the fixed point of the semiclassical model with zero oscillator amplitude and angular momentum pointing in the $-{\cal J}_z$ direction. We postulate that as the driving is increased form zero the ground state of the Hamiltonian Eq(\[drivenTC\]) is given by $$|{\cal E}_0\rangle= S(r)R(\theta)|j,-j\rangle_z\otimes|0\rangle_v$$ where $ |j,-j\rangle_z\otimes|0\rangle_v$ corresponds to all ions in the ground state and the vibrational mode in the ground state. The operator $S(r)$ is a squeezing operator defined by $$S^\dagger(r)aS(r)=\mu a+\nu a^\dagger$$ with $\mu=\cosh r,\ \ \nu=\sinh r$. The rotation operator $R(\theta)$ is defined by $$R(\theta)=e^{-\theta(\hat{J}_+-\hat{J}_-)}$$ and corresponds to a rotation of $2\theta$ around the $\hat{J}_y$ axis. Consider now $$H_I|{\cal E}_0\rangle=SR\left (R^\dagger S^\dagger H SR\right )|j,-j\rangle_z\otimes|0\rangle_v$$ If we now transform the Hamiltonian and require that $$R^\dagger S^\dagger H SR|j,-j\rangle_z\otimes|0\rangle_v=0$$ we find the following conditions, $$\begin{aligned} \nu(1+\cos2\theta) & = & \mu(1-\cos2\theta)\\ \Omega j\sin2\theta & = & E\end{aligned}$$ which requires that $$\cos2\theta=e^{-2r}$$ and the ground state energy is taken to be ${\cal E}_0=0$. The ground state is thus a product of a squeezed state for the vibrational mode and a rotated angular momentum state, rotated about the $\hat{J}_y$ axis. The above results are consistent with the semiclassical approximation. The mean amplitude of a squeezed vacuum state is zero, corresponding to the semiclassical fixed point at $\bar{X}=\bar{Y}=0$ while the rotation around the $\hat{J}_y$ axis corresponds to the semiclassical fixed point at $\bar{{\cal J}}_x=-\chi$. If we continue to increase $E$ above the threshold value the system adiabatically follows a zero energy state, although this is no longer a ground state. In fact the canonical transformation used in the semiclassical analysis can be applied to the quantum operator valued Hamiltonian. The result is the same as the semiclassical case, Eq (\[sc\_ham\]) with all variables replaced with the corresponding operators. The zero energy state then corresponds to the zero energy eigenstate of $\hat{Y}\cos\theta-\hat{X}\sin\theta$ with $\cos\theta=N\Omega/2E$. This is of course just a rotated, infinitely squeezed state. The electronic state is likewise a angular momentum eigenstate rotated from $|j,-j\rangle$ in the equatorial plane (orthogonal to $\hat{J}_z$). Thus above threshold the zero energy eigenstate deforms continuously from the state at threshold. Let us summarise these results. For no driving the ground state corresponds to the oscillator in the ground state and all ions in the ground state. As the driving is increased, but kept below threshold, this state deforms to a squeezed oscillator state while the collective spin system begins to rotate about the $\hat{J}_y$ axis. Note that the mean oscillator amplitude $\langle a\rangle$ remains zero as does the mean of the $y$-component of the collective spin. As the driving increases through the threshold value, this state changes its character so that a non zero value of $\hat{J}_y$ is acquired and the oscillator is infinitely squeezed in a direction at an angle $\cos\theta=N\Omega/2E$ to the below threshold squeezing. This morphological change of the state as the driving passes the semiclassical critical point is a quantum phase transition. The quantum phase transition can be seen in the mean value for $\hat{J}_y$ and $\hat{J}_z$ as shown in figure \[fig\_two\]. Below threshold the scaled mean values are given by $$\begin{aligned} \frac{\langle \hat{J}_y\rangle}{N/2} & = & 0\\ \frac{\langle \hat{J}_z\rangle}{N/2} & = & -\sqrt{1-x^2}\end{aligned}$$ and above threshold we have $$\begin{aligned} \frac{\langle \hat{J}_y\rangle}{N/2} & = & -\sqrt{1-\frac{1}{x^2}}\\ \frac{\langle \hat{J}_z\rangle}{N/2} & = & 0\end{aligned}$$ where $x=2E/N\Omega$. What are the experimental manifestations of this transition ? Needless to say no one is ever going to observe an infinitely squeezed state in an experiment. So what does happens at $2\theta=\pi/2$ when the electronic state is the $\hat{J}_x$ eigenstate $|j-j\rangle_x$ and the vibrational mode appears to be infinitely squeezed ? Is such a state physically possible ? Suppose for example we begin in the ground state of the Hamiltonian with no driving ($E=0$) which is simply $|j,-j\rangle_z\otimes|0\rangle_v$, and adiabatically increase the driving strength. It would appear that the system would then adiabatically evolve into the squeezed vibrational state described above. If we were ever able to reach the case $2\theta=\pi/2$ we would have reached an infinite energy state for the vibrational mode at a finite driving strength. Clearly this is not possible and to understand why it is useful to reconsider the semiclassical dynamics for this model. The adiabatic approximation requires that we vary the driving strength on a time scale slower than all other time scales in the system. The key time scale for the ground state variation is just the time scale associated with the hyperbolic unstable fixed point, $(N^2/4-\chi^2)^{-1/2}$, which goes to infinity as we approach $\frac{2E}{\Omega N}=1$. Thus the adiabatic increase of the driving must proceed infinitely slowly, that is it must be switched to the finite value $E=\frac{\Omega N}{2}$ in an infinite amount of time. This pumps an infinite amount of energy into the system and results in infinite squeezing in the centre of mass vibrational mode. Obviously in practice this cannot be achieved so the totally squeezed ground state is not possible. However it will still be possible to achieve some squeezing of the vibrational mode at smaller values of the driving. This would make an interesting observation for current ion trap experiments even with only a few ions. The squeezing of the vibrational mode can be observed using the dynamical method of reference [@Wineland-squeeze] In current ion trap experiments, laser cooling techniques allow the centre of mass mode to be prepared in the ground state. Unfortunately it does not stay there. Heating due to a variety of sources, including fluctuating linear potentials, lead to an irreversible evolution away from the ground state. If such heating is present during the coupling of the electronic and vibrational motions, irreversible dynamics will be spread to the collective spin degrees of freedom as well. As an example we consider what happens if we use the Tavis-Cummings interaction (excitation on first red sideband) in the presence of strong heating. Heating of the centre-of-mass mode due to fluctuating liner potentials may be described in the interaction picture by the master equation, $$\frac{dW}{dt}=-i\Omega[a\hat{J}_++a^\dagger\hat{J}_-,W]+\frac{\gamma}{2}\left ({\cal D}[a]+{\cal D}[a^\dagger]\right )W$$ where $W$ is the density operator for the spin and vibrational degrees of freedom and the superoperator ${\cal D}$ is defined by $ {\cal D}[A]\rho=2A\rho A^\dagger-A^\dagger A\rho-\rho A^\dagger A\ \ . $ The irreversible term corresponds to two point processes in which phonons are removed or added from centre of mass mode at the rates $\gamma\langle a^\dagger a\rangle$ and $\gamma\langle a a^\dagger\rangle$ respectively. This does not change any first order moments, however it does lead to a diffusion in energy as $ \frac{d\langle a^\dagger a\rangle}{dt} =\gamma\ \ . $ The effect of heating can be included in the semiclassical analysis by adding an appropriate stochastic term. In the Ito calculus[@Gardiner] the effect is to add to the equations for $X,Y$ terms of the form $$\begin{aligned} dX & = & (\ldots)+\sqrt{\gamma}dW_x(t)\\ dY & = & (\ldots)+\sqrt{\gamma}dW_y(t)\end{aligned}$$ where $dW_i(t)$ are independent Wiener processes. If the heating rate is small enough these terms can be neglected. However if they are large new steady states can occur in the semiclassical and quantum descriptions which will be described in a future publication. We would like to thank Howard Carmichael for useful discussions. [8.]{} D.F.Walls, P.D.Drummond,S.S.Hassan and H.J.Carmichael, Progress of Theoretical Physics, supplement no. 64 (1978). H. Reiger and A.P. Young, in [*Complex Behaviour of Glassy Systems* ]{} edt. Rubi, M., Perez-Vicente, C., Proceedings of the XIV Sitges Conference, Sitges, Barcelona, Spain, 10-14 June., Lecture Notes in Physics (Springer Verlag, Berlin 1996), (LANL arhive; cond-mat/9607005). D.F.Walls and G.J.Milburn, [ *Quantum Optics*]{}, (Springer, Heidelberg, 1994) Q. A. Turchette, C. J. Hood, W. Lange, H. Mabuchi, and H. J. Kimble, Phys. Rev. Lett, [**75**]{}, 4710, (1995). D.J. Wineland, C. Monroe, W.M. Itano, D. Leibfried, B.E. King, and D.M. Meekhof, Journal of Research of the National Institute of Standards and Technology [**103**]{},259 (1998). R.J. Hughes, D.F.V. James, J.J. Gomez, M.S. Gulley, M.H. Holzscheiter, P.G. Kwiat, S.K. Lamoreaux, C.G. Peterson, V.D. Sandberg, M.M. Schauer, C.M. Simmons, C.E. Thorburn, D. Tupa, P.Z. Wang, and A.G. White, Fortschritte der Physik [**46**]{}, 329 (1998). M.Tavis and F.W. Cummings, Phys. Rev. [**170**]{}, 379 (1968). Mark S.Gulley, Andrew white, and Daniel F.V.James, “A Raman approach to quantum logic in Calcium-like ions”, submitted to J. Opt. Soc. Amer. B (1999). P.D.Drummond, Phys. Rev. A [**22**]{}, 1179 (1980), and references therein. R.H.Dicke, Phys. Rev. [**93**]{}, 99 (1954). Private communication. P.Alsing and H.J.Carmichael, Quantum Optics, [**3**]{}, 13 (1991). D. M. Meekhof, C. Monroe, B. E. King, W. M. Itano, and D. J. Wineland, Phys. Rev.Lett.,[**76**]{}, 1796 (1996). C.W.Gardiner, [*Handbook of stochastic methods*]{} (Springer-Verlag, Berlin, 1983).
{ "pile_set_name": "ArXiv" }
--- abstract: 'In a recently proposed effective equation of motion for the $4D$- Einstein-Gauss-Bonnet gravity admits a static black hole solution with two horizons. Thus central singularity is, as that for a charged black hole, timelike instead of spacelike for the Schwarzschild black hole. It turns out that the rescaled Gauss-Bonnet coupling constant acts as a new ’gravitational charge’ with a repulsive effect to cause in addition to event horizon a Cauchy horizon. Thus it radically alters the causal structure of the black hole.' author: - Naresh Dadhich title: 'On causal structure of $4D$-Einstein-Gauss-Bonnet black hole' --- It is well known that the Lovelock theory [@lovelock], whose action is a homogeneous polynomial in Riemann curvature, is the most natural higher dimensional generalization of the Einstein gravity — general relativity (GR). It has the remarkable property that despite action being polynomial in Riemann curvature, yet the equation of motion remains second order. However the higher order terms in the action make non-zero contribution in the equation only in dimension $D>2N$ where $N$ is the degree of curvature polynomial in the action. GR is linear while Gauss-Bonnet (GB) is quadratic order Lovelock and so on. Thus Lovelock is quintessentially a higher dimensional natural generalization of GR.\ It is however possible to make higher order terms contribute in the equation in $4D$ by dilaton coupling — a scalar field coupled to higher order term in the action, see for instance [@kanti]. Recently a new proposal has been made [@glavan] wherein GB term is made to contribute in $4D$ without dilaton coupling. In that the GB coupling is scaled as $\alpha \rightarrow \alpha/(D-4)$ and thereby cancelling out $(D-4)$ factor in the equation, and then taking the limit $D\to4$. This results into an effective equation in $4D$ which is in fact the Einstein-Gauss-Bonnet (EGB) equation written for $D=4$ [^1]. Then it could be solved in spacetime with some specific symmetries for different situations, black holes and cosmology. Firstly this way of taking limit is rather contentious and most importantly, there is no corresponding $4D$-action for the equation. Not withstanding all this, it has instantly caught up like a wild fire as is evidenced by the runaway activity [@fire] which is still going stronger by the day as there continues to be a steady flow of papers on the arxiv.\ On the other hand there are some serious questions being posed on the overall acceptability of the limiting process, validity of the equation in $4D$ as well as absence of proper action and a consistent theory in $4D$ [@ai; @gurses; @mann; @clifton; @mukohyama; @mahapatra; @kurt; @cano]. In particular It is fair to say that the jury is out on this issue, and we have to wait for some time before the air is cleared.\ In this brief note we wish to take up the issue of the static black hole solution of the new proposed $4D$-EGB equation that admits two horizons instead of the usual one for the Schwarzschild solution. On the other hand the EGB equation has only one horizon like the Schwarzschild in $D\geq5$ which is the natural rightful playground for it. In transition to $4D$, it acquires an extra Cauchy horizon which indicates presence of a ’new charge’. How does that arise physically and how do we understand it ? These are the most pertinent and critical questions.\ Let’s begin by recalling the $4D$-EGB static black hole metric [@glavan] as given by $$ds^2 = - f(r)dt^2 + \frac{dr^2}{f(r)} + r^2 d\Omega^2$$ where $$f(r) = 1+\frac{r^2}{2\alpha}\left(1\pm\sqrt{1+\frac{8\alpha M}{r^3}}\right).$$ In the usual notation $d\Omega^2$ is the metric on unit $2$-sphere and $\alpha$ and $M$ are the rescaled GB coupling constant and mass of isolated body respectively. We have also set $G=c=1$. The negative sign is chosen in the above solution for gravity being attractive.\ The black hole horizon is given by $f(r)=0$ which solves to give horizons as $$r_{h\pm} = M(1 \pm \sqrt{1 - \alpha/M^2}) .$$ Thus black hole has two horizons, unless of course $\alpha<0$ [^2], with the condition $M^2\geq \alpha$. The two merge into one-another for $M^2=\alpha$, defining the extremality condition. For $\alpha=0, M^2$, the event horizon is respectively $r_h = 2M, M$. This is exactly like the Reissner-Nordström metric where $\alpha$ has replaced charge $Q^2$. That means it has acquired gravitational charge character which is rather very strange and queer — a coupling constant being a charge! It produces repulsive effect exactly like the Maxwell charge for charged black hole so as to create a Cauchy horizon. In a different context the same repulsive effect has also been found [@seyed].\ Note that we have the famous Boulware-Deser EGB black hole [@boulware] solution which could be written for any $D>4$ with $$f(r) = 1 + \frac{r^2}{2\alpha}\left(1\pm\sqrt{1+\frac{8\alpha M}{r^{D-1}}}\right)$$ where $\alpha$ is normalized for a $D$ dependent numerical factor. It is this solution which is written in Eq. (2) above for $D=4$ as the solution of $4D$-EGB equation. The horizon equation $f(r)=0$ would then take the form for a generic dimension $D$, $$r^{D-3} + \alpha r^{D-5} - 2M = 0.$$ Clearly this equation admits only one positive root for $D\geq5$ while two positive roots for $D=4$. That means EGB black hole has only one event horizon for $D\geq5$ while it has two — both event and Cauchy horizons for $D=4$. Further it is clear from Eq. (4) that the metric is regular at $r=0$ for $D=4, 5$ [@dadhich2] but curvatures diverge with a lesser power of $r$ — singularity is weakened. This is the GB effect.\ In transition from five to four dimension, spacetime structure has radically changed. In the former the central singularity is spacelike as for the Schwarzschild black hole in any dimension $D\geq4$ in GR. So is also the case for EGB black hole in $D\geq5$. This means static black hole in Einstein gravity for $D\geq4$ — $D$-dimensional Schwarzschild and in EGB gravity $D$-dimensional Boulware-Deser for $D\geq5$ share the same causal structure [@torii] having spacelike central singularity. In contrast $4D$-EGB black hole has radically different causal structure with central singularity being timelike. It shares the structure instead of the Schwarzschild with Reissner-Nordström charged black hole.\ The key questions that arise are: Since no additional matter field has been introduced, what is it that causes this radical change in spacetime structure? The Cauchy horizon is always caused by some new ’charge’ like Maxwell charge or rotation which has opposing repulsive contribution to mass. In here nothing of that sort has happened, and it has been left to the GB coupling $\alpha$ to do the job. A coupling constant should not serve as gravitational charge, it is the measure of field’s linkage to matter.\ One of the motivations for higher curvature theories, in particular Lovelock, is that when one probes high energy strong field limit of GR, it is pertinent to include higher powers of Riemann curvature in action [@dadhich2]. Yet if the equation of motion should not change its second order character, the Lovelock generalization is then unique and it plays out only in higher dimensions. This is how higher dimensions, $D>4$, are innately tied to the Lovelock gravity. Of course one would like to bring down higher curvature effects that rule in higher dimensions to four dimensions to confront them against observations. That has always been accomplished, as mentioned earlier, through dilaton coupling — a scalar field coupled to higher order Lovelock term in action. What is envisaged is that the effect in four dimension should appear as correction to GR rather than a radical departure from it. This is because GR is so well grounded and solid against all the observations that there is absolutely no room for any substantial departure, what to talk of change in causal structure.\ From this standpoint what ensues in $4D$-EGB is entirely different and afront. It radically alters the causal structure by letting the GB coupling behave like a ’charge’. This is rather queer and strange, and hard to understand. We do however have a similar situation in the brane-world gravity model [@randall]. There a black hole on the $3$-brane does acquire a new charge — the Weyl charge that arises from the Weyl curvature of bulk spacetime [@dmpr]. The metric is exactly that of the Reissner-Nordström charged black hole. Then the sign of the Weyl charge, analogue of $Q^2$ is taken as negative so as to have only one horizon. It was envisaged that modification ensuing from bulk should only produce ’correction’ to GR black hole without any significant deviation. The Weyl charge is sourced by Weyl curvature of bulk spacetime. It is through the Weyl charge that higher dimensional bulk geometry manifests in four dimension as a correction to GR.\ Despite there being imprint of two charges on the horizon, yet the spacetime asymptotically goes over to the Schwarzschild solution with mass alone as the parameter. At infinity it is only mass that shows up as gravitational charge. Since no matter field of any kind has been added, there is no charge that can come into play. It is the GB gravitational coupling constant that manifests as charge on the horizon. Conceptually this is the most discomforting aspect of this proposal. Gravity resides in spacetime geometry while gravitational charge in matter fields. Here there is no identifiable matter field that could be responsible for creation of the Cauchy horizon.\ In a very recent paper [@ge], it has been shown by causal structure analysis of bulk spacetime that the GB coupling is bounded from above in AdS space, $\alpha\leq0$;i.e. it is non-positive. In that case there will be only one horizon thereby recovering the Schwarzschild causal structure. But then the GB gravity would be repulsive in stark opposition to the Einstein gravity. This is certainly not acceptable. Gravity should have the same character at all Lovelock orders, it cannot suddenly turn repulsive for $N=2$. If true, this raises a clear and sharp question for overall tenability of the $4D$-EGB proposal.\ We would not however like to address and opine on the issues of validity of the limiting process or absence of a valid action for the equation or a consistent theory in four dimension, and so on. Ours are basic and elemental concerns of concept and principle. With that in view we have raised certain questions and we believe that the whole efficacy of this new edifice rests on adequately addressing them.\ Acknowledgement {#acknowledgement .unnumbered} =============== We thank Sumanta Chakraborty for useful discussion. [99]{} D. Lovelock, The Einstein tensor and its generalizations, J. Math. Phys. [**12**]{}, 498 (1971). P. Kanti, N. Mavromatos, J. Rizos, K. Tamvakis, E. Winstanley, Dilatonic black holes in higher curvature string theory, Phys. Rev. [**D54**]{}, 5094(1996), \[hep-th/9511071\]; K.-i Maeda, N. Ohta, Y. Sasagawa, Black hole solutions in string theory with Gauss-Bonnet curvature correction, Phys. Rev. [**D80**]{}, 104032 (2009), \[arxiv:0908.4151\]; T. Sotiriou, S. Zhou, Black hole hair in generalized scalar-tensor gravity, Phys. Rev. Lett. [**112**]{}, 251102 (2014), \[arxiv:1312.3622\]; P. Kanti, R. Gannouji, N. Dadhich, Early-time cosmological solutions in Einstein-scalar-Gauss-Bonnet theory, Phys. Rev. [**D92**]{}, 083524 (2015), \[arxiv:1506.04667\]. D. Glavan and C. Lin, Einstein-Gauss-Bonnet gravity in 4-dimensional space-time, Phys. Rev. Lett. [**124**]{}, 081301 (2020), \[arXiv:1905.03601\]. C. Liu, T. Zhu, and Q. Wu, Thin Accretion Disk around a four-dimensional Einstein-Gauss-Bonnet Black Hole, \[arXiv:2004.01662\]; M. Guo and P.-C. Li, The innermost stable circular orbit and shadow in the novel 4D Einstein-Gauss-Bonnet gravity, \[arXiv:2003.02523\]; S. Wei, Y. Liu, Testing the nature of Gauss-Bonnet gravity by four dimensional rotating black hole shadow, \[2003.07769\]; R. Kumar and S. Ghosh, Rotating black holes in the novel 4D Einstein-Gauss-Bonnet gravity, \[arXiv:2003.08927\]; R. Konoplya, A. Zinhailo, Quasinormal modes, stability and shadows of a black hole in the novel 4D Einstein-Gauss-Bonnet gravity, \[arXiv:2003.01188\]; M. Churilova, Quasinormal modes of the Dirac field in the novel 4D Einstein-Gauss-Bonnet gravity, \[arXiv:2004.00513\], B. Toshmatov, D. Malafarina, N. Dadhich, Dust collapse in $4D$ Einstein-Gauss-Bonnet gravity, \[arxiv:2004.07089\]. W. Ai, A note on the novel 4D Einstein-Gauss-Bonnet gravity, \[arXiv:2004.02858\]. M. Gurses, T. C. Sisman, and B. Tekin, Is there a novel Einstein-Gauss-Bonnet theory in four dimensions?, \[arXiv:2004.03390\]. R. Hennigar, D. Kubiznak, R. Mann, C. Pollack, On taking the $D\to4$ limit of Gauss-Bonnet gravity: Theory and solutions, \[arxiv:2004.09422\]. P. Fernandes, P. Carrilho, T. Clifton, D. Mulryn, Derivation of regularized equations for the Einstein-Gauss-Bonnet theory in four dimensions, \[arxiv:2004.08362\]. K. Aoki, M. Gorji, S. Mukohyama, A consistent theory of $D\to4$ Einstein-Gauss-Bonnet gravity, \[arxiv:2005.03959\]. S. Mahapatra, A note on the total action of $4D$ Gauss-Bonnet theory, \[arxiv:2004.09214\]. J. Banifacio, K. Hinterbicher, L. Johnson, Amplitudes and $4D$ Gauss-Bonnet theory, \[arxiv:2004.10716\]. J. Arrechea, A. Delhom, A. Jamenez-Cano, Yet another comment on four dimensional Einstein-Gauss-Bonnet gravity, \[arxiv:2004.12998\]. N. Dadhich, Probing universality of gravity, in Proceedings of the 11th Regional conference on Mathematical Physics, eds. S. Rahvar, N. Soodooghi, F. Shojai (World Scientific, 2005), \[gr-qc/0407003\]; Universalization as a physical guiding principle, \[gr-qc/0311028\]. S. Mansoori, Thermodynamic geometry of novel $4D$ Guass-Bonnet AdS black hole, \[arxiv:2003.13382\]. D. Boulware, S. Deser, Phys. Rev. Lett. [**55**]{}, 2656 (1985). N. Dadhich, On the Gauss-Bonnet gravity, Proceedings of the 12th Regional Conference on Mathematical Physics, M. J. Islam, F. Hussain, A. Qadir, R. Riazuddin and H. Saleem editors, pp.331 (World Scientific, 2006), \[gr-qc/0509126\]. T. Torii and H. Maeda, Spacetime structure of static solutions in Gauss-Bonnet gravity: Neutral case Phys. Rev. D [**71**]{}, 124002 (2005), \[hep-th/0504127\]. L. Randall, R. Sundrum, Phys. Rev. Lett. [**83**]{}, 3370,\[hep-th/9905221\] 4690,\[hep-th/9906064\] (1999). N. Dadhich, R. Maartens, P. Papadopoulos, V. Rezania, Phys. Lett. [**487**]{}, 1 (2000), \[hep-th/0003061\]. X-H. Ge, S-J. Sin, Causality of black holes in $4$-dimensional Einstein-Gauss-Bonnet-Maxwell theory, \[arxiv:2004.12191\]. [^1]: The equation is non-vacuous only in $D>4$ because of the multiplicative factor $(D-4)$, which has now been cancelled out by rescaling of $\alpha$. [^2]: Since gravity is universal and hence always attractive [@dadhich1], its coupling constant should always have the same sign for all Lovelock orders $N$. Therefore $\alpha$ cannot be negative. Since it is a dimensionful, its dimension will however depend on order $N$ and, consequently or otherwise also on spacetime dimension.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Recently, there has been quite a lot of interest in static, spherical wormhole spacetimes and the question of their stability with respect to time-dependent perturbations. The consideration of linearized perturbations usually leads to a master wave equation with effective potential which can then be analyzed using standard tools from quantum mechanics. However, in the wormhole case, particular care must be taken with the gauge conditions when formulating the master equation. A poor coordinate choice, based for example on fixing the areal radial coordinate, may lead to singularities at the throat which complicates the stability analysis or might even lead to erroneous conclusions regarding the stability of the underlying wormhole configuration. In this work, we present a general method for deriving a gauge-invariant wave system of linearized perturbation equations in the spherically symmetric case, assuming that the matter supporting the wormhole is a phantom scalar field, that is, a self-interacting scalar field whose kinetic energy has the reversed sign. We show how this system can be decoupled and reduced to a single master wave equation with a regular potential, with no intermediate steps involving singularities at the throat. Two applications of our formalism are given. First, we re-derive the master equation for the linearly perturbed Ellis-Bronnikov wormhole using our new, singularity-free method. As a second application, we derive the master equation describing the linear perturbations of a certain AdS-wormhole, provide a detailed analysis of the spectral properties of the underlying operator and prove that, as in the Ellis-Bronnikov case, this wormhole is linearly unstable and possesses a single unstable mode.' author: - Francesco Cremona - Livio Pizzocchero - Olivier Sarbach bibliography: - 'refs\_wormholes.bib' title: 'Gauge-invariant spherical linear perturbations of wormholes in Einstein gravity minimally coupled to a self-interacting phantom scalar field' --- Introduction ============ One of the most fascinating features of Einstein’s theory of General Relativity (GR) consists in the fact that spacetime may be curved and topologically non-trivial, describing intriguing objects like black holes and wormholes. Black hole spacetimes appear under rather natural conditions in GR, and they are expected to form in Nature, for instance, by the collapse of sufficiently massive stars at the end of their life. Furthermore, there is by now compelling evidence for their existence in our Universe which has recently been reinforced by the observation of gravitational waves from binary black hole mergers [@Abbott:2016blz] and the first image of the shadow of the supermassive black hole in the center of the galaxy M87 [@Akiyama:2019cqa]. In contrast to this, the occurrence of wormholes ([^1]) is much more speculative, and so far, there is no observational evidence for the existence of such structures. From the theoretical point of view, there are important constraints, such as the topological censorship theorem [@jFkSdW93]. This theorem implies that asymptotically flat, globally hyperbolic wormhole spacetimes (including those relevant to this paper whose Cauchy surfaces have topology ${\mathbb{R}}\times S^2$, representing a throat connecting two asymptotically flat ends) require the existence of “exotic” matter to support the throat, that is, they require matter whose stress-energy-momentum tensor violates the (averaged) null energy condition. Intuitively, the need for exotic matter can be understood by the fact that a light bundle that traverses a wormhole throat must focus as it approaches the throat, but then must expand again as it moves away from the throat, which is opposite to the focusing effect for light due to ordinary matter [@HawkingEllis-Book]. On the other hand, it has also been shown that an infinitesimally small quantity of matter violating the averaged null condition is sufficient to support the throat [@mVsKnD03]. This leads to the hope that quantum effects may give rise to a semiclassical theory in which wormhole spacetimes are allowed, in a similar way than quantum effects (Hawking radiation) induce black hole evaporation although an area decrease of the event horizon is forbidden in classical GR with matter fields satisfying the null energy condition [@HawkingEllis-Book]. Nevertheless, it remains to be seen whether or not such effects are strong enough to give rise to a traversable wormhole throat of macroscopic size [@eFrW96]. Instead of invoking quantum effects, an alternative way to violate the null energy condition (which has received important motivation from cosmology, see for example [@Caldwell:1999ew]) is the consideration of phantom scalar fields, i.e. scalar fields that have a negative kinetic energy (see for instance [@Bro2018] and references therein). Due to this property, such fields may lead to gravitational repulsion, and hence induce interesting effects like the accelerated expansion in the universe, universes with no particle horizon [@dFmGlP19] or the ability of supporting a wormhole throat [@hE73; @kB73]. On the other hand, the presence of unbounded negative kinetic energy might cast doubt on the possibility that any stationary solution found in this theory could ever be stable. ([^2]) Therefore, a pressing question regarding the relevance of static wormhole solutions in such theories (or other GR theories involving exotic matter fields) is their dynamical stability under small perturbations. The most widely studied wormhole models (including those analyzed in the present article) are based on static, spherically symmetric spacetimes in which the world sheet of the throat consists of spheres of minimal area [@mMkT88]. Within the context of phantom scalar fields, many such solutions have been found; the simplest ones are obtained for a real scalar field and are due to pioneering work by Ellis [@hE73] and by Bronnikov [@kB73]. Since then, these solutions have been generalized to arbitrary dimensions [@jEtZ08] and to the following supporting fields: a scalar with a self-interaction potential [@Dzhunushaliev:2008bq; @kBrKaZ12], a complex phantom scalar [@Dzhunushaliev:2017syc], a family of conventional and/or phantom scalars [@hSsH02; @oStZ12; @Carvente:2019gkd], a phantom scalar and an electromagnetic field [@jGfGoS09c], and, very recently, a k-essence scalar [@kBjFdR2019]. For the linear stability analysis of many of these solutions, see Refs. [@jGfGoS09a; @jGfGoS09c; @oStZ12; @kBjFaZ11; @kBrKaZ12; @fCfPlP19; @kBjFdR2019]. All these studies conclude that the static, spherically symmetric wormhole solutions are linearly unstable, with numerical simulations [@hSsH02; @jGfGoS09b; @aDnKdNiN09] revealing that the throat either collapses to a black hole or expands on timescales comparable to the light-crossing time of the radius of the throat. Therefore, finding a static, spherically symmetric wormhole solution in GR with exotic matter which can be shown to be linearly stable (or unstable with a large time scale associated with all the unstable modes) remains a challenging open problem. ([^3]) In this work, we focus on GR minimally coupled to a single, real phantom scalar field $\Phi$ under the influence of an arbitrary potential $V(\Phi)$ and provide a general framework to analyze the stability of static, spherically symmetric wormhole solutions in these theories. The first result of this paper is the derivation of a coupled, $2\times 2$ linear wave system subject to a constraint, describing the dynamics of linearized, time-dependent perturbations of such solutions in terms of two gauge-invariant linear combination of the linearized metric coefficients and of the scalar field. Apart from being gauge-invariant, a key feature of this system relies in the fact that it is *regular* at the throat, provided the scalar field does not have a critical point there. This is a non-trivial achievement, since most direct approaches to the linearized equations are based on fixing the radial coordinate and deriving a wave equation for the linearized scalar field. Due to the fact that the radial coordinate has a critical point at the throat, the effective potential appearing in this wave equation is necessarily *singular* at the throat. As explained in [@jGfGoS09a] (see also [@kBjFaZ11]) this yields an artificial (mirror-like) boundary condition at the throat which prevents perturbations from traversing the wormhole. This artificial boundary condition effectively restricts the class of physically admissible perturbations, and as it turns out, the unstable modes associated with the wormholes is precluded from this class, leading to the erroneous conclusion that the wormhole is linearly stable. To overcome these problems, a method for transforming the singular wave equation to a regular one was introduced in [@jGfGoS09a] to treat the linearized perturbations of the Ellis-Bronnikov wormhole; this approach was subsequently generalized and referred to as S-deformation method" in [@kBjFaZ11]. The present work proposes an alternative to the S-deformation method, based on the important assumption that the scalar field has no critical points, in which singularities (at the throat or elsewhere) do not appear at any stage of the derivation. The second result of our work is that, provided a non-trivial time-independent solution of the coupled $2\times 2$ system is known, it is possible to decouple the system, obtaining a single wave equation for an appropriate, gauge-invariant linear combination of the perturbed fields, from which all other perturbations can be reconstructed. In most situations, such a time-independent solution can be found by varying the parameters of a known family of static solutions. For the Ellis-Bronnikov wormhole, we show that the master equation obtained by our method agrees precisely with the one obtained in [@jGfGoS08; @jGfGoS09a; @fCfPlP19] by different means. Furthermore, we show that our gauge-invariant method for obtaining a master equation through the decoupling of the $2\times 2$ system also works for wormhole solutions whose stability has not been addressed so far. As an explicit example, we consider a static, spherically symmetric AdS-type wormhole which connects two asymptotic AdS ends, and prove that it is linearly unstable as well. Finally, we provide a detailed analysis for the behavior of the solution of the master equations in both the Ellis-Bronnikov and the AdS case, based on a rigorous spectral analysis of the Schrödinger operator appearing therein. A negative eigenvalue of the Schrödinger operator gives rise to a pair of modes, one exponentially growing and the other one exponentially decaying with respect to the time variable; a positive eigenvalue gives rise to a pair of oscillating modes, while a positive energy level lying in the continuous spectrum gives rise to a pair of non-normalizable oscillating modes, corresponding to generalized eigenfunctions of the Schrödinger operator. If zero is an eigenvalue it gives rise to a pair of normalizable modes, one of them constant and the other one linearly growing with time. We show that in the Ellis-Bronnikov case, the solution can be expanded in terms of an exponentially growing, an exponentially decaying, a constant, a linearly growing mode and a continuum of oscillators associated with non-normalizable modes. In contrast to this, in the AdS case the spectrum is a pure point spectrum giving rise to an exponentially growing, an exponentially decaying, and to an infinite, discrete set of oscillating normalizable modes. This is due to the Dirichlet-type boundary conditions imposed at the AdS boundary, which give rise to a regular Sturm-Liouville problem. The article is organized as follows. In section \[Sec:SphSymFEQ\] we specify our metric ansatz, make a few general comments regarding the coordinate conditions that will be relevant in this work and derive the field equations for a spherically symmetric, time-dependent configuration. In section \[Sec:Static\] we discuss two static wormhole solutions that will serve as examples and applications for our perturbation formalism and stability analysis: the Ellis-Bronnikov solution and a wormhole connecting two AdS universes. In section \[Sec:LinearPerturbation\] we derive the relevant set of linearized equations in a gauge-fixed setting in which the scalar field is held fixed. In section \[Sec:GaugeInvariant\] we introduce a set of combinations of the linearized fields which are invariant with respect to infinitesimal coordinate transformations, and the linearized field equations are cast into a constrained wave system for two of these gauge-invariant fields. In section \[Sec:Decoupling\] we show how to decouple this wave system, provided a static solution of the linearized field equations is available, in which case a single master wave equation is obtained. This method is then applied to the examples of section \[Sec:Static\], and it is shown that in each case the associated Schrödinger operator possesses a unique bound state with negative energy, implying that these wormholes are linearly unstable. In section \[Sec:Spectral\] we provide a detailed discussion on the spectral decomposition of the Schrödinger operator and the corresponding master equations (based on rigorous techniques from functional analysis) and contrast the Ellis-Bronnikov case with the one of the AdS wormhole. Conclusions, limitations and possible future applications of our method are given in section \[Sec:Conclusions\]. Technical details regarding the spectral theory of Schrödinger operators are given in the appendices. Throughout this work, we use the signature convention $(-,+,+,+,)$ and choose units in which $c=1$, $\hbar=1$. Spherically symmetric field equations and background {#Sec:SphSymFEQ} ==================================================== We consider a four-dimensional spacetime $(M,{\bf g})$ in which the gravitational field ${\bf g}$ is minimally coupled to a massless phantom scalar field $\Phi$, that is, a scalar field with the reversed sign in its kinetic term that self-interacts according to a potential $V(\Phi)$. The action functional of this system is $$S[\mathbf{g},\Phi]:=\int\bigg(\frac{R}{2\kappa} +\frac{1}{2} \nabla^\mu\Phi\cdot \nabla_\mu \Phi -V(\Phi)\bigg)dv\,,$$ where $\kappa = 8\pi G$ is the usual coupling constant while $R$ and $dv=\sqrt{|\textnormal{det}(g_{\mu\nu})|}\prod\limits_{\mu=0}^3 dx^{\mu}$ are the scalar curvature and the volume element associated with the metric $\mathbf{g}$. The corresponding field equations are $$\begin{aligned} R_{\mu\nu} &=& \kappa\left[ -\nabla_\mu\Phi \cdot \nabla_\nu\Phi + V(\Phi) g_{\mu\nu} \right]\,, \label{Eq:Einstein}\\ 0 &=& \nabla^\mu\nabla_\mu\Phi + V'(\Phi)\,, \label{Eq:KleinGordon}\end{aligned}$$ with $\nabla_\mu$ and $R_{\mu\nu}$ denoting the covariant derivative and Ricci tensor, respectively, associated with ${\bf g}$. In this work, we focus on spherically symmetric spacetimes $(M,{\bf g})$ of the form $M = \tilde{M}\times S^2$ with metric $${\bf g} = -\alpha(t,x)^2 dt^2 + \gamma(t,x)^2 \left( dx + \beta(t,x) dt \right)^2 + r(t,x)^2\left( d\vartheta^2 + \sin^2\vartheta\; d\varphi^2 \right)\,, \label{Eq:SphericalMetric}$$ which, in a general spherically symmetric coordinate system $(t,x,\vartheta,\varphi)$, is parametrized in terms of the four functions $\alpha,\beta,\gamma,r$ on the two-dimensional manifold $\tilde{M}$. Of course, the number of these functions can be reduced from four to two by an appropriate choice of the coordinates $(t,x)$ on $\tilde{M}$. There are several “natural" choices one can make. For example, given a smooth function $f: \tilde{M}\to {\mathbb{R}}$ with the property that its gradient is everywhere spacelike, one can choose an orthogonal coordinate system $(t,x)$ on $\tilde{M}$ such that $x = f$ and $\beta = 0$. (Likewise, if the gradient of $f$ is everywhere timelike one can choose $(t,x)$ such that $\beta = 0$ and $t = f$.) In particular, if the gradient of the areal radius $r$ is everywhere spacelike one can choose $f = r$ and one is left with the two functions $\alpha$ and $\gamma$ on $\tilde{M}$. Usually, however, the gradient of $r$ is not everywhere spacelike due to the presence of minimal or trapped surfaces, and the resulting coordinate system is only locally defined on $\tilde{M}$. The field equations (\[Eq:Einstein\],\[Eq:KleinGordon\]) for a spherically symmetric metric (\[Eq:SphericalMetric\]) in any gauge such that $\beta = 0$ can be written as $$\begin{aligned} \frac{\partial}{\partial t}\left( \frac{\dot{\gamma}}{\alpha} \right) - \frac{\partial}{\partial x}\left( \frac{\alpha'}{\gamma} \right) - \frac{\gamma}{\alpha}\frac{\dot{r}^2}{r^2} + \frac{\alpha}{\gamma}\frac{r'^2}{r^2} - \frac{\alpha\gamma}{r^2} &=& \frac{\kappa}{2}\left[ \frac{\gamma}{\alpha}\dot{\Phi}^2 - \frac{\alpha}{\gamma}\Phi'^2 \right]\,, \label{Eq:Ev1}\\ \frac{\partial}{\partial t} \left[ \frac{\gamma}{\alpha} r\dot{r} \right] - \frac{\partial}{\partial x} \left[ \frac{\alpha}{\gamma} r r' \right] &=& \alpha\gamma\left( \kappa r^2 V(\Phi)-1\right)\,, \label{Eq:Ev2}\\ \frac{\partial}{\partial t} \left[ \frac{\gamma}{\alpha} r^2\dot{\Phi} \right] - \frac{\partial}{\partial x}\left[ \frac{\alpha}{\gamma} r^2\Phi' \right] &=& \alpha \gamma r^2 V'(\Phi)\,, \label{Eq:Ev3}\end{aligned}$$ with the constraints $$\begin{aligned} {\cal H} &:=& \frac{\alpha}{\gamma}\left[ 2\frac{r''}{r} + \frac{r'}{r}\left( \frac{r'}{r} - 2\frac{\gamma'}{\gamma} \right) \right] - \frac{\gamma}{\alpha}\frac{\dot{r}}{r}\left( \frac{\dot{r}}{r} + 2\frac{\dot{\gamma}}{\gamma} \right) - \frac{\alpha\gamma}{r^2} -\frac{\kappa}{2}\left[ \frac{\gamma}{\alpha}\dot{\Phi}^2 + \frac{\alpha}{\gamma}\Phi'^2 \right] + \kappa\alpha \gamma V(\Phi) = 0\,, \label{Eq:Evh} \\ {\cal M} &:=& 2\frac{\dot{r}'}{r} - 2\frac{\dot{r}}{r}\frac{\alpha'}{\alpha} - 2\frac{r'}{r}\frac{\dot{\gamma}}{\gamma} -\kappa\dot{\Phi}\Phi' = 0\,. \label{Eq:Evm}\end{aligned}$$ Here and in the following, a dot and a prime refer to partial differentiation with respect to $t$ and $x$, respectively. In the conformally flat gauge, in which $\alpha=\gamma$, Eqs. (\[Eq:Ev1\],\[Eq:Ev2\],\[Eq:Ev3\]) yield a hyperbolic wave system for the quantities $(\alpha,r,\Phi)$ which is subject to the constraints (\[Eq:Evh\],\[Eq:Evm\]). This system (or slight variants thereof) is suitable for numerical time evolutions, see for instance [@jGfGoS09b]. Static wormhole solutions {#Sec:Static} ========================= In this section, we discuss two examples of static wormhole solutions that have been considered previously in the literature: the Ellis-Bronnikov wormhole connecting two asymptotically flat ends [@hE73; @kB73] and a reflection-symmetric wormhole connecting two AdS ends [@Bro2018]. Ellis-Bronnikov wormhole {#Subsection:Ellis} ------------------------ Let us assume a zero potential: $V(\Phi)=0$. In the static case, the functions $\alpha$, $\gamma$ and $r$ are $t$-independent and one can further adjust the coordinate $x$ so that $\alpha\gamma = 1$. In this case, the field equations can be reduced to the three differential equations $$[ \alpha^2 r^2]'' = 2\,,\qquad [ \alpha^2 r r']' = 1\,,\qquad [\alpha^2 r^2\Phi']' = 0\,,$$ which arise, respectively, from a recombination of Eqs. (\[Eq:Ev1\],\[Eq:Ev2\],\[Eq:Evh\]), from Eq. (\[Eq:Ev2\]) and from Eq. (\[Eq:Ev3\]). These can easily be integrated with the result $$\alpha = \gamma^{-1} = e^{\gamma_1 \arctan {x \over b}}\,,\qquad r^2 = (x^2 + b^2)\gamma^2\,,\qquad \Phi = \Phi_1\arctan {x \over b}\,. \label{Eq:StaticSolutions}$$ Here, $b > 0$, $\gamma_1$ and $\Phi_1$ are integration constants, and the Hamiltonian constraint ${\cal H} = 0$ enforces the relation $\kappa\Phi_1^2 = 2(1 + \gamma_1^2)$ (while the momentum constraint ${\cal M}=0$ is obviously satisfied). These solutions were obtained long time ago by Ellis [@hE73] and Bronnikov [@kB73], and they describe traversable wormholes whose throat is located at $x = \gamma_1 b$ (see also [@jGfGoS09a] for their physical properties). The reflection-symmetric case $\gamma_1 = 0$ for which $\alpha = \gamma = 1$ results in a particularly simple form of the wormhole metric which has been posed as an exercise in general relativity in the popular article by Morris and Thorne [@mMkT88]. A wormhole connecting two AdS universes {#Subsection:AdS} --------------------------------------- We now look for a static solution in the gauge $\alpha \gamma=1$, allowing $V(\Phi)$ to be nonzero. Let us show that a simple solution of this form can be obtained by putting as before $r = \sqrt{x^2+b^2}$, where $b>0$. With these choices it is easy to show that the combination (Eq. (\[Eq:Ev2\])$+r^2$Eq. (\[Eq:Evh\])) is satisfied if $\Phi = \sqrt{2/\kappa}\arctan(x/b)+\Phi_0$ with $\Phi_0$ a constant. With this expression for the scalar field, Eq. (\[Eq:Ev1\]) leads to $$\alpha=\sqrt{1-K \left(b^2+x^2\right)+M \left(b^2+x^2\right) \arctan\frac{x}{b} + b M x}\,,$$ where $K$ and $M$ are two constants. The remaining two equations Eq. (\[Eq:Ev2\]) (or, alternatively, Eq. (\[Eq:Evh\])) and Eq. (\[Eq:Ev3\]) can be solved putting $$V(\Phi(x))=\frac{ K(b^2+3 x^2) -M \left(b^2+3 x^2\right) \arctan\frac{x}{b} - 3 b M x}{\kappa\left(b^2+x^2\right)}\,.$$ Choosing, without loss of generality, $\Phi_0=0$, we obtain for $V(\Phi)$ $$V(\Phi)=\frac{K}{\kappa}\left[ 3 -2 \cos^2\left( \sqrt{\frac{\kappa}{2}}\Phi\right) \right] - \frac{M}{\kappa}\left\{ 3 \sin \left( \sqrt{\frac{\kappa}{2}}\Phi\right) \cos \left( \sqrt{\frac{\kappa}{2}}\Phi\right)+ \sqrt{\frac{\kappa}{2}}\Phi\left[ 3 -2 \cos ^2\left( \sqrt{\frac{\kappa}{2}}\Phi\right)\right]\right\}\,.$$ Actually, this solution is exactly the general solution given by Bronnikov and Fabris in [@kBjF06] and reconsidered in the recent survey [@Bro2018] (with some reparametrization of the involved constants). From here to the end of the paper we focus on the choice $$M = 0\,,\qquad K\equiv -k^2<0\,,$$ (with $k>0$) corresponding to a wormhole metric which is reflection-symmetric with respect to the throat. In this case, the solution simplifies to $$V(\Phi)= -\frac{k^2}{\kappa} \left[ 3 -2 \cos ^2\left( \sqrt{\frac{\kappa}{2}}\Phi\right) \right]\,,\quad \alpha = \gamma^{-1} = \sqrt{1 + k^2(x^2+b^2)}\,,\quad r = \sqrt{x^2 + b^2}\,,\quad \Phi = \sqrt{\frac{2}{\kappa}}\arctan\frac{x}{b}\,. \label{Eq:StaticSolutionsPot1}$$ In the limit case $b=0$ we should replace the third equality in (\[Eq:StaticSolutionsPot1\]) by $r=x>0$; the corresponding metric describes an Anti de Sitter (AdS) universe with cosmological constant $\Lambda = -3k^2$. From now on we intend $$b > 0\,,\qquad -\infty < x < +\infty\,;$$ since $r(x)\sim |x|$ for $x\to \pm \infty$, we can interpret the metric in (\[Eq:StaticSolutionsPot1\]) as describing a wormhole connecting two separate asymptotically AdS universes with the same cosmological constant $\Lambda = -3k^2$ and minimal areal radius $b$ at the throat. For this reason one could call the solution (\[Eq:StaticSolutionsPot1\]) an “AdS-AdS wormhole”; in the sequel this expression will be always shortened into “AdS wormhole”. Let us note that, for $k\to0$, the potential $V(\Phi)$ vanishes and the AdS wormhole (with $b$ fixed) becomes the reflection-symmetric Ellis-Bronnikov wormhole (as in Eq. (\[Eq:StaticSolutions\]), with $\gamma_1=0$). For further convenience, we introduce the change of variable $$t = \frac{s}{2k \sqrt{1+B^2}}\,,\qquad x=\frac{\sqrt{1+B^2}}{k}\tan{\frac{u}{2}}\,,\qquad B:=b k \,,\qquad s\in (-\infty,\infty)\,,\quad u\in (-\pi,\pi)\,,$$ so that in the new coordinates system the metric corresponding to the solution (\[Eq:StaticSolutionsPot1\]) is transformed into a metric of the form (\[Eq:SphericalMetric\]) with $(t,x)$ replaced by $(s,u)$ and $$\alpha=\gamma = \frac{1}{2 k \cos{\frac{u}{2}}}\,,\qquad \beta = 0\,,\qquad r= \frac{\sqrt{1+2 B^2-\cos u}}{\sqrt{2} k \cos \frac{u}{2}}\,,\qquad\Phi=\sqrt{\frac{2}{\kappa}}\arctan\left(\frac{\sqrt{1+B^2}\tan \frac{u}{2}}{B} \right) \label{Eq:StaticSolutionsPot2}$$ (of course, $V(\Phi)$ is still as in (\[Eq:StaticSolutionsPot1\])). Let us observe that the limits $x\to\pm\infty$, describing the far ends of the wormhole, are equivalent to the limits $u\to\pm \pi$. Linear perturbations and the $\delta\Phi = 0$ gauge {#Sec:LinearPerturbation} =================================================== In the sequel we consider, for an arbitrary potential $V(\Phi)$, a family of static solutions $(\alpha,\gamma,r,\Phi)$ of Eqs. (\[Eq:Ev1\]-\[Eq:Evm\]) (without necessarily assuming the gauge condition $\alpha\gamma = 1$). This family may depend on certain parameters (like the constants $b,\gamma_1$ in subsection \[Subsection:Ellis\] or the parameter $B$ in subsection \[Subsection:AdS\]). In addition, we consider a (non-static) perturbation $(\delta\alpha,\delta\gamma,\delta r,\delta\Phi)$ of this static solution, which is treated linearizing Eqs. (\[Eq:Ev1\]-\[Eq:Evm\]); let us recall that Eqs. (\[Eq:Ev1\]-\[Eq:Evm\]) assume $\beta = 0$ for the metric (\[Eq:SphericalMetric\]), so their linearization corresponds to taking $\delta \beta=0$. For the particular case in which the potential vanishes ($V = 0$) it can be shown (see e.g. Ref. [@jGfGoS09a]) that the linearized constraint equations $\delta {\cal H} = \delta {\cal M} = 0$ can be integrated. It turns out this is still the case for solutions with a non-trivial potential, yielding the conclusion that $$\sigma:={\alpha r \over \gamma}\left( \delta r'-\frac{\alpha'}{\alpha}\delta r -r'\frac{\delta\gamma}{\gamma}- \frac{\kappa}{2} r\Phi'\delta\Phi \right) \label{Eq:Pert1}$$ is a constant. This constant indeed describes a zero mode, that is, a perturbation corresponding to an infinitesimal variation of the parameters labelling the static solution (see section 3.1 of [@jGfGoS09a] for more details in the $V=0$ case). Since we are mainly interested in dynamical perturbations (rather than infinitesimal deformations along the static branch in the solution space), we assume from now on that $$\sigma = 0\,. \label{Eq:Sigma0}$$ For future use, it is advantageous to introduce the quantities ([^4]) $${\mathcal{D}}:= \frac{\delta\alpha}{\alpha}\,,\qquad {\mathcal{A}}:= \frac{\delta\gamma}{\gamma}\,,\qquad {\mathcal{C}}:= \frac{\delta r}{r}\,. \label{Eq:DAC}$$ Then Eqs. (\[Eq:Pert1\]-\[Eq:Sigma0\]) become $$\sigma=0\,,\qquad \sigma:={\alpha r^2 \over \gamma}\left[ {\mathcal{C}}' - \left(\frac{\alpha'}{\alpha} - \frac{r'}{r} \right) {\mathcal{C}}- \frac{r'}{r} {\mathcal{A}}- \frac{\kappa}{2}\Phi'\delta\Phi\right]\,; \label{Eq:Pert1Bis}$$ moreover, the linearization of Eqs. (\[Eq:Ev1\],\[Eq:Ev2\],\[Eq:Ev3\]) and the condition $\sigma=0$ give the following linear system of equations: $$\begin{aligned} \frac{\gamma}{\alpha}\ddot{{\mathcal{A}}} - \frac{\partial}{\partial x}\left( \frac{\alpha}{\gamma} {\mathcal{D}}' \right) - \frac{\alpha'}{\gamma}({\mathcal{D}}- {\mathcal{A}})' + 2\frac{\alpha}{\gamma}\frac{r'}{r} {\mathcal{C}}' - \frac{2\alpha\gamma}{r^2}({\mathcal{A}}- {\mathcal{C}}) + \kappa\frac{\alpha}{\gamma}\Phi'\delta\Phi' &=& 0\,, \label{Eq:A}\\ \frac{\gamma}{\alpha}\ddot{{\mathcal{C}}} - \frac{\partial}{\partial x}\left( \frac{\alpha}{\gamma} {\mathcal{C}}' \right) - \frac{\alpha}{\gamma}\frac{r'}{r}({\mathcal{D}}- {\mathcal{A}}+ 4{\mathcal{C}})' + \frac{2\alpha\gamma}{r^2}({\mathcal{A}}- {\mathcal{C}}) -\kappa\alpha\gamma\left[2V(\Phi) {\mathcal{A}}+ V'(\Phi)\delta\Phi \right] &=& 0\,, \label{Eq:C}\\ \frac{\gamma}{\alpha}\ddot{\delta\Phi} - \frac{\partial}{\partial x}\left( \frac{\alpha}{\gamma} \delta\Phi' \right) - 2\frac{\alpha}{\gamma}\frac{r'}{r}\delta\Phi' - \frac{\alpha}{\gamma}\Phi'( {\mathcal{D}}- {\mathcal{A}}+ 2{\mathcal{C}})' - \alpha\gamma\left[ 2V'(\Phi) {\mathcal{A}}+ V''(\Phi)\delta\Phi \right] &=& 0\,. \label{Eq:Phi}\end{aligned}$$ All the equations derived so far only assume the orthogonal gauge $\beta = 0$; at the linearized level there is still liberty which is related to the choice of a function $f$ on $\tilde{M}$, as explained in section \[Sec:SphSymFEQ\]. One possible choice is fixing the areal radius function $r(x)$ to its background form, such that $\delta r = 0$ and ${\mathcal{C}}= 0$. Eqs. (\[Eq:Pert1Bis\],\[Eq:C\]) then allow to express the metric fields ${\mathcal{A}}$ and ${\mathcal{D}}'$ in terms of $\delta\Phi$ and one obtains a master equation for the linearized scalar field $\delta\Phi$ (see e.g. [@Bro2018]). However, in this article, we are interested in deriving a master equation describing the dynamics of the linear perturbations of anyone of the two wormholes in the previous subsections \[Subsection:Ellis\] and \[Subsection:AdS\]. Since these solutions have $dr = 0$ at the wormhole throat, fixing the areal radius function $r(x)$ amounts in forcing the perturbations to vanish at the throat, which from a physical point of view is much too restrictive. At the mathematical level, enforcing the $\delta r = 0$ gauge results in a master equation for $\delta\Phi$ with a potential that is singular at the throat (see [@jGfGoS09a; @Bro2018] for more details). On the other hand, while $dr = 0$ at the wormhole throat, we note from Eq. (\[Eq:StaticSolutions\]) or Eq. (\[Eq:StaticSolutionsPot1\]) that $d\Phi =$ const$ \times dx/(x^2 + b^2)$ is everywhere spacelike, and hence the same will be true for sufficiently small perturbations of the static wormhole solution. As a consequence, we may choose the coordinates $(t,x)$ such that $\Phi$ is given by exactly the same expression as in Eq. (\[Eq:StaticSolutions\]) or Eq. (\[Eq:StaticSolutionsPot1\]), even for the perturbed spacetime. This implies, in particular, that $\delta\Phi = 0$. In this gauge, Eqs. (\[Eq:Pert1Bis\],\[Eq:Phi\]) reduce to $$\begin{aligned} && \sigma=0,\qquad \sigma:={\alpha r^2 \over \gamma}\left[ {\mathcal{C}}' - \left(\frac{\alpha'}{\alpha} - \frac{r'}{r} \right) {\mathcal{C}}- \frac{r'}{r} {\mathcal{A}}\right]\,, \label{Eq:Pert3a}\\ && {\mathcal{D}}' - {\mathcal{A}}' + 2{\mathcal{C}}' + 2 \gamma^2{ V'(\Phi)\over \Phi'} {\mathcal{A}}= 0\,; \label{Eq:Pert3b}\end{aligned}$$ using these equations in order to eliminate ${\mathcal{C}}'$ and ${\mathcal{D}}'$ and the static version of Eq. (\[Eq:Ev3\]) (from which one can eliminate the unperturbed quantity $\Phi''$), Eqs. (\[Eq:A\],\[Eq:C\]) reduce to $$\begin{aligned} \frac{\gamma}{\alpha}\ddot{{\mathcal{A}}} - \frac{\partial}{\partial x}\left[ \frac{\alpha}{\gamma}({\mathcal{A}}' - 2{\mathcal{C}}') \right] + 2\frac{\alpha}{\gamma}\left( \frac{\alpha'}{\alpha} + \frac{r'}{r} \right) {\mathcal{C}}' {- 2\frac{\alpha\gamma}{r^2}({\mathcal{A}}- {\mathcal{C}})} \nonumber\\ \qquad\qquad + {2\alpha\gamma V'(\Phi) \over \Phi'} {\mathcal{A}}' + 2\alpha\gamma\left[ \gamma^2{V'(\Phi)^2\over\Phi'^2} + \left( 3\frac{\alpha'}{\alpha} + 2\frac{r'}{r} \right) {V'(\Phi) \over \Phi'} + V''(\Phi) \right] {\mathcal{A}}&=& 0\,, \label{Eq:Pert4}\\ \frac{\gamma}{\alpha}\ddot{{\mathcal{C}}} - \frac{\partial}{\partial x}\left[ \frac{\alpha}{\gamma} {\mathcal{C}}' \right] - 2\frac{\alpha}{\gamma}\frac{r'}{r} {\mathcal{C}}' +2\frac{\alpha\gamma}{r^2}({\mathcal{A}}- {\mathcal{C}}) + 2\alpha\gamma \left[ \frac{r'}{r} {V'(\Phi)\over \Phi'} - \kappa V(\Phi) \right] {\mathcal{A}}&=& 0\,, \label{Eq:Pert5}\end{aligned}$$ which is still subject to the constraint given in Eq. (\[Eq:Pert3a\]). As a simple example, let us consider the reflection-symmetric subcase of the Ellis-Bronnikov wormhole, already mentioned at the end of the subsection \[Subsection:Ellis\], corresponding to the choice $V = 0$, $\alpha=\gamma=1$ and $r$ as in Eq. (\[Eq:StaticSolutions\]). In this case, the difference between Eqs. (\[Eq:Pert4\],\[Eq:Pert5\]), along with Eq. (\[Eq:Pert3a\]), gives $$\label{Eq:MasterEllis} \ddot{\chi} - \chi'' - \frac{3b^2}{r^4}\chi = 0\,,\qquad \chi := \frac{{\mathcal{A}}- {\mathcal{C}}}{r}\,,$$ which coincides with Eq. (15) in Ref. [@jGfGoS08]. We will return to this subcase at the end of section \[Sec:GaugeInvariant\]. The generalizations to the non reflection-symmetric Ellis-Bronnikov wormhole and to the AdS wormhole (subsections \[Subsection:Ellis\], \[Subsection:AdS\]) will be discussed in section \[Sec:Decoupling\]. Gauge-invariant reinterpretation {#Sec:GaugeInvariant} ================================ In this section we analyze the behavior of the perturbed fields under an infinitesimal coordinate transformation $$x^a \mapsto x^a + \delta x^a\,,\qquad (x^a) = (t,x)\,, \label{Eq:InfCT}$$ parametrized by a vector field $\mathbf{\delta x} = \delta x^a \partial_a = (\delta t) \partial_t + (\delta x) \partial_x$ on $\tilde{M}$, and try to rewrite the equations from the previous section in terms of fields which are manifestly gauge-invariant with respect to these transformations. Under the transformation (\[Eq:InfCT\]) the linear perturbations of the radial part of the metric, $\tilde{g}_{ab} dx^a dx^b := -\alpha^2 dt^2 + \gamma^2(dx + \beta dt)^2$, of the areal radius $r$, and of scalar field $\Phi$ transform according to $$\delta \tilde{g}_{ab} \mapsto \delta \tilde{g}_{ab} + \pounds_{\mathbf{\delta x}}\tilde{g}_{ab}\,,\qquad \delta r \mapsto \delta r + \pounds_{\mathbf{\delta x}} r\,,\qquad \delta\Phi \mapsto \delta\Phi + \pounds_{\mathbf{\delta x}}\Phi\,,$$ with $\pounds_{\mathbf{\delta x}}$ denoting the Lie derivative with respect to $\mathbf{\delta x}$. In terms of the parametrization (\[Eq:SphericalMetric\]) and assuming that the background is static, this yields $$\begin{aligned} \delta\alpha &\mapsto& \delta\alpha + \alpha'\delta x + \alpha \, \delta \dot{t}\,,\\ \delta\beta &\mapsto& \delta\beta + \delta \dot{x} - \frac{\alpha^2}{\gamma^2} \delta t'\,,\\ \delta\gamma &\mapsto& \delta\gamma + \left( \gamma\delta x \right)'\,,\\ \delta r &\mapsto& \delta r + r'\delta x\,,\\ \delta\Phi &\mapsto& \delta\Phi + \Phi'\delta x\end{aligned}$$ (where $\delta \dot{t}$, $\delta \dot{x}$ and $\delta t'$ refer to $\frac{\partial}{\partial t}(\delta t)$, $\frac{\partial}{\partial t}(\delta x)$ and $\frac{\partial}{\partial x}(\delta t)$, respectively; similar notations are used hereafter in relation to $\delta \beta$ and $\delta \Phi$). The following three quantities are invariant with respect to these transformations: $$\begin{aligned} A &:=& \frac{\delta\gamma}{\gamma} - \frac{1}{\gamma}\left( \gamma\frac{\delta\Phi}{\Phi'} \right)'\,, \label{Eq:AGI}\\ C &:=& \frac{\delta r}{r} - \frac{r'}{r}\frac{\delta\Phi}{\Phi'}\,, \label{Eq:CGI}\\ E &:=& \left( \frac{\delta\alpha}{\alpha} \right)' - \left( \frac{\alpha'}{\alpha}\frac{\delta\Phi}{\Phi'} \right)' + \frac{\gamma^2}{\alpha^2} \left( \delta\dot{\beta} - \frac{\delta\ddot{\Phi}}{\Phi'} \right)\,. \label{Eq:EGI}\end{aligned}$$ In the particular gauge used in the second half of the previous section, for which $\delta\beta = \delta\Phi = 0$, it turns out that $A={\mathcal{A}}$, $C={\mathcal{C}}$ and $E={\mathcal{D}}'$, where ${\mathcal{A}}$, ${\mathcal{C}}$ and ${\mathcal{D}}$ are defined by Eq. (\[Eq:DAC\]). Therefore, in this gauge, we may replace the quantities ${\mathcal{A}}$, ${\mathcal{C}}$ and ${\mathcal{D}}'$ in Eqs. (\[Eq:Pert3a\]–\[Eq:Pert5\]) by the quantities $A,C,E$ of the present section. Since the linearized field equations are gauge-invariant, the equations obtained in this way are valid in *any gauge*.\ Summing up, our gauge-invariant equations are $$\begin{aligned} && \sigma=0,\qquad \sigma=\frac{\alpha r^2}{\gamma} \left[ C'+\left({r'\over r}-{\alpha'\over\alpha}\right)C-{r'\over r}A \right]\,, \label{Eq:Pert3abis}\\ && E - A' + 2C' + 2\gamma^2{V'(\Phi)\over \Phi'} A = 0\,, \label{Eq:Pert3bbis}\\ && \frac{\gamma}{\alpha}\ddot{A} - \frac{\partial}{\partial x}\left[ \frac{\alpha}{\gamma}(A' - 2C') \right] + 2\frac{\alpha}{\gamma}\left( \frac{\alpha'}{\alpha} + \frac{r'}{r} \right) C' {- 2\frac{\alpha\gamma}{r^2}(A - C)} \nonumber\\ && \qquad\qquad + {2\alpha\gamma V'(\Phi) \over \Phi'} A' + 2\alpha\gamma\left[ \gamma^2{V'(\Phi)^2\over\Phi'^2} + \left( 3\frac{\alpha'}{\alpha} + 2\frac{r'}{r} \right) {V'(\Phi) \over \Phi'} + V''(\Phi) \right] A = 0\,, \label{Eq:Pert4bis}\\ && \frac{\gamma}{\alpha}\ddot{C} - \frac{\partial}{\partial x}\left[ \frac{\alpha}{\gamma} C' \right] - 2\frac{\alpha}{\gamma}\frac{r'}{r} C' +2\frac{\alpha\gamma}{r^2}(A - C) + 2\alpha\gamma \left[ \frac{r'}{r} {V'(\Phi)\over \Phi'} - \kappa V(\Phi) \right] A = 0\,. \label{Eq:Pert5bis}\end{aligned}$$ As a simple example, consider again the reflection-symmetric Ellis-Bronnikov wormhole, for which $V = 0$, $\alpha=\gamma=1$ and $r = \sqrt{x^2 + b^2}$, i.e. the same example as the one described at the end of section \[Sec:LinearPerturbation\] with Eqs. (\[Eq:Pert3a\]-\[Eq:Pert5\]) yielding Eq. (\[Eq:MasterEllis\]). However, now Eq. (\[Eq:MasterEllis\]) can be reinterpreted in a gauge-invariant framework where $\chi=(A-C)/r$. The interest of this equation is that it involves only one unknown function $\chi(t,x)$ and reduces the linear stability analysis of this wormhole to the spectral analysis of the Schrödinger operator $-d^2/dx^2-3b^2/(x^2 + b^2)^2$. Since this has one negative eigenvalue (see Refs. [@jGfGoS08; @jGfGoS09a]), one concludes that the wormhole is unstable. In Ref. [@jGfGoS08] an attempt was made to provide the present gauge-invariant formulation of the field equations in this particular subcase: while the two gauge-invariant quantities $A$ and $C$ were correctly defined, the quantity $D = \delta\alpha/\alpha$ defined in [@jGfGoS08] is only invariant with respect to the restricted set of gauge transformations for which $\delta \dot{t} = 0$. However, in general, this restricted set is not sufficient to achieve both conditions $\delta\Phi = 0$ and $\delta\beta = 0$ simultaneously, on which the derivation in Ref. [@jGfGoS08] was based. Finally, let us observe that the the gauge-invariant Eqs. (\[Eq:Pert3abis\],\[Eq:Pert4bis\],\[Eq:Pert5bis\]) (again in the present subcase $V = 0$, $\alpha=\gamma=1$ and $r = \sqrt{x^2 + b^2}$) are related to the results presented in [@fCfPlP19], which are based on the gauge $\delta\alpha = \delta\beta = 0$. For example, Eq. (\[Eq:Pert3abis\]), when choosing the gauge $\delta\alpha = \delta\beta = 0$, yields Eq. (3.9) in [@fCfPlP19]. The analysis of [@fCfPlP19] also yields a final equation of the form (\[Eq:MasterEllis\]), even though using a different approach related to the chosen gauge. ([^5] ) Decoupling of the pulsation equations {#Sec:Decoupling} ===================================== After considering the simple example of the reflection-symmetric Ellis-Bronnikov wormhole, let us return to the case of an arbitrary potential $V(\Phi)$. In this section we try to reduce the gauge-invariant equations of section \[Sec:GaugeInvariant\] to one involving only one unknown function $\chi(t,x)$ (generalizing the considerations which lead to Eq. (\[Eq:MasterEllis\]) in the reflection-symmetric Ellis-Bronnikov subcase). To this purpose we note the following: setting $${\cal F} := {A-C\over r}\,,\qquad{\cal G} := {C\over r} \label{Eq:FandG}$$ and performing a lengthy calculation, we can reformulate the system of Eqs. (\[Eq:Pert3abis\],\[Eq:Pert4bis\],\[Eq:Pert5bis\]) as the hyperbolic system of wave equations $$\left[ \frac{\partial^2}{\partial t^2} - \left( \frac{\alpha}{\gamma}\frac{\partial}{\partial x} \right)^2 + \left( \begin{array}{cc} Y_0 & Y_0 \\ 0 & 0 \end{array} \right)\frac{\alpha}{\gamma}\frac{\partial}{\partial x} + \frac{\alpha^2}{\gamma^2} \left( \begin{array}{cc} W_{11} & W_{12} \\ W_{21} & W_{22} \end{array} \right) \right] \left( \begin{array}{cc} {\cal F} \\ {\cal G} \end{array} \right) = 0\,, \label{Eq:WaveSystem}$$ subject to the constraint $${\cal G}' = \left( \frac{\alpha'}{\alpha} - \frac{r'}{r} \right){\cal G} + \frac{r'}{r}{\cal F}\, . \label{Eq:WaveConstraint}$$ Here, the functions $ Y_0$ and $W_{ij}$ are given by the following functions of the background quantities: $$\begin{aligned} Y_0 & =&2\alpha\gamma {V'(\Phi)\over{\Phi'}}\,, \label{Eq:Y0}\\ W_{11} &=& \frac{r'}{r}\left( 4\frac{\alpha'}{\alpha} + 3\frac{r'}{r} \right) - 3\frac{\gamma^2}{r^2} + Z_{11}\,,\\ W_{12} &=& 4\frac{\alpha'^2}{\alpha^2} +Z_{12}\,,\\ W_{21} &=& -4\frac{r'^2}{r^2} + 2\frac{\gamma^2}{r^2} + Z_{21}\,,\\ W_{22} &=& \frac{r'}{r}\left( -4\frac{\alpha'}{\alpha} + 3\frac{r'}{r} \right) - \frac{\gamma^2}{r^2} + Z_{22}\,,\end{aligned}$$ where $$\begin{aligned} Z_{11} &=& Z_{12} + \kappa\gamma^2V(\Phi)\,,\\ Z_{12} &=& 2\gamma^2\left[ \gamma^2{V'(\Phi)^2\over \Phi'^2} + \left( 3{\alpha'\over \alpha}+2{r'\over r} \right){V'(\Phi) \over \Phi'} + V''(\Phi) \right]\, ,\\ Z_{21} &=& 2\gamma^2\left[ -\kappa V(\Phi) + {r'\over r}{V'(\Phi)\over \Phi'} \right]\,,\\ Z_{22} &=& Z_{21} + \kappa\gamma^2 V(\Phi)\,.\end{aligned}$$ In deriving the wave system (\[Eq:WaveSystem\]) we have used the background equations $$\begin{aligned} \frac{r'}{r}\left( 2\frac{\alpha'}{\alpha} + \frac{r'}{r} \right) - \frac{\gamma^2}{r^2} + \frac{\kappa}{2}\Phi'^2 + \kappa\gamma^2 V(\Phi) &=& 0\,, \label{Eq:Background1}\\ \frac{\alpha''}{\alpha}- \frac{\alpha'}{\alpha}\left( \frac{\gamma'}{\gamma} - 2\frac{r'}{r} \right) + \kappa \gamma^2 V(\Phi) & = & 0\,, \label{Eq:Background2}\\ \frac{r''}{r} - \frac{r'}{r}\left( \frac{\gamma'}{\gamma} - \frac{\alpha'}{\alpha}- \frac{r'}{r} \right) - \frac{\gamma^2}{r^2} + \kappa \gamma^2 V(\Phi) & = & 0\,. \label{Eq:Background3}\end{aligned}$$ Observe that in the reflection-symmetric Ellis-Bronnikov subcase $V = 0$, $\alpha=\gamma=1$ and $r = \sqrt{x^2 + b^2}$ it follows that $Y_ 0 = W_{12} = 0$, such that the equation for ${\cal F}$ in the system (\[Eq:WaveSystem\]) decouples trivially from the remaining ones. In the following, we describe a general trick which allows one to decouple the wave system (\[Eq:WaveSystem\],\[Eq:WaveConstraint\]). Let us suppose that we know a static solution $({\cal F}_0(x),{\cal G}_0(x))$ of Eqs. (\[Eq:WaveSystem\],\[Eq:WaveConstraint\]) such that ${\cal G}_0(x)\neq 0$ for all $x$. In this case, based on Leibnitz’s product rule, it is not difficult to verify that the field $$\tilde\chi := {\cal F} - \frac{{\cal F}_0}{{\cal G}_0}{\cal G}$$ satisfies the decoupled wave equation $$\left[ \frac{\partial^2}{\partial t^2} - \left( \frac{\alpha}{\gamma}\frac{\partial}{\partial x} \right)^2 + Y_0\frac{\alpha}{\gamma} \frac{\partial}{\partial x} + \frac{\alpha^2}{\gamma^2}\tilde{\cal {V}} \right] \tilde\chi = 0\,, \label{Eq:Master1}$$ with the potential $$\tilde{\mathcal{ V}} = W_{11} - \frac{{\cal F}_0}{{\cal G}_0} W_{21} - 2\frac{r'}{r}\left( \frac{{\cal F}_0}{{\cal G}_0} \right)' +\frac{\gamma}{\alpha}\frac{r'}{r} Y_0\left( \frac{{\cal F}_0}{{\cal G}_0}+1 \right)\,.$$ Let us observe that it is possible to eliminate the first spatial derivative in Eq. (\[Eq:Master1\]). Indeed, let us define $$\label{Eq:ChiTrans} \chi:={\tilde\chi\over a}\,,\qquad a(x) = a_0 e^{\int_{x_0}^{x}\frac{Y_0(y)\gamma(y)}{2\alpha(y)}dy }\,,$$ where $a_0$ and $x_0$ are two constants. ([^6]) Then, it is found that $\chi$ satisfies the wave equation $$\label{Eq:Master2} \left[ \frac{\partial^2}{\partial t^2} - \left( \frac{\alpha}{\gamma}\frac{\partial}{\partial x} \right)^2 + \frac{\alpha^2}{\gamma^2}{\cal {V}} \right] \chi = 0\,,$$ with the potential ([^7]) $$\begin{aligned} {\cal{V}} &=& \tilde{\mathcal{ V}} + \frac{1}{4}\frac{\gamma^2}{\alpha^2} Y_0^2 - \frac{1}{2}\frac{\gamma}{\alpha} Y_0'\nonumber\\ &=& \frac{r'}{r}\left( 4\frac{\alpha'}{\alpha} + 3\frac{r'}{r} \right) - \frac{3\gamma^2}{r^2} + \gamma^2\left[ 2\gamma^2{V'(\Phi)^2\over \Phi'^2} + 4\left( \frac{\alpha'}{\alpha} + \frac{r'}{r} \right) {V'(\Phi) \over \Phi'} + V''(\Phi) + \kappa V(\Phi) \right] \nonumber\\ &+& \left[ 4\frac{r'^2}{r^2} - \frac{2\gamma^2}{r^2} + 2\kappa\gamma^2 V(\Phi) \right] \frac{{\cal F}_0}{{\cal G}_0} - 2\frac{r'}{r}\left( \frac{{\cal F}_0}{{\cal G}_0} \right)'\,. \label{Eq:Potential}\end{aligned}$$ We refer to Eq. (\[Eq:Master2\]) as the *master equation*; this reduces the linear stability analysis to the spectral analysis of the linear, Schrödinger-type operator $-\left( \frac{\alpha}{\gamma}\frac{d}{d x} \right)^2 + \frac{\alpha^2}{\gamma^2}\,{\cal {V}}$. Once the master equation has been solved for the field $\chi(t,x)$, it is possible to reconstruct the gauge-invariant quantities ${\cal F}$ and ${\cal G}$ by integrating the constraint equation (\[Eq:WaveConstraint\]). Using the definition of $\chi$ and the fact that $({\cal F}_0,{\cal G}_0)$ satisfy the constraint, one obtains $$\begin{aligned} {\cal G}(t,x) &=& {\cal G}_0(x)\int\limits_{x_0}^x \frac{r'(y)}{r(y)}\frac{a(y)}{{\cal G}_0(y)} \chi(t,y) dy\,,\\ {\cal F}(t,x) &=& a(x)\chi(t,x) + \frac{{\cal F}_0(x)}{{\cal G}_0(x)} {\cal G}(t,x)\,,\end{aligned}$$ and from this one can also reconstruct the gauge-invariant fields $A$ and $C$. Finally, the gauge-invariant field $E$ is obtained from Eq. (\[Eq:Pert3bbis\]). Let us repeat that the above approach requires the knowledge of a static solution $({\cal F}_0(x),{\cal G}_0(x))$ of Eqs. (\[Eq:WaveSystem\],\[Eq:WaveConstraint\]). A general strategy to obtain such a static solution is to make an infinitesimal variation along a static solution $(\alpha,\gamma,r,\Phi)$ of the Einstein-scalar equations with respect to its parameters. Since this linearization of the static solution automatically satisfies the linearized system (\[Eq:Ev1\]–\[Eq:Evm\]), the corresponding gauge-invariant fields $A$ and $C$ (defined by Eqs. (\[Eq:AGI\],\[Eq:CGI\])) fulfil the system (\[Eq:Pert4bis\],\[Eq:Pert5bis\]), provided that we have the vanishing condition (\[Eq:Pert3abis\]) for $\sigma$; obviously, under the same condition, the fields $\cal F$ and $\cal G$ associated with $A$ and $C$ represent a static solution of the wave system (\[Eq:WaveSystem\],\[Eq:WaveConstraint\]). In the following we will apply this general strategy to the cases of the Ellis-Bronnikov and AdS wormholes. Perturbed Ellis-Bronnikov wormhole {#SubSec:PertEllis} ---------------------------------- The Ellis-Bronnikov solution given in Eq. (\[Eq:StaticSolutions\]) can be linearized with respect to the parameters $b$ and $\gamma_1$, which yields $$\begin{aligned} \frac{\delta\gamma}{\gamma} &=& - \left(\arctan\frac{x}{b} \right)\delta\gamma_1 + \frac{\gamma_1 x}{x^2 + b^2}\delta b\,,\\ \frac{\delta r}{r} &=& -\left(\arctan \frac{x}{b} \right)\delta\gamma_1 + \frac{b + \gamma_1 x}{x^2 + b^2}\delta b\,,\\ \frac{\delta\Phi}{\Phi'} &=& \frac{\gamma_1}{1 + \gamma_1^2}(x^2 + b^2) \left(\arctan \frac{x}{b} \right)\frac{\delta\gamma_1}{b} - x\frac{\delta b}{b}\,.\end{aligned}$$ Introduced into Eqs. (\[Eq:AGI\],\[Eq:CGI\]) this gives rise to the gauge-invariant quantities $$\begin{aligned} A &=& -\frac{1}{1 + \gamma_1^2}\left[ \gamma_1 + \left( 1 + 2\gamma_1\frac{x}{b} \right) \arctan \frac{x}{b} \right] \delta\gamma_1 + \frac{\delta b}{b}\,, \label{Eq:AEllis}\\ C &=& -\frac{1 + \gamma_1\frac{x}{b}}{1 + \gamma_1^2}\left(\arctan \frac{x}{b} \right) \delta\gamma_1 + \frac{\delta b}{b}\, . \label{Eq:CEllis}\end{aligned}$$ As explained before, the fields $A$ and $C$ automatically satisfy the system of equations (\[Eq:Pert3abis\],\[Eq:Pert4bis\],\[Eq:Pert5bis\]). In this case the definition of $\sigma$ in Eq. (\[Eq:Pert3abis\]) gives $\sigma=-\delta b\gamma_1-b \delta \gamma_1=\delta(b\gamma_1)$, so that the condition $\sigma=0$ therein holds if $$\delta b=-{b \delta\gamma_1\over \gamma_1}\,. \label{Eq:SigmaEllis}$$ Inserting Eqs. (\[Eq:AEllis\],\[Eq:CEllis\],\[Eq:SigmaEllis\]) in the definition (\[Eq:FandG\]) of $\cal F$ and $\cal G$ (and omitting the proportionality factor $\delta\gamma_1$), one obtains the following time-independent solution of the constrained wave system (\[Eq:WaveSystem\],\[Eq:WaveConstraint\]): $$\left( \begin{array}{cc} {\cal F}_0 \\ {\cal G}_0 \end{array} \right) := \frac{1}{r}\left( \begin{array}{c} \frac{\gamma_1^2}{1 + \gamma_1^2}\left[ 1 + \frac{x}{b} \arctan\frac{x}{b} \right] \\ F(x) \end{array} \right)\,,\qquad F(x) := 1 + \frac{\gamma_1}{1 + \gamma_1^2}\left( 1 + \gamma_1\frac{x}{b} \right) \arctan \frac{x}{b}\,. \label{Eq:StaticSolutionWaveSystem}$$ Note that the function $F: {\mathbb{R}}\to {\mathbb{R}}$ is smooth and strictly positive. ([^8]) Based on these observations, we can apply the general method for decoupling the wave system (\[Eq:WaveSystem\],\[Eq:WaveConstraint\]), choosing the static solution $({\cal{F}}_0(x),{\cal{G}}_0(x))$ as in Eq. (\[Eq:StaticSolutionWaveSystem\]). Note that in this case we have $Y_0 = 0$ and can choose $a = 1$ as $V = 0$, which implies that $\mathcal{ V}=\tilde{\mathcal{ V}}$ and $\chi = \tilde{\chi}$. One can verify that the function $\frac{\alpha^2}{\gamma^2} \mathcal{ V}$ appearing in the master equation (\[Eq:Master2\]) agrees up to a rescaling with the potential defined in Eq. (32) of Ref. [@jGfGoS09a], denoted therein with $W$; more precisely, $$\label{Eq:W} \left(\frac{\alpha^2}{\gamma^2} \mathcal{V}\right)(x) =~{1 \over b^2} W\!\left(\frac{x}{b}\right)\,.$$ For future mention let us point out some features of this function, following from the analysis of $W$ in [@jGfGoS09a]. First of all $\frac{\alpha^2}{\gamma^2} \mathcal{V} : {\mathbb{R}}\to {\mathbb{R}}$ is a $C^\infty$ bounded function; moreover, if $\gamma_1 \neq 0$ one has $\left(\frac{\alpha^2}{\gamma^2} \mathcal{ V}\right)(x) \sim 2 e^{\pm 2 \pi \gamma_1}/x^2$ for $x \mapsto \pm \infty$. In the reflection-symmetric case $\gamma_1=0$, where $\alpha=\gamma=1$, one obtains $$\mathcal{V}(x) = - \frac{3 b^2}{(x^2 + b^2)^2} = - \frac{3 b^2}{r^4(x)}\,, \label{Eq:Vsim}$$ and the master equation (\[Eq:Master2\]) is found to coincide with Eq. (\[Eq:MasterEllis\]). Let us now sketch some spectral features of the Schrödinger operator $-\left({\alpha \over\gamma}{d\over dx}\right)^2 + {\alpha^2\over\gamma^2}\, \mathcal{V}$ appearing in the master equation (\[Eq:Master2\]) (which can be regarded as a selfadjoint operator in $L^2(\mathbb{R},{\gamma \over\alpha} dx)$); these features allow one to infer the linear instability of the Ellis-Bronnikov wormhole both for $\gamma_1\neq 0$ and for $\gamma_1=0$. As shown in [@jGfGoS09a], the zero energy" equation $\left[-\left({\alpha\over\gamma}{d\over dx}\right)^2 + {\alpha^2\over\gamma^2}\, \mathcal{V} \right]\chi_0 = 0$ has a solution $$\chi_0(x) = \frac{x - b\gamma_1}{r(x) F(x)}\,, \label{Eq:EBZeroMode}$$ which has precisely one zero in the interval $(-\infty,\infty)$. According to the Sturm comparison theorems (see for instance [@Weid-Book], [@bS05] and references therein) it follows that for each $\gamma_1$ including $\gamma_1 = 0$, the Schrödinger operator in the master equation possesses a single bound state with negative energy. Note that for $\gamma_1 \neq 0$ the function $\chi_0$ decays as $1/|x|$ for large $|x|$, so that it describes a bound state with zero energy, while for $\gamma_1 = 0$ it reduces to $\chi_0(x) = x/\sqrt{1 + x^2}$ which is not normalizable but still has a single zero. The existence of a single bound state with negative energy implies that the master equation (\[Eq:Master2\]) for each Ellis-Bronnikov wormhole possesses a unique mode diverging exponentially in time, a fact of course sufficient to infer the instability of the wormhole. In the next section we will give more details on the spectral properties of the Schrödinger operator and on the solution of the master equation (\[Eq:Master2\]) within a rigorous functional setting, also allowing comparison with the corresponding problem for the perturbed AdS wormhole. Perturbed AdS wormhole {#SubSec:Pertads} ---------------------- Next, we analyze the AdS wormhole in the coordinate system $(s,u)$, as described by Eq. (\[Eq:StaticSolutionsPot2\]) for arbitrary parameters $k,B > 0$, and apply the general framework presented in this section with $(s,u)$ in place of $(t,x)$. Although the static solution formally depends on two parameters $B$ and $k$, it is important to note that $k$ also appears in the potential function $V(\Phi)$ (see Eq. (\[Eq:StaticSolutionsPot1\])). However, since we regard the potential to be fixed in our perturbation analysis, we will exclude the possibility of varying $k$. In contrast to $k$, the parameter $B$ is free, and variation of the solution (\[Eq:StaticSolutionsPot2\]) with respect to it gives $$\begin{aligned} \label{Eq:Variations1} \delta \alpha &=&\delta \gamma=0\,,\\ \label{Eq:Variations2} {\frac{\delta r}{r}}&=& \frac{2B\delta B}{1+2B^2-\cos u}\,,\\ \label{Eq:Variations3} {\frac{\delta \Phi}{\Phi'}}&=&-\sin u\frac{ \delta B}{B(1+B^2)}\,.\end{aligned}$$ Eqs. (\[Eq:Variations1\],\[Eq:Variations2\],\[Eq:Variations3\]), introduced into Eqs. (\[Eq:AGI\],\[Eq:CGI\]), yields the following expressions for the gauge-invariant quantities $A$ and $C$: $$A = \frac{1+\cos u}{2 B(1+ B^2)} \delta B\,,\qquad C = \frac{{\delta B}}{B}\,. \label{Eq:ABAdS}$$ From here and from the definition of $\sigma$ in Eq. (\[Eq:Pert3abis\]) we see that $\sigma=0$, as required, for every choice of the perturbation $\delta B$. Inserting Eq. (\[Eq:ABAdS\]) in the definition (\[Eq:FandG\]) of $\cal F$ and $\cal G$ (and omitting the proportionality factor $\delta B$) one obtains, also in this case, a static solution of the system (\[Eq:WaveSystem\],\[Eq:WaveConstraint\]): $$\left( \begin{array}{cc} {\cal F}_0 \\ {\cal G}_0 \end{array} \right) := \frac{\sqrt{2}k}{B} \cos\left( \frac{u}{2} \right)\times \left( \begin{array}{c} -\frac{\sqrt{1+2B^2-{\cos u}}}{2(1+B^2)} \\ \frac{1}{\sqrt{1+2B^2-{\cos u}}} \end{array} \right)\,. \label{Eq:SolStat2}$$ Note that ${\cal G}_0$ is a strictly positive function of $u\in(-\pi,\pi)$, and that ${\cal F}_0/{\cal G}_0 = -(1 + 2B^2 - \cos u)/(2(1 + B^2))$. Having found the non-trivial solution, we can now obtain the master equation governing the spherical symmetric linearized perturbations of the AdS wormhole, following the general method explained before. We observe that $Y_0 = -2\tan\frac{u}{2}$ and that we can choose the constants $a_0$ and $x_0$ in Eq. (\[Eq:ChiTrans\]) such that $a = 1/\alpha^2$; therefore Eq. (\[Eq:ChiTrans\]) reads $$\chi := \left({\cal F} - \frac{{\cal F}_0}{{\cal G}_0}{\cal G}\right)\alpha^2$$ and the master equation (\[Eq:Master2\]) becomes (recalling that $\alpha/\gamma=1$ for the AdS wormhole in coordinates $(s,u)$) $$\left[ \frac{\partial^2}{\partial s^2} - \frac{\partial^2}{\partial u^2} + {\cal {V}} \right] \chi = 0\,, \label{Eq:MasterAdS}$$ with the potential $${\mathcal{V}}(u) \equiv {\mathcal{V}}_B(u) = -\frac{B^2 \left(2+B^2+\cos u\right) }{\left(1+2 B^2-\cos u\right)^2}\,. \label{Eq:PotentialB}$$ For the following, we assume Dirichlet boundary conditions at the two asymptotic AdS ends, that is, $$\chi(s,\pm \pi) = 0\,. \label{Eq:BounCond}$$ Since $$\chi = \frac{1}{\sqrt{2}\sqrt{1+2 B^2-\cos u}}\delta\gamma - \frac{1 + \cos u}{4(1 + B^2)(1 + 2B^2-\cos u)} \delta r$$ for $\delta \Phi=0$, a sufficient condition for Eq. (\[Eq:BounCond\]) to hold is that, in the gauge $\delta \Phi=0$, the perturbed functions $\delta r$ and $\delta \gamma$ vanish at the far ends $u=\pm\pi$ of the wormhole. For general considerations on boundary conditions for field theories on AdS spaces, see [@sAcIdS78][@cK13][@cDhFaM18]. Now (following the same scheme of the previous subsection) let us sketch some spectral features of the Schrödinger operator $-d^2/u^2+\mathcal{V}(u)$ with Dirichlet boundary conditions at $u=\pm \pi$ (to be regarded as a selfadjoint operator in $L^2((-\pi,\pi),du)$); these facts will allow us to infer the linear instability of the AdS wormhole. The zero energy Schrödinger equation $[-d^2/du^2 + \mathcal{V} ]\chi_0 = 0$ admits for each fixed $B > 0$ the general solution $$\chi_0(u) = C_1 {\sin {u \over 2} \over \sqrt{1+2B^2 -\cos u}} + C_2 {-2 u \sin{u \over 2} + 4 B^2 \cos{u \over 2}\over \sqrt{1+2B^2 -\cos u}}\,,\qquad -\pi < u < \pi\,, \label{Eq:AdSZeroMode}$$ with constants $C_1,C_2$. The Dirichlet boundary conditions $\chi_0(\pm\pi)=0$ are satisfied only in the trivial case $C_1=C_2=0$, which shows that none of these solutions is an eigenfunction of our Schrödinger operator. For $C_1 = -2\pi C_2\neq 0$ the zero energy solution satisfies the left boundary condition, i.e. $\chi_0(-\pi) = 0$, and since this solution has precisely one zero in the interval $(-\pi,\pi)$, ([^9]) it follows from the Sturm comparison theorem (e.g. Theorem 3.4 in [@bS05]) that our Schrödinger operator (with Dirichlet boundary conditions) has a single bound state with negative energy $E < 0$. This state gives rise to an exponentially in time growing mode solution of the master equation (\[Eq:MasterAdS\]) which is proportional to $e^{\sqrt{-E} t}$; this establishes the linear instability of the AdS wormhole. In the next section we analyze the spectral properties of the Schrödinger operator and their implications for the solutions of the master equation in a rigorous functional setting. Spectral representation of the solutions of the master equations {#Sec:Spectral} ================================================================ In this section we provide a rigorous analysis regarding the spectral properties of the Schrödinger-type operators involved in the master equations discussed so far. The next subsection and the related Appendix \[appebro\] concern the reflection-symmetric Ellis-Bronnikov case; the subsequent subsection sketches a similar analysis for the non-symmetric case. The third subsection and the related Appendix \[appeads\] treat the corresponding problem for the AdS case (not previously considered in the literature, to the best of our knowledge); in the same subsection we establish bounds on the eigenvalues of the Schrödinger operator. Brief comments regarding the time scale associated with the instability are made in the final subsection. Spectral decomposition of the master equation and instability of the Ellis-Bronnikov wormhole in the reflection-symmetric case {#Sec:SpectralA} ------------------------------------------------------------------------------------------------------------------------------ Let us consider the reflection-symmetric Ellis-Bronnikov wormhole and the corresponding master equation (\[Eq:MasterEllis\]), containing the potential $\mathcal{V}(x) = - 3 b^2/(x^2 + b^2)^2$; this equation can be written as $$\ddot{\chi}(t) + {H}\chi(t) = 0 \quad (t\in {\mathbb{R}}), \label{masterbro}$$ where $\chi(t)$ stands for the function ${\mathbb{R}}\ni x \mapsto \chi(t,x)$, and ${H}$ indicates the operator $- d^2/d x^2 + \mathcal{V}$. If we want a rigorous functional setting for Eq. (\[masterbro\]), we are led to consider the Hilbert space ([^10]) $$\label{hilbro} {\mathfrak{H}}:= L^2({\mathbb{R}}, d x)$$ made of the functions $f: {\mathbb{R}}\to {\mathbb{C}}$, $x \mapsto f(x)$ which are square integrable for the Lebesgue measure $d x$; we will write $\langle ~|~\rangle$ and $\|~\|$ for the natural inner product and norm of this space, defined by $\langle f | \ell \rangle := \int_{{\mathbb{R}}} d x \bar{f}(x) \ell(x)$ and $\| f \|^2 = \langle f | f \rangle$ for $f, \ell \in {\mathfrak{H}}$. ${H}$ can be regarded as a selfadjoint operator in ${\mathfrak{H}}$, if we give for it the precise definition $$\label{hbro} {H}:= -\frac{d^2}{d x^2} + \mathcal{V} : \mathfrak{D} \subset {\mathfrak{H}}\to {\mathfrak{H}}\,, \quad \mathfrak{D} := \{ f \in {\mathfrak{H}}~|~f_{x x} \in {\mathfrak{H}}\}$$ intending all $x$-derivatives in the distributional sense ([[^11]]{}). Due to general facts on Schrödinger operators [@Ber], and to a specific analysis performed in [@jGfGoS09a] for the potential $\mathcal{V}(x) = - 3 b^2/(x^2 + b^2)^2$, we can state that the spectrum of ${H}$ is the union of: - the point spectrum, which consists of a unique, simple eigenvalue $\mu_1 < 0$;\ - the continuous spectrum $[0,+\infty)$. One can construct a generalized orthonormal basis of the Hilbert space ${\mathfrak{H}}$, in the sense explained by Appendix \[appebro\] and by [@Ber], using: - a normalized eigenfunction $e_1$ for the eigenvalue $\mu_1$ ($e_1 \in \mathfrak{D}$, ${H}e_1 = \mu_1 e_1$, $\| e_1 \|=1$; $e_1$ is proved to be $C^\infty$); - two suitably chosen “improper eigenfunctions” $e_{i \lambda}$ ($i=1,2$) for each $\lambda \in (0,+\infty)$ (i.e., for each nonzero point $\lambda$ of the continuous spectrum); these are two linearly independent $C^\infty$ functions on ${\mathbb{R}}$ which fulfill $- d^2 e_{i \lambda}/ d x^2 + \mathcal{V} e_{i \lambda} = \lambda e_{i \lambda}$ but do not belong to ${\mathfrak{H}}$. Then, one can search for the solution ${\mathbb{R}}\ni t \to \chi(t)$ of Eq. (\[masterbro\]) with appropriate smoothness properties and with the initial conditions $$\chi(0) = q\,, \quad \dot{\chi}(0) = p\,, \label{cond}$$ where $q: x \mapsto q(x)$ and $p : x \mapsto p(x)$ are sufficiently regular functions. For all technical details, we refer again to Appendix \[appebro\]; here we introduce the selfadjoint operator $| {H}|^{1/2}: \mathfrak{D}^{1/2} \subset {\mathfrak{H}}\to {\mathfrak{H}}$ and indicate how to regard the domains $\mathfrak{D}^{1/2}$ and $\mathfrak{D}$ as Hilbert spaces with their own inner products. One can show that, for any $q \in \mathfrak{D}$ and $p \in \mathfrak{D}^{1/2}$, Eqs. (\[masterbro\],\[cond\]) have a unique solution $t \mapsto \chi(t)$ in $C({\mathbb{R}}, \mathfrak{D}) \cap C^1({\mathbb{R}}, \mathfrak{D}^{1/2}) \cap C^2({\mathbb{R}}, {\mathfrak{H}})$, which is as follows for all $t \in {\mathbb{R}}$: $$\chi(t) = \left[ \langle e_1 | q \rangle \cosh( |\mu_1|^{1/2} t) + \langle e_1 | p \rangle \frac{\sinh( |\mu_1|^{1/2} t)}{|\mu_1|^{1/2}} \right] e_1 + \sum_{i=1}^{2} \int_{0}^{+\infty} \hspace{-0.4cm} d \lambda \left[ \langle e_{i \lambda} | q \rangle \cos( \lambda^{1/2} t) + \langle e_{i \lambda} | p \rangle \frac{\sin( \lambda^{1/2} t)}{\lambda^{1/2}} \right] e_{i \lambda}\, . \label{solbro}$$ As explained in Appendix \[appebro\], the symbols $\langle \cdot | \cdot \rangle$ in the above formula indicate usual inner products in ${\mathfrak{H}}$, or suitably defined generalizations; the integrals over $\lambda$ are understood in a weak sense. Of course we are interested in the case where $\chi(t)$ is *real valued* for each $t$, which occurs if and only if the data $q, p$ are real valued functions. The coefficient of $e_1$ in Eq. (\[solbro\]) diverges exponentially both for $t \to -\infty$ and for $t \to + \infty$ (except for very special choices of $\langle e_1 | q \rangle$ and $\langle e_1 | p \rangle$ ([^12])); this suffices to infer the (linear) instability of the reflection-symmetric Ellis-Bronnikov wormhole [@jGfGoS08] [@jGfGoS09a] [@fCfPlP19]. In addition, let us remark that the integrals over $\lambda$ in Eq. (\[solbro\]) are superpositions of “non normalizable” oscillatory modes, living outside the space ${\mathfrak{H}}= L^2({\mathbb{R}}, d x)$ like the improper eigenfunctions $e_{i \lambda}$. Spectral decomposition of the master equation and instability of the Ellis-Bronnikov wormhole in the non-symmetric case {#Sec:SpectralB} ----------------------------------------------------------------------------------------------------------------------- Let us now pass to the non reflection-symmetric Ellis-Bronnikov wormhole (as in Eq. (\[Eq:StaticSolutions\]) with $\gamma_1 \neq 0$). In this case the master equation for $\chi(t,x)$ has the form (\[Eq:Master2\]), involving the operator $$- \left(\frac{\alpha}{\gamma} \frac{\partial}{\partial x}\right)^2 + \frac{\alpha^2}{\gamma^2} \mathcal{V}\,, \quad \frac{\alpha(x)}{\gamma(x)} = e^{2 \gamma_1 \arctan \frac{x}{b}}~~(x \in {\mathbb{R}})\,.$$ As noted in [@jGfGoS09a], the spectral analysis of this case can be simplified introducing the new coordinate $$\label{Eq:ro} \rho = \rho(x) := \int_{0}^x \frac{\gamma(y)}{\alpha(y)}\,;$$ note that the mapping $x \mapsto \rho(x)$ is a diffeomorphism of ${\mathbb{R}}$ to itself, and $\rho(x) \sim e^{\mp \pi \gamma_1} x$ for $x \to \pm \infty$. By construction $\frac{\alpha}{\gamma} \frac{\partial}{\partial x}$ $= \frac{\partial}{\partial \rho}$; so, writing $\chi(t,\rho)$ as an abbreviation for $\chi(t,x(\rho))$ we can rephrase the master equation (\[Eq:Master2\]) as $$\label{Eq:Master2reph} \left[ \frac{\partial^2}{\partial t^2} - \left( \frac{\partial}{\partial \rho} \right)^2 + \mathcal{U}(\rho) \right] \chi(t, \rho) = 0\,, \quad \mathcal{U}(\rho) := \left(\frac{\alpha^2}{\gamma^2} \mathcal{V}\right)(x(\rho)) \quad (t,\rho \in {\mathbb{R}}).$$ The function $\mathcal{U} : {\mathbb{R}}\to {\mathbb{R}}$ is $C^\infty$; due to the $x \to \pm \infty$ asymptotics of $\rho(x)$ (see after Eq. (\[Eq:ro\])) and $\left(\frac{\alpha^2}{\gamma^2} \mathcal{V}\right)(x)$ (see after Eq. (\[Eq:W\])), we have $\mathcal{U}(\rho) \sim 2/\rho^2$ for $\rho \to \pm \infty$. A precise functional setting for Eq. (\[Eq:Master2reph\]) can be obtained introducing the Hilbert space and the selfadjoint operator $$\label{hbrons} {\mathfrak{H}}:= L^2({\mathbb{R}}, d \rho)\,; \qquad {H}:= -\frac{d^2}{d \rho^2} + \mathcal{U} : \mathfrak{D} \subset {\mathfrak{H}}\to {\mathfrak{H}}\,, \quad \mathfrak{D} := \{ f \in {\mathfrak{H}}~|~f_{\rho \rho} \in {\mathfrak{H}}\}$$ (the $\rho$-derivatives are meant distributionally); $\langle~|~\rangle$ and $\|~\|$ indicate in the sequel the natural inner product and norm of ${\mathfrak{H}}$. ([^13]) After giving these prescriptions we write Eq. (\[Eq:Master2reph\]) in the form (\[masterbro\]), where $\chi(t)$ stands for the function $\rho \mapsto \chi(t,\rho)$; obviously enough, the treatment of this equation is reduced to a spectral analysis of the Schrödinger operator ${H}$ in Eq. (\[hbrons\]), which is rather similar to the discussion of the operator (\[hbro\]) for the reflection-symmetric wormhole. The main difference with respect to the symmetric case is that the operator ${H}$ in Eq. (\[hbrons\]) has a point spectrum consisting of *two* simple eigenvalues $\mu_1 < 0$ and $\mu_2 := 0$, see the comments below Eq. (\[Eq:EBZeroMode\]); the continuous spectrum is $(0,+\infty)$. Due to these facts there is a generalized orthonormal basis made of normalized eigenfunctions $e_1, e_2$ for the eigenvalues $\mu_1 < 0$ and $\mu_2 = 0$ ($e_1, e_2 \in \mathfrak{D}({H})$, ${H}e_1 = \mu_1 e_1$, ${H}e_2 = 0$, $\| e_1 \| = \| e_2 \| = 1$), plus two improper eigenfunctions $e_{i \lambda}$ ($i=1,2$) for each $\lambda$ in the continuous spectrum. As in the symmetric case, one can define Hilbert space structures for the domains $\mathfrak{D}$, $\mathfrak{D}^{1/2}$ of the operators ${H}$, $|{H}|^{1/2}$. For $q \in \mathfrak{D}$ and $p \in \mathfrak{D}^{1/2}$, the master equation  (\[masterbro\]) with initial conditions (\[cond\]) is proved again to possess a unique solution $t \mapsto \chi(t)$ in $C({\mathbb{R}}, \mathfrak{D}) \cap C^1({\mathbb{R}}, \mathfrak{D}^{1/2}) \cap C^2({\mathbb{R}}, {\mathfrak{H}})$; this has a representation similar to (\[solbro\]) with an additional term associated with the eigenvalue zero, namely: $$\chi(t) = \mbox{r.h.s. of Eq.~(\ref{solbro})} + \Big[ \langle e_2 | q \rangle + \langle e_2 | p \rangle t \Big] e_2\,. \label{solbrons}$$ So, besides the exponentially divergent term proportional to $e_1$, the expression of $\chi(t)$ contains a term diverging linearly for $t \to \pm \infty$ (if $\langle e_2 | p \rangle \neq 0$); in any case the wormhole is linearly unstable. Let us note that, as in Eq. (\[solbro\]), the present expression for $\chi(t)$ contains an integral over $\lambda$ of non normalizable oscillatory modes, proportional to the improper eigenfunctions $e_{i \lambda}$ which live outside ${\mathfrak{H}}$. Spectral decomposition of the master equation and instability of the AdS wormhole --------------------------------------------------------------------------------- In the AdS case we introduce the Hilbert space $$\label{hilbu} {\mathfrak{H}}:= L^2((-\pi,\pi), du)$$ formed by the functions $f : (-\pi,\pi) \to {\mathbb{C}}$, $u \mapsto f(u)$ which are square integrable with respect to the Lebesgue measure $d u$; from now on we denote by $\Braket{~|~}$ and $\|~\|$ the natural inner product and norm of ${\mathfrak{H}}$, so that $\langle f | \ell \rangle := \int_{-\pi,\pi} d u \bar{f}(u) \ell(u)$ and $\| f \|^2 = \langle f | f \rangle$ for $f, \ell \in {\mathfrak{H}}$. In addition, let us consider the potential ${\mathcal{ V} }$ appearing in Eq. (\[Eq:PotentialB\]) (a $C^\infty$ function on $[-\pi, \pi]$). A rigorous setting for the master equation (\[Eq:MasterAdS\]) with boundary conditions (\[Eq:BounCond\]) can be set up using the space (\[hilbu\]) and the selfadjoint operator $$\label{hB} {H}:= -{d^2 \over du^2}+ \mathcal{ V} : \mathfrak{D} \subset {\mathfrak{H}}\to {\mathfrak{H}}\,, \quad \mathfrak{D} := \{ f \in {\mathfrak{H}}~|~f_{u u} \in {\mathfrak{H}}~, f(\pm \pi) = 0 \}\,.$$ Here and in the sequel, the $u$-derivatives like $f_{u u }$ are understood distributionally; a function $f\in {\mathfrak{H}}$ with $f_{uu}\in {\mathfrak{H}}$ is in fact in $C^1([-\pi,\pi])$, so it can be evaluated at $u=\pm\pi$ ([^14]). As an operator in the Hilbert space ${\mathfrak{H}}$, ${H}$ has the following properties: - it is selfadjoint; - it is bounded from below; - it has a purely discrete spectrum. As known in general for Hilbert space operators fulfilling (i-iii), it is possible to represent the eigenvalues of ${H}$ as an increasing sequence $\mu_1 <\mu_2 < \ldots$ In addition, ${H}$ has the following properties: - any of its eigenfunctions is in the space $C^{\infty}([-\pi,\pi])$; - each one of its eigenvalues is simple. For future mention, let us recall that the operator ${H}^0:=-d^2/du^2$ with domain $\mathfrak{D}$ as above also has the properties (i-v); in this case the eigenvalues are $\mu^{0}_n :=n^2/4$, with normalized eigenfunctions $f^0_n(u):=(1/\sqrt{\pi})\sin[(n/2)(u+\pi)]$ ($n=1,2,\ldots$) ([^15]). In the remainder of this section the notations $\mathcal{V}$, ${\mathfrak{H}}$, $\mathfrak{D}$, ${H}$, $(\mu_n)_{n=1,2,\ldots}$ will always indicate, respectively, the potential $\mathcal{V}$ in Eq. (\[Eq:PotentialB\]), the Hilbert space in Eq. (\[hilbu\]), the domain and the operator in Eq. (\[hB\]), and the eigenvalues of this operator in increasing order. Sometimes it will be useful to emphasize that the potential $\mathcal{V}$ depends on the parameter $B \in (0,+\infty)$, thus originating in a similar dependence for the corresponding operator and its eigenvalues: $\mathcal{V} \equiv \mathcal{V}_B$, ${H}\equiv {{H}}_B$, $\mu_n \equiv \mu_n(B)$ ($n=1,2,\ldots$). As discussed in section \[SubSec:Pertads\], the existence of the zero-energy solutions in Eq. (\[Eq:EBZeroMode\]) implies that the ground state energy is negative, while all other eigenvalues are positive, such that $$\label{signs} \mu_1 < 0 < \mu_2 < \mu_3 < \ldots\,,$$ the negative eigenvalue $\mu_1$ being associated with an exponentially in time growing mode of the master equation (\[Eq:MasterAdS\]), whereas in contrast to this, the eigenvalues $\mu_n$ for $n\geqslant 2$ are associated with oscillatory modes. In what follows, we provide estimates for the eigenvalues of $\mu_n(B)$. We start with an upper bound for the ground state energy $\mu_1 \equiv \mu_1(B)$. According to the Rayleigh-Ritz variational characterization (see e.g. [@Schmudgen-Book], pages 265-266) one has $$\mu_1(B)=\inf_{f \in \mathfrak{D} \setminus \{0\}} {\Braket{f|{H}_B f} \over||f||^2} \, . \label{Eq:InfSpectrum}$$ Choosing in $\mathfrak{D}$ the function $$f(u) := \cos{u \over 2}$$ we get $${\Braket{f|{H}_B f} \over||f||^2}=\frac{1}{4}-B^2+\frac{\sqrt{1 + B^2} \left(4 B^2 -3\right)}{4B} =: \varepsilon(B)\,, \label{Eq:epsilonB}$$ which, together with Eq. (\[Eq:InfSpectrum\]), yields the estimate $$\mu_1(B)\leqslant \epsilon(B)$$ for each $B>0$. It can be checked that $B\mapsto\varepsilon(B)$ is a negative, monotonously increasing function on $(0,+\infty)$ with the properties $$\lim\limits_{B\to 0^+} \varepsilon(B) = -\infty\,,\qquad \lim\limits_{B\to +\infty} \varepsilon(B) = 0^-\,.$$ Therefore, we obtain the upper bound for the ground state energy, $$\mu_1(B)\leqslant \varepsilon(B) < 0 \label{Eq:mu1Bound}$$ which provides an independent proof for the fact that it is negative, and hence also for the linear instability of the AdS wormhole. Next, we provide two-sided bounds on the eigenvalues $\mu_n \equiv \mu_n(B)$ for arbitrary $n$. In order to achieve this, we check that for any fixed $B > 0$, one has $$\min_{u \in [-\pi,\pi]} {\mathcal{ V} }(u) = {\mathcal{ V} }(0) = - {1 \over 4} - {3 \over 4 B^2}\,, \qquad \max_{u \in [-\pi,\pi]} {\mathcal{ V} }_B(u) = {\mathcal{ V} }_B(\pm \pi) = - {1 \over 4} + {1 \over 4 (1 + B^2)}\,. \label{Eq:LimPot}$$ In the Hilbert space ${\mathfrak{H}}$, let us consider the operators ${H}=-{d^2\over du^2}+{\cal V}$, ${H}^- :=-{d^2\over du^2} - {1 \over 4} - {3 \over 4 B^2}$ and ${H}^+ :=-{d^2\over du^2} - {1 \over 4} + {1 \over 4 (1 + B^2)}$, all of them with the same domain $\mathfrak{D}$ as defined in Eq. (\[hB\]) (and all of them satisfying the properties (i-v) after the cited equation). Due to Eq. (\[Eq:LimPot\]) we have $\Braket{f|{H}^- f}\leqslant \Braket{f|{H}f}\leqslant\Braket{f|{H}^+ f}$ for all $f\in \mathfrak{D}$, and this implies (see e.g. [@Schmudgen-Book], pages 230 and 267) $\mu^{-}_n \leqslant \mu_n \leqslant \mu^{+}_n$ for $n=1,2,\ldots$, where $\mu^{\mp}_1 < \mu^{\mp}_2 < \ldots$ are the eigenvalues of ${H}^{\mp}$. On the other hand, the eigenvalues of ${H}^{\mp}$ are obtained shifting those of ${H}^0=-d^2/du^2$, i.e., $\mu^{-}_n ={n^2 \over 4} - {1 \over 4} - {3 \over 4 B^2}$ and $\mu^{+}_n ={n^2 \over 4} - {1 \over 4} + {1 \over 4 (1 + B^2)}$. In conclusion, the eigenvalues of ${H}$ satisfy the two-side bounds $$\label{Eq:boundsmun} {n^2 - 1 \over 4} - {3 \over 4 B^2} \leqslant \mu_n(B) \leqslant {n^2 - 1\over 4} + {1 \over 4 (1 + B^2)} \qquad (n=1,2,\ldots)\,.$$ Combining this result with Eq. (\[Eq:mu1Bound\]) one obtains the following two-side bound for the ground state energy: $$\label{Eq:boundsmu1} -\frac{3}{4B^2} \leqslant \mu_1(B) \leqslant \varepsilon(B) = -\frac{1}{2B^2} + {\cal O}\left( \frac{1}{B^4} \right)\,.$$ After these remarks concerning the spectral values of the Schrödinger operator ${H}$, we discuss the spectral decomposition of the master equation. To this purpose, we choose for each $n$ a normalized eigenfunction $e_n$ for the (simple) eigenvalue $\mu_n$: $$\label{basisads} e_n \in \mathfrak{D}~, \quad {H}e_n = \mu_n e_n~, \quad \| e_n \| = 1 \qquad (n=1,2,\ldots) \, .$$ Then $(e_n)_{n=1,2,\ldots}$ is an orthonormal basis of ${\mathfrak{H}}$ (in the ordinary sense), due to the spectral theorem for selfadjoint operators with purely discrete spectrum. In comparison with the previous analysis for the Ellis-Bronnikov wormhole, we do not have the technical complications associated with the continuous spectrum and to the related “improper” eigenfunctions. Next, we write the master equation (\[Eq:MasterAdS\]) in a form similar to (\[masterbro\]) and add initial conditions as in (\[cond\]); in this way we obtain the system $$\ddot{\chi}(s) + {H}\chi(s) = 0 ~~ (s \in {\mathbb{R}})\,, \quad \chi(0)=q\,,\quad \dot{\chi}(0) = p\,, \label{masterads}$$ where $\chi(s)$ refers to the function $u \mapsto \chi(s,u)$, the dots stand for $s$-derivatives and $q: u \mapsto q(u)$, $p: u \mapsto p(u)$ are functions with appropriate regularity. A technically precise framework for the discussion of the system (\[masterads\]) is provided by Appendix \[appeads\] where we introduce (similarly to the previous treatment for the Ellis-Bronnikov wormhole) the selfadjoint operator $| {H}|^{1/2}: \mathfrak{D}^{1/2} \subset {\mathfrak{H}}\to {\mathfrak{H}}$ and indicate how to regard the domains $\mathfrak{D}^{1/2}$ and $\mathfrak{D}$ as Hilbert spaces with appropriate inner products. It turns out that, for any $q \in \mathfrak{D}$ and $p \in \mathfrak{D}^{1/2}$, the system (\[masterads\]) has a unique solution $s \mapsto \chi(s)$ in $C({\mathbb{R}}, \mathfrak{D}) \cap C^1({\mathbb{R}}, \mathfrak{D}^{1/2}) \cap C^2({\mathbb{R}}, {\mathfrak{H}})$; using an orthonormal basis $(e_n)_{n=1,2,\ldots}$ as in Eq. (\[basisads\]), the solution can be written as follows for all $s \in {\mathbb{R}}$: $$\chi(s) = \left[ \langle e_1 | q \rangle \cosh( |\mu_1|^{1/2} s) + \langle e_1 | p \rangle \frac{\sinh( |\mu_1|^{1/2} s)}{|\mu_1|^{1/2}} \right] e_1 + \sum_{n=2}^{\infty} \left[ \langle e_{n} | q \rangle \cos( \mu^{1/2}_n s) + \langle e_{n} | p \rangle \frac{\sin( \mu^{1/2}_n s)}{\mu^{1/2}_n} \right] e_{n}\, . \label{solads}$$ The above function $\chi(s)$ is real valued for each $s$ if and only if the data $q, p$ are real valued functions. The coefficient of $e_1$ in Eq. (\[solads\]) diverges exponentially both for $s \to -\infty$, and for $s \to + \infty$ (except for very special choices of $\langle e_1 | q \rangle$ and $\langle e_1 | p \rangle$ ([^16])); so, the AdS wormhole is linearly unstable. For each $n \geqslant 2$, the $n$-th term in Eq. (\[solads\]) represents a “normalizable” oscillatory mode, living like $e_n$ inside the Hilbert space ${\mathfrak{H}}$ (indeed, inside the subspace $\mathfrak{D}\subset {\mathfrak{H}}$). This is a relevant difference with respect to the “non normalizable” oscillatory modes that we have found for the perturbed Ellis-Bronnikov wormhole, associated with the continuous spectrum and living outside the Hilbert space of the system (see the comments after Eqs. (\[solbro\]) and (\[solbrons\])). Instability times {#Sec:SpectralD} ----------------- In the Ellis-Bronnikov case it has been shown [@jGfGoS09a] that the time scale $\tau_\text{unstable}$ (measured with respect to proper time at the throat of the unperturbed solution) associated with the instable mode is of the order of the throat’s areal radius $r_\text{throat}$ divided by the speed of light. The estimates provided in Eq. (\[Eq:boundsmu1\]) allows us to estimate the corresponding time scale for the AdS wormhole, and yields $$\frac{1}{\sqrt{3}} \leq \frac{\tau_\text{unstable}}{r_\text{throat}} \leq \frac{1}{2B\sqrt{-\varepsilon(B)}}\,,$$ with the function $\varepsilon(B)$ defined in Eq. (\[Eq:epsilonB\]). Since $2B\sqrt{-\varepsilon(B)} \to \sqrt{2}$ for large $B$ and since for $B\to 0$ the AdS wormhole reduces to the reflection-symmetric Ellis-Bronnikov wormhole ([^17]), it follows also in this case that $\tau_\text{unstable}$ is of the order of the throat’s areal radius (divided by the speed of light in physical units). Conclusions {#Sec:Conclusions} =========== In this work we have analyzed the linear stability of a class of static, spherically symmetric wormhole solutions in GR minimally coupled with a self-interacting phantom scalar field. To this purpose, we have provided a gauge-invariant perturbation formalism that describes the dynamics of linearized, spherical but time-dependent perturbations of the metric and of the scalar field, resulting in a coupled $2\times 2$ linear wave system subject to a constraint (see Eqs. (\[Eq:WaveSystem\],\[Eq:WaveConstraint\])). Provided that a nontrivial, time-independent solution is known (as is usually the case when a family of static solutions is known) we have shown that this system can be decoupled to yield a master wave equation which is manifestly gauge-invariant and regular at the throat. This construction relies on a basic requirement (of course satisfied by the examples that we treat): the derivative $\Phi'$ of the (background) scalar field should vanish nowhere. The relevance of this condition in our approach is indicated by the almost ubiquitous presence of the reciprocal $1/\Phi'$ in the equations of sections \[Sec:LinearPerturbation\]-\[Sec:Decoupling\]. Based on our formalism we have re-derived the regular master equation first obtained in [@jGfGoS09a], describing linear spherical perturbations of the Ellis-Bronnikov wormhole in a fully gauge-invariant setting and without intermediate steps involving singularities at the throat. (For an alternative approach which treats the reflection-symmetric case in a fixed gauge, see [@fCfPlP19].) Furthermore, we have analyzed the linear stability of an AdS wormhole introduced in [@kBjF06], for which the scalar field is subject to a non-trivial self-interaction term, and we have shown that this solution is linearly unstable as well. In both examples the instability is characterized by a unique, exponentially in time growing mode associated with a bound state of negative energy of the Schrödinger operator arising in the master equation. As discussed in section \[Sec:SpectralD\] the associated instability times are rather short (of the order of a light-crossing time corresponding to the areal radius of the throat.) Based on spectral analysis, we have also provided a detailed and rigorous discussion for the mode decomposition of the solutions to the master wave equations in both the aforementioned examples, which revealed that besides the exponentially in time growing modes, there might be linearly growing modes, while all the remaining modes are oscillatory. In particular, the AdS wormhole has infinitely many normalizable, oscillatory modes in addition to the pair of exponential growing and decaying modes associated with the unique bound state of negative energy of the Schrödinger operator. Before concluding, we think it is necessary to return to the basic requirement of our approach already indicated above, i.e., the condition that $\Phi$ has no critical points. Removing this requirement would be interesting since, recently, a large class of new wormhole solutions of the Einstein-scalar equations has been found [@Carvente:2019gkd], generalizing previous work [@Dzhunushaliev:2017syc], in which the scalar field $\Phi$ has an extremum at the throat. Since $r$ has a global minimum at the throat and $r'$ converges to zero as fast or faster than $\Phi'$, it turns out the gauge-invariant quantity $C$ defined in Eq. (\[Eq:CGI\]) is still well-defined; unfortunately, it is unclear if a decoupled equation for $C$ can be obtained which is regular at the throat. In connection with this problem, one could try to recover the S-deformation method of [@jGfGoS09a; @kBjFaZ11] (see the discussion in the Introduction; the formulation of this method in [@kBjFaZ11] indeed considers the gauge-invariant quantity $C$). However, when the potential $V(\Phi)$ is non-zero, this method seems to require the numerical integration of a Riccati-type equation to find the regularized potential, and further one still needs to justify *a posteriori* the validity of the transformed equation at the throat. An alternative possibility consists in applying a variation of the approach discussed in this article, in which $\Phi'$ is absent from all denominators, thanks to the use of new gauge-invariant quantities in place of the functions $A,C,E$ of Eqs. (\[Eq:AGI\]-\[Eq:EGI\]); however, it is not clear whether this will be possible, and we prefer to leave the discussion of the problem to future works. Another problem that we think worthy of a future analysis is the following: is there a *deep* geometrical reason for which the present approach succeeds, in certain cases, in decoupling the perturbation equations (\[Eq:WaveSystem\]-\[Eq:WaveConstraint\]) and reducing them to a single, scalar master equation? Typically, the possibility of reducing to a simpler form a PDE or of a system of PDEs is due to the presence of a Lie group of symmetries; an interpretation of this kind could perhaps be given for our decoupling method. As already recalled, our approach uses a static solution of Eqs. (\[Eq:WaveSystem\]-\[Eq:WaveConstraint\]), arising from variations with respect to the parameters of a *family* of static wormhole solutions. The availability of such parametric families could perhaps be interpreted in terms of a Lie group of symmetries, acting on the static solutions of the Einstein-scalar system; if so, it would be interesting to understand the interplay of these symmetries with the linearized perturbation equations. F.C. and L.P. were supported by: INdAM, Gruppo Nazionale per la Fisica Matematica; Università degli Studi di Milano. LP was also supported by: INFN; MIUR, PRIN 2010 Research Project Geometric and analytic theory of Hamiltonian systems in finite and infinite dimensions.". O.S. was partially supported by a CIC grant to Universidad Michoacana. On the master equation for the reflection-symmetric Ellis-Bronnikov wormhole {#appebro} ============================================================================ We refer to the master equation for this wormhole in the formulation (\[masterbro\]), based on the Hilbert space ${\mathfrak{H}}:= L^2({\mathbb{R}}, dx)$ of Eq. (\[hilbro\]) and on the selfadjoint operator ${H}$ of Eq. (\[hbro\]), of domain $\mathfrak{D} \subset {\mathfrak{H}}$. We keep all notations introduced after these equations; in particular $\langle~|~\rangle$ and $\|~\|$ are the natural inner product and norm of ${\mathfrak{H}}$.\ \ **Relevant facts on the operator ${H}$ and its spectral features.** In the discussion following Eqs. (\[masterbro\]-\[hbro\]) we have mentioned a system made by a normalized eigenfunction $e_1$ for the unique eigenvalue $\mu_1 < 0$ of ${H}$, and by a pair of improper eigenfunctions $e_{i \lambda}$ ($i=1,2$) (lying in $C^\infty({\mathbb{R}})$ but not in ${\mathfrak{H}}$) for each $\lambda>0$. We have called this system a generalized orthonormal basis of ${\mathfrak{H}}$, which means that the following conditions (a-c) hold [@Ber]: - Consider the space $C_c({\mathbb{R}}) \equiv \mathfrak{C} \subset {\mathfrak{H}}$, made of the continuous functions $f : {\mathbb{R}}\to {\mathbb{C}}$ with compact support. For $f \in \mathfrak{C}$, consider the usual inner product $\langle e_1 | f \rangle = \int_{{\mathbb{R}}} d x \bar{e}_1(x) f(x)$ and define in addition a “generalized inner product” $\langle e_{i \lambda} | f \rangle := \int_{{\mathbb{R}}} d x \, \bar{e}_{i \lambda}(x) f(x)$ (the integral converges since $\bar{e}_{i \lambda} f \in \mathfrak{C}$); then, the maps $$(0,+\infty) \ni \lambda \mapsto \langle e_{i \lambda} | f \rangle \in {\mathbb{C}}\quad(i=1,2)$$ are both in $L^2((0,+\infty), d \lambda$). - For $i=1,2$, the linear map $\mathfrak{C} \subset {\mathfrak{H}}\to L^2((0,+\infty), d \lambda)$, $f \mapsto \Big( \lambda \mapsto \langle e_{i \lambda} | f \rangle \Big)$ is continuous with respect to the norms of the Hilbert spaces ${\mathfrak{H}}$ and $L^2((0,+\infty), d \lambda)$; thus, by the density of $\mathfrak{C}$ in ${\mathfrak{H}}$, this map has a unique continuous (and linear) extension to ${\mathfrak{H}}$, that we write as $${\mathfrak{H}}\to L^2((0,+\infty), d \lambda)\,, \quad f \mapsto \Big( \lambda \mapsto \langle e_{i \lambda} | f \rangle \Big)\,.$$ For each $f \in {\mathfrak{H}}$, the map $\lambda \mapsto \langle e_{i \lambda} | f \rangle$ is said to give the “generalized inner products” between the $e_{i \lambda}$’s and $f$. - Consider the direct sum Hilbert space ${\mathbb{C}}\oplus L^2((0,+\infty), d \lambda) \oplus L^2((0,+\infty), d \lambda)$ with its natural inner product; then the linear map $${\mathfrak{H}}\to {\mathbb{C}}\oplus L^2((0,+\infty), d \lambda) \oplus L^2((0,+\infty), d \lambda)\,, \quad f \mapsto \Big(\langle e_1 | f \rangle, \, \lambda \mapsto \langle e_{1 \lambda} | f \rangle, \, \lambda \mapsto \langle e_{2 \lambda} | f \rangle \Big) \label{unit}$$ is a unitary, i.e., it is one-to-one and preserves inner products: $$\langle f | \ell \rangle = \overline{\langle e_1 | f \rangle} \langle e_1 | \ell \rangle + \sum_{i=1}^2 \int_{0}^{+\infty} \hspace{-0.4cm} d \lambda \, \overline{\langle e_{i \lambda} | f \rangle} \langle e_{i \lambda} | \ell \rangle \quad (f,\ell \in {\mathfrak{H}})\,.$$ The forthcoming items (i-ii) describe some consequences of (a-c): - For $F \in L^2((0,+\infty), d \lambda)$ and $i = 1,2$ there is a unique element of ${\mathfrak{H}}$, indicated with $\int_{0}^{+\infty} d \lambda F(\lambda) e_{i \lambda}$, such that $$\label{Eq:wi} \left \langle \int_{0}^{+\infty} \hspace{-0.4cm} d \lambda F(\lambda) e_{i \lambda} \bigg| \ell \right \rangle = \int_{0}^{+\infty} \hspace{-0.4cm} d \lambda \overline{F}(\lambda) \langle e_{i \lambda} | \ell \rangle \quad \mbox{for all $\ell \in {\mathfrak{H}}$}$$ (note that the integral in the right hand side of Eq. (\[Eq:wi\]) exists, involving the product of two functions which are both in $L^2((0,+\infty), d \lambda)$). The element $\int_{0}^{+\infty} d \lambda F(\lambda) e_{i \lambda} \in {\mathfrak{H}}$ is called the weak integral of the function $\lambda \mapsto F(\lambda) e_{i \lambda}$. - The inverse of the unitary map (\[unit\]) can be expressed in terms of weak integrals; more precisely, such inverse is the map $${\mathbb{C}}\oplus L^2((0,+\infty), d \lambda) \oplus L^2((0,+\infty), d \lambda) \to {\mathfrak{H}}\,,\quad (\alpha, F_1, F_2) \mapsto \alpha e_1 + \sum_{i=1}^2 \int_{0}^{+\infty} \hspace{-0.4cm} d \lambda F_i(\lambda) e_{i \lambda}\,. \label{unitinv}$$ The fact that the composition of the maps (\[unit\],\[unitinv\]) is the identity map ${\mathfrak{H}}\to {\mathfrak{H}}$, $f \mapsto f$ can be written explicitly as follows: for each $f \in {\mathfrak{H}}$, $$f = \langle e_1 | f \rangle e_1 + \sum_{i=1}^2 \int_{0}^{+\infty} \hspace{-0.4cm} d \lambda \langle e_{i \lambda} | f \rangle e_{i \lambda}\,. \label{frep}$$ The identity (\[frep\]) is said to give the expansion of $f$ in terms of the generalized orthonormal basis under consideration. Up to now, we have not used the fact that $e_1$ is an eigenfunction of ${H}$ with eigenvalue $\mu_1 < 0$, nor the fact that $e_{i \lambda}$ is an “improper eigenfunction” with “eigenvalue” $\lambda$ for $i=1,2$ and all $\lambda > 0$. These facts yield the following representation for the operator ${H}$ and its domain $\mathfrak{D}$: $$\begin{gathered} \mathfrak{D} = \{ f \in {\mathfrak{H}}~|~\big(\lambda \mapsto \lambda \langle e_{i \lambda}| f \rangle \big) \in L^2((0,+\infty), d \lambda)~\mbox{for $i=1,2$} \}\,. \\ \mbox{For}~ f \in \mathfrak{D}: \quad \langle e_1 | {H}f \rangle = \mu_1\langle e_1 | f \rangle\,, ~\langle e_{i \lambda} | {H}f \rangle = \lambda\langle e_{i \lambda} | f \rangle\,, ~~\mbox{i.e.}\,, ~~{H}f = \mu_1 \langle e_1 | f \rangle e_1 + \sum_{i=1}^2 \int_{0}^{+\infty} \hspace{-0.4cm} d \lambda \, \lambda \langle e_{i \lambda} | f \rangle e_{i \lambda}\,. \end{gathered} \label{reph}$$ As well known, a functional calculus exists for selfadjoint Hilbert space operators (see, e.g., [@Schmudgen-Book]). This allows to define an operator $\mathscr{F}({H}) : \mathfrak{D}^{\mathscr{F}} \subset {\mathfrak{H}}\to {\mathfrak{H}}$ for each (Borel-) measurable function $\mathscr{F} : \sigma({H}) \to {\mathbb{C}}$, where $\sigma({H}) = \{ \mu_1 \} \cup [0,+\infty)$ is the spectrum of ${H}$ and $\mathfrak{D}^{\mathscr{F}}$ is a suitable domain, determined by (${H}$ and) $\mathscr{F}$; the operator $\mathscr{F}({H})$ is selfadjoint if $\mathscr{F}$ is real valued. Making reference to the previously mentioned, generalized orthonormal basis of eigenfunctions of ${H}$, one can prove the following statements: $e_1 \in \mathfrak{D}^{\mathscr{F}}$ and $\mathscr{F}({H}) e_1 = \mathscr{F}(\mu_1) e_1$; $\mathfrak{D}^{\mathscr{F}} = \{ f \in {\mathfrak{H}}~|~\big(\lambda \mapsto \mathscr{F}(\lambda) \langle e_{i \lambda}| f \rangle \big) \in L^2((0,+\infty), d \lambda)~\mbox{for $i=1,2$} \}$; for all $f \in \mathfrak{D}^{\mathscr{F}}$, one has $\langle e_1 | \mathscr{F}({H}) f \rangle = \mathscr{F}(\mu_1) \langle e_1 | f \rangle$ and $\langle e_{i \lambda} | \mathscr{F}({H}) f \rangle = \mathscr{F}(\lambda) \langle e_{i \lambda} | f \rangle$. For our purposes it is important to consider the choice $\mathscr{F}(\gamma) := | \gamma|^{1/2}$ for all $\gamma \in \sigma({H})$, producing a selfadjoint operator that we indicate with $$| {H}|^{1/2} : \mathfrak{D}^{1/2} \subset {\mathfrak{H}}\to {\mathfrak{H}}$$ and that behaves as follows in relation to our generalized orthonormal basis: $$e_1 \in \mathfrak{D}^{1/2}\,,\quad|{H}|^{1/2} e_1 = |\mu_1|^{1/2} e_1\,.$$ $$\begin{gathered} \mathfrak{D}^{1/2} = \{ f \in {\mathfrak{H}}~|~\big(\lambda \mapsto \lambda^{1/2} \langle e_{i \lambda}| f \rangle \big) \in L^2((0,+\infty), d \lambda)~\mbox{for $i=1,2$} \}\,.~\mbox{For}~ f \in \mathfrak{D}^{1/2}:\\ {~} \hspace{-0.5cm} \langle e_1 \big| |{H}|^{1/2} f \rangle = |\mu_1|^{1/2} \langle e_1 | f \rangle, \langle e_{i \lambda} \big| |{H}|^{1/2} f \rangle = \lambda^{1/2} \langle e_{i \lambda} | f \rangle,~\mbox{i.e.},~ |{H}|^{1/2} f = |\mu_1|^{1/2} \langle e_1 | f \rangle e_1 + \sum_{i=1}^2 \int_{0}^{+\infty} \hspace{-0.6cm} d \lambda \, \lambda^{1/2} \langle e_{i \lambda} | f \rangle e_{i \lambda}\,. \end{gathered} \label{reph12}$$ Finally, let us make explicit the Hilbert space structures for $\mathfrak{D}$ and $\mathfrak{D}^{1/2}$ mentioned before Eq. (\[solbro\]); these are provided by the (complete) inner products $\langle~|~\rangle_{\mathfrak{D}} : \mathfrak{D} \times \mathfrak{D} \to {\mathbb{C}}$ and $\langle~|~\rangle_{\mathfrak{D}^{1/2}} : \mathfrak{D}^{1/2} \times \mathfrak{D}^{1/2} \to {\mathbb{C}}$, where $$\langle f | \ell \rangle_{\mathfrak{D}} := \langle f | \ell \rangle + \langle {H}f | {H}\ell \rangle = (1 + \mu_1^2) \, \overline{\langle e_1 | f \rangle} \langle e_1 | \ell \rangle + \sum_{i=1}^2 \int_{0}^{+\infty} \hspace{-0.4cm} d \lambda \, (1 + \lambda^2) \, \overline{\langle e_{i \lambda} | f \rangle} \langle e_{1 \lambda} | f \rangle\,, \label{scadbro}$$ $$\langle f | \ell \rangle_{\mathfrak{D}^{1/2}} := \langle f | \ell \rangle + \langle |{H}|^{1/2} f \big| |{H}|^{1/2} \ell \rangle = (1 + |\mu_1|) \, \overline{\langle e_1 | f \rangle} \langle e_1 | \ell \rangle + \sum_{i=1}^2 \int_{0}^{+\infty} \hspace{-0.4cm} d \lambda \, (1 + \lambda) \, \overline{\langle e_{i \lambda} | f \rangle} \langle e_{1 \lambda} | f \rangle\,. \label{scad12bro}$$ **Solution of the master equation.** Let us consider Eq. (\[masterbro\]) with initial conditions (\[cond\]) i.e., $\ddot{\chi}(t) + {H}\chi(t) = 0$, $\chi(0)=q$, $\dot{\chi}(0) = p$; the unknown is a function ${\mathbb{R}}\ni t \mapsto \chi(t) \in \mathfrak{D}$. We first proceed formally, assuming that the initial data $q, p$ are in suitable spaces to be specified later. Applying $\langle e_1 |~\rangle$ and $\langle e_{i \lambda}|~\rangle$ to Eq. (\[masterbro\]) we obtain $(d^2/ d t^2 + \mu_1) \langle e_1 | \chi(t) \rangle = 0$ and $(d^2/ d t^2 + \lambda) \langle e_{i \lambda} | \chi(t) \rangle = 0$ for $i=1,2$ and all $\lambda > 0$. On account of the initial conditions (\[cond\]) (and recalling that $\mu_1 < 0$), these equations imply $$\langle e_1 | \chi(t) \rangle = \langle e_1 | q \rangle \cosh( |\mu_1|^{1/2} t) + \langle e_1 | p \rangle \frac{\sinh( |\mu_1|^{1/2} t)}{|\mu_1|^{1/2}} \,, \quad \langle e_{i \lambda} | \chi(t) \rangle = \langle e_{i \lambda} | q \rangle \cos( \lambda^{1/2} t) + \langle e_{i \lambda} | p \rangle \frac{\sin( \lambda^{1/2} t)}{\lambda^{1/2}}\,,$$ thus providing a formal justification for the expression (\[solbro\]) of $\chi(t)$. It can be checked a posteriori that, assuming $$q \in \mathfrak{D}\,, \quad p \in \mathfrak{D}^{1/2}\,, \label{assuqp}$$ all the previous manipulations make sense and Eq. (\[solbro\]) describes the unique solution $\chi: {\mathbb{R}}\ni t \mapsto \chi(t)$ of Eqs. (\[masterbro\],\[cond\]) such that $$\label{solspace} \chi \in C^2({\mathbb{R}}, {\mathfrak{H}}) \cap C^1({\mathbb{R}}, \mathfrak{D}^{1/2}) \cap C({\mathbb{R}}, \mathfrak{D})\,.$$ As an example of the necessary tests, let us consider any $t \in {\mathbb{R}}$ and show that $\chi(t)$ defined by Eq. (\[solbro\]) is an element of $\mathfrak{D}$. Due to the descriptions (\[unit\],\[unitinv\]) for ${\mathfrak{H}}$ and (\[reph\]) for $\mathfrak{D}$, $\chi(t)$ in Eq. (\[solbro\]) is in fact in $\mathfrak{D}$ if we are able to prove the following for $i=1,2$ (and for fixed $t$, as already indicated):\ Indeed, Eq. (\[ver1\]) follows noting that $$\lambda \mapsto \cos( \lambda^{1/2} t)\,,~\lambda \mapsto \frac{\sin( \lambda^{1/2} t)}{\lambda^{1/2}} \in L^\infty((0,+\infty), d \lambda)\,;\quad \lambda \mapsto \langle e_{i \lambda} | q \rangle\,,~\lambda \mapsto \langle e_{i \lambda} | p \rangle \in L^2((0,+\infty), d \lambda) \label{a18}$$ (the statements on $q, p$ in (\[a18\]) are correct, since Eq. (\[assuqp\]) obviously implies $q, p \in {\mathfrak{H}}$). Moreover, Eq. (\[ver2\]) follows noting that $$\lambda \mapsto \cos( \lambda^{1/2} t)\,,~\lambda \mapsto \sin( \lambda^{1/2} t) \in L^\infty((0,+\infty), d \lambda)\,;\quad \lambda \mapsto \lambda \langle e_{i \lambda} | q \rangle\,,~\lambda \mapsto \lambda^{1/2} \langle e_{i \lambda} | p \rangle \in L^2((0,+\infty), d \lambda) \label{a19}$$ (the statements on $q, p$ in (\[a19\]) are correct, due to the assumption (\[assuqp\]) $q \in \mathfrak{D}, p \in \mathfrak{D}^{1/2}$ and to the characterizations (\[reph\]) for $\mathfrak{D}$, (\[reph12\]) for $\mathfrak{D}^{1/2}$). On the master equation for the AdS wormhole {#appeads} =========================================== **Facts on the operator ${H}$.** Let us consider the Hilbert space of Eq. (\[hilbu\]), i.e., ${\mathfrak{H}}:= L^2((-\pi,\pi), du)$, with its natural inner product $\langle~|~\rangle$; this is the environment for the selfajoint operator of Eq. (\[hB\]), i.e., ${H}:= -{d^2 \over du^2}+ \mathcal{ V} : \mathfrak{D} \subset {\mathfrak{H}}\to {\mathfrak{H}}$ with domain $\mathfrak{D} := \{ f \in {\mathfrak{H}}~|~f_{u u} \in {\mathfrak{H}}~, f(\pm \pi) = 0 \}$. Let us recall that ${H}$ has purely discrete spectrum with simple eigenvalues $\mu_1 < 0 < \mu_2 < \mu_3 < \ldots$ (see Eq. (\[signs\])); due to Eq. (\[Eq:boundsmun\]), we have $\mu_n \sim {n^2 \over 4}$ for $n \mapsto + \infty$. In the sequel we frequently make use of an orthonormal basis $(e_n)_{n=1,2,\ldots}$ of ${\mathfrak{H}}$ as in Eq. (\[basisads\]), obtained choosing for each $n$ a normalized eigenfunction $e_n$ for the eigenvalue $\mu_n$. The fact that we have an orthonormal basis ensures that there is a one-to-one linear map $${\mathfrak{H}}\to \mathfrak{l}^2\,, \quad f \mapsto \big(\langle e_n| f \rangle\big)_{n=1,2,\ldots} \label{Eq:B1}$$ where $\mathfrak{l}^2$ is the Hilbert space of complex sequences $(a_n)_{n=1,2\ldots}$ such that $\sum_{n=1}^{\infty} |a_n|^2 < + \infty$, with its obvious inner product; moreover $\langle f | \ell \rangle = \sum_{n=1}^{\infty} \overline{\langle e_n| f \rangle} \langle e_n| \ell \rangle$ for all $f, \ell \in {\mathfrak{H}}$, i.e., the map (\[Eq:B1\]) is unitary. The fact that the orthonormal basis is formed by eigenfunctions of ${H}$ ensures the following representation for this operator and its domain $$\mathfrak{D} = \{ f \in {\mathfrak{H}}~|~\big(\mu_n \langle e_{n}| f \rangle \big)_{n=1,2,\ldots} \hspace{-0.2cm} \in \mathfrak{l}^2 \}\,.~ \mbox{For}~ f \in \mathfrak{D}:~\langle e_n | {H}f \rangle = \mu_n \langle e_{n}| f \rangle~(n=1,2,\ldots)\,,\mbox{i.e.}\,,~ {H}f = \sum_{n=1}^{+\infty} \mu_n \langle e_n | f \rangle e_n\,. \label{rephb}$$ In the previous Appendix \[appebro\] we have already mentioned the functional calculus for selfadjoint Hilbert space operators [@Schmudgen-Book]; this allows to define an operator $\mathscr{F}({H}) : \mathfrak{D}^{\mathscr{F}} \subset {\mathfrak{H}}\to {\mathfrak{H}}$ for each function $\mathscr{F} : \sigma({H}) \to {\mathbb{C}}$ where $\sigma({H}) = \{ \mu_1, \mu_2,\ldots\}$ is the spectrum of ${H}$; the operator $\mathscr{F}({H})$ is selfadjoint if $\mathscr{F}$ is real valued. With the choice $\mathscr{F}(\gamma) := |\gamma|^{1/2}$ we obtain a selfajoint operator indicated with $$| {H}|^{1/2} : \mathfrak{D}^{1/2} \subset {\mathfrak{H}}\to {\mathfrak{H}}\,,$$ which behaves as follows with respect to the previous orthonormal basis $(e_n)_{n=1,2,\ldots}$ of eigenfunctions of ${H}$: $$e_n \in \mathfrak{D}^{1/2}\,,\quad| {H}|^{1/2} e_n = | \mu_n |^{1/2} e_n ~~ (n=1,2,\ldots)\,,$$ $$\begin{gathered} \mathfrak{D}^{1/2} = \{ f \in {\mathfrak{H}}~|~\big(|\mu_n|^{1/2} \langle e_{n}| f \rangle \big)_{n=1,2,\ldots} \hspace{-0.2cm} \in \mathfrak{l}^2 \}\,. \\ \mbox{For}~ f \in \mathfrak{D}^{1/2}:~~\langle e_n | |{H}|^{1/2} f \rangle = |\mu_n|^{1/2} \langle e_{n}| f \rangle~(n=1,2,\ldots)\,,~~\mbox{i.e.}\,,~~ {H}f = \sum_{n=1}^{+\infty} |\mu_n|^{1/2} \langle e_n | f \rangle e_n \end{gathered} \label{rephb12}$$ (here and in the sequel, recall that $|\mu_n| = \mu_n$ for $n \geqslant 2$). Now, let us give Hilbert space structures to the domains $\mathfrak{D}$ and $\mathfrak{D}^{1/2}$; these are provided by the (complete) inner products $\langle~|~\rangle_{\mathfrak{D}} : \mathfrak{D} \times \mathfrak{D} \to {\mathbb{C}}$ and $\langle~|~\rangle_{\mathfrak{D}^{1/2}} : \mathfrak{D}^{1/2} \times \mathfrak{D}^{1/2} \to {\mathbb{C}}$, where ([[^18]]{}) $$\langle f | \ell \rangle_{\mathfrak{D}} := \langle {H}f | {H}\ell \rangle = \sum_{n=1}^{\infty} \mu^2_n \, \overline{\langle e_n | f \rangle} \langle e_n | \ell \rangle\,, \label{scadads}$$ $$\langle f | \ell \rangle_{\mathfrak{D}^{1/2}} := \langle |{H}|^{1/2} f \big| |{H}|^{1/2} \ell \rangle = \sum_{n=1}^{\infty} |\mu_n| \, \overline{\langle e_n | f \rangle} \langle e_n | \ell \rangle\,. \label{scad12ads}$$ **Solution of the master equation.** Let us consider the master equation as written in Eq. (\[masterads\]) with the initial conditions given therein, i.e., $\ddot{\chi}(s) + {H}\chi(s) = 0$, $\chi(0)=q$, $\dot{\chi}(0) = p$; the unknown is a function ${\mathbb{R}}\ni s \mapsto \chi(s) \in \mathfrak{D}$, and the spaces containing the data $q, p$ are to be specified. As in the previous Appendix we first proceed formally. Applying $\langle e_n |~\rangle$ to the differential equation in (\[masterads\]) we obtain $(d^2/ d s^2 + \mu_n) \langle e_n | \chi(s) \rangle = 0$ for $n=1,2,\ldots\,$; taking into account the initial conditions in (\[masterads\]) and the fact that $\mu_1 < 0 < \mu_2 < \mu_3 < \ldots$, we conclude $$\begin{gathered} \langle e_1 | \chi(s) \rangle = \langle e_1 | q \rangle \cosh( |\mu_1|^{1/2} s) + \langle e_1 | p \rangle \frac{\sinh( |\mu_1|^{1/2} s)}{|\mu_1|^{1/2}}\,, \\ \langle e_{n} | \chi(s) \rangle = \langle e_{n } | q \rangle \cos( \mu^{1/2}_n s) + \langle e_{n } | p \rangle \frac{\sin( \mu^{1/2}_n s)}{\mu^{1/2}_n} \quad (n=2,3,\ldots)\,, \end{gathered}$$ thus providing a formal justification for the expression (\[solads\]) of $\chi(s)$. It can be checked a posteriori that assuming $q \in \mathfrak{D}$ and $p \in \mathfrak{D}^{1/2}$, all the previous manipulations make sense and Eq. (\[solads\]) describes the unique solution ${\mathbb{R}}\ni s \mapsto \chi(s)$ of Eq. (\[masterads\]) in the space $C^2({\mathbb{R}}, {\mathfrak{H}}) \cap C^1({\mathbb{R}}, \mathfrak{D}^{1/2}) \cap C({\mathbb{R}}, \mathfrak{D})$. The verification of these statements relies on arguments similar to those exemplified after Eqs. (\[assuqp\],\[solspace\]) of the previous Appendix. [^1]: In this article, when talking about wormholes, we always refer to *traversable* Lorentzian wormhole spacetimes in a metric theory of gravity. [^2]: However, the presence of unbounded negative kinetic energy by itself does not imply that any stationary solution in the theory is *necessarily* unstable. For example, it turns out that the Minkowski spacetime is nonlinearly stable in Einstein theory minimally coupled to a massless scalar field irrespectively of the sign of the gravitational coupling constant (see the comments and references in appendix B.5 in [@mDiR08]). [^3]: See also [@pKbKjK11] for the construction of static, spherically symmetric wormholes in Einstein-Dilaton-Gauss-Bonnet theory, a modified gravity theory, which does not require exotic matter. However, a careful stability analysis has recently revealed that these solutions are linearly unstable as well [@mCrKaZ18]. [^4]: This choice of notation is somehow awkward; however the reason for it is to maintain compatibility with the notation used in Ref. [@jGfGoS08]. [^5]: In [@fCfPlP19], the variables $x$ and $b$ of the present paper are denoted with $\ell$ and $a$; the field $\mathscr{R}$ fulfilling the master equation (3.15) of the cited work is related to the present gauge-invariant quantities $C$ and $E$ by the relation $ \frac{\partial^2}{\partial t^2}\left[\mathscr{R}\left(\frac{t}{b},\frac{x}{b}\right)\right] =\frac{r^2}{b^3}\left(r \ddot{C} - r' E\right)$. Using the linearized field equations, this can also be rewritten as $\frac{\partial^2}{\partial t^2}\left[\mathscr{R}\left(\frac{t}{b},\frac{x}{b}\right)\right] = -\frac{1}{br}(A - C)$, which explains why $\mathscr{R}$ satisfies the same master equation as $\chi$. [^6]: Note that $a$ satisfies $$a' = \frac{Y_0 \gamma}{2 \alpha} a\,,\qquad a'' = \left[ \left(\frac{Y_0 \gamma}{2 \alpha}\right)' + \left(\frac{Y_0 \gamma}{2 \alpha}\right)^2 \right] a\,.$$ [^7]: Here $Y_0'$ can be computed by taking a derivative of Eq. (\[Eq:Y0\]) and eliminating $\Phi''$ via the static version of Eq. (\[Eq:Ev3\]). [^8]: For $\gamma_1 = 0$, $F = 1$ and the statement is trivial. When $\gamma_1\neq 0$ one has $F(x)\to +\infty$ for $x\to \pm\infty$, thus $F$ has a global minimum at some $x = x_0$, where $ 0 = (1 + \gamma_1^2) b F'(x_0) = \gamma_1\left[ \gamma_1\arctan(x_0/b) + (b + \gamma_1 x_0) b/(x_0^2 + b^2) \right]$. Eliminating the $\arctan$ term one obtains from this $(1 + \gamma_1^2) F(x_0) = (x_0 - b\gamma_1)^2/(x_0^2 + b^2)$. However, this minimum value must be strictly positive since otherwise $x_0 = b\gamma_1$ which would imply that $(1 + \gamma_1^2) b F'(x_0) = \gamma_1\left(\gamma_1\arctan\gamma_1 + 1 \right)$ which cannot be zero since $\gamma_1\neq 0$. [^9]: Let us justify this statement on the number of zeroes of $\chi_0$ for the special choice $C_1 = -2\pi C_2\neq 0$. In this case we can write $\chi_0(u) =\left(-2 C_2 \cos \frac{u}{2}\right) w(u)/\sqrt{1+2 B^2-\cos u}$ where $w : (-\pi,\pi)\to {\mathbb{R}}$, $u \mapsto w(u) := { (u+\pi ) \tan \frac{u}{2}-2 B^2}$. The zeroes of $\chi_0$ in $(-\pi,\pi)$ coincide with the zeroes of the function $w$. To find the zeroes of $w$, it is useful to note that this function has derivative $w'(u) = \left({1\over 2} \sec ^2\frac{u}{2}\right)\left(u+\sin u+\pi\right)>0$ for all $u \in (-\pi,\pi)$; from $w'>0$ it follows that $w$ is a strictly monotonic bijection of $(-\pi,\pi)$ to $(-2 B^2-2, + \infty)$, and thus possesses a unique zero. [^10]: Throughout the paper, the expression “Hilbert space” is an abbreviation for “complex, separable Hilbert space”. [^11]: The conditions $f \in {\mathfrak{H}}$ and $f_{x x} \in {\mathfrak{H}}$ imply $f_x \in {\mathfrak{H}}$, due to the Gagliardo-Nirenberg interpolation inequality (see e.g. [@Adams-Book]); $\mathfrak{D}$ is just the usual Sobolev space $W^{2, 2}({\mathbb{R}}) \equiv H^2({\mathbb{R}})$. Let us also remark that, for $f \in {\mathfrak{H}}$, one has automatically $\mathcal{V} f \in {\mathfrak{H}}$ due to the boundedness of $\mathcal{V}$. [^12]: \[footnote13\]For $\langle e_1 | q \rangle = \langle e_1 | p \rangle = 0$, the coefficient of $e_1$ in (\[solbro\]) vanishes. For $\langle e_1 | q \rangle = \xi \langle e_1 | p \rangle/|\mu_1|^{1/2} \neq 0$, with $\xi = \pm 1$, the coefficient of $e_1$ diverges for $t \to \xi (+\infty)$ and vanishes for $t \to \xi (-\infty)$. [^13]: \[footnote:Hilbert\] Note that, since $d\rho={\gamma\over \alpha}dx$, working with the operator and the Hilbert space defined in Eq. (\[hbrons\]) is equivalent to working directly with the operator $-\left({\alpha\over\gamma} {d\over d x}\right)^2+{\alpha^2\over\gamma^2}\mathcal{ V}$ in the Hilbert space $L^2(\mathbb{R},{\gamma\over\alpha} dx)$, which is the formulation considered at the end of subsection \[SubSec:PertEllis\]. [^14]: The conditions $f \in {\mathfrak{H}}$, $f_{u u} \in {\mathfrak{H}}$ imply $f_u \in {\mathfrak{H}}$, due to the already mentioned Gagliardo-Nirenberg interpolation inequality [@Adams-Book]. The space $\{ f \in {\mathfrak{H}}~|~f_{u u} \in {\mathfrak{H}}\}$ coincides with the standard Sobolev space $W^{2, 2}(-\pi,\pi) \equiv H^{2}(-\pi,\pi)$, which is contained in $C^1([-\pi,\pi])$ by the Sobolev embedding theorem (see again [@Adams-Book]). Let us also remark that, due to the boundedness of the function $\mathcal{V}$, for each $f \in {\mathfrak{H}}$ one has automatically $\mathcal{V} f \in {\mathfrak{H}}$. [^15]: Let us give more complete information on the above issues (i-v). For some general facts about Hilbert space operators with properties (i-iii) (including the possibility to arrange their eigenvalues in an increasing sequence), see e.g. [@Schmudgen-Book] (especially, pages 37, 178 and 265-67). To go on, let us recall the following regularity result: if $f$ is a distribution on an open interval $\Omega \subset {\mathbb{R}}$ (with derivatives $f^{(i)}$, $i=0,1,\ldots$) and $f$ fulfills a homogeneous linear ODE $f^{(k)} + \sum_{i=0}^{k-1} a_i f^{(i)}=0$ of any order $k \in \{1,2,\ldots\}$ with $C^\infty$ coefficients $a_i: \Omega \to {\mathbb{C}}$, then $f$ is a $C^\infty$ function on $\Omega$: this follows from Theorem IX in [@Schwartz-Book], page 130. The properties (i-v) of ${H}^0$ and the expressions given above for its eigenvalues and eigenfunctions are checked “by hand”, keeping in mind that the eigenfunctions are smooth due to the previously mentioned regularity result. Now consider any function $\mathcal{V}\in C^{\infty}([-\pi,\pi],\mathbb{R})$; then, due to the boundedness of this function, the multiplication operator by $\cal V$ is a bounded selfadjoint operator on ${\mathfrak{H}}$. As well known the properties (i), or (i-ii), or (i-iii) of an operator in an abstract Hilbert space are preserved by the addition of a bounded selfadjoint perturbation (see again [@Schmudgen-Book]); therefore the operator ${H}:= {H}^0+\mathcal{V}=-d^2/du^2+\mathcal{V}$ with domain $\mathfrak{D}$ fulfills (i-iii). The operator ${H}$ also has the properties (iv-v). For the proof of (iv) one can use again the cited regularity result for distributional, homogeneous linear ODEs; a derivation of (v) can be found e.g. in [@Poschel-Book], page 30. All the previous statements apply, in particular, with ${\cal V}$ as in Eq. (\[Eq:PotentialB\]). [^16]: See the footnote \[footnote13\] in the discussion after Eq. (\[solbro\]), which is readily adapted to the present framework. [^17]: See the comment on the limit $k\to 0$ after Eq. (\[Eq:StaticSolutionsPot1\]), keeping in mind that $B=b k$. [^18]: One could as well consider the alternative inner products $\langle~|~\rangle'_{\mathfrak{D}} : \mathfrak{D} \times \mathfrak{D} \to {\mathbb{C}}$ and $\langle~|~\rangle'_{\mathfrak{D}^{1/2}} : \mathfrak{D}^{1/2} \times \mathfrak{D}^{1/2} \to {\mathbb{C}}$, defined setting $$\langle f | \ell \rangle'_{\mathfrak{D}} := \langle f | \ell \rangle + \langle {H}f | {H}\ell \rangle = \sum_{n=1}^{\infty} (1 + \mu^2_n) \, \overline{\langle e_n | f \rangle} \langle e_n | \ell \rangle\,,\qquad \langle f | \ell \rangle'_{\mathfrak{D}^{1/2}} := \langle f | \ell \rangle + \langle |{H}|^{1/2} f \big| |{H}|^{1/2} \ell \rangle = \sum_{n=1}^{\infty} (1 + |\mu_n|) \, \overline{\langle e_n | f \rangle} \langle e_n | \ell \rangle\,;$$ these have structures more similar to those of the inner products for the spaces $\mathfrak{D}$ and $\mathfrak{D}^{1/2}$ in the previous Appendix, see Eqs. (\[scadbro\],\[scad12bro\]). However, in the present situation $\langle~|~\rangle'_{\mathfrak{D}}$ and $\langle~|~\rangle'_{\mathfrak{D}^{1/2}}$ are equivalent, respectively, to the inner products $\langle~|~\rangle_{\mathfrak{D}}$ and $\langle~|~\rangle_{\mathfrak{D}^{1/2}}$ of Eqs. (\[scadads\],\[scad12ads\]).
{ "pile_set_name": "ArXiv" }
--- abstract: 'We consider a particular one-parameter family of $q$-analogues of multiple zeta values. The intrinsic $q$-regularisation permits an extension of these $q$-multiple zeta values to negative integers. Renormalised multiple zeta values satisfying the quasi-shuffle product are obtained using an Hopf algebraic Birkhoff factorisation together with minimal subtraction.' address: - '[ICMAT, C/Nicolás Cabrera, no. 13-15, 28049 Madrid, Spain]{}. ' - 'Univ. Blaise Pascal, C.N.R.S.-UMR 6620, 3 place Vasarély, CS 60026, 63178 Aubière, France' - 'Department Mathematik, Friedrich-Alexander-Universität Erlangen-Nürnberg, Cauerstraße 11, 91058 Erlangen, Germany' author: - 'Kurusch Ebrahimi-Fard' - Dominique Manchon - Johannes Singer bibliography: - 'library.bib' title: 'renormalisation of $q$-regularised multiple zeta values' --- Introduction {#sec:intro} ============ For ${\operatorname{Re}}(s)>1$ the *Riemann zeta function* is defined by $$\begin{aligned} \zeta(s):=\sum_{n\geq 1} \frac{1}{n^s}, \end{aligned}$$ which can be meromorphically continued to ${\mathbb C}$ with a simple pole in $s=1$. It is well known that for even integers $k \in 2 \mathbb{N}$ we have $$\begin{aligned} {\zeta}(k)= -\frac{(2\pi i)^k B_k}{2k!} \end{aligned}$$ and for $k\in {\mathbb N}_0$ $$\begin{aligned} \label{eq:mero1} {\zeta}(-k) = -\frac{B_{k+1}}{k+1}, \end{aligned}$$ where $B_k\in {\mathbb Q}$ are the Bernoulli numbers defined by the generating series $$\begin{aligned} \label{eq:Bernoulli} \frac{te^t}{e^t-1}= \sum_{k\geq 0} B_k\frac{t^k}{k!}. \end{aligned}$$ For ${\mathbf{s}}:=(s_1,\ldots,s_n)\in {\mathbb C}^n$ with $\sum_{i=1}^j {{\operatorname{Re}}(s_i)}>j$ ($j=1,\ldots,n$) *multiple zeta values* (MZVs) are defined by the nested series $$\begin{aligned} \label{eq:MZV} {\zeta}(s_1,\ldots,s_n):=\sum_{m_1>\cdots >m_n>0} \frac{1}{m_1^{s_1} \cdots m_n^{s_n}}, \end{aligned}$$ which are natural generalisations of the Riemann zeta function ([@Krattenthaler07]). We call $n$ the *depth* and $|{\mathbf{s}}|:=s_1+\cdots+s_n$ the *weight* of ${\mathbf{s}}$. Usually MZVs are studied at positive integers $k_1,\ldots,k_n$ with $k_1\geq 2$. Especially the ${\mathbb Q}$-vector space $$\begin{aligned} {\mathcal M}:=\langle \zeta({\mathbf{k}})\colon {\mathbf{k}}\in {\mathbb N}^n, k_1\geq 2, n\in {\mathbb N}_0 \rangle_{{\mathbb Q}}\end{aligned}$$ is of great interest ([@Brown12; @Zagier12]) since it is an algebra with two non-compatible products – the *shuffle* product and the *quasi-shuffle* product. The latter is induced by the defining series applying the product rule of power series. The shuffle product is an application of the integration by parts formula for iterated Chen integrals using an appropriate integral representation of MZVs in terms of the 1-forms $\omega_j^{(1)}:=\frac{dt_j}{1-t_j}$ and $\omega_j^{(0)}:=\frac{dt_j}{t_j}$. Indeed, using the standard notation, $|{\mathbf{k}}_{(j)}|:=k_1+\cdots+k_j$, $j\in\{1,\ldots,n\}$, one can show that can be written as $$\label{eq:MZVint} \zeta(k_1,\ldots,k_n)=\int_{0}^{1} \bigg( \prod_{j=1}^{|{\mathbf{k}}_{(1)}|-1} \omega_j^{(0)}\bigg) \omega_{|{\mathbf{k}}_{(1)}|}^{(1)} \cdots \bigg( \prod_{j=|{\mathbf{k}}_{(n-1)}|+1}^{|{\mathbf{k}}_{(n)}|-1} \omega_j^{(0)}\bigg) \omega_{|{\mathbf{k}}_{(n)}|}^{(1)} .$$ This so-called double-shuffle structure gives rise to a great and intriguing variety of linear relations among MZVs ([@Hoffman97; @Ihara06]), for instance the identity $\zeta(4)=4\zeta(3,1)$. The function ${\zeta}$ is meromorphic on ${\mathbb C}^n$ with the subvariety ${\mathcal S}_n$ of singularities, where $$\begin{aligned} \label{eq:meroMZV} {\mathcal S}_n:=\left\{(s_1,\ldots,s_n)\in {\mathbb C}^n\colon \begin{array}{l} s_1 = 1 \text{~~or~~} s_1+s_2 = 2,1,0,-2,-4, \ldots \text{~~or~~} \\ s_1+ \cdots +s_j \in {\mathbb Z}_{\leq j}~\text{for}~j=3,\ldots,n \end{array} \right\}.\end{aligned}$$ In [@Akiyama01a] it was shown that for $k_1,k_2\in {\mathbb N}$ with $k_1+k_2$ odd we have $$\begin{aligned} \label{eq:mero2} {\zeta}(-k_1,-k_2) = \frac{1}{2}\frac{B_{k_1+k_2+1}}{k_1+k_2+1}.\end{aligned}$$ In contrast to the dimension one case the meromorphic continuation of MZVs does not provide sufficient information for arbitrary negative integer arguments. This lack of data resulted in an interesting phenomenon related to the program commonly known as renormalisation of multiple zeta values at negative arguments. Using the fact that the quasi-shuffle product of MZVs can be abstracted to a quasi-shuffle Hopf algebra [@Hoffman00], which is connected and graded, the authors in [@Guo08] and [@Manchon10] proposed renormalisation procedures that would allow the extension of any MZV to – strictly – negative arguments while preserving the quasi-shuffle product. Both approaches are based on the renormalisation Hopf algebra discovered by Connes and Kreimer [@Connes00; @Connes01], which permits to formulate the so-called BPHZ subtraction method in terms of an algebraic Birkhoff decomposition of regularised Hopf algebra characters. See [@Manchon08; @Manchon10b; @Panzer12] for introductions and more details. However, the two approaches, [@Guo08] and [@Manchon10], propose different ways, both elaborated and sophisticated, of regularising MZVs at negative arguments. The resulting methods for constructing renormalised MZVs attribute different values to MZVs at negative arguments, while preserving the quasi-shuffle product. We remark that in [@Ebrahimi15] the authors presented a way to renormalise MZVs while preserving the shuffle product. Again, this approach is also based on the use of an algebraic Birkhoff decomposition of characters defined on a connected graded Hopf algebra, which is, however, different from the ones used in [@Guo08] and [@Manchon10]. It is an interesting question of how to relate the resulting zoo of different values for MZVs at negative arguments, based on renormalisation methods, which share the same algebraic workings in terms of an Hopf algebraic Birkhoff decomposition, but differ by applying distinct regularisation schemes. Through the works [@Castillo13a; @Castillo13b] it became apparent that certain $q$-analogues of MZVs may exhibit double shuffle relations as well as an intrinsic regularisation that permits an extension of the resulting $q$-MZVs to negative arguments. A natural question that arises from this observation is whether such a $q$-regularisation can be used to renormalise MZVs while preserving – one of – the original shuffle-type products. In [@Ebrahimi15] this was shown to hold with respect to the shuffle product, for a certain $q$-analogue of MZVs known as the Ohno–Okuda–Zudilin (OOZ) model [@Ohno12]. In the light of the fact that there exist a variety of different $q$-analogues of MZVs, one may wonder whether a systematic approach is feasible, that characterises models of $q$-MZVs under the aspect of whether they provide a proper $q$-regularisation, which would permit a renormalisation of MZVs while preserving algebraic properties. The agenda of this note proposes a step in the aforementioned direction. Starting from a particular $q$-analogue of MZVs, its intrinsic $q$-regularisation permits its extension to negative arguments. The resulting $q$-MZVs at negative arguments are still represented as series of nested sums, which, moreover, exhibit a natural quasi-shuffle product. This fact allows us to present a rather transparent and simple construction of renormalised MZVs preserving the quasi-shuffle product. Moreover, our approach is enhanced by providing a one-parameter extension of this $q$-analogue, which results in infinitely many extensions rolled into a single approach. Regarding renormalisation of $q$-MZVs we remark that Zhao’s approach in [@Zhao08] is rather different from the point of view just presented. It extended the approach of [@Guo08] to a particular $q$-analog of MZVs, without attributing any regularisation properties to the $q$-parameter itself. The aim of the note is as follows: - We provide a one-parameter family of extensions of MZVs to negative integer arguments. Our approach is based on applying the renormalisation procedure of Connes and Kreimer combined with minimal subtraction to a particular $q$-regularisation of MZVs. - We show that the extended MZVs satisfy the quasi-shuffle product of MZVs, and coincide with the meromorphic continuation of MZVs whenever it is defined. - We exemplify that besides pure rational extensions of MZVs also irrational or complex values can appear as renormalised MZVs for certain deformation parameters, although the values given by the meromorphic continuation are always rational. The note is organised as follows. In Section \[sec:qMZVs\] we introduce a $q$-analogue of MZVs ($q$-MZVs) which is the starting point of this work. Section \[sec:quasi\] is devoted to a review of the well known aspects concerning the quasi-shuffle Hopf algebra. The main results are stated in Section \[sec:main\] and the proofs are given in Section \[sec:proofmain\]. Finally, in Section \[sec:num\] we provide some explicit numerical examples. [**[Acknowledgements]{}**]{}: The first author is supported by the Ramón y Cajal research grant RYC-2010-06995 from the Spanish government. He acknowledges support from the Spanish government under project MTM2013-46553-C3-2-P. The second author is supported by Agence Nationale de la Recherche (projet CARMA). We thank the referee for a subtle report, which helped improving the paper. $q$-analogues of MZVs {#sec:qMZVs} ===================== In this section we introduce the $q$-analogue of MZVs which plays the central role in our construction. Let $q$ be a real number and $k_1,\ldots,k_n$ are positive integers. The Schlesinger–Zudilin model is defined by $$\begin{aligned} {\zeta}_q^{SZ}(k_1,\ldots,k_n):=\sum_{m_1>\cdots >m_n>0}\frac{q^{k_1 m_1+\cdots + k_n m_n}} {[m_1]_q^{k_1}\cdots [m_n]_q^{k_n}} \in {\mathbb Q}[\![q]\!]\end{aligned}$$ with the $q$-number $[m]_q:=\frac{1-q^m}{1-q}$. For $0<q<1$ this series of nested sums is convergent [@Schlesinger01; @Zudilin03; @Singer15]. And if $k_1\geq 2$ we obtain MZVs defined in in the limit $q\nearrow 1$. Our approach, however, is based on the following modification of the above model: $$\begin{aligned} \label{eq:defSZ} {\zeta}_q(k_1,\ldots,k_n):=\sum_{m_1>\cdots >m_n>0}\frac{q^{|k_1| m_1+\cdots + |k_n| m_n}} {[m_1]_q^{k_1}\cdots [m_n]_q^{k_n}} \in {\mathbb Q}[\![q]\!].\end{aligned}$$ As a result, this permits to consider the $q$-parameter as a natural regularisation, such that may be defined for integer arguments $k_1,\ldots,k_n \in {\mathbb Z}$. In the following we refer to this as the *regularised Schlesinger–Zudilin* model. Let $D:=\{s\in {\mathbb C}\colon {\operatorname{Re}}(s)>0\}$. For $t\in D$ we define a one-parameter family of *modified* $q$-MZVs by $$\begin{aligned} \label{eq:defSZmod} {\overline{\zeta}}_q^{(t),\ast}(k_1,\ldots,k_n):=\sum_{m_1>\cdots>m_n>0}{\frac{q^{(|k_1|m_1+\cdots+|k_n|m_n)t}}{(1-q^{m_1})^{k_1}\cdots (1-q^{m_n})^{k_n}}}. \end{aligned}$$ Since $0<q<1$ and $t\in D$, convergence of the previous series is always ensured for any $k_1,\ldots,k_n\in {\mathbb Z}$ with $k_1\neq 0$. Indeed, since $0<q<1$ we observe that $(1-q)^k\leq (1-q^m)^k\leq 1$ for any $m\in {\mathbb N}$ and $k\in {\mathbb N}_0$. Furthermore $\left| q^{(|k_1|m_1+\cdots+|k_n|m_n)t} \right|\leq q^{|k_1|m_1 {\operatorname{Re}}(t)}$ because $k_1\neq 0$ and $t\in D$. Therefore $$\begin{aligned} \left|{\frac{q^{(|k_1|m_1+\cdots+|k_n|m_n)t}}{(1-q^{m_1})^{k_1}\cdots (1-q^{m_n})^{k_n}}}\right| \leq C q^{|k_1|m_1{\operatorname{Re}}(t)}, \end{aligned}$$ where $C$ is a constant depending on $k_1,\ldots,k_n$. All in all we obtain $$\begin{aligned} \left|{\overline{\zeta}}_q^{(t),\ast}(k_1,\ldots,k_n)\right| & \leq C \sum_{m_1>\cdots>m_n>0} q^{|k_1|m_1{\operatorname{Re}}(t)} = C \sum_{m_1,\ldots,m_n>0} q^{|k_1|(m_1+\cdots+m_n){\operatorname{Re}}(t)}\\ & = C\left(\frac{q^{|k_1|{\operatorname{Re}}(t)}}{1-q^{|k_1|{\operatorname{Re}}(t)}} \right)^n <\infty. \end{aligned}$$ See also [@Zhao07 Prop. 2.2]. Note that $ (1-q)^{|{\mathbf{k}}|}{\overline{\zeta}}_q^{(1),\ast}(k_1,\ldots,k_n) = {\zeta}_q(k_1,\ldots,k_n)$. The quasi-shuffle product {#sec:quasi} ========================= Recall that for integers $a,b>1$ multiplying the two sums $$\label{stuf} \sum_{m>0}\frac{1}{m^a}\sum_{n>0}\frac{1}{n^b} = \sum_{m>n>0}\frac{1}{m^an^b} + \sum_{n>m>0}\frac{1}{n^bm^a} + \sum_{m>0}\frac{1}{m^{a+b}},$$ which is known as Nielsen’s reflexion formula $\zeta(a)\zeta(b)=\zeta(a,b)+\zeta(b,a)+\zeta(a+b)$. Defining the weight of products of MZVs as the sum of the weights of the factors, one notes the matching of weights on both sides. However, the length is not preserved. This generalises to arbitrary MZVs, and is commonly refereed to as quasi-shuffle product of MZVs. As an example we state the product $$\zeta(a,b)\zeta(c) = \zeta(c,a,b) + \zeta(a,b,c) + \zeta(a,c,b) + \zeta(a,b+c) + \zeta(a+c,b).$$ for $a,c>1$ and $b>0$. The key observation, which is not hard to verify, and which underlies our approach, is the fact that this simple product carries over to regularised Schlesinger–Zudilin $q$-MZVs, that is, the series of nested sums in satisfy the quasi-shuffle product if all arguments are either strictly positive or strictly negative integers. For example the corresponding Nielsen reflexion formula for the regularised Schlesinger–Zudilin model defined in for negative integers is $$\begin{aligned} \lefteqn{{\zeta}_q(-a){\zeta}_q(-b) = \sum_{m>0}\left(q^{mt}[m]_q \right)^a\sum_{n>0}\left(q^{nt}[n]_q \right)^b}\\ & = \sum_{m>n>0}\left(q^{mt}[m]_q \right)^a\left(q^{nt}[n]_q \right)^b+ \sum_{n>m>0}\left(q^{nt}[n]_q \right)^b\left(q^{mt}[m]_q \right)^a +\sum_{m>0}\left(q^{mt}[m]_q \right)^{a+b}\\ & = {\zeta}_q(-a,-b)+{\zeta}_q(-b,-a)+{\zeta}_q(-a-b)\end{aligned}$$ for $a,b\in {\mathbb N}$. One should note that the quasi-shuffle product is not preserved if we allow for integer arguments with mixed signs in . It is common to abstract the quasi-shuffle product using the words approach. Let $Y:=\{y_n\colon n\in {\mathbb Z}\}$ be an alphabet. The set $Y^\ast$ of words defines a free monoid, with the empty word denoted by ${\mathbf{1}}$. The linear span ${\mathbb Q}\langle Y \rangle$ of words from $Y^*$ becomes a commutative algebra when equipped with the *quasi-shuffle product* $m_\ast\colon{\mathbb Q}\langle Y \rangle \otimes {\mathbb Q}\langle Y\rangle \to {\mathbb Q}\langle Y \rangle$, $m_*(u \otimes v)=: u * v$ [@Hoffman12]. The latter is defined iteratively for any words $u,v\in Y^\ast$ and letters $y_n,y_m \in Y$ by (i) ${\mathbf{1}}\ast u := u \ast {\mathbf{1}}:= u$, (ii) $y_n u \ast y_m v := y_n(u\ast y_m v) + y_m(y_n u\ast v) + y_{n+m}(u\ast v)$ ($n,m\in {\mathbb Z}$), and extended to ${\mathbb Q}\langle Y \rangle$ by distributivity. It is well-known that for ${\mathfrak{h}}^+:={\mathbb Q}\oplus \bigoplus_{n\geq 2}y_n{\mathbb Q}\langle Y^+\rangle$ with $Y^+:=\{y_n\colon n\in {\mathbb N}\}$ the map $\zeta^\ast\colon ({\mathfrak{h}}^+,m_\ast)\to ({\mathbb R},\cdot)$ defined by ${\zeta}^\ast(y_{k_1}\cdots y_{k_n}):={\zeta}(k_1,\ldots,k_n)$ and ${\zeta}^\ast({\mathbf{1}}):=1$ is an algebra morphism. For ${\mathfrak{h}}^-:={\mathbb Q}\langle Y^-\rangle$ with $Y^-:=\{y_n\colon n\in {\mathbb Z}_{<0}\}$ we observe the following simple fact: \[lem:char\] The map ${\overline{\zeta}}_q^{(t)}\colon ({\mathfrak{h}}^-,m_\ast)\to ({\mathbb C}[\![q]\!],\cdot)$ defined by ${\overline{\zeta}}_q^{(t)}(y_{k_1}\cdots y_{k_n}):={\overline{\zeta}}_q^{(t),\ast}(k_1,\ldots,k_n)$ for any $k_1,\ldots,k_n\in {\mathbb Z}_{<0}$ and ${\overline{\zeta}}_q^{(t)}({\mathbf{1}}):=1$ is a morphism of algebras for any $t\in D$. Following [@Hoffman12] we introduce the deconcatenation coproduct $\Delta \colon {\mathbb Q}\langle Y \rangle \to {\mathbb Q}\langle Y \rangle \otimes {\mathbb Q}\langle Y \rangle$ $$\begin{aligned} \label{eq:coproduct} \Delta(w) = \sum_{uv=w}u\otimes v\end{aligned}$$ for $w \in Y^\ast$. Together with the quasi-shuffle product, $({\mathbb Q}\langle Y \rangle,m_\ast, \Delta)$ becomes a Hopf algebra. This construction directly applies to $Y^-$ and ${\mathfrak{h}}^-$ and therefore we observe: \[cor:Hopf\] The triple $({\mathfrak{h}}^{-}, m_\ast,\Delta)$ is a graded, connected Hopf algebra, in which the grading is given by the weight ${\operatorname{wt}}(y_{k_1}\cdots y_{k_n}):= |k_1|+\cdots + |k_n|$. The reader is referred to [@Manchon08; @Manchon10b] for details on connected graded Hopf algebras and related topics relevant to the forthcoming presentation. Renormalisation of MZVs {#sec:main} ======================= In order to extract renormalised MZVs from regularised expressions we use a theorem of Connes and Kreimer introduced in perturbative quantum field theory. See also [@Manchon08; @Ebrahimi07a] for related results. Regularisation refers to a procedure of introducing a (family of) formal parameter(s) that renders an otherwise divergent expression formally finite. This step displays some freedom as how to regularise MZVs at negative arguments, and it may alter algebraic properties, which makes it therefore a nontrivial part of the renormalisation program. \[theo:ConKre\] Let $({\mathcal{H}},m_{{\mathcal{H}}},\Delta)$ be a graded, connected Hopf algebra and ${\mathcal{A}}$ a commutative unital algebra equipped with a renormalisation scheme ${\mathcal{A}}={\mathcal{A}}_{-}\oplus {\mathcal{A}}_{+}$ and the corresponding idempotent Rota–Baxter operator $\pi$, where ${\mathcal{A}}_{-}=\pi({\mathcal{A}})$ and ${\mathcal{A}}_{+}=({\operatorname{Id}}-\pi)({\mathcal{A}})$. Further let $\psi\colon {\mathcal{H}}\to {\mathcal{A}}$ be a Hopf algebra character, i.e., a multiplicative linear map from ${\mathcal{H}}$ to ${\mathcal{A}}$. Then the character $\psi$ admits a unique decomposition $$\begin{aligned} \label{eq:birk} \psi=\psi_{-}^{\star{(-1)}}\star \psi_{+} \end{aligned}$$ called *algebraic Birkhoff decomposition*, in which $\psi_{-}\colon {\mathcal{H}}\to {\mathbb Q}\oplus {\mathcal{A}}_{-}$ and $\psi_{+}\colon {\mathcal{H}}\to {\mathcal{A}}_{+}$ are characters. The product on the right hand side of is the convolution product defined on the vector space $L({\mathcal{H}}, {\mathcal{A}})$ of linear maps from the ${\mathcal{H}}$ to ${\mathcal{A}}$. Recall that the vector space $L({\mathcal{H}},{\mathcal{A}})$ together with the convolution product $\phi \star \psi := m_{{\mathcal{A}}} \circ (\phi \otimes \psi) \circ \Delta : {\mathcal{H}}\to {\mathcal{A}}$, where $\phi,\psi \in L({\mathcal{H}},{\mathcal{A}})$, is an unital associative algebra. The set of characters is denoted by $G_{\mathcal{A}}$ and forms a (pro-unipotent) group for the convolution product with (pro-nilpotent) Lie algebra $g_{{\mathcal{A}}}$ of infinitesimal characters. The latter are linear maps $\xi \in L({\mathcal{H}}, {\mathcal{A}})$ such that for elements $x, y \in {\mathcal{H}}$, both different from $\mathbf{1}$, $\xi(xy) = 0$. The exponential map $\exp^{\star}$ restricts to a bijection between $ g_{{\mathcal{A}}}$ and $G_{{\mathcal{A}}}$. The inverse of a character $\psi \in G_{{\mathcal{A}}}$ is given by composition with the Hopf algebra antipode $S: {\mathcal{H}}\to {\mathcal{H}}$, e.g., $\psi_{-}^{\star{(-1)}} = \psi_- \circ S$. As a regularisation scheme we choose the commutative algebra ${\mathcal{A}}:={\mathbb Q}[z^{-1},z]\!]$ with ${\mathcal{A}}= {\mathcal{A}}_{-}\oplus {\mathcal{A}}_{+},$ where ${\mathcal{A}}_{-}:=z^{-1}{\mathbb Q}[z^{-1}]$ and ${\mathcal{A}}_{+}:={\mathbb Q}[\![z]\!]$. On ${\mathcal{A}}$ we define the corresponding projector $\pi: {\mathcal{A}}\to {\mathcal{A}}_{-}$ by $$\begin{aligned} \pi\left(\sum_{n=-k}^\infty a_n z^n \right):= \sum_{n=-k}^{-1}a_n z^n\end{aligned}$$ with the common convention that the sum over the empty set is zero. Then $\pi$ and ${\operatorname{Id}}-\pi: {\mathcal{A}}\to {\mathcal{A}}_{+}$ are Rota–Baxter operators of weight $-1$ (see [@Connes00; @Ebrahimi02; @Ebrahimi07]). For the renormalisation of MZVs we perform two main steps. Firstly, we construct a regularised character $\psi\colon {\mathfrak{h}}^- \to {\mathcal{A}}$ which means that we have to deform the divergent MZVs to a meromorphic function in the regularisation parameter $z$. Secondly, we apply an appropriate subtraction scheme. The latter is naturally given by Equation , which essentially relies on the convolution product induced by the Hopf algebra $({\mathfrak{h}}^{-},m_\ast,\Delta)$. The maps $\psi_+$ and $\psi_-$ of Equation are recursively given by $$\begin{aligned} \psi_-(x) & =-\pi \left(\psi(x)+\sum_{(x)}\psi_{-}(x')\psi(x'') \right), \label{eq:psim} \\ \psi_+(x) & =({\operatorname{Id}}-\pi) \left(\psi(x)+\sum_{(x)}\psi_{-}(x')\psi(x'') \right),\label{eq:psip}\end{aligned}$$ for $x \in {\mathcal{H}}$, ${\operatorname{wt}}(x) >0$, and $\psi_\pm \in G_{\mathcal{A}}$. Note that we used Sweedler’s notation for the reduced coproduct $\Delta'(x):=\sum_{(x)}x'\otimes x'':=\Delta(x)-{\mathbf{1}}\otimes x - x\otimes {\mathbf{1}}$. The map $\psi^{(t)}\colon ({\mathfrak{h}}^-,m_\ast) \to {\mathbb C}[z^{-1},z]\!]$ defined by $$\begin{aligned} \label{eq:char} \psi^{(t)}(y_{-k_1}\cdots y_{-k_n})(z):= \frac{(-1)^{k_1+\cdots+k_n}}{z^{k_1+\cdots+ k_n}} {\overline{\zeta}}_{e^z}^{(t)}(y_{-k_1}\cdots y_{-k_n})\end{aligned}$$ for $k_1,\ldots,k_n\in {\mathbb N}$ is an algebra morphism (see Lemma \[lem:defexp\] below). Applying Theorem \[theo:ConKre\] we define *renormalised MZVs* ${\zeta}_+^{(t)}$ by $$\begin{aligned} {\zeta}_+^{(t)}(-k_1,\ldots,-k_n):=\lim_{z\to 0} \psi_+^{(t)}(y_{-k_1}\cdots y_{-k_n})\end{aligned}$$ for any $k_1,\ldots,k_n\in {\mathbb N}$. Now we present the main theorem of the paper which will be proven in the next section. \[theo:main\] Let $t\in D$. a) The renormalised MZVs ${\zeta}_+^{(t)}$ verify the meromorphic continuation of MZVs , i.e., for any ${\mathbf{k}}\in ({\mathbb Z}_{<0})^n\setminus {\mathcal S}_n$ we have ${\zeta}_+^{(t)}({\mathbf{k}})={\zeta}({\mathbf{k}})\in {\mathbb Q}$. b) The renormalised MZVs ${\zeta}_+^{(t)}$ satisfy the quasi-shuffle product. c) The renormalised MZVs ${\zeta}_+^{(t)}$ are rational functions in $t$ over ${\mathbb Q}$ without singularities in $D$. As a consequence we obtain: \[coro:main\] a) Let $t\in D\cap{\mathbb Q}$ then ${\zeta}_+^{(t)}({\mathbf{k}})\in {\mathbb Q}$ for any ${\mathbf{k}}\in ({\mathbb Z}_{<0})^n$, $n\in {\mathbb N}$. b) Let $t\in D\cap {\mathbb R}$ be transcendental over ${\mathbb Q}$. Then there exists a ${\mathbf{k}}\in ({\mathbb Z}_{<0})^n$ such that ${\zeta}_+^{(t)}({\mathbf{k}})\in {\mathbb R}\setminus {\mathbb Q}$. The first claim follows from Theorem \[theo:main\] c). For the proof of b) see Section \[sec:num\]. Proof of Theorem \[theo:main\] {#sec:proofmain} ============================== In this section we implicitly use the following corollary of the fundamental theorem of algebra. Let $p(t)\in {\mathbb C}[t]$ be a polynomial over ${\mathbb C}$ with $p(n)=0$ for countably infinitely many $n\in {\mathbb N}$. Then $p=0$. Therefore some proofs in the subsequent paragraph – concerning polynomial identities in $t$ – are provided for $t\in {\mathbb N}$ which then can be extended to $t\in D$ using this argument. \[lem:defexp\] Let $k_1,\ldots,k_n\in {\mathbb N}$. Explicitly we have $$\begin{aligned} \psi^{(t)}(y_{-k_1}\cdots y_{-k_n})(z) = \sum_{m_1,\ldots,m_n\geq 0}\frac{B_{m_1}}{m_1!} \cdots \frac{B_{m_n}}{m_n!} C^{k_1,\ldots,k_n}_{m_1,\ldots,m_n}(t) z^{m_1+\cdots+m_n-(k_1+\cdots+k_n)-n} \end{aligned}$$ with $$\begin{aligned} \label{eq:contC} C_{m_1,\ldots,m_n}^{k_1,\ldots,k_n}(t):=\sum_{l_1=0}^{k_1} \cdots \sum_{l_n=0}^{k_n} \prod_{j=1}^n\binom{k_j}{l_j} (-1)^{l_j+k_j+1}(l_1+k_1t+\cdots + l_j+k_jt)^{m_j-1}. \end{aligned}$$ Furthermore $\psi^{(t)}\colon ({\mathfrak{h}}^-,m_\ast) \to {\mathbb C}[z^{-1},z]\!]$ is an algebra morphism. In Equation we defined $\psi^{(t)}$ by the following composition of maps $$\begin{aligned} \begin{tabular}{ccccccc} $({\mathfrak{h}}^-,m_\ast)$ & $\longrightarrow$ & $({\mathbb C}[\![q]\!],\cdot)$ &$\longrightarrow$ & $({\mathbb C}[z^{-1},z]\!],\cdot)$ & $\longrightarrow$ & $({\mathbb C}[z^{-1},z]\!],\cdot)$ \\ $y_{k_1}\cdots y_{k_n}$ & $\longmapsto$ & ${\overline{\zeta}}_q^{(t)}(y_{k_1} \cdots y_{k_n})$ &$\longmapsto$ & ${\overline{\zeta}}_{e^z}^{(t)}(y_{k_1}\cdots y_{k_n})$ & $\longmapsto$ & $(-z)^{|{\mathbf{k}}|}{\overline{\zeta}}_{e^z}^{(t)}(y_{k_1} \cdots y_{k_n})$, \end{tabular}\end{aligned}$$ where $k_1,\ldots,k_n\in {\mathbb Z}_{<0}$. By Lemma \[lem:char\] the first map is an algebra morphism and the substitution map $q\mapsto e^z$ preserves this property. Since the quasi-shuffle product $m_\ast$ preserves the weight, the multiplication map is an algebra morphism, too. Hence, $\psi^{(t)}$ is a character. Performing the composition of maps we explicitly obtain for $k_1,\ldots,k_n \in {\mathbb N}$ Now we rewrite the last expression by substituting $e^z$ for $q$ such that where we used . Multiplying the last expression by $(-z)^{-|{\mathbf{k}}|}$ yields which concludes the proof. [[ The particular choice of the map $(-\log(q))^{|{\mathbf{k}}|}{\overline{\zeta}}_q^{(t),\ast}({\mathbf{k}})$ in the definition of the character $\psi^{(t)}$ is necessary in order to assure that the renormalised MZVs verify the meromorphic continuation. Indeed, for the – at first glance – more natural looking choice $(1-q)^{|{\mathbf{k}}|}{\overline{\zeta}}_q^{(t),\ast}({\mathbf{k}})$ in the definition of the character $\psi^{(t)}$ we observe a contradiction to the meromorphic continuation, e.g., we would obtain $\zeta_+^{(t)}(-1)=-\frac{1}{12}\frac{t^2+t-1}{t^2+t}\neq -\frac{1}{12}$ for any $t\in D$. ]{}]{} \[lem:meroconst\] Let $k\in {\mathbb N}$. Then In the case $m=0$ the claim is exactly Equation . For $m\geq 1$ we observe which concludes the proof. \[lem:mero1\] For $k\in {\mathbb N}$ we have $$\begin{aligned} \psi^{(t)}_-(y_{-k})(z) = -C_0^k(t) z^{-k-1} \hspace{0.5cm} \text{~and~} \hspace{0.5cm} \psi^{(t)}_+(y_{-k})(z) = -\frac{B_{k+1}}{k+1} + O_t(z), \end{aligned}$$ where $O_t(z)$ is the standard Landau notation in which the subindex $t$ denotes the $t$-dependence of the coefficients of higher order terms. Since $y_{-k}$ is primitive in the Hopf algebra $({\mathfrak{h}}^-,m_\ast,\Delta)$, Equations and imply respectively $$\begin{aligned} \psi^{(t)}_-(y_{-k})(z) = -\pi \psi^{(t)}(y_{-k})(z) \hspace{0.5cm}\text{~and~}\hspace{0.5cm} \psi^{(t)}_+(y_{-k})(z) = ({\operatorname{Id}}-\pi)\psi^{(t)}(y_{-k})(z). \end{aligned}$$ From Lemma \[lem:meroconst\] we deduce $$\begin{aligned} \psi^{(t)}_-(y_{-k})(z) &= -\pi\left( \sum_{m\geq 0}\frac{B_m}{m!}C^k_m(t) z^{m-k-1} \right) = -\sum_{m= 0}^k\frac{B_m}{m!}C^k_m(t) z^{m-k-1} = -C_0^k(t) z^{-k-1}\end{aligned}$$ and $$\begin{aligned} \psi^{(t)}_+(y_{-k})(z) &= ({\operatorname{Id}}-\pi)\left(\sum_{m\geq 0}\frac{B_m}{m!}C^k_m(t) z^{m-k-1} \right) = \sum_{m\geq k+1}\frac{B_m}{m!}C^k_m(t) z^{m-k-1} =- \frac{B_{k+1}}{k+1} + O_t(z). \end{aligned}$$ \[lem:mero2\] For $k_1,k_2 \in {\mathbb N}$ with $k_1+k_2$ odd we have $$\begin{aligned} \psi_+^{(t)}(y_{-k_1}y_{-k_2})(z) = \frac{1}{2} \frac{B_{k_1 + k_2 + 1}}{k_1 + k_2 + 1} + O_t(z). \end{aligned}$$ First we remark that $$\begin{aligned} \Delta(y_{-k_1}y_{-k_2}) = {\mathbf{1}}\otimes y_{-k_1}y_{-k_2} +y_{-k_1}\otimes y_{-k_2} + y_{-k_1}y_{-k_2} \otimes {\mathbf{1}}\end{aligned}$$ and therefore Equation implies $$\begin{aligned} \psi^{(t)}_+(y_{-k_1}y_{-k_2}) = ({\operatorname{Id}}-\pi)\left( \psi^{(t)}(y_{-k_1}y_{-k_2})+ \psi^{(t)}_-(y_{-k_1})\psi^{(t)}(y_{-k_2}) \right). \end{aligned}$$ Since $k_1+k_2+2$ is odd and $B_s=0$ for odd $s\geq 2$ we obtain with $B_1=\frac{1}{2}$ $$\begin{aligned} ({\operatorname{Id}}-\pi)\psi^{(t)}(y_{-k_1}y_{-k_2})(z) &= \sum_{m_1+m_2=k_1+k_2+2} \frac{B_{m_1}}{m_1!}\frac{B_{m_2}}{m_2!} C_{m_1,m_1}^{k_1,k_2}(t) + O_t(z) \\ & = \frac{1}{2}\frac{B_{k_1+k_2+1}}{(k_1+k_2+1)!} \left( C^{k_1,k_2}_{k_1+k_2+1,1}(t)+C^{k_1,k_2}_{1,k_1+k_2+1}(t) \right) +O_t(z). \end{aligned}$$ On the one hand $\sum_{l_2=0}^{k_2}\binom{k_2}{l_2}(-1)^{l_2}=0$ and therefore $$\begin{aligned} C_{k_1+k_2+1,1}^{k_1+k_2}(t)=\sum_{l_1=0}^{k_1} \sum_{l_2=0}^{k_2}\binom{k_1}{l_1} \binom{k_2}{l_2}(-1)^{l_1+l_2+k_1+k_2}(l_1+k_1t)^{k_1+k_2}=0. \end{aligned}$$ On the other hand we get and therefore $$\begin{aligned} ({\operatorname{Id}}-\pi)\psi^{(t)}(y_{-k_1}y_{-k_2})(z) = \frac{1}{2}\frac{B_{k_1+k_2+1}}{k_1+k_2+1} +O_t(z).\end{aligned}$$ Moreover, since $B_{k_1+k_2+2}=0$ we observe with Lemma \[lem:defexp\] and \[lem:mero1\] $$\begin{aligned} ({\operatorname{Id}}-\pi)\left( \psi_-^{(t)}(y_{-k_1})(z)\psi^{(t)}(y_{-k_2})(z)\right) & = -({\operatorname{Id}}-\pi)\left( C^{k_1}_0(t) z^{-k_1-1} \cdot \sum_{m\geq 0} \frac{B_m}{m!}C_m^{k_2}(t)z^{m-k_2-1}\right)\\ &=O_t(z). \end{aligned}$$ To summarise, we have shown that $$\begin{aligned} \psi_+^{(t)}(y_{-k_1}y_{-k_2})(z) &=({\operatorname{Id}}-\pi)\left( \psi^{(t)}(y_{-k_1}y_{-k_2})+ \psi^{(t)}_-(y_{-k_1})\psi^{(t)}(y_{-k_2}) \right)(z)\\ &= \frac{1}{2}\frac{B_{k_1+k_2+1}}{k_1+k_2+1} + O_t(z), \end{aligned}$$ which concludes the proof. From Lemma \[lem:mero1\] we deduce for $k\in {\mathbb N}$ $$\begin{aligned} {\zeta}_+^{(t)}(-k) = \lim_{z\to 0}\psi_+^{(t)}(y_{-k})(z) = -\frac{B_{k+1}}{k+1}, \end{aligned}$$ which coincides with and for $k_1,k_2\in {\mathbb N}$ with $k_1+k_2$ odd we observe from Lemma \[lem:mero2\] that $$\begin{aligned} {\zeta}_+^{(t)}(-k_1,-k_1) = \lim_{z\to 0}\psi_+^{(t)}(y_{-k_1}y_{-k_2})(z) = \frac{1}{2} \frac{B_{k_1+k_2+1}}{k_1+k_2+1}, \end{aligned}$$ which is consistent with . For length greater than three the meromorphic continuation does not provide any information for negative integer arguments since $({\mathbb Z}_{<0})^n\subseteq {\mathcal S}_n$ for $n \geq 3$ (see ). Therefore a) is established. Next we prove b). As shown in Lemma \[lem:defexp\] $\psi^{(t)}$ is an algebra morphism with respect to the quasi-shuffle product $m_\ast$. Therefore Theorem \[theo:ConKre\] implies that $\psi_+^{(t)}$ is also an algebra morphism and hence is ${\zeta}_+^{(t)}$. For c) we remark that Equation of Lemma \[lem:defexp\] shows that $\psi^{(t)}$ is a meromorphic function in $z$, whose coefficients are rational functions in $t$ over ${\mathbb Q}$. Equations and indicate that the power series $\psi_+^{(t)}$ is obtained from $\psi^{(t)}$ by subtractions and projections. Therefore ${\zeta}_+^{(t)}$ is – as the constant coefficient of the power series $\psi_+^{(t)}$ – a rational function in $t$ over ${\mathbb Q}$. Numerical examples {#sec:num} ================== In this subsection we provide some explicit numerical examples of the renormalisation process introduced in the previous sections. First we provide an explicit example of how to compute renormalised MZVs. Let us calculate the renormalised MZV ${\zeta}_+^{(1)}(-1,-3)$. Using we observe $$\begin{aligned} \Delta(y_{-1}y_{-3}) = {\mathbf{1}}\otimes y_{-1}y_{-3}+ y_{-1}\otimes y_{-3}+ y_{-1}y_{-3}\otimes {\mathbf{1}}. \end{aligned}$$ Consequently, formula implies $$\begin{aligned} \psi_+^{(1)}(z)=({\operatorname{Id}}-\pi) \left(\psi^{(1)}(y_{-1}y_{-3})(z)+\psi_{-}^{(1)}(y_{-1})(z)\psi^{(1)}(y_{-3})(z) \right).\end{aligned}$$ Using $$\begin{aligned} \psi^{(1)}(y_{-1})(z)&= \frac{1}{2}{z}^{-2}-\frac{1}{12}+\frac{7}{720}{z}^{2}- \frac{31}{30240}{z}^{4}+ O({z}^{5}), \\ \psi^{(1)}(y_{-3})(z)& = \frac{1}{60}{z}^{-4}+\frac{1}{120}-\frac{41}{1008}{z}^{2}+ \frac{2203}{28800}{z}^{4}+O({z}^{5}),\\ \psi^{(1)}(y_{-1}y_{-3})(z) &= \frac{3}{560}{z}^{-6}+\frac{1}{560}{z}^{-5}+\frac{47}{11200}{z}^{-2}-\frac{5377}{282240}+\frac{1}{84}z+O({z}^{2}), \end{aligned}$$ we obtain $$\begin{aligned} ({\operatorname{Id}}-\pi)\psi^{(1)}(y_{-1}y_{-3})(z) = -\frac{5377}{282240}+\frac{1}{84}z+O({z}^{2}). \end{aligned}$$ Since $y_{-1}$ is a primitive element of the Hopf algebra $({\mathfrak{h}}^{-}, m_\ast,\Delta)$ we have $\psi_-^{(1)}(y_{-1})(z)= -\frac{1}{2}{z}^{-2}$ by Equation and therefore $$\begin{aligned} ({\operatorname{Id}}-\pi) \left(\psi_{-}^{(1)}(y_{-1})(z)\psi^{(1)}(y_{-3})(z) \right) = \frac{41}{2016} +O(z^2).\end{aligned}$$ All in all we get $$\begin{aligned} \psi_+^{(1)}(z)=({\operatorname{Id}}-\pi) \left(\psi^{(1)}(y_{-1}y_{-3})(z)+\psi_{-}^{(1)}(y_{-1})(z)\psi^{(1)}(y_{-3})(z) \right) = \frac{121}{94080}+\frac{1}{84}z+O(z^2),\end{aligned}$$ which results in $$\begin{aligned} {\zeta}_+^{(1)}(-1,-3)=\lim_{z\to 0} \psi_+^{(1)}(z) = \frac{121}{94080}. \end{aligned}$$ In Table \[table1\] we list the renormalised MZVs in the case $t=1$ for depth two. For depth one the renormalised MZVs are always rational as well as for $k_1,k_2\in {\mathbb N}$ with $k_1+k_2$ odd, due to Theorem \[theo:main\] a). Because of the quasi-shuffle relation $$\begin{aligned} {\zeta}_+^{(t)}(-k){\zeta}_+^{(t)}(-k) = 2 {\zeta}_+^{(t)}(-k,-k) + {\zeta}_+^{(t)}(-2k)\end{aligned}$$ for $k\in {\mathbb N}$ the diagonal entries ${\zeta}_+^{(t)}(-k,-k)$ are also always rational and do not depend on the parameter $t$. The first case for which we obtain a non-constant rational function in $t$ over ${\mathbb Q}$ is the case ${\zeta}_+^{(t)}(-1,-3)$. The quasi-shuffle product implies $$\begin{aligned} {\zeta}_+^{(t)}(-1){\zeta}_+^{(t)}(-3) = {\zeta}_+^{(t)}(-1,-3)+{\zeta}_+^{(t)}(-3,-1) + {\zeta}_+^{(t)}(-4). \end{aligned}$$ Therefore a priori ${\zeta}_+^{(t)}(-1,-3)$ and ${\zeta}_+^{(t)}(-3,-1)$ are not explicitly given by that relation. They only have to satisfy $ {\zeta}_+^{(t)}(-1,-3)+{\zeta}_+^{(t)}(-3,-1) = {\zeta}_+^{(t)}(-1){\zeta}_+^{(t)}(-3)$. Using and we find This example also proves the second claim of Corollary \[coro:main\]. Indeed, since not all renormalised MZVs ${\zeta}_+^{(t)}$ are constant as a rational function in $t$ over ${\mathbb Q}$ the choice of a transcendental $t\in D$ leads to irrational values for ${\zeta}_+^{(t)}$. It is easily seen that also complex numbers can appear as renormalised MZVs. Conclusion {#conclusion .unnumbered} ---------- Extensions of MZVs to strictly negative arguments by renormalisation, i.e., by using subtraction methods on properly regularised nested sums, implied by a Hopf algebraic Birkhoff decomposition, were introduced in [@Guo08] and [@Manchon10]. The two approaches differ in the regularisations used to obtain formally well-defined expressions. This results in different values associated to certain MZVs at negative arguments, while both approaches preserve the quasi-shuffle product. In this paper, we have shown how to renormalise MZVs using minimal subtraction in an intrinsically regularised $q$-analogue of MZVs. The subtraction method is implied by a Hopf algebraic Birkhoff decomposition, and the quasi-shuffle product is preserved. Regarding the parameter $t$ of ${\zeta}_+^{(t)}$, it is natural to ask for the relation between different sets of renormalised MZVs corresponding to different parameters. Let $t_1,t_2 \in D$. Recent work shows that we have ${\zeta}_+^{(t_1)} = \alpha_{t_1,t_2} \star {\zeta}_+^{(t_2)}$, where $\alpha_{t_1,t_2}$ belongs to the *renormalisation group of MZVs* [@Ebrahimi15c], which is a particular subgroup of the group of characters of the quasi-shuffle Hopf algebra. Similarly one can compare the renormalised MZVs in this work with those values obtained in the literature [@Guo08; @Manchon10]. The reader is referred to [@Ebrahimi15c] for more details. Moreover, in the light of the results in [@Ebrahimi15] one may wonder whether the approach presented here can be further developed by characterising different $q$-analogues of MZVs as distinct $q$-regularisation methods for MZVs. A possible prescription would be to start by replacing the summation variables in the nested series by $q$-integers, and then to introduce in the nominator a particular polynomial in the $q$-regularisation parameter. The choice of the latter is crucial since it determines the algebraic as well as analytic properties of the resulting $q$-MZVs. Several $q$-analogues of MZVs have been described in the literature, and it would be rather interesting to understand them from the point of view of extending MZVs to negative arguments.
{ "pile_set_name": "ArXiv" }
--- author: - | par Daniel Barlet[^1],\ avec un appendice de Claude Sabbah[^2] date: '06/09/05' title: | Interaction de strates consécutives pour les cycles évanescents III :\ Le cas de la valeur propre 1. --- Abstract. {#abstract. .unnumbered} ========= This text is a study of the missing case in our article \[B.91\], that is to say the eigenvalue 1 case. Of course this is a more involved situation because the existence of the smooth stratum for the hypersurface  $\lbrace f = 0 \rbrace $  forces to consider three strata for the nearby cycles. And we already know that the smooth stratum is always “tangled” if it is not alone (see \[B.84b\] and the introduction of \[B.03\]). The new phenomenon is the role played here by a “new” cohomology group, denote by  $ H^n_{c\,\cap\,S}(F)_{=1} $, of the Milnor’s fiber of  $f$  at the origin. It has the same dimension as  $H^n(F)_{=1}$   and  $H^n_c(F)_{=1}$[^3] and it leads to a non trivial factorization of the canonical map  $ can : H^n_c(F)_{=1} \rightarrow H^n(F)_{=1}$, and to a monodromic isomorphism of variation  $ var : H^n_{c\,\cap\,S}(F)_{=1} \rightarrow H^n_c(F)_{=1}$. It gives a canonical hermitian form $$\mathcal{H} : H^n_{c\,\cap\,S}(F)_{=1}\times H^n(F)_{=1} \rightarrow \mathbb{C}$$ which is non degenerate[^4]. This generalizes the case of an isolated singularity for the eigenvalue 1 (see \[B.90\] and \[B.97\] ). The “overtangling” phenomenon for strata associated to the eigenvalue 1 implies the existence of triple poles at negative integers (with big enough absolute value) for the meromorphic continuation of the distribution  $ \int_X \vert f \vert^{2\lambda} \Box $   for functions  $f$  having semi-simple local monodromies at each singular point of  $ \lbrace f = 0 \rbrace .$ [**AMS Classification 2000**]{} : 32-S-25, 32-S-40, 32-S-50. [**Key Words**]{} : Hypersurface, Non isolated Singularity, Vanishing Cycles,Tangling of Strata. Résumé. {#résumé. .unnumbered} ======== Ce texte étudie le cas manquant de notre article \[B.91\], à savoir le cas de la valeur propre 1. C’est évidemment un cas plus compliqué que celui d’une valeur propre $\not= 1$  puisque la présence de la strate des points lisses de l’hypersurface  $\lbrace f = 0 \rbrace $  oblige à considérer trois strates pour les cycles proches. Et l’on sait déjà que cette strate lisse est toujours “emmêlée” en présence d’une autre strate (voir \[B.84b\] et l’introduction de \[B.03\]). Le phénomène nouveau est le rôle joué ici par un “nouveau” groupe de cohomologie, noté  $ H^n_{c\,\cap\,S}(F)_{=1} $, de la fibre de Milnor de  $f$  à l’origine, qui est de même dimension que  $H^n(F)_{=1}$  et  $H^n_c(F)_{=1}$[^5] et qui donne lieu à une factorisation non triviale de l’application canonique  $ can : H^n_c(F)_{=1} \rightarrow H^n(F)_{=1}$, et à un isomorphisme monodromique de variation  $ var : H^n_{c\,\cap\,S}(F)_{=1} \rightarrow H^n_c(F)_{=1}$. On en déduit une forme hermitienne canonique $$\mathcal{H} : H^n_{c\,\cap\,S}(F)_{=1}\times H^n(F)_{=1} \rightarrow \mathbb{C}$$ qui est non dégénérée[^6]. Ceci généralise le cas d’une singularité isolée pour la valeur propre 1 (voir \[B.90\] et \[B.97\] ). Le phénomène de “suremmêlement” de strates pour la valeur 1 donne l’existence de pôles triples aux entiers négatifs (assez grands en valeur absolue) pour le prolongement méromorphe de la distribution  $ \int_X \vert f \vert^{2\lambda} \Box $  en présence de monodromies locales semi-simples pour  $f$  en chaque point du lieu singulier de  $ \lbrace f = 0 \rbrace .$ [**AMS Classification 2000**]{}: 32-S-25, 32-S-40, 32-S-50. [**Mots Clefs**]{} : Hypersurface, Singularité non isolée, Cycles Évanescents, Emmêlement de Strates. Introduction. {#introduction. .unnumbered} ============= Le but du présent article est de compléter l’étude du phénomène\ d’emmêlement de strates (consécutives) pour les cycles évanescents, faite dans \[B.91\][^7] pour une valeur propre de la monodromie différente de 1 par l’étude du cas de la valeur propre 1 de la monodromie. La situation que nous considérons ici est le cas d’un germe non constant de fonction holomorphe  $ \tilde{f} : (\mathbb{C}^{n+1}, 0) \rightarrow (\mathbb{C}, 0)$  vérifiant la propriété suivante : Il existe un germe de courbe  $(S, 0)$  à l’origine de  $\mathbb{C}^{n+1}$  tel qu’en chaque point  $x$  en dehors de cette courbe la monodromie locale de  $f$  agissant sur la cohomologie réduite de la fibre de Milnor de  $f$  en  $x$  ne présente pas la valeur propre 1. Bien sur, le germe de courbe  $(S, 0)$  est “à priori” contenu dans le lieu singulier de l’hypersuface  $f^{-1}(0)$  mais ce lieu singulier sera en général de dimension plus grande que 1. De même que pour le cas d’une valeur propre  $\not= 1$  on rencontre cette situation de la fa[ç]{}on suivante : On considère un germe non constant de fonction holomorphe $$\tilde{F} : (\mathbb{C}^{N}, 0) \rightarrow (\mathbb{C}, 0)$$ et on denote par  $\Sigma_0 \subset \Sigma_1 $  les deux plus grosses strates du lieu singulier de  $ \tilde{F}^{-1}(0) $  le long desquelles la valeur propre 1 de la monodromie apparait (sur la cohomologie [^8]). Supposons que la codimension de  $\Sigma_0$  soit égale à  $n+1$  et que celle de  $\Sigma_1$  soit égale à  $n$[^9]. Effectuons une section plane transverse en un point générique de  $\Sigma_0 $  par un plan  $P$  de dimension  $n+1$. Un tel plan “générique” est non caractéristique pour le  $D-$module correspondant au faisceau pervers  $\psi_{\tilde{F, 1}}$  et la restriction  $ \tilde{f} : = \tilde{F}\vert_P $  vérifie nos hypothèses et décrit la situation locale pour  $\psi_{\tilde{F, 1}}$  dans un voisinage du point générique de  $\Sigma_0 $  que l’on a considéré[^10]. Les hypothèses standards pour la valeur propre 1. ------------------------------------------------- Soit  $X$  un voisinage ouvert de Stein connexe de l’origine dans  $\mathbb{C}^{n+1}$  et supposons que  $ n \geq 2 $[^11]. Soit  $ f : X \rightarrow \mathbb{C}$  une fonction holomorphe non constante. Notons pour  $p \geq 0$  par  $H^p(0)$  le faisceau constructible sur  $ Y : = f^{-1}(0)$  donné par le sous-faisceau spectral de la monodromie de  $f$  pour la valeur propre 1, agissant sur le  $p-$ième faisceau de cohomologie du complexe  $\psi_f$  des cycles évanescents. Donc le germe en chaque point  $y \in Y$  de  $H^p(0)$  est le  sous-espace spectral pour la valeur propre 1 de la monodromie agissant sur le $p-$ième groupe de la cohomologie de la fibre de Milnor de  $f$  en  $y$. Nous ferons les hypothèses suivantes : - [(i)]{} Le faisceau  $H^n(0)$  est concentré à l’origine. - [(ii)]{} Le faisceau  $H^{n-1}(0) $  est concentré sur une courbe  $S$  admettant l’origine pour unique point singulier. De plus, chaque composante irréductible de  $S$  contient l’origine et est un disque topologique. - [(iii)]{} La restriction de  $H^{n-1}(0) $  à  $S^* : = S \setminus \lbrace 0 \rbrace$  est un système local. - [(iv)]{} Pour tout  $ i \geq 2$  on suppose que  $ H^{n-i}(0) \equiv 0 .$ ### Remarques. - Seule la condition (iv) est réellement restrictive dans notre situation locale d’un germe à l’origine de  $\mathbb{C}^{n+1}$  en raison de la perversité de  $ \psi_{f,1}.$ - Bien sur, ces hypothèses sont satisfaites (quitte à se restreindre à un voisinage ouvert de l’origine assez petit) dès que le lieu singulier est de dimension inférieur ou égal à 1. Pour  $n=2$   c’est le cas dès que l’on considère un germe non constant de fonction holomorphe qui est réduit. - Pour  $n = 3$  ces hypothèses seront satisfaites (quitte à nouveau à se restreindre ...) pour un germe réduit dès que les singularités transverses aux composantes irréductibles de dimension 2 du lieu singulier sont des courbes réduites et irréductibles[^12]. Les résultats. -------------- Décrivons les principaux résultats que nous obtenons. La nouveauté principale par rapport au cas d’une valeur propre différente de  $1$, traité dans \[B.91\], est l’apparition d’une “cohomologie intermédiaire” pour la partie spectrale relative à la valeur propre  $1$  de la monodromie, pour la fibre de Milnor de  $f$  à l’origine. Nous la noterons par  $H^n_{c\,\cap\,S}(0)$  et elle est introduite au paragraphe 4. Elle correspond à des cycles évanescents dont la limite à l’origine rencontre le sous-ensemble analytique  $S$  suivant un compact (mais la limite peut rencontrer le bord de  $f^{-1}(0)$  loin de  $S$). Ceci est précisé dans l’appendice I. On a alors deux applications naturelles d’élargissement de support $$H^n_c(0) \overset{can_{c}}{\rightarrow} H^n_{c\,\cap\,S}(0) \overset{can_{c\,\cap\,S}} {\rightarrow} H^n(0)$$ dont la composée est l’application canonique usuelle $$can : H^n_c(0)\rightarrow H^n(0) .$$ Dans le cas d’une singularité isolée pour la valeur propre 1, l’application  $ can_{c\,\cap\,S} : H^n_{c\,\cap\,S}(0) \rightarrow H^n(0) $  est un isomorphisme. Ce n’est plus le cas sous les hypothèses standards précisées plus haut et nous dirons que le phénomène de [**suremmêlement pour la valeur propre 1**]{} se produit dans cette situation précisément quand cette application n’est pas un isomorphisme. En suivant la même ligne de preuve que dans \[B.91\] nous commencerons par établir le ### Théorème 1. [*Sous les hypothèses standards pour la valeur propre 1, on a, pour  $n\geq 3$, la suite exacte :* ]{} $$0 \rightarrow H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) \overset{i}{\rightarrow} H^n_{c\,\cap\, S}(0) \overset{can_{c\,\cap\,S}}{\longrightarrow} H^n(0) \overset{\theta}{\rightarrow} H^1(S^*, H^{n-1}(0)) .$$ Ce théorème est prouvé au paragraphe 4, ainsi qu’un résultat analogue pour  $n = 2$. Maintenant l’existence d’une forme hermitienne localement constante et non dégénérée sur le système local  $H^{n-1}(0)$  déduite de la forme hermitienne canonique des sections hyperplanes transverses à  $S^*$  qui présentent une singularité isolée pour la valeur propre  $1$  (voir \[B.90\]), donne l’inégalité[^13] $$\dim H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) \leq \dim H^1(S^*, H^{n-1}(0)) .$$ On en déduit l’inégalité $$\dim H^n_{c\,\cap\,S}(0) \geq \dim H^n(0) .$$ Le résultat le plus difficile sera de démontrer que l’on a, en fait, toujours égalité de ces deux dimensions sous les hypothèses standards, pour  $n \geq 3 $. ### Théorème 2. Notre preuve passe par la construction d’une application de variation, compatible aux monodromies $$var : H^n_{c\,\cap\,S}(0) \rightarrow H^n_c(0) \ ,$$ construction qui sera menée à bien au paragraphe 6. Nous prouverons ensuite au paragraphe 8 que cette variation est injective. Ceci impliquera l’égalité cherchée puisque la dualité de Poincaré sur la fibre de Milnor à l’origine donne l’égalité des dimensions de  $H^n_c(0)$  et  $H^n(0)$. L’injectivité de la variation (qui est donc finalement un isomorphisme grâce à l’égalité des dimensions), donne également la non dégénérescence de la forme hermitienne[^14] canonique généralisée : $$\mathcal{H} : H^n_{c\,\cap\,S}(0) \times H^n(0) \rightarrow \mathbb{C} .$$ qui est compatible aux monodromies et reliée, comme dans le cas d’une singularité isolée (voir \[Loe. 86\]) et dans le cas d’une singularité isolée pour la valeur propre  $1$  (voir \[B.97\]), à la dualité de Poincaré (hermitienne), notée  $ < , > $, par la formule $$\mathcal{H}(e,e') = < \tilde{var}(e) , e' > \quad \forall (e,e') \in H^n_{c\,\cap\,S}(0) \times H^n(0)$$ où  $ \tilde{var} = var\circ \Theta $ , l’automorphisme  $\Theta$  de  $H^n_{c\,\cap\,S}(0)$  étant défini par $$\Theta : = \sum_{k=0}^{\infty} (-1)^k \frac{(T-1)^k}{k+1} = \frac{Log\, T}{T}$$ où  $T$  dénote la monodromie. Ceci est montré au paragraphe 7. Nous concluons en étendant les théorèmes 13 et 14 de \[B.91\] au cas de la valeur propre 1. On obtient ainsi les deux théorèmes suivants qui, je l’espère, seront suffisants pour donner une approche “filtrée” (donc non localisée en  $f$) du phénomène d’emmêlement de strates dans le cas d’une fonction holomorphe à lieu singulier de dimension 1, au moins dans le cas considéré dans \[B.04\]. ### Théorème 3. *Sous les hypothèses standards pour la valeur propre 1, considérons  $e \in H^n(0) $  vérifiant  $\mathcal{N}^k(e) = 0 $  avec  $k \geq k_0$[^15]. Alors les conditions suivantes sont équivalentes :* - [1)]{}  $\tilde{ob}_k(e) = 0$ . - [2)]{} *Si  $w \in \Gamma(Y, Ker\,\delta^n(k))$  vérifie  $ r^n(k)(w) = e $, pour chaque  $j \in \mathbb{Z} $, les fonctionnelles analytiques $$P_{k+l} \big(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge \bar{w}_k \wedge \square \big)$$ sont nulles pour  $l \geq 2$.* *De plus, pour vérifier 2) il suffit de le faire pour  $l = 2$  et pour  $j \in [0,n] $.* ### Théorème 4. *Sous les hypothèses standards pour la valeur propre 1, supposons que le prolongement méromorphe de  $ \int_X \vert f \vert^{2\lambda} \square $  admette un pôle d’ordre  $ \geq k $  en un entier négatif, avec  $ k \geq \sup{(k_0, k_1)} + 2$  où  $k_0$  et  $k_1$  sont respectivement les ordres de nilpotence de la monodromie agissant sur le système local  $ H^{n-1}(0)_{\vert S^*} $  et sur l’espace vectoriel  $H^n(0)$.* *Alors l’ordre des pôles aux entiers négatifs assez grand est exactement égal à  $k$  et l’on a  $ k_0 \leq k_1 = k - 2 $  avec  $\tilde{ob}_{k_1} \not\equiv 0 .$* L’étude de “réciproques” analogues aux théorèmes 15 et 16 de \[B.91\] n’est pas abordée ici. Ces questions ne semblent pas du tout triviales à résoudre pour la valeur propre 1. D’ailleurs les résultats de ce type dans \[B.91\] sont probablement les plus délicats de cet article. Nous n’avons pas, non plus, abordé ici les raffinements de \[B.91\] qui sont obtenus dans \[B.03\].\ Nous terminons cette étude par deux appendices.\ Le premier donne une description topologique na[ï]{}ve (à la Milnor) du groupe  $H^n_{c \cap S}(0)$. Le second, écrit par Claude Sabbah, donne une interprétation en terme des complexes de cycles évanescents et de cycles proches des phénomènes topologiques sous-jacents à cet article. Voici le plan de cet article - [1.]{} Définitions et rappels. - [2.]{} Le cas d’une singularité isolée pour la valeur propre 1. - [3.]{} La forme hermitienne canonique sur le système local  $ H^{n-1}(0)\vert_{S^*}.$ - [4.]{} L’espace vectoriel monodromique  $H^n_{c\,\cap\,S}(0).$ - [5.]{} La suite exacte longue de J. Leray généralisée. - [6.]{} Construction de  $\tilde{var} : H^n_{c\,\cap\,S}(0) \rightarrow H^n_c(0).$ - [7.]{} La forme hermitienne canonique  $\mathcal{H} : H^n_{c\,\cap\,S}(0)\times H^n(0) \rightarrow \mathbb{C}.$ - [8.]{} Injectivité de la variation pour  $n \geq 3$. Le cas  $ n = 2$. - [9.]{} Applications et un exemple pour conclure. - [Appendice I :]{} Description topologique de  $H^n_{c\,\cap\,S}(0).$ - [Appendice II :]{} Interprétation en termes de complexes de cycles proches et évanescents [*par Claude Sabbah*]{}. Définitions et rappels. ======================== Sous les hypothèses standards, introduisons les complexes de faisceaux   $(\Omega^{\bullet}(k), \delta_0 )$  et  $(\mathcal{E}^{\bullet}(k),\delta_0^{\bullet} )$. Rappelons que  $\Omega^{\bullet}$ (resp. $\mathcal{E}^{\bullet} $ ) désigne le faisceau des formes méromorphes (resp. semi-méromorphes) de degré $^{\bullet}$ sur  $X$  à pôles dans  $Y$  restreint (topologiquement) à  $Y$, que $$\Omega^{\bullet}(k) : = \Omega^{\bullet} \otimes \mathbb{C}^k , \quad \mathcal{E}^{\bullet}(k) : = \mathcal{E}^{\bullet} \otimes \mathbb{C}^k$$ et que $$\delta_0(w) : = dw - \frac{df}{f}\wedge\ _kN(w)$$ où  $ _kN \in End(\mathbb{C}^k ) $  est défini dans la base canonique  $ e_1, \cdots , e_k $  par\ $ _kN (e_j) = e_{j+1} , j \in [1,k] $  avec la convention  $ e_{k+1} = 0 $ ; cet endomorphisme agit sur  $ \Omega^{\bullet}(k) $\ (resp. sur  $\mathcal{E}^{\bullet}(k)$ )  par  $_k\mathcal{N} : = Id \otimes\, _kN$. Ceci se “visualise” de la fa[ç]{}on suivante : $$\delta_0(w) = \delta_0 \begin{pmatrix} {w_k} \\ {\vdots }\\ {w_2} \\ {w_1} \end{pmatrix} = \begin{pmatrix}{dw_k - \frac{df}{f}\wedge w_{k-1}} \\ {\vdots }\\ {dw_2 -\frac{df}{f}\wedge w_1} \\{dw_1} \end{pmatrix} .$$ De plus l’inclusion évidente de  $(\Omega^{\bullet}(k), \delta_0 )$  dans   $(\mathcal{E}^{\bullet}(k),\delta_0^{\bullet} )$  est un quasi-isomorphisme (voir \[B.91\], prop.1, p. 444.) On a les morphismes de complexes de degré 0 suivants, pour tout couple  $(k, k')$  d’entiers : $$\begin{aligned} \pi_{k+k', k} : (\mathcal{E}^{\bullet}(k+k'), \delta_0^{\bullet}) \longrightarrow (\mathcal{E}^{\bullet}(k), \delta_0^{\bullet}) \\ j_{k, k+k'} : (\mathcal{E}^{\bullet}(k), \delta_0^{\bullet}) \longrightarrow (\mathcal{E}^{\bullet}(k+k'), \delta_0^{\bullet}) \\\end{aligned}$$ donnés par les projections et injections évidentes : $$\begin{aligned} \mathbb{C}^{k+k'} \simeq \mathbb{C}^k \oplus \mathbb{C}^{k'} \rightarrow \mathbb{C}^k \\ \mathbb{C}^k \hookrightarrow \mathbb{C}^{k+k'} \simeq \mathbb{C}^k \oplus \mathbb{C}^{k'} . \end{aligned}$$ Ils se “visualisent” de la fa[ç]{}on suivante : $$\pi_{k+k', k}\begin{pmatrix} {w_{k+k'}}\\ {\vdots}\\ {w_k}\\ {\vdots} \\ {w_1} \end{pmatrix} = \begin{pmatrix} {w_k} \\ {\vdots} \\ {w_1} \end{pmatrix} \quad \quad j_{k, k+k'} \begin{pmatrix} {w_k} \\ {\vdots} \\ {w_1} \end{pmatrix} = \begin{pmatrix} {w_k} \\ {\vdots} \\ {w_1} \\ {0} \\ {\vdots} \\ {0} \end{pmatrix} ,$$ où le dernier vecteur colonne comprend  $k'$  zéros. On définit également, pour  $k \geq 1 $, le morphisme de complexe de degré  $0$ : $$_k \mathcal{N} : (\mathcal{E}^{\bullet}(k), \delta_0^{\bullet}) \rightarrow (\mathcal{E}^{\bullet}(k), \delta_0^{\bullet})$$ en posant  $ _k \mathcal{N} = j_{k-1,k}\circ \pi_{k,k-1} $ ; ce qui se “visualise” comme suit : $$_k \mathcal{N}\begin{pmatrix} {w_k}\\ {\vdots} \\ {w_2} \\{w_1} \end{pmatrix} = \begin{pmatrix} {w_{k-1}}\\ {\vdots} \\ {w_1} \\ {0} \end{pmatrix} .$$ On a, bien sur,  $ (_k \mathcal{N})^k = 0 $  sur  $\mathcal{E}^{\bullet}(k).$ La monodromie est alors définie sur ce complexe par  $ T : = exp(-2i\pi. _k\mathcal{N}).$ Enfin nous utiliserons également les morphismes de complexes[^16] de degré +1 $$\tau_{k,k'} : (\mathcal{E}^{\bullet}(k), \delta_0^{\bullet}) \rightarrow (\mathcal{E}^{\bullet}(k'), \delta_0^{\bullet})$$ donnés par : $$\tau_{k,k'}\begin{pmatrix} {w_k}\\ {\vdots} \\ {w_2} \\ {w_1} \end{pmatrix} = \begin{pmatrix}{0} \\ {\vdots} \\ {0} \\ {\frac{df}{f}\wedge w_k} \end{pmatrix} \Bigg\rbrace k' .$$ Notons par  $h^i(k)$  le i-ème faisceau de cohomologie du complexe  $(\mathcal{E}^{\bullet}(k),\delta_0 )$. Le morphisme de complexe $$r^{\bullet} : (\Omega^{\bullet}(k), \delta_0 ) \rightarrow (\Omega^{\bullet}_{X/f}[f^{-1}], d_{/f})\vert_{Y}$$ défini par  $ \tilde{r}^{\bullet}(w) = w_k $  pour  $ w \in \Omega^{\bullet}(k) $, donne la suite exacte courte suivante, pour tout  $p \geq 1 $ (voir la prop.1 p. 427-428 de \[B.91\] et la proposition (1.9) pour le lien entre  $\tilde{r}^{\bullet}$  et  $r^{\bullet}$) : $$0 \rightarrow H^{p-1}(0)\big/ Im \mathcal{N}^k_{p-1} \rightarrow h^p(k) \overset{r^p(k)}{\rightarrow} Ker \mathcal{N}^k_p \rightarrow 0$$ oú  $ -2i\pi\mathcal{N}_q $  désigne le logarithme nilpotent de la monodromie de  $f$  agissant[^17] sur le faisceau  $ H^q(0) $  pour  $ q \geq 0 $. On prendra garde que pour  $q=0$  on considère la cohomologie réduite (donc les “vrais” cycles évanescents et non les cycles proches.) Sous les hypothèses standards la suite exacte (1) montre que les seuls faisceaux de cohomologie non nuls du complexe  $(\mathcal{E}^{\bullet}(k),\delta_0 )$  sont les  $h^i(k)$  pour  $ i =0,1,n-1,n,n+1 .$ Supposons d’abord  $ n \geq 3$. On a alors des isomorphismes monodromiques  $\underline{\mathbb{C}}_Y \simeq h^0(k) $  et  $\underline{\mathbb{C}}_Y \simeq h^1(k) $  donnés par $$\lambda \in \underline{\mathbb{C}}_Y \rightarrow \begin{pmatrix} {\lambda} \\ {0} \\ {.} \\ {.} \\ {.} \\ {0} \end{pmatrix} \quad {\rm et} \quad \lambda \in \underline{\mathbb{C}}_Y \rightarrow \begin{pmatrix} {0} \\ {.} \\ {.} \\ {.} \\ {0} \\ { \lambda \frac{df}{f} } \end{pmatrix} .$$ Le morphisme  $\tau_k $  induit l’isomorphisme évident . De plus, grace à la nullité de  $H^{n-2}(0)$, on a un isomorphisme $$r^{n-1}(k) : h^{n-1}(k) \rightarrow Ker\,\mathcal{N}_{n-1}^k \subset H^{n-1}(0) \quad \forall k \geq 1$$ ainsi que la suite exacte pour  $k' \gg 1$ $$0 \rightarrow h^{n-1}(k') \overset{\tau_{k',k}}{\rightarrow} h^n(k) \rightarrow Ker\,\mathcal{N}_n^k \rightarrow 0 \quad \forall k \geq 1 .$$ Comme les faisceaux  $ h^{n-1}(0) $  et  $h^n(0)$  sont à support dans  $S$  et que  $H^n(0)$  est à support l’origine, on en déduit que sur  $ S^*$  on a un isomorphisme pour  $ k \gg 1 $ $$\tau_k : h^{n-1}(k)\vert_{S^*} \simeq h^n(k)\vert_{S^*} .$$ Alors, la suite exacte pour  $ k \gg 1$ $$0 \rightarrow H^{n-1}(0) \overset{i^{n-1}}{\rightarrow} h^n(k) \overset{\tau_k}{\rightarrow}h^{n+1}(k) \rightarrow 0$$ montre que  $h^{n+1}(k) $  est à support l’origine et isomorphe à  $H^n(0)$ (pour  $k \gg1$). Pour  $ n = 2 $, on a la suite exacte $$0 \rightarrow h^0(k) \overset{\tau_k}{\rightarrow} h^1(k) \overset{r^1(k)}{\rightarrow} Ker\, \mathcal{N}_1^k \rightarrow 0 \quad \forall k \geq 1 .$$ Comme  $\tau_k $  induit un isomorphisme sur  $Y \setminus S$  ( et les faisceaux  $h^i(k)\vert_{Y\setminus S} $  sont isomorphes à  $ \underline{\mathbb{C}}_{Y\setminus S} , \ i = 0,1$ ), on aura dans ce cas des isomorphismes monodromiques  $ h^1(k)\big/ \tau_k h^0(k) \simeq Ker\,\mathcal{N}_1^k \subset H^1(0) $  ainsi que  $ h^0(k) \simeq \underline{\mathbb{C}}_{Y} .$ Pour  $n = 2$  on remplacera l’isomorphisme  $ r^{n-1}(k) : h^{n-1}(k) \rightarrow H^{n-1}(0) $  dû à la nullité de l’application  $\tau_k : h^{n-2}(k) \rightarrow h^{n-1}(k) $  que l’on a pour  $n \geq 3$  par l’isomorphisme $$\underset{k \rightarrow \infty}{lim} \ h^1(k) \rightarrow H^1(0)$$ qui résulte de l’isomorphisme monodromique (plus concret) pour  $k \geq k_0$ $$r^{1}(k+1) : j_{k,k+1}(h^{1}(k)) \rightarrow H^1(0).$$ Comme l’application  $\tau_k$  n’est pas compatible à la limite inductive sur  $k$, il sera plus commode d’utiliser directement l’isomorphisme  $ \underset{k \rightarrow \infty}{lim} \ h^1(k) \rightarrow H^1(0) $  que la suite exacte $$0 \rightarrow h^0(k) \overset{\tau_k}{\rightarrow} h^1(k) \overset{r^1(k)}{\rightarrow} Ker\, \mathcal{N}_1^k \rightarrow 0 \quad \forall k \geq 1 .$$ Ce point est important puisqu’il permettra de traiter le cas  $ n = 2$  de fa[ç]{}on assez analogue au cas “général”  $ n \geq 3 $. On déduit alors facilement de la constructibilité des faisceaux  $ h^i(k) $  et de la contractibilité de  $Y$ la propriété suivante[^18] : $$\mathbb{H}^n(Y, (\mathcal{E}^{\bullet}(k), \delta^{\bullet})) \big/ \tau_k \big(\mathbb{H}^{n-1}(Y, (\mathcal{E}^{\bullet}(k), \delta^{\bullet}))\big)\quad \simeq Ker\, \mathcal{N}_n^k \subset H^n(0)$$ Nous allons établir directement une assertion analogue pour les supports compacts ; cette approche directe s’applique également aux supports fermés. Par contre la dégénérescence de la suite spectrale évoquée dans ce cas n’est plus vraie pour les supports compacts. Par ailleurs ceci motivera notre définition de l’espace vectoriel monodromique  $H^n_{c\,\cap\, S}(0)$  et rendra “évidente” les applications “naturelles”  $H^n_c(0) \rightarrow H^n_{c\,\cap\, S}(0)$  et  $ H^n_{c\,\cap\, S}(0) \rightarrow H^n(0)$. **Proposition.** ---------------- Pour chaque  $k \geq 1$, on a une application surjective $$r^n_c(k) : \mathbb{H}_c^n(Y, (\mathcal{E}(k)^{\bullet}, \delta^{\bullet})) \rightarrow Ker\, \mathcal{N}_{n,c}^k \subset H_c^n(0)$$ dont le noyau contient  $ \tau_k \ \mathbb{H}_c^{n-1}(Y, (\mathcal{E}(k)^{\bullet}, \delta^{\bullet}))$.\ Ces applications vérifient  $ r^n_c(k) \circ\, _k\mathcal{N} = \mathcal{N}_{n,c} \circ r^n_c(k) $  et elles sont compatibles aux applications  $ j_{k,k+k'} $. De plus, ces applications induisent un isomorphisme compatible aux monodromies entre $$\underset{k \rightarrow \infty}{lim}\ \mathbb{H}_c^n(Y, (\mathcal{E}(k)^{\bullet}, \delta^{\bullet}))$$ et  $H^n_c(0).$ *Preuve.* En fait il est plus facile de définir l’application $$\tilde{r}^n_c(k) : \mathbb{H}_c^n(Y, (\mathcal{E}(k)^{\bullet}, \delta^{\bullet})) \rightarrow Ker\, \mathcal{N}_{n,c}^k \subset H_c^n(0)$$ qui est reliée à  $r^n_c(k)$  par la formule  $ r^n_c(k) = exp(-\mathcal{N}_{n,c}.Log (s_0) )\circ \tilde{r}^n_c(k)$, où  $ F : = f^{-1}(s_0) $  est la fibre de Milnor de  $f$  à l’origine, et où  $s_0 \in D \cap \mathbb{R}^{+*} $  est un point base (et on a fixé  $ Log (s_0) \in \mathbb{R} .$) Comme  $exp(-\mathcal{N}_{n,c}.Log (s_0) )$  est un automorphisme de  $ H^n_c(0)$  compatible avec la monodromie  $ T_c = exp(-2i\pi.\mathcal{N}_{n,c}) $  il suffit de prouver la proposition pour l’application  $\tilde{r}^n_c(k)$[^19]. Pour  $w \in \Gamma_{c/f}(X,\mathcal{E}^n(k)) \cap Ker\,\delta $  posons $$\tilde{r}^n_c(k)[w] = [w_k\vert_F ] \in H^n_c(0) .$$ En effet, la relation  $ \delta w = 0 $  donne  $ dw_k = \frac{df}{f}\wedge w_{k-1} $  ce qui montre que la restriction de  $w$  à  $F$  est bien d-fermée (et à support compact). Pour voir la surjectivité, il suffit de constater que si  $e \in H^n_c(0) $  vérifie  $ \mathcal{N}^k_c(e)= 0 $  alors  $ \varepsilon : = \sum_0^{k-1} \frac{(Log s )^j}{j!} \mathcal{N}^j_c(e) $  est une section uniforme du fibré de Gauss-Manin à support propre de  $f$  qui se calcule à l’aide du complexe  $ (\mathcal{E}^{\bullet},\delta^{\bullet}) $ (voir \[B.91\] Prop.1 p.444) ce qui permet de trouver un antécédent à  $e$  de la même fa[ç]{}on que dans le cas à support fermé. Pour démontrer la relation  $ r^n_c(k) \circ\,_k\mathcal{N} = \mathcal{N}_{n,c} \circ r^n_c(k) $  nous devons montrer que, si  $ \delta w = 0 $  on a  $ [w_{k-1}\vert_F] = \mathcal{N}_c([w_k\vert_F]) .$ Soit  $ \tilde{X} : = X \underset{D}{\times} H $  où  $ H \overset{exp}{\rightarrow} D^* $  est le revêtement universel . D’après Milnor \[Mi.68\] on a un isomorphisme  $\mathcal{C}^{\infty}$  au dessus de  $D^*$  $\Phi : \tilde{X} \rightarrow F \times H $. La n-forme  $\mathcal{C}^{\infty}$ $$\Omega : = \sum_0^{k-1} \frac{(-1)^j}{j!}.(\zeta - Log (s_0))^j .\tilde{w}_{k-j}$$ où  $ \tilde{w_h} = ((\Phi)^{-1})^*(w_h) $  vérifie  $ d\Omega = 0 $  sur  $ F \times H$. Elle vérifie également  $ \Omega\vert_{F\times \lbrace Log (s_0) \rbrace} = w_k\vert_F .$ On en déduit que $$T_c(w_k\vert_F) = exp(-2i\pi.\mathcal{N}_c)(w_k\vert_F) = \sum_0^{k-1} \frac{(-1)^j}{j!} (2i\pi)^j w_{k-j}\vert_F .$$ On obtient alors facilement (par récurrence) la relation désirée. La compatibilité des applications  $ r^n_c(k)$  avec les morphismes  $ j_{k,k+k'} $  est immédiate. Nos assertions sur le noyau se déduisent facilement du lemme suivant. **Lemme.** ---------- On utilise les notations introduites ci-dessus. Soit  $ w \in \Gamma_c(Y, \mathcal{E}^n(k)) \cap Ker\,\delta $. Supposons que l’on ait  $ r^n_c(k)([w]) = 0 $  et qu’il existe  $ v \in \Gamma_c(Y, \mathcal{E}^{n-1}(k)) $  vérifiant  $ _k\mathcal{N}(w) = \delta v $. Alors $$j _{k,k+1}(w) = \begin{pmatrix}{w_k} \\ {\vdots} \\ {w_1}\\{0} \end{pmatrix} \in \delta \Gamma_c(Y, \mathcal{E}^{n-1}(k+1)) .$$ *Preuve.* Commen[ç]{}ons par traiter le cas  $k =1$. La condition sur  $\mathcal{N}_c(w)$  est vide, et on a seulement  $ dw_1 = 0 $  et  $ [w_1\vert_F ] = 0 $. Mais alors  $w_1$  est une section horizontale du fibré de Gauss-Manin à support propre de  $f$  qui est nulle en  $ s_0$. Elle est donc nulle, et on peut écrire $$w_1 = \frac{df}{f}\wedge \alpha + d\beta$$ où  $ \alpha, \beta \in \Gamma_{c/f}(X, \mathcal{E}^{n-1}(1)) $. On a alors  $ \frac{df}{f}\wedge d\alpha = 0 $, ce qui montre que  $\alpha$  est une section du fibré de Gauss-Manin à support propre de  $f$ de degré  $n-1$. Mais ce fibré (méromorphe) est nul, car on a  $ H^{n-1}_c(F) = 0 $  puisque  $F$  est une variété de Stein de dimension  $n$. On peut donc écrire $$\alpha = \frac{df}{f}\wedge \gamma + d\xi$$ où  $ \gamma, \xi \in \Gamma_{c/f}(X, \mathcal{E}^{n-2}(1)) $. On a donc $$w_1 = \frac{df}{f} \wedge d\xi + d\beta = d(\beta - \frac{df}{f}\wedge\xi ).$$ On obtient alors  $ w_1 = \delta \tilde{\beta}$  où  $ \tilde{\beta} \in \Gamma_{c/f}(X, \mathcal{E}^{n-1}(1)) .$ Passons au cas général. La relation  $_k\mathcal{N}(w) = \delta v $  donne $$w_{k-1} = dv_k -\frac{df}{f}\wedge w_{k-2}$$ et  $ \delta w = 0 $  donne  $ dw_k = \frac{df}{f}\wedge w_{k-1} $. On aura donc  $ dw_k = \frac{df}{f}\wedge dv_k $. La $n$-forme semi-méromorphe  $ w_k + \frac{df}{f}\wedge v_k$  est donc $d-$fermée à support $f-$propre et induit $0$ dans  $ H^n_c(0) $. D’après le cas  $k = 1 $  on peut écrire $$w_k = -\frac{df}{f}\wedge v_k + d\tilde{\beta} .$$ On aura alors $$j_{k,k+1}(w) = \begin{pmatrix}{w_k} \\{w_{k-1}}\\ {\vdots} \\ {w_1}\\{0} \end{pmatrix} = \delta \begin{pmatrix}{\tilde{\beta}}\\ {v_k }\\ {\vdots} \\ {v_2}\\ {v_1} \end{pmatrix}$$ ce qui achève la preuve du lemme et de la proposition (1.9). $\hfill \blacksquare $ **Remarque.** ------------- Comme on a  $ j_{k,k+k'} \circ \tau_k = 0 $  pour  $k' \geq k$  il est inutile de quotienter par  $\tau_k \big(\mathbb{H}^{n-1}_c(Y, (\mathcal{E}^{\bullet}(k), \delta^{\bullet}))\big)$  avant de prendre la limite inductive sur  $k$ . Nous considérerons également le complexe de faisceaux sur  $Y$  noté  $(\mathcal{F}^{\bullet}(k), \delta_0^{\bullet} )$, défini pour  $n \geq 3$  comme suit : $$(\mathcal{F}^{\bullet}(k), \delta_0^{\bullet} ) : = \ \mathcal{E}^0(k) \overset{\delta_0}{ \rightarrow } Ker\, \delta_0^1 \rightarrow 0 \rightarrow 0 \cdots$$ On a une inclusion naturelle de complexes  $ (\mathcal{F}(k)^{\bullet},\delta^{\bullet}_0) \rightarrow (\mathcal{E}^{\bullet}(k),\delta^{\bullet}_0 )$ Nous noterons par  $(\mathcal{\check{E}}^{\bullet}(k),\delta_0^{\bullet} )$ le complexe quotient ; ce qui donne de fa[ç]{}on explicite : $$(\mathcal{\check{E}}^{\bullet}(k),\delta^{\bullet}_0 ) : = \ 0 \rightarrow \mathcal{E}^1(k)/Ker\,\delta_0^1 \overset{\delta_0}{\rightarrow} \mathcal{E}^2(k) \overset{\delta_0}{\rightarrow} \mathcal{E}^3(k) \overset{\delta_0}{\rightarrow}\cdots$$ Le complexe  $(\mathcal{\check{E}}^{\bullet}(k),\delta^{\bullet}_0 )$   n’aura sur  $Y^*$  que deux faisceaux de cohomologie donc nuls, comme c’était le cas pour le complexe  $(\mathcal{E}^{\bullet}(k),\delta^{\bullet}_u )$   pour une valeur propre différente de 1. Ceci nous permettra d’utiliser sur  $S^*$   la proposition 2 p. 435 de \[B.91\]. Pour  $n = 2$  nous définirons le complexe  $(\mathcal{F}^{\bullet}(k),\delta_0^{\bullet} )$  en rempla[ç]{}ant dans la définition précédente le faisceau  $Ker\, \delta_0^1$  par le sous-faisceau $$\delta_0 \mathcal{E}^{0}(k) + \tau_k h^{0}(k) \hookrightarrow Ker\,\delta_0^{1}(k) .$$ De même nous remplacerons dans la définition de  $(\mathcal{\check{E}}^{\bullet}(k),\delta^{\bullet}_0 )$  le quotient  $\mathcal{E}^1(k)/Ker\,\delta_0^1 $  par $$\mathcal{E}^1(k)\big/ \delta_0 \mathcal{E}^{0}(k) + \tau_k h^{0}(k) .$$ Ceci conduit à nouveau au fait que le complexe  $(\mathcal{\check{E}}^{\bullet}(k),\delta^{\bullet}_0 )$  ait deux faisceaux de cohomologie non nuls sur  $Y^*$ isomorphes à  $h^{1}(k)/\tau_k (h^{0}(k) $  et  $h^2(k)$  en degrés  $ n-1 = 1 $  et  $ n = 2$  respectivement. Le quotient  $ \mathcal{E}^1(k)\big/ \delta_0 \mathcal{E}^{0}(k) + \tau_k h^{0}(k) $  n’ a pas la cohomologie sur  $S^*$ en degrés positifs. En effet, on a $$\delta_0\mathcal{E}^0(k)\,\cap\, \tau_k(h^0(k)) \simeq 0$$ On en déduit facilement que  $ \forall i \geq 1$ $$H^i(S^*, \mathcal{E}^1(k)\big/ \delta_0\mathcal{E}^{0}(k) + \tau_k h^{0}(k) ) \simeq H^{i+1}(S^*, h^0(k)) \oplus H^{i+1}(S^*, \delta_0\mathcal{E}^{0}(k)) \simeq 0 .$$ On pourra donc encore utiliser sur  $S^*$  la proposition 2 p. 435 de \[B.91\] pour décrire l’hypercohomologie sur  $S^*$  du complexe  $(\mathcal{\check{E}}^{\bullet}(k),\delta^{\bullet}_0 )$. **Le cas d’une singularité isolée pour la valeur propre 1 .** ============================================================= Considérons ici le cas où on a  $n \geq 1 $  et où l’on suppose que les faisceaux  $H^p(0)$ sont tous nuls pour  $p \not= n $. Quitte à choisir le représentant de Milnor de notre germe assez petit on peut également supposer que le faisceau  $H^n(0) $  est concentré à l’origine. On a alors les résultats suivants. **[Théorème.]{}(voir \[B.90\]).** --------------------------------- [*Soit  $\tilde{f} : (\mathbb{C}^{n+1}, 0) \rightarrow (\mathbb{C}, 0)$  un germe de fonction holomorphe présentant une singularité isolée à l’origine pour la valeur propre 1 de la monodromie[^20]. Il existe une forme hermitienne  $h$  sur  $H^n(0)$, non dégénérée, invariante par la monodromie, définie de la fa[ç]{}on suivante :*]{} [*Soient  $ e, e' \in H^n(0) $  vérifiant  $\mathcal{N}^k(e) =\mathcal{N}^k(e') = 0 $  où on a noté par  $-2i\pi.\mathcal{N}$  le logarithme nilpotent de la monodromie (unipotente) agissant sur  $H^n(0)$. Soient  $ w,w' \in H^n(\Gamma(X, \mathcal{E}^{\bullet}(k)), \delta_0^{\bullet}) $  vérifiant  $ r^n(k)(w) = e$  et  $ r^n(k)(w') = e'$. Alors on a* ]{} $$(2i\pi)^{n+1}. h(e,e') = P_2\big(\lambda = 0, \int_X \vert f \vert^{2\lambda} \rho.\frac{df}{f}\wedge w_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w'}_k \big)$$ [*pour  $\rho \in \mathcal{C}^{\infty}_c(X) $  vérifiant  $ \rho \equiv 1 $  près de l’origine.*]{} On a noté  $ P_2(\lambda = 0, F(\lambda)) $  le coefficient de  $\frac{1}{\lambda^2}$  dans le développement de Laurent en  $ \lambda = 0 $  de la fonction méromorphe  $F.$ On utilisera dans cet article, comme on l’a déjà fait dans \[B.97\], les complexes  $(\Omega^{\bullet}(k), \delta^{\bullet}_0)$  pour le cas méromorphe et  $(\mathcal{E}^{\bullet}(k), \delta^{\bullet}_0 )$  pour le cas semi-méromorphes au lieu de ceux correspondants à la differentielle  $ \delta^{\bullet}_1$  de la différentielle (utilisés dans \[B.90\]) ce qui explique que l’énoncé ci-dessus diffère de celui de \[B.90\] (en particulier par “ $ \lambda = 0$ ”). Nous renvoyons à \[B.91\], aux rappels donnés dans \[B.97\] paragraphe 2 ou au paragraphe 1 ci-dessus pour plus de précisions sur ces complexes et sur l’application  $r^n(k)$. La forme hermitienne définie ci-dessus dans le cas d’une singularité isolée pour la valeur propre 1 de la monodromie sera appelée la forme hermitienne canonique. Dans le cas d’une singularité isolée, la forme hermitienne canonique a été introduite dans \[B.85\]. Fran[ç]{}ois Loeser \[Loe.86\] a montré dans ce cas, en utilisant des résultats de théorie de Hodge, comment construire la forme hermitienne canonique à partir de la variation et de la dualité (hermitienne) de Poincaré sur la fibre de Milnor. Ceci a été généralisé au cas d’une singularité isolée pour la valeur propre 1 (sans théorie de Hodge, car elle ne semble pas facilement applicable dans ce contexte plus général) dans \[B.97\]. Rappelons les énoncés correspondants. **[Théorème.]{} [(voir \[B.97\]).]{}** -------------------------------------- *Soit  $\tilde{f} : (\mathbb{C}^{n+1}, 0) \rightarrow (\mathbb{C}, 0)$  un germe de fonction holomorphe présentant une singularité isolée à l’origine pour la valeur propre 1 de la monodromie. Il existe une application linéaire “naturelle” de variation  $ var : H^n(0) \rightarrow H^n_c(0) $  vérifiant les propriétés suivantes :* - [0)]{} *[Si la singularité de  $f$  est isolée[^21], on retrouve la variation “classique” (voir \[A.V.G.\]).]{}* - [1)]{}*[ L’application “var” est topologique et on a ]{} $$\begin{aligned} can \circ var = T - 1 \quad {\rm sur} \quad H^n(0) \\ var \circ can = T_c - 1 \quad {\rm sur} \quad H^n_c(0) \end{aligned}$$ $$\mathcal{I}(a,b) = \frac{1}{(2i\pi)^n} \int_F a \wedge \bar{b} .$$ *où  $F$  désigne la fibre de Milnor de*  $f$  en  $0$.* - [3)]{} Soit  $ \Theta : = \sum_{k=0}^{\infty} (-1)^k \frac{(T-1)^k}{k+1} .$*[ C’est un automorphisme de  $H^n(0)$  puisque  $T$  est unipotent. Posons]{}  $ \tilde{var} : = var \circ \Theta $. *[Alors on a ]{} $$\forall e, e' \in H^n(0), \quad h(e,e') = \mathcal{I}(\tilde{var}(e), e') .$$** On déduit facilement des propriétés ci-dessus que l’on a $$\begin{aligned} \quad & Im\ can = Im\ (T-1) \qquad &{\rm dans} \quad H^n(0) \\ {\rm et} \qquad & Ker\ can = Ker\ (T_c -1) \qquad &{\rm dans} \quad H^n_c(0) .\end{aligned}$$ Comme nous aurons besoin plus loin de construire une généralisation de l’application de variation, nous allons rappeler ici brièvement l’expression de l’application  $ \tilde{var}$  donnée dans \[B.97\] en terme de formes différentielles semi-méromorphes. Soit donc  $e \in H^n(0)$  vérifiant  $\mathcal{N}^k(e) = 0 $. Soit  $ w \in \Gamma(Y, \mathcal{E}^n(k)) $  satisfaisant  $\delta_0(w) = 0 $  et  $ r^n(k) = e .$ L’hypothèse de singularité isolée pour la valeur propre 1 permet alors de trouver  $ u \in \Gamma(Y^*, \mathcal{E}^{n-1}(k)) $  vérifiant $$_kN\, w \vert_{Y^*} = \delta_0\, u .$$ Soit  $\rho \in \mathcal{C}^{\infty}_{c/f}(X) $  vérifiant  $ \rho \equiv 1 $  près de l’origine. Posons  $ \xi = (1-\rho).u $ . Alors on a  $ \xi \in \Gamma_c(Y, \mathcal{E}^{n-1}(k))$  et  $ v =\ _kN\, w - \delta_0\,\xi $  est dans  $\Gamma_c(Y, Ker\, \delta_0^n(k)) .$ On a donc, pour  $j \in [1,k]$, avec la convention  $w_0 = 0 = \xi_0 $ : $$w_{j-1} = v_j + d\xi_j - \frac{df}{f}\wedge \xi_{j-1} \ ;$$ et comme  $ \delta_0\, w = 0 $  implique  $ dw_k = \frac{df}{f} \wedge w_{k-1} $  on aura $$d(w_k + \frac{df}{f} \wedge \xi_k ) = \frac{df}{f} \wedge v_k .$$ Choisissons maintenant une fonction  $ \sigma \in \mathcal{C}^{\infty}_{c/f}(X) $  vérifiant  $ \sigma \equiv 1 $  sur un ouvert de  $Y$  contenant  $ (Supp\, v)\, \cap\, Y $. Alors le théorème de Leray généralisé au cas d’une hypersurface non nécessairement lisse mais ne présentant pas la valeur propre 1 pour la monodromie[^22], permet d’écrire au voisinage du bord de la boule de Milnor $$d\sigma \wedge (w_k + \frac{df}{f} \wedge \xi_k ) = \frac{df}{f}\wedge \eta + \omega + d\alpha$$ avec  $\eta$  et  $\omega$  d-fermées et  $ \mathcal{C}^{\infty}$  à support  $f-$propres de degrés respectifs  $n$  et  $n+1$, et où  $\alpha$  est semi-méromorphe de degré  $n$  à pôles dans  $Y$  et à support  $f-$propre. On obtient alors que la section  $V(w) \in \Gamma_c(Y, \mathcal{E}^n(k))$  définie par $$\begin{aligned} & V(w)_k = v_k + \eta \\ & V(w)_j = v_j \quad \quad {\rm pour} \quad j \in [1,k-1]\end{aligned}$$ est  $\delta_0-$fermée et on a  $ r^n(k)(V(w)) = \tilde{var}(e)$. Pour plus de détails, on consultera \[B.97\], ou bien la construction de la variation généralisée donnée au paragraphe 6. L’objectif de ce troisième paragraphe est de généraliser à notre situation la proposition 11 de \[B.91\]. **La forme hermitienne canonique sur le système local  $H^{n-1}(0)\vert_{S^*}.$** --------------------------------------------------------------------------------- ### Considérons donc le système local  $H^{n-1}(0)$  sur  $S^*$. Sa fibre en chaque point  $\sigma \in S^*$  s’identifie au sous-espace spectral pour la valeur propre 1 du  $(n-1)-$ième groupe de cohomologie de la fibre de Milnor de la restriction de  $f$ à une section hyperplane transverse à  $S^*$  en  $\sigma$. Comme la restriction de  $f$  à un tel hyperplan a une singularité isolée pour la valeur propre 1 en  $\sigma$, ce sous-espace spectral pour la valeur propre 1 de ce  $(n-1)-$ième groupe de cohomologie est muni d’une forme hermitienne canonique qui est non dégénérée et invariante par la monodromie de  $f$. Ceci permet, en raisonnant comme dans \[B.91\] p.456, de munir le système local (d’espaces vectoriels monodromiques)  $H^{n-1}(0)\vert_{S^*}$  d’une forme hermitienne localement constante, non dégénérée et invariante par la monodromie[^23]. Nous la noterons par $$h : H^{n-1}(0)\vert_{S^*} \times H^{n-1}(0)\vert_{S^*} \rightarrow \underline{\mathbb{C}}_{S^*} .$$ On montre comme au lemme 1 de \[B.91\] p.456 que cette forme hermitienne est intrinsèque. La proposition suivante se démontre de fa[ç]{}on analogue à la proposition 1 de \[B.91\] p.457. ### **Proposition.** On se place sous les hypothèses standards pour la valeur propre 1. Soit  $V$  un ouvert de  $X^* : = X \setminus \lbrace 0 \rbrace$ et soient  $w$  et  $w'$  deux sections sur  $V$  du faisceau  $\mathcal{E}^{n-1}(k)$  vérifiant  $ \delta(w) = \delta(w') \equiv 0 $. Notons par  $e$  et  $e'$  les sections correspondantes de  $H^{n-1}(0)$  sur  $U : = V\, \cap \, S^* $ (via le morphisme  $r^{n-1}(k)$). Alors la fonction  $h(e,e')$  est localement constante sur  $U$. Pour chaque forme différentielle  $ \varphi \in \mathcal{C}^{\infty}(V)$  de degré 2 vérifiant : - [i)]{} $ Supp\,\varphi \,\cap\, S^* $  est compact, - [ii)]{}  $d\varphi = 0 $  au voisinage de  $U$  dans  $X^*$, on aura la formule : $$(2i\pi)^{n} \int_U h(e,e').\varphi = P_2 \big( \lambda = 0, \int_X \vert f \vert^{2\lambda} w_k \wedge \bar{w'}_k \wedge \varphi \wedge \frac{df}{f} \wedge \frac{\bar{df}}{\bar{f}} \big).$$ ### Quitte à remplacer la formule (1) p.409 de \[B.91\] donnant l’expressions de la forme (hermitienne) d’intersection comme résidu dans le prolongement analytique de la distribution  $ \vert f \vert^{2\lambda}$  par la formule du théorème (0.6), la preuve de cette proposition est analogue à celle de la proposition 1 de \[B.91\] p.457-460 . ### Remarque. Pour  $n = 2$  il n’est pas à priori automatique de pouvoir représenter une section sur  $U : = V\, \cap \, S^* $ du faisceau  $H^1(0)$  par un élément de  $\Gamma(V, Ker\,\delta^1(k))$  pour  $ k \gg 1$, car cela nécessite la surjectivité des applications $$H^0(V, Ker\,\delta^1(k)) \rightarrow H^0(V, h^1(k)) \rightarrow H^0(V, H^1(0)) .$$ L’isomorphisme  $ h^1(k)\big/\tau_k(h^0(k)) \rightarrow H^1(0) $  pour  $ k \gg 1$, montre que l’on a une obstruction à la surjectivité de la seconde flèche. En fait le décalage sur  $k$  suffit à lever cette obstruction puisque  $ \underset{k \rightarrow \infty}{lim} h^1(k) \rightarrow H^1(0) $  est un isomorphisme. La surjectivité de la première flèche est realisée dès que l’on a  $ H^1(V, \delta\mathcal{E}^0(k)) = 0 $, ce qui sera toujours vrai quitte à prendre pour  $V$  un voisinage ouvert assez petit de  $U$, car on a  $H^1(V, \delta\mathcal{E}^0(k)) \simeq H^2(V, h^0(k)) \simeq H^2(U, \mathbb{C}) = 0 .$ ### En utilisant le cycle fondamental de  $S^*$  qui est l’élément de  $H_1(S^*, \mathbb{Z})$  obtenu en coupant  $S^*$  par une sphère centrée à l’origine de rayon assez petit, on obtient, un accouplement sesquilinéaire non dégénéré $$\tilde{h} : H^0(S^*, H^{n-1}(0)) \times H^1(S^*, H^{n-1}(0) \rightarrow \mathbb{C} .$$ Pour donner une formule analytique donnant le nombre  $\tilde{h}(e,e')$  analogue à la proposition 11 de \[B.91\], nous devons commencer par définir l’analogue des applications  $\tilde{\theta}_k$  et  $\tilde{ob}_k $  de loc. cit. **Les applications  $\tilde{\theta}_k$  et  $\tilde{ob}_k $.** -------------------------------------------------------------- ### Définissons l’entier  $k_0$  comme l’ordre de nilpotence de la monodromie agissant sur le système local  $H^{n-1}(0)\vert_{S^*}$[^24]. ### [**Définition.**]{} [*Sous les hypothèses standards pour la valeur propre 1 et pour  $n \geq 3$, notons par  $-2i\pi\mathcal{N}_n$ le logarithme nilpotent de la monodromie agissant sur  $H^n(0)$. Pour chaque  $k \geq k_0 $  on a deux applications linéaires naturelles* ]{} $$\begin{aligned} \tilde{ob}_k : & \ Ker\, \mathcal{N}_n^k \rightarrow H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) \\ \tilde{\theta}_k : & \ Ker\, \mathcal{N}_n^k \rightarrow H^1(S^*, H^{n-1}(0)) \simeq H^2_{\lbrace 0 \rbrace}(S, H^{n-1}(0))\end{aligned}$$ [*qui sont définies de la fa[ç]{}on suivante :*]{} - [*La composée*]{} $$\mathbb{H}^n(Y, (\mathcal{E}^{\bullet}(k), \delta^{\bullet})) \rightarrow \mathbb{H}^n(S^*, (\mathcal{\check{E}}^{\bullet}(k), \delta^{\bullet})) \overset{ob_k}{\rightarrow} H^0(S^*, H^{n-1}(0)) \overset{\partial}{\rightarrow} H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0))$$ [*passe au quotient par*]{}  $\tau_k \ \mathbb{H}^{n-1}(Y, (\mathcal{E}^{\bullet}(k), \delta^{\bullet}))$  [*et donne*]{}  $\tilde{ob_k}$[^25]. - [*La composée*]{} $$\mathbb{H}^n(Y, (\mathcal{E}^{\bullet}(k), \delta^{\bullet})) \rightarrow \mathbb{H}^n(S^*, (\mathcal{\check{E}}^{\bullet}(k), \delta^{\bullet})) \overset{\theta_k}{\rightarrow} H^1(S^*, H^{n-1}(0))$$ [*passe également au quotient par*]{}  $\tau_k \ \mathbb{H}^{n-1}(Y, (\mathcal{E}^{\bullet}(k), \delta^{\bullet}))$  [*et donne* ]{} $\tilde{\theta_k}.$ Ces résultats sont analogues à ceux des pages 410 et 411 de \[B.91\], le complexe  $(\mathcal{\check{E}}^{\bullet}(k), \delta^{\bullet}))$  ayant les mêmes propriétés dans notre cas que le complexe “usuel” pour une valeur propre différente de 1 (puisqu’il n’a que deux faisceaux de cohomologie non nuls sur  $Y^*$, et on a l’acyclicité sur  $S^*$  des faisceaux  $\mathcal{\check{E}}^{\bullet}(k)$.) On prendra garde qu’ici à nouveau les applications  $\tilde{\theta}_k$  sont compatibles quand  $k$  augmente, alors que l’on a  $\tilde{ob}_{k+1} = \tilde{ob}_k\circ \mathcal{N}_n.$ Nous définirons alors l’application  $ \theta : H^n(0) \rightarrow H^2_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) $  comme étant  $\tilde{\theta_k}$  pour  $k \gg 1$. ### **Remarque.** Toujours en supposant  $n \geq 3$, le faisceau  $ \check{\mathcal{E}}^1(k) \simeq \mathcal{E}^1(k) \big/ Ker\,\delta^1 $  n’a pas de cohomologie en degré positif sur  $S^*$. En effet on a les suites exactes $$\begin{aligned} 0 \rightarrow Ker\,\delta^1 \rightarrow \mathcal{E}^1(k) \rightarrow \check{\mathcal{E}}^1(k) \rightarrow 0 \\ 0 \rightarrow Im\,\delta^0 \rightarrow Ker\,\delta^1 \rightarrow h^1(k) \rightarrow 0 \\ 0 \rightarrow h^0(k) \rightarrow \mathcal{E}^0(k) \rightarrow Im\,\delta^0 \rightarrow 0\end{aligned}$$ qui donnent successivement les annulations pour tout  $k \geq 1$ $$\begin{aligned} & H^i(S^*, Im\,\delta^0(k)) = 0 , & \forall i \geq 1 \\ & H^j(S^*, Ker\,\delta^1(k)) = 0 , & \forall j \geq 2 \\ & H^l(S^*, \check{\mathcal{E}}^1(k) ) = 0 , & \forall l \geq 1 \end{aligned}$$ On a donc la dégénérescence de la première suite spectrale donnant l’hypercohomologie sur  $S^*$  du complexe  $ (\check{\mathcal{E}}^{\bullet}(k) ,\delta^{\bullet}) $  et donc un isomorphisme $$\mathbb{H}^n(S^*, (\mathcal{\check{E}}(k)^{\bullet}, \delta^{\bullet})) \simeq H^n(\Gamma(S^*, \mathcal{\check{E}}(k)^{\bullet}), \delta^{\bullet}) .$$ Nous pouvons étendre la définition des applications  $\tilde{\theta}_k$  et  $\tilde{ob}_k $  aux éléments  $\delta-$fermés de  $ \Gamma(S^*, \mathcal{E}^n(k))$  pour  $k \geq k_0$ : ils induisent des classes dans  $\mathbb{H}^n(S^*, (\mathcal{\check{E}}(k)^{\bullet}, \delta^{\bullet}))$  qui s’envoie dans  $\mathbb{H}^n(Y^*, (\mathcal{\check{E}}(k)^{\bullet}, \delta^{\bullet})) \simeq \mathbb{H}^n(S^*, (\mathcal{\check{E}}(k)^{\bullet}, \delta^{\bullet})) $  puisque les faisceaux de cohomologie non nuls du complexe  $(\mathcal{\check{E}}(k)^{\bullet}, \delta^{\bullet})$  sont à supports dans  $S$. Pour éviter toute confusion, nous noterons ces applications par $$\widehat{ob}_k : \Gamma(S^*, \mathcal{E}^n(k)) \cap Ker\,\delta \rightarrow H^0(S^*, H^{n-1}(0))$$ et par $$\widehat{\theta}_k : \Gamma(S^*, \mathcal{E}^n(k)) \cap Ker\,\delta \rightarrow H^1(S^*, H^{n-1}(0)).$$ On notera que  $\widehat{ob}_k $  est simplement donnée par la composée des morphismes de faisceaux sur  $S^*$ $$(\tau_k)^{-1} : h^n(k) \rightarrow h^{n-1}(k) \hookrightarrow H^{n-1}(0) .$$ ### **Le cas  $n = 2$.** La définition des applications  $\tilde{ob}_k $  et  $\theta$  ne pose pas de problème sérieux pour  $n = 2$. Une manière de le voir directement (sans utiliser \[B.91\]) est de transformer l’écriture locale $$w\vert_{X_{\sigma}} = \tau_k(\alpha_{\sigma}) + \delta\beta_{\sigma}$$ correspondant à la surjection sur  $S^*$  $\tau_k ; h^1(k) \rightarrow h^2(k) $, en $$j_{k,2k}(w)\vert_{X_{\sigma}} = \delta \begin{pmatrix}\beta_{\sigma}\\ \alpha_{\sigma}\end{pmatrix} .$$ Alors $$\begin{pmatrix} \beta_{\sigma} - \beta_{\sigma'} \\ \alpha_{\sigma} - \alpha_{\sigma'} \end{pmatrix}$$ induira un élément de  $H^1(S^*, h^1(2k))$  dont l’image dans  $H^1(S^*, H^1(0))$  (pour  $k \gg1$ ) sera  $\theta([w])$. La cochaîne  $\alpha_{\sigma}$  donnant une section globale sur  $S^*$ du faisceau quotient  $h^1(k)\big/\tau_k(h^0(k) \simeq H^1(0)$  pour  $ k \geq k_0 $, ce qui définit  $\tilde{ob}_k([w]).$ Donnons maintenant l’analogue de la proposition 11 de \[B.91\]. ### **Théorème.** [*Sous les hypothèses standards pour la valeur propre 1, considérons des entiers* ]{} $k \geq k_0$  et  $k' \geq k_0$. [*Soient* ]{} $ v\in \Gamma(S^*, \mathcal{E}^n(k)) \cap Ker\, \delta $  et  $ w\in \Gamma(S^*, \mathcal{E}^n(k)) \cap Ker\, \delta $. [*Notons par*]{}  $ e : = \widehat{ob}_k(v)\in H^0(S^*, H^{n-1}(0))$  et par  $ \eta : = \widehat{\theta}_k (w)\in H^1(S^*, H^{n-1}(0))$. [*Alors on a* ]{} $$(2i\pi)^{n+1}.\tilde{h}(e,\eta) = \sum_{a=1}^{k_0 } (-1)^a P_{a+1} \big(\lambda = 0, \int_X \vert f \vert^{2\lambda} \frac{df}{f}\wedge \bar{w'}_{k'}\wedge v_a \wedge \gamma \big)$$ [*où  $\gamma$  est une forme  $\mathcal{C}^{\infty}$  de degré 1 au voisinage de  $S^*$  vérifiant*]{}[^26] - [i)]{} $Supp\, \gamma \cap \, S^* $  [*est compact ;*]{} - [ii)]{}  $ d\gamma = 0$  [*au voisinage de*]{}  $S^*$ ; - [iii)]{}  $\gamma$  [*induit la classe du cycle fondamental de  $S^*$  dans*]{}  $H^1_c(S^*, \mathbb{C}).$ ### **Corollaire 1.** Sous les hypothèses du théorème précédent, supposons que  $v$  et  $w$   soient restrictions à  $S^*$  d’ éléments de  $ \Gamma(S, \mathcal{E}^n(k)) \cap Ker\, \delta$. Alors on a $ \tilde{h}(e,\eta) = 0 .$ Ce corollaire est analogue au corollaire 1 de la proposition 11 de \[B.91\]. ### **Corollaire 2.** Soient  $k,k' \geq k_0 $  et soient  $w \in H^n((\Gamma(Y, \mathcal{E}^{\bullet}(k)), \delta))$  et  $ w' \in H^n((\Gamma(Y, \mathcal{E}^{\bullet}(k')), \delta))$. Posons  $ e : = r^n(k)(w)$  et  $e' : = r^n(k')(w')$. Alors on a $$(2i\pi)^{n+1}.\tilde{h}(\tilde{ob}_k(e), \theta (e')) = (-1)^k P_{k+2}\big(\lambda = 0, \int_X \vert f \vert^{2\lambda} \frac{df}{f}\wedge w_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w'}_{k'} \big).$$ Ce corollaire est analogue au corollaire 2 de la proposition 11 de \[B.91\]. ### **Corollaire 3.** Sous les hypothèses standards, l’image de  $\theta$  est contenue dans l’orthogonal pour  $\tilde{h}$  du sous-espace  $ H^0(S, H^{n-1}(0))$  de  $H^0(S^*, H^{n-1}(0)).$ Ce dernier corollaire est analogue à la première partie de la proposition 12 de \[B.91\]. Il se déduit immédiatemment du corollaire 1 (voir la preuve du corollaire 1 de la proposition 11 de \[B.91\] p.466). ### **Remarque.** Soient  $k \geq k_0$ , et  $e \in Ker \mathcal{N}_n^k $  tels que  $\tilde{ob}_k(e) \not= 0 $. Pour conclure grace au corollaire 2 que l’on aura alors un pôle d’ordre  $k+2$   aux entiers négatifs dans le prolongement méromorphe de  $ \vert f \vert ^{2\lambda} $   (voir le théorème 13 de \[B.91\] et notre théorème 3), il est essentiel d’avoir l’égalité entre l’image de  $\theta $  dans  $ H^1(S^*, H^{n-1}(0)) $  et l’orthogonal pour la forme hermitienne  $\tilde{h}$  de  $H^0(S, H^{n-1}(0))$. C’est à dire de savoir que l’inclusion donnée au corollaire 3 est en fait une égalité. Ce qui revient à montrer que  $\tilde{h}$  établit une dualité hermitienne entre  $H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) $  et  $ Im \ \theta $. Ceci sera notre objectif dans ce qui suit. Il demandera encore beaucoup de travail puisqu’il ne sera atteint qu’ à la fin du paragraphe 8. Ce résultat sera également la clef de la non-dégénérescence de la forme hermitienne canonique “généralisée” que nous allons introduire plus loin. **L’espace vectoriel monodromique  $H^n_{c\,\cap\,S}(0)$.** =========================================================== Introduisons sur  $X$  et  $Y$  les familles paracompactifiantes de supports suivants -  $ c/f$   la famille des fermés  $f-$propres de  $X$. Elle donne par restriction à  $Y$  la famille des compacts de  $Y$  que nous noterons simplement “c”. -  $ c \cap S$  la famille des fermés de  $X$  (ou de  $Y$ ) qui rencontrent  $S$  suivant un compact. -  $mod\,S$ la famille des fermés de  $X$  (ou de  $Y$) qui ne rencontrent pas  $S$. Nous utiliserons également ces familles de supports pour un ouvert de  $X$  ou de  $Y$. Comme les faisceaux  $ \mathcal{E}^{\bullet}(k)$  sont fins et la famille  $c\,\cap\,S $  paracompactifiante, le  $j-$ième groupe de cohomologie du complexe $$\big( \Gamma_{c\,\cap\,S}(Y, \mathcal{E}^{\bullet}(k)), \delta^{\bullet}\big)$$ est isomorphe à l’hypercohomologie $$\mathbb{H}_{c\,\cap\,S}^j(Y, \mathcal{E}^{\bullet}(k)), \delta^{\bullet}).$$ Comme  $\tau_k $  est un morphisme de complexe de degré +1 il induit une flèche $$\tau_k : \mathbb{H}_{c\,\cap\,S}^{j-1}(Y, \mathcal{E}^{\bullet}(k)), \delta^{\bullet}) \rightarrow \mathbb{H}_{c\,\cap\,S}^j(Y, \mathcal{E}^{\bullet}(k)), \delta^{\bullet}) .$$ Mais l’égalité[^27] $$j_{k,k+k'}\circ\tau_k = \tau_{k+k'}\circ ( _{k+k'}\mathcal{N}^{k'})\circ j_{k,k+k'} , \tag{@}$$ entre morphismes de faisceaux  $ h^{\bullet}(k) \rightarrow h^{\bullet}(k+k') $   permet de voir qu’il est inutile de passer au quotient par  $ \tau_k \big(\mathbb{H}_{c\,\cap\,S}^{j-1}(Y, \mathcal{E}^{\bullet}(k)), \delta^{\bullet})\big) $  et l’on considerera donc simplement le système inductif $$\big( \mathbb{H}_{c\,\cap\,S}^j(Y, \mathcal{E}^{\bullet}(k)), \delta^{\bullet}) , j_{k,k+k'} \big)$$ De plus les endomorphismes de complexes  $ _k\mathcal{N}$  déduit des endomorphismes  $ _kN \in End(\mathbb{C}^k) $  sont compatibles avec les  $ j_{k,k+k'}$. **Définition.** --------------- [*Pa[ç]{}ons-nous sous les hypothèses standards pour la valeur propre 1. Nous définirons l’espace vectoriel  $ H^j_{c\,\cap\,S}(0) $  comme la limite inductive quand  $ k \rightarrow \infty$  du système inductif ci-dessus. L’action de  $ _k\mathcal{N} $  sur ce système inductif permet de définir une action monodromique (unipotente) sur sa limite inductive qui sera donnée par  $T_{c\,\cap\,S} = exp(-2i\pi. \mathcal{N}_{c\,\cap\,S}).$* ]{} Les limites inductives $$\underset{k \rightarrow \infty}{lim} \ \mathbb{H}^n_{c}(Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet})$$ et $$\underset{k \rightarrow \infty}{lim} \ \mathbb{H}^n(Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet}))$$ sont respectivement isomorphes à  $H^n_c(0) $  et  $H^n(0)$   (voir (1.8) et (1.9) ) ce qui nous fournit des applications  $\mathbb{C}-$linéaires monodromiques “naturelles” $$can_c^{c\,\cap\,S} : H^n_c(0) \rightarrow H^n_{c\,\cap\,S}(0) \quad {\rm et} \quad can_{c\,\cap\,S} : H^n_{c\,\cap\,S} \rightarrow H^n(0)$$ dont la composée est l’application usuelle d’oubli de support $$can : H^n_c(0) \rightarrow H^n(0) .$$ On remarquera que si la singularité de  $f$  est isolée pour la valeur propre 1 (et donc $S = \lbrace 0 \rbrace$), alors  $H^n_{c\,\cap\,S}(0)$  est isomorphe à  $H^n(0)$  via  $can_{c\,\cap\,S}$. Notre premier résultat va montrer que pour  $ n \geq 3$  le noyau  $Ker\,\theta$  est simplement l’image de l’application  $ H^n_{c\,\cap\, S}(0) \overset{can_{c\,\cap\,S}}{\longrightarrow} H^n(0)$  que nous avons introduite plus haut. Le cas  $n = 2$  sera traité séparément au (4.12). **Théorème 1.** --------------- [*Sous les hypothèses standarts, on a, pour  $n \geq 2$, la suite exacte :* ]{} $$0 \rightarrow H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) \overset{i}{\rightarrow} H^n_{c\,\cap\, S}(0) \overset{can_{c\,\cap\,S}}{\longrightarrow} H^n(0) \overset{\theta}{\rightarrow} H^1(S^*, H^{n-1}(0)) .$$ *.* Nous utiliserons les lemmes suivants : **Lemme.** ---------- Supposons  $n\geq 2$  et les hypothèses standards vérifiées pour la valeur propre 1 de la monodromie de  $f$. Alors on a $$\begin{aligned} \quad & H^j(S^*, \delta\mathcal{E}^i(k)) = 0 , \quad \quad \forall i \in [0,n-2] , \forall j \geq 1 \\ \quad & H^j(S, \delta\mathcal{E}^i(k)) \simeq H^j(Y, \delta\mathcal{E}^i(k)) = 0 , \quad \quad \forall i \geq 0 , \forall j \geq 1 \end{aligned}$$ *Preuve.* Nous utiliserons les annulations suivantes pour les faisceaux de cohomologie  $ h^j(k)$  du complexe  $( \mathcal{E}^{\bullet}(k),\delta^{\bullet})$  qui sont non nuls seulement pour  $ j = 0,1,n-1,n,n+1$ et qui sont décrits au (1.2) : $$\begin{aligned} H^q(S^*, h^j(k)) = 0 , \qquad \forall j \geq 0 , \forall q \geq 2 \\ H^p(S, h^j(k)) \simeq H^p(Y, h^j(k)) = 0 , \qquad \forall j \geq 0 , \forall p \geq 1 .\end{aligned}$$ Pour  $i = 0$  considérons la suite exacte de faisceaux : $$0 \rightarrow h^0(k) \rightarrow \mathcal{E}^0(k) \rightarrow \delta\mathcal{E}^0(k) \rightarrow 0 . \tag{$A_0(k)$}$$ Comme  $ \mathcal{E}^0(k)$  est un faisceau fin, on obtient (3) et (4) pour  $ i = 0 .$ Pour  $i = 1$  on considère la suite exacte de faisceaux : $$0 \rightarrow \delta\mathcal{E}^0(k) \rightarrow Ker\,\delta^1(k) \rightarrow h^1(k) \rightarrow 0 . \tag{$B_1(k)$}$$ Elle donne  $ H^q(S^*, Ker\,\delta^1(k)) = 0 , \forall q \geq 2 $  ainsi que l’annulation de\ $H^p(S, Ker\,\delta^1(k))$  et de  $H^p(Y, Ker\,\delta^1(k)), \ \forall p \geq 1, $  grace à (3) et (4)\ pour  $i = 0$. La suite exacte de faisceaux : $$0 \rightarrow Ker\,\delta^1(k) \rightarrow \mathcal{E}^1(k) \rightarrow \delta\mathcal{E}^1(k) \rightarrow 0 \tag{$A_1(k)$}$$ donne alors les annulations (3) et (4) pour  $i = 1$  puisque  $\mathcal{E}^1(k) $  est fin. Supposons que l’on a montré les annulations (3) et (4) pour  $ i \in [1, n-3]$ ; nous allons en déduire ces mêmes annulations pour  $i+1$, ce qui prouvera par récurrence les annulations (3) et (4) pour  $i\in [0,n-2]$. Considèrons la suite exacte de faisceaux : $$0 \rightarrow \delta \mathcal{E}^i(k) \rightarrow \mathcal{E}^{i+1}(k) \rightarrow \delta \mathcal{E}^{i+1}(k) \rightarrow 0 .$$ Elle permet de conclure aux annulations (3) et (4) pour  $i+1$  grace à la finesse du faisceau  $ \mathcal{E}^{i+1}(k) .$ Pour prouver les annulations (4) pour  $i \geq n-1 $  on utilise successivement les suites exactes : $$\begin{aligned} & 0 \rightarrow \delta\mathcal{E}^{q-1}(k) \rightarrow Ker\,\delta^{q}(k) \rightarrow h^{q}(k) \rightarrow 0 \tag{$B_q(k)$} \\ & 0 \rightarrow Ker\,\delta^{q}(k) \rightarrow \mathcal{E}^{q}(k) \rightarrow \delta\mathcal{E}^{q}(k) \rightarrow 0 \tag{$A_q(k)$}\end{aligned}$$ pour  $q \geq n-1 $. $\hfill \blacksquare $ **Lemme.** ---------- Sous les hypothèses standarts avec  $n \geq 4$   et  $k \gg1$  la flèche naturelle $$e_k : = \frac{H^0(S^*, Ker\,\delta^{n-1}(k))}{H^0(S, Ker\,\delta^{n-1}(k)) + \delta H^0(S^*, \mathcal{E}^{n-2}(k))} \rightarrow H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0))$$ est un isomorphisme d’espaces vectoriels monodromiques. Pour  $n = 3 $  la flèche analogue est surjective et de noyau isomorphe à  $H^1(S^*, h^1(k)).$ Donc on obtient encore un isomorphisme monodromique, quitte à passer à la limite inductive sur  $k$[^28]. Pour  $n = 2 $  la flèche analogue, après passage à la limite inductive, est encore surjective et a un noyau isomorphe à  $H^1(S^*, \mathbb{C})$. [*Preuve.*]{} Comme on a supposé  $n \geq 3 $  l’application $$r^{n-1}(k) : h^{n-1}(k) \rightarrow H^{n-1}(0)$$ est un isomorphisme pour  $ k \gg 1 $  d’après (1.6). La suite exacte $$0 \rightarrow \delta\mathcal{E}^{n-2}(k) \rightarrow Ker\,\delta^{n-1}(k) \rightarrow h^{n-1}(k) \rightarrow 0 \tag{$B_{n-1}(k)$}$$ donne, puisque  $ H^1(S^*, \delta \mathcal{E}^{n-2}(k)) = 0 $  d’après (4.7), la surjectivité de la flèche  $q : H^0(S^*, Ker\,\delta^{n-1}(k)) \rightarrow H^0(S^*, h^{n-1}(k)) $, d’où une flèche naturelle et surjective $$H^0(S^*, Ker\,\delta^{n-1}(k)) \rightarrow H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0))$$ puisque l’on a $$H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) \simeq \frac{H^0(S^*, H^{n-1}(0))}{H^0(S, H^{n-1}(0))}$$ grâce à l’annulation de  $ H^1(S, H^{n-1}(0)) $[^29]. Il nous reste donc à identifier le noyau de cette surjection pour conclure la preuve. Comme les espaces  $ H^0(S, Ker\,\delta^{n-1}(k))$  et  $\delta H^0(S^*, \mathcal{E}^{n-2}(k))$  ont des images qui sont clairement dans le noyau, il suffit de montrer que tout élément  $ \alpha \in H^0(S^*, Ker\,\delta^{n-1}(k))$  dont l’image dans  $ H^0(S^*, h^{n-1}(k)) $  se prolonge à $H^0(S, H^{n-1}(0))$  est de cette forme. Comme l’application $$H^0(S, Ker\,\delta^{n-1}(k)) \rightarrow H^0(S, h^{n-1}(k))$$ est surjective[^30] il existe  $\beta \in H^0(S, Ker\,\delta^{n-1}(k))$  dont l’image dans  $H^0(S, h^{n-1}(k)) $  prolonge à  $S$  l’image de  $\alpha$  dans  $H^0(S^*, h^{n-1}(k)) $. Donc la restriction de  $\beta$  à  $S^*$  est égale à  $\alpha$  modulo le noyau  $H^0(S^*, \delta\mathcal{E}^{n-2}(k))$. On conclut pour  $n \geq 4$  en remarquant que l’égalité  $ Ker\,\delta^{n-2}(k) = \delta \mathcal{E}^{n-3}(k)$  (qui est vraie pour  $n \geq 4$)  donne, grace à l’annulation de  $ H^1(S^*, \delta \mathcal{E}^{n-3}(k)) $  montrée au (4.7), la surjectivité de l’application naturelle $$\delta : H^0(S^*, \mathcal{E}^{n-2}(k)) \rightarrow H^0(S^*, \delta \mathcal{E}^{n-2}(k))$$ d’où notre assertion. Pour  $n = 3 $  la suite exacte de cohomologie de la suite exacte de faisceaux : $$0 \rightarrow Ker\,\delta^1(k) \rightarrow \mathcal{E}^{1}(k)) \rightarrow \delta \mathcal{E}^{1}(k))\rightarrow 0 \tag{$A_1(k)$}$$ montre que le conoyau de la flèche $$\delta : H^0(S^*, \mathcal{E}^{1}(k)) \rightarrow H^0(S^*, \delta \mathcal{E}^{1}(k))$$ est égal à  $ H^1(S^*, Ker\,\delta^1(k))$  qui est isomorphe à   $H^1(S^*, h^1(k))$  comme on le constate en utilisant la suite exacte longue de cohomologie de la suite exacte : $$0 \rightarrow \delta\mathcal{E}^0(k) \rightarrow Ker\,\delta^1(k)) \rightarrow h^1(k) \rightarrow 0 \tag{$B_1(k)$}$$ et (4.7). D’où notre assertion pour  $n = 3$, puisque  $ \underset{k \rightarrow \infty} {\lim}h^1(k) = 0 $  dans ce cas. Pour  $n = 2 $   on a  $ \underset{k \rightarrow \infty}{lim} \ h^1(k) \simeq H^1(0) $  et la suite exacte $$0 \rightarrow \delta\mathcal{E}^0(k) \rightarrow Ker\,\delta^1(k)\rightarrow h^1(k) \rightarrow 0$$ donne la surjectivité de  $ H^0(S^*, Ker\,\delta^1(k)) \rightarrow H^0(S^*, h^1(k)) $  puisque $$H^1(S^*, \delta\mathcal{E}^0(k)) \simeq H^2(S^*, h^0(k)) \simeq 0 .$$ Donc la flèche  $\underset{k \rightarrow \infty}{lim} \ (e_k)$  est surjective. Son noyau va être égal à la limite inductive du quotient $$\frac{H^0(S^*, \delta\mathcal{E}^0(k))}{\big(H^0(S, Ker\,\delta^1(k)) + \delta H^0(S^*, \mathcal{E}^0(k)) \big)\cap H^0(S^*, \delta\mathcal{E}^0(k)) } .$$ Mais on a  $ H^0(S, Ker\,\delta^1(k)) \cap H^0(S^*, \delta\mathcal{E}^0(k)) \simeq H^0_{\lbrace 0 \rbrace}(S, h^1(k)) $  qui a une limite inductive nulle[^31]; la limite inductive du dénominateur co[ï]{}ncide donc avec celle de  $\delta H^0(S^*, \mathcal{E}^0(k))$. La suite exacte $$0 \rightarrow h^0(k) \rightarrow \mathcal{E}^0(k)) \rightarrow \delta\mathcal{E}^0(k)) \rightarrow 0$$ montre alors que le noyau est isomorphe à $$\underset{k \rightarrow \infty} \lim \ \frac{ H^0(S^*, \delta\mathcal{E}^0(k))}{\delta H^0(S^*, \mathcal{E}^0(k))} \simeq \underset{k \rightarrow \infty} \lim H^1(S^*, h^0(k)) \simeq H^1(S^*, \mathbb{C}).$$ D’où notre assertion pour  $n = 2 $. $ \hfill \blacksquare$ Construction de l’application  $i$. ----------------------------------- Comme nous supposons  $n \geq 3$  il nous suffit, d’après le lemme 4.8 , de construire une application linéaire monodromique $$H^0(S^*, Ker\,\delta^{n-1}(k)) \rightarrow H^n_{c\,\cap \,S}(0)$$ et de vérifier que son noyau co[ï]{}ncide avec  $ \delta H^0(S^*, \mathcal{E}^{n-2}(k)) + H^0(S, Ker\,\delta^{n-1}(k))$[^32]  et que son image co[ï]{}ncide avec  $ Ker\,can_{c\,\cap\,S} $. Soit donc  $\alpha \in H^0(S^*, Ker\,\delta^{n-1}(k))$  et fixons une fonction  $\chi \in \mathcal{C}^{\infty}(X) $  qui vérifie  $ \chi \equiv 1 $  près de  $ S \setminus B(0,r) $  et qui est nulle dès que l’on s’éloigne de  $ S \setminus B(0,r) $  de sorte que  $ Supp\,( d\chi )\cap S^* $  soit compact. Posons alors  $ i(\alpha) : = d\chi\wedge \alpha = \delta(\chi.\alpha)$. Il est immédiat de vérifier que la classe ainsi définie dans  $H^n_{c\,\cap \,S}(0) $  ne dépend ni de  $r \in ]0,\varepsilon[$[^33] ni du choix de  $\chi$. De plus, si on a  $\alpha = \delta\beta$, avec  $\beta \in H^0(S^*, \mathcal{E}^{n-2}(k))$, on aura $$\delta (\chi.\alpha) = \delta(-d\chi\wedge \beta)$$ et  $ d\chi\wedge \beta $  est à support dans  $ c\,\cap\,S $. Si  $\alpha$  est la restriction à  $S^*$  de  $\tilde{\alpha} \in H^0(S, Ker\,\delta^{n-1}(k))$  on aura  $ \delta(\chi.\alpha) = \delta((1-\chi).\tilde{\alpha}) $  et  $(1-\chi).\tilde{\alpha} $  est à support dans  $ c\,\cap\,S $. Ceci montre que l’application  $i$  est bien définie pour  $n \geq 3$. Elle est monodromique car on a $$d\chi \wedge \mathcal{N}(\alpha) = \mathcal{N}(d\chi\wedge\alpha).$$ Pour montrer l’injectivité de  $i$, considérons  $\alpha \in H^0(S^*, Ker\,\delta^{n-1}(k))$  tel que $i(\alpha) = \delta(\chi.\alpha) = \delta \beta$  avec  $\beta \in \Gamma_{c\,\cap\,S}(Y, \mathcal{E}^{n-1}(k)) $[^34]. Alors   $ \chi.\alpha -\beta$  est dans  $ \Gamma(Y, Ker\,\delta^{n-1}(k)) $  et on obtient ainsi un prolongement de  $\alpha$  à  $H^0(S, Ker\,\delta^{n-1}(k))$. Pour identifier l’image de  $i$  commen[ç]{}ons par remarquer que  $ can_{c\,\cap\,S}\circ i = 0 $. En effet, par définition,  $i(\alpha) = \delta(\chi.\alpha)$  et  $ \chi.\alpha \in \Gamma(Y, \mathcal{E}^{n-1}(k))$. D’où l’inclusion de l’image de  $i$  dans  $ Ker\,can_{c\,\cap\,S}$. Réciproquement, soit  $ u$  dans  $ \Gamma(Y, \mathcal{E}^{n-1}(k)) $   telle que le support de  $\delta u $  soit dans  $ c\, \cap\,S $. On a donc un compact  $ K$   de  $S$  tel qu’au voisinage de  $S \setminus K$   on ait  $\delta u = 0$. Alors  $u$  définit un élément de $$H^0(S\setminus K, h^{n-1}(k)) \simeq H^0(S^*, H^{n-1}(0))$$ pour  $ k \gg 1 $. L’image par  $i$  de cette classe est représentée par  $ \delta(\chi.u) .$ Mais  $ \delta u - \delta( \chi.u) = \delta((1-\chi).u) $  et on a  $ (1-\chi).u \in \Gamma_{c\,\cap\,S}(Y, \mathcal{E}^{n-1}(k))$ , ce qui montre que la classe initiale  $\delta u$  est bien dans l’image de  $i$. Le noyau de l’application  $\theta$. ------------------------------------ Rappelons maintenant comment calculer de l’application  $\theta.$ D’après \[B.91\] p. 428 on a la suite exacte de faisceaux suivante, pour  $k \gg1$ $$h^{n-2}(k) \overset{\tau_k}{\rightarrow} h^{n-1}(k) \overset{\tau_k}{\rightarrow} h^{n}(k) \rightarrow H^n(0) \rightarrow 0$$ Soit  $ w \in \Gamma(Y, Ker\,\delta^n(k))$. On peut donc écrire, localement le long de  $S^*$, puisque le faisceau  $H^n(0)$  est concentré à l’origine, $$w\vert_{X_{\sigma}} = \tau_{k}(\alpha_{\sigma}) + \delta \beta_{\sigma}$$ avec  $ \delta \alpha_{\sigma}= 0 $ (comparer avec \[B.91\] p.439). On obtient, comme dans loc. cit. que les  $\alpha_{\sigma}$  se recollent[^35] pour donner une section sur  $S^*$  du faisceau  $h^{n-1}(k)$  dont l’image dans  $H^1_{\lbrace 0 \rbrace}(S, h^{n-1}(k)) $  est indépendante des choix effectués. En utilisant ce qui précède, on peut trouver  $ \alpha \in \Gamma(S^*, Ker\,\delta^{n-1}(k)) $  telle que  $w - \tau_k(\alpha) $  soit localement  $\delta-$exacte le long de  $S^*$. A partir des écritures locales $$w - \tau_k(\alpha)\vert_{X_{\sigma}} = \delta\beta_{\sigma}$$ on définit  $\theta(w)$  par la classe dans $$H^1(S^*, h^{n-1}(k))\simeq H^1(S^*, H^{n-1}(0))$$ pour  $ k \gg 1$, du 1-cocycle  $ \beta_{\sigma} - \beta_{\sigma'} $. On vérifie facilement que ceci est indépendant des choix effectués. Nous voulons montrer maintenant que le noyau de  $\theta$  est l’image de  $can_{c\,\cap\,S}$. D’abord si  $w$  a son support dans  $c\,\cap\,S$  on peut choisir pour chaque  $ \sigma \in S \setminus K , \alpha_{\sigma} = 0 $  ainsi que   $ \beta_{\sigma} = 0 .$ La section  $\alpha$  sera donc nulle sur  $S\setminus K$  et comme  $H^{n-1}(0) $  est un système local sur  $S^*$, on aura  $\alpha = 0 $  et donc  $\theta[w]$  sera nul dans  $ H^1(S\setminus K, H^{n-1}(0)) $. A nouveau le fait que  $H^{n-1}(0)$  soit un système local sur  $S^*$  donne la nullité de  $\theta[w]$. Réciproquement, supposons que  $\theta[w] = 0 $. Utilisons maintenant un  $k$  assez grand pour assurer que  $\tilde{ob}_k \equiv 0 $[^36]. Cela signifie que l’on peut écrire globalement sur  $S^*$ $$w = \delta\beta$$ avec  $ \beta \in \Gamma(S^*, \mathcal{E}^{n-1}(k))$. Soit $\chi\in\mathcal{C}^{\infty}(Y)$  vérifiant  $\chi \equiv 1 $  près de  $S \setminus B(0,r)$  et s’annulant identiquement dès que l’on s’éloigne de  $S \setminus B(0,r)$. Alors  $ w - \delta(\chi.\beta) $  est un représentant de la classe de  $w$  qui est à support dans  $ c\,\cap\,S $. Donc cette classe est bien dans l’image de  $can_{c\,\cap\,S}$. Ceci achève la preuve du théorème 4.6. $\hfill \blacksquare$ Une conséquence importante mais immédiate pour  $n \geq 3$  du théorème 4.6 et du corollaire 3 du théorème 3.3.1 est l’égalité de la codimension de  $Im (\theta) $  dans l’orthogonal pour  $\tilde{h}$  de  $H^0(S, H^{n-1}(0)) $ avec  $\dim H^n_{c\, \cap\,S}(0) - \dim H^n(0)$. Ceci montre déja que la dimension de l’espace vectoriel  $ H^n_{c\, \cap\,S}(0)$  est au moins égale à celle de  $H^n(0)$. Notre objectif va être maintenant de montrer que l’on a en fait égalité de ces deux dimensions. Ceci n’est pas évident et sera obtenu via la construction d’une application de variation injective (pour  $n \geq 3$  au moins) commutant aux monodromies : $$var : H^n_{c\,\cap\,S}(0) \rightarrow H^n_c(0) .$$ On en déduira alors que  $ \dim H^n_{c\, \cap\,S}(0) \leq \dim H^n_c(0) = \dim H^n(0) $  pour  $n \geq 3$   cette dernière égalité résultant de la dualité de Poincaré sur la fibre de Milnor de  $f$  à l’origine. On en conclura, toujours pour  $n \geq 3$, que, de plus, cette application de variation est bijective, ce qui sera la clef de la non-dégénérescence de la forme hermitienne canonique “généralisée” : $$\mathcal{H} : H^n_{c\,\cap\,S}(0) \times H^n(0) \rightarrow \mathbb{C}$$ qui sera définie plus loin.\ Ceci montrera également que  $Im(\theta)$  co[ï]{}ncide avec l’orthogonal dans  $ H^1(S^*, H^{n-1}(0))$  pour  $\tilde{h}$  de  $H^0(S, H^{n-1}(0))$. Le cas  $n =2$. --------------- ### Définissons sous les hypothèses standards pour la valeur propre 1 et pour  $n =2$ $$\mathcal{K} : = \underset{k \rightarrow\infty}{lim} \ \frac{H^0(S^*, Ker\,\delta^{1}(k))}{H^0(S, Ker\,\delta^{1}(k)) + \delta H^0(S^*, \mathcal{E}^{0}(k))} .$$ Il résulte du Lemme (4.8) que l’on a une suite exacte monodromique : $$0 \rightarrow H^1(S^*, \mathbb{C}) \overset{j}{\rightarrow} \mathcal{K} \overset{e}{\rightarrow} H^1_{\lbrace 0 \rbrace}(S, H^1(0)) \rightarrow 0$$ obtenue par passage à la limite inductive sur  $k$. On remarquera que le sous-espace  $H^1(S^*, \mathbb{C}) $  est dans la partie invariante par la monodromie de  $\mathcal{K}$. Ceci résulte du fait que la monodromie agit comme l’identité sur les faisceaux  $h^0(k)$. ### Théorème 1, cas  $n = 2$. *Sous les hypothèses standarts pour la valeur propre 1, on a pour  $n = 2$  la suite exacte monodromique : $$0 \rightarrow \mathcal{K} \overset{i}{\rightarrow} H^2_{c\,\cap\,S}(0)\overset{can_{c\,\cap\,S}}{ \longrightarrow} H^2(0) \overset{\theta}{\rightarrow} H^1(S^*, H^1(0))$$ où l’application  $i$  est déduite du lemme (4.8).* La preuve est tout à fait analogue à celle donnée dans le cas  $n \geq 3$  modulo les adaptations que l’on vient de faire pour tenir compte du résultat différent du lemme (4.8) dans ce cas, et modulo les modifications suivantes : - Pour montrer que  $ Ker\,can_{c\,\cap\,S} \subset Im(i) $  on considère  $ u \in \Gamma(Y, \mathcal{E}^{1}(k)) $  qui est telle que le support de  $\delta u $  soit dans  $ c\, \cap\,S $. On a donc un compact  $ K$   de  $S$  tel qu’au voisinage de  $S \setminus K$   on ait  $\delta u = 0$. Alors  $u$  définit un élément de  $ H^0(S\setminus K, Ker\,\delta^{1}(k))$. Comme on a $$H^0(S\setminus K, H^1(0)) \simeq H^0(S^*, H^1(0))$$ son image dans  $H^0(S\setminus K, H^1(0))$   donne une section de  $H^0(S^*, H^1(0))$  ce qui nous fournit une section  $ v \in H^0(S^*, h^1(k))$  qui prolonge l’image de  $u$  dans  $H^0(S\setminus K, h^1(k))$  pour  $k \gg1 $. Comme on a  $H^1(S^*, \delta\mathcal{E}^0(k)) = 0 $  d’après le lemme (4.8), on en déduit l’existence de  $u_1 \in H^0(S^*, Ker\, \delta^1(k))$  dont la restriction à  $ S\setminus K $  est égale à  $u$. Comme on a  $H^0(S\setminus K, h^0(k)) \simeq H^0(S^*, h^0(k))$  on peut modifier  $u_1$  pour que sa restriction à  $ S\setminus K $  soit exactement égale à  $u$. On conclut alors comme dans le cas  $n \geq 3.$ - Pour l’étude du noyau de  $\theta$, on doit remplacer la suite exacte considérée pour  $ n \geq 3$   par la suite exacte : $$0 \rightarrow \frac{h^1(k)}{\tau_k h^0(k)} \rightarrow h^2(k) \rightarrow H^2(0) \rightarrow 0 .$$ et utiliser l’isomorphisme sur  $S^* : \frac{h^1(k)}{\tau_k h^0(k)} \simeq H^1(0) $. Localement près de  $\sigma \in S^*$  on a encore une écriture $$w\vert_{X_{\sigma}} - \tau_k(\alpha_{\sigma}) = \delta\beta_{\sigma}$$ avec  $\delta(\alpha_{\sigma}) = 0$. Mais maintenant le recollement des  $\alpha_{\sigma}$  n’a lieu sur  $S^*$  que dans le quotient  $\frac{h^1(k)}{\tau_k h^0(k)}$. On ne peut donc, en général, trouver un élement  $\alpha \in \Gamma(S^*, Ker\,\delta^1(k)) $  tel que  $ w - \tau_k(\alpha)$  soit localement  $\delta-$exact le long de  $S^*$. Mais ceci ne change rien à la preuve de l’inclusion  $ Im(can_{c\,\cap\,S})\subset Ker\,\theta .$ - Pour voir que  $ Ker\,\theta \subset Im(can_{c\,\cap\,S})$  l’argument permettant d’écrire $ w = \delta \beta$  globalement près de  $S^*$  avec  $\beta \in \Gamma(S^*, \mathcal{E}^1(k))$ n’est plus correct. L’annulation de  $\tilde{ob}_k $  pour  $k \gg 1$  reste valable, mais donne seulement l’existence, pour  $k \gg 1$, de  $\gamma_{\sigma} \in \Gamma(X_{\sigma}, Ker\,\delta^0(k))$, vérifiant  $\alpha_{\sigma} = \tau_k( \gamma_{\sigma})$  dans  $\Gamma(X_{\sigma}, Ker\,\delta^1(k))$  quitte à modifier le choix des  $\beta_{\sigma}$. Mais quitte à changer  $k$  en  $k+1$  en appliquant  $j_{k,k+1}$, on rend  $\tau_k(\gamma_{\sigma}) \ \delta-$exact, et donc  $j_{k,k+1}(w)$  devient localement  $\delta-$exacte le long de  $S^*$. Alors l’argument du cas  $n \geq 3$  s’applique et permet de conclure à l’égalité  $ Ker\,\theta = Im(can_{c\,\cap\,S}).$ ### Montrons que dans le cas  $n = 2$  le théorème 1 permet de prouver une inégalité de dimension analogue à celle du cas  $ n \geq 3$[^37] $$\dim_{\mathbb{C}} \ H^2_{c\,\cap\,S}(0) \geq \dim_{\mathbb{C}}\ H^2(0) + \dim_{\mathbb{C}}\ H^1(S^*, \mathbb{C}).$$ En effet la suite exacte du théoreme donne $$\dim\, H^2_{c\,\cap\,S}(0) + \dim\, Im(\theta) = \dim\, \mathcal{K} + \dim\, H^2(0)$$ et la suite exacte du (4.12.1) donne $$\dim\, H^1(S^*, \mathbb{C}) + \dim\, H^1_{\lbrace 0 \rbrace}(S, H^1(0)) = \dim\, \mathcal{K} .$$ L’inclusion de  $Im(\theta)$  dans l’orthogonal pour  $\tilde{h}$  de  $ H^0(S, H^1(0))$  donne, puisque  $H^0(S^*, H^1(0))$  est le dual de  $ H^1(S^*, H^1(0))$, l’inégalité $$\dim\, Im(\theta) + \dim\, H^0(S, H^1(0)) \leq \dim\, H^1(S^*, H^1(0)) .$$ D’où notre assertion, puisque $$H^1_{\lbrace 0 \rbrace}(S, H^1(0)) \simeq \frac{H^0(S^*, H^1(0))}{H^0(S, H^1(0))}.$$ La suite exacte longue de J. Leray généralisée. =============================================== Le premier ingrédient dans la construction de notre variation sera une généralisation encore un peu plus sophistiquée que celle de \[B.97\] du résidu de J. Leray. Théorème. ---------- *Soit  $Z$  une variété complexe connexe de dimension  $n+1$  et soit  $Y $  une hypersurface fermée d’intérieur vide dans  $Z$. Soit  $S$  un sous-ensemble analytique fermé et d’intérieur vide de  $Y$  tel qu’en chaque point  $y$  de  $Y\setminus S$, l’hypersurface  $Y$  admette une équation locale réduite  $f_y$  dont la monodromie n’admet pas la valeur propre 1 (dans son action sur la cohomologie réduite de la fibre de Milnor de  $f_y$  en  $y$ ).* *Notons par  $mod\,S $  la famille (paracompactifiante) des fermés de  $Z$  qui ne rencontrent pas  $S$.* *Alors on a la suite exacte longue “ de Leray” à supports  $mod\,S$ :* $$\cdots \rightarrow H^{q}_{mod\,S}(Z, \mathbb{C}) \overset{i^q}{\rightarrow} H^{q}_{mod\,S}(Z\setminus Y, \mathbb{C}) \overset{Res^q}{\rightarrow} H^{q-1}_{mod\,S}(Y, \mathbb{C}) \overset{\partial^q}{\rightarrow} H^{q+1}_{mod\,S}(Z, \mathbb{C}) \rightarrow \cdots$$ *Si on dispose d’une équation glabale réduite  $f \in \mathcal{O}(Z)$  de  $Y$  dans  $Z$, pour toute classe  $[w] \in H^q_{mod\,S}(Z\setminus Y, \mathbb{C})$  représentée par une forme semi-méromorphe d-fermée  $w$, à pôles dans  $Y$  et à support dans  $mod\,S$, il existe un voisinage ouvert  $\mathcal{W}$  de  $Y$  dans  $Z$  et des formes  $\eta \in \mathcal{C}^{\infty}(\mathcal{W})^{q-1} $  et  $\omega \in \mathcal{C}^{\infty}(\mathcal{W})^q$   d-fermées sur  $\mathcal{W}$  à supports dans  $mod\,S \vert_{\mathcal{W}}$  et une forme semi-méromorphe  $\alpha$  de degré  $q-1$  à pôles dans  $Y$  et à support dans  $mod\,S \vert_{\mathcal{W}}$  vérifiant :* $$w =\frac{df}{f} \wedge \eta + \omega + d\alpha \quad {\rm sur} \quad \mathcal{W}\setminus Y \tag{@}$$ *La classe  $ [\eta\vert_{Y}] \in H^{q-1}_{mod\,S}(Y, \mathbb{C})$  est alors l’image par  $Res^q$  de la classe* $$[w] \in H^q_{mod\,S}(Z \setminus Y, \mathbb{C}).$$ Remarque. --------- Le seul point non trivial dans la suite exacte ci-dessus est bien sur l’identification du groupe de cohomologie à support  $ H^{q+1}_{Y, mod\,S}(Z, \mathbb{C}) $  avec le groupe  $H^{q-1}_{mod\,S}(Y, \mathbb{C}) $  qui est analogue à ce qui se passe dans le cas d’une hypersurface lisse. La présence de la famille de supports  $ mod\,S $   est relativement anodine puisqu’elle est paracompactifiante et que sa restriction à  $Y$  est constituée des fermés de  $Y$  qui sont dans $ mod\,S $. On notera cependant que cet énoncé peut s’appliquer en présence d’un lieu singulier pour l’hypersurface  $Y$  qui est bien plus gros que  $S$. Démonstration. -------------- La preuve repose sur deux faits : - 1\) Le quasi-isomorphisme, dû à A. Grothendieck (voir \[Gr. 65\]) : $$Rj_*j^* \underline{\mathbb{C}}_Z \simeq ( \mathcal{C}^{\infty}_Z[*Y]^{\bullet}, d^{\bullet} )$$ qui montre que le complexe des formes semi-méromorphes à pôles dans  $Y$  donne une incarnation “fine” du complexe  $ Rj_*j^* \underline{\mathbb{C}}_Z$ . - 2\) La dégénérescence de la suite spectrale : $$E_2^{p,q} : = H^q_{mod\,S}(Z, \underline{H}_Y^p(\mathbb{C}))$$ qui résulte du fait que, sur  $ Z\setminus S $, on a  $ \underline{H}^p_Y(\mathbb{C}) = 0 $   pour  $p \not= 2 $  et  $ \underline{H}^2_Y(\mathbb{C}) \simeq \underline{\mathbb{C}}_Y .\frac{df}{f} $ , combiné avec le fait que si le faisceau  $\mathcal{F}$  est nul sur  $Z\setminus S$  alors on a  $ H^p_{mod\,S}(Z, \mathcal{F}) = 0 , \forall p \geq 0 $. Le calcul de  $ \underline{H}^p_Y(\mathbb{C}) $  sur  $Z\setminus S$  vient de l’absence de la valeur propre 1 pour la monodromie locale (voir la remarque finale de \[B.97\] ) d’une équation locale réduite  $f$  de  $Y$, qui donne l’annulation des groupes  $ H^i(D^*, R^j f_*(\mathbb{C}_Z)) $  pour tout  $ i \geq 0 $   quand  $ j \not= 0$. La suite exacte longue de Leray en résulte alors d’après \[Go.58\] th. 4.10.1 puisque la dégénérescence de la suite spectrale donne les isomorphismes $$H^{q+1}_{Y, mod\,S}(Z, \mathbb{C}) \simeq H^{q-1}_{mod\,S}(Y, \mathbb{C}) .$$ Pour prouver  $(@)$  représentons  $Res^q[w] \in H^{q-1}_{mod\,S}(Y, \mathbb{C}) $ par une forme  $\eta \in \mathcal{C}^{\infty}$  de degré  $q-1$  d-fermée à support dans  $mod\,S$[^38] sur un voisinage ouvert  $\mathcal{W}$  de  $Y$  assez petit. Alors la classe  $ [w - \frac{df}{f} \wedge \eta ] \in H^q_{mod\,S}(\mathcal{W}\setminus Y, \mathbb{C}) $  aura un résidu nul d’après la suite exacte de Leray sur  $\mathcal{W}$. On en déduit l’existence de  $ \omega \in \mathcal{C}^{\infty}_{mod\,S}(\mathcal{W})^q $  vérifiant  $ d\omega = 0 $  et dont la restriction à  $Z\setminus Y$  induit la classe  $ [w - \frac{df}{f} \wedge \eta ] \in H^q_{mod\,S}(\mathcal{W}\setminus Y, \mathbb{C}) $. On en alors déduit l’existence de  $\alpha$  donnant  $(@)$. $\hfill \blacksquare$ Construction de  $ \widetilde{var} : H^n_{c\,\cap\,S}(0) \rightarrow H^n_c(0) .$ ================================================================================= Considérons donc, pour  $ k \geq 1 $  donné,  $w\in\Gamma_{c\,\cap\,S}(Y, \mathcal{E}^n(k)) \cap Ker\,\delta $. Notons par  $X'$  la trace sur  $X$  d’une boule centrée à l’origine et de rayon  $\varepsilon' < \varepsilon$  mais assez proche de  $\varepsilon$  pour avoir  $ Supp(\,w) \cap\, S \subset Y' $  où  $ Y' : = Y \cap X' $. Notons alors par  $ Z$  un voisinage ouvert de  $ Y \setminus \bar{Y}' $. Sur  $Z$, le morphisme de complexe $$( \mathcal{F}^{\bullet}_{mod\,S}(k), \delta^{\bullet}) \rightarrow (\mathcal{E}^{\bullet}_{mod\,S}(k), \delta^{\bullet})$$ est un quasi isomorphisme[^39]. Comme la monodromie agit comme l’identité[^40] sur les faisceaux de cohomologies du complexe  $( \mathcal{F}^{\bullet}_{mod\,S}(k), \delta^{\bullet}) $  on peut trouver  $ u\in\Gamma_{mod\,S}(Z, \mathcal{E}^{n-1}(k))$  vérifiant $$_k\mathcal{N}(w) = \delta u \quad {\rm sur} \quad Z .$$ Soit  $\rho\in\mathcal{C}^{\infty}_{c/f}(X)$  valant identiquement 1 au voisinage de  $\bar{X}'$. Alors posons  $\xi : = (1-\rho).u$  et  $v : = _k\mathcal{N}(w) - \delta \xi $. On a alors  $v \in \Gamma_{c/f}(Y, \mathcal{E}^n(k)) \cap Ker\, \delta $. Cela donne explicitement $$w_{j-1} = v_j + d\xi_j - \frac{df}{f}\wedge \xi_{j-1} \qquad \forall j\in [1,k]$$ avec la convention  $ w_0 = 0 = \xi_0 $. On aura en particulier, puisque  $\delta w = 0 $ $$d(w_k + \frac{df}{f}\wedge \xi_k) = \frac{df}{f}\wedge v_k .$$ Choisissons maintenant une boule centrée à l’origine de rayon  $ \varepsilon'' \in ]\varepsilon', \varepsilon[$  et assez proche de  $\varepsilon$  pour que la trace  $X''$  de cette boule sur  $X$  contienne  $Supp\, (v)$. Posons alors  $\tilde{Y} : = Y \setminus\bar{Y}'' $. Alors la forme semi-méromorphe  $w_k + \frac{df}{f}\wedge \xi_k$  de degré  $n$  est  $d-$fermée sur un voisinage ouvert assez petit  $\tilde{Z}$  de  $\tilde{Y}$  et a un support qui ne rencontre pas  $S \cap \tilde{Y} $. On peut donc lui appliquer le théorème (5.1) et l’écrire sur  $\tilde{Z}$, quitte à localiser autour de  $f^{-1}(0)$ : $$w_k + \frac{df}{f}\wedge \xi_k = \frac{df}{f}\wedge\eta + \omega + d\alpha \tag{@@}$$ où les formes  $\mathcal{C}^{\infty}$  $\eta$  et  $\omega$  sont  $\mathcal{C}^{\infty}$  sur un voisinage ouvert  $\mathcal{V}(\tilde{Y})$  de  $\tilde{Y}$  dans  $\tilde{Z}$, de degrés respectifs  $n-1$  et  $n$, vérifient : - [i)]{}  $d\eta = 0 , \ d\omega = 0 $ - [ii)]{} $ Supp\, \eta \cap S = \emptyset , \ Supp\, \omega \cap S = \emptyset .$ et où la  $(n-1)-$forme semi-méromorphe  $\alpha$  sur  $\mathcal{V}(\tilde{Y})$, est à pôles dans  $\tilde{Y}$  et à support  $mod\,S$. Soit maintenant  $\sigma\in\mathcal{C}^{\infty}_{c/f}(X)$  valant identiquement 1 au voisinage de  $\bar{X}'' $. Alors la forme semi-méromorphe de degré  $n+1$  à poles dans  $\tilde{Y}$  et à support  $f-$propre dans  $\tilde{Z}$ $$W : = d\sigma \wedge (w_k + \frac{df}{f}\wedge \xi_k)$$ est égale sur  $\tilde{Z}$, grace à (@@), à $$W = -\frac{df}{f}\wedge(d\sigma\wedge\eta) + d\sigma\wedge \omega - d(d\sigma\wedge\alpha)$$ où les formes   $d\sigma\wedge\eta$  et  $d\sigma \wedge\omega$  sont  $\mathcal{C}^{\infty}$  sur un voisinage ouvert  $\mathcal{V}(\tilde{Y})$  de  $\tilde{Y}$  dans  $\tilde{Z}$, de degrés respectifs  $n$  et  $n+1$, à supports  $f-$propres dans  $\mathcal{V}(\tilde{Y})$  vérifient “à fortiori” - [i)]{}  $d(d\sigma\wedge\eta) = 0 , \ d(d\sigma\wedge\omega) = 0 $ - [ii)]{} $ Supp(d\sigma\wedge\eta) \cap S = \emptyset , \ Supp(d\sigma\wedge\omega) \cap S = \emptyset .$ et où la  $n-$forme semi-méromorphe  $d\sigma\wedge\alpha$  sur  $\mathcal{V}(\tilde{Y})$, est à pôles dans  $\tilde{Y}$  et à support  $f-$propre $\cap\,mod\,S$  également. [**Nous définirons alors**]{} $$2i\pi. \widetilde{var}([w]) : = r^n_c(k)(\tilde{v}) \quad {\rm avec}$$ $$\tilde{v}_k : = v_k + d\sigma\wedge\eta, \quad {\rm et} \quad \tilde{v}_j : = v_j \quad \forall j\in[1,k-1].$$ Il reste évidemment à vérifier que tout ceci définit bien une application linéaire $$\widetilde{var} : H^n_{c\,\cap\,S}(0) \rightarrow H^n_c(0)$$ commutant aux monodromies. Commen[ç]{}ons par remarquer que si l’on change  $w$  par son image par  $j_{k,k+k'}$  le résultat de notre construction ne change pas. Supposons maintenant que  $ w = \delta (\beta)$  avec  $ \beta \in \Gamma_{c\,\cap\,S}(Y, \mathcal{E}^{n-1}(k)) $. Alors on peut choisir  $ u = \mathcal{N}( \beta)$  et on aura $$w_k + \frac{df}{f}\wedge \xi_k = d\beta_k - \rho \frac{df}{f}\wedge \beta_{k-1} .$$ Sur  $\tilde{Z}$  on aura donc  $ w_k + \frac{df}{f}\wedge \xi_k = d\beta_k $. Ceci montre que l’on peut prendre  $ \eta = 0 , \omega = 0 \ {\rm et} \ \alpha = \beta_k $. Comme on a  $ v = \delta(\rho.\mathcal{N}(\beta)) $  qui induit la classe nulle dans  $H^n_c(0)$, ceci montre bien que la modification de  $w$  par un cobord ne change pas le résultat de notre construction. On en déduit immédiatement que si  $ w = \tau_k (\gamma) $  avec  $ \gamma \in \Gamma_{c\,\cap\,S}(Y, Ker \delta^{n-1}(k)) $, alors on trouve  $0$  dans  $H^n_c(0)$. En effet, on peut remplacer  $w_k$  par  $ j_{k,k+k'}(w_k) $  d’après ce qui précède, et alors on aura d’après l’égalité (@) du (4.2) $$j_{k,k+k'}(\tau_k(\gamma)) = \tau_{k+k'}(_{k+k'}\mathcal{N}^{k'}(j_{k,k+k'}(\gamma)))$$ ce qui est dans  $ \delta \big(\Gamma_{c\,\cap\, S}(Y, \mathcal{E}^{n-1}(k+k')\big)$  pour  $ k' \geq k $. Les changements de choix des fonctions  $\rho$  et  $\sigma$  sont laissés en exercice au lecteur. Le changement du choix de  $\eta $  consiste, d’après le théorème (5.1) à remplacer  $\eta$  par  $ \eta + d\theta $  où  $ \theta \in \mathcal{C}^{\infty}_{mod\,S}(\tilde{Z})^{n-2} $. Mais on ajoute alors à  $ \tilde{v}$  le cobord $$\delta (j_{1,k}(d\sigma \wedge \theta )= j_{1,k}(-d\sigma \wedge d\theta)$$ ce qui ne change pas l’image dans  $H^n_c(0)$. Il nous reste à voir l’action de la monodromie. Si on part de  $_k\mathcal{N}(w) $  on peut prendre  $_k\mathcal{N}(u)$  et donc  $_k\mathcal{N}(\xi)$  dans notre construction. On doit alors prendre l’image de  $_k\mathcal{N}(\tilde{v}) =\ _k\mathcal{N}(v) $  dans  $H^n_c(0)$  ce qui donne bien le résultat attendu. En effet, l’égalité $$w_{k-1} + \frac{df}{f} \wedge \xi_{k-1} = v_k + d\xi_k$$ permet de voir que sur  $\tilde{Z}$, on a  $ w_{k-1} + \frac{df}{f} \wedge \xi_{k-1} = d\xi_k $. On peut donc choisir  $ \eta = 0 , \omega = 0 \ {\rm et} \ \alpha = \xi_k $. Ceci termine nos vérifications sur cette construction. La forme hermitienne canonique\ $ \mathcal{H} : H^n_{c\,\cap\,S}(0)\times H^n(0) \rightarrow \mathbb{C}.$ ========================================================================= Nous adopterons ici le point de vue de \[B.90\] qui est repris dans \[B.97\]. Théorème-Définition. -------------------- *Sous les hypothèses standards définissons la forme hermitienne canonique $$\mathcal{H} : H^n_{c\,\cap\,S}(0) \times H^n(0) \rightarrow \mathbb{C}$$ de la fa[ç]{}on suivante : pour  $ e \in H^n_{c\,\cap\,S}(0) $  et  $e' \in H^n(0)$  représentés par  $w \in \Gamma_{c\,\cap\,S}(Y, \mathcal{E}^n(k)) \cap Ker\,\delta $  et  $w' \in \Gamma(Y, \mathcal{E}^n(k)) \cap Ker\,\delta $  vérifiant  $ [w] = e $  et  $ r^n(k)(w') = e' $  posons $$\mathcal{H}(e,e') = \frac{1}{(2i\pi)^{n+1}} P_2\big(\lambda = 0 , \int_X \vert f \vert^{2\lambda} \rho.\frac{df}{f} \wedge w_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k \big)$$ où  $ \rho \in \mathcal{C}_c^{\infty}(X) $  vaut identiquement 1 au voisinage du compact  $ Supp\,w \,\cap\,S $.* *Notons par  $ \mathcal{I} : H^n_c(0) \times H^n(0) \rightarrow \mathbb{C} $  la dualité (hermitienne) de Poincaré sur la fibre de Milnor  $F$  de  $f$  l’origine, donnée par $$\mathcal{I}(a,b) : = \frac{1}{(2i\pi)^n} \int_F a \wedge \bar{b} .$$ Alors on a  $ \mathcal{H}(e,e') = \mathcal{I}(\widetilde{var}(e), e') $  où l’application $$\widetilde{var} : H^n_{c\,\cap\,S}(0) \rightarrow H^n_c(0)$$ a été construite plus haut.* Remarque. --------- Une conséquence immédiate du théorème (7.1) sera l’équivalence entre les deux propriétés suivantes : - La variation est bijective (ou bien  $\widetilde{var}$  est bijective). - La forme hermitienne canonique est non dégénérée. Nous prouverons la première de ces assertions pour tout  $n \geq 3$  à la fin du paragraphe 8, ce qui donnera donc également la non dégénérescence de  $\mathcal{H}$. Pour  $ n = 2$  la forme hermitienne canonique  $ \mathcal{H}$  passe au quotient par le sous-espace (T-invariant)  $j(H^1(S^*, \mathbb{C})$  de  $H^2_{c\,\cap\,S}(0)$  et induit une dualité hermitienne entre   $H^2_{c\,\cap\,S}(0) \big/ j(H^1(S^*, \mathbb{C})$  et  $H^2_{c\,\cap\,S}(0)$. Donc, à nouveau, on a un résultat analogue au cas  $n \geq 3$  quitte à remplacer  $H^2_{c\,\cap\,S}(0)$  par son quotient par  $j(H^1(S^*, \mathbb{C})$. La démonstration du théorème (7.1) occupera le reste de ce paragraphe 7. Commen[ç]{}ons par vérifier que le nombre  $\mathcal{H}(e,e')$  est bien défini, c’est à dire est indépendant des divers choix effectués. - [Indépendance du choix de  $\rho$]{}\ Soit donc  $\rho'\in \mathcal{C}_c^{\infty}(X) $  valant aussi identiquement 1 au voisinage du compact  $ Supp\,w \,\cap\,S $. Alors on a  $Supp\,(\rho-\rho')\,\cap\,S = \emptyset $. Donc le prolongement méromophe de $$\int_X \vert f \vert^{2\lambda}( \rho- \rho').\frac{df}{f} \wedge w_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k$$ n’a, au pire, qu’un pôle simple en  $\lambda = 0 $, ce qui donne l’indépendance désirée. - [Indépendance du choix de  $w'$  représentant  $e'$]{}\ Il s’agit en fait de montrer que si on a  $ w' = \delta v' $  avec  $ v'\in \Gamma(Y, \mathcal{E}^{n-1}(k))$  alors on a $$P_2\big(\lambda = 0 , \int_X \vert f \vert^{2\lambda} \rho.\frac{df}{f} \wedge w_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{\delta v}'_{k} \big) = 0 .$$ Comme on a  $ \frac{df}{f} \wedge w'_k = - d(\frac{df}{f}\wedge v'_k )$  ceci va résulter de la formule de Stokes et du fait que  $Supp\,d\rho\,\cap\,Supp\,w\,\cap \, S = \emptyset $, le prolongement méromorphe de  $\vert f \vert^{2\lambda} $  sur  $ X \setminus S$  n’ayant, au pire, que des pôles simples aux entiers négatifs. - [Indépendance du choix de  $w$  représentant  $e$]{}\ Comme ci-dessus, il s’agit en fait de montrer que si on a  $ w= \delta v$  avec  $ v\in \Gamma_{c\,\cap\, S}(Y, \mathcal{E}^{n-1}(k))$  alors on a $$P_2\big(\lambda = 0 , \int_X \vert f \vert^{2\lambda} \rho.\frac{df}{f} \wedge\delta v_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{ w}'_{k} \big) = 0 .$$ La preuve est analogue au cas précédent, en choisissant la fonction  $\rho$  de fa[ç]{}on qu’elle soit identiquement égale à 1 sur un voisinage ouvert de  $Supp\,v\,\cap\,S $. Ceci est possible grace à la condition de support sur  $v$  et à l’indépendance du choix de  $\rho$  prouvée ci-dessus. Montrons maintenant l’invariance de  $\mathcal{H}$  par la monodromie, c’est à dire que l’on a $$\mathcal{H}(\mathcal{N}(e), e') = \mathcal{H}(e,\mathcal{N}(e')) \quad \forall e\in H^n_{c\,\cap\,S}(0), \quad \forall e'\in H^n(0)$$ où  $-2i\pi.\mathcal{N}$  désigne le logarithme (nilpotent) de la monodromie agissant sur  $H^n_{c\,\cap\,S}(0)$  ou bien sur  $ H^n(0) $. Compte tenu des indépendances de choix prouvés ci-dessus, ceci revient à établir la formule : $$\begin{aligned} & P_2\big(\lambda = 0 , \int_X \vert f \vert^{2\lambda} \rho.\frac{df}{f} \wedge w_{k-1} \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k \big) = \\ & P_2\big(\lambda = 0 , \int_X \vert f \vert^{2\lambda} \rho.\frac{df}{f} \wedge w_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_{k-1} \big) \end{aligned}$$ puisque  $_k\mathcal{N}(w)$  et respectivement  $_k\mathcal{N}(w')$  induisent  $\mathcal{N}(e)$  et  $\mathcal{N}(e')$  dans  $H^n_{c\,\cap\,S}(0)$  et  $H^n(0)$. Mais pour  $\Re(\lambda) \gg 0 $  on a $$\begin{aligned} d\big(\vert f \vert^{2\lambda}\rho.(\frac{df}{f} + \frac{\bar{df}}{\bar{f}})\wedge w_k \wedge \bar{w}'_k \big) = \\ \vert f \vert^{2\lambda}.d\rho \wedge(\frac{df}{f} + \frac{\bar{df}}{\bar{f}})\wedge w_k \wedge \bar{w}'_k \ + \\ -\vert f \vert^{2\lambda}\rho.(\frac{df}{f} + \frac{\bar{df}}{\bar{f}})\wedge dw_k \wedge \bar{w}'_k \ + \\ (-1)^{n+1} \vert f \vert^{2\lambda}\rho.(\frac{df}{f} + \frac{\bar{df}}{\bar{f}})\wedge w_k \wedge d\bar{w}'_k .\end{aligned}$$ Comme on a  $Supp\,d\rho\,\cap\,Supp\,w\,\cap \, S = \emptyset $  le prolongement méromorphe du premier terme du membre de droite de l’égalité ci-dessus n’aura, au pire, qu’un pôle simple en  $\lambda = 0$. Par ailleurs le fait que  $w$  et  $w'$  soit  $\delta-$fermées donne $$dw_k = \frac{df}{f}\wedge w_{k-1} \quad {\rm et} \quad dw'_k = \frac{df}{f}\wedge w'_{k-1} .$$ La formule de Stokes et le prolongement analytique donnent alors la formule désirée. Il nous reste à prouver la formule $$\mathcal{H}(e,e') = \mathcal{I}(\widetilde{var}(e), e')$$ pour achever la preuve du théorème (7.1). Reprenons les notations utilisées en (6.1) lors de la construction de  $\widetilde{var}(e)$. Pour  $\Re{\lambda} \gg 0$  on aura $$\begin{aligned} d\big((\vert f \vert^{2\lambda}\sigma.(w_k + \frac{df}{f}\wedge\xi_k) \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k \big) = & \ \lambda.\vert f \vert^{2\lambda}\sigma.\frac{df}{f} \wedge w_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k \\ & + \vert f \vert^{2\lambda} d\sigma \wedge (w_k + \frac{df}{f}\wedge\xi_k) \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k \\ & + \vert f \vert^{2\lambda} \frac{df}{f} \wedge v_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k \end{aligned}$$ puisque  $ d(w_k + \frac{df}{f}\wedge\xi_k) = \frac{df}{f} \wedge v_k $  et  $\sigma.v_k = v_k $. D’après la formule de Stokes, l’intégrale du membre de gauche n’aura aucun pôle, et, par prolongement analytique, la nullité du résidu en  $\lambda = 0 $  du membre de droite donnera la relation $$\begin{aligned} - P_2\big(\lambda = 0, \int_X \vert f \vert^{2\lambda}\sigma.\frac{df}{f} \wedge w_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k\big) = & \ Res\big(\lambda = 0, \int_X \vert f \vert^{2\lambda}W \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k\big) \\ + &Res\big(\lambda = 0, \int_X \vert f \vert^{2\lambda}\frac{df}{f} \wedge v_k \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k \big) .\end{aligned}$$ Mais on a $$W = -\frac{df}{f}\wedge(d\sigma\wedge\eta) + d\sigma\wedge \omega - d(d\sigma\wedge\alpha) .$$ Montrons qu’alors $$\begin{aligned} Res\big(\lambda = 0, \int_X \vert f \vert^{2\lambda}d\sigma\wedge \omega \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k\big) = 0 \\ Res\big(\lambda = 0, \int_X \vert f \vert^{2\lambda}d(d\sigma\wedge\alpha) \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k\big) = 0 .\end{aligned}$$ Pour le premier résidu c’est conséquence du fait que la forme  $\omega$  est  $\mathcal{C}^{\infty}$  et donc qu’il n’y a pas de puissance négative de  $f$  dans cette intégrale. On conclut en utilisant que les racines du polynôme de Bernstein de  $f$  sont strictement négatives (\[K.76\]). Pour montrer que le second résidu est nul, on utilise à nouveau la formule de Stokes et le prolongement analytique pour obtenir $$Res\big(\lambda = 0, \int_X \vert f \vert^{2\lambda}d(d\sigma\wedge\alpha) \wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k\big) = P_2\big(\lambda = 0, \int_X \vert f \vert^{2\lambda}\sigma. \frac{df}{f}\wedge d\alpha\wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k\big).$$ Mais comme le support de  $d\alpha$  ne rencontre pas  $S$  le membre de droite est nul. Il nous reste donc finalement l’égalité : $$(2i\pi)^{n+1}. \mathcal{H}(e,e') = - Res\big(\lambda = 0, \int_X \vert f \vert^{2\lambda}\frac{df}{f} \wedge (v_k + d\sigma \wedge \eta)\wedge \frac{\bar{df}}{\bar{f}} \wedge \bar{w}'_k \big) .$$ Mais la fonction $$s \rightarrow \big(\frac{i}{2\pi}\big)^n \int_{f=s} (v_k + d\sigma\wedge\eta)\wedge\bar{w}'_k$$ est un polynôme de degré au plus  $k-1$  en  $Log (\vert\frac{s}{s_0}\vert^2) $  dont le terme constant est  $ \mathcal{I}(\widetilde{var}(e),e')$. Par transformation de Mellin on obtient la formule désirée.  $\hfill \blacksquare$ Injectivité de la variation. ============================ Pour montrer que la variation est injective il nous suffit, d’après le théorème (7.1) de montrer que pour toute classe non nulle  $ e \in H^n_{c\,\cap\,S}(0)$  il existe  $e'\in H^n(0)$  telle que l’on ait  $\mathcal{H}(e,e') \not= 0 .$ En fait, comme la variation commute à la monodromie, on peut se contenter de montrer cela pour une classe  $e$  qui est invariante par la monodromie. Soit donc  $v \in\Gamma_{c\,\cap\,S}(Y, \mathcal{E}^n(k)) \cap Ker\,\delta $  définissant la classe  $e$  considérée. Comme on suppose que  $T(e) = e$, ce qui équivaut à  $\mathcal{N}(e) = 0 $, on pourra trouver, quitte à choisir  $k$  assez grand,  $ u\in\Gamma_{c\,\cap\,S}(Y, \mathcal{E}^{n-1}(k)) $  vérifiant  $ _k\mathcal{N}(v) = \delta u $  ce qui se traduit par les égalités : $$v_{j-1} = du_j - \frac{df}{f}\wedge u_{j-1}\quad \forall j\in[1,k] \qquad {\rm avec} \quad v_0 = u_0 = 0 .$$ On en déduit que l’on a  $ d\hat{w} = 0 $  où l’on a posé  $ \hat{w} = v_{k} + \frac{df}{f}\wedge u_{k-1} .$ De plus on a l’ égalité $$j_{1,k+1}(\hat{w}) = j_{k,k+1}(v) -\delta \tilde{u}$$ où l’on définit  $ \tilde{u}_{k+1} = 0$  et  $ \tilde{u}_j = u_j \quad \forall j \in [1,k] .$ Ceci montre que la classe  $e$  est représentée par la forme semi-méromorphe  $d-$fermée  $\hat{w}$  qui a son support dans  $c\,\cap\,S$. Remarquons par ailleurs que l’image dans  $H^n(0)$  de la classe  $e$   peut également être représentée par une forme méromorphe $d-$fermée  $ w \in \Gamma(Y, \Omega^n)$  grâce au quasi-isomorphisme  $(\Omega^{\bullet}(k), \delta^{\bullet}) \simeq (\mathcal{E}^{\bullet}(k), \delta^{\bullet}) $  pour  $k = 1$. On aura alors l’existence de  $\gamma \in \Gamma(Y, \mathcal{E}^{n-1}(1))$  vérifiant  $ w = \tilde{w} + d\gamma $  au voisinage de  $Y$. La généralisation suivante du résultat de contribution “sureffective” démontrée dans \[B.84 b\] va nous permettre de conclure dans le cas où l’image de la classe  $e$  dans  $H^n(0)$   est non nulle. Nous traiterons le cas où cette image est nulle à la fin du paragraphe 8 (voir (8.10) à (8.22)). Théorème. ---------- *Soit  $f : X \rightarrow D $  un représentant de Milnor d’un germe non constant de fonction holomorphe à l’origine de  $\mathbb{C}^{n+1}$. Soit  $ Y = f^{-1}(0)$  et soit  $S_1$  un sous-ensemble analytique fermé de  $Y$  tel qu’en chaque point  $y$  de  $Y\setminus S_1$  la monodromie locale de  $f$  en  $y$  agissant sur la cohomologie (réduite) de la fibre de Milnor de  $f$  en  $y$   ne présente pas la valeur propre 1.* *On suppose que l’on a  $ H^n(X\setminus S_1, \mathbb{C}) = 0 $.* Remarques. ----------- - [1)]{} La conclusion du théorème est, bien sur, indépendante du choix de la fonction  $\rho \in \mathcal{C}^{\infty}_c(X)$  valant identiquement  $1$  au voisinage de  $K$  puisque si  $\sigma \in \mathcal{C}^{\infty}_c(X)$  vaut identiquement  $0$  au voisinage de  $K$  le support de  $\sigma.\tilde{w} $  ne rencontre plus  $S_1$  ce qui implique que le prolongement méromorphe de  $ \vert f \vert^{2\lambda}.\sigma,\tilde{w} $  ne présente que des pôles d’ordre  $\leq 1 $  aux entiers. - [2)]{} Si la dimension du sous-ensemble analytique (fermé)  $S_1$  de Y est  $\leq p$  avec  $ 2p < n+1 $  alors la condition  $H^n(X\setminus S_1, \mathbb{C}) = 0 $  est automatiquement réalisée[^41]. Nous utiliserons essentiellement le cas  $ p = 1$  et  $n \geq 2 $  de ce théorème. *Démonstration.* Posons pour  $ j \in \mathbb{N} $ : $$\mathcal{T}^{n,0}_j : = \ Res(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} w \wedge \Box) \ + \ P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} \frac{df}{f}\wedge \gamma^{n-1,0}\wedge \Box )$$ et montrons déja que l’on a  $ d'\mathcal{T}^{n,0}_j = 0 $  sur  $ V : = X \setminus K $  pour tout  $j \in \mathbb{N}$. On a $$d' Res(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} w \wedge \Box )\ = \ P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge w \wedge \Box )$$ puisque  $ dw = d'w = 0 $. Par ailleurs $$d'P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} \frac{df}{f}\wedge \gamma^{n-1,0}\wedge \Box ) \ = \ - P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} \frac{df}{f}\wedge d' \gamma^{n-1,0}\wedge \Box )$$ ce qui donne notre assertion puisque les pôles aux entiers négatifs du prolongement méromorphe de  $\vert f \vert^{2\lambda} $  sont simples au plus le long de  $Y\setminus S_1$  et que l’on a  $ w - d'\gamma^{n-1,0} = \tilde{w}^{n,0} = 0 $  au voisinage de  $ V \cap S_1 $. Supposons maintenant que pour  $ j = n $  le courant  $\mathcal{T}^{n,0}_n$   soit  $d'-$exact sur  $V$. C’est à dire qu’il existe un courant  $ U^{n-1,0}_n $  de type  $(n-1,0)$  sur  $V$  vérifiant $$\mathcal{T}^{n,0}_n = d' U^{n-1,0}_n \quad {\rm sur} \quad V .$$ Alors, comme on a $$\bar{f}. \mathcal{T}^{n,0}_{j+1} = \mathcal{T}^{n,0}_j \quad \forall j \in \mathbb{N}$$ on obtient, en posant  $ U^{n-1,0}_j = \bar{f}^{n-j}.U^{n-1,0}_n \quad \forall j\in [1,n] $ $$\mathcal{T}^{n,0}_j = d' U^{n-1,0}_j \quad {\rm sur} \quad V \quad \forall j\in [1,n].$$ Montrons qu’alors les courants sur  $V$  définis pour  $ j \in [1,n-1] $ $$\begin{aligned} \mathcal{T}^{n-1,1}_j : = & \ d''U^{n-1,0}_j + j.d\bar{f}\wedge U^{n-1,0}_{j+1} \ + \\ & - P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} \frac{\bar{df}}{\bar{f}}\wedge \gamma^{n-1,0}\wedge \Box )\ + \\ & + P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} \frac{df}{f}\wedge \gamma^{n-2,1}\wedge \Box )\end{aligned}$$ sont  $d'-$fermés sur  $V$. Pour cela calculons  $ d''\mathcal{T}^{n,0}_j + j.d\bar{f}\wedge \mathcal{T}^{n,0}_{j+1} $. On trouve $$\begin{aligned} d'' \mathcal{T}^{n,0}_j + j.d\bar{f}\wedge \mathcal{T}^{n,0}_{j+1} = & \ P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge w \wedge \Box ) + \\ & + P_3(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \frac{df}{f}\wedge \gamma^{n-1,0} \wedge \Box ) + \\ & - P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge d'' \gamma^{n-1,0} \wedge \Box )\end{aligned}$$ Utilisons maintenant les relations  $ w = d'\gamma^{n-1,0} $   et  $ d'' \gamma^{n-1,0} + d' \gamma^{n-2,1} = 0 $  valables au voisinage de  $ V \cap S_1 $. Elles donnent, puisque l’on a $$\begin{aligned} & d' P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \gamma^{n-1,0} \wedge \Box ) = \\ & - P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge d' \gamma^{n-1,0} \wedge \Box ) + \\ & - P_3(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \frac{df}{f}\wedge \gamma^{n-1,0} \wedge \Box ) \end{aligned}$$ ainsi que $$d' P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} \frac{df}{f}\wedge \gamma^{n-2,1}\wedge \Box )\ = \ - P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} \frac{df}{f}\wedge d' \gamma^{n-2,1}\wedge \Box )$$ l’égalité $$\begin{aligned} d'' \mathcal{T}^{n,0}_j + j.d\bar{f}\wedge \mathcal{T}^{n,0}_{j+1} = & - d' P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \gamma^{n-1,0} \wedge \Box )\ + \\ & + d'P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j} \frac{df}{f}\wedge \gamma^{n-2,1}\wedge \Box ) \end{aligned}$$ ce qui prouve bien la  $d'-$fermeture sur  $V$  des courants  $\mathcal{T}^{n-1,1}_j $  pour  $j \in [1, n-1]$. Comme on a l’annulation des groupes  $ H^{p+1}(V, \Omega^q) $  pour tout  $ p \in [0,n-2]$  et tout  $q \in \mathbb{N}$, on peut trouver un courant  $ U^{n-2,1}_{n-1} $  sur  $V$  vérifiant  $ d' U^{n-2,1}_{n-1} = \mathcal{T}^{n-1,1}_{n-1} $. Posons  $ U^{n-2,1}_{j} : = \bar{f}^{n-j-1}.U^{n-2,1}_{n-1}$  pour  $j\in [1,n-1] $. Alors les hypothèses du Lemme suivant sont vérifiées pour  $p = 0 , a = 1 , b = n-1 $. Lemme. ------ On suppose donnés sur  $V$  des courants  $U_j^{n-p-1,p}$  pour  $j \in [a,b+1]$  et  $U_j^{n-p-2,p+1} $  pour  $j \in [a,b]$   où les  $U_j^{n-p-1,p}$  vérifient  $ \bar{f}.U_{j+1}^{n-p-1,p} = U_j^{n-p-1,p} $  pour  $j \in [a,b]$  et où les courants  $U_j^{n-p-2,p+1}$  vérifient  $ \bar{f}.U_{j+1}^{n-p-2,p+1} = U_j^{n-p-2,p+1} $  pour  $j \in [a,b-1]$. Supposons également donnée sur  $V$  une forme semi-méromorphe  $\gamma$  de degré  $n-1$  à pôles dans  $\lbrace f = 0 \rbrace$  telle que  $d\gamma$  soit de type  $(n,0)$. Supposons que l’on ait sur  $V$  les égalités de courants suivantes pour  $ j \in [a,b]$ $$\begin{aligned} \mathcal{T}^{n-p-1,p+1}_j : = &\ d''U_j^{n-p-1,p} + j.\bar{df}\wedge U_{j+1}^{n-p-1,p} + \\ \quad & - P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \gamma^{n-p-1,p}\wedge\Box ) + \\ \quad & \ P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge \gamma^{n-p-2,p+1}\wedge\Box ) \\ \quad = & \ d'U_j^{n-p-2,p+1} \end{aligned}$$ Alors les courants $$\begin{aligned} \mathcal{T}^{n-p+2,p+2}_j : = & \ d''U_j^{n-p-2,p+1} + j.\bar{df}\wedge U_{j+1}^{n-p-2,p+1} + \\ \quad & - P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \gamma^{n-p-2,p+1}\wedge\Box ) + \\ \quad & + P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge \gamma^{n-p-3,p+2}\wedge\Box ) \end{aligned}$$ sont  $d'-$fermés sur  $V$  pour  $ j \in [a,b-1] $. Preuve du lemme. ---------------- On a pour  $j \in [a,b]$ : $$\begin{aligned} d'\big(d'' U_j^{n-p-2,p+1} \big) & = - d'' \big(d' U_j^{n-p-2,p+1} \big) \\ \quad & = - d'' \big(j.\bar{df}\wedge U_{j+1}^{n-p-1,p}\big) \ + \\ \quad & + d''\big(P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \gamma^{n-p-1,p}\wedge\Box )\big) \\ \quad & - d''\big( P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge \gamma^{n-p-2,p+1}\wedge\Box ) \big) \end{aligned}$$ et donc $$\begin{aligned} d'\big(d'' U_j^{n-p-2,p+1} \big) & = j.\bar{df}\wedge d'' U_{j+1}^{n-p-1,p} \qquad \tag{1} \\ \quad & - P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge d''\gamma^{n-p-1,p}\wedge\Box ) \qquad \tag{2} \\ \quad & - P_3(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge\frac{df}{f}\wedge\gamma^{n-p-2,p+1}\wedge\Box ) \quad \tag{3} \\ \quad & j.P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge\frac{df}{f}\wedge \gamma^{n-p-2,p+1}\wedge\Box ) \quad \tag{4} \\ \quad & + P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge d''\gamma^{n-p-2,p+1}\wedge\Box ) \quad \tag{5} \end{aligned}$$ puisque l’on a $$\begin{aligned} & d'' \big(\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge\gamma^{n-p-2,p+1}\wedge\Box \big) = \\ &\qquad \qquad \qquad (\lambda - j).\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge\frac{df}{f}\wedge \gamma^{n-p-2,p+1}\wedge\Box \quad + \\ &\qquad \qquad \qquad - \int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge d''\gamma^{n-p-2,p+1}\wedge\Box \end{aligned}$$ Ensuite on a $$\begin{aligned} & d'\big( j.\bar{df}\wedge U_{j+1}^{n-p-2,p+1}\big) = -j.\bar{df}\wedge d''U_{j+1}^{n-p-1,p} \ + \tag{6} \\ & -j.\bar{df}\wedge P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j-1}\frac{df}{f}\wedge \gamma^{n-p-2,p+1}\wedge\Box ) \quad \tag{7} ; \end{aligned}$$ puis $$\begin{aligned} & - d' P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \gamma^{n-p-2,p+1}\wedge\Box ) = \\ & - P_3(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge\frac{\bar{df}}{\bar{f}}\wedge\gamma^{n-p-2,p+1}\wedge\Box ) \quad \tag{8} \\ & + P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge d'\gamma^{n-p-2,p+1}\wedge\Box ) \quad \tag{9} \end{aligned}$$ puisque $$\begin{aligned} & d' \int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \gamma^{n-p-2,p+1}\wedge\Box = \\ & \lambda.\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge\frac{\bar{df}}{\bar{f}}\wedge\gamma^{n-p-2,p+1}\wedge\Box \ + \\ & - \int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge d'\gamma^{n-p-2,p+1}\wedge\Box . \end{aligned}$$ Et enfin $$\begin{aligned} & d'\big(P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge \gamma^{n-p-3,p+2}\wedge\Box ) \big) = \\ & - P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge d' \gamma^{n-p-3,p+2}\wedge\Box ) \quad \tag{10}\end{aligned}$$ Maintenant les relations $$\begin{aligned} & d''\gamma^{n-p-1,p} + d'\gamma^{n-p-2,p+1} = 0 \\ & d''\gamma^{n-p-2,p+1} + d'\gamma^{n-p-3,p+2} = 0\end{aligned}$$ qui résultent de notre hypothèse sur  $\gamma$  ainsi que les égalités $$(1) + (6) = 0 , (2)+ (9) = 0 , (3) + (8) = 0 , (4) + (7) = 0 , (5) + (10) = 0$$ permettent de conclure.  $\hfill \blacksquare$ Remarque. --------- Pour peu que l’on ait  $ H^{p+2}(V, \Omega^{n-p+2}) = 0 $  on peut trouver, sous l’hypothèse du Lemme, un courant  $ U^{n-p+1,p+2}_{b-1}$  sur  $V$  vérifiant  $ d'U^{n-p+1,p+2}_{b-1} = \mathcal{T}^{n-p+2,p+2}_{b-1} $  et poser  $ U^{n-p+1,p+2}_{j} = \bar{f}^{b-j-1}. U^{n-p+1,p+2}_{b-1} $  pour  $j\in [a,b-a -1] $  et se retrouver à nouveau dans l’hypothèse du Lemme pour  $ p +1, a \ {\rm et} \ b-1 $. Fin de la démonstration du théorème (8.2). ------------------------------------------ Puisque l’on dispose de l’annulation de $ H^{p+1}(V, \Omega^q) $  pour tout  $ p \in [0,n-2]$  et tout  $q \in \mathbb{N}$, la remarque ci-dessus permet de continuer à appliquer le Lemme  $(n-1)-$fois (pour  $p\in [0,n-2]$) et d’obtenir un courant $d'-$fermé de type  $(0,n)$ $$\begin{aligned} \mathcal{T}^{0,n}_1 : = & d''U^{0,n-1}_1 + \bar{df}\wedge U^{0,n-1}_2 + \\ \quad & - P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \gamma^{0,n-1}\wedge\Box ) . \\ \end{aligned}$$ Donc  $ \mathcal{T}^{0,n}_1$  est une forme antiholomorphe de degré  $n$  sur  $V$. De plus on a $$d'' \mathcal{T}^{0,n}_1 + \frac{\bar{df}}{\bar{f}} \wedge \mathcal{T}^{0,n}_1 = 0$$ sur  $V\setminus Y$  car  $ d''\gamma^{0,n-1} = 0 $  montre que  $ P_2(\lambda = 0,\int_V \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{\bar{df}}{\bar{f}}\wedge \gamma^{0,n-1}\wedge\Box )$  est annulé par  $\wedge \bar{df} $  et  $d$. Par Hartogs, elle se prolonge en une forme antiholomorphe  $\bar{\Omega_1} $  sur  $X$  vérifiant  $ d(f.\Omega_1) = 0 $  sur  $X$. De plus on aura des courants  $\mathcal{U}$  et  $\mathcal{V}$  sur  $V$  de degré $n-1$  vérifiant l’égalité suivante  sur  $V\setminus S_1 = X \setminus S_1$[^42] : $$\mathcal{T}^{n,0}_1 - \bar{\Omega}_1 = d\mathcal{U} + \bar{df}\wedge \mathcal{V} .$$ Remarquons déja que la  $n-$forme holomorphe $d-$fermée  $\Omega_0 : = f.\Omega_1$  est $d-$exacte sur  $X$. Donc elle induit la classe nulle dans  $H^n(F,\mathbb{C})$. Il en sera donc de même pour  $ \Omega_1$ et pour une $n-$forme holomorphe  $A$  vérifiant  $ dA = df \wedge \Omega_1 $, d’après le théorème de positivité de Malgrange (voir \[M.74\] ou l’appendice de \[B.84 b\]). Posons $$T : = Pf(\lambda = 0 ,\int_X \vert f \vert^{2\lambda} w \wedge \Box ).$$ Montrons que l’on a  $ dT = \bar{df}\wedge \mathcal{T}^{n,0}_1 $  sur  $X$. En effet, on a $$d'T = Res(\lambda = 0 ,\int_X \vert f \vert^{2\lambda}\frac{df}{f}\wedge w \wedge \Box ) = 0$$ car l’absence de puissance de  $\bar{f}$  en dénominateur montre que la fonction méromorphe dont on prend le résidu à l’origine n’a pas de pôles en ce point[^43]. De plus on a $$d''T = Res(\lambda = 0 ,\int_X \vert f \vert^{2\lambda}\frac{\bar{df}}{\bar{f}}\wedge w \wedge \Box )$$ ce qui co[ï]{}ncide bien avec  $ \bar{df}\wedge \mathcal{T}^{n,0}_1 $  car on a $$\begin{aligned} & P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \frac{\bar{df}}{\bar{f}}\wedge \frac{df}{f}\wedge \gamma^{n-1,0}\wedge \Box ) = \\ & \qquad \qquad d''Res(\lambda = 0, \int_X \vert f \vert^{2\lambda} \frac{df}{f}\wedge \gamma^{n-1,0}\wedge \Box ) \ + \\ & \qquad \qquad + Res(\lambda = 0, \int_X \vert f \vert^{2\lambda} \frac{df}{f}\wedge d'' \gamma^{n-1,0}\wedge \Box ) \\ \quad & \qquad \qquad = \ 0 \end{aligned}$$ car à nouveau les residus à l’origine sont nuls pour la même raison que ci-dessus. On obtient donc l’égalité suivante de courants sur  $X \setminus S_1$ $$d\big(T - \bar{A} + \bar{df}\wedge \mathcal{U} \big) = 0$$ Comme on a, par hypothèse,  $H^n(X\setminus S_1, \mathbb{C}) = 0 $  on en conclut à l’existence d’un courant  $ \mathcal{W}$   de degré $n-1$  sur  $X \setminus S_1$  tel que $$T = \bar{A} + \bar{df}\wedge \mathcal{U} + d\mathcal{W} .$$ Alors le Lemme (D) de \[B.84 a\] donne que  $w$  induit la classe nulle dans  $H^n(F, \mathbb{C})$  contredisant notre hypothèse. Donc le courant  $\mathcal{T}^{n,0}_n$  n’est pas  $d'-$exact sur  $V$. La dualité entre\ $H^n(X\setminus K, \mathcal{O}) $  et  $\Gamma(K,\Omega^{n+1})$[^44] et la densité de l’image de la restriction  $ \Gamma(X,\Omega^{n+1}) \rightarrow \Gamma(K, \Omega^{n+1})$  permettent alors de trouver une forme holomorphe  $\omega\in \Gamma(X,\Omega^{n+1})$  telle que l’on ait  $< \mathcal{T}^{n,0}_n , d'\rho\wedge \bar{\omega} > \not= 0 $  ce qui donne $$\begin{aligned} & \ Res(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-n} w \wedge d'\rho\wedge \bar{\omega} ) \ +\\ \qquad & P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-n} \frac{df}{f}\wedge \gamma^{n-1,0}\wedge d'\rho\wedge \bar{\omega} ) \not= 0\end{aligned}$$ Mais on a $$\begin{aligned} d\big( \vert f \vert^{2\lambda} \bar{f}^{-n} .w \wedge \rho. \bar{\omega} ) = & \lambda \vert f \vert^{2\lambda} \bar{f}^{-n}.\frac{df}{f}\wedge w \wedge \rho. \bar{\omega} \ + \\ \qquad \qquad & (-1)^n \vert f \vert^{2\lambda} \bar{f}^{-n} . w \wedge d'\rho\wedge \bar{\omega}\end{aligned}$$ ce qui donne : $$\begin{aligned} & \ Res(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-n} w \wedge d'\rho\wedge \bar{\omega} ) = \\ & \qquad \qquad (-1)^{n+1} P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-n} \frac{df}{f}\wedge w \wedge\rho. \bar{\omega} )\end{aligned}$$ Comme on a $$\begin{aligned} & P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-n} \frac{df}{f}\wedge \gamma^{n-1,0}\wedge d'\rho\wedge \bar{\omega} ) = \\ &\qquad \qquad (-1)^{n} P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-n} \frac{df}{f}\wedge d' \gamma^{n-1,0}\wedge \rho\ . \bar{\omega} ) \end{aligned}$$ on obtient finalement $$P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-n} \frac{df}{f}\wedge(w - d' \gamma^{n-1,0})\wedge \rho\ . \bar{\omega} ) \not= 0$$ c’est à dire, vus les types $$P_2(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-n} \frac{df}{f}\wedge\tilde{w} \wedge \rho\ . \bar{\omega} ) \not= 0 .$$ Ceci achève la preuve du théorème (8.2).  $\hfill \blacksquare$ Pour conclure à l’injectivité de la variation dans le cas où l’image de la classe  $ e \in H^n_{c\,\cap\,S}(0) $  dans  $H^n(0)$  n’est pas nulle, il suffit de remarquer que si l’on écrit  $ \omega = \frac{df}{f}\wedge w' $  la décomposition dans le système de Gauss-Manin (localisé) en degré  $n$   de  $f$  de la forme méromorphe  $ \frac{w'}{f^n} $  va fournir une classe  $e' \in H^n(0)$  vérifiant  $\mathcal{H}(e,e') \not= 0 $. En effet, les formes méromorphes induisant des classes dans les sous espaces spectraux de la monodromie pour des valeurs propres différentes de  $1$  ne donneront pas de pôles doubles aux entiers négatifs, pas plus que les formes du type  $ \frac{df}{f}\wedge du $  où  $u$  est méromorphes de degré  $n-1$  dans le prolongement méromorphe de $$\int_X \vert f \vert^{2\lambda} \rho \frac{df}{f}\wedge \tilde{w}\wedge \bar{\Box} .$$ Le cas plus délicat est celui où la classe  $e$  considérée est dans le noyau de l’application  $ can_{c\,\cap\,S} : H^n_{c\,\cap\,S}(0) \rightarrow H^n(0) .$ D’après le théorème 1 du (4.6) il existe  $[\alpha] \in H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) $  vérifiant  $ i([\alpha]) = e $. De plus, comme l’application  $i$  est injective, on peut supposer que l’on a  $ T([\alpha]) = [\alpha] $  dans  $H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) .$ Les lemmes suivants vont nous fournir un représentant de la classe  $[\alpha]$  considérée qui met en évidence la propriété d’invariance par la monodromie. Lemme. ------- Supposons  $n \geq 3$. Soit  $ [\alpha] \in H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0))$  vérifiant  $T(i([\alpha])) = i([\alpha])$  dans  $ H^n_{c\,\cap\,S}(0) $. Alors, pour  $k$  assez grand, il existe un représentant\  $ \hat{\alpha}\in \Gamma(S^*, Ker\,\delta^{n-1}(k +1)) $   tel que l’on ait $$\mathcal{N}(\hat{\alpha}) \in \Gamma(Y, Ker\,\delta^{n-1}(k +1))$$ et dont l’image dans  $H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) $  est  $[\alpha]$. *Preuve.* Pour  $k \gg 1 $  on a un isomorphisme  $ r^{n-1}(k) : h^{n-1}(k) \rightarrow H^{n-1}(0) .$ Puisque l’on a  $H^1(S,H^{n-1}(0)) = 0 $   et l’annulation de  $ H^1(S^*, \delta\mathcal{E}^{n-2}) $  prouvée au lemme (4.7), on peut relever  $[\alpha]$  en une section  $\alpha' \in \Gamma(S^*, Ker\,\delta^{n-1}(k)) .$ Grace à l’annulation de  $ H^1(Y, \delta\mathcal{E}^{n-2}) $  prouvée au lemme (4.7), l’hypothèse  $ T[\alpha] = [\alpha]$  dans  $ H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0))$  permet de trouver  $ \beta \in \Gamma(Y, Ker\,\delta^{n-1}(k)) $  et  $ \Gamma \in \Gamma(S^*, \delta\mathcal{E}^{n-2}(k))$  vérifiant $$\mathcal{N} \alpha' = \beta + \Gamma \quad {\rm au \ voisinage \ de} \ S^* .$$ Ceci se déduit d’une “chasse” facile sur le diagramme commutatif exact d’espaces vectoriels monodromiques : $$\xymatrix{\quad & \quad & 0 & \quad \\ \quad & \quad &\quad H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0))\ar[u] & \quad \\ H^0(S^*, \delta\mathcal{E}^{n-2}(k)) \ar[r] & H^0(S^*, Ker\,\delta^{n-1}(k)) \ar[r] & H^0(S^*, H^{n-1}(0)) \ar[r] \ar[u]& 0 \\ H^0(Y, \delta\mathcal{E}^{n-2}(k)) \ar[r] \ar[u]& H^0(Y, Ker\,\delta^{n-1}(k)) \ar[r] \ar[u]& H^0(Y, H^{n-1}(0)) \ar[r] \ar[u] & 0 }$$ La suite exacte $$0 \rightarrow Ker\,\delta^{n-2}(k) \rightarrow \mathcal{E}^{n-2}(k) \rightarrow \delta\mathcal{E}^{n-2}(k) \rightarrow 0$$ donne, pour  $ n \geq 4$  la surjectivité de la flèche $$\delta : H^0(S^*, \mathcal{E}^{n-2}(k)) \rightarrow H^0(S^*, \delta\mathcal{E}^{n-2}(k))$$ puisque, pour  $ n \geq 4$, on a  $ Ker\,\delta^{n-2}(k) \simeq \delta\mathcal{E}^{n-3}(k)$  et que l’annulation du  $H^1(S^*, \delta\mathcal{E}^{n-3}(k))$  est donnée par (4.7). On peut donc trouver\ $ \gamma \in \Gamma(S^*, \mathcal{E}^{n-2}(k)) $  vérifiant  $ \Gamma = \delta\gamma $. Pour  $n = 3 $  on a  $ H^1(S^*, Ker\,\delta^1(k)) \simeq H^1(S^*, h^1(k)) \not= 0 $. Mais l’application $$j_{k,k+1} : H^1(S^*, h^1(k)) \rightarrow H^1(S^*, h^1(k+1))$$ est nulle (voir (1.6)) et quitte à changer  $k$  en  $k+1$  on peut lever l’obstruction et trouver également  $ \gamma \in \Gamma(S^*, \mathcal{E}^{n-2}(k+1)) $  vérifiant  $ \Gamma = \delta\gamma $. Définissons alors  $ \hat{\alpha}\in \Gamma(S^*, Ker\,\delta^{n-1}(k +1)) $   en posant $$\hat{\alpha}_{k+1} = \alpha'_k + \frac{df}{f}\wedge \gamma_k \quad {\rm et} \quad \hat{\alpha}_j = \beta_j \quad \forall j\in [1,k] .$$ L’égalité  $\delta\hat{\alpha} = 0 $  est immédiate et on a  $ \mathcal{N}(\hat{\alpha}) = j_{k,k+1}(\beta)$  ce qui achève la preuve. $\hfill \blacksquare $ Variante pour  $n = 2 $. ------------------------ Si  $\alpha \in \mathcal{K} $  vérifie  $ \mathcal{N}([\alpha]) = 0 $  dans  $ H^1_{\lbrace 0 \rbrace}(S, H^1(0))$, il existe  $ k \gg 1 $  et  $ \hat{\alpha} \in \Gamma(S^*, Ker\,\delta^1(k)) $  tel que  $\mathcal{N}(\hat{\alpha}) $  soit la restriction à  $ S^*$  d’un élément de  $ \Gamma(Y,, Ker\,\delta^1(k)) $, et induisant la classe de  $\alpha$  dans $$\frac{\mathcal{K}}{H^1(S^*, \mathbb{C})} \simeq H^1_{\lbrace 0 \rbrace}(S, H^1(0)) .$$ *Preuve.* Pour  $ k \gg 1 $  on peut représenter  $\alpha$  par  $ \tilde{\alpha} \in \Gamma(S^*, Ker\,\delta^1(k)$ ; de plus on peut également trouver  $ \beta \in \Gamma(S, Ker\,\delta^1(k)) $  et  $\Gamma \in \Gamma(S^*, \delta\mathcal{E}^0(k) $  vérifiant $$\mathcal{N}(\tilde{\alpha}) = \beta + \Gamma .$$ Comme  $ \underset{k \rightarrow \infty}{\lim} h^1(k) \simeq H^1(0) $  est à support dans  $S$, on peut supposer également que  $\beta$  est la restriction à  $S$  de  $\tilde{\beta} \in \Gamma(Y, Ker\,\delta^1(k)$. Comme la monodromie est l’identité sur  $ H^1(S^*, \mathbb{C}) $, on peut trouver  $ \gamma \in \Gamma(S^*, \mathcal{E}^0(k)) $  vérifiant  $ \mathcal{N}(\Gamma) = \delta \gamma .$ On aura alors $$\tilde{\Gamma} : = \begin{pmatrix} {\Gamma_k + \frac{df}{f} \wedge \gamma_k } \\ {0} \\ {\vdots} \\ {0 } \end{pmatrix} = \begin{pmatrix} {\Gamma_k} \\ {\Gamma_{k-1} } \\ {\vdots} \\ {0} \end{pmatrix} - \delta \begin{pmatrix} {0} \\ {\gamma_k} \\ {\vdots} \\ {\gamma_1} \end{pmatrix} .$$ Donc, quitte à ajouter un cobord (et à changer  $k$  en  $k+1$) on peut supposer que $$\tilde{\alpha} = \tilde{\beta} + \tilde{\Gamma}$$ avec  $ \Gamma_j = 0 $  pour  $ j \not= k $. Posons alors  $ \hat{\alpha}_{k+1} = \alpha_k - \tilde{\Gamma_k} $  et  $ \hat{\alpha}_j = \beta_j \quad \forall j \in [1,k] $. Alors  $\hat{\alpha}$  vérifie bien les propriétés requises. $\hfill \blacksquare$ Lemme. ------- Dans la situation du lemme (8.11) ci-dessus, il existe (quitte à restreindre  $X$ ) une $n-$forme semi-méromorphe  $\hat{w}$  qui est définie et  $d-$fermée dans  $X^* = X \setminus\lbrace 0 \rbrace$  et à support dans  $mod\, S$, qui, restreinte à  $X\setminus \bar{X}'\setminus Y $, induit l’image de la restriction de  $i[\alpha]$  dans  $H^n_{mod\, S}(X \setminus \bar{X}' \setminus Y, \mathbb{C}) $. *Preuve.* Fixons un voisinage ouvert  $ \mathcal{W}$  de  $S^*$  dans  $X^*$ et supposons que, quitte à restreindre le disque de centre  $0$  dans  $\mathbb{C}$  qui définit  $X$, la section  $\beta$  construite au lemme (8.11) est définie et  $\delta-$fermée sur  $X$. Soit  $\chi \in \mathcal{C}^{\infty}(\mathcal{W})$  valant identiquement 1 près de  $S^*$  et identiquement nulle près de  $\partial\mathcal{W}$ . Nous la prolongerons par  $0$  sur  $X^*$. Posons alors $$\begin{aligned} \hat{w} = d\chi\wedge \hat{\alpha}_{k+1} - (1 - \chi)\frac{df}{f}\wedge\beta_k \\ \beta'_j = \beta_j \quad \forall j\in[1,k] \quad {\rm et} \quad \beta'_{k+1} = 0 .\end{aligned}$$ On a alors $$d\chi\wedge \hat{\alpha} = j_{1,k+1}(\tilde{w}) + \delta ( (1 - \chi).\beta' ) .$$ Comme  $(1 - \chi).\beta' $  est à support dans  $mod\, S$ , on en déduit que la $n-$forme semi-méromorphe  $\hat{w}$  vérifie les propriétés demandées. $\hfill \blacksquare $ [**Remarque importante.**]{} Si la classe définie par  $\hat{w}$  dans  $H^n_{mod\,S}(X^*\setminus Y^*, \mathbb{C})$  est nulle, on peut trouver une $(n-1)-$forme semi-méromorphe  $\varepsilon$  sur  $X^*$   à pôles dans  $Y^*$  et à support dans  $mod\,S$[^45] et vérifiant $$\hat{w} = d\varepsilon \quad {\rm sur} \ X^* .$$ Posons alors $$\gamma : = \chi.\hat{\alpha} - (1 - \chi).\beta' - j_{1,k+1}(\varepsilon) .$$ Comme  $\delta \gamma = 0 $  cela définit une classe dans $$\mathbb{H}^{n-1}(X^*, (\mathcal{E}^{\bullet}(k+1), \delta^{\bullet})$$ qui s’identifie au $(n-1)-$ième groupe de cohomologie de la fibre de Milnor de  $f$  à l’origine. On a donc un élément de  $ H^0(S, H^{n-1}(0))$. Mais ceci montre, puisque  $\varepsilon \ {\rm et} \ (1-\chi)$  sont nulles au voisinage de  $S^*$  que la section correspondante sur  $S$  du faisceau  $H^{n-1}(0)$  prolonge la section “initiale”  $\alpha \in H^0(S^*, H^{n-1}(0))$. Ceci montre que la nullité de  $\hat{w}$  dans l’espace  $H^n_{mod\,S}(X^*\setminus Y^*, \mathbb{C})$  implique celle de la classe de  $\alpha$  dans  $H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0))$. Variante pour  $n =2$. ---------------------- L’énoncé est ici le même qu’en (8.13) sauf que l’on obtient dans l’espace  $H^2_{mod\, S}((X \setminus \bar{X}' )\setminus Y, \mathbb{C}) $  qu’une égalité [**modulo**]{} l’image dans cet espace du sous-espace  $i\circ j(H^1(S^*, \mathbb{C}))$. La preuve est analogue. La remarque importante qui suit le (8.13) est valable pour  $n = 2$. Dans ce cas, elle donne que la classe initiale de  $\mathcal{K}$  dont on part, est dans  $ j(H^1(S^*, \mathbb{C}))$  dès que la classe de  $\hat{w}$  dans  $H^n_{mod\,S}(X^*\setminus Y^*, \mathbb{C})$  est nulle. On remarquera que l’image dans  $H^2_{c\,\cap\,S}(0)$  de  $ j(H^1(S^*, \mathbb{C}))$   (modulo laquelle on travaille) est l’image par la restriction naturelle de  $H^2_{mod\, S}(X^*, \mathbb{C})$. Ceci résulte de notre construction et de l’isomorphisme $$H^2_{mod\, S}(X^*, \mathbb{C}) \simeq H^1(S^*, \mathbb{C})$$ qui découle facilement des annulations des  $ H^i(X^*, \mathbb{C}) $  pour  $i = 1,2 $. Notre stratégie, pour terminer la preuve de l’injectivité de la variation pour  $n \geq 3$, consiste à considérer le diagramme commutatif suivant dont les lignes sont exactes d’après le théorème (5.1)[^46] : $$\xymatrix{H^n_{mod\,S}(X^*) \ar[r] & H^n_{mod\,S}(X^*\setminus Y^*) \ar[d]^{r} \ar[r]^{Res} & H^{n-1}_{mod\,S}(Y^*) \ar[d]^{r'} \ar[r] & H^{n+1}_{mod\,S}(X^*) \\ \cdots \ar[r] & H^n_{mod\,S}(X\setminus \bar{X}'\setminus Y) \ar[r]^{Res'} & H^{n-1}_{mod\,S}(Y\setminus \bar{Y}') \ar[d]^{\partial} \ar[r] & \cdots \\ \quad & \quad & H^n_c(0) & \quad}$$ et à montrer les propriété suivantes qui permettent facilement de conclure que si  $ \partial\circ Res' $  annule l’image de  $\hat{w}$   dans  $H^n_{mod\,S}(X\setminus \bar{X}'\setminus Y)$   alors c’est que cette image est déja nulle ; mais alors l’image de  $ d\chi\wedge \tilde{\alpha}$  est également nulle dans ce même espace. Le théorème (4.6) donne alors que la classe de  $ \alpha \in H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0)) $  est nulle. Ceci achève donc la preuve de l’injectivité de la variation pour  $n \geq 3$, modulo la démonstration des assertions suivantes: - [1)]{} La restriction de  $\hat{w}$  est dans l’image de l’application de restriction notée  $r$ . - [2)]{} L’application de résidu  $Res$  est un isomorphisme, grace à l’annulation des groupes  $H^i_{mod\,S}(X^*) \quad {\rm pour} \quad i = n, n+1$  pour  $n \geq 3 $. - [3)]{} La composée  $\partial\circ r' $  est injective. Les lemmes (8.11) et le (8.13) donnent déja la propriété 1). Montrons donc les annulations qui donnent la seconde propriété pour  $n \geq 3$. Comme on a  $ H^i(X^*) = 0 \quad {\rm pour} \quad i = n, n+1$  il nous suffit de montrer que si $ \psi \in \mathcal{C}^{\infty}(X^*) $  est de degré  $n-1$  ou  $n$  et a une différentielle identiquement nulle au voisinage de  $S^*$  il existe  $ \chi \in \mathcal{C}_{mod\,S}^{\infty}(X^*)$  telle que  $d\chi = d\psi$  sur  $X^*$. Comme on a  $n \geq 3$  on a une base de voisinages ouverts  $\mathcal{W}$   de  $S^*$  dans  $X^*$  qui vérifient  $ H^i(\mathcal{W}) = 0 $   pour  $ i \geq 2 $. Ceci permet aisément de conclure. Il nous reste à montrer l’injectivité de la flèche $$\partial\circ r' : H^{n-1}_{mod\,S}(Y^*) \rightarrow H^n_c(0)$$ qui est définie comme suit : à  $\psi \in \mathcal{C}^{\infty}_{mod\,S}(\mathcal{V}(Y^*))$  vérifiant  $d\psi = 0 $  on associe la restriction à  $\lbrace f=s \rbrace$  pour  $s$  assez voisin de  $0$  de la forme  $ d\rho \wedge \psi $  où  $ \rho \in \mathcal{C}^{\infty}_c(X)$  vaut identiquement au voisinage de l’origine. Cette injectivité va résulter d’un théorème général. Théorème. --------- *Notons respectivement par  $ \mathbb{H}^p_c(k) $  et  $ \mathbb{H}^p_{c\,\cap\,S}(k)$  les groupes d’hypercohomologie à supports  $ \mathbb{H}^p_c(Y, \mathcal{E}(k)^{\bullet}, \delta^{\bullet})$  et  $ \mathbb{H}^p_{c\,\cap\,S}(Y, \mathcal{E}(k)^{\bullet}, \delta^{\bullet})$. Alors on a, pour chaque  $ k \geq 1 $  la suite exacte longue d’espaces vectoriels monodromiques :* $$\rightarrow \mathbb{H}^{p-1}_{mod\,S}(\partial Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet}) \overset{j}{\rightarrow} \mathbb{H}^p_c(k) \overset{can_{c,S}}{\longrightarrow} \mathbb{H}^p_{c\,\cap\,S}(k) \overset{\eta}{\rightarrow} \mathbb{H}^{p}_{mod\,S}(\partial Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet})\rightarrow$$ *où l’on a noté par  $ \mathbb{H}^{p}_{mod\,S}(\partial Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet})$  la limite inductive quand  $ \varepsilon' \rightarrow \varepsilon$  de  $ \mathbb{H}^{p}_{mod\,S}( Y\setminus\bar{Y}', \mathcal{E}^{\bullet}(k), \delta^{\bullet})$  avec  $ Y' = Y \,\cap\,X' $  où  $X'$  est la trace sur  $X$  de la boule de centre  $0$  et de rayon*  $\varepsilon'$[^47]. Démonstration. -------------- Commen[ç]{}ons par définir l’application  $j$. Fixons  $ X': = B(0, \varepsilon') \cap f^{-1}(D) \subset X = B(0, \varepsilon) \cap f^{-1}(D), \varepsilon - \varepsilon' \ll\varepsilon \ll 1 $. Soit  $\varphi$  une  section de  $\mathcal{E}^{p-1}(k)$  sur  $ X\setminus \bar{X'} $  identiquement nulle au voisinage de  $S$  et  $\delta-$fermée. Soit, de plus, $\rho \in \mathcal{C}^{\infty}(X) $, une fonction identiquement égale à 1 au voisinage de  $\bar{X'}$ et à support  $f-$propre. Alors  $ j(\varphi) = d\rho \wedge \varphi = \delta((\rho - 1).\varphi)$  est dans  $ \Gamma_c(Y, Ker\,\delta \cap \mathcal{E}^p(k))$; elle définit donc un élément de  $ \mathbb{H}^p_c(k)$. Montrons que ceci définit bien une application linéaire $$j : \mathbb{H}^{p-1}_{mod\,S}(\partial Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet}) \rightarrow \mathbb{H}^p_c(k) .$$ Si  $ \varphi$  est l’image par  $\delta$ d’une section  $ \psi \in \Gamma_{mod\, S}(\partial Y, \mathcal{E}^{p-2}(k)) $  alors on aura $$d\rho \wedge\delta \psi = \delta(-d\rho\wedge\psi) .$$ où l’on a  $ d\rho\wedge \psi \in \Gamma_c(Y, \mathcal{E}^{p-1}(k)) $, en prolongeant par  $0$. On a clairement  $j\circ _k\mathcal{N} = _k\mathcal{N}\circ j $  ce qui montre l’aspect monodromique de l’application linéaire  $j$. On remarquera que pour  $k = 1$  cette application est bien compatible avec l’application  $\partial$  considérée plus haut. Identifions le noyau de  $j$ ; si on a  $ d\rho \wedge \varphi = \delta\gamma$  où  $\gamma \in \Gamma_c(Y, \mathcal{E}^{p-1}(k)) $, on aura, quitte à augmenter la taille de  $ X' $, que la restriction  $\tilde{\gamma}$  de  $\gamma$  à  $X\setminus \bar{X'}$  sera dans  $ \Gamma_{mod\,S}(\partial Y, \mathcal{E}^{p-1}(k)) $  et vérifiera  $ d\rho \wedge \varphi = \delta\tilde{\gamma} .$ Alors  $ \gamma + (1-\rho).\varphi $  définira un élément de  $ \mathbb{H}^{p-1}_{c\,\cap\,S}(Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet}) .$ La construction de l’analogue $\eta$   en degré  $(p-1)$  de l’application  $\eta$  que nous allons définir maintenant montrera que le noyau de  $j$  est bien l’image de  $ \mathbb{H}^{p-1}_{c\, \cap \,S}(k) $  par  $\eta$. La construction de l’application  $\eta$  est très simple. Si  $ w \in \Gamma_{c\,\cap\,S}(Y, Ker\,\delta \cap \mathcal{E}^{p}(k))$  pour un choix de  $X'$  assez gros, la restriction de  $w$  à  $Y \cap (X\setminus\bar{X'})$  est à support dans  $mod\,S$  et donne une classe dans  $ \mathbb{H}^{p}_{mod\,S}(\partial Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet}) $. On vérifie immédiatement que ceci passe au quotient. L’application est donc bien définie et elle est clairement linéaire et monodromique. De plus, il est clair que si  $w$  est à support compact, son image par  $\eta$  est nulle. Réciproquement, si  $\eta[w] = 0 $, alors on peut écrire sur  $\partial Y$ : $$w = \delta v$$ avec  $v \in \Gamma_{mod\,S}(\partial Y, \mathcal{E}^{p-1}(k)) $. Alors pour  $\rho \in \mathcal{C}^{\infty}_{c/f}(X) $  valant identiquement sur  $\bar{X'}$  avec  $X'$  assez gros, on aura  $ w' : = w - \delta((1-\rho).v) $  qui représentera la classe  $[w] \in H^p_{c\,\cap\,S}(0)$  et sera à support compact sur  $Y$. Donc  $[w]$  est bien dans l’image de  $can_{c,S} .$ Pour achever la démonstration il nous reste à montrer que le noyau de l’application $$can_{c,S} : \mathbb{H}^p_c(k) \rightarrow \mathbb{H}^p_{c\,\cap\,S}(k)$$ est bien l’image de $j$. Soit donc  $v \in \Gamma_{c\,\cap\,S}(Y, Ker\,\delta \cap \mathcal{E}^{p-1}(k)) $  tel que  $w = \delta v $  soit à support compact dans  $Y$. Alors  $v$  définit par restriction à  $\partial Y$  un élément  $[v]$ de  $ \mathbb{H}^{p-1}_{mod\,S}(\partial Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet})$. Pour calculer  $j([v])$  choisissons  $\rho$  comme ci-dessus[^48]. Alors on aura  $ j([v]) = [d\rho\wedge v] = \delta((\rho - 1).v) $. Mais on a $$\delta v - \delta(\rho .v) = \delta((1-\rho).v)$$ avec  $(1-\rho).v \in \Gamma_c(Y, \mathcal{E}^{p-1}(k)) $  et donc la classe de  $j([v])$  dans  $\mathbb{H}^p_c(k)$  coïncide bien avec celle de  $ w = \delta v . \hfill \blacksquare$ Pour terminer la preuve de l’injectivité de la variation (c’est à dire pour déduire l’assertion 3) du (8.15) du théorème (8.17)), il nous suffit de montrer que pour  $n \geq 3$  la flèche  $\eta$  en degré  $n-1$  de la suite exacte longue de (8.17) est nulle, ce qui résulte de la proposition suivante. Proposition. ------------ Sous les hypothèses standards pour  $n \geq 3$  l’application canonique $$can_{c,S} : \underset{k \rightarrow \infty}{lim} \mathbb{H}^{n-1}_c(k) \rightarrow \underset{k \rightarrow \infty}{lim} \mathbb{H}^{n-1}_{c\,\cap\,S}(k)$$ est un isomorphisme. De plus, pour  $n \geq 4 $  ces deux groupes sont nuls. *Preuve.* Commen[ç]{}ons par montrer le lemme suivant Lemme. ------ Pour  $ n \geq 4 $  on a  $ H^{n-1}_{c\,\cap\,S}(Y, \mathbb{C}) = 0 $. Pour  $ n = 2 , 3 $  on a l’isomorphisme  $ H^{n-1}_{c\,\cap\,S}(Y, \mathbb{C}) \simeq H^{n-1}_{\lbrace 0 \rbrace}(S, \mathbb{C} )$. *Preuve.* Comme on peut remplacer  $Y$  par la limite inductive de ses voisinages ouverts, on peut calculer  $ H^{n-1}_{c\,\cap\,S}(Y, \mathbb{C})$  à l’aide du complexe de de Rham des germes de formes différentielles à support dans  $ c\,\cap\,S $. Soit donc  $\varphi $  un tel germe d-fermé de degré  $n-1$. Comme   $H^{n-1}(Y, \mathbb{C}) = 0 $  puisque  $Y$  est contractible, on peut trouver un germe  $\psi$  de degré  $n-2$  vérifiant  $ d\psi = \varphi $. Si  $K$  est un compact de  $S$  assez gros, on aura  $ d\psi = 0 $  au voisinage de  $S\setminus K$. On constate alors que si la classe de cohomologie définie par  $\psi$  dans  $H^{n-2}(S\setminus K, \mathbb{C})$   est nulle pour  $n \geq 3$  ou prolongeable à  $S$  pour  $n = 2$  alors la classe définie par  $\varphi$  est nulle dans  $ H^{n-1}_{c\,\cap\,S}(Y, \mathbb{C})$. Ceci permet alors de construire un isomorphisme  $ H^{n-1}_{c\,\cap\,S}(Y, \mathbb{C}) \simeq H^{n-1}_{ \lbrace 0 \rbrace}(S, \mathbb{C} )$  ce qui achève la preuve puisque pour  $ n \geq 4 $  le groupe  $ H^{n-1}_{\lbrace 0 \rbrace}(S, \mathbb{C} )$  est nul. $\hfill \blacksquare$ Preuve de la proposition (8.19). -------------------------------- Supposons maintenant  $ n \geq 3 $  et notons par  $\Phi $  une famille paracompactifiante de fermés de  $Y$  égale soit à  $c$  ou bien à  $c\,\cap\,S $. Comme les faisceaux  $\mathcal{E}^{\bullet}(k) $  sont fins, on aura $$\mathbb{H}^{n-1}_{\Phi}(Y, \mathcal{E}^{\bullet}(k), \delta^{\bullet}) \simeq \Gamma_{\Phi}(Y, Ker\,\delta^{n-1}(k))\big/ \delta \Gamma_{\Phi}(Y, \mathcal{E}^{n-2}(k)) .$$ Soit maintenant  $w \in \Gamma_{\Phi}(Y, Ker\,\delta^{n-1}(k)) $. Alors si  $K$  est un compact de  $Y$  assez gros, $w$  est identiquement nulle au voisinage de  $S\setminus K$. La suite exacte $$0 \rightarrow \delta \mathcal{E}^{n-2}(k) \rightarrow Ker\,\delta^{n-1} \rightarrow H^{n-1}(0) \rightarrow 0 .$$ donne alors $$0 \rightarrow \Gamma_{\Phi}(Y, \delta \mathcal{E}^{n-2}(k)) \rightarrow \Gamma_{\Phi}(Y, Ker\,\delta^{n-1} ) \overset{r}{\rightarrow} \Gamma_{\Phi}(Y, h^{n-1}(k)) \rightarrow H^1_{\Phi}(Y, \delta \mathcal{E}^{n-2}(k)) \cdots$$ Et comme, pour  $n \geq 3$  le faisceau  $ h^{n-1}(k)$  est à support dans  $S$, on aura $$r(w)\vert_{S\setminus K} \equiv 0 \quad {\rm dans} \quad \Gamma_{\Phi}(S\setminus K, h^{n-1}(k)) \simeq \Gamma_{\Phi}(Y\setminus K, h^{n-1}(k)).$$ Comme, de plus,  $ h^{n-1}(k)$  est un système local sur  $S^*$  et n’admet pas de section non nulle à support l’origine, on aura  $r(w) = 0 $  et on en conclut que  $ w \in \Gamma_{\Phi}(Y, \delta \mathcal{E}^{n-2}(k)).$ Les suites exactes $$0 \rightarrow \delta\mathcal{E}^{q-1}(k) \rightarrow \mathcal{E}^{q}(k)\rightarrow \delta\mathcal{E}^{q}(k) \rightarrow 0$$ pour  $ 2 \leq q \leq n-2 $  conduisent aux isomorphismes, puisque les faisceaux  $ \mathcal{E}^q(k)$  sont fins et la famille de supports  $ \Phi $  est paracompactifiante $$\Gamma_{\Phi}(Y, \delta \mathcal{E}^{n-2}(k))\big/ \delta \Gamma_{\Phi}(Y, \mathcal{E}^{n-2}(k)) \simeq H^1_{\Phi}(Y, \delta \mathcal{E}^{n-3}(k)) \simeq \cdots \simeq H^{n-3}_{\Phi}(Y, \delta \mathcal{E}^{1}(k)).$$ Bien sur pour  $n=3 $  on doit remplacer  $ H^{n-3}_{\Phi}(Y, \delta \mathcal{E}^{1}(k))$  par son quotient par  $ \delta H^0_{\Phi}(Y, \mathcal{E}^1(k)) .$ La suite exacte $$0 \rightarrow Ker\,\delta^1(k) \rightarrow \mathcal{E}^1(k) \rightarrow \delta \mathcal{E}^{1}(k)) \rightarrow 0$$ donne alors l’isomorphisme  $ H^{n-3}_{\Phi}(Y, \delta \mathcal{E}^{1}(k)) \simeq H^{n-2}_{\Phi}(Y, Ker\,\delta^1(k)) .$ Les suites exactes $$\begin{aligned} 0 \rightarrow \delta \mathcal{E}^0(k) \rightarrow Ker\,\delta^1(k) \rightarrow h^1(k) \rightarrow 0 \\ 0 \rightarrow h^0(k) \rightarrow \mathcal{E}^0(k) \rightarrow \delta \mathcal{E}^0(k) \rightarrow 0 \end{aligned}$$ combinées avec le fait que  $ j_{k,k+1} : h^1(k) \rightarrow h^1(k+1) $  est nulle, donnent un isomorphisme $$\begin{aligned} \underset{k \rightarrow \infty}{lim} \mathbb{H}^{n-1}_{\Phi}(k) \simeq \underset{k \rightarrow \infty}{lim} H^{n-2}_{\Phi}(Y, Ker\,\delta^1(k)) \simeq \underset{k \rightarrow \infty}{lim} H^{n-2}_{\Phi}(Y, \delta\mathcal{E}^0(k)) \\ \simeq \underset{k \rightarrow \infty}{lim} H^{n-1}_{\Phi}(Y, h^0(k)) \simeq H^{n-1}_{\Phi}(Y, \mathbb{C}). \end{aligned}$$ Pour terminer la preuve de la proposition (8.17), il suffit donc de voir que l’application canonique $$H^{n-1}_{c\,\cap\,S}(Y, \mathbb{C}) \rightarrow H^{n-1}_c(Y, \mathbb{C})$$ est un isomorphisme pour  $ n \geq 3 $, et entre groupes nuls pour  $n \geq 4 $. Ceci résulte du lemme (8.20).  $ \hfill \blacksquare $ Ceci achève donc la preuve de l’injectivité de la variation pour  $n \geq 3$. On a donc également établi la non dégénérescence de la forme hermitienne canonique (voir (7.1)) et l’égalité de dimension $$\dim_{\mathbb{C}}\, \big( Im\, \theta\big) = \dim_{\mathbb{C}}\,\big( H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0))\big).$$ Cette égalité implique que  $ Im\, \theta$  co[ï]{}ncide avec l’orthogonal dans  $ H^1(S^*, H^{n-1}(0))$  du sous espace  $ H^0(S, H^{n-1}(0))$  de  $H^0(S^*, H^{n-1}(0))$  pour l’accouplement non dégénéré déduit de la forme hermitienne canonique sur le système local  $ H^{n-1}(0))$ (voir (3.1)). Le cas  $n = 2$. ---------------- Nous allons montrer que les résultats précédents s’étendent “mutadis mutandis” au cas  $n =2 $  quitte à remplacer l’espace vectoriel monodromique  $H^2_{c\,\cap\,S}(0) $  par le quotient $$\frac{H^2_{c\,\cap\,S}(0)}{H^1(S^*, \mathbb{C})}$$ où l’inclusion  $j$  de  $H^1(S^*, \mathbb{C})$[^49]a été définie au Lemme (4.8) (voir aussi (4.12.1)). Nous allons seulement préciser les points de la démonstration du cas  $n \geq 3$  qui sont à modifier de fa[ç]{}on significative dans ce cas. ### Commen[ç]{}ons par remarquer que toute la première partie de la preuve de l’injectivité de la variation, à savoir les sections (8.1) à (8.10), reste valable telle quelle[^50] Il n’en n’est pas de même pour la seconde partie de cette preuve. - [1)]{} Les lemmes (8.11) et (8.13) doivent être remplacés par les variantes (8.12) et (8.14) qui les suivent. - [2)]{} L’assertion 2) de (8.15) doit être remplacée par l’assertion suivante: L’application  $Res$  est surjective et son noyau a pour image par  $r$  l’image injective par la restriction à  $ (X \setminus \bar{X}' )\setminus Y $  du sous espace  $j(H^1(S^*, \mathbb{C}))$  de  $H^2_{c\,\cap\,S}(0)$. Ceci est une conséquence facile de la remarque qui conclut le (8.14) - [3)]{} On a pour  $n = 2$,  $H^1_c(0) \simeq 0 $  et  $ H^1_{c\,\cap\,S}(0) \simeq H^1_{\lbrace 0 \rbrace} (S,\mathbb{C}) $. De plus, un élément de  $\mathcal{K}\big/ j(H^1(S^*, \mathbb{C})$  invariant par la monodromie donne une classe nulle par l’application  $Res$  du diagramme de (8.17) si et seulement si cet élément est nul. ### Preuve de l’assertion 3). L’annulation de  $H^1_c(0)$  résulte du fait que la fibre de Milnor  $F$  de  $f$  à l’origine est une variété de Stein de dimension 2 ; donc  $H^3(F, \mathbb{C}) \simeq 0$   et par dualité de Poincaré on a  $H^1_c(F, \mathbb{C}) \simeq 0 .$ L’isomorphisme $$\frac{ H^0(S^*, \mathbb{C}) }{ H^0(S, \mathbb{C})}\simeq H^1_{\lbrace 0 \rbrace} (S, \mathbb{C}) \rightarrow H^1_{c\,\cap\,S}(Y, \mathbb{C})$$ est donné par la flèche $$\varepsilon : H^0(S^*, \mathbb{C}) \rightarrow H^1_{c\,\cap\,S}(Y, \mathbb{C})$$ définie de la fa[ç]{}on suivante : à la fonction localement constante  $\varphi$  au voisinage de  $S^*$  on associe la classe  $[\varphi.d\chi ]$  où  $\chi \in \mathcal{C}^{\infty}(Y) $  vaut identiquement au voisinage de  $\partial S $  et s’annule identiquement dès que l’on s’éloigne de  $\partial S$. On constate facilement que ceci ne dépend pas du choix de la fonction  $\chi$  et passe au quotient par  $H^0(S, \mathbb{C})$[^51]. L’application induite par  $\varepsilon$  est injective : si  $ \varphi.d\chi = d\psi $  où  $ \psi \in \mathcal{C}^{\infty}_{c\,\cap\, S}(Y)$, on aura  $ d(\psi - \chi.\varphi) = 0 $  sur  $Y$, et donc  $\psi - \chi.\varphi $  sera localement constante (donc constante) sur  $Y$. Comme  $\psi$  est nulle et  $\chi$  vaut identiquement  $1$  près de  $\partial S$, ceci montre que  $\varphi$  se prolonge en une constante sur  $Y$  et donc induit la classe nulle dans  $H^1_{\lbrace 0 \rbrace} (S, \mathbb{C})$. L’application induite par  $\varepsilon$  est surjective : si  $\alpha \in \mathcal{C}^{\infty}_{c\,\cap\, S}(Y)^1 $  est  $d-$fermée, on peut l’écrire, puisque  $Y$  est contractible,  $ \alpha = d\beta $  avec  $ \beta \in \mathcal{C}^{\infty}(Y)^0$. On en déduit que  $\beta$  est localement constante au voisinage de  $\partial S$. Soit  $\varphi$  la fonction localement constante au voisinage de  $S^*$  qu’elle définie. Alors l’égalité  $\alpha - d(\chi.\varphi) = d\big((1 - \chi).\beta\big) $  montre, puisque  $ (1 - \chi).\beta $  est à support dans  $ c\,\cap\, S$  que l’on a  $\varepsilon([\varphi]) = [\alpha]$. Il nous reste à montrer que l’application évidente $$H^1_{c\,\cap\,S}(Y, \mathbb{C}) \rightarrow H^1_{c\,\cap\,S}(0)$$ est un isomorphisme. Elle est injective : si  $\alpha \in \mathcal{C}^{\infty}_{c\,\cap\, S}(Y)^1 $  est  $d-$fermée et si  $ j_{1,k}(\alpha) = \delta \beta $  avec  $ \beta \in \Gamma_{c\,\cap\,S}(Y,\mathcal{E}^0(k))$, on a $$\begin{aligned} \quad \quad & \alpha = d\beta_k - \frac{df}{f} \wedge \beta_{k-1} \\ {\rm et} \quad & d\beta_j - \frac{df}{f} \wedge \beta_{j-1} = 0 \quad \forall j \in [1,k-1]\end{aligned}$$ avec la convention  $\beta_0 : = 0$. Alors  $ d\beta_1 = 0 $  et donc  $\beta_1 = 0 $  puisque c’est une fonction localement constante et nulle près de  $\partial S$. On en déduit de même que les  $\beta_j$  sont nulles pour  $ j \in [1,k-1]$. Il nous reste alors seulement l’équation sur  $Y$ $$\alpha = d\beta_k .$$ Mais une fonction semi-méromorphe dont la différentielle est  $\mathcal{C}^{\infty}$  est  $\mathcal{C}^{\infty}$[^52]. On a donc  $[\alpha] = 0 $  dans  $ H^1_{c\,\cap\,S}(Y, \mathbb{C})$. Elle est surjective : si  $ \gamma \in \Gamma_{c\,\cap\,S}(Y, Ker\,\delta^1(k))$, comme on a la nullité de  $H^0_{c\,\cap\,S}(Y, H^1(0)) $  ainsi que l’isomorphisme  $ \underset{k \rightarrow \infty}{\lim} \ h^1(k) \simeq H^1(0)$, quitte à augmenter  $k$, on peut supposer que  $ \alpha \in \Gamma_{c\,\cap\,S}(Y, \delta\mathcal{E}^0(k)).$ Mais la suite exacte $$0 \rightarrow h^0(k) \rightarrow \mathcal{E}^0(k)\rightarrow \delta \mathcal{E}^0(k)\rightarrow 0$$ montre que l’on a un isomorphisme $$\partial : \frac{ \Gamma_{c\,\cap\,S}(Y, \delta\mathcal{E}^0(k))}{\delta \Gamma_{c\,\cap\,S}(Y, \mathcal{E}^0(k))} \rightarrow H^1_{c\,\cap\,S}(Y, h^0(k)) \simeq H^1_{c\,\cap\,S}(Y, \mathbb{C}).$$ On vérifie alors facilement que l’application considérée co[ï]{}ncide avec  $ \partial^{-1}$. Vérifions enfin la dernière partie de l’assertion 3). On doit donc examiner le cas où la $2-$ forme fermée  $\hat{w}$  peut être choisie  $\mathcal{C}^{\infty} $  sur  $X^*$. Mais comme on a,  $ H^2_{mod\,S}(X\setminus\bar{X}', \mathbb{C}) \simeq H^1(S^*, \mathbb{C})$, cela signifie que l’on peut choisir  $ \hat{w} = d\chi \wedge \varphi $  où  $\varphi$  est une $1-$forme  $\mathcal{C}^{\infty} $  et  $d-$fermée au voisinage de  $S^*$  et où la fonction  $\chi$ est identiquement égale à 1 près de  $S^*$. Mais alors la classe initiale est dans  $j(H^1(S^*,\mathbb{C})$  d’après la remarque importante du (8.13) et de sa variante dans (8.14) Terminons ce paragraphe 8 en redonnant l’énoncé de notre résultat : Théorème 2. ----------- Applications et un exemple pour conclure. ========================================= Il n’est pas difficile de déduire les analogues pour la valeur propre  $1$  des théorèmes 13 et 14 de  $[B.91]$  de l’égalité entre  $ Im(\theta) $  et l’orthogonale pour  $\tilde{h}$  de  $H^0(S, H^{n-1}(0))$  dans  $H^1(S^*, H^{n-1}(0))$. On obtient les théorèmes suivants dont la ligne de démonstration, maintenant que l’on sait que  $Im(\theta)$  est exactement l’orthogonal de  $H^0(S, H^{n-1}(0))$  dans  $ H^1(S^*, H^{n-1}(0))$  pour  $\tilde{h}$, suit pas à pas celle de \[B.91\]. Théorème 3. ----------- *Sous les hypothèses standards pour la valeur propre 1, considérons  $e \in H^n(0) $  vérifiant  $\mathcal{N}^k(e) = 0 $  avec  $k \geq k_0$[^53]. Alors les conditions suivantes sont équivalentes :* - [1)]{}  $\tilde{ob}_k(e) = 0$ . - [2)]{} *Si  $w \in \Gamma(Y, Ker\,\delta^n(k))$  vérifie  $ r^n(k)(w) = e $, pour chaque  $j \in \mathbb{Z} $, les fonctionnelles analytiques $$P_{k+l} \big(\lambda = 0, \int_X \vert f \vert^{2\lambda} \bar{f}^{-j}\frac{df}{f}\wedge \bar{w}_k \wedge \square \big)$$ sont nulles pour  $l \geq 2$.* *De plus, pour vérifier 2) il suffit de le faire pour  $l = 2$  et pour  $j \in [0,n] $.* Théorème 4. ----------- *Sous les hypothèses standards pour la valeur propre 1, supposons que le prolongement méromorphe de  $ \int_X \vert f \vert^{2\lambda} \square $  admette un pôle d’ordre  $ \geq k $  en un entier négatif, avec  $ k \geq \sup{(k_0, k_1)} + 2$  où  $k_0$  et  $k_1$  sont respectivement les ordres de nilpotence de la monodromie agissant sur le système local  $ H^{n-1}(0)_{\vert S^*} $  et sur l’espace vectoriel  $H^n(0)$.* *Alors l’ordre des pôles aux entiers négatifs assez grand est exactement égal à  $k$  et l’on a  $ k_0 \leq k_1 = k - 2 $  avec  $\tilde{ob}_{k_1} \not\equiv 0 .$* Exemple. -------- Considérons la fonction  $ f(x,y,z) = x^2.(x^2 + y^2) + z^4 .$ Comme cet exemple est très similaire à celui étudié en détail à la fin de \[B.91\], nous allons seulement esquisser son étude. Tout d’abord l’homogéne[ï]{}té de  $f$  nous assure que la monodromie agit de fa[ç]{}on semi-simple sur  $H^2(0).$ Cependant, nous allons montrer que le prolongement méromorphe de $$\int_{X} \vert f \vert^{2\lambda}\rho. \vert z \vert^2 dx\wedge d\bar{x}\wedge dy\wedge d\bar{y}\wedge dz\wedge d\bar{z}$$ admet en  $ \lambda = -1 $  un pôle triple, si  $\rho \in \mathcal{C}^{\infty}_c(X)$  vaut identiquement 1 près de l’origine. Pour cela il suffit, par transformation de Mellin complexe, de montrer que le développement asymptotique, quand  $ \vert s \vert \rightarrow 0$  de l’intégrale $$\int_{f = s} \rho. \vert z \vert^2\frac{ dx\wedge dy}{z^2}\wedge \frac{ d\bar{x}\wedge d\bar{y}}{\bar{z}^2}$$ commence par un terme non nul en  $ (Log\vert s\vert )^2$, puisque l’on a sur  $ \lbrace f = s \rbrace$ $$4.z.dx\wedge dy\wedge dz \big/ df = \frac{ dx\wedge dy}{z^2} .$$ En rempla[ç]{}ant la fonction de troncature  $\rho$  par la fonction caractéristique du polydisque  $ (\vert x \vert \leq 1) \cap( \vert y \vert \leq 2) $ (voir \[B.91\] pour justifier que cela ne modifiera pas le pôle triple cherché) on est ramené à étudier le premier terme du développement asymptotique de la fonction $$\varphi(s) = \int_{(\vert x \vert \leq 1 )\cap (\vert y \vert \leq 2)} \frac{dx\wedge d\bar{x}\wedge dy\wedge d\bar{y}}{\vert s - x^2(x^2 + y^2)\vert} .$$ Après le changement de variable  $ x = u.s^{\frac{1}{4}} , y = t.u.s^{\frac{1}{4}} $  on trouve $$\varphi(s) = c.\int_{(\vert u\vert \leq \vert s \vert^{-\frac{1}{4}}) \cap (\vert t.u \vert \leq 2.\vert s \vert^{-\frac{1}{4}})} \frac{\vert u \vert^2 .du\wedge d\bar{u}\wedge dt \wedge d\bar{t}}{\vert 1 - u^4.(1 + t^2)\vert}$$ où  $c$  est une constante non nulle. Pour  $ u$  fixé, on a $$\int_{(\vert t.u \vert \leq 2.\vert s \vert^{-\frac{1}{4}})} \frac{dt\wedge d\bar{t}}{\vert 1 - u^4.(1 + t^2)\vert} \simeq c'. \vert u \vert^{-4}.Log(\vert u.s^{\frac{1}{4}}\vert)$$ ce qui donne $$\varphi(s) \simeq c''.(Log\vert s \vert )^2$$ avec  $c'$  et  $c"$  des constantes non nulles. Ceci montre que l’on a bien un pôle triple. Il est facile de donner un élément non nul de  $H^1_{\lbrace 0 \rbrace}(S, H^1(0)) $  dans cet exemple: Fixons  $ 0 < \alpha \ll \varepsilon $. Près d’un point du cercle  $ \lbrace x = z = 0\ ; \vert y \vert = \alpha \rbrace $, on peut choisir comme coordonnées locales  $ \xi = x.\sqrt{y^2 + x^2} , y , z $  de sorte que la fonction  $f$  dans ces coordonnées se réduit à  $ \xi^2 + z^4 .$ Posons $$v : = 2x.d\xi - \xi.dx .$$ Alors la forme différentielle holomorphe  $z.v$, vérifie  $ d(z.v) = \frac{df}{f}\wedge z.v $  et elle induit une section globale, uniforme et non nulle[^54]du système local  $H^1(0)$  le long de ce cercle. Le lecteur se convaincra facilement qu’elle n’est pas prolongeable à l’origine et donne donc un élément non nul de  $H^1_{\lbrace 0 \rbrace}(S, H^1(0)) $. Appendice I : Description topologique de  $H^n_{c\,\cap\,S}(0).$ {#appendice-i-description-topologique-de-hn_ccaps0. .unnumbered} ================================================================ Soit  $ \mathbb{H} \overset{exp(2i\pi.\Box)}{\longrightarrow} D^* $  le revêtement universel du disque pointé, et posons $$\hat{X} : = X \underset{D}{\times} \mathbb{H} .$$ Pour chaque voisinage ouvert  $\mathcal{U}$  de  $ S\, \cap\, \partial X $  notons par  $\hat{\mathcal{U}}$  l’image réciproque sur  $\hat{X}$  de  $\mathcal{U}$. Définissons la famille paracompactifiante  $\Phi$  de fermés de  $\hat{X}$  en posant $$G \in \Phi \quad {\rm ssi} \quad \exists \ \mathcal{U} \supset S \cap \partial X \quad / \quad G \cap \hat{\mathcal{U}} = \emptyset .$$ On remarquera que cette famille de supports est invariante par l’action de l’automorphisme de monodromie de  $\hat{X}$  induite par l’automorphisme\ $\zeta \rightarrow \zeta +1 $  du demi-plan  $\mathbb{H}$. Nous allons montrer la [**Proposition.**]{} *On a un isomorphisme “naturel” d’espaces vectoriels monodromiques* $$\varepsilon : H^n_{c\,\cap\, S}(0) \rightarrow H^n_{\Phi}(\hat{X}, \mathbb{C})_{=1}.$$ *Démonstration.* Commen[ç]{}ons par remarquer que l’on a un difféomorphisme compatible à  $f$[^55], grace à J. Milnor \[Mi.68\] $$\hat{X} \rightarrow F \times \mathbb{H}$$ où  $F$  dénote la fibre de Milnor de  $f$  à l’origine. En particulier on a un isomorphisme “naturel” d’espaces vectoriels monodromiques  $ H^n(\hat{X}, \mathbb{C}) \simeq H^n(F, \mathbb{C}) $  qui donne, pour la valeur propre  $1$  l’isomorphisme monodromique $ H^n(\hat{X}, \mathbb{C})_{=1} \simeq H^n(0).$ Définissons  $\varepsilon$. Soit donc  $ w \in \Gamma_{c\,\cap\,S}(Y, \mathcal{E}^n(k))\,\cap \,Ker\,\delta $  et définissons sur  $\hat{X}$  la  $n-$forme  $\mathcal{C}^{\infty} $ $$W : = \sum_{j=0}^{k-1} \frac{(-\zeta)^j}{j!}.p^*(w_{k-j})$$ où  $ p : \hat{X} \rightarrow X $  est la projection et où  $\zeta \in \mathbb{H}.$ On a  $dW = 0$  car on a supposé que  $\delta w = 0 $, et on a sur  $\hat{X}$  l’égalité  $ p^*(\frac{df}{f}) = d\zeta .$ De plus, le fait que le support de  $w$  rencontre  $S$  suivant un compact montre que  $W$  est à support dans  $\Phi$. On en déduit facilement que l’application  $\varepsilon$  est bien définie en posant $$\varepsilon ([w]) = [W] .$$ Pour vérifier que  $\varepsilon$  commute aux monodromies, il nous suffit alors d’établir la relation $$\varepsilon(\mathcal{N}(w)) = - \frac{\partial W}{\partial\zeta} \qquad \qquad (@)$$ puisque d’après la formule de Taylor, on a pour un polynôme  $P$ $$exp(\frac{\partial}{\partial\zeta})(P)(\zeta) = P(\zeta + 1).$$ La vérification de (@) est immédiate. On en conclut que la classe  $[W]$  est bien dans  $ H^n_{\Phi}(\hat{X}, \mathbb{C})_{=1}.$ Montrons maintenant l’injectivité de  $\varepsilon$. Il suffit de considérer le cas d’une classe  $ [w]$  invariante par la monodromie qui donne  $0$  dans  $ H^n_{\Phi}(\hat{X}, \mathbb{C})_{=1}$, grace à l’aspect monodromique de  $\varepsilon$. Mais dans ce cas, l’image par l’application d’oubli de support de  $[W]$  dans  $H^n(\hat{X}, \mathbb{C})_{=1} \simeq H^n(0) $  montre que  $[w] $  est dans le noyau de  $can_{c\,\cap\,S}$. On a donc seulement à traiter le cas où  $ [w] = i([\alpha]) = [\tilde{\alpha}\wedge d\chi ]$  avec  $\alpha \in H^1_{\lbrace 0 \rbrace}(S, H^{n-1}(0))$  d’après le théorème 1[^56] Soit donc une telle classe invariante  $[w] = [\tilde{\alpha}\wedge d\chi ]$  vérifiant  $\varepsilon([w]) = 0 $. Ceci signifie que l’on peut trouver une  $(n-1)-$forme  $v \in \mathcal{C}_{\Phi}^{\infty}(\hat{X}) $  vérifiant  $dv = p^*(w) $. Fixons  $\mathcal{U}$  un voisinage ouvert de  $ S\, \cap\, \partial X $  assez petit pour que l’on ait  $ Supp(v)\, \cap\, \hat{\mathcal{U}} = \emptyset $. Supposons maintenant que  $[w] \not= 0 $  dans  $H^n_{c\,\cap\,S}(0)$  et  $ n \geq 3$. Comme on sait que  $ [w] = i[\alpha] $  le choix de la fonction  $\chi $[^57] permet de supposer que l’on a  $ Supp(\hat{w})\,\cap \,S \subset \mathcal{U} $  où  $\hat{w} = \tilde{\alpha}\wedge d\chi $  est également un représentant de la classe  $[w]$. La forme sesquilinéaire  $\mathcal{H}$  étant non dégénérée, on peut trouver une classe  $[w'] \in H^n(0) $  telle que l’on ait $$P_2 \big(\lambda = 0, \int_X \vert f \vert^{2\lambda} \frac{df}{f}\wedge \frac{\bar{df}}{\bar{f}}\wedge w \wedge \bar{w'}_k \big) \not= 0.$$ Soit  $\sigma \in \mathcal{C}_c^{\infty}(\mathcal{U}) $  valant identiquement  $1$  au voisinage de  $ S\,\cap \, Supp(w ) $. Remarquons que, quitte à choisir  $\mathcal{U}$  plus petit, on peut toujours supposer que  $ Supp(\sigma)\,\cap\,\mathcal{U} = \emptyset $. La formule de Stokes et le prolongement analytique donne alors $$P_1 \big(\lambda = 0, \int_X \vert f \vert^{2\lambda} d\sigma \wedge \frac{\bar{df}}{\bar{f}}\wedge w \wedge \bar{w'}_k \big) \not= 0.$$ ce qui signifie que dans le développement asymptotique en  $ s = 0 $  de l’intégrale fibre $$s\frac{\partial}{\partial s} \int_{f = s} \sigma.w\wedge\bar{w'}_k$$ le terme constant est non nul. Mais ceci est absurde car  $p^*(w) = dv $  et  $Supp(\sigma) \subset \mathcal{U} $  montrent que l’ intégrale $$\int_{f = s} \sigma.w\wedge\bar{w'}_k = - \int_{f = s} d\sigma \wedge v \wedge\bar{w'}_k$$ est  $\mathcal{C}^{\infty}$  près de  $ s = 0$[^58] puisque  $ Supp(v)\,\cap \hat{\mathcal{U}} = \emptyset . $ Terminons la preuve de l’injectivité de  $\varepsilon$  pour  $ n = 2$. Il suffit de voir que  $j(H^1(S^*, \mathbb{C})) $  s’injecte dans  $ H^2_{\Phi}(\hat{X}, \mathbb{C})_{=1}$. Si  $ \varphi$  est $d-$fermée au voisinage de  $S^*$  et  $\chi \equiv 1$  près de  $S^*$  et s’annule dès que l’on s’éloigne, on a  $ j[\varphi] = [ d\chi \wedge \varphi] $  et  $ \varepsilon[d\chi \wedge \varphi] = [ p^*(d\chi \wedge \varphi)] .$ Mais si  $ p^*(d\chi \wedge \varphi) = d\lambda $  où  $ \lambda \in \mathcal{C}^{\infty}_{\Phi}(\hat{X})^{n-1} $  on aura  $ p^*(\chi.\varphi) - \lambda $  qui sera $d-$fermée et localement $d-$exacte près de  $ \partial S $. La classe qu’elle définie dans  $ H^{n-1}(F, \mathbb{C}) $  est donc nulle[^59]. On en conclut que  $\varphi$  est globalement $d-$exacte au voisinage de  $\partial S$  et donc induit la classe nulle de  $H^1(S^*, \mathbb{C}).$ Montrons maintenant la surjectivité pour  $ n \geq 2$, en montrant par récurrence sur  $ k \geq 1$, que si  $ [W] \in H^n_{\Phi}(\hat{X}, \mathbb{C})_{=1}$  vérifie  $ (T - 1)^k [W] = 0 $, elle est bien dans l’image de  $\varepsilon$. Pour  $k = 1$, on a  $ T[W] = [W] $  et on peut trouver une $(n-1)-$forme  $V \in \mathcal{C}^{\infty}_{\Phi}(\hat{X}) $  qui vérifie  $ -\frac{\partial}{\partial \zeta} W = dV .$ On écrit alors[^60] $ V = -\frac{\partial}{\partial \zeta} U $  avec  $ U \in \mathcal{C}^{\infty}_{\Phi}(\hat{X})^{n-1}$. Alors  $ W - dU$  est un représentant $T-$invariant de la classe  $[W]$  et il existe une forme  $ w \in \mathcal{C}^{\infty}_{c\,\cap\,S}(X \setminus f^{-1}(0))^n $  vérifiant  $ dw = 0 $  et  $p^*(w) = W - dU .$ On conclut en utilisant le théorème de A.Grothendieck \[Gr.65\] pour trouver une $n-$forme semi-méromorphe $d-$fermée  $\tilde{w}$  à pôles dans  $f^{-1}(0)$  et à support  $c\,\cap\,S$  induisant la même classe que  $w$  dans  $ H^n_{c\,\cap\,S}(X \setminus f^{-1}(0), \mathbb{C}).$ Supposons maintenant le résultat montré pour  $ k \leq k_0$  avec  $ k_0 \geq 1$  et montrons-le pour  $ k = k_0 +1.$ Alors d’après l’hypothèse de récurrence, on peut écrire  $ -\frac{\partial}{\partial \zeta} W = \tilde{W} + dV $  où  $[ \tilde{W}]$  est dans l’image de  $ \varepsilon$  et où  $ V \in \mathcal{C}^{\infty}_{\Phi}(\hat{X})^{n-1}. $ On écrit à nouveau  $ V = -\frac{\partial}{\partial \zeta} U $  avec  $ U \in \mathcal{C}^{\infty}_{\Phi}(\hat{X})^{n-1}$  et arrive à  $ -\frac{\partial}{\partial \zeta}(W - dU) = \tilde{W} = \varepsilon(\tilde{w}) .$ Mais alors la $n-$forme $$\int_0^{\zeta} \tilde{W}(z).dz + (W - dU)$$ est invariante par la monodromie. Elle est donc image réciproque d’une forme  $\mathcal{C}^{\infty}$  et  $d-$fermée sur  $ X \setminus f^{-1}(0) $  à support dans  $ c\,\cap\,S$. On conclut alors facilement grâce au même argument que dans le cas  $ k = 1$, en intégrant terme à terme la formule donnant  $ \varepsilon(\tilde{w})$. $\hfill \blacksquare$ Appendice II : Interprétation en termes de complexes de cycles proches et évanescents.\ [*Par Claude Sabbah.*]{} {#appendice-ii-interprétation-en-termes-de-complexes-decycles-proches-et-évanescents.-par-claude-sabbah. .unnumbered} ======================================================================================= Soit $f:(\mathbb{C}^{n+1},0)\to(\mathbb{C},0)$ un germe de fonction holomorphe ($n\geq1$). Soit $\psi_f{\mathbb{Q}}$ (resp. $\phi_f{\mathbb{Q}}$) le complexe des cycles proches (resp. évanescents) de $f$ pour le faisceau constant ${\mathbb{Q}}$ (voir \[D. 73\] et aussi \[K.S. 90\]). Il est muni d’un opérateur de monodromie $T$. À un décalage de $n$ près, c’est un faisceau pervers (voir \[Br. 86\] et aussi \[K.S. 90\]). Il admet une décomposition suivant les valeurs propres $$\psi_f{\mathbb{Q}}=\oplus_{u\in{\mathbb{Q}}\cap[0,1[}\psi_{f,\exp(-2i\pi u)}{\mathbb{Q}}\quad \text{resp.~}\phi_f{\mathbb{Q}}=\oplus_{u\in{\mathbb{Q}}\cap[0,1[}\phi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$$ dans la catégorie des faisceaux pervers (voir *loc. cit.*). L’opérateur $T-\exp(-2i\pi u)\operatorname{Id}$ est nilpotent sur $\psi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$ (resp. $\phi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$). Pour chaque $u$, on dispose d’un morphisme “canonique” $$\operatorname{can}:\psi_{f,\exp(-2i\pi u)}{\mathbb{Q}}\to\phi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$$ et d’un morphisme “variation” $$\operatorname{var}:\phi_{f,\exp(-2i\pi u)}{\mathbb{Q}}\to\psi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$$ dont les composés $\operatorname{var}\circ\operatorname{can}$ et $\operatorname{can}\circ\operatorname{var}$ sont égaux à $T-\operatorname{Id}$ sur les complexes correspondants. En un point $x$ de $f^{-1}(0)$, le germe en $x$ du faisceau de cohomologie $H^p(u):={\mathcal{H}}^p\psi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$ est l’espace propre généralisé $H^p(F_x,{\mathbb{Q}})_{\exp(-2i\pi u)}$ de la monodromie, si $F_x$ est la fibre de Milnor (d’un bon représentant) de $f$ en $x$. Lorsque $u\neq0$, ces deux morphismes sont des isomorphismes, et on identifie $\psi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$ et $\phi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$ *via* $\operatorname{can}$, de sorte que $\operatorname{var}$ devient égal à $T-\operatorname{Id}$. Soit $i_{f^{-1}(0)}:\{f=0\}\hookrightarrow {\mathbb{C}}^{n+1}$ l’inclusion. Lorsque $u=0$, on a un triangle de la catégorie dérivée des complexes bornés à cohomologie constructible sur $f^{-1}(0)$: $$\tag{$*$} i_{f=0}^{-1}{\mathbb{Q}}\to\psi_{f,1}{\mathbb{Q}}\to\phi_{f,1}{\mathbb{Q}}\To{+1}$$ qui dit que les faisceaux de cohomologie de $\phi_{f,1}{\mathbb{Q}}$ sont les faisceaux de cohomologie réduite de $\psi_{f,1}{\mathbb{Q}}$. *Hypothèse*. Fixons une valeur propre $\exp(-2i\pi u)$ ($u\in[0,1[\cap{\mathbb{Q}}$). Nous faisons l’hypothèse que le complexe de faisceaux $\phi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$ est à support sur une *courbe* $(S,0)$. Nous noterons $$i:\{0\} \hookrightarrow S,\quad j:S\setminus \{0\}\hookrightarrow S,\quad i_S:S \hookrightarrow \{f=0\},$$ les inclusions. Pour simplifier, nous noterons $\psi_{f,\exp(-2i\pi u)},\phi_{f,\exp(-2i\pi u)}$ les complexes $\psi_{f,\exp(-2i\pi u)}{\mathbb{Q}},\phi_{f,\exp(-2i\pi u)}{\mathbb{Q}}$. **1.** La perversité de $\phi_{f,\exp(-2i\pi u)}$ implique que $\phi_{f,\exp(-2i\pi u)}$ n’a de cohomologie qu’en degré $n$ et $n-1$. Le faisceau ${\mathcal{H}}^n\phi_{f,\exp(-2i\pi u)}$, noté ${\mathcal{H}}^n$, est à support l’origine, et le faisceau ${\mathcal{H}}^{n-1}\phi_{f,\exp(-2i\pi u)}$, noté ${\mathcal{H}}^{n-1}$, est à support sur $S$. On en déduit (sans hypothèse si $u\neq0$ et sous l’hypothèse $n\geq2$ si $u=0$, en utilisant le triangle $(*)$) que, - ${\mathcal{H}}^n$ est la partie correspondant à la valeur propre $\exp(-2i\pi u)$ du $n$-ième groupe de cohomologie de la fibre de Milnor de $f$ en $0$, - ${\mathcal{H}}^{n-1}$ est le faisceau analogue pour la cohomologie en degré $n-1$. Comme le faisceau constant ${\mathbb{Q}}$ sur ${\mathbb{C}}^{n+1}$ est auto-dual (par dualité de Poincaré-Verdier) à décalage près de $2(n+1)$, on déduit, puisque le foncteur des cycles évanescents commute à la dualité à un décalage de $1$ près (voir \[Br. 86\] et \[K.S. 90\]), que $\phi_{f,\exp(-2i\pi u)}$ est dual de $\phi_{f,\exp(2i\pi u)}$ à un décalage de $2n$ près. D’un autre côté, le foncteur $i^{-1}$ est transformé en $i^!=R\Gamma_{0}$ à un décalage de $2$ près. Le complexe $i^{-1}\phi_{f,\exp(-2i\pi u)}$ est un complexe d’espaces vectoriels ayant deux groupes de cohomologie ${\mathcal{H}}^n$ et ${\mathcal{H}}^{n-1}_0:=i^{-1}{\mathcal{H}}^{n-1}$ (dans les degrés indiqués). Par dualité de Poincaré-Verdier, on en déduit que $i^!\phi_{f,\exp(2i\pi u)}$ n’a que deux groupes de cohomologie, qui sont $${\mathcal{H}}^n(i^!\phi_{f,\exp(2i\pi u)})\simeq{{\mathcal{H}}^n}^\vee\quad\text{et}\quad{\mathcal{H}}^{n+1}(i^!\phi_{f,\exp(2i\pi u)})\simeq{{\mathcal{H}}^{n-1}_0}^\vee$$ (où ${}^\vee$ désigne le dual des espaces vectoriels). En particulier, on a un accouplement non dégénéré ${\mathcal{H}}^n(i^!\phi_{f,\exp(2i\pi u)}) \otimes{\mathcal{H}}^n\to{\mathbb{Q}}$. **2.** Puisque $\phi_{f,\exp(-2i\pi u)}$ n’a que deux faisceaux de cohomologie, on a un triangle $${\mathcal{H}}^{n-1}\phi_{f,\exp(-2i\pi u)}[-(n-1)]\to\phi_{f,\exp(-2i\pi u)}\to{\mathcal{H}}^n[-n]\To{+1}$$ et donc, en appliquant $i^!$, on a un triangle correspondant, qui donne une suite exacte longue $$\begin{gathered} 0\rightarrow H_{0}^{1}(S,{\mathcal{H}}^{n-1})\rightarrow{\mathcal{H}}^n(i^!\phi_{f,\exp(-2i\pi u)})\rightarrow{\mathcal{H}}^n\rightarrow H_{0}^{2}(S,{\mathcal{H}}^{n-1})\\ \rightarrow{\mathcal{H}}^{n+1}(i^!\phi_{f,\exp(-2i\pi u)})\rightarrow0\end{gathered}$$ avec $0$ aux deux bouts puisque ${\mathcal{H}}^n$ est à support l’origine. Pour la même raison, le morphisme du milieu $\operatorname{can}:{\mathcal{H}}^n(i^!\phi_{f,\exp(-2i\pi u)})\to{\mathcal{H}}^n$ est le $H^n$ du morphisme de complexes $i^!\phi_{f,\exp(-2i\pi u)}\to i^{-1}\phi_{f,\exp(-2i\pi u)}$ (oubli des supports). **3. Interprétation de ${\mathcal{H}}^n(i^!\phi_{f,\exp(-2i\pi u)})$.** J’interprète le groupe $H^n_{c\cap S}(u)$ comme le groupe de cohomologie de degré $n$ du complexe $i^!i_S^{-1}\psi_{f,\exp(-2i\pi u)}$. Si $u\neq0$, on a donc ${\mathcal{H}}^n(i^!\phi_{f,\exp(-2i\pi u)})=H^n_{c\cap S}(u)$. Supposons maintenant que $u=0$. Après application de $i_S^{-1}$, le triangle $(*)$ devient, puisque $\phi_{f,1}$ est à support dans $S$ (donc $i_S^{-1}\phi_{f,1}=\phi_{f,1}$) $${\mathbb{Q}}_S\to i_S^{-1}\psi_{f,1}\to\phi_{f,1}\To{+1}$$ et, après application de $i^!$, $$i^!{\mathbb{Q}}_S\to i^!i_S^{-1}\psi_{f,1}\to i^!\phi_{f,1}\To{+1}.$$ La cohomologie du complexe $i^!{\mathbb{Q}}_S=R\Gamma_{0}{\mathbb{Q}}_S$ n’existe qu’en degré $2$. Donc $${\mathcal{H}}^n(i^!\phi_{f,1})= \begin{cases} H^n_{c\cap S}(0)&\text{si }n\geq3,\\ H^n_{c\cap S}(0)/\text{image}\big(H^2_0({\mathbb{Q}}_S)\big)&\text{si }n=2, \end{cases}$$ où $H^2_0({\mathbb{Q}}_S)\simeq H^1(S^*,{\mathbb{Q}})$. Bibliographie. ============== - [\[A.V.G\]]{} Arnold, V., Varchenko, A. et Goussein-Zadé, S. *Singularités des applications différentiables*, vol.2 Monodromie et comportement asymptotique des intégrales, Moscou 1986 (traduction fran[ç]{}aise, Ed. Mir). - [\[B.84.a\]]{} Barlet, D. *Contribution effective de la monodromie aux développements asymptotiques*, Ann. Scient. ENS 4-ième série, 17 (1984) p.239-315. - [\[B.84.b\]]{} Barlet, D. *Contribution du cup-produit de la fibre de Milnor aux pôles de  $\vert f \vert^{2\lambda}$* , Ann. Inst. Fourier (Grenoble), 34 (1984) p.75-107. - [\[B.85\]]{} Barlet, D. *La forme hermitienne canonique sur la cohomologie de la fibre de Milnor d’une hypersurface à singularité isolée*, Invent. Math. 81 (1985) p.115-153 . - [\[B.90\]]{} Barlet, D. *La forme hermitienne canonique pour une singularité presque isolée*, in Complex Analysis (K.Diederich eds) Vieweg,Wuppertal (1990) p.20-28 . - [\[B.91\]]{} Barlet, D. *Emmêlements de strates consécutives pour les cycles évanescents*, Ann. Scient. ENS 4-ième série 24 (1991) p.401-506. - [\[B.97\]]{} Barlet, D. *La variation pour une hypersurface à singularité isolée relativement à la valeur propre 1* , Revue de l’Inst. E. Cartan (Nancy) 15 (1997) p.1-29 . - [\[B.02\]]{} Barlet, D. *Singularités réelles isolées et développements asymptotiques d’intégrales oscillantes*, in Séminaire et Congrès 9 (Actes des journées mathématiques à la mémoire de J. Leray) Guillopé, L. et Robert, D. éditeurs, Société Mathématique de France (2004), p. 25-50. - [\[B.03\]]{} Barlet, D. *Interaction de strates consécutives II* , Publ. du RIMS Kyoto Univ., vol. 41 (2005), p. 139-173. - [\[B.04\]]{} Barlet, D. *Sur certaines singularités non isolées d’hypersurfaces I*, preprint Institut E.Cartan (Nancy) 2004/$n^{\circ}$ 03. - [\[B.M.89\]]{} Barlet, D. et Maire, H.-M. *Asymptotic expansion of complex integrals via Mellin transform*, Journ. Funct. Anal. 83 (1989) p.233-257. - [\[Br. 86\]]{} Brylinski, J.-L. *Transformations canoniques, dualité projective, théorie de [Lefschetz]{}, transformation de [Fourier]{} et sommes trigonométriques*, in [*Géométrie et analyse microlocales*]{}, Ast[é]{}risque, vol. 140-141, Soci[é]{}t[é]{} Math[é]{}matique de France, 1986, p.3-134. - [\[D. 73\]]{} Deligne, P. *Le formalisme des cycles évanescents (exposés 13 et 14)*, in [*SGA 7 II*]{}, Lect. Notes in Math., vol.340, Springer-Verlag, 1973, p.82-173. - [\[Go.58\]]{} Godement, R. *Théorie des faisceaux*, Hermann (1958) Nancago. - [\[Gr.65\]]{} Grothendieck, A. *On the De Rham cohomology of algebraic varieties*, Publ. Math. IHES, 29 (1966) p.96-103. - [\[K.76\]]{} Kashiwara, M. *b-Function and Holonomic Systems, Rationality of Roots of b-Functions*, Invent. Math. 38 (1976) p.33-53. - [\[K.84\]]{} Kashiwara, M. *The Riemann-Hilbert Problem for Holonomic Systems*, Publ. RIMS Kyoto Univ. 20 (1984) p.319-365. - [\[K. S. 90\]]{} Kashiwara, M. and Schapira, P. *Sheaves on Manifolds*, [*Grundlehren der mathematischen Wissenschaften*]{}, vol.292, Springer-Verlag, 1990. - [\[Le.59\]]{} Leray, J. *Le problème de Cauchy III* , Bull.Soc. Math. France 87 (1959) p.81-180 . - [\[Lo.86\]]{} Loeser, F. *A propos de la forme hermitienne canonique d’une singularité isolée d’hypersurface*, Bull. Soc. Math. France, 114 (1986) p.385-392. - [\[Ma.74\]]{} Malgrange, B. *Intégrales asymptotiques et monodromie*, Ann. Scient. ENS 4-ième serie 7 (1974) p.405-430 . - [\[Mi.68\] ]{} Milnor, J. *Singular Points of Complex Hypersurfaces*, Ann. of Math. Studies 61 (1968) Princeton . [^1]: Université Henri Poincaré et Institut Universitaire de FranceInstitut Elie Cartan (Nancy) UMR 7502 UHP-CNRS-INRIABP 239 - F - 54506 Vandoeuvre-lès-Nancy Cedex.France.e-mail : [email protected] [^2]: Centre de Mathématiques Laurent Schwartz, Ecole Polytechnique, UMR 7640 CNRSe-mail [email protected] [^3]: The case  $n = 2$, that is to say the case of an hypersurface in  $\mathbb{C}^3$, is analoguous but special. [^4]: Here again the  $n = 2$  is analoguous but special. [^5]: Le cas  $n = 2 $, c’est à dire le cas d’une hypersurface de  $\mathbb{C}^3$, est analogue mais spécial. [^6]: La encore le cas  $n = 2$ est analogue mais spécial. [^7]: Voir également \[B.03\]. [^8]: Ce sont donc les supports des deux premiers faisceaux de cohomologie non nuls du complexe à cohomologie constructible des cycles évanescents de  $\tilde{F}$  relatif à la valeur propre 1 de la monodromie, que l’on notera  $ \psi_{\tilde{F}, 1}.$ [^9]: C’est dans cette hypothèse, qui est “générique”, que l’adjectif “consécutif” prend son sens dans l’expression “strates consécutives”:  $\Sigma_0$  est de codimension 1 dans  $\Sigma_1$. [^10]: Grace au théorème de restriction non caratéristique (voir \[B.M.89\]) et à la correspondance de Riemann Hilbert (voir \[K.84\]). [^11]: Dans le cas  $n=1$  le phénomène d’emmêlement de strates pour la valeur propre 1 que nous allons étudier n’apparaît pas, même pour une fonction non réduite. Le cas où  $n=1$  sera considéré dans la suite sous l’hypothèse plus forte d’une singularité isolée pour la valeur propre 1 (notion déjà étudiée dans \[B.90\] et \[B.97\] ). Ce cas sera essentiel quand on étudiera la situation d’une section hyperplane transverse en un point générique de la courbe  $S$. [^12]: En effet, pour une courbe plane réduite et irréductible, la valeur propre 1 de la monodromie n’apparait pas. [^13]: Puisque  $H^{n-1}(0)$  n’a pas de section non nulle à support l’origine. [^14]: Elle est sesquilinéaire, et  $( id\times can_{c\,\cap\,S})^*(\mathcal{H})$  est hermitienne sur  $ H^n_{c\,\cap\,S}(0) \times H^n_{c\,\cap\,S}(0) $. [^15]: Rappelons que, par définition,  $k_0$  est l’ordre de nilpotence de la monodromie agissant sur le système local  $H^{n-1}(0)$  sur  $S^*$. [^16]: En fait on a  $ \tau_{k,k'}\circ \delta_0 + \delta_0 \circ \tau_{k,k'} = 0.$ [^17]: de fa[ç]{}on unipotente. [^18]: Nous omettrons dorénavant l’indice de la différentielle  $ \delta = \delta_0$. [^19]: Pour une discussion plus détaillée de la relation entre  $r^n(k)$  et  $\tilde{r}^n(k)$   on pourra consulter la Remarque/Erratum qui suit le Corollaire 1 du Théorème 1 de \[B.02\]. [^20]: Donc  $H^p(0) = 0 $  pour  $ p \not= n$  puisque, par perversité, seul le faisceau  $H^n(0)$  peut avoir un support de dimension  $0$. [^21]: Au sens “usuel”, c’est à dire que l’on a  $ df_x \not= 0 $  pour  $ x \not= 0 .$ [^22]: Ce point est détaillé dans \[B.97\] ; il sera repris et généralisé plus loin; voir le 5. [^23]: On prendra garde qu’à l’origine, cette construction n’a, à priori, pas de sens. [^24]: Comme l’application de restriction  $ \Gamma(S, H^{n-1}(0)) \rightarrow \Gamma(S^*, H^{n-1}(0))$  est injective (perversité),  $k_0$  majore également l’ordre de nilpotence de la monodromie agissant sur la fibre à l’origine de  $H^{n-1}(0)$. [^25]: Grace aux isomorphismes rappelés plus haut ; voir (1.8). [^26]: Comme dans \[B.91\], si  $\rho \in \mathcal{C}^{\infty}_c(X)$  vaut identiquement 1 près de l’origine,  $ \gamma = d\rho$  convient. [^27]: En fait le lemme 2 p. 442 de \[B.91\] donne une homotopie  $t_k : \mathcal{E}^{\bullet}(k) \rightarrow \mathcal{E}^{\bullet}(k)$  vérifiant  $\tau_k \circ _k\mathcal{N} - _k\mathcal{N}\circ \tau_k = \delta \circ t_k - t_{k}\circ \delta $  qui, combinée avec l’égalité évidente (dans  $ Hom(\mathcal{E}^{\bullet}(k), \mathcal{E}^{\bullet}(k+k')$ ) : $$j_{k,k+k'} \circ \tau_k = ( _{k+k'}\mathcal{N}^{k'})\circ \tau_{k+k'}\circ j_{k,k+k'}$$ permet de raisonner directement et redonne (@). [^28]: En fait, de fa[ç]{}on plus précise, l’application  $ j_{k,k+1}$  induit  $0$  sur  $h^1(k)$. [^29]: Rappelons que  $H^{n-1}(0)$  est un système local sur  $S^*$. [^30]: Grace à l’annulation de  $ H^1(S, \delta\mathcal{E}^{n-2}(k))$  qui est prouvée au (4.7). [^31]: Car  $H^1(0)$  n’a pas de section non nulle à support l’origine. [^32]: quitte à passer à la limite inductive sur  $k$  pour  $n = 3$. [^33]: Où ici  $\varepsilon$  est le rayon de la boule de Milnor  $X$  que l’on considère. [^34]: Quitte à choisir  $k \gg1 $. [^35]: Pour  $n \geq 4$ ; pour  $n = 3$  le recollement a lieu seulement dans  $H^2(0)$. [^36]: Comme on l’a déja fait remarquer plus haut la formule  $ \tilde{ob}_{k+k'} = \tilde{ob}_k \circ \mathcal{N}^{k'} $  de \[B.91\] est encore valable ici ; donc pour  $k$  assez grand on a  $\tilde{ob}_k \equiv 0 .$ [^37]: En fait, dans le cas  $n = 2 $  on doit systématiquement remplacer l’espace  $H^2_{c\,\cap\,S}(0) $  par son quotient par  $H^1(S^*, \mathbb{C})$  et donc  $ \mathcal{K}$  par son quotient par  $H^1(S^*, \mathbb{C})$  pour avoir les mêmes résultats que pour  $n \geq 3 $. [^38]: Ceci est possible grace à \[Go.58\] th. 4.11.1 et à de Rham. [^39]: On a noté ici par  $\mathcal{G}_{mod\,S}$  le sous faisceau du faisceau  $\mathcal{G}$  des sections qui sont nulles au voisinage de  $S$. On a donc, par définition la suite exacte courte : $$0 \rightarrow \mathcal{G}_{mod\,S} \rightarrow \mathcal{G} \rightarrow j_! j^*(\mathcal{G}) \rightarrow 0$$ où  $j : X \setminus S \hookrightarrow X $  désigne l’inclusion. [^40]: L’opérateur  $_k\mathcal{N}$  est nul sur  $h^0(k)$  et, en dehors de  $S$, sur  $h^1(k)$. Attention, pas sur  $S$  pour  $n=2$ ! [^41]: On peut s’en convaincre en montrant par récurrence sur  $p$  que pour un sous-ensemble analytique fermé de dimension  $p$  de  $\mathbb{C}^{n+1}$   les faisceaux de cohomologie à support  $\underline{H}^q_{S}(\mathbb{C}) $  sont nuls pour  $ q < 2n -2p +1 $. [^42]: Rappelons que les supports des parties polaires d’ordre  $\geq 2$  aux entiers du prolongement méromorphes de  $\vert f \vert^{2\lambda}$  sont contenus dans  $S_1$. [^43]: On utilise à nouveau le fait que les racines du polynôme de Bernstein de  $f$  sont strictement négatives ! Voir \[K.76\] [^44]: On peut choisir pour  $K$  la trace sur  $S_1$  d’une boule assez grosse. [^45]: ce qui signifie que l’adhérence de son support dans  $X^*$   ne rencontre pas  $S^*$. [^46]: Touts ces groupes de cohomologie sont à valeurs dans  $ \mathbb{C}.$ [^47]: Rappelons que  $X$  est l’intersection de la boule de centre $0$  et de rayon  $\varepsilon$  avec  $f^{-1}(D_{\eta}), \quad \eta \ll \varepsilon.$ [^48]: et tel que l’on ait  $Supp(d\rho) \cap Supp(\delta v) = \emptyset $. [^49]: muni de la monodromie triviale, c’est à dire égale à l’identité. [^50]: Grace, entre autres, au fait que le sous-espace  $j(H^1(S^*, \mathbb{C}))$ de  $H^2_{c\,\cap\,S}(0)$  est formé de vecteurs invariants par la monodromie. [^51]: Si  $\varphi$  est constante au voisinage de  $S$  on peut prolonger la constante à  $Y$  et alors  $ \varphi.d\chi = d\big((1 - \chi).\varphi \big) $  montre que la classe  $ [\varphi.d\chi] $  est nulle dans  $H^1_c(0)$. [^52]: On se ramène au cas méromorphe par Dolbeault (local) et on conclut par le lemme de de Rham holomorphe. [^53]: Rappelons que, par définition,  $k_0$  est l’ordre de nilpotence de la monodromie agissant sur le système local  $H^{n-1}(0)$  sur  $S^*$. [^54]: puisque le monôme  $z$  n’est pas dans l’idéal jacobien de  $ \xi^2 + z^4 .$ [^55]: Où  $f$  est définie sur  $\hat{X}$  comme la composée de la projection sur  $\mathbb{H}$  avec la fonction  $exp(2i\pi.\Box)$. [^56]: L’adaptation de ceci pour  $n = 2$  ne présente pas de difficulté et elle est laissée au lecteur. [^57]: voir la définition de l’application  $i$  au (4.9). [^58]: Plus exactement, la partie dans  $ \sum_{j = 0}^n \mathbb{C}[[s,\bar{s}]].(Log\vert s \vert^2)^j $  de son développement asymptotique en  $s = 0$  est en fait dans  $ \mathbb{C}[[s,\bar{s}]].$ [^59]: Car une section globale du faisceau  $H^{n-1}(0) $  qui est nulle près de  $\partial S$  est nulle. [^60]: C’est toujours possible : on utilise ici le complexe  $(\mathcal{E}^{\bullet}[Log\, f], d^{\bullet})$  où  $-2i\pi. \zeta = Log\,f $ sur  $\hat{X}$, pour calculer la partie spectrale pour la valeur propre 1 de la monodromie des cycles évanescents de  $f$.
{ "pile_set_name": "ArXiv" }
--- author: - | \ University of Hamburg\ E-mail: bibliography: - 'BKMRW.bib' title: 'Determining MSSM parameters via chargino production at the LC: a one-loop analysis' --- Introduction {#sec:1} ============ A linear collider (LC) is ideally suited to high precision studies of physics beyond the standard model (BSM). Light electroweakinos, motivated by naturalness arguments ref. [@Hall:2011aa] and GUT motivated SUSY models [@Brummer:2011yd], could well be within reach of a first stage LC. This would allow the nature of the underlying SUSY model to be probed via the determination of the fundamental parameters. LO studies (see ref. [@Desch:2003vw] and references therein) have shown that the U(1) parameter $M_1$, the SU(2) parameter $M_2$, the higgsino parameter $\mu$ and $\tan\beta$, the ratio of the vacuum expectation values of the two neutral Higgs doublet fields, can be determined at the percent level via chargino and neutralino pair-production. However, as one-loop effects in the MSSM can be large, higher order calculations are crucial in order to perform an accurate assessment. These corrections mean that additional MSSM parameters become relevant, such as the masses of the stops which are so far only weakly constrained by the LHC. Fitting experimental results to loop corrected predictions, calculated in the on-shell scheme, it should therefore be possible to extract the parameters of the chargino–neutralino sector of the MSSM Lagrangian, as well as to indirectly gain insight into other sectors. In sec. \[sec:one-loop\] we discuss the strategy used to fit the MSSM parameters and briefly describe the calculation of NLO corrections. Then in sec. \[sec:num-res\] we present our numerical results and finally in sec. \[sec:conc\] we discuss the implications of these results. Fit strategy and the calculation of NLO corrections {#sec:one-loop} =================================================== In the chargino and neutralino sectors there are four real parameters which we fit to: $M_1$, $M_2$, $\mu$ and $\tan\beta$. If beyond the direct reach of the LC, the sneutrino mass is also included in the fit. Depending on the scenario, only limited knowledge about the remaining MSSM parameters, e.g. for the stop sector, which will all contribute at the loop level, may be available. Our analysis also offers unique indirect sensitivity to such parameters at the LC. In the fit we use the polarised cross-sections and forward backward asymmetry for chargino production as well as the ${\tilde{\chi}}_1^{\pm},{\tilde{\chi}}_2^{\pm}$ and ${\tilde{\chi}^0_{1}}, {\tilde{\chi}^0_{2}}, {\tilde{\chi}^0_{3}}$ masses, calculated at NLO in an on-shell scheme as described below. We assume the masses to have been measured at the LC using the threshold scan method, however we also investigate how the fit precision changes if the masses were obtained from the continuum [@AguilarSaavedra:2001rg]. Further details of the fit method and errors are given in ref. [@Bharucha:2012ya]. We carry out the fit for two scenarios, S1 and S2, shown in tab. \[tab:s1\].[^1] Due to the current status of direct LHC searches [@atlas:2012rz; @*cms:2012jx], in both scenarios we require first and second generation squarks and gluinos to be heavy, but on the other hand assume the stop sector to be light. Further, indirect limits, checked using [`micrOmegas 2.4.1`]{} [@Belanger:2006is; @*Belanger:2010gh], lead us to choosing mixed gaugino higgsino scenarios, favoured by the relic density measurements [@Komatsu:2010fb] and relatively high pseudoscalar Higgs masses, preferred by flavour physics constraints i.e. the branching ratio $\mathcal{B}(b\to s\gamma)$ and $\Delta(g_\mu-2)/2$. Our scenarios are chosen in order to assess the fit sensitivity for different possible choices of parameters e.g. the pseudoscalar Higgs mass $M_{A^0}$. Note that for S2, $m_h$ (=125 GeV) is compatible with the recent Higgs results from the LHC [@ATLAS:2012gk; @*CMS:2012gu]. The one-loop corrections to the polarised cross-section and forward backward asymmetry for $e^+e^-\to \tilde{\chi}^+_1\tilde{\chi}^-_1$ are calculated in full within the MSSM, following [@bfmw; @*Bharucha:2012re], using the program `FeynArts` [@Denner:1992vza], which allowed an automated generation of the Feynman diagrams and amplitudes. Together with the packages `FormCalc` [@FormCalc2] and `LoopTools` [@Hahn:1998yk] we derived the final matrix elements and loop integrals. The regularisation method we use is dimensional reduction (see e.g. [@0503129]), which ensures that SUSY is preserved, as in e.g. ref. [@delAguila:1998nd]. The results at NLO depend on many MSSM parameters beyond the small subset at tree level. Finite results at one-loop order are obtained by adding counterterm diagrams, also generated by `FeynArts`, however expressions for the counterterms which renormalise the tree-level couplings are required as input, as given in detail in ref. [@Bharucha:2012ya]. Using this prescription to renormalise the vertices we obtain UV-finite results. Soft and collinear radiation must be included to obtain a result free of infra-red and collinear singularities. In the regions $E<\Delta E$ and $\theta<\Delta \theta$ where $\Delta E$ and $\Delta \theta$ denote the cut-offs, the radiative cross-section can be factorised into analytically integrable expressions proportional to the tree-level cross-section $\sigma^{\rm tree}(e^+e^-\to \tilde{\chi}_i^+\tilde{\chi}_j^-)$. The soft contribution can easily be incorporated using `FormCalc`, however the collinear contribution must be added explicitly, this is done as described in ref. [@Bharucha:2012ya]. Numerical results {#sec:num-res} ================= [lclclclc]{} &\ \ &\ ------------------------------------------------------------------------ Parameter & Value & Parameter & Value& Parameter & Value & Parameter & Value\ ------------------------------------------------------------------------ $M_1$ & 125& $M_2$ & 250&$M_1$ & 106& $M_2$ & 212\ $\mu$ & 180 & $M_{A^0}$ & 1000&$\mu$ & 180 & $M_{A^0}$ & 500\ $M_3$ & 700 & $\tan\beta$ & 10&$M_3$ & 1500 & $\tan\beta$ & 12\ $M_{e_{1,2}}$ & 1500 & $M_{e_{3}}$ & 1500&$M_{e_{1,2}}$ & 125 & $M_{e_{3}}$ & 106\ $M_{l_{i}}$ & 1500 & $M_{q_{1,2}}$ & 1500 &$M_{l_{i}}$ & 180 & $M_{q_{i}}$ & 1500\ ------------------------------------------------------------------------ $M_{{q/u}_{3}}$ & 400 & $A_f$ & 650&$M_{{u}_{3}}$ & 450 & $A_f$ & -1850\ In S1, only the charginos and three neutralinos will be accessible at the LC. It should be possible to probe the supersymmetric QCD sector, with masses of $\sim$1.5 TeV, at the LHC. As input for the fit we therefore use: the masses of the charginos (${\tilde{\chi}}_1^{\pm},\,{\tilde{\chi}}_2^{\pm}$) and three lightest neutralinos (${\tilde{\chi}^0_{1}},\,{\tilde{\chi}^0_{2}},\,{\tilde{\chi}^0_{3}}$); the production cross section $\sigma({\tilde{\chi}}^+_1{\tilde{\chi}}^-_1)$ with polarised beams at $\sqrt{s} = 350$ and $500 {\ \mathrm{GeV}}$; the forward-backward asymmetry $A_{FB}$ at $\sqrt{s} = 350$ and $500 {\ \mathrm{GeV}}$; the branching ratio $\mathcal{B}(b\to s\gamma)$ calculated using [`micrOmegas 2.4.1`]{} [@Belanger:2006is; @*Belanger:2010gh], see tab. \[tab:inputsc1\]. Note that $\mathcal{B}(b\to s\gamma)$ increases sensitivity to the third generation squark sector, and the estimated experimental precision is $0.3\cdot 10^{-4}$ [@O'Leary:2010af]. As the impact of the muon anomalous magnetic moment is negligible it is not included in the fit. For S2, the input for the fit, as summarised in tab. \[tab:inputsc1\], is the same as in S1, however we consider a centre of mass energy $\sqrt{s}=400{\ \mathrm{GeV}}$ instead of $350{\ \mathrm{GeV}}$, and this is supplemented by the Higgs boson mass, $m_h$, with a theoretical uncertainty of $\sim$1 GeV, calculated using `FeynHiggs 2.9.1` [@Heinemeyer:1998yj; @*Degrassi:2002fi; @*Frank:2006yh]. The estimated experimental precision at the LC for $m_h$, taken from ref. [@AguilarSaavedra:2001rg], is adopted. The anomalous muon magnetic moment $\Delta (g_\mu -2)/2$, calculated using [`micrOmegas 2.4.1`]{} [@Belanger:2006is; @*Belanger:2010gh], plays a role and is therefore included in the fit, with a projected experimental error of $3.4\cdot10^{-10}$ [@Carey:2009zzb], which we assume would dominate over the theoretical uncertainty. We assume that the sneutrino mass has been measured. [@lr@[.]{}lr@[.]{}lr@[.]{}lr@[.]{}l@r@[.]{}lr@[.]{}lr@[.]{}lr@l@]{} &&\ \ &&\ ------------------------------------------------------------------------ Observable & & & & & & & &\ ${m_{\tilde{\chi}^\pm_{1}}}$ & $149$& $6$ & & $0$ & $1\ (0.2)$ & & $139$ & $3$ & & $0$ & $1$ & $-$\ ${m_{\tilde{\chi}^\pm_{2}}}$ & $292$ & $3$ & & $0$ & $5\ (2)$ & & $266$ & $2$ & & $0$ & $5$ & $-$\ ${m_{\tilde{\chi}^0_{1}}}$ & $106$ & $9$ & & $0$ & $2$ & & $92$ & $8$ & & $0$ & $2$ & $-$\ ${m_{\tilde{\chi}^0_{2}}}$ & $164$ & $0$ & $2$ & $0$ & $0$ & $5\ (1)$ & $0$ & $5$ & $148$ & $5$ & $\qquad2$ & $4$ & $0$ & $5$ & $0.5$\ ${m_{\tilde{\chi}^0_{3}}}$ & $188$ & $6$ & $-1$ & $5$ & $0$ & $5\ (1)$ & $0$ & $5$ & $189$ & $7$ & $-7$ & $3$ & $0$ & $5$ & $0.5$\ $\sigma({\tilde{\chi}}^+_1{\tilde{\chi}}^-_1)^{350/400}_{(-0.8,0.6)}$ & $2347$ & $5$ & $-291$ & $3$ & $8$ & $7$ & $2$ & $0$ & $709$ & $7$ & $-85$ & $1$ & $0$ & $7$ & $-$\ $\sigma({\tilde{\chi}}^+_1{\tilde{\chi}}^-_1)^{350/400}_{(0.8,-0.6)}$ & $224$ & $4 $ & $7$ & $6 $ & $2$ & $7$ & $0$ & $5$ & $129$ & $8 $ & $20$ & $0 $ & $0$ & $3$ & $-$\ $\sigma({\tilde{\chi}}^+_1{\tilde{\chi}}^-_1)^{500}_{(-0.8,0.6)}$ & $1450$ & $6 $ & $-24$ & $4$ & $6$ & $7$ & $2$ & $0$ & $560$ & $0 $ & $ -70$ & $1$ & $0$ & $7$ & $-$\ $\sigma({\tilde{\chi}}^+_1{\tilde{\chi}}^-_1)^{500}_{(0.8,-0.6)}$ & $154$ & $8 $ & $ 12$ & $7 $ & $2$ & $0$ & $0$ & $5$ & $97$ & $1 $ & $ 16$ & $4 $ & $0$ & $3$ & $-$\ $A_{FB}^{350/400}(\%)$ & $-2$&2& 6&8&0&8&0&1& 24&7&$-2$&8& 1&4 & $0.1$\ $A_{FB}^{500}(\%)$ &$-2$&6 &5&3&1&0&0&1 &39&2&$-5$&8 & 1&5 & $0.1$\ The results for S1, given in tab. \[tab:ressc1\], show the fit to the 8 MSSM parameters: $M_1$, $M_2$, $\mu$, $\tan\beta$, $m_{\tilde{\nu}}$, $\cos\theta_{\tilde{t}}$, $m_{\tilde{t}_1}$, and $m_{\tilde{t}_2}$. We find that the gaugino and higgsino mass parameters are determined with an accuracy better than 1%, while $\tan\beta$ is determined with an accuracy of $5\%$, and 2-3% for the sneutrino mass. The limited access to the stop sector (tab. \[tab:ressc1\]) could nevertheless lead to hints allowing a well-targeted search at the LHC. In tab. \[tab:ressc1\] we also compare the fit results obtained using masses of the charginos and neutralinos from threshold scans to those obtained using masses from the continuum. For the latter, the fit quality deteriorates, clearly indicating the need to measure these masses via threshold scans. For S2, the heavy Higgs boson mass is added to and the sneutrino mass is removed from the set of fit parameters. The results seen in tab. \[tab:ressc1\] show that the electroweakino parameters are precisely determined, and that the fit is sensitive to $m_{\tilde{t}_2}$, with an accuracy better than $20\%$. In addition, an upper limit on the mass of the heavy Higgs boson can be placed at 1000 GeV, at the 2$\sigma$ level. [lr@l@[ ]{}l r@l@[ ]{}l r@l@[ ]{}l ]{}&&\ \ &&\ ------------------------------------------------------------------------ Parameter & & &\ $M_1$ & $125 $ & $\pm 0.3 $ & $ (\pm 0.7) $ & $125 $ & $\pm 0.6 $ & $ (\pm 1.2)$ & $106$ &$\pm 0.3$ & $(\pm 0.5) $\ $M_2$ & $250 $ & $\pm 0.6 $ & $ (\pm 1.3) $ & $250 $ & $\pm 1.6 $ & $ (\pm 3)$ & $212$ &$\pm 0.5 $ & $ (\pm 1.0) $\ $\mu$ & $180 $ & $\pm 0.4 $ & $ (\pm 0.8) $ & $180 $ & $\pm 0.7 $ & $ (\pm 1.3)$ & $180$ &$\pm 0.4 $ & $ (\pm 0.9) $\ $\tan\beta$ & $10 $ & $ \pm 0.5 $ & $ (\pm 1) $ & $10 $ & $\pm 1.3 $ & $ (\pm 2.6) $ & $12$ &$\pm 0.3 $ & $ (\pm 0.7) $\ $m_{\tilde{\nu}}$ & $1500 $ & $\pm 24 $ & $ (^{+60}_{-40}) $ & $1500 $ & $\pm 20 $ & $ (\pm 40) $&\ $\cos\theta_{\tilde{t}}$ & $0.15$&$^{+0.08}_{-0.06} $ & $ (^{+0.16}_{-0.09}) $ & $0 $ & $\pm 0.15 $ & $ (^{+0.4}_{-0.3}) $ &\ $m_{\tilde{t}_1}$ & $400 $ & $^{+180}_{-120} $ & $ (^{\textrm{at limit}}_{\textrm{at limit}}) $ & & $430$&$^{+200}_{-130} $ & $ (^{+300}_{-400}) $\ $m_{\tilde{t}_2}$ & $800$ & $^{+300}_{-170} $ & $ (^{+1000}_{-290}) $ & $800$ & $^{+350}_{-220} $ & $ (^{\textrm{at limit}}_{\textrm{at limit}}$) & $1520$&$^{+200}_{-300} $ & $ (^{+300}_{-400}) $\ $m_{A^0}$ &&& & $ (<1000) $\ Conclusion {#sec:conc} ========== We have extended previous analyses to extract fundamental MSSM parameters via chargino production at the LC, by incorporating loop corrections. The polarised cross-sections and forward backward asymmetry for chargino production as well as the ${\tilde{\chi}}_1^{\pm},{\tilde{\chi}}_2^{\pm}$ and ${\tilde{\chi}^0_{1}}, {\tilde{\chi}^0_{2}}, {\tilde{\chi}^0_{3}}$ masses, are calculated at NLO in an on-shell scheme. When $M_1$, $M_2$ and $\mu$ are light, we determine them at $\mathcal{O}(\%)$, and $\tan\beta$ to $<5\%$. Masses obtained from the continuum as opposed to threshold scans result in doubled uncertainties on the fundamental parameters. For our 125 GeV Higgs mass compatible scenario we obtain an accuracy of better than $20\%$ on the mass of the heavy, left-handed stop. In summary, incorporating NLO corrections is shown to be required for the precise determination of the fundamental electroweakino parameters at the LC, and to provide sensitivity to the parameters describing particles contributing via loops. [^1]: Note that S2 corresponds to S3 in ref. [@Bharucha:2012ya].
{ "pile_set_name": "ArXiv" }
--- abstract: 'We consider a multivariate non-linear Hawkes process in a multi-class setup where particles are organised within two populations of possibly different sizes, such that one of the populations acts excitatory on the system while the other population acts inhibitory on the system. The goal of this note is to present a class of Hawkes Processes with stable dynamics without assumptions on the spectral radius of the associated weight function matrix. This illustrates how inhibition in a Hawkes system significantly affects the stability properties of the system.' address: - 'Mads Bonde Raad: Department of Mathematical Sciences, University of Copenhagen, Universitetsparken 5, 2100 Copenhagen, Danemark.' - 'E. Löcherbach: Universit[é]{} de Paris 1 Panthéon-Sorbonne, SAMM, EA 4543 et FR FP2M 2036 CNRS, 75013 Paris, France.' author: - Mads Bonde Raad - 'Eva L[ö]{}cherbach' title: Stability for Hawkes processes with inhibition --- [*Key words*]{} : Multivariate nonlinear Hawkes processes, Stability, Piecewise deterministic Markov processes, Lyapunov functions.\ [*MSC 2000*]{} : 60G55; 60G57; 60J25; 60Fxx Introduction and main result {#sec:1} ============================ We consider a system of interacting Hawkes processes structured within two populations. We shall label the two populations with “$+$” or “$-$” signaling that the population acts excitatory or inhibitory on the system, respectively. Let $N_+,N_{-} \in {\mathbb {N}}$ be the number of units in each population. Introduce weight functions given by $$\begin{aligned} \label{eq:weightfunctions} h_{++} ( t)& =&\frac{c_{++}}{N_{+}} e^{ -\nu_{+}t}, \quad h_{+-}( t)=\frac{c_{+-}}{N_{+}}e^{ -\nu_{+}t},\\ h_{-+}( t)&=&\frac{c_{-+}}{N_{-}} e^{ -\nu_{-}t},\quad h_{--}( t)=\frac{c_{--}}{N_{-}} e^{ -\nu_{-}t}, \end{aligned}$$ for $ t \geq 0.$ In the above formula, $h_{+-}$ indicates the weight function from a unit in the excitatory group “$+$” to a unit in the inhibitory group “$-$”, and so on. The coefficients of the system of interacting Hawkes processes are the exponential leakage terms $ \nu_+ > 0 , \nu_- > 0 $ and the weights $ c_{++} , c_{+-} , c_{-+}, c_{--} $ satisfying that $$c_{++} \geq 0, \; c_{+-} \geq 0, \; c_{--} \leq 0, \; c_{-+} \leq 0.$$ The multivariate linear Hawkes process with these parameters is given as $$\begin{aligned} \label{eq:dyn} Z_{+}^i (t) &= & \int_0^t \int_0^\infty \indiq_{ \{ z \leq \psi^{i}_+ ( X_{+} ({s-})) \}} \pi_{+}^i (ds, dz) , 1 \le i\leq N_{+}, \\ Z_{-}^j (t) &= & \int_0^t \int_0^\infty \indiq_{ \{ z \leq \psi^{j}_- ( X_{-} ({s-})) \}} \pi_{-}^j (ds, dz) ,1 \le j\leq N_{-}, \\ X_+ (t) &= & e^{- \nu_+ t } X_+ (0) + \frac{c_{++}}{N_+} \sum_{i=1}^{N_{+}}\int_0^t e^{- \nu_+ ( t- s) } Z_{+}^i (d s) + \frac{c_{-+}}{N_-} \sum_{j=1}^{N_{-}}\int_0^t e^{- \nu_+ ( t- s) } Z_{-}^j (ds) ,\\ X_{-} (t) &= & e^{- \nu_- t } X_{-} (0 )+ \frac{c_{+-}}{N_+} \sum_{i=1}^{N_{+}}\int_0^t e^{- \nu_- ( t- s) } Z_{+}^i (ds) +\frac{ c_{--}}{N_-} \sum_{j=1}^{N_{-}}\int_0^t e^{- \nu_- ( t- s) } Z_{-}^j (ds) ,\end{aligned}$$ where the jump rate functions $\psi^i_+ : {\mathbb {R}}\to {\mathbb {R}}_+ , \psi^i_- : {\mathbb {R}}\to {\mathbb {R}}_+ $ are given by $$\label{eq:aplus} \psi^i_{\pm} (x) = a^i_{\pm} + \max (x, 0), \; \mbox{ where } a^i_{\pm } > 0 ,$$ and where the $ \pi^i_{\pm} , i \geq 1, $ are i.i.d. Poisson random measures on $ {\mathbb {R}}_+ \times {\mathbb {R}}_+ $ having intensity $ dt dz.$ Notice that the process $ ( X_+, X_{-} ) $ is a piecewise deterministic Markov process having generator $$\begin{gathered} A g (x, y ) = - \nu_+ x \partial_x g (x,y ) - \nu_- y \partial_y g (x, y ) + \sum_{i=1}^{N_+} \psi^i_+ ( x) [ g ( x + \frac{c_{++}}{N_+} , y+ \frac{c_{+-}}{N_+} )- g(x,y ) ] \\ +\sum_{j=1}^{N_-} \psi_-^j (y ) [ g ( x + \frac{c_{-+}}{N_-} , y + \frac{c_{--}}{N_-} ) - g(x, y ) ] ,\end{gathered}$$ for sufficiently smooth test functions $g.$ Classical stability results for multivariate nonlinear Hawkes processes found e.g. in [@bm] or in the recent paper [@manonetal], which is devoted to the study of the stabilising effect of inhibitions, are stated in terms of an associated weight function matrix $\Lambda,$ imposing that the spectral radius of $\Lambda$ is strictly smaller than one. In this case the process is termed to be [*subcritical*]{}. This spectral radius stability condition has a natural interpretation in terms of a multitype branching process with immigration which is spatially structured and where each jump of a given type ($+ $ or $-$) gives rise to future jumps of the same or of the opposite type, see [@ho]. The subcriticality condition ensures the recurrence of this process (see [@kaplan]). In our system, the weight function matrix is given by $$\label{eq:Lambda} \Lambda = \left( \begin{array}{cc} \frac{c_{++}}{\nu_+} & \frac{|c_{-+}|}{\nu_+} \\ \frac{c_{+-}}{\nu_-} & \frac{|c_{--}|}{\nu_-} \end{array} \right) .$$ Notice that in , negative synaptic weights do only appear through their absolute values. This is due to the fact that using the Lipschitz continuity of the rate functions leads automatically to considering absolute values and does not enable us to make profit from the inhibitory action of $c_{-+} $ and $ c_{--}. $ Obviously, having sufficiently fast decay, that is, $ \min (\nu_+ , \nu_-) >> 1, $ is a sufficient condition fo subcriticality. The purpose of this note is to show how the presence of sufficiently high (in absolute value) negative weights helps stabilising the process without imposing such a subcriticality condition, in particular, without imposing $ \nu_+, \nu_- $ being large. To the best of our knowledge, only few results have been obtained on this natural question in the literature. [@bm] gives an attempt in this direction but does only deal with the case when $ c_{+- } $ and $ c_{-+}$ are of the same sign (see Theorem 6 in [@bm]), and [@manonetal] do only work with the positive part of the weight functions, without profiting from the explicit inhibitory part within the system. Our approach is based on the construction of a convenient Lyapunov function using the inhibitory part of the dynamics. As such, this approach is limited to the present Markovian framework where the weight functions are decreasing exponentials. In the following, we shall write $$c_{++}^* := c_{++} - \nu_+ , \; c_{-- }^* := c_{--} - \nu_{-} .$$ Notice that $ c_{++}^* $ could be interpreted as the net increase of $ X_+ $ due to self-interactions of $X_+ $ with itself. $ c_{--}^* $ is always negative. \[aslol\] We assume the following inequalities. $$\begin{aligned} \label{eq:stab} c_{++}^* +c_{--}^* &<& 0 ,\\ ( c_{++}^*- c_{--}^* )^2 &< &4 c_{+- } | c_{-+}| ,\\ c_{++}^* - c_{--}^* &>& 0.\end{aligned}$$ This assumption ensures that the system is balanced. Notice that Assumption \[aslol\] does not imply - nor is implied by - that the spectral radius of $\Lambda$ is strictly smaller than $1$. For example, if Assumption \[aslol\] is satisfied for some parameters $ ( c_{++},c_{+-},c_{-+},c_{--},\nu,\nu ) , $ i.e., $\nu_+ = \nu_- = \nu, $ such that additionally $ c_{++} + c_{--} < 0, $ then for all $C>1 $ and all $\varepsilon>0,$ the set of parameters $( Cc_{++},Cc_{+-},Cc_{-+},Cc_{--},\varepsilon\nu,\varepsilon\nu ) $ satisfies Assumption \[aslol\] as well. But the associated offspring matrix $\Lambda_{C,\varepsilon}$ of the scaled parameters is equal to $(C/\varepsilon) \Lambda , $ and thus the spectral radius is also scaled by $C/\varepsilon$. \[ass:2d\] We assume that either $ \nu_+ \neq \nu_- $ or $ \nu_+ = \nu_- $ and $( c_{++},c_{+-} ) ,( c_{-+},c_{--}) $ are linearly independent. We are now able to state our main result. It states that under Assumptions \[aslol\] and \[ass:2d\], the process $ X = (X_+, X_{-} ) $ is positive Harris recurrent, together with a strong mixing result. To state our result, for any $ t > 0 $ and for $ z = (x,y ) \in {\mathbb {R}}^2 ,$ we write $ P_t ( z, \cdot )$ for the transition semigroup of the process, defined through $P_t (z, A) = E_z ( 1_A (X (t)) ) .$ Moreover, for any pair of probability measures $\mu_1, \mu_2 $ on $ {\mathcal B} ({\mathbb {R}}^2)$ and for any function $ V : {\mathbb {R}}^2 \to [1, \infty [, $ we put $$\| \mu_1- \mu_2 \|_{ V} := \sup_{ g : |g| \le V } | \mu_1 ( g) - \mu_2 (g) | .$$ \[theo:harris\] Grant Assumptions \[aslol\] and \[ass:2d\].\ 1) Then the process $ X = (X_+, X_- ) $ is positive recurrent in the sense of Harris, and its unique invariant probability measure $ \mu $ possesses a Lebesgue continuous part.\ 2) There exists a function $V (x, y ) : {\mathbb {R}}^2 \to [1, \infty [ $ such that $\lim_{ |x| + |y| \to \infty } V ( x,y ) = \infty $ and there exist $ c_1, c_2 > 0 $ such that for all $z \in {\mathbb {R}}^2$ and all $ t \geq 0, $ $$\label{eq:last} \| P_t(z , \cdot ) - \mu\|_{ V} \le c_1 V (z) e^{ - c_2 t} .$$ Notice that if Assumption \[ass:2d\] is not satisfied, that is, if $ \nu_+ = \nu_- $ and if $$\left( \begin{array}{c} c_{-+}\\ c_{--} \end{array}\right) \in H:= {\mathbb {R}}\left( \begin{array}{c} c_{++}\\ c_{+-} \end{array}\right),$$ then it is easily shown that almost surely, $ dist ( X (t) , H) \to 0 $ as $t \to \infty $ and that $H$ is invariant under the dynamics. Moreover, the restriction of the dynamics to $H$ is Harris recurrent, having a unique invariant measure $ \mu $ which is absolutely continuous with respect to the Lebesgue measure on $H.$ However, it is easy to show that the original process $X,$ defined on $ {\mathbb {R}}^2, $ is not Harris in this case, since it is not $ \mu-$irreducible. Proof of Theorem \[theo:harris\] ================================ This section is devoted to the proof of Theorem \[theo:harris\]. A Lyapunov function for $X$ --------------------------- We start this section with the following useful property. \[prop:Feller\] The process $ X$ is a Feller process, that is, for any $f : {\mathbb {R}}^2 \to {\mathbb {R}}$ which is bounded and continuous, we have that ${\mathbb {R}}^2 \ni (x, y ) = z \mapsto E_{z} f (X (t) ) = P_t f (z) $ is continuous. The proof of this result follows from classical arguments, see e.g. the proof of Proposition 4.8 in [@evaflow], or [@ikeda1966]. The next result shows that if the cross-interactions, that is, influence from $ X_+ $ to $X_{-} $ and vice versa, are sufficiently strong, then – under mild additional assumptions – it is possible to construct a Lyapunov function for the system that does mainly profit from the inhibitory part of the jumps. \[prop:lyapunov\] Grant Assumption \[aslol\] and put $$V ( x, y ) := \left\{ \begin{array}{ll} V_{++} ( x, y ) : = c_{+- } x^2 -c_{-+}y^2 - (c_{++}^* - c_{--}^*) xy & x \in {\mathbb {R}}_+ , y\in {\mathbb {R}}_+ \\ V_{+-} (x,y ) := c_{+- } x^2 + q y^2 - (c_{++}^* - c_{--}^*) xy & x\in {\mathbb {R}}_+ , y \in {\mathbb {R}}_- \\ V_{-+} (x,y ) := px^2 -c_{-+}y^2 - (c_{++}^* - c_{--}^*) xy & x \in {\mathbb {R}}_- , y\in {\mathbb {R}}_+ \\ V_{--} (x,y ) := p x^2 + qy^2 - (c_{++}^* - c_{--}^*) xy & x \in {\mathbb {R}}_- , y\in {\mathbb {R}}_- \end{array} \right\} ,$$ with $p$ so small such that $$- (c_{++}^* - c_{--}^* ) (c_{--} - \nu_+ - \nu_- ) + 2 p c_{-+} > 0$$ and $q$ so large such that $$(c_{++}^* - c_{--}^* ) [ \nu_+ + \nu_- - c_{++} ] + 2 q c_{+- } > 0 \mbox{ and } 4 pq > (c_{++}^* - c_{--}^*)^2 .$$ Then $\lim_{ |x| + |y| \to \infty } V ( x,y ) = \infty $ and there exist $ \kappa, c, K > 0 $ such that $$A V (x,y ) \le - \kappa V ( x, y ) + c 1_{\{ | x| + |y| \geq K\}} .$$ We calculate $ A V ( x, y ) = A^1 V (x, y ) + A^2 V (x, y ) , $ with $$A^1 V ( x, y ) = - \nu_+ \partial_x V (x,y ) - \nu_- \partial_y V (x, y )$$ and $ A^2 $ the jump part of the generator. [**Part 1.1**]{} Suppose first that $ x \geq |c_{-+}|/ N_- , y \geq | c_{--} |/N_- . $ Then $$A V (x, y ) = A^1 V_{++} (x, y) + A^2 V_{++} (x,y ) = a_{++} x^2 + b_{++} xy + d_{++} y^2 + L_{++} (x,y ) ,$$ where $L_{++} $ is a polynomial of degree $1.$ A straightforward calculus shows that $$\begin{aligned} a_{++} &=& c_{+-} (c_{++}^* + c_{--}^* ) , \\ b_{++} &=& - (c_{++}^* - c_{--}^* ) (c_{++}^* + c_{--}^* )\\ d_{++} & =& - c_{-+} (c_{++}^* + c_{--}^* ) ,\end{aligned}$$ proving that $$A V( x, y ) = (c_{++}^* + c_{--}^* ) V (x,y) + L_{++} (x,y ) .$$ This implies that there exist $K, \kappa > 0 $ such that $$A V ( x, y ) \le - \kappa V ( x,y )$$ for all $ x > K , y > K,$ since $ c_{++}^* + c_{--}^* < 0 $ by assumption. [**Part 1.2**]{} Suppose now that $0 \le x < |c_{-+} |/N_- $ and $y \geq | c_{--} |/N_- .$ Then a jump of one of the inhibitory neurons will lead to a change $ x \mapsto x + c_{-+}/N_- < 0 .$ In this case we obtain $$A V (x,y) = A V_{++} ( x,y ) + \sum_{j=1}^{N_-}(a^{j}_- + y ) ( V_{-+} ( x + \frac{c_{-+}}{N_-}, y + \frac{c_{--}}{N_-} ) - V_{++} ( x + \frac{c_{-+}}{N_-}, y + \frac{c_{-+}}{N_-} )) .$$ But $$|V_{-+} ( x + \frac{c_{-+}}{N_-}, y + \frac{c_{--}}{N_-} ) - V_{++} ( x + \frac{c_{-+}}{N_-}, y + \frac{c_{-+}}{N_-} )| \le C ,$$ since $ | x| < |c_{-+} |,$ and therefore $$A V (x,y) \le A V_{++} ( x,y ) + L (y) ,$$ where $ L(y) $ is a monomial in $ y.$ The other case $0 \le y < |c_{--} |/N_- $ and $x \geq | c_{-+} |/N_- $ is treated analogously. [**Part 2.1**]{} Suppose now that $ x \geq |c_{-+}|/N_- , y \leq - c_{+-} /N_+ . $ Then $$A V (x, y ) = A^1 V_{+-} (x, y) + A^2 V_{+-} (x,y ) = a_{+-} x^2 + b_{+-} xy + d_{+-} y^2 + L_{+-} (x,y ) ,$$ where $L_{+-} $ is a polynomial of degree $1.$ We obtain $$\begin{aligned} a_{+-} &=& c_{+-} (c_{++}^* + c_{--}^* ) , \\ b_{+-} &=& (c_{++}^* - c_{--}^* ) (\nu_+ + \nu_- - c_{++} ) + 2 q c_{+- } \\ d_{+- } & =& - 2 \nu_- q .\end{aligned}$$ Since $ b_{+-} > 0 $ by choice of $q,$ this implies that for a suitable positive constant $ \kappa > 0 ,$ $$A V (x,y) \le - \kappa V(x,y) + L_{+-} (x,y) ,$$ which allows to conclude as before. [**Part 2.2**]{} The cases $ x \geq |c_{-+}|/N_- , 0 \geq y > - c_{+-}/N_+ $ or $ 0 \le x < |c_{-+}|/N_- , y \leq - c_{+-}/N_+ $ are treated analogously to Part 1.2. [**Part 3**]{} Suppose now that $ x \le - c_{++} /N_+ , y \geq - c_{- -} /N_- . $ Then $$A V (x, y ) = A^1 V_{-+} (x, y) + A^2 V_{-+} (x,y ) = a_{-+} x^2 + b_{-+} xy + d_{-+} y^2 + L_{-+} (x,y ) ,$$ where $L_{-+} $ is a polynomial of degree $1$ and where $$\begin{aligned} a_{-+} &=& -2 \nu_+ p , \\ b_{-+} &=& (c_{++}^* - c_{--}^* ) ( \nu_+ + \nu_- - c_{--} ) + 2 p c_{-+} \\ d_{-+} & =& - c_{-+} (c_{++}^* + c_{--}^* ) .\end{aligned}$$ Notice that by choice of $p,$ $ b_{-+} > 0 .$ The conclusion of this part follows analogously to the previous parts 1.1 and 2.1. [**Part 4**]{} Suppose finally that $ x \le - c_{++}/N_+ , y \le - c_{+ -} /N_+ . $ Then $$A V (x, y ) = A^1 V_{--} (x, y) + A^2 V_{--} (x,y ) = a_{--} x^2 + b_{--} xy + d_{--} y^2 + L_{--} (x,y ) ,$$ where $L_{--} $ is a polynomial of degree $1$ and where $$\begin{aligned} a_{--} &=& -2 \nu_+ p , \\ b_{--} &=& (c_{++}^* - c_{--}^* ) (\nu_+ + \nu_- ) \\ d_{--} & =& - 2 \nu_- q ,\end{aligned}$$ leading to the same conclusion as in the previous parts. As a consequence of Proposition \[prop:lyapunov\], the process $X$ is stable in the sense that it necessarily possesses invariant probability measures, maybe several of them. The uniqueness of the invariant probability measure together with the Harris recurrence will follow from the following local Doeblin type lower bound. \[thm:Doeblin\] For all $ T> 0 $ and for all $z_* = (x_*, y_*) \in {\mathbb {R}}^2 $ the following holds. There exist $R > 0 , $ an open set $ I \subset {\mathbb {R}}^2 $ with strictly positive Lebesgue measure and a constant $\beta \in (0, 1), $ depending on $I , R$ and the coefficients of the system with $$\label{doblinminorization} P_{T} (z , dz' ) \geq \beta 1_C (z) \nu ( dz') ,$$ where $ C = B_R ( z_* ) $ is the (open) ball of radius $R$ centred at $z_* ,$ and where $ \nu $ is the uniform probability measure on $ I.$ We start with the case $ \nu_+ \neq \nu_- ,$ under the assumption that $ c_{++}, c_{--}, c_{+-}, c_{-+} \neq 0 . $ In this case, [@Clinet] in the proof of their Lemma 6.4 establish the lower bound for the four-dimensional Markov process $ \bar X = (X_{++}, X_{+-}, X_{-+}, X_{--} ) $ given by $$X_{++} (t) = e^{- \nu_+ t } X_{++} (0) + \frac{c_{++}}{N_+} \sum_{i=1}^{N_{+}}\int_0^t e^{- \nu_+ ( t- s) } Z_{+}^i (d s) ,$$ $$X_{-+} (t) = e^{- \nu_+ t } X_{-+} (0)+ \frac{c_{-+}}{N_-} \sum_{j=1}^{N_{-}}\int_0^t e^{- \nu_+ ( t- s) } Z_{-}^j (ds),$$ $$X_{+-} (t) = e^{- \nu_- t } X_{+-} (0) + \frac{c_{+-}}{N_+} \sum_{i=1}^{N_{+}}\int_0^t e^{- \nu_- ( t- s) } Z_{+}^i (ds) ,$$ $$X_{--} (t) = e^{- \nu_- t } X_{--} (0)+ \frac{ c_{--}}{N_-} \sum_{j=1}^{N_{-}}\int_0^t e^{- \nu_- ( t- s) } Z_{-}^j (ds) ,$$ where $ X_{++} (0) + X_{-+} (0) = X_+ (0) , X_{--} (0) + X_{+-} (0) = X_- (0) .$ More precisely, they show that for any $ \bar z_* \in {\mathbb {R}}^4 ,$ there exist $\bar R > 0 , $ an open rectangle $ \bar I \subset {\mathbb {R}}^4 $ with strictly positive Lebesgue measure and a constant $\bar \beta \in (0, 1), $ such that $$\bar P_{T} (\bar z , d\bar z' ) \geq \bar \beta 1_{\bar C} (\bar z) \bar \nu ( d \bar z') ,$$ where $ \bar C = B_R ( \bar z_* ) $ is the (open) ball of radius $\bar R$ centred at $\bar z_* ,$ and where $ \bar \nu $ is the uniform probability measure on $ \bar I.$ The above formula can be interpreted in the following way: For any $ \bar z \in \bar C, $ with probability $\bar \beta, $ the law of $ \bar X (T) $ is equal to the law of $ U = (U_1, U_2, U_3, U_4) $ where $ U $ is a uniform random vector on $ \bar I .$ Since $ \bar I $ is supposed to be a rectangle, this implies in particular the independence of its coordinates $ U_1, \ldots , U_4.$ Notice that we have $ X (T) = A \bar X (T) ,$ where $$A = \left( \begin{array}{cccc} 1&0&1&0\\ 0&1&0&1 \end{array} \right) .$$ We now show how the above result implies the local lower bound for the original process $X.$ For that sake let $ z_* \in {\mathbb {R}}^2 $ be arbitrary and fix any $ \bar z_* \in {\mathbb {R}}^4 $ such that $A \bar z_* = z_* .$ Let $\bar R$ be the associated radius and choose $R$ such that $ B_R ( z_* ) \subset A B_{\bar R} ( \bar z_*) .$ Then for all $ z \in B_R (z_* ) $ and $ \bar z \in B_{\bar R} ( z_* ) $ with $ A \bar z = z, $ $$P_z ( X (T) \in \cdot ) = P_{\bar z} (A \bar X (T) \in \cdot ) \geq \bar \beta \P ( A U \in \cdot ) .$$ Since $$A U = \left( \begin{array}{c} U_1 + U_3 \\ U_2 + U_4 \end{array} \right) ,$$ by independence of the coordinates $ U_1, \ldots , U_4,$ this implies the desired result for the two-dimensional Markov process $X $ as well. We finally deal with the case $ \nu_+ = \nu_- $ and $( c_{++},c_{+-} ) ,( c_{-+},c_{--}) $ linearly independent. Fix $ z_* = (x_*, y_* ) $ and $ M> |x_*|+ |y_*| $ arbitrarily and let $ H := \{ z = (x, y ) : |x| \le M, |y|\le M \} .$ Recall and introduce finally the event $E$ given by - $\pi^1_{+} ( [ 0,T] \times [ 0,a^1_{+}]) =1,$ - $\pi_{+}^1 ( [ 0,T] \times ]a^1_{+}, a^1_{+}+c_{++} + M ) =0,$ - $\pi_{+}^i ( [ 0,T] \times [ 0, a^i_{+}+c_{++} + M ) =0$ for all $ 2 \le i \le N_+,$ - $\pi^1_{-} ( [ 0,T] \times [ 0,a^1_{-}]) =1,$ - $\pi_{-}^1 ( [ 0,T] \times ] a^1_{-}, a^1_{-}+c_{+-} + M ) =0,$ - $\pi_{-}^j ( [ 0,T] \times [ 0, a^j_{-}+c_{+-} + M ) =0$ for all $ 2 \le j \le N_- .$ Define the substochastic kernel $$Q^T_{z} ( A) =P_z( E\cap \{ X ({T}) \in A\} )=P( E) P_z( X ({T}) \in A | E) .$$ The conditional law of $ X ({T}) $ given $ E,$ under $P_z,$ is equal to the law of $$Y_z ({T}) =ze^{-\nu_+ T}+e^{-\nu_+ U_{+}}\left(\begin{array}{c}c_{++}/N_+ \\c_{+-}/N_+ \end{array}\right)+ e^{-\nu_+ U_{-}}\left( \begin{array}{c}c_{-+}/N_-\\c_{--}/N_- .\end{array}\right) ,$$ where the two jump-times $U_{+},U_{-}$ are independent uniform variables on $[ 0,T] .$ Since $$C=\left( \begin{array}{cc} c_{++}/N_+ & c_{-+}/N_-\\c_{+-}/N_+& c_{--}/N_- \end{array} \right)$$ is invertible and the law of $( e^{-\nu_+ U_{+}},e^{-\nu_+ U_{-}})$ is equivalent with the Lebesgue measure on $[ e^{-\nu_+ T},1] ^2, $ the law of $ Y_z ({T})$ has density $$f_{z}:v\mapsto | det\; C|^{-1} f\circ C^{-1}( v-ze^{-\nu_+ T}),$$ where $f$ is the density of $( e^{-\nu_+ U_{+}}, e^{-\nu_+ U_{-}})$. The density is positive on the interior of its support $$supp( Y_z ({T}))= e^{-\nu_+ T}z +C [ e^{-\nu_+ T},1]^2 .$$ Since $C$ is a homeomorphism, it is an open mapping. Thus we can find balls $B_r ( v_{0})\subset B_{2r} ( v_{0}) \subset C [ e^{- \nu_+ T},1]^2 $ for all $T>1.$ Take now $T$ so large that $e^{-\nu_+ T}\sup_{v \in H} \| v \| <r.$ For such $T$ and all $ z \in H$ we have $$\overline{B}_r ( v_{0}) \subset e^{-\nu_+ T} z + B_{2r} ( v_{0})\subset supp( Y_z ({T})) .$$ Note now that $H\times \overline{B}_r ( v_{0}) \ni (z,v) \mapsto f_{z}( v) $ is continuous, so the positivity of the density gives $\inf_{z \in H,v\in \overline{B}_r ( v_{0}) } f_{z}( v):=\alpha>0.$ We therefore conclude that $$Q^T_{z}( A)\geq P( E) \cdot \alpha \cdot \lambda ( A\cap B_r ( v_{0}) ),$$ for all $z \in H,$ where $ \lambda $ denotes the Lebesgue measure on ${\mathbb {R}}^2.$ This proves the desired result. We do now dispose of all ingredients to conclude the proof of Theorem \[theo:harris\]. 1\) We apply Proposition \[thm:Doeblin\] with $z_* = 0 .$ Let $R$ be the associated radius. By Proposition \[prop:lyapunov\], we know that for a suitable compact set $K \subset {\mathbb {R}}^2, $ $X $ comes back to $K $ infinitely often almost surely. For $ z = ( x, y ), $ write $$\label{eq:flowy} \varphi_t (z) = ( \varphi^{(1)}_t ( x) ,\varphi^{(2)}_t (y) ) = (e^{ - \nu_+ t}x , e^{- \nu_- t} y)$$ for the flow of the process in between successiv jumps and let $ \| z\|_1 := |x| + |y|.$ Then $$\sup_{z \in K, t \geq 0} \| \varphi_t (z) \|_1 := F < \infty \; \; \mbox{ and } \; \; \sup_{z \in K} \| \varphi_t (z) \|_1 \to 0$$ as $t \to \infty .$ Therefore there exists $t_* $ such that $\varphi_t (z) \in B_{R } ( 0) $ for all $t \geq t_* , $ for all $ z \in K .$ Hence, $$\inf_{z\in K} P_z ( X ({t_* + s }) \in B_R ( 0 ), 0 \le s \le 2T ) > 0 .$$ Consequently, the Markov chain $(X ({kT}))_{k \in {\mathbb {N}}} $ visits $ B_{R } ( 0 )$ infinitely often almost surely.\ The standard regeneration technique (see e.g. [@dashaeva]) allows to conclude that $(X ({kT}))_{k \in {\mathbb {N}}} $ and therefore $(X(t))_t $ are Harris recurrent. This concludes the proof of the Harris recurrence of the process. 2\) The sampled chain $ (X ({kT }))_{k \geq 0 }$ is Feller according to Proposition \[prop:Feller\]. Moreover it is $ \nu-$irreducible, where $ \nu $ is the measure introduced in Proposition \[thm:Doeblin\], associated with the point $z_* = (0,0) .$ Since $\nu$ is the uniform measure on some open set of strictly positive Lebesgue measure, the support of $\nu $ has non-empty interior. Theorem 3.4 of [@MT1992] implies that all compact sets are ‘petite’ sets of the sampled chain. The Lyapunov condition established in Proposition \[prop:lyapunov\] allows to apply Theorem 6.1 of [@MT1993] which implies the second assertion of the theorem. [99]{} Stability of nonlinear Hawkes processes. , 24(3) (1996) 1563-1588. Statistical inference for ergodic point processes and application to Limit Order Book. , 127 (2017), 1800-1839. Renewal in Hawkes processes with self-excitation and inhibition. , 2018. Hawkes processes on large networks. 26 (2016), 216–261. Multi-class oscillating systems of interacting neurons. 127 (2017), 1840–1869. A cluster process representation of a self-exciting process. 11 (1974), 493–503. R. H[ö]{}pfner and E. L[ö]{}cherbach. Statistical models for [B]{}irth and [D]{}eath on a [F]{}low: Local absolute continuity and likelihood ratio processes. , 26(1):107–128, 1999. N. Ikeda, M. Nagasawa, and S. Watanabe. A construction of [M]{}arkov processes by piecing out. , 42(4):370–375, 1966. E. Löcherbach and D. Loukianova. On [N]{}ummelin splitting for continuous time [H]{}arris recurrent [M]{}arkov processes and application to kernel estimation for multi-dimensional diffusions. , 118:1301–1321, 2008. The Multitype Galton-Watson Process with Immigration. , 6:947–953, 1973. S.P. Meyn and R.L. Tweedie. Stability of [Markovian processes I : Criteria for discrete-time chains.]{} , 24:542–574, 1992. S.P. Meyn and R.L. Tweedie. Stability of [Markovian processes III : Foster-Lyapunov]{} criteria for continuous-time processes. , 25:487–548, 1993.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Transport process in molecular chain in nonequilibrium stationary state is theoretically investigated. The molecule is interacting at its both ends with thermal baths which has different temperatures, while no dissipation mechanism is contained inside the molecular chain. We have first obtained the nonequilibrium stationary state outside the Hilbert space in terms of the complex spectral representation of Liouvillian. The nonequilibrium stationary state is obtained as an eigenstate of the Liouvillian which is constructed through the collision invariant of the kinetic equation. The eigenstate of the Liouvillian contains an information of spatial correlation between the molecular chain and the thermal baths. While energy flow in the nonequilibrium state which is due to the first order correlation can be described by Landauer formula, the particle current due to the second order correlation cannot be described by the Landauer formula. The present method provides a simple and perspective way to evaluate the energy transport of molecular chain under the nonequilibrium situation.' author: - Satoshi - Kazuki - Tomio Petrosky title: Nonequilibrium Transport of Quantum Molecular Chain in terms of the Complex Liouvillian Spectrum --- Introduction ============ Recently there have been paid much attention to nonequilibrium transport processes through molecular wire junctions.[@JortnerBook; @Cahill03] As the system is shrunk down to a smaller size than the mean free path of a carrier, such as an electron, a phonon, or an exciton, the ballistic quantum transport exhibits a characteristically different feature from the corresponding bulk properties. For example, it was discovered that the electronic or thermal conductance of sub-micron size one-dimensional chain is quantized at low temperature.[@Rego98; @Schwab00; @Wees88; @Wharam88] When the system size is further shrunk down to the nanometer size, a discretized energy level structure of a nanowire plays an important role in transport processes through the resonance effect. Recent ultrafast nonlinear optical spectroscopy has revealed the temporal behavior of relaxation process of the photo-excited molecular chain.[@Schwarzer04; @Wang07] Also in biomolecules, such as alpha-helix protein and DNA strands, which consists of molecular chain structures, the transport process in nonequilibrium state have been extensively studied in order to clarify the biofunctions of these molecules from a microscopic view point.[@Berlin02; @Botan07; @Nguyen10; @Kawai10; @Tanaka09] In order to understand the transport processes in the nonequilibrium stationary state, we need to know first of all how to describe the nonequilibrium stationary state with which a physical quantity can be obtained as an expectation value of an observable. While it is obvious that a stationary state is described by a canonical distribution in a thermal equilibrium, it is not so easy to find the explicit form of the stationary state for nonequilibrium situation. Meanwhile a phenomenological Landauer formula has been extensively used as a useful tool to evaluate a transport coefficient in a nonequilibrium stationary state without explicitly describing the stationary state. There have been enormous works trying to make clear its microscopic foundation and its applicability in terms of nonequilibrium Green’s function method, quantum Langevin method, and so on. [@JortnerBook; @Landauer57; @Rieder67; @Zurcher90; @Segal01; @Segal03; @Petrov05; @Galperin07; @Subotnik09] On the other hand, Prigogine and his coworkers have developed a theory of the nonequilibrium statistical mechanics in terms of the complex spectral representation of the Liouville-von Neumann super-operator, or simply called [*Liouvillian*]{}.[@PrigogineText; @Petrosky99] They have clarified that the eigenvalue problem of the Liouvillian is classified into independent subspaces according to the order of correlations based on the concept of the [*dynamics of the correlation*]{} where the fundamental object is a correlation. In a classical gas system, for example, each correlation is characterized by the dependence of the distribution function on the wave number which characterizes the spatial correlation: The inhomogeneity is a correlation component which has a single non-vanishing wave number of a particle, while [*inter-particle correlation*]{} is a component which has several non-vanishing wave numbers of different particles.[@Petrosky97; @Petrosky02; @Petrosky10] This classification in terms of the correlation enables to transform the eigenvalue problem of the Liouvillian to the eigenvalue problem of the collision operator of the kinetic equation in a correlation subspace. It is well known that the collision operator plays a central role in nonequilibrium statistical mechanics, as seen in Boltzmann equation or Fokker-Planck equation.[@PrigogineText; @ResiboisBook] A striking finding of the theory is that the spectrum of the collision operator is identical with that of the Liouvillian, which signifies the direct link between the microscopic dynamics governed by the Liouvillian and the phenomenological kinetic theory.[@Petrosky96; @Petrosky10; @Petrosky97] Since the collision operator is non-Hermite operator, the eigenvalues can take complex values which reflects the dissipation of the system. The eigenstate of the Liouvillian is obtained from the eigenstate of the collision operator by operating a creation-of-correlation operator onto the eigenstate of the collision operator which represents the transition from a privileged correlation subspace to other non-privileged correlation subspaces.[@Petrosky96; @Petrosky97] The eigenstate of the Liouvillian is then represented as a series of the correlation generated by the dynamics of correlation. There have been a few applications of the theory to real physical systems,[@Petrosky99; @Petrosky02; @Tay06] but the transport processes in nonequilibrium situations have not been fully investigated yet in the context of dynamics of correlation. Our aim here is to apply the theory to the real mesoscopic system under nonequilibrium situation and to systematically derive the transport quantities in a stationary state. In the present work, we consider a molecular chain coupled with different thermal baths at the both ends, where a quantum particle is confined in the molecule and it ballistically transfers within the molecule. An exchange of energy occurs between the molecule and the thermal baths, while there is no particle exchange between them as shown in Fig.\[Fig1\]. The nonequilibrium stationary state is obtained as the zero eigenstate of the Liouvillian which is represented by superposition of different order of correlations. A leading term is the non-correlation component, or called [*vacuum of correlation*]{}, followed by the higher order correlation components. As will be shown, the inhomogeneity component of a single particle is created by the second order interaction from the vacuum of correlation. This is contrast to the case of a well-known gas system where the inhomogeneity component is disconnected from the vacuum of correlation subspace by the interaction, thus is not created from the vacuum of correlation.[@PrigogineText] Once the nonequilibrium stationary state is obtained as a zero eigenstate of the Liouvillian, a physical quantity is evaluated as an expectation value of a corresponding observable in the stationary state. Since the nonequilibrium stationary state is an eigenstate of the Liouvillian, the expression of a physical quantity is justified from the microscopic dynamics without any phenomenological assumptions. We have derived an expression of the energy flow with use of the nonequilibrium stationary state, and found that the first order correlation is responsible for the energy flow. We shall show that the expression of the energy flow is cast into the Landauer formula with a characteristic transmission function, because the first oder correlation is directly related to a collision operator which gives a transition probability between the molecular states, represented by Fermi’s golden rule. On the other hand, the physical quantity described by the higher order correlation cannot be reduced in the Landauer formula. As an example, we consider the particle current which is induced by external thermal force, as in the mechanical force when a molecule is subjected to an external electric field. However, it is well-known that the treatment of the thermal force is in general much more complex than a mechanical force, because the thermal force stems from a many-body dissipative effect.[@BalescuBook] We shall reveal that the particle current is attributed to the second order correlation component. Since there is no direct connection between the second order correlation and the collision operator, we cannot cast it into the Landauer formula unlike the case of the energy flow in this problem. In Section \[Sec:Model\] we present a model Hamiltonian of the molecular chain coupled with different thermal baths at its both ends. The eigenvalue problem of the Liouvillian is presented in Section \[Sec:Eigen\] to obtain the nonequilibrium stationary state as zero eigenstate of the Liouvillian. Energy flow is obtained with use of the nonequilibrium stationary state and Landauer formula is derived in Section \[Sec:EFlow\] where we explain why the energy flow can be cast into the Landauer formula. In Section \[Sec:Pflow\], the induced polarization, or its conjugate current, is evaluated as an example of the higher order correlation which cannot be reduced to Landauer formula. We illustrate in Section \[Sec:Discussion\] an application to DNA molecular chain. In Section \[Sec:Remarks\] we give some concluding remarks. For readers who are not familiar with the complex spectral representation of the Liouvillian in Appendix \[App:LiouvilleSpace\] we introduce the Liouville space and show the explicit expression of the interaction in terms of the Liouville basis, respectively. The complex spectral representation of Liouvillian is summarized in Appendix \[App:Complex\] where the eigenstate of the Liouvillian is given as a functional of the eigenstate of the collision operator. With use of the explicit expression of the interaction, we derive the collision operator in Appendix \[App:DerivKinEq\]. In Appendix \[App:BathChange\] we derive a formula which we use in Section \[Sec:Eigen\]. Model {#Sec:Model} ===== ![One-dimensional molecular chain composed of ${\cal N}$ molecular units each of which possesses a bound state. The left- and right-end states have the energies of $\varepsilon_{1}=\varepsilon_{L}$ and $\varepsilon_{\cal N}=\varepsilon_{R}$, respectively, while the other states have the same energy of $\varepsilon_{m}=\varepsilon_{0}$ for $m=2,\dots, {\cal N}-1$. A particle transfers between these states with the transfer integral of $J$. The molecular chain interacts with the two thermal baths with different temperatures of $T_{L}$ and $T_{R}$ at the both ends of the molecule. []{data-label="Fig1"}](Molecule_10.pdf){width="8cm"} We consider a one-dimensional molecular chain consisting in $\cal N$-molecular units each of which contains a single bound state. This molecular chain is coupled with two thermal baths at the both ends. The total Hamiltonian is written as $$\label{Htot} H=H_M+H_{B}+g H_{MB} \;,$$ where $H_{M}$ and $H_{B}$ describe the molecular chain and the thermal baths, respectively, and the interaction is represented by $H_{MB}$ with a dimensionless coupling constant $g$. We consider that a quantum particle transfers in the molecular chain. The molecular Hamiltonian is then described by a one-dimensional tight binding Hamiltonian: $$\label{HM} H_{M}= \sum_{m=1}^{\cal N} \varepsilon_{m}|m\>\<m| - \sum_{\<m,m'\>} J_{m,m'} |m\>\<m'| ,$$ where $\varepsilon_{m}$ is an energy of the bound state $|m\>$ at the $m$-th molecular unit as shown in Fig.\[Fig1\]. We denote the left end state as $|L\>\equiv |1\>$ and the right end state as $|R\>\equiv|N\>$. We assume that $\varepsilon_{m}=\varepsilon_{0}$ in the middle of the chain for $m=2,\dots,{\cal N}-1$, while the left and the right end states have different energies of $\varepsilon_{L}$ and $\varepsilon_{R}$. The second term represents the particle transfer where we take into account only the nearest neighbor transfer; $\<m,m'\>$ in the second term denotes taking a sum of the nearest neighbor bound states. In the present work, we assume a constant value for the transfer integrals of $J_{m,m'}=J$ for any $m$ and $m'$, and we take $J=1$ as an energy unit which also becomes a unit of temperature. The eigenvalue problem of the molecular Hamiltonian $H_{M}$ with the $\cal N$ dimension is solved to obtain the $\cal N$ eigenstates as $$\label{HmEVP} H_M| E_{\bar j}\>=E_{\bar j}|E_{\bar j}\> \quad ( \bar j=1,\cdots,{\cal N}) \;,$$ where an eigenstate $|E_{\bar j}\>$ is represented by $$\label{Barj} |E_{\bar{j}}\>=\sum_{m=1}^{\cal N} c_{m,\bar j} |m\> \;.$$ Hereafter in order to avoid a heavy notation, we simply describe $|E_{\bar j}\>$ as $|\bar j\>$. Both of the left and right thermal baths are assumed to be a three-dimensional harmonic crystal described by a Debye model in a large box of volume $L_{B}^{3}$. The Hamiltonian of the thermal bath systems reads $$\begin{aligned} \label{HB} H_B=\sum_{r=L,R}\sum_{\bf q}\hbar\omega_{r,\bf q}b^\dagger_{r,\bf q}b_{r,\bf q} \;,\end{aligned}$$ where $b_{r,{\bf q}}$ $(r=L,R)$ are the annihilation operators of the phonons of the thermal baths with the energy dispersion given by $\omega_{r,{\bf q}}=c|{\bf q}|$. We take the box normalization with a periodic boundary condition for the thermal bath systems which gives the discrete wave vectors for the phonon mode as $${\bf q}_{j}={\bf j}\triangle q \;,$$ where ${\bf j}$ is a three-dimensional integer vector and $\triangle q\equiv 2\pi/L_{B}$. In the large volume limit for the bath systems ($L_{B}\to\infty$), we have $$\label{LargeLim} {1\over \Omega}\sum_{{\bf q}_{j}}\cdots \longrightarrow \int d{\bf q} \cdots \;, \; \Omega \delta_{{\bf q},0} \longrightarrow \delta({\bf q}) \;,$$ where $\Omega\equiv (1/\triangle q)^{3}$. With use of the density of states of the thermal phonon system per volume ${\cal D}_{ph}(\omega)$, we can change the integral of $\bf q$ to the integral of $\omega$ as $$\label{IntDOS} {1\over \Omega}\sum_{{\bf q}_{j}} \xrightarrow{\Omega\to\infty} \int d{\bf q} \cdots =\int_{0}^{\infty} d\omega {\cal D}_{ph}(\omega) \cdots \;,$$ where ${\cal D}_{ph}(\omega)$ is given by $$\begin{aligned} \label{phonon DOS} {\cal D}_{ph}(\omega)= \begin{cases} { 4\pi \omega^{2}\over c^{3} } & \mbox{for } 0\le \omega \le \omega_{D} \\ 0 \; &\mbox{otherwise} \;. \end{cases}\end{aligned}$$ In Eq.(\[phonon DOS\]) Debye frequency $\omega_{D}$ is given by $$\omega_{D}=c\left( {3 N_{B} \over 4\pi \Omega} \right)^{1/3} \;,$$ where $N_{B}$ is a number of the normal phonon modes. In the present work, we take $\omega_{D}$ to be large enough compared to the other parameters so that the value of $\omega_{D}$ will not affect the results. The unperturbed Hamiltonian $H_{0}$ is then defined as a sum of $H_{M}$ and $H_{B}$: $$\label{H0} H_{0}=H_{M}+H_{B}\;.$$ We consider the interaction between the molecule and the thermal baths represented by $$\begin{aligned} \label{HMB} g H_{MB}&=&{g\over \sqrt{\Omega}} \sum_{\bf q} v_{L,{\bf q}} |L\>\<L|( b_{L,{\bf q}}+b_{L,{\bf q} }^{\dagger}) \nonumber\\ &&+ {g\over \sqrt{\Omega}}\sum_{\bf q} v_{R,{\bf q}} |R\>\<R|( b_{R,{\bf q}}+b_{R,{\bf q}}^{\dagger}) \nonumber \\ &&\equiv g ( H_{MB}^{L}+H_{MB}^{R} ) \;,\end{aligned}$$ where $v_{r,{\bf q}}$ $(r=L,R)$ are the interaction potentials, and we assume that $v_{r,{\bf q}}$ is independent of $\Omega$ and ${\bf q}$, i.e., $v_{r,{\bf q}}=v_{r}$. We consider that in Eq.(\[HMB\]) the molecular chain is interacting with a thermal bath only at the both ends of the chain so that the energies of the end states $|L\>$ and $|R\>$ are fluctuated by the interaction. With use of the eigenstates of $H_{M}$, the interaction Hamiltonian $ H_{MB}$ is represented by $$\<\bar j'|g H_{MB}|\bar j\>={g \over \sqrt{\Omega}}\sum_{r=L,R}\sum_{\bf q} v_r c_{r,\bar j'} c_{r,\bar j}(b_{r,{\bf q}}+b_{r,{\bf q}}^{\dagger} ) \;.$$ In Fig.\[Fig:Scheme\], we draw the molecular eigenstates $|\bar j\>$ of $H_{M}$ with the energy $E_{\bar j}$. These states are coupled with thermal baths through the coupling of the end state components. ![Molecular level and the interaction scheme. The molecular eigenstates $|\bar j\>$ are coupled with the thermal baths.[]{data-label="Fig:Scheme"}](Scheme2.pdf){width="8cm"} Nonequilibrium state as an eigenstate of Liouvillian {#Sec:Eigen} ==================================================== The time evolution of the total system obeys the Liouville-von Neumann equation $$i \frac{\partial}{\partial t} \rho (t) =\mathcal{L}\rho(t) \;, \label{Liouville}$$ where $\rho(t)$ is a density matrix of the total system and $\mathcal{L}$ is the Liouvillian defined by $ \mathcal{L}\rho \equiv [H,\rho]/\hbar . $ We assume that in the initial state the left and right thermal baths are in thermal equilibrium with different temperatures $T_{L}$ and $T_{R}$, respectively: $$\begin{aligned} \label{EqBath} \rho_{r}^{eq}=\frac{\exp\big[-\sum_q \hbar\omega_{r,q}b^\dagger_{r,q}b_{r,q} / k_{B}T_{r}\big]}{Z_{r}} \quad (r=L,R) \;,\end{aligned}$$ where $ Z_{r} =\prod_q \big(1-\exp[ -\hbar\omega_{r,q}/k_{B}T_{r}]\big)^{-1} $ with $k_{B}$ the Boltzmann constant. Hereafter we take $k_{B}=1$. The thermal distribution $\rho_{r}^{eq}$ gives the Planck’s distribution $$\label{Planck} n_{r}(\omega)={1\over \exp[-\hbar\omega/T_{r}]-1 } \;.$$ In the present work, we shall use the Liouville space representation which is briefly summarized in Appendix \[App:LiouvilleSpace\]. As a basis set for the particle system we shall introduce a Wigner representation defined by $$\label{WignerPtcl} |{\eta},{Y}\>\!\>\equiv|Y+{\eta\over 2};Y-{\eta \over 2}\>\!\>\equiv |Y+{\eta\over 2}\>\<Y-{\eta \over 2}| \;,$$ where $|Y+{\eta / 2}\>$ and $\<Y-{\eta / 2}| $ are the eigenstates of $H_{M}$: $|\bar j\>=|Y+\eta/ 2\>$ and $|\bar j'\>=|Y-\eta/ 2\>$. Then $|\eta,Y\>\!\>$ becomes an eigenstate of ${\cal L}_{M}$: $$\label{LMEV} {\cal L}_{M}|\eta, Y\>\!\>=\Delta_{\eta,Y}|\eta,Y\>\!\>$$ with the eigenvalue of $$\label{DelDef} \Delta_{\eta,Y}\equiv {1\over \hbar}\left( E_{Y+{\eta\over2}}-E_{Y-{\eta\over 2}} \right) \;.$$ Note the property $$\label{PropDelta} \Delta_{-\eta,Y}=-\Delta_{\eta,Y} \;.$$ By this definition, $\eta=0$ and $\eta\neq 0$ represent the diagonal and the off-diagonal components of the density matrix of the particle, respectively. Note that the representation of $|\eta,Y\>\!\>$ corresponds to the Wigner basis representation of $|k,P\>\!\>$ of a Boltzmann gas system where $k$ and $P$ correspond to $\eta$ and $Y$, respectively. Similarly the Wigner representation is defined for the phonon systems, we write the usual Wigner representation for a $(r,{\bf q})$ phonon mode as $$\label{WignerBath} |\nu_{r,{\bf q}}, N_{r,{\bf q}}\>\!\>\equiv |n_{r,{\bf q}};n'_{r,{\bf q}} \>\!\> \;,$$ where $$\nu_{r,{\bf q}}\equiv n_{r,{\bf q}}-n'_{r,{\bf q}} \;, \; N_{r,{\bf q}}\equiv { n_{r,{\bf q}}+n'_{r,{\bf q}} \over 2 } \;.$$ The Wigner basis $|\{\nu \},\{N\} \>\!\>$ satisfies $${\cal L}_{B}|\{\nu \},\{N\} \>\!\> =\nu\omega |\{\nu \},\{N\} \>\!\> \;,$$ where $\{\cdots\}$ denotes a set of all the phonon normal modes and $\nu\omega\equiv \sum_{r=L,R}\sum_{\bf q}\nu_{r,{\bf q}}\omega_{r,{\bf q}}$. The eigenstates of the unperturbed Liouvillian is then represented by a tensor products of $|\eta, Y\>\!\>$ and $|\{\nu\},\{N\}\>\!\>$ as $$\label{L0EV} {\cal L}_{0}|\eta, Y\>\!\>\!\otimes\! |\{\nu\},\{N\}\>\!\>=\left( \Delta_{\eta,Y}+\nu\omega \right) |\eta, Y\>\!\>\!\otimes\! |\{\nu\},\{N\}\>\!\> \;.$$ In terms of the Wigner basis of the eigenstates of $\mathcal{L}_{0}$, we can classify the Liouville space according to the order of correlations. For that purpose, we introduce the projection operators that specify the correlation components, such as a one-particle distribution of the particle $\hat{\cal P}^{(\eta)}$, correlation between the particle and the phonon systems $\hat{\cal P}^{(\eta,\nu_{r,\bf q})}$, and so on: \[Pclass\] $$\begin{aligned} && \hat{\cal P}^{(\eta)}\equiv\sum_{\{N\}}\sum_{Y} |\eta, Y\>\!\>{\<\!\<}\eta,Y| \otimes |\{0\},\{N\} \>\!\>\<\!\<\{ 0\}, \{N\}| \; , \nonumber\\ && \\ &&\hat{\cal P}^{(\eta,\nu_{r,{\bf q}})}=\sum_{\{N\}}\sum_{Y}|\eta,Y\>\!\>{\<\!\<}\eta,Y|\nonumber\\ &&\otimes |\nu_{r,{\bf q}},\{0\}'_{\bf q},N_{r,{\bf q}},\{N\}_{\bf q}' \>\!\>\<\!\<\nu_{r,{\bf q}},\{0\}'_{\bf q},N_{r,{\bf q}},\{N\}'_{\bf q}| \; , \nonumber\\ && \\ &&\hat{\cal P}^{(\eta,\nu_{r,{\bf q}_{1}},\nu_{r,{\bf q}_{2}})}=\sum_{\{N\}}\sum_{Y}|\eta,Y\>\!\>{\<\!\<}\eta,Y|\nonumber\\ &&\otimes |\nu_{r,{\bf q}_{1}},\nu_{r,{\bf q}_{2}},\{0\}'_{{\bf q}_{1},{\bf q}_{2}},N_{r,{\bf q_{1}}},N_{r,{\bf q_{2}}},\{N\}'_{{\bf q}_{1},{\bf q}_{2}} \>\!\> \nonumber\\ &&\qquad \<\!\<\nu_{r,{\bf q}_{1}},\nu_{r,{\bf q}_{2}},\{0\}'_{{\bf q}_{1},{\bf q}_{2}},N_{r,{\bf q_{1}}},N_{r,{\bf q_{2}}},\{N\}'_{{\bf q}_{1},{\bf q}_{2}}| \; , \nonumber\\ && \qquad \dots \;,\end{aligned}$$ where $\{\cdots\}'_{{\bf q}_{1},{\bf q}_{2},\cdots}$ means a set of all the normal modes other than ${\bf q}_{1},{\bf q}_{2},\cdots$. In order to consider the nonequilibrium stationary state which is obtained in the long time limit, it is appropriate to consider eigenvalue problem of ${\cal L}$, instead of solving Eq.(\[Liouville\]) as an initial value problem. For the system coupled with the thermal baths with an infinite degrees of freedoms, $\cal L$ may have a complex spectrum. As shown in Appendix \[App:Complex\], the complex eigenvalue problem of the Liouvillian is classified by the correlations: $$\label{LEVProb} \mathcal{L}|F_\alpha^{(\mu)})\!)=Z_\alpha^{(\mu)}|F_\alpha^{(\mu)})\!) \; , \; (\!(\tilde{F}_\alpha^{(\mu)}|\mathcal{L}=(\!(\tilde{F}_\alpha^{(\mu)}|Z_\alpha^{(\mu)} \;,$$ where $(\mu)$ is a combined index of $(\eta, \{\nu\})$ in Eqs.(\[Pclass\]) and $\alpha$ is an index of an eigenstate of a $(\mu)$-subspace. The eigenstates of $|F_\alpha^{(\mu)})\!)$ and $(\!(\tilde{F}_\alpha^{(\mu)}|$ are the left- and right-eigenstates of ${\cal L}$ with the complex eigenvalues $Z_\alpha^{(\mu)}$ for the total system composed of the particle and phonon systems. In Eq.(\[LEVProb\]), we have used $|\cdot )\!)$ notation instead of $|\cdot\>\!\>$ for the total system consisting in the particle and phonon systems. We briefly summarize the complex spectral representation of Liouvillian in Appendix \[App:Complex\], and the reader could consult Refs.[@Petrosky97; @Petrosky99; @Petrosky10] for the details. Since the nonequilibrium stationary state is time independent and is achieved in the long time, we should seek for the zero eigenstate of ${\cal L}$. As have shown in Appendix \[App:Complex\], by acting the projection operators on Eq.(\[LEVProb\]), we have the eigenvalue problem of a collision operator given by $$\label{App:EVPCol} \hat{\Psi}^{(\mu)} (Z_j^{(\mu)})|u_j^{(\mu)})\!) = Z_j^{(\mu)} | u_j^{(\mu)} )\!) \; ,$$ where $$\label{App:uj} |u_j^{(\mu)} )\!) = (N_j^{(\mu)})^{-1/2} \hat{\cal P}^{(\mu)} |F_j^{(\mu)} )\!)$$ is a [*privileged*]{} component of $|F_{j}^{(\mu)})\!)$ and $N_j^{(\mu)}$ is a normalization constant which is given by Eq.(\[App:Nnorm\]). Here, $\hat{\Psi}^{(\mu)}$ is the [*collision operator*]{} familiar to nonequilibrium statistical mechanics.[@PrigogineText; @Resibois67; @ResiboisBook; @BalescuBook] This operator is associated to [*diagonal transitions*]{} between two states corresponding to the same projection operator $\hat{\cal P}^{(\mu)}$: $$\label{App:Psi} \hat{\Psi}^{(\mu)}(z)={\hat{\cal P}}^{(\mu)} \mathcal{L}_0 \hat{\cal P}^{(\mu)} +\hat{\cal P}^{(\mu)}g\mathcal{L}_{MB}\hat{\cal Q}^{(\mu)}\hat{\mathcal{C}}^{(\mu)}(z) \hat{\cal P}^{(\mu)} \; ,$$ where $\hat{\cal Q}^{(\mu)}\equiv 1-\hat{\cal P}^{(\mu)}$. In Eq.(\[App:Psi\]) $$\label{App:Creation} \hat{\mathcal{C}}^{(\mu)}(z)=\frac{1}{z-\hat{\cal Q}^{(\mu)}\mathcal{L}\hat{\cal Q}^{(\mu)}}\hat{\cal Q}^{(\mu)}g\mathcal{L}_{MB}\hat{\cal P}^{(\mu)}$$ is called the creation-of-correlation operator, or simply the [*creation operator*]{}.[@Petrosky97] Eqs.(\[LEVProb\]) and (\[App:EVPCol\]) show that the spectra of $\cal L$ and $\hat\Psi$ are identical. This directly links the microscopic dynamics to macroscopic kinetic processes. We can expand the creation operator in a series expansion of $g$ as $$\begin{aligned} \label{Cexpand} \hat{\mathcal{C}}^{(\mu)}(z)&=&{1 \over z-\hat{\cal Q}^{(\mu)}\mathcal{L}_0 \hat{\cal Q}^{(\mu)}} \nonumber\\ &&\times \sum_{\xi=0}^\infty g^{\xi} \left ( \hat{\cal Q}^{(\mu)} \mathcal{L}_{MB} \hat{\cal Q}^{(\mu)} {1 \over z-\hat{\cal Q}^{(\mu)}\mathcal{L}_0 \hat{\cal Q}^{(\mu)}} \right)^\xi \nonumber\\ &&\times g \left ( \hat{\cal Q}^{(\mu)}\mathcal{L}_{MB}\hat{\cal P}^{(\mu)} \right ) \nonumber \\ &\equiv& \sum_{\xi=0}^\infty g^{\xi + 1} {\cal C}_{\xi +1}^{(\mu)}(z) \;.\end{aligned}$$ The lower order expansion of the creation operator reads \[C1C2\] $$\begin{aligned} \hat{\mathcal{C}}^{(\mu)}_1(z)&=&{1 \over z-\hat{\cal Q}^{(\mu)}\mathcal{L}_0 \hat{\cal Q}^{(\mu)}} \hat{\cal Q}^{(\mu)}\mathcal{L}_{MB}\hat{\cal P}^{(\mu)} \;, \\ \hat{\mathcal{C}}_2^{(\mu)}(z)&=&{1 \over z-\hat{\cal Q}^{(\mu)}\mathcal{L}_0 \hat{\cal Q}^{(\mu)}} \hat{\cal Q}^{(\mu)} \mathcal{L}_{MB} \hat{\cal Q}^{(\mu)} \nonumber\\ &&\times {1 \over z-\hat{\cal Q}^{(\mu)}\mathcal{L}_0 \hat{\cal Q}^{(\mu)}} \hat{\cal Q}^{(\mu)}\mathcal{L}_{MB}\hat{\cal P}^{(\mu)} \; . \end{aligned}$$ Substituting Eq.(\[Cexpand\]) into Eq.(\[App:Psi\]) we have the expansion of $\hat{\Psi}^{(\mu)}(z)$ as $$\hat{\Psi}^{(\mu)}(z)={\hat{\cal P}}^{(\mu)} \mathcal{L}_0 \hat{\cal P}^{(\mu)} +g^2 \hat\psi_2^{(\mu)}(z) +O(g^4) \;,$$ where $$\label{psi2} \hat{\psi}^{(\mu)}_2(z)= \hat{\cal P}^{(\mu)}\mathcal{L}_{MB}\hat{\cal Q}^{(\mu)} \hat{\mathcal{C}}^{(\mu)}_1(z) \;.$$ On the other hand, the non-privileged component ($\hat{\cal Q}^{(\mu)}$-component) is obtained as a functional of the privileged component by using the creation operator as shown in Eq.(\[QF\]). The eigenstate $|F_{j}^{(\mu)})\!)$ is then written as $$\label{App:Fright} |F_{j}^{(\mu)})\!)=\sqrt{N_{j}^{(\mu)}} \Big(\hat{\cal P}^{(\mu)}+\hat{\mathcal{C}}^{(\mu)}(Z_{j}^{(\mu)}) \Big) |u_{j}^{(\mu)})\!) \; .$$ Since the nonequilibrium stationary state $|F_{0}^{(0)})\!)$ is a zero eigenstate of $\cal L$, it is represented by the zero eigenstate of the collision operator $|u_{0}^{(\mu)})\!)$ as $$\label{Fright} |F_0^{(0)})\!)=\sqrt{N_0^{(0)}}\Big( \hat {\cal P}^{(0)}+\hat{\mathcal C}^{(0)}(+i0)\Big)|u_{0}^{(\mu)})\!) \; ,$$ where the direction of the analytic continuation of $\hat{\cal C}^{(0)}$ is indicated by $z=+i0$, which is consistent with the fact that the approach to equilibrium is achieved in our future. In the weak coupling case considered here, the eigenstate $|F_0^{(0)})\!)$ is represented by the expansion of the interaction up to the second order as $$\label{F00} |F_0^{(0)})\!)=\sqrt{N_0^{(0)}} \left( 1+g\hat{\mathcal C}_{1}^{(0)}(+i0)+g^{2}\hat{\mathcal C}_{2}^{(0)}(+i0)\right) |u_{0}^{(\mu)})\!) \;,$$ where $\hat{\mathcal C}_{1}^{(0)}$ and $\hat{\mathcal C}_{2}^{(0)}$ are given in Eqs.(\[C1C2\]). As shown in Appendix \[App:Complex\], the left eigenstate is similarly determined by the expansion of the interaction up to the second order as $$\label{F00Left1} (\!(\tilde F_0^{(0)}|=\sqrt{N_0^{(0)}} (\!(\tilde v_{0}^{(0)}|\left( 1+g\hat{\mathcal D}_{1}^{(0)}(+i0)+g^{2}\hat{\mathcal D}_{2}^{(0)}(+i0)\right) \;,$$ where $\hat{\cal D}^{(0)}_{1}(z)$ and $\hat{\cal D}^{(0)}_{2}(z)$ are respectively the first and second order destruction-of-correlation operators which are determined by the series expansion of $g$ from $\hat{\cal D}^{(0)}(z)$ in Eq.(\[App:Destruction\]). Moreover, $(\!(\tilde v_{0}^{(0)}|$ is the left eigenstate of the collision operator shown in Eq.(\[App:EVPColLeft\]), and the normalization constant $N_{0}^{(0)}$ is given by Eq.(\[App:Nnorm\]). In order to determine $|u_{0}^{(\mu)})\!)$, we shall now solve the eigenvalue problem of the collision operator Eq.(\[App:EVPCol\]). Since the number of degrees of freedom of the thermal phonon system is infinitely larger than the number of molecular system $\cal N$, it can be shown that the $\hat{\cal P}^{(0)}$-component of the phonon systems does not change in time and stay in their initial canonical distributions characterized by the initial temperatures $T_{L}$ and $T_{R}$. Therefore we search for the eigenstate in the form of the tensor product of the density matrices of the molecular system and the phonon systems as $$\label{uj0} |u_{j}^{(0)})\!)=|\varphi_{j}^{(0)}\>\!\>|\rho_{L}^{eq}\rho_{R}^{eq}\>\!\> \;,$$ where $|\rho_{L}^{eq}\rho_{R}^{eq}\>\!\>$ is a tensor product of the Liouville space vector of the left and right thermal phonon equilibrium distributions represented by Eq.(\[EqBath\]). In Eq.(\[uj0\]), $|\varphi_{j}^{(0)}\>\!\>$ is an eigenstate of the reduced collision operator for the molecular system defined by $$\bar{\Psi}^{(0)}(z)\equiv {\rm Tr}_{L\otimes R}\left[\hat\Psi^{(0)}(z) |\rho_{L}^{eq}\rho_{R}^{eq}\>\!\>\right] \;,$$ where ${\rm Tr}_{L\otimes R}$ stands for taking a partial trace of the thermal phonon systems, and therefore $\bar{\Psi}^{(0)}(z)$ is still a super-operator working on the molecular system. Up to the second order of the interaction, the reduced collision operator is represented by $$\begin{aligned} \label{Psi2Expression} \bar{\Psi}^{(0)}_{2}&=&g^{2} {\mathop\mathrm{Tr}\nolimits}_{L\otimes R} \left[\hat{\cal P}^{(0)}\mathcal{L}_{MB}\hat{\cal Q}^{(0)} \frac{1}{i0^+-\mathcal{L}_0} \right. \nonumber \\ &&\times \left. \hat{\cal Q}^{(0)} \mathcal{L}_{MB}\hat{\cal P}^{(0)}|\rho_{L}^{eq}\rho_{R}^{eq}\>\!\>\right] \;.\end{aligned}$$ There are eight diagrams for $\bar{\Psi}^{(0)}_{2}$ in our model as shown in Appendix \[App:DerivKinEq\]. In the present model, the matrix element of $\bar{\Psi}^{(0)}_{2}$ is explicitly represented in terms of the transition probabilities as \[RedPsi2\] $$\begin{aligned} &&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y\>\!\>=-i\sum_{r=L,R}\sum_{\eta >0} \Big(k_{r}^{Y-\eta,Y} +k_{r}^{Y+\eta,Y} \Big) \;,\nonumber \\ &&\\ && \<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y+\eta\>\!\>=i k_{r}^{Y,Y+\eta} \;,\\ &&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y-\eta\>\!\> = i k_{r}^{Y,Y-\eta} \;,\end{aligned}$$ where $k_{r}^{Y,Y\pm\eta}$ is a transition probability from $|0,Y\pm\eta\>\!\>$ ($\eta>0$) to the $|0,Y\>\!\>$ given by Eqs.(\[App:tranProb1\]) and (\[App:tranProb2\]). When we define the reduced density operator as $$\label{RhoM} f(t) \equiv \mathrm{Tr_{L\otimes R}}\Big[|\rho(t)\>\!\>\Big] \;,$$ it is found that for the weakly coupled system the $f(t)$ obeys the following Pauli master equation of $$\begin{aligned} \label{dtfj} \frac{d}{dt}f_{Y}^{(0)}(t)&=&-\sum_{\eta \neq 0}\sum_{r=L,R}\Big\{ k_{r}^{Y+\eta,Y}f_{Y}^{(0)}(t) \nonumber\\ && - k_r^{Y,Y+\eta} f_{Y+\eta}^{(0)}(t) \Big\} \;,\end{aligned}$$ where we have defined as $$\label{fY0Def} f_{Y}^{(0)}(t)\equiv \<\!\<0,Y|f(t)\>\!\> \;.$$ It has been known that this type of the master equation has a unique zero eigenstate and the other eigenstates have negative imaginary values of their eigenvalues.[@Keizer72; @Schnakenberg76; @ModernThermo; @HakenBook] The zero value right eigenstate of $\bar{\Psi}_{2}^{(0)}$ is called [*collision invariant*]{} which is represented by $$\label{ColInv} |\varphi_0^{(0)}\>\!\> =\sum_{Y}\phi_{Y}|0,Y\>\!\> \;.$$ Here we take the normalization condition for $\phi_{Y}$ as $$\sum_{Y}\phi_{Y}=1 \;,$$ which gives from Eq.(\[uj0\]) that $\mathrm{Tr}\Big[ |u_{0}^{(0)})\!) \Big] =1$. Correspondingly the collision invariant for the left eigenstate is given by $$\label{v00-2} (\!(\tilde v_{0}^{(0)}|=\sum_{Y}\sum_{\{N\}}\<\!\<0,Y|\<\!\<\{0\},\{N\}| \;,$$ which satisfies $(\!(\tilde v_{0}^{(0)}|u_{0}^{(0)})\!)=1$. Substituting Eq.(\[ColInv\]) into Eqs.(\[uj0\]) and (\[F00\]), the nonequilibrium stationary state is obtained up to the second order of the interaction as $$\begin{aligned} \label{F00-2} |F_0^{(0)})\!)&=&\sqrt{N_0^{(0)}}\left( 1+g\hat{\mathcal C}_{1}^{(0)}(+i0)+g^{2}\hat{\mathcal C}_{2}^{(0)}(+i0)\right) \nonumber\\ &&\times \sum_{Y}\phi_{Y}|0,Y\>\!\>\prod_{r=L,R}|\rho_{r}^{eq}\>\!\> \;.\end{aligned}$$ Similarly the left eigenstate for the stationary state can be obtained as $$\begin{aligned} \label{F00Left} (\!( \tilde F_{0}^{(0)}|&=&\sqrt{N_0^{(0)}}\sum_{Y}\<\!\<0,Y|\sum_{\{N\}}\<\!\<\{0\},\{N\}| \nonumber\\ &&\times\left( 1+g\hat{\mathcal D}_{1}^{(0)}(+i0)+g^{2}\hat{\mathcal D}_{2}^{(0)}(+i0)\right) \;,\end{aligned}$$ which satisfies $ (\!(\tilde F_{0}^{(0)}|F_{0}^{(0)})\!)=1 $ and $$\label{F00Norm3} \mathrm{Tr}\Big[F_{0}^{(0)}\Big]=\sqrt{N_0^{(0)}} \;.$$ First we shall consider the case for ${\cal N}=2$ where the nonequilibrium stationary state is analytically obtained. The nonequilibrium stationary population $\phi_{Y}$ ($Y=1, 2$) are obtained by \[N2 Collision Invariant\] $$\begin{aligned} \phi_{1}&=&{v_L^2 ( n_{L}(\Delta)+1) +v_R^2 ( n_{R}(\Delta)+1) \over v_L^2 (2 n_{L}(\Delta)+1) +v_R^2 (2 n_{R}(\Delta)+1) } \;, \\ \phi_{2}&=&{v_L^2 n_{L}(\Delta) +v_R^2 n_{R}(\Delta) \over v_L^2 (2 n_{L}(\Delta)+1) +v_R^2 (2 n_{R}(\Delta)+1) }\;,\end{aligned}$$ where $\Delta$ is given as an energy differnce between the two molecular states of $|E_{\bar 2}\>$ and $|E_{\bar 1}\>$ (See Eq.(\[DelDef\])): $$\label{DeltaDef} \Delta\equiv\Delta_{{1},{3\over2}}={ E_{\bar 2}-E_{\bar 1}\over \hbar }= {1\over\hbar}\sqrt{(\varepsilon_{L}-\varepsilon_{R})^{2}+4 J^{2}} \;.$$ ![(a) The molecular level structure for the chain with ${\cal N}=10$, where the parameters are $\varepsilon_{L}=\varepsilon_{R}=0, J=1.0, v_{R}=v_{L}=v=1.0$, and $g=0.1$. (b) The spectrum of the collision operator $Z_{\alpha}^{(0)}$ as a function of $T_{R}$ changing from 0 to 4.0 with $T_{L}=2.0$ fixed, where the vertical axis denotes $-\rm{Im}Z_{\alpha}^{(0)}$. (c) The state population of the nonequilibrium stationary state $\phi_{Y}$ as a function of $T_{R}$. The dotted lines are the population for the average temperature of $T_{M}=(T_{L}+T_{R})/2$. []{data-label="Fig:Kspec"}](Fig2.pdf){width="8cm"} For a longer molecular chain, it is difficult to analytically obtain the nonequilibrium stationary state. Instead we have numerically solved the eigenvalue problem of the collision operator and obtained the collision invariant. As an example, we show in Fig.\[Fig:Kspec\] the results for a molecule with ${\cal N}=10$: the molecular level structure, the eigenstates of the molecular Hamiltonian (a), the spectrum of the collision operator (b), and the population of the collision invariant (c). In Fig.\[Fig:Kspec\](b) and (c) we have fixed at $T_{L}=2.0$ and changed $T_{R}$ and have taken $ v_{L}=v_{R}=1$ and $g=0.1$. Since there are ten basis states belonging to $P^{(0)}$ subspace ( $|0,Y\>\!\>$ with $Y={1}$ to ${10}$ ), we have ten eigenstates of the collision operator. For any $T_{R}$, there always exists an eigenstate with $Z_{0}^{(0)}=0$, [*i.e.*]{} a collision invariant, while other eigenstates are decaying states with $\rm{Im}Z_{\alpha(\neq 0)}^{(0)}<0$, as mentioned above. In Fig.\[Fig:Kspec\](c), we have also shown by the dotted lines the canonical distribution for the average temperature $T_{M}\equiv (T_{L}+T_{R})/2$. At the low temperature, the nonequilibrium state population is different from the canonical distribution for the average temperature, while both agree well at the higher temperature. Furthermore, at the low temperature case where the discrete molecular level structure plays a key role, the nonequilibrium state population cannot be represented by the canonical distribution for any unique temperature. This suggests that due to the quantum effect the local temperature of the molecule cannot be identified at the low temperature case, in contrast to the classical system.[@Rego98] Energy Flow {#Sec:EFlow} =========== In this section, we consider the energy flow in the nonequilibrium stationary state in terms of the zero eigenstate of $\cal L$ obtained in Section \[Sec:Eigen\]. The energy flow through the molecule is evaluated by considering the energy change of the molecule. The energy change of the molecule in the nonequilibrium stationary state is obtained by $$\label{Wdef} w(t)\equiv \frac{d}{dt}\<H_M\>_t= \frac{d}{dt}{\rm Tr}\left[H_{M}\rho(t)\right] \;.$$ Using Eq.(\[Liouville\]), $w(t)$ is represented by \[WtRW\] $$\begin{aligned} w(t) &=&-i\<\!\<H_M|{\rm Tr}_{L\otimes R}\big[{\cal L}\rho(t)\big] \>\!\> \\ &=&-i g \<\!\< H_M|{\rm Tr}_{L\otimes R}\big[{\cal L}_{MB} \rho(t)\big] \>\!\> \;,\end{aligned}$$ where we have used the fact that $\left( {\cal L}_{M} H_{M}\right)^{\dagger}=0$ and ${\rm Tr}_{L\otimes R}[{\cal L}_{B}\rho(t)]=0$ in the second equality. Taking into account ${\cal L}_{MB}={\cal L}_{MB}^{L}+{\cal L}_{MB}^{R}$ corresponding to $H_{MB}=H_{MB}^{L}+H_{MB}^{R}$ in Eq.(\[HMB\]), we can divide $w(t)$ into two contributions due to the interactions with the left and the right thermal baths: $$\begin{aligned} \label{WLWR} w(t)=w^{L}(t)+w^R(t) \;,\end{aligned}$$ where $$\label{WBdef} w^{r}(t)\equiv-i g \<\!\<H_M|{\rm Tr}_{L\otimes R}\left[{\cal L}_{MB}^{r} \rho(t)\right] \>\!\> \quad (r=L,R) \;.$$ The sign of $w^{r}(t)$ is positive when energy flow comes into the molecule from the thermal bath $r=L$ or $R$. Since ${\cal L}F_{0}^{(0)}=0$ in the nonequilibrium stationary state, we find that $w (\infty)=-i\<\!\<H_M|{\rm Tr}_{L\otimes R}\big[{\cal L}F_{0}^{(0)}\big] \>\!\>=0$, resulting in $w^{L}(\infty )=-w^{R}(\infty )$ which guarantees that influx and outflow of molecular energy are balanced in the stationary state. We then define the energy flow from the left to the right thermal bathes going through the molecule as $$\label{Energy Flow Def} {\cal U}(t)\equiv {1\over 2}\left( w^{L}(t)-w^{R}(t)\right) \;.$$ Since in the long time Eq.(\[F00Norm3\]) indicates that $$\label{rhoLim} |\rho(t) )\!) \xrightarrow{t\rightarrow\infty} {1\over \sqrt{N_0^{(0)}}}|F_{0})\!) \;,$$ it is found from Eq.(\[F00-2\]) that ${\cal U}(\infty)$ can be represented up to the second order of ${\cal L}_{MB}$ as $$\label{Wst2} {\cal U}(\infty)= -i g^{2}\<\!\< H_M|\mathrm{Tr}_{L\otimes R} \left[{\cal L}_{MB}^{L}\hat{\cal C}_{1}^{(0)}(+i0)\varphi_0^{(0)} \rho_{L}^{eq}\rho_{R}^{eq}\right] \>\!\> \;.$$ It should be noted that the first order correlation created from the collision invariant contributes to the energy flow in the nonequilibrium stationary state.[@PrigogineText; @ResiboisBook; @Resibois67] Furthermore, by comparison with Eq.(\[psi2\]), it is found that the energy flow is related to the collision operator which is expressed by the transition probability given in Eq.(\[RedPsi2\]). Inserting Eqs.(\[ColInv\]) and (\[C1C2\]a) into (\[Wst2\]) and using the matrix representation of ${\cal L}_{MB}^{L}$ given in Eqs.(\[App:LMBcomp\]), we can explicitly represent ${\cal U}^{st}\equiv{\cal U}(\infty)$ as a sum of any pair of the states of $|0,Y\>\!\>$ and $|0,Y+\eta\>\!\>$: \[WstExp\] $$\begin{aligned} &&{\cal U}^{st}\equiv \sum_{\eta>0}\sum_{Y} {\cal U}_{Y+\eta,Y} \nonumber\\ &&=\frac{4\pi g^2 v_L^2}{\hbar} \sum_{\eta>0}\sum_{Y}{\cal D}^{ph}(\Delta_{\eta,Y+{\eta\over2}})\big|c_{L,Y+\eta}^{*}c_{L,Y}\big|^2 \Delta_{\eta,Y+{\eta\over2}} \nonumber\\ &&\times\bigg\{ \phi_{Y}n_L(\Delta_{\eta,Y+{\eta\over2}})-\phi_{Y+\eta} \big( n_L(\Delta_{\eta,Y+{\eta\over2}})+1 \big) \bigg\} \nonumber\\ &&\\ &&=2 \sum_{\eta>0}\sum_{Y}\hbar \Delta_{\eta,Y+{\eta\over2}}\Big\{ k_L^{Y+\eta,Y} \phi_{Y}-k_L^{Y,Y+\eta} \phi_{Y+\eta} \Big\} \;, \nonumber \\ \end{aligned}$$ where $2\textstyle\sum_{\eta>0}$ in Eq.(\[WstExp\]b) can be replaced by $\textstyle\sum_{\eta\neq0}$. Here we would like to make a comment that the same formula can be derived by evaluating energy change of the thermal bath as shown in Appendix \[App:BathChange\]. We note that there is a striking correspondence between Eq.(\[WstExp\]b) and Eq.(\[dtfj\]): The right hand side of the two equations involve a common factor of $( k_L^{Y+\eta,Y} \phi_{Y}-k_L^{Y,Y+\eta} \phi_{Y+\eta} )$ which is the transition rate per time between the states of $|0,Y\>\!\>$ and $|0,Y+\eta\>\!\>$ due to the coupling with the left thermal bath. The energy flow coming in from the left thermal bath is obtained by multiplying with it the energy difference between the states of $|0,Y\>\!\>$ and $|0,Y+\eta\>\!\>$ states, $\hbar\Delta_{\eta,Y+\eta/2}$. This correspondence is originated in the fact that the collision operator $\hat\psi_{2}^{(0)}(z)$ can be represented by the first order creation operator $\hat{\cal C}_{1}^{(0)}(z)$ as shown in Eq.(\[psi2\]). Therefore in the weak coupling case the energy flow can be derived by using the solution of the eigenvalue problem of the collision operator: Simply multiplying the energy difference of the states and the transition rate between them. This correspondence naturally leads it to a Landauer formula which has been widely used to interpret the carrier flow of the mesoscopic system in nonequilibrium situation.[@Landauer57; @Rego98; @Schwab00; @Wharam88] Landauer formula reads $$\label{wst} {\cal U}^{st}=\int_{0}^{\infty} \hbar \omega{\cal T}(\omega ) \big(n_{L}(\omega)-n_{R}(\omega)\big) d\omega\;,$$ where ${\cal T}(\omega )$ is a transmission function determined by using Fermi’s golden rule. In the simplest approximation, ${\cal T}(\omega)=1$ is assumed, which results in a quantization of thermal conductance.[@Rego98] In more elaborate works, they have estimated the $\omega$ dependence of ${\cal T}(\omega)$ reflecting a resonance effect due to the discretized molecular level structure.[@Segal01; @Segal03] In the present model, the equation (\[WstExp\]) can be cast into the form of Eq.(\[wst\]) in terms of the transmission function defined by $$\begin{aligned} \label{TransmissionF} &&{\cal T}(\omega)= 2 \sum_{\eta>0}\sum_{Y} \delta (\omega-\Delta_{\eta,Y+{\eta\over 2}}) \nonumber\\ & & \times{ (k_L^{Y+\eta,Y}-k_R^{Y+\eta,Y} )\phi_{Y}- (k_L^{Y,Y+\eta}-k_R^{Y,Y+\eta} )\phi_{Y+\eta} \over n_L(\Delta_{\eta,Y+\eta})- n_R(\Delta_{\eta,Y+\eta}) } \;. \nonumber\\ &&\end{aligned}$$ The transmission function has a strong resonance characteristic for the molecular level structure, which is reflected in the delta function of Eq.(\[TransmissionF\]): $\omega=(E_{Y+\eta}-E_{Y})/\hbar$. While the energy flow due to the first order correlation can be represented by the Landauer formula as shown above, we shall show in the next section a physical quantity due to the higher order correlation, such as an induced polarization, cannot be represented by the Landauer formula. It is worthwhile to note that the Landauer formula is derived in our approach through the resonance effects between the molecular chain and the thermal baths, i.e., the dissipation occurs at the edges of the system contacting to the baths. Before going to the next section, we shall show some examples of the energy flow in the nonequilibrium stationary state of the molecule. The energy flow is analytically obtained for a molecule with the length of ${\cal N}=2$ by using Eq.(\[N2 Collision Invariant\]). In this case, we have $$\begin{aligned} \label{T2site} {\cal T}(\omega) &=&\delta(\omega-\Delta){2\pi g^{2} \over \hbar^2} {4 J^2 \over (\varepsilon_L-\varepsilon_R)^2 + 4 J^2} \; {\cal D}_{ph}(\Delta) \nonumber\\ &\times&{v_L^2 v_R^2\over (2 n_L(\Delta)+1)v_L^2+(2 n_R(\Delta)+1)v_R^2 } \;,\end{aligned}$$ yielding $$\begin{aligned} \label{N2 energy flow} {\cal U}^{st}&=&{2\pi g^2\over \hbar} { J^2 \over (\varepsilon_L-\varepsilon_R)^2 + 4 J^2} \; {\cal D}_{ph}(\Delta)\nonumber\\ &\times& {2 v_L^2 v_R^2 (n_L(\Delta)-n_R(\Delta))\over (2 n_L(\Delta)+1)v_L^2+(2 n_R(\Delta)+1)v_R^2 }\end{aligned}$$ where $\Delta$ is given by Eq.(\[DeltaDef\]). ![The energy flow ${\cal U}^{st}$ of a molecule with the length of ${\cal N}=10$ in the $T_{L}$-$T_{R}$ plane. The parameters for the calculation are the same as used in Fig.\[Fig:Kspec\]. The temperature range is taken from 0 to 10 in (a), and from 0 to 1 in (b). The energy flow as a function of $T_{R}$ for several fixed values of $T_{L}$ are shown in (c) and (d).[]{data-label="Fig:EFlow"}](EFlowTRTL7.pdf){width="8cm"} For a longer molecule we have numerically calculated the energy flow. As an example, we have shown in Fig.\[Fig:EFlow\] the energy flow through the same molecule as studied in Section \[Sec:Eigen\]. We show the energy flow ${\cal U}^{st}$ in the bird’s-eye view in the $T_{L}$-$T_{R}$ plane in Fig.\[Fig:EFlow\](a) and (b). The wide temperature region of ${\cal U}^{st}$ is shown in (a), and the low temperature region in an expanded scale in (b). We also show the energy flow ${\cal U}^{st}$ as a function of $T_{R}$ for various values of $T_{L}$ in (c) and (d). It should be noted that while the energy flow is linearly proportional to the temperature difference near thermal equilibrium $T_{L}\simeq T_{R}$, this linear relation breaks down in far-from equilibrium. ![The dependence of the energy flow ${\cal U}^{st}$ on $\varepsilon_{L}$ and $\varepsilon_{R}$ with $T_{L}=2.0$ and $T_{R}=0.01$ fixed, where the other parameters are the same as used in Fig.\[Fig:EFlow\]. The bird’s eye view and the contour plot are shown in (a) and (b), respectively. []{data-label="Fig:ELER"}](ELER2.pdf){width="8cm"} The energy flow ${\cal U}^{st}$ also depends on $\varepsilon_{L}$ and $\varepsilon_{R}$. We show in Fig.\[Fig:ELER\] the energy flow ${\cal U}^{st}$ as a function of $\varepsilon_{L}$ and $\varepsilon_{R}$ for $T_{L}=2.0$ and $T_{R}=0.01$, where the other parameters are the same as in Fig.\[Fig:EFlow\]. The bird’s eye view and the contour plot are shown in (a) and (b), respectively. The ${\cal U}^{st}$ is maximized at $\varepsilon_{L}-\varepsilon_{R}=0$ as a function of $\varepsilon_{L}-\varepsilon_{R}$, because a transfer between the end states $|L\>$ and $|R\>$ through the molecule occurs most effectively. For different values of $\varepsilon_{L}$ and $\varepsilon_{R}$, the energy transfer is allowed to occur due to the energetically spread molecular states as shown in Fig.\[Fig:Scheme\]. As temperature increases, the energy flow increases. Induced Polarization, Particle current {#Sec:Pflow} ====================================== In spite of the fact that the thermal force stems from complicated many-body dissipative effect, when the polarizable molecule is subject to an external thermal force under nonequilibrium condition, a polarization is induced, just like a simple mechanical force with an external electric field. We shall show that an induced polarization within the molecular states is attributed to the second order correlation so that the quantity cannot be reduced in a Landauer formula unlike the energy flow. Polarization operator is represented by $$\label{Polarization} e \hat x=\sum_{m=1}^{\cal N} e x_{m} |m\>\<m| \;,$$ when $x_{m}=m d$ with a lattice constant $d$ and the electric charge $e$. In the present work, we take $e=1$. The conjugate current to the polarization is defined as a time derivative of the polarization: $${\cal I}(t)\equiv \frac{d}{dt}\<\hat{x}\>_t= {d\over dt}{\rm Tr}[ \hat x \rho(t) ] \;.$$ Using Eq.(\[Liouville\]), ${\cal I}(t)$ is then represented as $$\begin{aligned} \label{Current} {\cal I}(t)\equiv \frac{d}{dt}\<\hat{x}\>_t&=&-i \<\!\< \hat x|{\rm Tr}_{L\otimes R}[\mathcal L \rho(t) ] \>\!\> \nonumber\\ &=&-i\<\!\< \hat x| \mathcal L_{M} |f(t) \>\!\> \;,\end{aligned}$$ where we have used the fact that $\<\!\< \hat x|{\rm Tr}_{L\otimes R} [{\cal L}_{B}+g {\cal L}_{MB}]=0$, and $f(t)$ is a reduced density operator of the particle given by Eq.(\[RhoM\]). When we define the current operator as $$\hat{\cal I}\equiv i[H_{M},\hat x] \;,$$ then the current ${\cal I}(t)$ is given by ${\cal I}(t)=\mathrm{Tr}[\hat{\cal I} f(t)]$. Note that the current operator $\hat{\cal I}$ may be represented by using the site basis $\{ |m\>\}$ as $$\begin{aligned} \label{CurOp} &&\hat{\cal I} =\frac{i}{\hbar}\sum_{m=1}^{N} x_m\big[ H_M,|m\>\<m|\big] \nonumber \\ &&= - \frac{i d}{\hbar} \sum_{m=1}^{N-1} J_{m,m+1} \left(|m+1\>\<m|-|m\>\<m+1|\right) \;,\nonumber\\\end{aligned}$$ which agrees with the ordinary definition of the particle current in a one-dimensional discrete lattice.[@MahanBook] Considering that $\rho(t)$ in the long time limit is given by $F_{0}^{(0)}$ Eq.(\[rhoLim\]), we then have $$\label{TotalI} {\cal I}^{st}\equiv -i \<\!\< \hat x|{\rm Tr}_{L\otimes R} [\mathcal LF_{0}^{(0)}] \>\!\> =0 \;,$$ i.e., the total particle current in the nonequilibrium stationary state vanishes because ${\cal L}F_{0}^{(0)}=0$. Substituting the spectral representation of ${\cal L}_{M}$ given by Eq.(\[LMEV\]) into Eq.(\[Current\]), we have \[CurrentComp\] $$\begin{aligned} {\cal I}(t)&=&-i\sum_{\eta(\neq 0),Y}\<\!\<\hat x|\eta,Y\>\!\>\Delta_{\eta,Y}\<\!\<\eta,Y|f(t)\>\!\> \nonumber\\ &&\\ &=&-i\sum_Y\sum_{\eta> 0}\<\!\<\hat x|\eta,Y\>\!\>\Delta_{\eta,Y} \nonumber\\ &&\quad\times \big(\<\!\<\eta,Y|-\<\!\<-\eta,Y| \big) |f(t)\>\!\>\\ &\equiv&\sum_Y\sum_{\eta> 0} {\cal I}_ {\eta,Y}(t) \;,\end{aligned}$$ where we have defined a particle current component ${\cal I}_{\eta,Y}(t)$. The current component ${\cal I}_{\eta,Y}(t)$ is a product of the three factors: i) Transition dipole moment between $|Y+\eta\>$ and $|Y-\eta\>$ molecular states, $\<\!\<\hat x|\eta,Y\>\!\> =\<Y-\eta/2|\hat x|Y+\eta/2\>$, ii) Transition frequency between these two states, $\Delta_{\eta,Y}$, and iii) Off diagonal matrix element of the reduced density matrix, $(\<\!\<\eta,Y|-\<\!\<-\eta,Y|)|f(t)\>\!\> $. The product of the factors i) and ii) represents that the particle current is attributed to the transition of the particle between $|Y-\eta/2\>$ and $|Y+\eta/2\>$ molecular states. Furthermore, since the factor iii) is written as $$\begin{aligned} \label{ftoffstate} &&(\<\!\<\eta,Y|-\<\!\<-\eta,Y|\Big)|f(t)\>\!\>\nonumber\\ &&=\big\langle Y+{\eta\over2}\big|f(t)\big|Y-{\eta\over2}\big\rangle- \big\langle Y-{\eta\over2}|f(t)\big|Y+{\eta\over2}\big\rangle \;, \nonumber\\\end{aligned}$$ the reduced density matrix $f(t)$ should be non-Hermitian in order to obtain a non-vanishing ${\cal I}_{\eta,Y}(t)$. This is quite contrast to a thermal equilibrium at $\beta=1/T$ where the density matrix is given by a Hermitian matrix $\rho_{eq}=\exp[- \beta H]$ which leads to a vanishing particle current component, while ${\cal I}_{\eta,Y}^{st}$ take finite values in the nonequilibrium stationary state. Substituting $F_{0}^{(0)}$ into Eqs.(\[CurrentComp\]), the current component is expressed by $$\begin{aligned} \label{Jrep1} {\cal I}_{\eta,Y}^{st}&=&-i\<\!\<\hat x|\eta,Y\>\!\>\Delta_{\eta,Y}\big(\<\!\<\eta,Y|-\<\!\<-\eta,Y| \big) \nonumber\\ &\times& g^{2}\big|{\rm Tr}_{L\otimes R}\big[ \hat{\cal C}_{2}^{(0)}(+i0)\varphi_{0}^{(0)}\rho_{ph}^{eq}\big] \>\!\> \textrm{, for }\eta>0 \;.\end{aligned}$$ It should be noted that the contribution to the current component ${\cal I}_{\eta,Y}^{st}$ is attributed to the second order creation operator. Inserting Eqs.(\[ColInv\]) and (\[C1C2\]) into Eq.(\[Jrep1\]), the explicit expression of ${\cal I}_{\eta,Y}^{st}$ is obtained as $$\begin{aligned} \label{Icomp1} &&{\cal I}_{\eta,Y}^{st}=\frac{2\pi }{\hbar^2}\<\!\<\hat x|\eta,Y\>\!\> \sum_{r=L,R} g^2 v_r^2 c_{r,Y+{\eta\over 2}}^{*}c_{r,Y-{\eta\over2}} \nonumber\\ && \times\sum_{\xi>0}\bigg\{ {\cal D}_{ph}(\Delta_{\xi,Y-\eta-\xi}) |c_{r,Y-{\eta\over2}-\xi}|^2 \big\{\phi_{Y-{\eta\over2}}\big(n_r(\Delta_{\xi,Y-\eta-\xi})+1\big)-\phi_{Y-{\eta\over2}-\xi}n_r(\Delta_{\xi,Y-\eta-\xi})\big\} \nonumber\\ && \qquad \quad+ {\cal D}_{ph}(\Delta_{\xi,Y-\eta+\xi}) |c_{r,Y-{\eta\over2}+\xi}|^2 \big\{\phi_{Y-{\eta\over2}}n_r(\Delta_{\xi,Y-\eta+\xi})-\phi_{Y-{\eta\over2}+\xi}\big(n_r(\Delta_{\xi,Y-\eta+\xi}) +1\big)\big\} \nonumber\\ && \qquad \quad+ {\cal D}_{ph}(\Delta_{\xi,Y+\eta-\xi}) |c_{r,Y+{\eta\over2}-\xi}|^2 \big\{\phi_{Y+{\eta\over2}}\big(n_r(\Delta_{\xi,Y+\eta-\xi})+1\big)-\phi_{Y+{\eta\over2}-\xi}n_r(\Delta_{\xi,Y+\eta-\xi})\big\} \nonumber\\ && \qquad \quad+ {\cal D}_{ph}(\Delta_{\xi,Y+\eta+\xi}) |c_{r,Y+{\eta\over2}+\xi}|^2 \big\{\phi_{Y+{\eta\over2}}n_r(\Delta_{\xi,Y+\eta+\xi})-\phi_{Y+{\eta\over2}+\xi}\big(n_r(\Delta_{\xi,Y+\eta+\xi}) +1\big)\big\} \;, \end{aligned}$$ or by using the transition probabilities of $k_{r}^{Y,Y+\eta}$ given by Eqs.(\[App:tranProb1\]), ${\cal I}_{\eta,Y}^{st}$ can be written as $$\begin{aligned} \label{Icomp2} {\cal I}_{\eta,Y}^{st}= \big\langle Y-{\eta\over2}\big|\hat x\big|Y+{\eta\over2}\big\rangle \sum_{r=L,R} \sum_{\xi\neq 0}&&\bigg\{{ c_{r,Y+{\eta\over2}}\over c_{r,Y-{\eta\over2}}} \big( k_r^{Y+{\eta\over2}+\xi,Y-{\eta\over2}}\phi_{Y-{\eta\over2}} - k_r^{Y-{\eta\over2},Y+{\eta\over2}+\xi}\phi_{Y+{\eta\over2}+\xi} \big) \nonumber\\ &&+ { c_{r,Y-{\eta\over2}}\over c_{r,Y+{\eta\over2}}} \big( k_r^{Y+{\eta\over2}+\xi,Y+{\eta\over2}}\phi_{Y+{\eta\over2}} - k_r^{Y+{\eta\over2},Y+{\eta\over2}+\xi}\phi_{Y+{\eta\over2}+\xi} \big) \bigg\} \;. \end{aligned}$$ As seen in Eq.(\[Icomp2\]), in the calculation of the current component ${\cal I}_{\eta,Y}^{st}$ due to the pair of the molecular states between $|Y+\eta/2\>$ and $|Y-\eta/2\>$ we have to consider the transitions between the other molecular state $|Y+\eta'\>$ and $|Y\>$ expressed by $k_r^{Y+\eta',Y}$ or $k_r^{Y,Y+\eta'}$, and between $|Y+\eta'\>$ and $|Y+\eta\>$ expressed by $k_r^{Y+\eta',Y+\eta}$ or $k_r^{Y+\eta,Y+\eta'}$. As a result, the particle current cannot be cast into the Landauer formula contrary to the energy flow, where there is no other molecular states involved in the calculation of the energy flow ${\cal U}_{Y+\eta,Y}^{st}$. ![The particle current components ${\cal I}_{\eta(>0),Y}^{st}$ for ${\cal N}=10$, $v_{L}=v_{R}=1.0$, $g=0.1$, $T_{L}=2.0$, and $T_{R}=0.01$. In (a), all the components of ${\cal I}_{\eta,Y}^{st}$ are drawn, and in (b) and (c) the components of $\eta=1$ and $\eta=3$ are shown, respectively. []{data-label="Fig:Jmat1"}](PtclCrnt10.pdf){width="8cm"} ![The temperature dependence of some current components shown in Fig.\[Fig:Jmat1\]. The current components as a function of $T_{R}$ for a fixed value of $T_{L}=2.0$ are shown. The ${\cal I}_{\eta,Y}$ for $(\eta,Y)=(1,5/2), (1,9/2), (1,13/2), (1,17/2)$ and $(\eta,Y)=(3,7/2), (3,9/2), (3,13/2), (3,17/2)$ are shown in (a) and (b), respectively. []{data-label="Fig:IcomTR"}](IcomTR.pdf){width="8cm"} As an example of the particle current we show in Fig.\[Fig:Jmat1\] the particle current components ${\cal I}_{\eta,Y}^{st}$ of the same molecule which has been investigated in Section \[Sec:EFlow\], where we have taken ${\cal N}=10$, $v_{L}=v_{R}=1.0$, $g=0.1$, $T_{L}=2.0$, and $T_{R}=0.01$. In (a), all the components of ${\cal I}_{\eta,Y}^{st}$ are drawn, and in (b) and (c) the components for fixed values of $\eta=1$ and $\eta=3$ are shown, respectively, where the horizontal axis is $Y$. When $\varepsilon_{L}=\varepsilon_{R}$, ${\cal I}_{\eta,Y}^{st}=0$ for an even integer of $\eta$, because $\<\!\<\eta,Y|\hat x\>\!\>=0$ due to the symmetry. It is found that ${\cal I}_{\eta,Y}^{st}$ takes a large value when $\eta=1$ which suggests that the particle current is large for a pair of adjacent molecular states in energy, and the ${\cal I}_{\eta,Y}^{st}$ becomes small as $\eta$ increases. It is also found that ${\cal I}_{\eta,Y}^{st}$ is positive for a large $Y$ while it is negative for a small $Y$: The induced polarization due to the molecular states with a higher energy is directed from high temperature side to the low temperature side. We also show the temperature dependence of ${\cal I}_{\eta,Y}^{st}$ of Fig.\[Fig:IcomTR\] as a function of $T_{R}$ with a fixed value of $T_{L}=2.0$. In (a) are shown ${\cal I}_{\eta,Y}^{st}$ for $(\eta,Y)=(1,5/2), (1,9/2), (1,13/2), (1,17/2)$, and in (b) for $(\eta,Y)=(3,7/2), (3,9/2), (3,13/2), (3,17/2)$. The particle current vanishes at $T_{R}=T_{L}$, and they linearly depends on the temperature difference around $T_{R}=T_{L}$, while they change nonlinearly in the low temperature region reflecting the discrete molecular level structures. This behavior corresponds to that of the energy flow shown in Fig.\[Fig:EFlow\], because both the energy flow and the particle current are born out from the unique collision invariant by the actions of $\hat{\cal C}_{1}^{(0)}$ and $\hat{\cal C}_{2}^{(0)}$ on $|u_{0}^{(0)})\!)$. ![Space correlation in the molecule. All the parameters are the same as in Fig.\[Fig:Jmat1\]; (a) All the components of $f_{m,M}$, and (b) the dependence on the relative distance $m$ for $M=5$ and $M=5.5$. []{data-label="Fig:SpaceCor"}](SpaceCor.pdf){width="8cm"} So far we have investigated the particle current components attributed to a quantum correlation between a pair of molecular states. With use of the representation in terms of the site basis given in Eq.(\[Barj\]), we may reveal the quantum correlation in space which is generated in the nonequilibrium stationary state. Similarly to Eq.(\[ftoffstate\]), we shall investigate the non-symmetricity of the off-diagonal elements of $f(t)$ in terms of the site basis as $$\begin{aligned} \label{ftm1m2} f_{m,M}\equiv\big(\<\!\<m,M|-\<\!\<-m,M|\big) |f(t)\>\!\> \;,\end{aligned}$$ where $|m,M\>\!\>$ is the Wigner basis in terms of the site basis defined similarly to Eq.(\[WignerPtcl\]) by $$\begin{aligned} &&|\pm m,M\>\!\>\equiv |M\pm{m\over 2};M\mp{m\over 2}\>\equiv |M\pm{m\over 2}\>\<M\mp{m\over 2}| \;, \nonumber\\ &&\hspace{3cm} \textrm{(double sign in same order)} \end{aligned}$$ with use of site basis $|M\pm{m/ 2}\>$. We show in Fig.\[Fig:SpaceCor\] the imaginary part of $f_{m,M}$ for the same case of Fig.\[Fig:Jmat1\]: In (a) we show the all the components of $f_{m,M}$ and in (b) the dependence on the relative distance $m$ for $M=5$ and $M=5.5$. As seen from the figures, the values of $f_{m,M}$ does not depend on the central position of $M$ but on the relative distance of $m$, suggesting that the quantum correlation in space decreases, as $m$ increases. We also found that $f_{m,M}=0$ at $T_{L}=T_{R}$: The quantum correlation in space comes to appear only at nonequilibrium situation. Application to one-dimensional DNA chain {#Sec:Discussion} ======================================== As an application of our result, let us give an example of a real physical system, a hole transfer in one-dimensional chain of DNA base pairs.[@Berlin02] In this case, a hole (or an electron) is an energy carrier and the molecular system has been described by the tight binding Hamiltonian given by Eq.(\[HM\]). It has been known that in a chain of DNA bases, the site energy of a guanine (G) and cytosine (C) base pair is higher than that of a thymine (T) and adenine (A) pair: $\varepsilon_{GC}-\varepsilon_{TA}\simeq 0.5$eV, and the transfer between the base pairs have been estimated to be $J\simeq 0.4$eV. We show in Fig.\[Fig:DNA1\](a) a typical result of the energy flow when we put two GC pairs at the both ends of a molecular chain with its length of ${\cal N}=10$, where we have taken $g=0.5$. It is found that the energy flow increases nonlinearly with temperature; Note that $T_{L,R}=300$K corresponds to $T_{L,R}=0.06$ in our unit. The hole current in the molecular states is also shown in Fig.\[Fig:DNA1\](b) and the schematic picture of the energy flow and the particle current is shown in Fig.\[Fig:DNA1\](c), where we fix $T_{L}=0.07$ and $T_{R}=0.06$ corresponding to 350K and 300K, respectively. It is found that the energy is transported by a particle current in the lower molecular states. ![The energy flow (a) and the particle current (b) of ome-dimensional base pairs of DNA molecule with 10 bases. In (b) $T_{L}=0.07$ (350K) and $T_{R}=0.06$ (300K) are used. []{data-label="Fig:DNA1"}](DNA2_TP.pdf){width="8cm"} concluding remarks {#Sec:Remarks} ================== We have obtained the nonequilibrium stationary state under thermal force as the zero eigenstate of the Liouvillian of a molecular chain which is weakly coupled with different thermal baths at the both ends. The zero eigenstate is represented in the expansion of the order of correlation following the principle of the dynamics of correlation. The physical quantity in the nonequilibrium stationary state is derived by taking an expectation value of an observable with respect to the stationary state. The energy flow and the particle current are attributed to the first order and second order correlations created from the vacuum of correlation, respectively. Since the first order correlation is directly related to the collision operator in kinetic theory, the energy flow can be cast into the form of Landauer equation with a transmission function with a strong resonance which reflects a discrete level structure of the molecular states. It is also found that the thermal force induces an polarization, or its conjugate particle current, which cannot be cast into the Landauer type formula, because the particle current is attributed to the second order correlation. Even so, there is a correspondence between them in their temperature dependences. Here we have dealt with the nonequilibrium transport process of a molecule coupling with a thermal phonon bath with very broad spectral width. It is interesting to investigate how the energy flow will be changed if we modify the phonon density of states in such a way that a particular molecular states be resonantly excited. By replacing a phonon field with a radiation field in the present work, we can investigate how the energy flow behaves under a monochromatic light excitation in a nonequilibrium stationary state. These further extensions of the present model will be shown elsewhere. The authors thank Profs. N. Hatano and H. Hayakawa, and Mr. R. Tatsumi for fruitful discussions. This work was supported by the Grant-in-Aid for Scientific Research from the Ministry of Education, Science, Sports, and Culture of Japan and partially supported by Yukawa International Program for Quark-Hadron Sciences YIPQS. Representation in Liouville space and the expression of $\mathcal{L}_{MB}$ {#App:LiouvilleSpace} ========================================================================== In this section, we shall briefly review the Liouville space representation of a Hilbert space operator. The Liouville space is spanned by linear operators in $A, B, \cdots$ in the ordinary wave function space.[@Petrosky97] As usual, the inner product of the Liouville space is defined by $$\label{App:InPrdct} \<\!\< A|B\>\!\> = {\mathop\mathrm{Tr}\nolimits}(A^{\dagger}B) \;,$$ where $A$ and $B$ are linear operators acting on wave functions, and $A^{\dagger}$ is a Hermite conjugate of $A$. For the case where the wave function space is spanned by a complete orthonormal basis, $$\label{App:Ortho} \sum_{\alpha}|\alpha\>\<\alpha |=1\;, \<\alpha|\beta\> =\delta_{\alpha,\beta} \;,$$ the Liouville space is spanned by a complete orthonormal basis of the dyads $ |\alpha ; \beta \>\!\>\equiv |\alpha \>\<\beta| \;, $ i.e., $$\sum_{\alpha,\beta} |\alpha ; \beta \>\!\> \<\!\< \alpha ; \beta | =1 \; , \; \<\!\<\alpha ; \beta |\alpha' ; \beta' \>\!\>=\delta_{\alpha,\alpha'}\delta_{\beta,\beta'} \;.$$ The matrix element of the usual operator $A$ in the wave function space is given by $$\<\!\<\alpha ; \beta |A\>\!\>=\<\alpha|A|\beta\> \;.$$ The Liouville basis is constructed of the tensor product of the eigenstates of the unperturbed Hamiltonian. For the molecular system and the thermal bath systems, the Liouville basis are written by $$|\bar j_{1};\bar j_{2}\>\!\>\equiv |\bar j_{1}\>\<\bar j_{2}| \;,$$ where $|\bar j\>$ denotes the eigenstate of $H_{M}$ given by Eq.(\[Barj\]), and $$|n_{r,{\bf q}};n'_{r,{\bf q}}\>\!\>\equiv |n_{r,{\bf q}}\>\<n'_{r,{\bf q}}| \;,$$ where $|n_{r,{\bf q}}\>$ is the number state for a thermal bath normal mode $(r,{\bf q})$ of $H_{B}$. In order to clarify the order of correlation, we prefer to use the Wigner basis defined by Eqs.(\[WignerPtcl\]) and (\[WignerBath\]): $$\label{App:WignerPtcl} | \eta,Y \>\!\> \equiv |\bar j_{1};\bar j_{2}\>\!\> \; ,$$ where $$\eta\equiv \bar j_{1}-\bar j_{2} \;,\; Y\equiv {\bar j_{1}+\bar j_{2}\over 2} \;,$$ and $$\label{App:WignerBath} |\nu_{r,{\bf q}}, N_{r,{\bf q}}\>\!\>\equiv |n_{r,{\bf q}};n'_{r,{\bf q}} \>\!\> \;,$$ where $\nu_{r,{\bf q}}$ and $N_{r,{\bf q}}$ are defined in Eq.(\[WignerBath\]). The Wigner basis is the eigenstate of ${\cal L}_{0}$ as shown in Eq.(\[L0EV\]). The Wigner basis then form the complete orthonormal basis satisfying $$\begin{aligned} \label{App:WignerComp} &&\<\!\< \eta,Y|\eta',Y'\>\!\>\!\otimes\<\!\<\{\nu\},\{N\}|\{\nu'\},\{N'\}\>\!\>=\delta_{\eta,\eta'}\delta_{Y,Y'} \nonumber\\ && \qquad\qquad \cdot\delta_{\{\nu\},\{\nu'\}}\delta_{\{N\},\{N'\}} \;,\\ &&\sum_{\eta,Y}| \eta,Y\>\!\>\<\!\<\eta,Y|\sum_{\{\nu\},\{N\}}|\{\nu\},\{N\}\>\!\>\<\!\<\{\nu\};\{N\}|=1 \;, \nonumber\\\end{aligned}$$ where $|\{\nu\},\{N\}\>\!\>\equiv \prod_{r,{\bf q}}|\nu_{r,{\bf q}},N_{r,{\bf q}}\>\!\>$. Now we consider the matrix element of the interaction Liouvillian ${\cal L}_{MB}\equiv H_{MB}\times 1 -1\times H_{MB}$ in terms of these Wigner basis. The calculation of the matrix elements of ${\cal L}_{MB}$ can be done in a straightforward manner, yielding \[App:LMBcomp\] $$\begin{aligned} &&\<\!\<\eta,Y|\<\!\<\{\nu\};\{N\}|g{\cal L}_{MB}|\eta',Y'\>\!\>|\{\nu'\},\{N'\}\>\!\> ={g\over\hbar\sqrt{\Omega}}\sum_{r=L,R}\sum_{\bf q} v_{r}\nonumber \\ &&\times\bigg[ \<Y+{\eta\over 2}|r\>\<r|Y'+{\eta'\over 2}\>\delta_{Y'-{\eta'\over2},Y-{\eta\over2}}\sqrt{N_{r,{\bf q}}+{\nu_{r,{\bf q}}\over2}+1} \;e^{{1\over2}{d\over dN_{r,{\bf q}}}}\delta_{\nu'_{r,{\bf q}},\nu_{r,{\bf q}}+1} \delta'_{\{\nu\},\{\nu'\}} \delta_{\{N\},\{N'\}} \\ && - \<Y'-{\eta'\over 2}|r\>\<r|Y-{\eta\over 2}\>\delta_{Y'+{\eta'\over2},Y+{\eta\over2}}\sqrt{N_{r,{\bf q}}-{\nu_{r,{\bf q}}\over2}} \;e^{-{1\over2}{d\over dN_{r,{\bf q}}}}\delta_{\nu'_{r,{\bf q}},\nu_{r,{\bf q}}+1} \delta'_{\{\nu\},\{\nu'\}} \delta_{\{N\},\{N'\}} \\ && + \<Y'+{\eta'\over 2}|r\>\<r|Y+{\eta\over 2}\>\delta_{Y'-{\eta'\over2},Y-{\eta\over2}}\sqrt{N_{r,{\bf q}}+{\nu_{r,{\bf q}}\over2}} \;e^{-{1\over2}{d\over dN_{r,{\bf q}}}}\delta_{\nu'_{r,{\bf q}},\nu_{r,{\bf q}}-1} \delta'_{\{\nu\},\{\nu'\}} \delta_{\{N\},\{N'\}} \\ && - \<Y-{\eta\over 2}|r\>\<r|Y'-{\eta'\over 2}\>\delta_{Y'+{\eta'\over2},Y+{\eta\over2}}\sqrt{N_{r,{\bf q}}-{\nu_{r,{\bf q}}\over2}+1} \;e^{{1\over2}{d\over dN_{r,{\bf q}}}}\delta_{\nu'_{r,{\bf q}},\nu_{r,{\bf q}}-1} \delta'_{\{\nu\},\{\nu'\}} \delta_{\{N\},\{N'\}} \bigg] \;,\end{aligned}$$ where $\delta'$ stands for the product of the Kronecker delta except for the interaction normal mode of $(r,{\bf q})$. We draw the diagram of the correlation corresponding to these four terms in Fig.\[Fig:LMBdiag\]. In the figures, the dotted line denotes the correlation, $\nu_{r,{\bf q}}$, of the phonon mode involved in the interaction, and the solid or double solid lines denote the correlation, $\eta$ or $\eta'$, of the molecular state. The filled circle stands for the vertex of the interaction whose matrix element is written at the vertex. The diagrams of (a) and (c) correspond to Eqs.(\[App:LMBcomp\]a) and (\[App:LMBcomp\]c), each of which is attributed to the $b_{r,{\bf q}}$ (phonon absorption) and $b_{r,{\bf q}}^{\dagger}$ (phonon emission) terms in $H_{MB}\times 1$ term of ${\cal L}_{MB}$, where the phonon line appearing on the left side of the vertex implies that the transition of the particle ket state of $|Y'+\eta'/2\>$ to $|Y+\eta/2\>$ occurs by the interaction. On the other hand, the diagrams of (b) and (d) correspond to Eqs.(\[App:LMBcomp\]b) and (\[App:LMBcomp\]d), each of which is attributed to the $b_{r,{\bf q}}^{\dagger}$ (phonon emission) and $b_{r,{\bf q}}$ (phonon absorption) terms in $1\times H_{MB}$ term of ${\cal L}_{MB}$, where the phonon line appearing on the right side of the vertex implies that the transition of the particle bra state of $\<Y'-\eta'/2|$ to $\<Y-\eta/2|$ occurs by the interaction. The reader may be referred to the textbook about the way of writing the correlation lines.[@PrigogineText] ![The diagram of the interaction Liouvillian $L_{MB}$, where the lines represent the correlation.[@PrigogineText; @Petrosky97] The dotted line denotes the correlation, $\nu_{r,{\bf q}}$, of the phonon mode involved in the interaction, and the solid or double solid lines denote the correlation, $\eta$ or $\eta'$, of the molecular state. The filled circle stands for the vertex of the interaction whose matrix element is written at the vertex. []{data-label="Fig:LMBdiag"}](LMBdiag2.pdf){width="12cm"} The Complex Spectral Representation of Liouvillean and Subdynamics {#App:Complex} =================================================================== In this section, we shall summarize the complex spectral representation of Liouvillian.[@Petrosky97] Useful formula for this paper are listed without proof. The reader may refer to some references for detail.[@Petrosky97; @Petrosky02; @Petrosky10] In the complex spectral representation of Liouvillian, we consider the eigenvalue problem for each correlation subspace $(\mu)=(\eta,\nu)$, where $\eta$ and $\nu$ represent the order of the correlation of the particle and phonon, respectively. The complex eigenvalue problem is written as $$\begin{aligned} \label{App:EigenEq} \mathcal{L}|F_{j}^{(\mu)} )\!) = Z_{j}^{(\mu)}|F_{j}^{(\mu)} )\!) \;, \; (\!( {\tilde F}_{j}^{(\mu)}| \mathcal{L} =(\!( {\tilde F}_{j}^{(\mu)}| Z_{j}^{(\mu)} \;,\end{aligned}$$ where the Liouvillian can have complex eigenvalues $\mathrm{Im} Z_{j}^{(\mu)}\neq 0$. It has been shown that the time evolution splits into two semigroups; one is oriented toward our future $t>0$ with $\mathrm{Im} Z_{j}^{(\mu)}<0$ (equilibrium is approached for $t\rightarrow \infty$), while the other is oriented toward our past $t<0$ with $\mathrm{Im} Z_{j}^{(\mu)}>0$. All irreversible processes have the same time orientation. To be self-consistent we choose the semigroup oriented toward our future, which determines the direction of the analytic continuation of the eigenfunction of $\mathcal{L}$.[@Petrosky96; @Petrosky97] Now we introduce the projection operators defined in (\[Pclass\]) which satisfy $$\begin{aligned} &&\mathcal{L}_{0}\hat{\cal P}^{(\mu)}=\hat{\cal P}^{(\mu)}\mathcal{L}_{0} \; , \\ && \hat{\cal P}^{(\mu)}\hat{\cal P}^{(\mu')}=\hat{\cal P}^{(\mu)}\delta_{\mu,\mu'} \;, \\ && \sum_{\mu}\hat{\cal P}^{(\mu)}=1 \;.\end{aligned}$$ We also introduce the projection operators $$\label{App:Qk} \hat{\cal Q}^{(\mu)}=1-\hat{\cal P}^{(\mu)}$$ which are orthogonal to $\hat{\cal P}^{(\mu)}$. We solve the eigenvalue problem (\[App:EigenEq\]) for the perturbed system with $g\neq 0$ under the boundary conditions for the unperturbed case: $$\begin{aligned} &&|F_{j}^{(\mu)} )\!) =\hat{\cal P}^{(\mu)} |F_{j}^{(\mu)} )\!) \;, \nonumber\\ &&(\!({\tilde F}_{j}^{(\mu)}| =(\!({\tilde F}_{j}^{(\mu)}| \hat{\cal P}^{(\mu)} \quad{\mathrm{for}}\; g=0 \;.\end{aligned}$$ Hence, $\hat{\cal Q}^{(\mu)}|F_{j}^{(\mu)} )\!)=0$ for $g=0$. The $\hat{\cal Q}^{(\mu)}$ components are created through the interaction for $g\neq 0$. The right and left eigenstates, $|F_{j}^{(\mu)} )\!)$ and $(\!({\tilde F}_{j}^{(\mu)}|$, are biorthonormal sets satisfying $$\label{App:BiOrthoF} (\!(\tilde{F}_{j}^{(\mu)}|F_{j'}^{(\mu')})\!) =\delta_{j,j'}\delta_{\mu,\mu'}\; , \sum_{\mu,j}|F_{j}^{(\mu)})\!)(\!(\tilde{F}_{j}^{(\mu)}|=1 \; .$$ Applying the projection operators $\hat{\cal P}^{(\mu)}$ and $\hat{\cal Q}^{(\mu)}$ in (\[Pclass\]) and (\[App:Qk\]) to the (\[App:EigenEq\]), we derive the set of equations: \[App:PQF\] $$\begin{aligned} &&\hat{\cal P}^{(\mu)}\mathcal{L}\Big(\hat{\cal P}^{(\mu)}|F_{j}^{(\mu)})\!)+\hat{\cal Q}^{(\mu)}|F_{j}^{(\mu)})\!) \Big)=Z_{j}^{(\mu)}\hat{P}^{(\mu)}|F_{j}^{(\mu)}{\>\!\>} \;, \nonumber\\ &&\\ &&\hat{\cal Q}^{(\mu)}\mathcal{L}\Big(\hat{\cal P}^{(\mu)}|F_{j}^{(\mu)})\!)+\hat{\cal Q}^{(\mu)}|F_{j}^{(\mu)})\!) \Big)=Z_{j}^{(\mu)}\hat{\cal Q}^{(\mu)}|F_{j}^{(\mu)})\!) \;.\nonumber\\\end{aligned}$$ Equation (\[App:PQF\]b) leads to $$\label{QF} \hat{\cal Q}^{(\mu)}|F_j^{(\mu)} )\!) =\hat{\mathcal{C}}^{(\mu)}(Z_j^{(\mu)}) \hat{\cal P}^{(\mu)} | F_j^{(\mu)})\!) \;,$$ where $$ \hat{\mathcal{C}}^{(\mu)}(z)=\frac{1}{z-\hat{\cal Q}^{(\mu)}\mathcal{L}\hat{\cal Q}^{(\mu)}}\hat{\cal Q}^{(\mu)}g\mathcal{L}_{MB}\hat{\cal P}^{(\mu)}$$ is called the creation-of-correlation operator, or simply the [*creation operator*]{}.[@Petrosky97] Substituting (\[QF\]) into (\[App:PQF\]a), we obtain $$\label{App:EVPCol2} \hat{\Psi}^{(\mu)} (Z_j^{(\mu)})|u_j^{(\mu)})\!) = Z_j^{(\mu)} | u_j^{(\mu)} )\!) \; ,$$ where $$\label{App:uj2} |u_j^{(\mu)} )\!) = (N_j^{(\mu)})^{-1/2} \hat{\cal P}^{(\mu)} |F_j^{(\mu)} )\!)$$ and $N_j^{(\mu)}$ is a normalization constant which will be determined later. Here, $\hat{\Psi}^{(\mu)}$ is the [*collision operator*]{} familiar to nonequilibrium statistical mechanics.[@PrigogineText; @Resibois67; @ResiboisBook; @BalescuBook] This operator is associated to [*diagonal transitions*]{} between two states corresponding to the same projection operator $\hat{P}^{(\mu)}$: $$\begin{aligned} \hat{\Psi}^{(\mu)}(z)&=&{\hat{\cal P}}^{(\mu)} \mathcal{L}_0 \hat{\cal P}^{(\mu)} +\hat{\cal P}^{(\mu)} g\mathcal{L}_{MB}\hat{\cal P}^{(\mu)} \nonumber\\ &+&\hat{\cal P}^{(\mu)}g\mathcal{L}_{MB}\hat{\cal Q}^{(\mu)}\hat{\mathcal{C}}^{(\mu)}(z) \hat{\cal P}^{(\mu)} \; .\end{aligned}$$ Note that (\[App:EVPCol2\]) is a nonlinear equation in the same sense of the Brillouin-Wigner perturbation method, i.e., the eigenvalue $Z_{j}^{(\mu)}$ appears in the collision operator. Assuming completeness in the space $\hat{\cal P}^{(\mu)}$, we may always construct a set of states $\{(\!(\tilde{u}_{j}^{(\mu)}|\}$ biorthogonal to $\{|{u}_{j}^{(\mu)})\!)\}$, i.e., $$\label{App:CompUj} (\!(\tilde{u}_{j}^{(\mu)}|u_{j'}^{(\mu')})\!) =\delta_{j,j'}\delta_{\mu,\mu'}\; , \sum_{\mu,j}|u_{j}^{(\mu)})\!)(\!(\tilde{u}_{j}^{(\mu)}|=1 \; .$$ The equation (\[App:EVPCol2\]) combined with (\[App:uj2\]) shows that the $\hat{\cal P}^{(\mu)}$ component of $|F_{j}^{(\mu)})\!)$ (which is called “[*privileged component*]{}” of $|F_{j}^{(\mu)})\!)$) is an eigenstates of the collision operator, which has the same eigenvalue $Z_{j}^{(\mu)}$ as the Liouvillean. The solution of the eigenvalue problem of the Liouvillean for our class of singular functions has unique features. The privileged components satisfy closed equations and the $\hat{\cal Q}^{(\mu)}$ components are “driven” by the privileged components \[See (\[QF\])\]. Combining (\[QF\]) with (\[App:uj\]), we obtain the right eigenstates of the Liouvillean given by $$\label{App:Fright2} |F_{j}^{(\mu)})\!)=\sqrt{N_{j}^{(\mu)}} \Big(\hat{\cal P}^{(\mu)}+\hat{\mathcal{C}}^{(\mu)}(Z_{j}^{(\mu)}) \Big) |u_{j}^{(\mu)})\!) \; .$$ Similarly, we obtain for the left eigenstates given by $$\label{App:Fleft} (\!( \tilde{F}_{j}^{(\mu)}|=(\!( \tilde{v}_{j}^{(\mu)}| \Big(\hat{\cal P}^{(\mu)}+\hat{\mathcal{D}}^{(\mu)}(Z_{j}^{(\mu)}) \Big)\sqrt{N_{j}^{(\mu)}} \; ,$$ where the operator $\hat{\mathcal{D}}^{(\mu)}(Z_{j}^{(\mu)})$ is called the destruction-of-correlation operator, or the [*destruction operator*]{} for short, and is defined by \[cf. (\[App:Creation\])\] $$\label{App:Destruction} \hat{\mathcal{D}}^{(\mu)}(z)=\hat{\cal P}^{(\mu)}g\mathcal{L}_{MB}\hat{\cal Q}^{(\mu)}\frac{1}{z-\hat{\cal Q}^{(\mu)}\mathcal{L}\hat{\cal Q}^{(\mu)}}\hat{\cal Q}^{(\mu)} \;.$$ Again $ \hat{\mathcal{D}}^{(\mu)}(z)$ corresponds to the off-diagonal transitions; $ \hat{\mathcal{D}}^{(\mu)}(z)=\hat{\cal P}^{(\mu)} \hat{\mathcal{D}}^{(\mu)}(z)\hat{\cal Q}^{(\mu)}$. Using $\hat{\mathcal{D}}^{(\mu)}(z)$, the collision operator $\hat{\Psi}^{(\mu)}(z)$ is also written as $$\begin{aligned} \label{App:PsiUsingD} \hat{\Psi}^{(\mu)}(z)&=&{\hat{\cal P}}^{(\mu)} \mathcal{L}_0 \hat{\cal P}^{(\mu)} +\hat{\cal P}^{(\mu)} g\mathcal{L}_{MB}\hat{\cal P}^{(\mu)} \nonumber\\ &+&\hat{\cal P}^{(\mu)}\hat{\mathcal{D}}^{(\mu)}(z) \hat{\cal Q}^{(\mu)} g\mathcal{L}_{MB} \hat{\cal P}^{(\mu)} \; .\end{aligned}$$ The normalization constant $N_{j}^{(\mu)}$ is determined by Eq.(\[App:BiOrthoF\])with Eqs.(\[App:Fright2\]) and (\[App:Fleft\]) as $$\label{App:Nnorm} \left( N_{j}^{(\mu)}\right)^{-1}=(\!(\tilde v_{j}^{(\mu)}|[\hat{\cal P}^{(\mu)}+\hat{\cal D}^{(\mu)}(Z_{j}^{(\mu)})\hat{\cal C}^{(\mu)}(Z_{j}^{(\mu)})]|u_j^{(\mu)})\!)\;.$$ The states $(\!(\tilde{v}_{j}^{(\mu)}|$ are the left eigenstates of the collision operator $\hat{\Psi}^{(\mu)}$, $$\label{App:EVPColLeft} (\!( \tilde{v}_{j}^{(\mu)}| \hat{\Psi}^{(\mu)}(Z_{j}^{(\mu)})=(\!( \tilde{v}_{j}^{(\mu)}| Z_{j}^{(\mu)} \;.$$ We have revealed the correspondence between the eigenvalue problems of the Liouvillean $\mathcal{L}$ and the collision operator $\hat{\Psi}^{(\mu)}$. Collision operator in the vacuum of correlation subspace {#App:DerivKinEq} ======================================================== In this section, we shall give the explicit expression of the collision operator of $\bar{\Psi}^{(0)}_{2}$ by applying Eqs.(\[App:LMBcomp\]) into Eq.(\[Psi2Expression\]). We show the diagrams of $\bar\Psi_{2}^{(0)}(+i0)$ in Fig.\[Fig:CollisionDiag\]. The diagrams (a), (b), (c), and (d) represent the loss of the state $|0,Y\>\!\>$ which are written by $$\begin{aligned} \label{App:Collision term loss} &(a):&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y\>\!\>_{(a)}={g^2\over \hbar^2} {1\over \Omega}\sum_q \sum_{r=L,R}\sum_{\eta} |v_r|^2|c_{r,Y}^* c_{r,Y-\eta}|^2 {1\over i 0^+ - (\Delta_{-\eta,Y-\eta/2}+\omega_{r,q}) } \left(n_{r}(\omega_{r,q})+1\right) \;,\\ &(a'):&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y\>\!\>_{(a')}={g^2\over \hbar^2} {1\over \Omega}\sum_q \sum_{r=L,R} \sum_{\eta} |v_r|^2|c_{r,Y}^* c_{r,Y+\eta}|^2 {1\over i 0^+ - (\Delta_{\eta,Y+\eta/2}-\omega_{r,q}) } n_{r} (\omega_{r,q}) \;,\\ &(b):&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y\>\!\>_{(b)}={g^2\over \hbar^2} {1\over \Omega}\sum_q \sum_{r=L,R}\sum_{\eta} |v_r|^2|c_{r,Y}^* c_{r,Y+\eta}|^2 {1\over i 0^+ - (\Delta_{-\eta,Y+\eta/2}+\omega_{r,q}) } n_{r} (\omega_{r,q}) \;,\\ &(b'):&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y\>\!\>_{(b')}={g^2\over \hbar^2} {1\over \Omega}\sum_q \sum_{r=L,R}\sum_{\eta} |v_r|^2|c_{r,Y}^* c_{r,Y-\eta}|^2 {1\over i 0^+ - (\Delta_{\eta,Y-\eta/2}-\omega_{r,q}) }\left( n_{r} (\omega_{r,q})+1 \right) \;.\end{aligned}$$ where $n_{r}(\omega)$ is Planck’s distribution function given by Eq.(\[Planck\]). ![ Diagram of the collision operator $\overline{\Psi}_{2}^{(k)}(+i0)$. []{data-label="Fig:CollisionDiag"}](CollisionDiag2.pdf){width="12cm" height="10cm"} On the other hand, the diagrams (c), (c’), (d), and (d’) represent the gain of the $|0,Y\>\!\>$ state from the $|0,Y\pm \eta\>\!\>$ states which are written as $$\begin{aligned} \label{App:Collision term gain} &(c):&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y+\eta\>\!\>_{(c)}=-{g^2\over \hbar^2} {1\over \Omega}\sum_q \sum_{r=L,R} |v_r|^2|c_{r,Y}^* c_{r,Y+\eta}|^2 {1\over i 0^+ - (\Delta_{-\eta,Y+\eta/2}+\omega_{r,q}) } \left(n_{r}(\omega_{r,q})+1\right) \;,\\ &(c'):&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y-\eta\>\!\>_{(c')}=-{g^2\over \hbar^2} {1\over \Omega}\sum_q \sum_{r=L,R} |v_r|^2|c_{r,Y}^* c_{r,Y-\eta}|^2 {1\over i 0^+ - (\Delta_{\eta,Y-\eta/2}-\omega_{r,q}) } n_{r}(\omega_{r,q}) \;,\\ &(d):&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y-\eta\>\!\>_{(d)}=-{g^2\over \hbar^2} {1\over \Omega}\sum_q \sum_{r=L,R} |v_r|^2|c_{r,Y}^* c_{r,Y-\eta}|^2 {1\over i 0^+ - (\Delta_{-\eta,Y-\eta/2}+\omega_{r,q}) } n_{r}(\omega_{r,q}) \;,\\ &(d'):&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y+\eta\>\!\>_{(d')}=-{g^2\over \hbar^2} {1\over \Omega}\sum_q \sum_{r=L,R} |v_r|^2|c_{r,Y}^* c_{r,Y+\eta}|^2 {1\over i 0^+ - (\Delta_{\eta,Y+\eta/2}-\omega_{r,q}) }\left( n_{r}(\omega_{r,q}) +1 \right) \;, \end{aligned}$$ where $n_{r}(\omega)$ is the Planck’s distribution given by Eq.(\[Planck\]). The summation for the thermal phonon modes $q$ is replaced by the integral with use of the density of states given by Eq.(\[phonon DOS\]). When we sum up the contributions of these diagrams, we have for $\eta>0$ $$\begin{aligned} \<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y\>\!\>&=&-2\pi i {g^2\over \hbar^2} \sum_{r=L,R}\sum_{\eta >0} |v_r|^2 \Big\{ |c_{r,Y}^* c_{r,Y-\eta}|^2 {\cal D}_{ph}(\Delta_{\eta,Y-\eta/2}) (n_{r}(\Delta_{\eta,Y-\eta/2})+1 ) \nonumber\\ && + |c_{r,Y}^* c_{r,Y+\eta}|^2 {\cal D}_{ph}( \Delta_{\eta,Y+\eta/2}) n_{r} ( \Delta_{\eta,Y+\eta/2}) \Big\} \;, \\ \<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y+\eta\>\!\>&=&2\pi i {g^2\over \hbar^2} \sum_{r=L,R} |v_r|^2|c_{r,Y}^* c_{r,Y+\eta}|^2 {\cal D}_{ph}(\Delta_{\eta,Y+\eta/2}) \left(n_{r}(\Delta_{\eta,Y+\eta/2})+1\right) \;,\\ \<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y-\eta\>\!\>&=&2\pi i{g^2\over \hbar^2} \sum_{r=L,R} |v_r|^2|c_{r,Y}^* c_{r,Y-\eta}|^2 {\cal D}_{ph}(\Delta_{\eta,Y-\eta/2}) n_{r}(\Delta_{\eta,Y-\eta/2}) \;,\end{aligned}$$ where we have used Eqs.(\[IntDOS\]) and (\[PropDelta\]). When we define the transition probabilities as $$\begin{aligned} \label{App:tranProb1} k_{r}^{Y,Y+\eta}&\equiv& 2\pi {g^2\over \hbar^2} |v_r|^2|c_{r,Y}^* c_{r,Y+\eta}|^2 {\cal D}_{ph}(\Delta_{\eta,Y+\eta/2}) \left(n_{r}(\Delta_{\eta,Y+\eta/2})+1\right) \;,\\ k_{r}^{Y,Y-\eta}&\equiv &2\pi {g^2\over \hbar^2} |v_r|^2|c_{r,Y}^* c_{r,Y-\eta}|^2 {\cal D}_{ph}(\Delta_{\eta,Y-\eta/2}) n_{r}(\Delta_{\eta,Y-\eta/2}) \;,\end{aligned}$$ where $\eta >0$, we can write \[App:tranProb2\] $$\begin{aligned} &&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y\>\!\>=-i\sum_{r=L,R}\sum_{\eta >0} \left(k_{r}^{Y-\eta,Y} +k_{r}^{Y+\eta,Y} \right) \;, \\ &&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y+\eta\>\!\>= i\sum_{r=L,R} k_{r}^{Y,Y+\eta} \;,\\ &&\<\!\<0,Y|\bar\Psi_2^{(0)}|0,Y-\eta\>\!\>= i\sum_{r=L,R} k_{r}^{Y,Y-\eta} \;.\end{aligned}$$ For a ${\cal N}=2$ molecule, the eigenstates of $H_{M}$ are obtained as \[App:EigSt2\] $$\begin{aligned} &&|\bar 1\>=c_{L,\bar 1}|L\>+c_{R,\bar 1}|R\> \;,\\ &&|\bar 2\>=c_{L,\bar 2}|L\>+c_{R,\bar 2}|R\> \;,\end{aligned}$$ with $$\begin{aligned} &&c_{L,\bar 1}=\cos \theta \;,\; c_{R,\bar 1}=\sin \theta \;,\\ &&c_{L,\bar 2}=-\sin \theta \;,\; c_{R,\bar 2}=\cos \theta \;,\end{aligned}$$ where $\tan 2\theta = 2J/(E_{R}-E_{L})$. The eigenvalues of these eigenstates are given by $$\begin{aligned} &&E_{\bar 1}={ \varepsilon_{L}+\varepsilon_{R}-\sqrt{(\varepsilon_{L}-\varepsilon_{R})^{2}+4J^{2}}\over 2} \;,\\ &&E_{\bar 2}={ \varepsilon_{L}+\varepsilon_{R}+\sqrt{(\varepsilon_{L}-\varepsilon_{R})^{2}+4J^{2}} \over 2}\;.\end{aligned}$$ The transition probabilities are then given by $$\begin{aligned} k_{L}^{1,2}&=& {\pi g^2 v_{L}^2 \over 2\hbar^2} \sin^2 2\theta \;{\cal D}_{ph}(\Delta) \left(n_{L}(\Delta)+1\right) \;,\\ k_{R}^{1,2}&=& {\pi g^2 v_{R}^2 \over 2\hbar^2} \sin^2 2\theta \;{\cal D}_{ph}(\Delta) \left(n_{R}(\Delta)+1\right) \;,\\ k_{L}^{2,1}&=& {\pi g^2 v_{L}^2 \over 2\hbar^2} \sin^2 2\theta \;{\cal D}_{ph}(\Delta) n_{L}(\Delta) \;,\\ k_{R}^{2,1}&=& {\pi g^2 v_{R}^2 \over 2\hbar^2} \sin^2 2\theta \;{\cal D}_{ph}(\Delta) n_{R}(\Delta) \;.\end{aligned}$$ Energy Flow in terms of a energy change of the thermal bath {#App:BathChange} =========================================================== In this section, we shall show that the formula of the energy flow Eq.(\[WstExp\]) is also obtained by investigating the energy change of the thermal bath systems. We can denote the energy flow from the left thermal bath as a energy change of the left thermal bath: $$\label{App:wLdef} w^L(t)={d\over dt}\<H_B^L\>_t=\sum_{\bf q}\omega_{L,{\bf q}}{\rm Tr}[\hat N_{L,{\bf q}}{\cal L}_{MB}\rho(t) ] \;,$$ where $\hat N_{L,{\bf q}}\equiv b_{L,{\bf q}}^\dagger b_{L,{\bf q}}$. This definition coincides with the ordinary definition of the particle current from a particle bath to a mesoscopic system.[@Comment] Using Eqs.(\[App:LMBcomp\]) and replacing $\rho(t)$ with $F_{0}^{(0)}$, we have the expression of the energy flow from the left bath in the nonequilibrium stationary state as $$\label{App:wLst} w_L^{st}={g v_L\over\sqrt{\Omega}}\sum_{\bf q}\sum_{\eta,Y}c_{L,Y+\eta}c_{L,Y}\rm{Tr}\big[(b_{L,{\bf q}}^\dagger-b_{L,{\bf q}})F_0^{(0)}\big] \;.$$ Substituting Eq.(\[F00-2\]) into Eq.(\[App:wLst\]) and after some calculation, we obtain $$\begin{aligned} w_L^{st}&=&\frac{2\pi g^2 v_L^2}{\hbar} \sum_{\eta>0}{\cal D}^{ph}(\Delta_{\eta,Y+\eta/2})\big|c_{L,Y+\eta}^{*}c_{L,Y}\big|^2 \Delta_{\eta,Y+\eta/2} \nonumber\\ &\times&\bigg\{ \phi_{Y}n_L(\Delta_{\eta,Y+\eta/2})-\phi_{Y+\eta} \big( n_L(\Delta_{\eta,Y+\eta/2})+1 \big) \bigg\} \;. \nonumber\\\end{aligned}$$ The energy flow going out to the right thermal bath has been calculated similarly. By summing up the two contributions to obtain the same results as given in Eq.(\[WstExp\]). These derivations mentioned here is to clarify that the energy flow which is carried by a phonon particle flowing from the left bath to the right one. This is the reason why the energy flow in the present case is described by the Landauer formula which is mostly used to represent the electronic current in which an electron flows in from an electron reservoir at one end and going out to the other end through a one-dimensional mesescopic system. J. Jortner, M. Ratner, [*Molecular Electronics*]{}, (Blackwell Science, 1997). D. G. Cahill, W. K. Ford, K. E. Goodson, G. D. Mahan, A. Majumdar, H. J. Maris, R. Merlin, and S. R. Phillpot, J. App. Phys. [**93,**]{} 793 (2003). B. J. Wees, H. va HOuten, C. W. J. Beenaker, J. G. Williamson, L. P. Kouwenhoven, D. van der Marel, and C. T. Foxon, Phys. Rev. Lett. [**60,**]{} 848 (1988). D. A. Wharam, T. J. Thornton, R. Newbury, M. Pepper, and H. Ahmed, J. Phys. C: Solid State Phys. [**21,**]{} L209 (1988). K. Schwab, E. A. Henriksen, J. M. Worlock, and M. L. Roukes, Nature [**404,**]{} 974 (2000). L. G. C. Rego and G. Kirczenow, Phys. Rev. Lett. [**81,**]{} 232 (1998). D. Schwarzer, P. Kutne, C. Schr[ö]{}der, and J. Troe, J. Chem. Phys. [**121,**]{} 1754 (2004). Z. Wang, J. A. Carter, A. Lagutchev, Y. Kan Koh, N.-H. Seong, D. G. Cahill, D. D. Dlott, Science [**317,**]{} 787 (2007). Y. A. Berlin, A. L. Burin, M. Ratner, Chem. Phys. [**275,**]{} 61 (2002). V. Botan, E. H. G. Backus, R. Pfister, A. Moretto, M. Crisma, C. Toniolo, P. H. Nguyen, G. Stock, and P. Hamm, PNAS [**104**]{}, 12749 (2007). P. H. Nguyen, S.-M. Park, and G. Stock, J. Chem. Phys. [**132,**]{} 025102 (2010). K. Kawai, H. Kodera, and T. Majima, J. Am. Chem. Soc. [**132,**]{} 627 (2010). S. Tanaka, K. Kanki, and T. Petrosky, Phys. Rev. B [**80,**]{} 094304 (2009). R. Landauer, IBM J. Res. Dev. [**1,**]{} 223 (1957). Z. Rieder, J. L. Lebowitz, and E. Lieb, J. Math. Phys. [**8,**]{}1073 (1967). U. Z[ü]{}rcher and P. Talkner, Phys. Rev. A [**42,**]{} 3278 (1990). D. Segal and A. Nitzan, Chem. Phys. [**268,**]{} 315 (2001). D. Segal, A. Nitzan, and P. H[ä]{}nggi, J. Chem. Phys. [**119,**]{} 6840 (2003). E. G. Petrov, V. May, P. H[ä]{}nggi, Chem. Phys. [**319,**]{} 380 (2005). M. Galperin, A. Nitzan, and M. A. Ratner, Phys. Rev. B [**75,**]{} 155312 (2007). J. E. Subotnik, T. Hanse, M. A. Ratner, and A. Nitzan, J. Chem. Phys. [**130,**]{} 144105 (2009). I. Prigogine, [*Nonequilibrium statistical mechanics*]{} (John Willey & Sons, 1962). T. Petrosky and I. Prigogine, Foundations Phys. [**29,**]{} 1417 (1999), [*ibid*]{} 1581 (1999). T. Petrosky and I. Prigogine, Adv. Chem. Phys. [**99,**]{} 1 (1997). T. Petrosky and V. Barsegov, Phys. Rev. E[**65,**]{} 46102 (2002). T. Petrosky, Prog. Theor. Phys. [**123,**]{} 395 (2010). P. R[é]{}sibois and M. de Leener, [*Classical Kinetic Theory of Fluids*]{} (John Willey & Sons, 1977). T. Petrosky and I. Prigogine, Chaos, Solitons & Fractals [**7,**]{} 441 (1996). B. A. Tay and G. Ordonetz, Phys. Rev. E [**73,**]{} 016120 (2006). R. Balescu, [*Equilibrium and Nonequilibrium Statistical Mechanics,*]{} (John-Wiley & Sons, New York, 1975). P. R[é]{}sibois, in [*Physics of Many-Particle Systems,*]{} ed. by E. Meeron (Gordon and Breach, New York, 1967). J. Keizer, J. Stat. Phys. [**6,**]{} 67 (1972). J. Schnakenberg, Rev. Mod. Phys. [**48,**]{} 571 (1976). D. Kondepudi and I. Prigogine,[*Modern Thermodynamics*]{} (Wiley, 1998). H. Haken, [*Synergetics: An introduction* ]{}, (Springer, Berlin, Heidelberg, New York, Tokyo,1983) G. D. Mahan, [*Many-Particle Physics*]{} 2nd ed. (Plenum, 1993) See Eq.(12.8) in pp.162 in [*Quantum Kinetics in Transport and Optics of Semiconductors*]{}, ed. by H. Haug and A. -P. Jouho. (Springer-Verlag, Berlin Heidelberg, 1996).
{ "pile_set_name": "ArXiv" }
--- abstract: 'A Hermitian structure on a manifold is called locally conformally Kähler (LCK) if it locally admits a conformal change which is Kähler. In this survey we review recent results of invariant LCK structures on solvmanifolds and present original results regarding the canonical bundle of solvmanifolds equipped with a Vaisman structure, that is, a LCK structure whose associated Lee form is parallel.' address: - 'FaMAF-CIEM, Universidad Nacional de Córdoba, Ciudad Universitaria, 5000 Córdoba, Argentina' - 'FaMAF-CIEM, Universidad Nacional de Córdoba, Ciudad Universitaria, 5000 Córdoba, Argentina, and KU Leuven Kulak, BE-8500 Kortrijk, Belgium' author: - 'A. Andrada' - 'M. Origlia' title: 'Locally conformally Kähler solvmanifolds: a survey' --- Introduction ============ Solvmanifolds and nilmanifolds, i.e., compact quotients of a simply connected solvable (respectively, nilpotent) Lie group by a discrete subgroup, have been used to provide many examples and counterexamples of manifolds with different kind of geometric structures and it is known that they possess a rich structure. For instance, it was shown in [@pal] that the total space of a principal torus bundle over a torus is a 2-step nilmanifold. In complex geometry, we can mention that it was proved in [@BG] that a non-toral nilmanifold cannot admit Kähler structures. This was later generalized to completely solvable solvmanifolds by Hasegawa in [@Has], where he also proves that Kähler solvmanifolds are finite quotients of complex tori which have the structure of a complex torus bundle over a torus. Also, the first example of a compact symplectic manifold that does not admit any Kähler metric happens to be a nilmanifold, the well known Kodaira-Thurston manifold, which is a primary Kodaira surface ([@Kod; @Th]). It is well known that the de Rham cohomology of nilmanifolds and completely solvable solvmanifolds can be computed in terms of invariant forms (according to Nomizu [@N] and Hattori [@Hat], respectively), and it is conjectured that the Dolbeault cohomology of nilmanifolds can also be computed using invariant forms (this conjecture has been proved for several particular cases, see [@CFGU; @CF; @Rol; @FRR]). In [@BDV] it was shown that any nilmanifold with an invariant complex structure has holomorphically trivial canonical bundle, while it is known that this does not happen generally for solvmanifolds (see [@FOU]). Other results regarding complex geometric structures on nilmanifolds and solvmanifolds can be found in [@AV; @CG; @K1; @MPPS; @U; @UV], among many others. It is then a natural question to study the existence of (invariant) *locally conformally Kähler structures* on nilmanifolds and solvmanifolds. We recall that a locally conformally Kähler structure (LCK) on a manifold is a Hermitian structure such that each point has a neighborhood where the metric can be rescaled into a Kähler metric. The family of LCK manifolds is very important and very rich (for instance, diagonal Hopf manifolds, Kodaira surfaces, Kato surfaces, some Oeljeklaus-Toma manifolds admit such structures), and it has been widely studied lately. In particular, there are several recent results about LCK structures on solvmanifolds, and it is our attempt to recollect many of them in this survey. The outline of this article is as follows. In section $2$ we recall the notion and properties of LCK metrics as well as some very well known results on solvmanifolds. In section $3$ we report general facts about invariant LCK structures on solvmanifolds, with emphasis on the case of nilmanifolds. We also recall all the compact complex surfaces that admit a LCK structure which can be seen as solvmanifolds. In section $4$ we recall results of LCK structures on solvmanifolds of special kinds: namely, LCK metrics on solvmanifolds with abelian complex structures, and LCK structures on almost abelian Lie groups. Next, in section $5$ we review the construction of the Oeljeklaus-Toma manifolds and in particular we exhibit them as solvmanifolds, according to the work of Kasuya. In section $6$ we review some recent results on Vaisman solvmanifolds and we also present original results about the triviality of the canonical bundle of such a manifold (see Theorem \[canonical\]). Finally, in section $7$ we discuss briefly locally conformally symplectic solvmanifolds and their properties. **Acknowledgements.** The authors would like to thank Isabel Dotti for useful suggestions during the preparation of this manuscript. Preliminaries ============= Let $(M,J,g)$ be a $2n$-dimensional Hermitian manifold, where $J$ is a complex structure and $g$ is a Hermitian metric, and let $\omega$ denote its fundamental $2$-form, that is, $\omega(X,Y)=g(JX,Y)$ for any $X,Y$ vector fields on $M$. The manifold $(M,J,g)$ is called [*locally conformally Kähler*]{} (LCK) if $g$ can be rescaled locally, in a neighborhood of any point in $M$, so as to be Kähler, i.e., there exists an open covering $\{ U_i\}_{i\in I}$ of $M$ and a family $\{ f_i\}_{i\in I}$ of $C^{\infty}$ functions, $f_i:U_i \to \R$, such that each local metric $$\label{gi} g_i=\exp(-f_i)\,g|_{U_i}$$ is Kähler. These manifolds are a natural generalization of the class of Kähler manifolds, and they have been much studied since the work of I. Vaisman in the ’70s. The main references for locally conformally Kähler geometry are [@DO], the more recent reports [@O; @OV1] and the bibliography therein. An equivalent characterization of a LCK manifold can be given in terms of the fundamental form $\omega$. Indeed, a Hermitian manifold $(M,J,g)$ is LCK if and only if there exists a closed $1$-form $\theta$ globally defined on $M$ such that $$\label{lck} d\omega=\theta\wedge\omega.$$ This closed $1$-form $\theta$ is called the *Lee form* (see [@L]). Furthermore, the Lee form $\theta$ is uniquely determined by the following formula: $$\label{tita} \theta=-\frac{1}{n-1}(\delta\omega)\circ J,$$ where $\delta$ is the codifferential operator and $2n$ is the dimension of $M$. It follows from [@Ga] that $(1-n)\theta$ can be identified with the trace of the torsion of the Chern connection. A Hermitian manifold $(M,J,g)$ is called *globally conformally Kähler* (GCK) if there exists a $C^{\infty}$ function, $f:M\to\R$, such that the metric $\exp(-f)g$ is Kähler, or equivalently, the Lee form is exact. Therefore a simply connected LCK manifold is GCK. There is yet another equivalent definition of LCK manifolds (see for instance [@DO]). Let $\tilde M$ be the universal covering of the complex manifold $(M,J)$. Then $M$ has a compatible LCK metric if and only if there is a representation $\chi:\pi_1(M)\to \R_{>0}$ and a Kähler metric $\tilde g$ on $\tilde M$ such that $$\gamma^*(\tilde{g})=\chi(\gamma)\tilde{g}, \qquad \text{for any } \gamma\in \pi_1(M),$$ i.e., the fundamental group of $M$ (seen as the group of deck transformations of $\tilde M$) acts by homotheties. Note that the LCK metric metric obtained is GCK if and only if $\chi\equiv 1$. It is well known that LCK manifolds belong to the class $\mathcal{W}_4$ of the Gray-Hervella classification of almost Hermitian manifolds [@GH]. Also, an LCK manifold $(M,J,g)$ is Kähler if and only if $\theta=0$. Indeed, $\theta\wedge\omega=0$ and $\omega$ non degenerate imply $\theta=0$. It is known that if $(M,J,g)$ is a Hermitian manifold with $\dim M\ge 6$ such that holds for some $1$-form $\theta$, then $\theta$ is automatically closed, and therefore $M$ is LCK. In contrast to the Kähler class, the LCK class is not stable under small deformations. Indeed, it follows from [@Bel] that some Inoue surfaces do not admit LCK structures but they are complex deformations of other Inoue surfaces which do admit LCK metrics [@Tr]. On the other hand, just as in the Kähler case, the LCK class is stable under blowing-up points (see [@Tr; @Vu1]). The Hopf manifolds are examples of LCK manifolds, and they are obtained as a quotient of $\C^n-\{0\}$ with the Boothby metric by a discrete subgroup of automorphisms. These manifolds are diffeomorphic to $S^1\times S^{2n-1}$ and for $n\geq 2$ they have first Betti number equal to 1, so that they do not admit any Kähler metric. The LCK structures on these Hopf manifolds have a special property, as shown by Vaisman in [@V2]. Indeed, the Lee form is parallel with respect to the Levi-Civita connection of the Hermitian metric. The LCK manifolds sharing this property form a distinguished class, which has been much studied since Vaisman’s seminal work [@Bel; @GMO; @KS; @OV3; @OV4; @V2; @V3]. Indeed, LCK manifolds with parallel Lee form are nowadays called *Vaisman manifolds*. Vaisman manifolds satisfy stronger topological properties than general LCK manifolds. For instance, a compact Vaisman non-Kähler manifold $(M,J,g)$ has $b_1(M)$ odd ([@KS; @V3]), which implies that such a manifold cannot admit Kähler metrics. Moreover, it was proved in [@OV3 Structure Theorem] and [@OV4 Corollary 3.5] that any compact Vaisman manifold admits a Riemannian submersion to a circle such that all fibers are isometric and admit a natural Sasakian structure. It was shown in [@Ve] that any compact complex submanifold of a Vaisman manifold is Vaisman, as well. In [@Bel] the classification of compact complex surfaces admitting a Vaisman structure is given. It is known that a homogeneous LCK manifold is Vaisman when the manifold is compact ([@GMO; @HK2]) and, more generally, when the manifold is a quotient of a reductive Lie group such that the normalizer of the isotropy group is compact ([@ACHK]). In [@AHK] it was proved that a simply connected homogeneous Vaisman manifold $G/H$ with $G$ unimodular is isomorphic to a product $\R\times M_1$, where $M_1$ is a simply connected homogeneous Sasakian manifold of a unimodular Lie group, which in turn is a certain $\R$-bundle or $S^1$-bundle over a simply connected homogeneous Kähler manifold of a reductive Lie group. Associated to any LCK metric on a Hermitian manifold there is a cohomology $H^*_\theta(M)$ which can be defined as follows. Since the corresponding Lee form $\theta$ is closed, we can deform the de Rham differential $d$ to obtain the adapted differential operator $$d_\theta \alpha= d\alpha -\theta\wedge\alpha.$$ This operator satisfies $d_\theta^2=0$, thus it defines the *Morse-Novikov cohomology* $H_\theta^*(M)$ of $M$ relative to the closed $1$-form $\theta$, also called the Lichnerowicz cohomology or the adapted cohomology. It is known that if $M$ is a compact oriented $n$-dimensional manifold, then $H_\theta^0(M)= H_\theta^n(M)=0$ for any non exact closed $1$-form $\theta$ (see for instance [@GL; @Ha]). For any LCK structure $(\omega,\theta)$ on $M$, the $2$-form $\omega$ defines a cohomology class $[\omega]_\theta\in H_\theta^2(M)$, since $d_\theta\omega=d\omega-\theta\wedge \omega=0$. It was proved in [@LLMP] that the Morse-Novikov cohomology associated to the Lee form of a Vaisman structure on a compact manifold vanishes. Some facts on solvmanifolds --------------------------- Recall that a discrete subgroup $\Gamma$ of a simply connected Lie group $G$ is called a *lattice* if the quotient $\Gamma\backslash G$ is compact. According to [@Mi], if such a lattice exists then the Lie group must be unimodular. The quotient $\Gamma\backslash G$ is known as a solvmanifold if $G$ is solvable and as a nilmanifold if $G$ is nilpotent, and it is known that $\pi_1(\Gamma\backslash G)\cong \Gamma$. Moreover, the diffeomorphism class of solvmanifolds is determined by the isomorphism class of the corresponding lattices, as the following results show: [@R Theorem 3.6] Let $G_1$ and $G_2$ be simply connected solvable Lie groups and $\Gamma_i$, $i=1,2$, a lattice in $G_i$. If $f:\Gamma_1\to\Gamma_2$ is an isomorphism, then there exists a diffeomorphism $F:G_1\to G_2$ such that 1. $F|_{\Gamma_1}=f$, 2. $F(\gamma g)=f(\gamma)F(g)$, for any $\gamma\in \Gamma_1$ and $g\in G_1$. [@Mo] \[mostow\] Two solvmanifolds with isomorphic fundamental groups are diffeomorphic. It is not easy to determine whether a given unimodular solvable Lie group admits a lattice. However, there is such a criterion for nilpotent Lie groups. Indeed, it was proved by Malcev in [@Ma] that a nilpotent Lie group admits a lattice if and only if its Lie algebra has a rational form, i.e. there exists a basis of the Lie algebra such that the corresponding structure constants are all rational. More recently, in [@B], the existence of lattices in simply connected solvable Lie groups up to dimension $6$ was studied. A general result proved by Witte in [@Wi Proposition 8.7] states that only countably many nonisomorphic simply connected Lie groups admit lattices. In the case when $G$ is completely solvable, i.e., it is a solvable Lie group such that the endomorphisms $\ad_X$ of its Lie algebra $\g$ have only real eigenvalues for all $X\in \g$, the de Rham and the Morse-Novikov cohomology of $\Gamma\backslash G$ can be computed in terms of the cohomology of $\g$. Indeed, Hattori proved in [@Hat] that if $V$ is a finite dimensional triangular[^1] $\g$-module, then $\overline{V}:=C^\infty(\Gamma\backslash G)\otimes V$ is a $\mathfrak{X}(\Gamma\backslash G)$-module and there is an isomorphism $$\label{kill} H^*(\g,V) \cong H^*(\mathfrak{X}(\Gamma\backslash G), \overline{V}).$$ Therefore: - If $V=\R$ is the trivial $\g$-module, then the right-hand side in gives the usual de Rham cohomology of $\Gamma\backslash G$, so that $$\label{deRham} H^*(\g) \cong H^*_{dR}(\Gamma\backslash G).$$ - If $V=V_\theta$, where $\theta$ is a closed left-invariant 1-form and $V_{\theta}$ is a $1$-dimensional $\g$-module with action given by $$Xv=-\theta(X)v, \quad X\in\g,\, v\in V_{\theta},$$ then we can identify $\overline{V}$ with $C^\infty(\Gamma\backslash G)$ and the action of $\mathfrak{X}(\Gamma\backslash G)$ on $C^\infty(\Gamma\backslash G)$ is given by $$X\cdot f= Xf-\theta(X)f, \qquad X\in\g, \, f\in C^\infty(\Gamma\backslash G).$$ Here we are using that there is a natural inclusion $\g \hookrightarrow \mathfrak{X}(\Gamma\backslash G)$ and a bijection $C^\infty(\Gamma\backslash G)\otimes \g \to \mathfrak{X}(\Gamma\backslash G)$ given by $f\otimes X \mapsto fX$. As a consequence, in this case becomes (cf. [@Mil Corollary 4.1]) $$\label{kill_adapted} H^*_\theta(\g)\cong H^*_\theta(\Gamma\backslash G).$$ In particular, $H^*_{dR}(\Gamma\backslash G)$ and $H^*_\theta(\Gamma\backslash G)$ do not depend on the lattice $\Gamma$. In the general case, i.e. when $G$ is not necessarily completely solvable, there is always an injection $i^*:H^*_\theta(\g)\to H^*_\theta(\Gamma\backslash G)$, where $\theta$ is any closed left-invariant 1-form, as shown in [@K]. LCK solvmanifolds ================= Let $G$ be a Lie group with a left invariant Hermitian structure $(J,g)$. If $(J,g)$ satisfies the LCK condition , then $(J,g)$ is called a [*left invariant LCK structure*]{} on the Lie group $G$. Clearly, the fundamental $2$-form is left invariant and, using , it is easy to see that the corresponding Lee form $\theta$ on $G$ is also left invariant. This fact allows us to define LCK structures on Lie algebras. We recall that a [*complex structure J*]{} on a Lie algebra $\g$ is an endomorphism $J: \g \to \g$ satisfying $J^2=-\I$ and $$N_J=0, \quad \text{where} \quad N_J(x,y)=[Jx,Jy]-[x,y]-J([Jx,y]+[x,Jy]),$$ for any $x,y \in \g$. Let $\g$ be a Lie algebra, $J$ a complex structure and $\pint$ a Hermitian inner product on $\g$, with $\omega\in\alt^2\g^*$ the fundamental $2$-form. We say that $(\g,J,\pint)$ is [*locally conformally Kähler*]{} (LCK) if there exists $\theta \in \g^*$, with $d\theta=0$, such that $$\label{g-lck-0} d\omega=\theta\wedge\omega.$$ We will assume from now on that $\theta\neq 0$, so that we are excluding the Kähler case. Recall that if $\alpha\in\g^*$ and $\eta\in\alt^2\g^*$, then their exterior derivatives $d\alpha\in\alt^2\g^*$ and $d\eta\in\alt^3\g^*$ are given by $$d\alpha(x,y)= -\alpha([x,y]), \qquad d\eta(x,y,z)=-\eta([x,y],z)-\eta([y,z],x)-\eta([z,x],y),$$ for any $x,y,z\in\g$. We have the following orthogonal decomposition for a Lie algebra $\g$ with an LCK structure $(J,\pint)$, $$\g=\mathbb{R}A \oplus \ker\theta$$ where $\theta$ is the Lee form and $\theta(A)=1$. Since $d\theta=0$, we have that $\g'=[\g,\g]\subset\ker\theta$. Note also that since $\theta\neq0$, $\g$ cannot be a semisimple Lie algebra. It is clear that $JA\in\ker\theta$, but more can be said when $\g$ is unimodular. Indeed, we proved in [@AO] that $JA$ is in the commutator ideal $[\g,\g]$ of $\g$, using . We note that any left invariant LCK structure $(J,g)$ on a Lie group $G$ induces naturally a LCK structure on a quotient $\Gamma\backslash G$ of $G$ by a discrete subgroup $\Gamma$. The induced LCK structure on the quotient will be called *invariant*. LCK nilmanifolds ---------------- The first example of an LCK nilmanifold was given in [@CFL]. We recall their construction: \[heisenberg\] Let $\g=\R\times\h_{2n+1}$, where $\h_{2n+1}$ is the $(2n+1)$-dimensional Heisenberg Lie algebra. There is a basis $\{X_1,\dots,X_n,Y_1,\dots,Y_n,Z_1,Z_2\}$ of $\g$ with Lie brackets given by $[X_i,Y_i]=Z_1$ for $i=1,\dots,n$ and $Z_2$ in the center. We define an inner product $\pint$ on $\g$ such that the basis above is orthonormal. Let $J_0$ be the almost complex structure on $\g$ given by: $$J_0X_i=Y_i, \quad J_0Z_1=-Z_2 \; \; \; \text{for $i=1,\dots,n$}.$$ It is easily seen that $J_0$ is a complex structure on $\g$ compatible with $\pint$. If $\{x^i,y^i,z^1,z^2\}$ denote the $1$-forms dual to $\{X_i,Y_i,Z_1,Z_2\}$ respectively, then the fundamental $2$-form is: $$\omega=\sum_{i=1}^n(x^i\wedge y^i) - z^1\wedge z^2.$$ Thus, $$d\omega=z^2\wedge\omega,$$ and therefore $(\g,J_0,\pint)$ is LCK. It is known that $\g$ is the Lie algebra of the Lie group $\R\times H_{2n+1}$, where $H_{2n+1}$ is the $(2n+1)$-dimensional Heisenberg group. The Lie group $H_{2n+1}$ admits a lattice $\Gamma$ and therefore the nilmanifold $N= S^1 \times \Gamma\backslash H_{2n+1}$ admits an invariant LCK structure which is Vaisman. The nilmanifold $N$ is a primary Kodaira surface and it cannot admit any Kähler metric, due to [@BG]. In [@U], it was proved that if $(J,g)$ is a Hermitian structure on a $6$-dimensional nilmanifold $\Gamma\backslash G$ such that $J$ is invariant, then this structure is LCK if and only if $G$ is isomorphic to $H_5\times \R$, and the complex structure $J$ is equivalent to $J_0$. Moreover, the LCK structure is actually Vaisman. Based on this result, Ugarte states in the same article the following conjecture: **Conjecture:** A $(2n+2)$-dimensional nilmanifold admitting an LCK structure (not necessarily invariant) is diffeomorphic to a product $\Gamma\backslash H_{2n+1}\times S^1$, where $\Gamma$ is a lattice in $H_{2n+1}$. We mention next some recent advances towards the solution of this conjecture. In [@S], H. Sawai proves the conjecture in the case when the complex structure on the Hermitian nilmanifold is invariant, that is, he extends the result of Ugarte to any even dimension. Let $(M,J)$ be a non-toral compact nilmanifold with a left invariant complex structure. If $(M,J)$ admits a locally conformally Kähler metric, then $(M,J)$ is biholomorphic to a quotient of $(H_{2n+1}\times \R ,J_0)$. In his proof, Sawai uses the fact that the Morse-Novikov cohomology $H^k_{\theta}(\g)$ of a nilpotent Lie algebra $\g$ is trivial for any closed $1$-form $\theta$ and $k\geq 2$. Another weaker version of the conjecture was proved by G. Bazzoni in [@Baz]. He considers nilmanifolds equipped with Vaisman structures and he proves: \[nilpotent\] A $(2n+2)$-dimensional nilmanifold admitting a Vaisman structure (not necessarily invariant) is diffeomorphic to a product $\Gamma\backslash H_{2n+1}\times S^1$, where $\Gamma$ is a lattice in $H_{2n+1}$. The idea of the proof goes as follows. If $\Gamma\backslash G$ is a compact Vaisman nilmanifold then, due to results of Ornea-Verbitsky, there exists a Sasakian manifold $S$ equipped with a Sasakian automorphism $\varphi: S\to S$ such that $\Gamma\backslash G$ is diffeomorphic to the mapping torus $S_{\varphi}$. Then it is proved that this Sasakian manifold has the same minimal model as a compact nilmanifold $\Lambda\backslash H$, where $\Lambda=\pi_1(S)$. Using some arguments concerning minimal models of nilpotent spaces appearing in [@CDMY], the author proves that the nilpotent Lie group $H$ is isomorphic to $H_{2n+1}$. Standard results on Lie groups and their lattices imply the result. 4-dimensional LCK solvmanifolds ------------------------------- In [@ACFM], it is given the first known example of a compact solvmanifold equipped with an LCK structure, which is not diffeomorphic to a nilmanifold. Their construction is given as follows: \[ejemplo-4d\] Let $\g$ be the $4$-dimensional Lie algebra with basis $\{A,X,Y,Z\}$ with Lie brackets $$[A,X]=X,\quad [A,Z]=-Z, \quad [X,Z]=-Y.$$ Define an inner product on $\g$ in such a way that the basis above is orthonormal, and define an almost complex structure $J$ on $\g$ by $$JA=Z, \quad JX=Y.$$ It is easy to see that $J$ is integrable, thus it is a complex structure on $\g$. If $\{ \alpha,x,y,z\}$ denotes the dual basis of $\g^*$, then the fundamental $2$-form $\omega$ is given by $\omega=\alpha\wedge z+x\wedge y$, and hence $d\omega=-\alpha\wedge \omega$. Consequently, $\g$ admits an LCK structure with Lee form $\theta=-\alpha$. Note that it is not Vaisman since $(\nabla_X \theta)(X)=1$. It is well known that the corresponding simply connected Lie group $G$ has non-nilpotent lattices, and therefore if $\Gamma\backslash G$ is any associated solvmanifold, then it carries a non-Vaisman LCK structure. It was proved by Kamishima in [@Kam] that this invariant LCK structure on the compact complex surface $\Gamma\backslash G$ coincides with the LCK structure constructed by Tricerri on the Inoue surface of type $S^+$ in [@Tr]. As mentioned before, the solvmanifold from Example \[ejemplo-4d\] can be identified with an Inoue surface of type $S^+$. More generally, it was proved in [@Ha1 Theorem 1] that a compact complex surface $X$ diffeomorphic to a solvmanifold $\Gamma\backslash G$ is either (1) a complex torus, (2) a primary Kodaira surface, (3) a secondary Kodaira surface, (4) an Inoue surface of type $S^0$, (5) an Inoue surface of type $S^+$, or (6) a hyperelliptic surface; moreover, the complex structure on $X$ can be seen to be induced by a left invariant one on $G$. In each case we recall the structure equations for the Lie algebra of $G$, which has a basis $\{A,X,Y,Z\}$: 1. complex torus: all brackets vanish, 2. primary Kodaira surface: $[X,Y]=Z$, 3. secondary Kodaira surface: $[X,Y]=Z,\, [A,X]=Y,\, [A,Y]=-X$, 4. Inoue surface of type $S^0$: $[A,X]=-\frac12 X+bY,\, [A,Y]=-bX-\frac12 Y,\, [A,Z]=Z$, $b\in\R$, 5. Inoue surface of type $S^+$: $[X,Z]=-Y,\, [A,X]=X,\, [A,Z]=-Z$, 6. hyperelliptic surface: $[A,X]=-Y,\, [A,Y]=X$. According to [@HK], all of the $4$-dimensional compact complex surfaces above admit a LCK structure, except for the complex torus and the hyperelliptic surface. More precisely, consider for the cases (2)-(5) the complex structure $J$ given by $JX=Y,\, JA=Z$ and the metric such that the basis $\{A,X,Y,Z\}$ is orthonormal. If $\{\alpha,x,y,z\}$ denotes the dual basis of this basis then the fundamental $2$-form $\omega$ is given by $\omega=\alpha\wedge z+x\wedge y$, and the Lee form is $\theta=\alpha$ for cases (2)-(4) and $\theta=-\alpha$ for the case (5). More recently, in [@AngO] a complete classification of locally conformally Kähler structures on $4$-dimensional solvable Lie algebras (not only unimodular) up to linear equivalence was given. For the sake of completeness, we recall that LCK structures on $4$-dimensional reductive Lie algebras were studied in [@ACHK]. The authors prove that if a reductive Lie algebra admits a LCK structure then the Lie algebra is isomorphic to either $\mathfrak{gl}(2,\R)$ or $\mathfrak{u}(2)$. In particular, such examples occur only in dimension four. A compact quotient of $\operatorname{GL}(2,\R)$ by a discrete subgroup corresponds to a properly elliptic surface, while $\operatorname{U}(2)$ with an invariant complex structure corresponds to a Hopf surface. Special cases ============= In this section we review results concerning special cases of LCK structures on solvable Lie algebras. First we consider LCK structures where the complex structure is abelian, and later we analyze the existence of LCK structures on almost abelian Lie algebras. LCK structures with abelian complex structure --------------------------------------------- An almost complex structure $J$ on $\g$ is called *abelian* if $$[JX,JY]=[X,Y], \, \, \text{for all} \, \, X,Y\in \g.$$ Note that an abelian almost complex structure is automatically integrable, and therefore it will be called an abelian complex structure. These complex structures have had interesting applications in differential geometry. For instance, a pair of anticommuting abelian complex structures on $\g$ gives rise to an invariant weak HKT structure on any Lie group $G$ associated to $\g$ (see [@DF] and [@GP]). In [@CF] it has been shown that the Dolbeault cohomology of a nilmanifold with an abelian complex structure can be computed algebraically. Also, deformations of abelian complex structures on nilmanifolds have been studied in [@CFP; @MPPS]. More recently, it was shown in [@UV] that the holonomy group of the Bismut connection of an invariant Hermitian structure on a $6$-dimensional nilmanifold reduces to a proper subgroup of $SU(3)$ if and only if the complex structure is abelian. This result was later extended in [@AV] to solvmanifolds of any dimension equipped with abelian balanced Hermitian structures. Some properties of this kind of complex structures are stated in the following result (see [@ABD; @BD] for their proofs). \[prop\] Let $\g$ be a Lie algebra with $\z(\g)$ its center and $\g':=[\g,\g]$ its commutator ideal. If $J$ is an abelian complex structure on $\g$, then 1. $J\z(\g)=\z(\g)$. 2. $\g' \cap J\g'\subset \z(\g'+J\g')$. 3. The codimension of $\g'$ is at least $2$, unless $\g$ is isomorphic to ${\mathfrak{aff}}(\R)$ (the only $2$-dimensional non-abelian Lie algebra). 4. $\g'$ is abelian, therefore $\g$ is $2$-step solvable. We recall next the main result from [@AO], concerning Lie algebras equipped with a LCK structure with abelian complex structure. Before stating the main result, we consider the following variation of Example \[heisenberg\]. Recall that $\R\times\mathfrak{h}_{2n+1}$ has a basis $\{X_1,\dots,X_n,Y_1,\dots,Y_n,Z_1,Z_2\}$ such that $[X_i,Y_i]=Z_1$ for $i=1,\dots,n$, and that this Lie algebra admits an abelian complex structure $J_0$ given by $J_0X_i=Y_i,\, J_0Z_1=-Z_2$. For any $\lambda>0$, consider the metric $\pint_\lambda$ such that the basis above is orthogonal, with $|X_i|=|Y_i|=1$ but $|Z_1|^2=|Z_2|^2=\frac{1}{\lambda}$. It is easy to see (just as in Example \[heisenberg\]) that $(J_0,\pint_\lambda)$ is an LCK structure, in fact, it is Vaisman. Furthermore, the metrics $\pint_\lambda$ are pairwise non-isometric, since the scalar curvature of $\pint_\lambda$ is $-\frac{n\lambda^2}{2}$. Let $(J,\pint)$ be an LCK structure on $\g$ with abelian complex structure $J$. If $\g$ is unimodular then $\g \simeq \R\times\h_{2n+1}$, where $\mathfrak{h}_{2n+1}$ is the $(2n+1)$-dimensional Heisenberg Lie algebra, and $(J,\pint)$ is equivalent to $(J_0,\pint_\lambda)$ for some $\lambda>0$. In consequence, up to isometry and rescaling, the only examples of LCK structures on Lie algebras with abelian complex structures are those from Example \[heisenberg\], and they are Vaisman. LCK structures on almost abelian solvmanifolds ---------------------------------------------- We recall that a Lie group $G$ is said to be *almost abelian* if its Lie algebra $\g$ has a codimension one abelian ideal. Such a Lie algebra will be called almost abelian, and it can be written as $\g= \R f_1 \ltimes_{\ad_{f_1}} \mathfrak{u}$, where $\mathfrak u$ is an abelian ideal of $\g$, and $\R$ is generated by $f_1$. Accordingly, the Lie group $G$ is a semidirect product $G=\R\ltimes_\varphi \R^d$ for some $d\in\mathbb N$, where the action is given by $\varphi(t)=e^{t\ad_{f_1}}$. In [@AO1] we considered the existence of invariant LCK structures on solvmanifolds associated to almost abelian Lie groups. Firstly, we showed that there are plenty of almost abelian Lie algebras which admit LCK structures. Indeed, they are completely characterized in the next result: \[lcK\] Let $\g$ be a $(2n+2)$-dimensional almost abelian Lie algebra and $(J,\pint)$ a Hermitian structure on $\g$, and let $\g'$ denote the commutator ideal $[\g,\g]$ of $\g$. 1. If $\dim \g'=1$, then $(J,\pint)$ is LCK if and only if $\g$ is isomorphic to $\mathfrak{h}_3 \times\R$ or $\mathfrak{aff}(\R)\times \R^2$ as above. 2. If $\dim\g'\geq 2$, then $(J,\pint)$ is LCK if and only if $\g$ can be decomposed as $\g={\mathfrak a}^\perp\ltimes \mathfrak a$, a $J$-invariant orthogonal sum with a codimension $2$ abelian ideal $\mathfrak a$, and there exists an orthonormal basis $\{f_1,f_2\}$ of ${\mathfrak a}^\perp$ such that $$[f_1,f_2]=\mu f_2, \quad f_2=Jf_1, \quad \ad_{f_2}|_{\mathfrak a}=0 \text{ and } \ad_{f_1}|_{\mathfrak a}= \lambda I + B,$$ for some $\mu,\lambda\in\R$, $\lambda\neq 0$, and $B\in \mathfrak{u}(n)$. The corresponding Lee form is given by $\theta=-2\lambda f^1$. Furthermore, the Lie algebra $\g$ is unimodular if and only if $\lambda=-\frac{\mu}{2n}$. We note that the left invariant LCK structures obtained on the Lie groups corresponding to the Lie algebras in Theorem \[lcK\] (2) or $\mathfrak{aff}(\R)\times \R^2$ from Theorem \[lcK\] (1) are never Vaisman. This follows from Lemma \[ad\_A-antisim\] below. On the other hand, any LCK structure on $\mathfrak{h}_3\times \R$ from Theorem \[lcK\] (1) is Vaisman (see [@AO; @S]). However, almost abelian solvmanifolds equipped with invariant LCK structures are very scarce, since we have shown in [@AO1] that they only occur in dimension 4. \[dim-6\] If $G$ is as above with $\mu\neq0$ and $\dim G \geq 6$, i.e. $n\geq 2$, then $G$ admits no lattice. This theorem is a consequence of the following result about the roots of a certain class of polynomials with integer coefficients. \[integer-polynomial\] Let $p$ be a polynomial of the form $$p(x)=x^{2n+1}-m_{2n}x^{2n}+m_{2n-1}x^{2n-1}+\cdots+m_1x-1$$ with $m_j\in\Z$ and $n\geq 2$, and let $x_0,\dots,x_{2n}\in \C$ denote the roots of $p$. If $x_0\in\R$ is a simple root and $|x_1|=\dots=|x_{2n}|$, then $x_0=1$ and $|x_j|=1$, $j=1,\ldots, 2n$. Indeed, the proof of Theorem \[dim-6\] goes as follows. If $\g=\R f_1\ltimes \R^{2n+1}$, then the Lie group $G$ is a semidirect product $G=\R\ltimes_{\varphi} \R^{2n+1}$, where $\varphi(t)=\exp(t\ad_{f_1}|_{\R^{2n+1}})$, $t\in\R$. Let us assume that $G$ admits lattices. According to [@B], there exists a $t_0\neq 0$ such that $\varphi(t_0)$ can be conjugated to an integer matrix. Therefore, taking into account that $\dim G\geq 6$, it follows from Theorem \[lcK\] that the characteristic polynomial of $\varphi(t_0)$ satisfies the hypothesis of Lemma \[integer-polynomial\], hence its only simple root $e^{t_0\mu}$ is equal to 1, equivalently, $t_0\mu=0$. Since $\mu\neq 0$ (otherwise the Hermitian structure would be Kähler) we obtain a contradiction. In dimensions at least 6 there are no almost abelian solvmanifolds equipped with invariant LCK structures. In dimension 4, a unimodular almost abelian Lie algebra admitting an LCK structure is isomorphic either to $\h_3\times\R$ or to $\g_b$, for $b\geq 0$, where $\g_b=\R f_1\ltimes \R^3$ and the adjoint action is given by $$\ad_{f_1}|_{\R^3}=\begin{pmatrix} 1 \\ & -\frac{1}{2}& -b\\ & b & -\frac{1}{2} \end{pmatrix}.$$ It was shown in [@AO1] that the associated simply connected almost abelian Lie groups $G_b$ admit lattices for countably many values of the parameter $b$. As mentioned before, it is known that the solvmanifolds associated to $\g_b$ for these values of $b$ are Inoue surfaces of type $S^0$ [@Kam; @Tr] (see [@S1] for an explicit construction of a lattice on some of the Lie groups $G_b$). Oeljeklaus-Toma manifolds ========================= In this section we review the construction of the Oeljeklaus-Toma manifolds (OT manifolds for short), which appeared in [@OT]. These non-Kähler compact complex manifolds arise from certain number fields, and they can be considered as generalizations of the Inoue sufaces of type $S^0$. We recall briefly their construction. Let $K$ denote a number field of degree $n=[K:\Q]$ with $s>0$ real embeddings $\sigma_1,\ldots,\sigma_s$ and $2t>0$ complex embeddings $\sigma_{s+1},\ldots, \sigma_{s+2t}$ into $\C$, with $\sigma_{s+i+t}=\overline{\sigma}_{s+i}$ for $i=1,\ldots,t$; therefore $n=s+2t$. For any choice of natural numbers $s,t$, such a field always exists, according to [@OT Remark 1.1]. Let $\mathcal{O}_K$ denote the ring of algebraic integers of $K$, $\mathcal{O}^\times_K$ its multiplicative group of units and set $$\mathcal{O}^{\times,+}_K:=\{ a\in \mathcal{O}^\times_K \mid \sigma_i(a)>0,\, i=1,\ldots,s\}.$$ If $\H$ denotes the upper complex half-plane, Oeljeklaus and Toma proved, using the Dirichlet’s units theorem, that $\mathcal{O}^{\times,+}_K\ltimes \mathcal{O}_K$ acts freely on $\H^s\times \C^t$ by $$\begin{aligned} (a,b)\cdot & (x_1+iy_1,\ldots,x_s+iy_s,z_1,\ldots,z_t) = \\ & = (\sigma_1(a)x_1+\sigma_1(b)+i\sigma_1(a)y_1,\ldots, \sigma_s(a)x_s+\sigma_s(b)+i\sigma_s(a)y_s, \\ & \qquad \quad \sigma_{s+1}(a)z_1+\sigma_{s+1}(b),\ldots,\sigma_{s+t}(a)z_t+\sigma_{s+t}(b)).\end{aligned}$$ Moreover, there exists an *admissible* subgroup $U\subset \mathcal{O}^{\times,+}_K$ such that the action of $U\ltimes \mathcal{O}_K$ is properly discontinuous. For $t=1$, every subgroup of finite index in $\mathcal{O}^{\times,+}_K$ is admissible. The manifold $(\H^s\times \C^t)/U\ltimes \mathcal{O}_K$ is called an OT manifold of type $(s,t)$ and it is usually denoted $X(K,U)$. It is a compact complex manifold of (complex) dimension $s+t$. These manifolds have many interesting properties; for instance, it was proved in [@OT] that the following holomorphic bundles over an OT manifold $X=X(K,U)$ are flat and admit no global holomorphic section: the bundle of holomorphic $1$-forms $\Omega^1_X$, the holomorphic tangent bundle $\mathcal{T}_X$ and any positive power $K_X^{\otimes k}$ of the canonical bundle. Moreover, the Kodaira dimension of any OT manifold is $-\infty$ and they can never be Kähler. Concerning Hermitian metrics compatible with the complex structure of OT manifolds, the following result is well known when $t=1$: For $s>0$, any OT manifold of type $(s,1)$ admits LCK metrics. Indeed, consider the following potential $$F: \H^s\times \C\to \R,\qquad F(z):=\frac{1}{\prod_{j=1}^s(i(z_j-\overline{z}_j))}+|z_{s+1}|^2.$$ Then $\omega:=i\partial\overline{\partial}F$ gives the desired Kähler form on $\H^s\times \C$. Note that for $s = t = 1$ and $U= \mathcal{O}^{\times,+}_K$, the OT manifold $X(K,U)$ is an Inoue surface of type $S^+$ with the metric given in [@Tr]. In [@K] Kasuya proved another interesting feature of OT manifolds. Indeed, he proved: OT manifolds are solvmanifolds. In order to prove this, Kasuya considers the Lie group $G=\R^s\ltimes_\phi(\R^s\times \C^t)$, where the action of $\phi$ can be described as follows. For $a\in U$, set $t_i=\log|\sigma_i(a)|$, $i=1,\ldots,s$; it can be shown that there exist $b_{ik},\,c_{ik}\in \R$ such that $$|\sigma_{s+k}(a)|=\exp\left(\frac12 \sum_ib_{ik}t_i\right) \quad \text{and } \quad \sigma_{s+k}(a)=\exp\left(\frac12 \sum_ib_{ik}t_i+\sum_i c_{ik}t_i\right).$$ Now, for $(t_1,\ldots,t_s)\in\R^s$, we have $$\phi(t_1,\ldots,t_s)=\operatorname{diag}(e^{t_1},\ldots,e^{t_s},e^{\psi_1+i\varphi_1},\ldots,e^{\psi_t+i\varphi_t}),$$ where $\psi_k=\frac12 \sum_ib_{ik}t_i$, $\varphi_k=\sum_i c_{ik}t_i$. It is clear that $G$ is a solvable Lie group. The subgroup $U\ltimes \mathcal{O}_K$ can be embedded as a lattice in $G$; moreover, under the correspondence $$\begin{aligned} \H^s\times \C^t & \ni (x_1+iy_1,\ldots,x_s+iy_s,z_1,\ldots,z_t) \\ & \mapsto (x_1,\log y_1,\ldots,x_s,\log y_s,z_1,\ldots,z_t) \in \R^s\times \R^s\times \C^t,\end{aligned}$$ the action of $U\ltimes \mathcal{O}_K$ on $\H^s\times \C^t $ can be identified with the left action of the lattice $U\ltimes \mathcal{O}_K$ on $G$. Therefore, the OT manifold $X(K,U)=(\H^s\times \C^t)/U\ltimes \mathcal{O}_K$ may be considered as the solvmanifold $U\ltimes \mathcal{O}_K\backslash G$. Furthermore, it can be seen that the natural complex structure on $G$ is left invariant and the induced complex structure on $X(K,U)$ coincides with the complex structure constructed in [@OT]. If $\g$ denotes the Lie algebra of $G$, then $\g$ has real dimension $2(s+t)$ and there is a basis $\{\alpha_1,\ldots,\alpha_s,\beta_1,\ldots,\beta_s,\gamma_1,\ldots,\gamma_{2t}\}$ of $\g$ such that the differential is given by $$\begin{aligned} d\alpha_j & =0, \\ d\beta_j & =-\alpha_j\wedge\beta_j,\\ d\gamma_{2k-1} & =\frac12 \sum_j b_{jk}\alpha_j\wedge\gamma_{2k-1}+\sum_j c_{jk}\alpha_j\wedge\gamma_{2k},\\ d\gamma_{2jk} & =-\sum_j c_{jk}\alpha_j\wedge\gamma_{2k-1}+\frac12 \sum_j b_{jk}\alpha_j\wedge\gamma_{2k},\end{aligned}$$ for any $j=1,\ldots,s$, $k=1,\ldots,t$. Accordingly, $\g$ has a basis $\{A_1,\ldots,A_s,B_1,\ldots,B_s,C_1,\ldots,C_{2t}\}$ such that the Lie bracket is expressed as $$\begin{aligned} \label{OT} [A_j,B_j] &=B_j,\nonumber \\ [A_j,C_{2k-1}] &=-\frac12 b_{jk}C_{2k-1}+c_{jk}C_{2k},\\ [A_j,C_{2k}] &=-c_{jk} C_{2k-1}-\frac12 b_{jk}C_{2k},\nonumber \end{aligned}$$ for any $j=1,\ldots,s$, $k=1,\ldots,t$. In terms of these bases, the complex structure is given by $$JA_j=B_j,\qquad JC_{2k-1}=C_{2k},$$ and a basis of $(1,0)$-forms is given by $\{\alpha_j+i\beta_j, \gamma_{2k-1}+i\gamma_{2k}\mid j=1,\ldots,s$, $k=1,\ldots,t\}$. For $t=1$, consider the closed $1$-form $\theta\in\g^*$ given by $\theta=\alpha_1+\cdots+\alpha_s$, and the $2$-form $\omega\in\alt^2\g^*$ given by $$\omega=2\sum_i \alpha_i\wedge\beta_i + \sum_{i\neq j}\alpha_i\wedge \beta_j + \gamma_1\wedge \gamma_2.$$ Then it can be verified that $d\omega=\theta\wedge\omega$ and, therefore, for $g=\omega(\cdot,J\cdot)$, we have that $(J,g)$ is a left invariant LCK structure on $G$. In [@S4] H. Sawai gives an explicit construction of a 6-dimensional OT solvmanifold carrying LCK metrics. He considers the $6$-dimensional group $G$ of matrices given by $$G=\left\{ \begin{pmatrix} \alpha_1^{t_1}{\alpha'}_1^{t_2} & 0 & 0 & 0 & x_1 \\ 0 & \alpha_2^{t_1}{\alpha'}_2^{t_2} & 0 & 0 & x_2 \\ 0 & 0 & \beta^{t_1}\beta'^{t_2} & 0 & z \\ 0 & 0 & 0 & \overline{\beta}^{t_1}\overline{\beta}'^{t_2} & \overline{z} \\ 0 & 0 & 0 & 0 & 1 \end{pmatrix}\, : t_1,t_2,x_1,x_2\in\R,\, z\in \C\right\},$$ where $\alpha_1,\alpha_2\in\R$ and $\beta,\overline{\beta}\in\C$ are the different roots of the polynomial $f_1(x) = x^4-2x^3-2x^2 + x + 1$, and $\alpha'_i=\alpha_i^{-1}+\alpha_i^{-2}$, $\beta'= \beta^{-1}+\beta^{-2}$. It can be seen that $\alpha'_1,\alpha'_2,\beta',\overline{\beta}'$ are the different roots of the polynomial $f_2(x) = x^4-4x^3 +4x^2-3x +1$. Sawai shows the existence of lattices in $G$, by carefully studying properties of the polynomials $f_1$ and $f_2$, which are the characteristic polynomials of the following unimodular matrices: $$B_1=\begin{pmatrix} 0 & 0 & 0 & -1 \\ 1 & 0 & 0 & -1 \\ 0 & 1 & 0 & 2 \\ 0 & 0 & 1 & 2 \end{pmatrix}, \quad B_2=\begin{pmatrix} 2 & 0 & 1 & 0 \\ 2 & 2 & 1 & 1 \\ -1 & 2 & 0 & 1 \\ 0 & -1 & 0 & 0 \end{pmatrix}.$$ Indeed, it is easy to see that $B_1B_2=B_2B_1$ and therefore these matrices are simultaneously diagonalizable over $\C$, and let $P\in \operatorname{GL}(4,\C)$ such that $PB_1P^{-1}$ and $PB_2P ^{-1}$ are both diagonal. Sawai then shows that $$\Gamma:=\left\{ \begin{pmatrix} PB_1^{s_1}B_2^{s_2}P^{-1} & P\begin{pmatrix} x_1\\x_2\\y_1\\y_2 \end{pmatrix} \\ 0 \quad 0 \quad 0 \quad 0 & 1 \end{pmatrix} \, : \, s_j,x_j,y_j\in \Z\right \}$$ is a lattice in $G$. It can be shown that the Lie brackets of the Lie algebra $\g$ of $G$ are isomorphic to the Lie brackets with $t=1$ for some constants $b_{jk},\,c_{jk}$, thus the solvmanifold $\Gamma\backslash G$ is in fact an OT solvmanifold carrying an invariant LCK structure. OT manifolds have been intensely studied lately; see for instance the survey [@OVu]. We summarize next some of their properties: - The LCK structure on an OT manifold with $s=2,t=1$ is a counterexample to a conjecture formulated by I. Vaisman, which stated that any compact LCK manifold has odd first Betti number. Indeed, it was shown in [@OT] that this 6-dimensional LCK manifold has first Betti number $b_1=2$ (more generally, $b_1=s$ for any OT manifold of type $(s,t)$). Using the description of the OT manifolds as solvmanifolds, Kasuya proved in [@K1] that the Betti numbers of an OT manifold of type $(s,1)$ are given by $b_p=b_{2s+2-p}=\binom{s}{p}$ for $1\leq p\leq s$ and $b_{s+1}=0$. Therefore, for even $s$ these manifolds also provide counterexamples to the Vaisman’s conjecture. - Kasuya proved in [@K] that OT manifolds do not admit any Vaisman metric, by studying the Morse-Novikov cohomology $H^*_\theta$ of solvmanifolds $\Gamma\backslash G$ equipped with LCS structures, where $\theta$ denotes the corresponding Lee form of the LCS structure. - Ornea and Verbitsky proved in [@OV5] that OT manifolds of type $(s,1)$ contain no non-trivial complex submanifolds. This generalizes the fact that Inoue surfaces carry no closed analytic curve. - In [@Z] the Chern-Ricci flow, an evolution equation of Hermitian metrics, is studied on a family of OT manifolds. It is shown that, after an initial conformal change, the flow converges, in the Gromov-Hausdorff sense, to a torus with a flat Riemannian metric determined by the OT-manifolds themselves. - In [@IO] both the de Rham cohomology and the Morse-Novikov cohomology of OT manifolds of general type are studied. In fact, the cohomology groups $H_{dR}^*(X(K,U))$ and $H_{\theta}^*(X(K,U))$ (for any closed 1-form $\theta$ on $X(K,U)$) are computed in terms of invariants associated to the background number field $K$. We would like to mention that OT manifolds are the only known examples of non-Vaisman LCK solvmanifolds in dimensions greater than $4$. It would be very interesting to find new examples. As mentioned previously, OT manifolds with $t=1$ admit LCK metrics. The natural question is to ask about the existence of such metrics on OT manifolds with $t>1$. Oeljeklaus and Toma proved already in [@OT Proposition 2.9] that an OT manifold with $s=1$ and $t\geq 2$ admits no LCK metric. Moreover, it follows from the proof of this result that if a LCK metric exists on a OT manifolds of type $(s,t)$ then the complex embeddings satisfy the condition $$\label{igual_modulo} |\sigma_{s+1}(u)|=\cdots =|\sigma_{s+t}(u)|, \quad \forall u \in U.$$ This fact was reobtained by V. Vuletescu in [@Vu], where he also proves that an OT manifold of type $(s,t)$ admits no LCK metric whenever $1<s\leq t$. Later, in an appendix of [@D] written by L. Battisti, it was shown that condition is also sufficient for the existence of a LCK metric. Therefore there is the following result: Let $X$ be an OT manifold of type $(s,t)$. Then $X$ admits a LCK metric if and only if holds. The proof of the sufficiency of condition follows by exhibiting a Kähler potential on $\mathbb{H}^s \times \C^t$, given by $$\varphi(z)= \left(\prod_{j=1}^s \frac{i}{z_j-\overline{z}_j} \right)^\frac{1}{t} +\sum_{k=1}^t |z_{s+k}|^2 .$$ The corresponding Kähler form $\omega:=i\partial \overline{\partial}\varphi$ on $\mathbb{H}^s \times \C^t$ gives rise to a LCK metric on the associated OT manifold. The main result of [@D] shows that if an OT manifold with $t\geq 2$ admits an LCK metric, then the natural numbers $s$ and $t$ satisfy a strong condition. Indeed, one has: If an OT manifold with $s\geq 1$, $t\geq 2$, admits an LCK metric then there exist integers $m\geq 0$ and $q\geq 2$ such that $$s = (2t + 2m)q - 2t.$$ In particular, $s$ is even and $s\geq 2t$. Recently, Istrati and Otiman determined in [@IO] the Betti numbers of OT manifolds carrying LCK metrics. Indeed, they show that the Betti numbers $b_j=\dim_\C H_{dR}^j(X,\C)$ of an OT manifold $X$ of type $(s,t)$ admitting some LCK metric are given by $$\begin{gathered} b_j=b_{2(s+t)-j}=\binom{s}{j}, \quad \text{if } 0\leq j\leq s,\\ b_j=0, \quad \text{if } s<j<s+2t. \end{gathered}$$ Regarding the Morse-Novikov cohomology of OT manifolds carrying LCK metrics, it is also shown in [@IO] that if $X=X(K,U)$ is an OT manifold of type $(s,t)$ then there exists at most one Lee class of a LCK metric. Furthermore, if $\theta$ is the Lee form of a LCK structure on $X$, then the corresponding twisted Betti numbers $ b_j^\theta= \dim_\C H_{\theta}^j(X,\C)$ are given by: $$b_j^\theta= t\binom{s}{j-2}, \qquad \text{for } 0\leq j\leq 2(s+t).$$ In particular, this shows again that OT manifolds do not admit Vaisman metrics, since $H^*_\theta(X)$ does not vanish (see [@LLMP]). Recalling the description of OT manifolds as solvmanifolds $\Gamma\backslash G$, one may wonder if the isomorphisms and still hold for this class of solvmanifolds, even if $G$ is not completely solvable. For the de Rham cohomology, it is shown in [@K1] that holds in the case of OT manifolds of type $(s,1)$. For the Morse-Novikov cohomology, it was shown in [@AOT] that the isomorphism holds for a special class of OT manifolds of type $(s,1)$, namely, those satisfying the *Mostow condition*. However, using the description of the Morse-Novikov cohomology of OT manifolds given in [@IO], Istrati and Otiman show that holds for any OT manifold of type $(s,1)$, even if the Mostow condition does not hold. Vaisman solvmanifolds ===================== In this section we will consider solvmanifolds equipped with invariant Vaisman structures, which will be called simply Vaisman solvmanifolds. Recall that a LCK structure where the Lee form is parallel with respect to the Levi-Civita connection of the Hermitian metric is called Vaisman. Firstly we state some structural results proved in [@AO2], and secondly we use these results in order to exhibit Vaisman solvmanifolds whose underlying complex manifold has holomorphically trivial canonical bundle. Structure results ----------------- In this section we present some results from [@AO2] concerning solvmanifolds equipped with invariant Vaisman structures. Since this structure is invariant, we may work at the Lie algebra level, and therefore we will consider a Vaisman structure $(J,\pint)$ on a Lie algebra $\g$, with associated Lee form $\theta$. We may assume that $|\theta|=1$. Therefore, if $A\in\g$ denotes the metric dual to $\theta$, we may decompose $\g$ as $\g=\R A\ltimes \ker\theta$, where $\theta(A)=1$ and $\ker\theta$ is an ideal since $d\theta=0$, which implies $\g'\subseteq \ker\theta$. The Vaisman condition $\nabla \theta=0$ is equivalent to $\nabla A=0$, and due to $d\theta=0$, this is in turn equivalent to $A$ being a Killing vector field (considered as a left invariant vector field on the associated Lie group with left invariant metric). Recalling that a left invariant vector field is Killing with respect to a left invariant metric if and only if the corresponding adjoint operator on the Lie algebra is skew-symmetric, we have: \[ad\_A-antisim\] Let $(J,\pint)$ be an LCK structure on $\g$ and let $A\in\g$ be as above. Then $(J,\pint)$ is Vaisman if and only if $\ad_A$ is a skew-symmetric endomorphism of $\g$. Other properties of Vaisman Lie algebras are given in the next result: \[vaisman-properties\] Let $(\g,J,\pint)$ be a Vaisman Lie algebra, then 1. $[A,JA]=0$, 2. $J\circ\ad_A=\ad_A\circ J$, 3. $J\circ\ad_{JA}=\ad_{JA}\circ J$, 4. $\ad_{JA}$ is skew-symmetric. Let us denote $\k:=\text{span}\{A,JA\}^\perp$, so that $\ker \theta= \R JA\oplus \k$ and $\g=\R A\oplus \R JA\oplus \k$. Set $\xi:=JA$, $\eta:=-J\theta|_{\ker\theta}$ and define an endomorphism $\phi\in\operatorname{End}(\ker\theta)$ by $\phi(a\xi+x)=Jx$ for $a\in\R$ and $x\in \k$. It follows from general results of I. Vaisman that the quadruple $(\pint|_{\ker\theta},\phi,\eta,\xi)$ determines a Sasakian structure on $\ker\theta$ (up to certain constants). Let us further assume that the Vaisman Lie algebra $\g$ is solvable and unimodular. In any unimodular LCK Lie algebra we have that the vector field $JA$ is in the commutator ideal $\g'$ of $\g$ (see [@AO2]). Therefore, since $\g$ is solvable, the endomorphism $\ad_{JA}$ is nilpotent. On the other hand, it follows from Proposition \[vaisman-properties\] that $\ad_{JA}$ is skew-symmetric. These two conditions imply that $\ad_{JA}=0$, that is, $JA$ is a *central* element of $\g$. Moreover, it was proved in [@AO2] that the dimension of the center of $\g$ is at most 2, since it is contained in $\text{span}\{A,JA\}$. It follows that the Sasakian Lie algebra $\ker \theta$ has a one-dimensional center. Using results from [@AFV], it follows that the $J$-invariant subspace $\k$ admits a Lie bracket $[\,,]_\k$ such that $(\k,[\,,]_\k,J|_\k,\pint|_\k)$ is a Kähler Lie algebra. Indeed, the Lie bracket $[\,,]_\k$ is defined as follows: for $x,y\in\k$, $[x,y]_\k$ denotes the component in $\k$ of $[x,y]$. Moreover, it is easy to show that we actually have $$\label{corchete_k} [x,y]= \omega(x,y) JA+[x,y]_\k, \quad x,y\in\k.$$ This means that $\ker\theta$ is isomorphic to the one-dimensional central extension of $(\k,[\,,]_\k)$ by the cocycle $\omega|_{\k}$, which coincides with the Kähler form of $(J|_\k,\pint|_\k)$. From now on, $\k$ will be considered as a Lie algebra with Lie bracket $[\,,]_\k$. It is easy to verify that $\k$ is also unimodular. Since $\k$ is Kähler, it follows from a classical result by Hano [@Hano] that the metric $\pint|_\k$ is *flat*. In what follows a Lie algebra equipped with a flat metric will be called a flat Lie algebra. We note moreover that if we denote $D:=\ad_A$, then $D$ satisfies: $D(\k)\subseteq \k$ and $D|_{\k}$ is a skew-symmetric derivation of $\k$ that commutes with $J|_{\k}$. This procedure can be reversed, and the main result in [@AO2] is: \[Vaisman-KF\] There is a one to one correspondence between unimodular solvable Lie algebras equipped with a Vaisman structure and pairs $(\mathfrak k, D)$ where $\mathfrak k$ is a Kähler flat Lie algebra and $D$ is a skew-symmetric derivation of $\mathfrak k$ which commutes with the complex structure. We continue our study of unimodular solvable Vaisman Lie algebras. Applying Milnor’s result on flat Lie algebras [@Mi], which was later refined in [@BDF], we obtain that $\k$ can be decomposed orthogonally as $\k=\z\oplus\h\oplus \k'$, where $\z$ denotes the center of $\k$, $\h$ is an abelian subalgebra and $\k'$ denotes the commutator ideal of $\k$, which happens to be abelian and even-dimensional. Moreover, the following conditions hold, where $\nabla$ denotes the Levi-Civita connection on $\k$, which is flat: 1. $\ad^\k :\h\to\mathfrak{so(k')}$ is injective, 2. $\ad^\k_x=\nabla_x$ for any $x\in\z\oplus\h$, 3. $\nabla_x=0$ if and only if $x\in\z\oplus\k'$. Furthermore, since $\k$ is Kähler, we have 1. $J(\z\oplus\h)=\z\oplus\h \quad \text{and} \quad J(\k')=\k'$, 2. $[H,Jx]_\k=J[H,x]_\k$ for any $H\in\h,\,x\in\k'$, i.e. $\ad^\k :\h\to\mathfrak{u(k')}$. It was proved in [@AO2] that any unitary derivation of a Kähler flat Lie algebra acts as the zero endomorphism on $\h + J\h$, therefore we obtain that $[A,H]=[A,JH]=0$ for all $H\in\h$. As a consequence of these results, we were able to prove that the class of unimodular solvable Lie algebras that admit Vaisman structures is quite restricted: \[Vaisman-imag.puras\] If the unimodular solvable Lie algebra $\g$ admits a Vaisman structure, then the eigenvalues of the operators $\ad_x$ with $x\in\g$ are all imaginary (some of them are 0). As a consequence of Theorem \[Vaisman-imag.puras\], together with results in [@S2] and Theorem \[nilpotent\], we reobtain the following result, proved by Sawai in [@S3]. Recall that a completely solvable Lie group $G$ is a solvable Lie group such that for all $x$ in its Lie algebra the operators $\ad_x$ have only real eigenvalues. Let $G$ be a simply connected completely solvable Lie group and $\Gamma\subset G$ a lattice. If the solvmanifold $\Gamma\backslash G$ admits a Vaisman structure $(J,g)$ such that the complex structure $J$ is induced by a left invariant complex structure on $G$, then $G=H_{2n+1}\times\R$, where $H_{2n+1}$ denotes the $(2n+1)$-dimensional Heisenberg Lie group. As stated above, the center of any unimodular solvable Lie algebra which admits a Vaisman structure is non trivial and, moreover, its dimension is at most 2. If we consider unimodular solvable Lie algebras with general LCK structures, there are examples with trivial center, but in all the examples known so far the dimension of the center is still bounded by 2. In [@MP] the authors construct examples of compact Vaisman manifolds which are obtained as the total spaces of a principal $S^1$-bundle over coKähler manifolds. They also show that these Vaisman manifolds are diffeomorphic to solvmanifolds. Yet another description of unimodular Vaisman Lie algebras was given in [@AHK]. Indeed, in that article it is shown that applying suitable modifications to any unimodular Vaisman Lie algebra one obtains one of the following Lie algebras: $\h_{2n+1}\times\R$, $\mathfrak{su}(2)\times \R$ or $\mathfrak{sl}(2,\R)\times\R$ (see [@AHK] for the relevant definitions). Canonical bundle of Vaisman solvmanifolds ----------------------------------------- In this section we present original results concerning the canonical bundle of Vaisman solvmanifolds. Indeed, using the description of unimodular solvable Lie algebras given in the previous section, we exhibit in Theorem \[canonical\] necessary and sufficient conditions for the existence of an *invariant* nowhere vanishing holomorphic $(n,0)$-form on such a solvmanifold, where $n$ is half its real dimension. As a consequence, in this case the canonical bundle of the complex solvmanifold is holomorphically trivial. We recall first that the complex manifold underlying the Vaisman nilmanifold $\Gamma\backslash H_{2n+1}\times S^1$, where $\Gamma$ is a lattice in the Heisenberg group $H_{2n+1}$, has holomorphically trivial canonical bundle. Indeed, it was shown in [@BDV] and [@CG] that any nilmanifold equipped with an invariant complex structure has trivial canonical bundle. A proof follows by using a special basis of left invariant $(1,0)$-forms on the nilmanifold determined by S. Salamon in [@Sal]. On the other hand, the main examples of Vaisman manifolds, that is, the Hopf manifolds $S^1\times S^{2n-1}$, have non trivial canonical bundle since its Kodaira dimension is $-\infty$. This follows from the fact that this Vaisman structure is toric (as shown in [@Pi Example 4.8]), together with the fact that any compact toric LCK manifold has Kodaira dimension $-\infty$ ([@MMP Theorem 6.1]). In what follows we will show that some Vaisman solvmanifolds have holomorphically trivial canonical bundle. Indeed, using the description of Vaisman Lie algebras given in the previous section, we will determine which of these Lie algebras admit a holomorphic $(n,0)$-form, where $2n$ is the real dimension of the Lie algebra. This holomorphic form will induce a nowhere vanishing section of the canonical bundle of any associated solvmanifold. Let $M=\Gamma\backslash G$ be a $2n$-dimensional solvmanifold equipped with an invariant Vaisman structure $(J,g)$. If $\g$ denotes the Lie algebra of $G$, which is unimodular and solvable, then $(J,g)$ determines a Vaisman structure on $\g$ with Lee form $\theta\in\g^\ast$. If $A\in\g$ denotes the metric dual to $\theta$, then we know that $\g=\R A\oplus \ker\theta$, $JA$ is a central element of $\g$, and the $J$-invariant subspace $\k=\text{span}\{A,JA\}^\perp\subset \ker\theta$ admits a Lie bracket that turns it into a Kähler Lie algebra with flat metric. Moreover, $\k$ can be decomposed orthogonally as $\k= \z\oplus \h\oplus \k'$ as before, where $\z\oplus \h$ and $\k'$ are $J$-invariant subspaces and $\ad_A$ vanishes on $\h+ J\h$. Recall that $\ad_A$ commutes with $J$ and is skew-symmetric, so that $\ad_A\in\u(\g)$; we also have that $\ad_H|_{\k'}\in\u(\k')$. \[canonical\] With notation as above, the $2n$-dimensional Vaisman solvmanifold $M=\Gamma\backslash G$ admits an invariant nowhere vanishing holomorphic $(n,0)$-form if and only if $$\ad_A\in\mathfrak{su}(\g), \quad \text{ and } \quad \ad_H|_{\k'}\in\mathfrak{su}(\k')$$ for all $H\in\h$. In particular, $M$ has holomorphically trivial canonical bundle. As $\ad_A=0$ on $\h+J\h$, we will decompose orthogonally the $J$-invariant subspace $\z\oplus \h$ as follows: $\z\oplus \h=(\z\cap J\z)\oplus (\h+J\h)$. According to [@AO2], there exists an orthonormal basis $\{e_1,\ldots,e_{2m}\}$ of $\k'$ such that $Je_{2i-1}=e_{2i}$ and for $A\in\g$ and any $H\in\h+J\h$ we have $$\ad_A|_{\k'}=\left(\begin{array}{ccccc} 0 & -a_1 & & & \\ a_1 & 0 & & & \\ & & \ddots & & \\ & & & 0 & -a_m\\ & & & a_m & 0 \\ \end{array}\right), \quad \ad_H|_{\k'}=\left(\begin{array}{ccccc} 0 & -\lambda_1(H) & & & \\ \lambda_1(H) & 0 & & & \\ & & \ddots & & \\ & & & 0 & -\lambda_m(H)\\ & & & \lambda_m(H) & 0 \\ \end{array}\right)$$ for some $a_i\in\R$ and $\lambda_i\in (\h+J\h)^*$, $i=1,\ldots,m$. Since $\ad_A|_{\z\cap J\z}$ is unitary, there exists an orthonormal basis $\{z_1,\ldots,z_{2r}\}$ of $\z\cap J\z$ such that $Jz_{2i-1}=z_{2i}$ and $$\ad_A|_{\z\cap J\z}=\left(\begin{array}{ccccc} 0 & -c_1 & & & \\ c_1 & 0 & & & \\ & & \ddots & & \\ & & & 0 & -c_r\\ & & & c_r & 0 \\ \end{array}\right),$$ for some $c_i\in\R$. Finally, we may choose an orthonormal basis $\{H_1,\ldots,H_{2s}\}$ of $\h+J\h$ such that $JH_{2i-1}=H_{2i}$, and we denote $\lambda_i^j=\lambda_i(H_j)$. To sum up, setting $B:=JA$ and taking into account, we may describe the Lie bracket of $\g$ as follows: $$\begin{gathered} [z_{2i-1},z_{2i}]=B, \quad [H_{2j-1},H_{2j}]=B, \quad [e_{2k-1},e_{2k}]=B, \\ [A,z_{2i-1}]=c_i z_{2i}, \quad [A, z_{2i}]=-c_i z_{2i-1},\\ [A,e_{2k-1}]=a_k e_{2k}, \quad [A, e_{2k}]=-a_k e_{2k-1},\\ [H_{j},e_{2k-1}]=\lambda_k^{j} e_{2k}, \quad [H_{j},e_{2k}]=-\lambda_k^{j} e_{2k-1}, $$ for $i=1,\ldots,r$, $j=1,\ldots,2s$ and $k=1,\ldots,m$. If $\{\alpha, \beta, z^1,\ldots,z^{2r}, H^1,\ldots,H^{2s}, e^1,\ldots,e^{2m}\}$ denotes the dual basis of $\{A, B\}\cup\{z_i\}\cup\{ H_j\}\cup\{e_k\}$, then the Lie bracket can be encoded in terms of their differentials as follows: $$\begin{aligned} d\alpha & =0,\\ d\beta & = -\sum_{i=1}^r z^{2i-1}\wedge z^{2i} -\sum_{j=1}^s H^{2j-1}\wedge H^{2j}-\sum_{k=1}^m e^{2k-1}\wedge e^{2k} ,\\ dz^{2i-1} & = c_i\alpha\wedge z^{2i},\\ dz^{2i} & = -c_i\alpha\wedge z^{2i-1},\\ dH^{2j-1} & = 0,\\ dH^{2j} & = 0,\\ de^{2k-1} & = \gamma_k\wedge e^{2k},\\ de^{2k} & = -\gamma_k\wedge e^{2k-1}, \end{aligned}$$ where $\gamma_k = a_k\alpha + \sum_{l=1}^s\lambda_k^l H^l$ for any $k=1,\ldots,m$. Let us define the complex $1$-forms $$\begin{aligned} \mu & = \alpha + i\beta ,\\ \xi_j & = z^{2j-1} + iz^{2j},\\ \delta_j & = H^{2j-1} + iH^{2j},\\ \omega_j & = e^{2j-1} + ie^{2j}. \end{aligned}$$ These 1-forms are all of type $(1,0)$, so that $$\eta:=\mu\wedge\xi_1\wedge\ldots\wedge\xi_r\wedge\delta_1\wedge\ldots\wedge\delta_s\wedge\omega_1\wedge\ldots\wedge\omega_m$$ is a $(n,0)$-form which generates $\alt^{(n,0)}\g^*$, where $n=1+r+s+m$. Therefore, we only have to check when $\eta$ is holomorphic or, equivalently, $\eta$ is closed. It can be easily verified that $d\eta=0$ if and only if $$\label{sumas} \sum_{j=1}^r c_j + \sum_{j=1}^m a_j=0 \quad \text{and} \quad \sum_{j=1}^m \lambda_j^k=0 \quad \forall k=1,\ldots,s.$$ It is easy to see that equations hold if and only if $\ad_A\in \mathfrak{su}(\g)$ and $\ad_H|_{\k'}\in \mathfrak{su}(\k')$ for all $H\in\h$ (using that $\h+J\h \subseteq \h\oplus\z$), and this completes the proof. When the $2n$-dimensional unimodular solvable Vaisman Lie algebra $\g=Lie (G)$ admits a non zero holomorphic $(n,0)$-form, then any solvmanifold $\Gamma\backslash G$ has holomorphically trivial canonical bundle, that is, this property is independent of the choice of lattice. \[fibrado\] It is well known that the only compact complex surfaces with holomorphically trivial canonical bundle are the $K3$ surfaces and the primary Kodaira surfaces. Moreover, the former do not admit Vaisman metrics and are not solvmanifolds, whereas the latter do admit Vaisman metrics and are in fact nilmanifolds. Next, using Theorem \[canonical\], we will exhibit examples of Vaisman solvmanifolds with holomorphically trivial canonical bundle in any dimension greater than or equal to $6$, which are not nilmanifolds. Indeed, in [@AO2] we determined all Vaisman Lie algebras $\g$ such that the associated Kähler flat Lie algebra $\k$ is abelian, $\k=\R^{2n-2}$. The Vaisman Lie algebra $\g$ can be decomposed as $\g=\R A \ltimes_D \h_{2n-1}$ where the adjoint action of $\R$ on $\h_{2n-1}$ is given by $$\label{D} D=\left(\begin{array}{cccccc} 0 & & & & & \\ & 0 & -a_1 & & & \\ & a_1 & 0 & & & \\ & & & \ddots & & \\ & & & & 0 & -a_{n-1}\\ & & & & a_{n-1} & 0 \\ \end{array}\right),$$ in a basis $\{B=JA,e_1,\dots,e_{2n-2}\}$ of $\h_{2n-1}$ such that $[e_{2i-1},e_{2i}]=B$ and $Je_{2i-1}=e_{2i}$, for some $a_i\in\R$ (and not all of them zero, otherwise we obtain a nilpotent Lie algebra). We proved that whenever $a_i\in\Q$ the corresponding simply connected Lie group admits non-nilpotent lattices. It follows from Theorem \[canonical\] that if the constants $\{a_i\}$ satisfy $\sum_{i=1}^{n-1} a_i=0$, then the associated solvmanifolds have holomorphically trivial canonical bundle and they are not diffeomorphic to a nilmanifold. For instance, when $n=3$, $a_1=1$ and $a_2=-1$, we reobtain the Lie algebra $\g_6=A_{6,82}^{0,1,1}$ from [@FOU], where the classification of $6$-dimensional solvable Lie algebras equipped with a non-zero holomorphic $(3,0)$-form was given. If in the construction given in Example \[fibrado\] we take $n=1$ and $a_1=1$, the associated Lie group admits a lattice such that the corresponding $4$-dimensional solvmanifold is a secondary Kodaira surface. Therefore this secondary Kodaira surface is a Vaisman solvmanifold and it is well known that its canonical bundle is not holomorphically trivial (see for instance [@BHPV]). We recall that, according to [@To], a $2n$-dimensional compact complex manifold has holomorphically trivial canonical bundle if and only if it admits a Hermitian metric whose Chern connection has (unrestricted) holonomy group contained in $SU(n)$. Locally conformally symplectic solvmanifolds ============================================= A natural generalization of LCK manifolds are the locally conformally symplectic (LCS) manifolds, that is, a manifold $M$ with a non degenerate $2$-form $\omega$ such that there exists an open cover $\{U_i\}$ and smooth functions $f_i$ on $U_i$ such that $$\omega_i=\exp(-f_i)\omega$$ is a symplectic form on $U_i$. This condition is equivalent to requiring that $$\label{lcs} d\omega=\theta\wedge\omega$$ for some closed $1$-form $\theta$, called the Lee form. Moreover, $M$ is called globally conformally symplectic (GCS) if there exist a $C^{\infty}$ function, $f:M\to\R$, such that $\exp(-f)\omega$ is a symplectic form. Equivalently, $M$ is a GCS manifold if there exists a exact $1$-form $\theta$ globally defined on $M$ such that $d\omega=\theta\wedge\omega$. The pair $(\omega, \theta)$ will be called a LCS structure on $M$. Note that if $\theta=0$ then $(M,\omega)$ is a symplectic manifold. The unique vector field $V$ on $M$ satisfying $i_V\omega=\theta$ is called the Lee vector field of the LCS structure. These manifolds were considered by Lee in [@L] and they have been thoroughly studied by Vaisman in [@V]. Some recent results can be found in [@Ba; @Baz1; @BM1; @Ha; @HR; @LV], among others. LCS manifolds play an important role in mathematics and in physics as well, for example in Hamiltonian mechanics, since they provide a generalization of the usual description of the phase space in terms of symplectic geometry (see [@V]). We recall a definition due to Vaisman ([@V]) concerning different kinds of LCS structures. Given a LCS structure $(\omega, \theta)$ on $M$, a vector field $X$ is called an infinitesimal automorphism of $(\omega,\theta)$ if $\textrm{L}_X\omega=0$, where $\textrm{L}$ denotes the Lie derivative. It follows that $\textrm{L}_X\theta=0$, as a consequence, $\theta(X)$ is a constant function on $M$. The subspace $\mathfrak{X}_\omega(M)=\{X\in\mathfrak{X}(M): \textrm{L}_X\omega=0\}$ is a Lie subalgebra of $\mathfrak{X}(M)$, and the map $\theta : \mathfrak{X}_\omega(M) \to \R$ is a well defined Lie algebra morphism called the Lee morphism. If there exists an infinitesimal automorphism $X$ such that $\theta(X)\neq 0$, the LCS structure $(\omega,\theta)$ is said to be of [*the first kind*]{}, and it is of [*the second kind*]{} otherwise. Much more is known about manifolds carrying LCS structures of the first kind. For instance, in [@V] it was shown the existence of distinguished foliations on such manifolds, while in [@Ba] it was shown that a compact manifold equipped with a LCS structure of the first kind fibers over the circle and that the fiber inherits a contact structure. It is well known that the LCS structure underlying any Vaisman manifold is of the first kind (see for instance [@CDMY1]). The relation between LCK and LCS manifolds is analogous to the Kähler vs symplectic manifolds. In [@OV] the following problem was formulated: “Give examples of compact locally conformally symplectic manifolds which admit no locally conformally Kähler metric.” The first example was given in [@BK]; this example is a product $M\times S^1$, where $M$ is certain $3$-dimensional compact contact manifold. More recently, in [@BM1] another example was constructed, this example is a $4$-dimensional nilmanifold and it is not a product of a $3$-dimensional compact manifold with $S^1$. This motivates the study of nilmanifolds or, more generally, solvmanifolds, equipped with invariant LCS structures. Given a Lie group $G$, a LCS structure on $G$ is called left invariant if $\omega$ is left invariant, and this easily implies that $\theta$ is also left invariant using condition . This fact allows us to define a LCS structure on a Lie algebra. We say that a Lie algebra $\g$ admits a *locally conformally symplectic* (LCS) structure if there exist $\omega\in\alt^2\g^*$ and $\theta \in \g^*$, with $\omega$ non degenerate and $\theta$ closed, such that is satisfied. LCS structures on Lie algebras can also be distinguished in two different kinds, just as LCS structures on manifolds. Indeed, given a LCS structure $(\omega,\theta)$ on a Lie algebra $\g$, let us denote by $\g_\omega$ the set of infinitesimal automorphisms of the LCS structure, that is, $$\label{autom} \g_\omega = \{x\in\g: \textrm{L}_x\omega=0\} = \{x\in\g: \omega([x,y],z)+\omega(y,[x,z])=0 \; \text{for all} \; y,z\in\g\}.$$ Note that $\g_\omega \subset \g$ is a Lie subalgebra, and the restriction of $\theta$ to $\g_\omega$ is a Lie algebra morphism called the Lee morphism. The LCS structure $ (\omega,\theta)$ is said to be *of the first kind* if the Lee morphism is surjective, and *of the second kind* if it is identically zero. Note that a LCS structure of the first kind on a Lie algebra naturally induces a LCS structure of the first kind on any quotient of the associated simply connected Lie group by a discrete subgroup. There is another way to distinguish LCS structures. Indeed, a LCS structure $(\omega,\theta)$ on a Lie algebra $\g$ is said to be *exact* if $[\omega]_\theta =0$, and *non exact* if $[\omega]_\theta\neq0$. It was shown in [@BM] that on any unimodular Lie algebra a LCS structure is of the first kind if and only if it is exact. Recently, in [@BM] and [@ABP] many interesting results on LCS structures of the first kind on Lie algebras and solvmanifolds were established. We review some of them next. Firstly, in [@BM] it is shown a relation between Lie algebras equipped with LCS structures of the first kind and Lie algebras with contact structures. Recall that a $1$-form $\eta$ on a $(2n+1)$-dimensional Lie algebra $\h$ is called a contact form if $\eta\wedge (d\eta)^n\neq0$. This form $\eta$ is called the contact form and the unique vector $V\in\h$ satisfying $\eta(V)=1$ and $i_V d\eta=0$ is called the Reeb vector of the contact structure. \[lcs\_contact\] There is a one to one correspondence between $(2n+1)$-dimensional contact Lie algebras $(\h,\eta)$ endowed with a derivation $D$ such that $\eta\circ D=0$ and $(2n+2)$-dimensional LCS Lie algebras of the first kind $(\g,\omega,\theta)$. The relation is given by $\h=\ker\theta$, $\omega=d_\theta \eta$ and $D=\ad_U$ where $U$ satisfies $\eta=-i_U\omega$. When the Lee vector is in the center of the Lie algebra, a more refined result can be obtained, relating the LCS structure with a symplectic structure on a lower dimensional Lie algebra (cf. Theorem \[Vaisman-KF\]): \[lcs\_symplectic\] There is a one to one correspondence between Lie algebras of dimension $2n +2$ admitting a LCS structure of the first kind with central Lee vector and symplectic Lie algebras $(\mathfrak{s}, \beta)$ of dimension $2n$ endowed with a derivation $E$ such that $\beta(EX,Y)+\beta(X,EY)=0$ for all $X,Y\in\mathfrak s$. Considering LCS structures on nilpotent Lie algebras, it was shown in [@BM] that any such structure is of the first kind and, moreover, the Lee vector is central. Combining this with the previous theorem the next result follows: Every LCS nilpotent Lie algebra of dimension $(2n+2)$ with non-zero Lee form is a suitable extension of a $2n$-dimensional symplectic nilpotent Lie algebra $\mathfrak{s}$ equipped with a symplectic nilpotent derivation. Moreover, the symplectic nilpotent Lie algebra $\mathfrak s$ may obtained by a sequence of $(n-1)$ symplectic double extensions by nilpotent derivations from the abelian Lie algebra of dimension $2$, according to [@MR; @DM]. On the other hand, LCS structures of the second kind are less understood. Examples of LCS structures of the second kind on the $4$-dimensional solvmanifold constructed in [@ACFM] are given in [@Ba2]. More precisely, these LCS structures are non exact. In [@ABP] properties of non exact LCS structures were investigated, producing new examples. LCS structures on almost abelian Lie algebras were studied in [@AO1] and, among other results, it was proved that on any such Lie algebra of dimension at least $6$ any LCS structure is of the second kind. Moreover, for each $n\geq 2$ a $2n$-dimensional simply connected almost abelian Lie group with left invariant LCS structures of the second kind are constructed; these groups admit lattices and we obtain in this way solvmanifolds with LCS structures, moreover, it can be seen that these structures are of the second kind. In [@ABP] a classification of LCS structures on $4$-dimensional Lie algebras, up to Lie algebra automorphism, is given and LCS structures on compact quotients of all $4$-dimensional simply connected solvable Lie groups are constructed. They also give structure results for LCS Lie algebras inspired by the results of [@Ov] and [@BM]. The first result is about exact lcs structures, not necessarily of the first kind. [@ABP Theorem 1.4] There is a one-to-one correspondence between exact LCS Lie algebras $(\g,\omega=d\eta-\theta\wedge\eta,\theta)$, $\dim \g=2n$, such that $\theta(U)\neq0$, where $\eta=-i_U\omega$ and contact Lie algebras $(\h,\eta)$, $\dim \h=2n-1$, with a derivation $D$ such that $D^*\eta=\alpha\eta$, $\alpha\neq1$. The second structure result is about Lie algebras with a non exact LCS structure, establishing a relation with *cosymplectic* Lie algebras. We recall that a $(2n-1)$-dimensional Lie algebra $\h$ is said to be cosymplectic if there exist two closed $1$-forms $\eta,\beta\in\h^*$ such that $\eta\wedge \beta^{n-1}\neq 0$. [@ABP Proposition 1.8] Let $(\h,\eta,\beta)$ be a cosymplectic Lie algebra of dimension $2n-1$, endowed with a derivation $D$ such that $D^*\beta=\alpha\beta$ for some $\alpha\neq0$. Then $\g=\h\ltimes_D \R$ admits a natural LCS structure. The Lie algebra $\g$ is unimodular if and only if $\h$ is unimodular and $D^*\eta=-\alpha(n-1)\eta+\delta$ for some $\delta\in \h^*$ with $\delta(R)=0$ where $R$ denotes the Reeb vector of the cosymplectic structure. If $\h$ is unimodular then the LCS structure on $\g$ is not exact. The third result is related with the cotangent extension problem in the LCS setting; the symplectic case was studied in [@Ov]. [@ABP Proposition 1.17] Let $(\g,\omega,\theta)$ be a $2n$-dimensional LCS Lie algebra with a Lagrangian ideal $j\subset \ker\theta$. Then $\g$ is a solution of the cotangent extension problem, that is, there exists a $n$-dimensional Lie algebra $\h$ with a closed 1-form $\mu\in\h^*$ such that $\g\cong \h^*\oplus\h$, and $(\omega,\theta)$ is equivalent to a canonical LCS structure on $\h^*\oplus\h$ constructed using $\mu$. Furthermore, in [@ABP] it is shown that every LCS structure on a $4$-dimensional Lie algebra can be recovered by one of the three constructions detailed above. [99]{}, Homogeneous locally conformally Kähler and Sasaki manifolds. *Internat. J. Math.* [**6**]{} (2015), 1541001 (29 pp). , Homogeneous Sasaki and Vaisman manifolds of unimodular Lie groups, preprint 2018, arXiv:1810.01095. , Abelian Hermitian geometry, *Differential Geom. Appl*. **30** (2012), 509–519. , A class of Sasakian 5-manifolds, *Transform. Groups* **14** (2009), 493–512. , Locally conformally Kähler structures on unimodular Lie groups, *Geom. Dedicata* **179** (2015), 197–216. , Lattices in almost abelian Lie groups with locally conformal Kähler or symplectic structures, *Manuscripta Math.* **155** (2018), 389–417. , Vaisman solvmanifolds and relations with other geometric structures, preprint 2017, arXiv:1709.01567. , Abelian balanced Hermitian structures on unimodular Lie algebras, *Transform. Groups*, **21** (2016), 903–927. , Examples of four dimensional locally conformal Kähler solvmanifolds, *Geom. Dedicata* **29** (1989), 227–232. , Structure of locally conformally symplectic Lie algebras and solvmanifolds, to appear in *Ann. Sc. Norm. Super. Pisa Cl. Sci. (5)*, arXiv:1704.01197. , Locally conformally Kähler structures on four-dimensional solvable Lie algebras, preprint 2018, arXiv:1809.08149. , Cohomologies of locally conformally symplectic manifolds and solvmanifolds, *Ann. Global Anal. Geom,* **53** (2018), 67–96. , Moser stability for locally conformally symplectic structures, *Proc. Amer. Math. Soc.* [**137**]{} (2009), 2419–2424. , On the geometry of locally conformal symplectic manifolds, in *Infinite dimensional Lie groups in geometry and representation theory*, World Sci. Publ., River Edge, NJ, (2002), 79–91. , Examples of non $d_\theta$-exact locally conformal symplectic forms, *J. Geom.* **87** (2007), 1–13. , Abelian complex structures on solvable Lie algebras, *J. Lie Theory* **14** (2004), 25–34. Hyper-Kähler quotients of solvable Lie groups, *J. Geom. Phys.* **56** (2006), 691–711. , Canonical bundles of complex nilmanifolds, with applications to hypercomplex geometry, *Math. Res. Lett.* **16** (2009), 331–347. , *Compact complex surfaces*. Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics, **4**, Springer-Verlag Berlin Heidelberg, 2004. , Vaisman nilmanifolds, *Bull. Lond. Math. Soc.* **49** (2017), 824–830. , Locally conformally symplectic and Kähler geometry, *EMS Surv. Math. Sci.* **5** (2018), 129–154. , Locally conformal symplectic nilmanifolds with no locally conformal Kähler metrics, *Complex Manifolds* **4** (2017), 172–178. , On locally conformal symplectic manifolds of the first kind, *Bull. Sci. Math.* **143** (2018), 1–57. , On the metric structure of non-Kähler complex surfaces, *Math. Ann.* **317** (2000), 1–40. , Kähler and symplectic structure on nilmanifolds, *Topology* **27** (1988), 513–518. , On low-dimensional solvmanifolds, *Asian J. Math.* **20** (2016), 199–262. , Sasakian nilmanifolds, *Int. Math. Res. Not. IMRN* **2015** (2015), 6648–6660. , Hard Lefschetz theorem for Vaisman manifolds, to appear in *Trans. Amer. Math. Soc.*, arXiv:1510.04946. , Generalized complex structures on nilmanifolds, *J. Symplectic Geom.* **2** (2004), 393–410. , Dolbeault cohomology of compact nilmanifolds, *Transform. Groups* **6** (2001), 111–124. , Stability of abelian complex structures, *Internat. J. Math.* **17** (2006), 401–416. , Compact nilmanifolds with nilpotent complex structures: Dolbeault cohomology, *Trans. Amer. Math. Soc.* **352** (2000), 5405–5433. , Compact locally conformal Kähler nilmanifolds, *Geom. Dedicata* **21** (1986), 187–192. , Double extension symplectique d’un groupe de Lie symplectique, *Adv. Math.* **117** (1996) 208–227. , Hyper-Kähler torsion structures invariant by nilpotent Lie groups, *Classical Quantum Gravity* **19** (2002), 551–562. , *Locally conformal Kähler geometry*. Progress in Mathematics **155**, Birkhäuser Boston, 1998. , Nonreciprocal units in a number field with an application to Oeljeklaus-Toma manifolds, *New York J. Math.* **20** (2014), 257–274. , Six-dimensional solvmanifolds with holomorphically trivial canonical bundle, *Int. Math. Res. Not. IMRN* **2015**, no.24 (2015), 13757–13799. , Dolbeault cohomology of complex nilmanifolds foliated in toroidal groups, preprint 2018, arXiv:1808.08090. , La $1$-forme de torsion d’une variété hermitienne compacte, *Math. Ann.* **267** (1984), 495–518. , Compact homogeneous lcK manifolds are Vaisman, *Math. Ann.* **361** (2015), 1043–1048. , Geometry of Hyper-Kähler connections with torsion, *Comm. Math. Phys.* **213** (2000), 19–37. , The sixteen classes of almost Hermitian manifolds and their linear invariants, *Ann. Mat. Pura Appl.* **123** (1980), 35–58. , Géométrie des algébres de Lie locales de Kirillov, *J. Math. Pures Appl.* [**63**]{} (1984), 407–484. , Some properties of locally conformal symplectic manifolds, in *Infinite dimensional Lie groups in geometry and representation theory*, World Sci. Publ., River Edge, NJ, (2002), 92–104. , Reduction for locally conformal symplectic manifolds, *J. Geom. Phys.* **37** (2001), 262–271. , On Kaehlerian homogeneous spaces of unimodular Lie groups, *Amer. J. Math.* **79** (1957), 885–900. , Complex and Kähler structures on compact solvmanifolds, *J. Symplectic Geom.* **3** (2005), 749–767. , A note on compact solvmanifolds with Kähler structures, *Osaka J. Math.* **43** (2006), 131–135. , Locally conformally Kähler structures on homogeneous spaces, in *Geometry and Analysis on Manifolds*, Progress in Mathematics, vol. 308. Birkhäuser, Cham (2015), 353–372. , Compact homogeneous locally conformally Kähler manifolds, *Osaka J. Math.* **53** (2016), 683–703. , Spectral sequence in the de Rham cohomology of fibre bundles, *J. Fac. Sci. Univ. Tokyo Sect. I* **8** (1960), 289–331. , De Rham and twisted cohomology of Oeljeklaus-Toma manifolds, to appear in *Ann. Inst. Fourier (Grenoble)*, arXiv:1711.07847. , Note on locally conformal Kähler surfaces, *Geom. Dedicata* **84** (2001), 115–124. , On harmonic forms in compact locally conformal Kähler manifolds with the parallel Lee form, *Ann. Fac. Sci. Univ. Nat. Zaïre (Kinshasa) Sect. Math.-Phys.* **6** (1980), 17–29. , Vaisman metrics on solvmanifolds and Oeljeklaus-Toma manifolds, *Bull. Lond. Math. Soc.* **45** (2013), 15–26. , Minimal models, formality, and hard Lefschetz properties of solvmanifolds with local systems, *J. Differential Geom.* **93** (2013), 269–297. , On the structure of compact complex analytic surfaces I, *Amer. J. Math.* **86** (1964), 751–798. , Cohomology theories on locally conformal symplectic manifolds, *Asian J. Math.* **19**, (2015), 45–82. , A kind of even dimensional differential geometry and its application to exterior calculus, *Amer. J. Math.* **65** (1943), 433–438. , On the computation of the Lichnerowicz-Jacobi cohomology, *J. Geom. Phys.* **44** (2003), 507–522. , Deformations of 2-step nilmanifolds with abelian complex structures, *J. Lond. Math. Soc.* **73** (2006), 173–193. , On toric locally conformally Kähler manifolds, *Ann. Global Anal. Geom.* **51** (2017), 401–417. , On solvable Lie algebras, *Bull. Acad. Sci. URSS. Sér. Math. \[Izvestia Akad. Nauk SSSR\]* **9** (1945), 329–356. , Compact generalized Hopf and cosymplectic solvmanifolds and the Heisenberg group $H(n,1)$, *Israel J. Math.* **101** (1997), 189–204. , Groupes de Lie à structure symplectique invariante, in *Symplectic Geometry, Groupoids, and Integrable Systems*, Math. Sci. Res. Inst. Publ., vol. 20, Springer, New York. (1991), 247–266. , Cohomology of solvable Lie algebras and solvmanifolds, *Math. Notes* [**77**]{} (2005), 61–71. , Curvatures of left invariant metrics on Lie groups, *Adv. Math.* **21** (1976), 293–329. , Factor spaces of solvable groups, *Ann. of Math.* **60** (1954), 1–27. , On the cohomology of homogeneous spaces of nilpotent Lie groups, *Ann. of Math.* **59** (1954), 531–538. , Non-Kähler compact complex manifolds associated to number fields, *Ann. Inst. Fourier (Grenoble)* **55** (2005), 161–171. , Locally conformal Kähler manifolds. A selection of results, *Lect. Notes Semin. Interdiscip. Mat.* **IV** (2005), 121–152. , Locally conformal Kähler manifolds with potential, *Math. Ann.* **348** (2010), 25–33. , A report on locally conformally Kähler manifolds, in *Harmonic maps and differential geometry*, Contemp. Math. **542** (2011), 135–149. , Structure theorem for compact Vaisman manifolds, *Math. Res. Lett.* **10** (2003), 799–805. , Oeljeklaus-Toma manifolds admitting no complex subvarieties, *Math. Res. Lett.* **18** (2011), 747–754. , LCK rank of locally conformally Kähler manifolds with potential, *J. Geom. Phys.* **107** (2016), 92–98. , Oeljeklaus-Toma manifolds and locally conformally Kähler metrics. A state of the art, *Stud. Univ. Babeş-Bolyai Math.* **58** (2013), 459–468. , Complex, symplectic and Kähler structures on four dimensional Lie groups, *Rev. Un. Mat. Arg.* **45** (2004), 55–68. , Torus bundles over a torus, *Proc. Amer. Math. Soc.* **12** (1961), 26–29. , Toric Vaisman manifolds, *J. Geom. Phys.* **107** (2016), 149–161. , *Discrete subgroups of Lie groups*, Springer, Berlin, 1972. , Dolbeault cohomology of nilmanifolds with left-invariant complex structure, in *Complex and Differential Geometry*, Springer Proceedings in Mathematics, vol. 8, Springer Berlin Heidelberg, (2011), 369–392. , Complex structures on nilpotent Lie algebras, *J. Pure Appl. Algebra* **157** (2001), 311–333. , Locally conformal Kähler structures on compact nilmanifold with left-invariant complex structures, *Geom. Dedicata* **125** (2007), 93–101. , Locally conformal Kähler structures on compact solvmanifolds, *Osaka J. Math.* **49** (2012), 1087–1102. , Vaisman structures on compact solvmanifolds, *Geom. Dedicata* **178** (2015), 389–404. , Structure theorem for Vaisman completely solvable solvmanifolds, *J. Geom. Phys.* **114** (2017), 581–586. , Example of a six-dimensional LCK solvmanifold, *Complex Manifolds* **4** (2017), 37–42. , Some simple examples of symplectic manifolds, *Proc. Amer. Math. Soc.* **55** (1976), 467–468. , Non-Kähler Calabi-Yau manifolds, in *Analysis, complex geometry, and mathematical physics: in honor of Duong H. Phong*, Contemp. Math. vol. 644, Amer. Math. Soc., Providence, RI, (2015), 261–277. , Some examples of locally conformal Kähler manifolds, *Rend. Semin. Mat. Univ. Politec. Torino* **40** (1982), 81–92. , Hermitian structures on six-dimensional nilmanifolds, *Transform. Groups* **12** (2007), 175–202. , Balanced Hermitian geometry on 6-dimensional nilmanifolds, *Forum Math.* **27** (2015), 1025–1070. , On locally conformal almost Kähler manifolds, *Israel J. Math.* **24** (1976), 338–351. , Locally conformal Kähler manifolds with parallel Lee form, *Rend. Mat., VI Ser.* **12** (1979), 263–284. , Generalized Hopf manifolds, *Geom. Dedicata* **13** (1982), 231–255. , Locally conformal symplectic manifolds, *Internat. J. Math. Math. Sci.* [**12**]{} (1985), 521–536. , Theorems on the vanishing of cohomology for locally conformally hyper-Kähler manifolds, *Proc. Steklov Inst. Math.* **246** (2004), 54–78. , Blowing-up points on l.c.K. manifolds, *Bull. Math. Soc. Sci. Math. Roumanie (N.S.)* **52 (100)** (2009), 387–390. , LCK metrics on Oeljeklaus-Toma manifolds versus Kronecker’s theorem, *Bull. Math. Soc. Sci. Math. Roumanie (N.S.)* **57 (105)** (2014), 225–231. , Superrigidity of lattices in solvable Lie groups, *Invent. Math.* [**122**]{} (1995), 147–193. , The Chern-Ricci flow on Oeljeklaus-Toma manifolds, *Canad. J. Math.* **69** (2017), 220–240. [^1]: A $\g$-module $V$ is called triangular if the endomorphisms of $V$ defined by $v\mapsto Xv$ have only real eigenvalues for any $X\in\g$.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Fluctuation Theorem(FT) has been studied as far from equilibrium theorem, which relates the symmetry of entropy production. To investigate the application of this theorem, especially to biological physics, we consider the FT for tilted rachet system. Under, natural assumption, FT for steady state is derived.' address: 'Department of Applied Physics ,Waseda University,3-4-1 Okubo, Shinjuku-ku,Tokyo 169-8555,Japan ' author: - 'T. Monnai' title: Fluctuation Theorem in Rachet System --- Fluctuation theorem ,Rachet ,random walk 05.10.Gg ,05.40.Jc ,05.60.Cd INTRODUCTION ============ Since the discovery by Evans.et.al.[@EvansCohen], Fluctuation Theorem(FT) has been studied in many situations, both stochastic[@Crooks],[@Kurchan],[@Spohn1999] and deterministic[@EvansCohen][@Jarzynski], though systems and interpretations differ, FT has a universal form : $$\frac{Prob(\Delta S=A)}{Prob(\Delta S=-A)}\simeq e^{\Delta S}$$ here, $\Delta S$ is the entropy generated, and $Prob(\Delta S =A)$ is the probability for $\Delta S =A$. It is interesting that FT kind relation universally holds, but its direct experimental application is not known.(Though Jarzynski equality[@JarzynskiEq] is closely related to FT and has been used in the experiment, this equality is not FT itself, in this sense application of Jarzynsky equality is not direct.) So, we would like to derive FT for rachet system, because it relates biological physics, and direct application of FT is expected. The analysis is based on Langevin treatment. Kurchan showed that for Langevin system, if the initial state is Gibbsian, FT holds generally[@Kurchan]. Though we consider Langevin treatment, the result does not depend on the initial condition and in this sense our interest is different from that of Kurchan. RACHET SYSTEM ============= In this paper we consider the rachet system, namely a particle on a periodic potential $V(x+L)=V(x)$ is dragged by a constant load force $F$. Then, the effective potential the particle feels is $V_{eff}\equiv V(x)-x F$. If, the potential barrier is high enough, the particle can be considerd to be almost at equilibrium and may satisfy the detailed balance relation. Indeed, the quantitative argument is given in the review by Reimann[@ReimannP]. We summarize the work by Reimann. We consider a system that a particle on a periodic potential(period $L$) is dragged by a constant load force $F$. The motion of this particle can be described by the Langevin eq. : $$\begin{aligned} \eta\frac{d}{dt}x(t)=&&-\frac{d}{dx}V_{eff}(x)+\xi(t) \nonumber \\ V_{eff}(x)\equiv&&V(x)-F x \nonumber \\ \left\langle\xi(t) \xi(s)\right\rangle =&&\sqrt{2 \eta k_B T}\delta(t-s) \nonumber\end{aligned}$$ The corresponding Fokker-Planck equation becomes : $$\frac{\partial}{\partial t} P(x,t) = \frac{\partial}{\partial x} (\frac{dV_{eff}}{d x}\frac{1}{\eta}+\frac{k_B T}{\eta}\frac{\partial}{\partial x} ) P(x,t)= -\frac{\partial }{\partial x}J(x,t)$$ Here, $J(x,t)$ is the probability current. Then at a steady state, the reduced probability current : $$\hat{J}(x,t)\equiv \Sigma_{n=-\infty}^{\infty}J(x+n L,t)$$ becomes : $$\begin{aligned} \hat{J}^{st}=&&N (1-e^{-\frac{F L}{k_B T}}) \nonumber \\ N=&&\frac{k_B T}{\eta}(\int_0^L dx \int_x^{x+L}dy e^{\frac{V_{eff}(y)-V_{eff}(x)}{k_B T}})^{-1}\end{aligned}$$ If in each period $L$ there exists one minimum $x_{min}$ and one maximum $x_{max}$, and the system is in the weak noise regime $k_B T\ll \Delta V_{eff}\equiv V_{eff}(x_{max})-V_{eff}(x_{min})$, the saddle point approximation gives the following probability current : $$\begin{aligned} \hat{J}^{st}=&&k_+-k_- \nonumber \\ k_+=&&\frac{\mid \frac{d^2}{d x^2}V_{eff}(x_{max})\frac{d^2}{d x^2}V_{eff}(x_{min})\mid^{\frac{1}{2}}} {2 \pi \eta}e^{-\frac{\Delta V_{eff}}{k_B T}} \nonumber \\ k_- =&& k_+ e^{-\frac{F L}{k_B T}}\end{aligned}$$ Here, $k_+$ and $k_-$ are the Kramers escape rate, the probability per unit time for a particle near a potential minimum escape to the right minimum and left minimum respectively. These $k_+$ and $k_-$ satisfy the detaled balance relation. $$\frac{k_+}{k_-}=e^{\frac{F L}{k_B T}}\label{ratio}$$ DERIVATION OF STEADY STATE FLUCTUATION THEOREM ============================================== Because of the assumption $k_B T\ll \Delta V_{eff}$, we can describe the motion of a particle in terms of biased random walk. Namely, it is sufficient to observe only the local minmum of effective potential and on what minimum the particle (approximately) exists at each discrete time $t=0,\tau, 2\tau,...$. It is natural to assume that a particle hops only to the nearest-neighbor local minima : to the right minimum with probability $k_+$ , to the left minimum with $k_-$, and it does not move with probability $1-(k_++k_-)$. Then the motion of a particle follows the biased random walk. A particle hops to right, left, and remains with probabilities $k_+$, $k_-$, and $1-(k_++k_-)$. Next we consider the probability $\pi_n(k)$ which is the probability that after n unit time, particle moves k toward right. Here, we can show the equation: $$\frac{\pi_n(k)}{\pi_n(-k)}=(\frac{k_+}{k_-})^k$$ One can prove this equation via following one to one correspondence.: Consider any process that causes the movement k after n unit time. The process is constituted by movements : $k+l$ times toward right, $l$ times toward left, and $n-k-2l$ times remaining. This process occurs with probability $$\frac{n!}{(k+l)!l!(n-k-2l)!}k_+^{k+l}k_-^l(1-k_+-k_-)^{n-k-2l}$$ The corresponding opposite process : $l$ times toward right, $k+l$ times toward left, and $n-k-2l$ times remaining. This process occurs with probability $$\frac{n!}{(k+l)!l!(n-k-2l)!}k_+^lk_-^{k+l}(1-k_+-k_-)^{n-k-2l}$$ So, the ratio of the probability that these two opposite processes occurs is $(\frac{k_+}{k_-})^k$. This is valid for any $l$, and the equation is derived. Then we substitute the detaled balance relation for Kramers escape rate $\frac{k_+}{k_-}=e^{\frac{F L}{k_B T}}$ into the above equation, we obtain following expression: $$\begin{aligned} \frac{\pi_n(k)}{\pi_n(-k)}=e^{\frac{F L k}{k_B T}}\end{aligned}$$ We define the entropy generated $\Delta S\equiv \frac{\Delta Q}{k_B T}$, where $\Delta Q=F L k$ is the Joule heat. $\Delta S$ can be considered as the entropy generated, since the work done on the system extenally will be entirely dissipated, and there are no contribution to the internal energy. Finally, we take the continuous limit, $n\rightarrow t, L k\rightarrow x$, we obtain the steady state Fluctuation Theorem: $$\frac{\pi_t(\Delta S)}{\pi_t(-\Delta S)}=e^{\Delta S}$$ DISCUSSION ========== FT for rachet system is derived. FT derived here is applicable to any system which obeys a biased random walk and satisfies the detailed balance relation. Especially, a complete example was recently given by Nishiyama.et.al[@Higuchi]. They performed an experiment where a single kinesin externally forced moves along a microtuble obeying a biased random walk with regular $8nm$ steps. And they measured the ratio of the forward to backward movements at each load force $1\sim 9pN$, the result well agrees with the detailed balance relation. FT derived in this letter would therefore valid for this system. Because the derivation of FT consists of two parts, identity (\[ratio\]) and detailed balance relation, there is the possibility that under some condition detailed balance does not hold and FT should be modified. This modification of FT is an open problem. Acknowledgement =============== The author is grateful to Professor P.Gaspard, and Professor S.Tasaki for fruitful discussions. [30]{} P.Reimann, Physics Reports 361 (2002) 57-265 H.Risken The Fokker-Planck Equation second edition, Springer C.R.Doering, W.Horsthemke, and J.Riordan (1994), Phys.Rev.Lett.72 19. 2984 M.Nishiyama, H.Higuchi, and T.Yanagida (2002) Nature Cell Biology.vol4.October.790 D.J. Evans, E.G.D. Cohen and G.P. Morriss (1993) Phys. Rev. Lett. 71. 2401. C. Jarzynski, Phys. Rev. Lett. 78 (1996) 2690; J. Stat. Phys. 95 (1999) 367. G. Gallavotti and E.G.D. Cohen (1995) Phys. Rev. Lett. 74. 2694; G.Gallavotii and E.G.D. Cohen (1995) J. Stat. Phys. 80. 931; E.G.D. Cohen and G. Gallavotti, J.Stat.Phys. 96 (1999) 1343. D.J. Evans and D. Searles (1994) Phys. Rev. E 50. 1645. D. Searles and D.J. Evans, J. Chem. Phys. 112 (2000) 9727; [*ibid.*]{} 113 (2000) 3503; D.J. Evans and D. Searles, Adv. Phys. 51 (2002) 1529; E. Mittag and D.J. Evans, Phys. Rev. E 67 (2003) 026113. J. Kurchan (1998) J. Phys. A 31. 3719. J.L. Lebowitz and H. Spohn (1999) J. Stat. Phys. 95. 333. C. Maes (1999) J. Stat. Phys. 95. 367. C. Jarzynski (2000) J. Stat. Phys. 98. 77. G.E. Crooks (1999) Phys. Rev. E 60. 2721.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Polarization profiles of several hundred pulsars have been published recently (eg. Gould & Lyne 1998, GL98; Weisberg et al. 1999). In this report, we summarize the characteristics of circular and high linear polarizations of pulsar profiles, based on all previously published data.' author: - 'J.L. Han$^{1,2}$, R.N. Manchester$^2$, G.J. Qiao$^3$' title: Polarization Characteristics of Pulsar Profiles --- Circular Polarization ===================== Han et al. (1998) systematically studied the circular polarization of pulsar profiles. They found that circular polarization is common in profiles but diverse in nature. Circular polarization is about 9% in average, weaker than linear polarization (15%) in general. We emphasize the following points about circular polarization (CP): [**(1). CP not unique to core emission:**]{} One misleading concept is that circular polarization alway accompanies core emission. It is generally strongest in the central or ‘core’ regions of a profile, but is by no means confined to these regions. Circular polarization has been detected from conal components of many pulsars, for example, conal-double pulsars. [**(2). Sense reversal not unique to core:**]{} Circular polarization often changes sense near the middle of the profile. But sense reversals have been observed at other longitudes, e.g. conal components of PSRs B0834+06, B1913+16, B2020+28, B1039$-$19, J1751$-$4657, and B0329+54 in abnormal mode. [**(3). No PA correlation for core emission:**]{} There is no correlation between the sense of the sign change of circular polarization and the sense of variation of linear polarization angle (PA), in contrast to earlier conclusions on this issue. [**(4). Correlation for cone-dominated pulsars:**]{} We found a strong correlation between between the sign of PA variation and sense of circular polarization in conal-double pulsars, with right-hand (negative) circular polarization accompanying increasing PA and vice versa. No good examples contrary to this trend have been found. [**(5). Variation with frequency:**]{} Circular polarization generally does not vary systematically with frequency. Multifrequency profiles of many pulsars show similar CP across a wide frequency range (eg. PSRs B0329+54, B0525+21). However, we have found several examples with significant variations (Table 5 of Han et al 1998). For example, PSR B1749$-$28 now has been confirmed by the data of GL98 to change from dominant right hand (“$-$”) CP at low frequencies to a sense change “$-/+$” at high frequencies, while PSRs B1859+03 and B1900+01 change from “$+$” to “$+/-$”. High Linear Polarization ======================== Strong linear polarization is an outstanding characteristics of pulsars. Almost all pulsars with log $\dot{E} > 34.5$, if a polarization profile is available, have (at least) one highly linearly polarized component. Using our profile database, we found that the highly linearly polarized components do not have to be associated with high $\dot{E}$. Several types of pulsar profiles have highly polarized components: [**(1). Leading-polarized component:**]{} The prototype of this kind of pulsars is PSR B0355+54. The leading component is almost 100% polarized. The component does not dominate the profile below several hundred MHz, but becomes stronger towards high frequency. It may be emitted from a different region. Other examples are PSRs B0450+55, B1842+14, B2021+51, B0809+74, B0626+24, B1822$-$09. [**(2). Trailing-polarized component:**]{} PSR B0559$-$05 is a mirror-symmetrical type to PSR B0355+54. Its trailing component is highly linearly polarized, and becomes stronger with increasing frequency. Good examples are PSR B2224+65, J1012+5307 and J1022+1001. The latter two are millisecond pulsars. [**(3). Polarized multicomponents:**]{} The prototype PSR B0740$-$28 has 7 Gaussian components fitted to a high time resolution profile. These pulsars have a sharp leading edge and more gradual trailing edge, and are highly polarized. Good examples are PSRs J0538+2817, B0540+23, B0833$-$45, B0950+08i, J1359$-$6038, B1929+10m. PSRs B0538$-$75, J0134$-$2937 and the postcursor of B0823+26 have just mirror-symmetrical profiles to these examples above. [**(4). Polarized single component:**]{} Almost 100% polarized single components may be conal emission emitted from the very outer edge of beam. Two dozen examples have been found. The best examples are PSRs B0105+65, B0611+22, B0628$-$28, B1322+83, B1556$-$44, J1603$-$5657, B1706$-$44, B1828 $-$10, B1848+13, B1913+10, B1915+13. [**(5). Interpulse pulsars:**]{} PSR B0906$-$49 and B1259$-$63 are young, energentic ($\dot{E}>35$), interpulse pulsars with extremely highly polarization. These may be pulsars where both sides of a wide conal beam from a single pole are observed. PSR J0631+1036 may be an off-centre cut through a wide cone. The polarization characteristics of the mean pulse profile provide a framework for understanding the emission processes in pulsars. The characteristics of pulsar circular polarization summarized by Han et al. (1998) should be considered by all emission models. The points on highly linear polarization we make here should be included into future pulsar classifications and geometrical studies of pulsar emission beam. [**Acknowledgements  **]{} JLH thanks financial support from the National Natural Science Foundation (NNSF) and the Educational Ministry of China. Gould D.M., Lyne A.G., 1998, MNRAS 301, 235 Han J.L., Manchester R.N., Xu R.X., Qiao G.J., 1998, MNRAS 300, 373 Weisberg J.M., Cordes J.M., Lundgren S.C., et al., 1999, ApJS 121, 171
{ "pile_set_name": "ArXiv" }
--- abstract: 'To any generic curve in an oriented surface there corresponds an oriented chord diagram, and any oriented chord diagram may be realized by a curve in some oriented surface. The genus of an oriented chord diagram is the minimal genus of an oriented surface in which it may be realized. Let $g_n$ denote the expected genus of a randomly chosen oriented chord diagram of order $n$. We show that $g_n$ satisfies: $$g_n = \frac{n}{2} - \Theta(\ln n).$$' address: - 'School of Computer Science and Engineering, The Hebrew University of Jerusalem, Jerusalem 91904, Israel' - 'Department of Mathematics, Bar-Ilan University, Ramat-Gan 52900, Israel' author: - Nathan Linial - Tahl Nowik date: 'April 27, 2009' title: The expected genus of a random chord diagram --- \[section\] \[thm\][Lemma]{} \[thm\][Proposition]{} \[thm\][Corollary]{} \[thm\][Definition]{} \[thm\][Remark]{} \[thm\][Example]{} \[thm\][Rule]{} Introduction {#int} ============ The study of plane curves dates back to C.F. Gauss. Gauss in [g]{} has attached to any plane curve with $n$ double points a $2n$ letter word, as follows. To each double point attach a letter, and then register the letters you encounter as you travel along the curve. One obtains a word of length $2n$, where each of the $n$ letters appears precisely twice. Such a word is called a *Gauss word*. Clearly not any Gauss word may be realized by a plane curve, and Gauss has pointed out a necessary condition for it to be realizable. One can enhance the Gauss word of a curve with a mark on one of the two occurrences of each letter, signifying that the corresponding strand of the curve crosses the curve at the given double point from right to left. Various necessary and sufficient conditions for a Gauss word to be realizable in the plane have been eventually given, both in the marked and unmarked settings. See [dot]{},[lom]{},[rr]{},[r]{},[t]{} and references therein. To avoid the arbitrariness of the assignment of letters to the different double points, one can replace Gauss words with chord diagrams. To a marked Gauss word corresponds an oriented chord diagram, which is by definition a division of the set $\{1,\dots,2n\}$ into $n$ ordered pairs. This may be represented as a circle with $2n$ designated points, and $n$ oriented chords connecting pairs of these points. Though not every oriented chord diagram may be realized by a curve in the plane, it may be realized by a curve in some oriented surface. J.S. Carter in [c]{} has given a direct construction for the minimal genus surface in which a given diagram may be realized, as follows. Take an annulus, which is thought of as a regular neighborhood of the curve, and identify $n$ pairs of regions along the annulus according to the prescription of the diagram. One obtains an orientable surface $F$ with some $d$ boundary components. Capping off the boundary components with discs produces the required minimal genus surface. The genus of this surface is $g = \frac{1}{2}(n+2-d)$ and so $0 \leq g \leq \frac{1}{2}(n+1)$. We will refer to this minimal genus as the genus of the given diagram. A diagram being realizable in the plane is equivalent to it having genus 0. For fixed $n$, we are interested in the distribution of genera of the oriented chord diagrams with $n$ chords, and we ask: What is the expected genus $g_n$ of a randomly chosen diagram? We show that the expected genus is very close to the maximal possible genus $\frac{1}{2}(n+1)$, in fact, we show that $$g_n = \frac{n}{2} - \Theta(\ln n).$$ Since $g = \frac{1}{2}(n+2-d)$, this is equivalent to showing that the expected number of boundary components is $\Theta(\ln n)$. Though the above description is geometric, counting the number of boundary components may be described in a purely combinatorial manner. Traveling along a boundary component of our surface corresponds to a walk on the chord diagram according to the following rule. When moving along the circle of the diagram, and arriving at an end of a chord, continue your motion along the chord to its other side. If your motion along the chord is in the direction (respectively, against the direction) of its orientation, then continue your motion along the circle in the same (respectively, opposite) direction as you have moved before entering the chord. So, to identify a boundary component, one travels along the diagram according to the above rule, until returning to the starting point. Repeating this process, one obtains all boundary components. This combinatorial walk along a chord diagram is reminiscent of the walk along the cycles of a permutation. The distribution of the cycles of a random permutation on $n$ letters is well understood, and the expected number of cycles is also $\Theta(\ln n)$. Indeed, it is for similar reasons that the expected number of cycles in a permutation and the expected number of cycles of the walk along a chord diagram are both $\Theta(\ln n)$, though as will be seen, the setting of chord diagrams is substantially more complicated. Since this problem may be formulated both in a topological and a purely combinatorial manner, it may be of interest to both topologists and combinatorialists, and indeed the text is aimed for both audiences. This work may be viewed as part of the recent expanding interest in probabilistic questions in topology, as appears in [bhk]{},[dut]{},[lim]{},[mw]{},[ps]{}. Definitions and statement of result {#def} =================================== Let $F$ be an oriented surface. A *generic curve* in $F$ is an immersion $c: S^1 \to F$ for which the only self intersections are transverse double points. We fix $n$ once and for all, and call a point of $S^1$ a *dot* if it is one of the $2n$ points $\{ e^{\pi i k / n} \ : \ 1 \leq k \leq 2n \}$. If a generic curve $c$ has $n$ double points, then there are $2n$ points in $S^1$ mapped into them, and we will always assume that these $2n$ points are precisely our $2n$ dots. A generic curve with such $n$ double points will be called an $n$-curve. An oriented chord diagram of order $n$ is a division of the set of dots into $n$ disjoint ordered pairs. One can represent an oriented chord diagram in the plane, by drawing an oriented chord connecting each ordered pair, where the orientation of the chord represents the order of the pair. An oriented chord diagram will also be called simply a *diagram*. We denote the set of all diagrams of order $n$ by ${{\mathcal D}}_n$, and we have $|{{\mathcal D}}_n| = \frac{(2n)!}{n!}$. Any $n$-curve determines a diagram $D(c) \in {{\mathcal D}}_n$ as follows. The double points of $c$ divide the $2n$ dots into pairs, and the orientation of the surface $F$ induces an ordering on each pair, in the following way. If $c(a)=c(b)$ for dots $a,b$, and $c'(b),c'(a)$ is a positive basis with respect to the orientation of the surface, then the ordered pair $(a,b)$ is taken. See Figure [f1]{}. Any diagram may be realized by a curve on some oriented surface, and a regular neighborhood of the curve in all such surfaces is the same, as we now explain (compare [c]{}). Given a diagram $D \in {{\mathcal D}}_n$, take the annulus $A = S^1 \times [-{\epsilon},{\epsilon}]$ with a fixed orientation, and identify $S^1$ with $S^1 \times \{ 0 \}$. For each dot $a$ let $S_a {\subseteq}A$ be the $2{\epsilon}\times 2{\epsilon}$ square centered at $(a,0)$. We now identify pairs of dots $a,b$ according the prescription of $D$, and we identify the corresponding squares $S_a,S_b$ with a positive or negative $\frac{\pi}{2}$ rotation, so that the self intersection of the curve at the identified square will be as prescribed by the orientation of the chord between $a$ and $b$ in $D$. We obtain an oriented surface with $d(D) \geq 1$ boundary components, which we denote $F(D)$. Clearly, the embedding $S^1 \to A$ given by $z \mapsto (z,0)$, composed with the quotient map $A \to F(D)$, is an $n$-curve $S^1 \to F(D)$ which realizes $D$, and a regular neighborhood of any curve in any surface realizing $D$, is identical to $F(D)$. It follows that the genus of the closed surface obtained by capping off the $d(D)$ boundary components of $F(D)$ with discs, is the minimal genus of a surface in which $D$ may be realized. Given a diagram $D$, we define $g(D)$, the genus of $D$, to be the minimal genus of a closed oriented surface admitting a curve $c:S^1 \to F$ with $D(c)=D$. The image in $F(D)$ of our $n$-curve is a graph with $n$ vertices and $2n$ edges, and so $2-2g(D) = n - 2n + d(D)$, or, $g(D) = \frac{1}{2}(n+2-d(D))$. Since $d(D) \geq 1$, we deduce that $0 \leq g(D) \leq \frac{1}{2}(n+1)$. In this work we study the following question: What is the expected genus of a randomly chosen diagram $D \in {{\mathcal D}}_n$? We will show that the expected genus is close to the maximal possible genus $\frac{1}{2}(n+1)$. More precisely we will show: \[main\] The expected genus $g_n$ of a random oriented chord diagram $D \in {{\mathcal D}}_n$ satisfies: $$g_n = \frac{n}{2} - \Theta(\ln n).$$ We think of $g(D)$ and $d(D)$ as random variables defined on a randomly chosen $D \in {{\mathcal D}}_n$. That is, our sample space is ${{\mathcal D}}_n$, each diagram having equal probability $\frac{n!}{(2n)!}$. We denote the expected values by $g_n = E[g]$ and $d_n = E[d]$. We will show that $d_n = \Theta(\ln n)$, from which Theorem [main]{} follows via $g_n = \frac{1}{2}(n+2-d_n)$. The random procedure {#ran} ==================== We label the dot $e^{\pi i k /n}$, and its corresponding square, simply by $k$, and so $k \pm 1$ will mean addition mod $2n$. The interval along the boundary of the annulus $A$ between two adjacent squares will be called an *edge*, so we have $4n$ edges. We orient the edges according to the orientation induced on ${{\partial}}A$ from that of $A$, and we denote the oriented edge from square $a$ to square $b$ by $[a,b]$. The edge $[b,a]$ will then be the parallel edge in the other boundary component. So, all edges in $S^1 \times \{ {\epsilon}\}$ are of the form $[a,a+1]$, and they will be called *positive* edges, and all edges in $S^1 \times \{ -{\epsilon}\}$ are of the form $[a+1,a]$ and will be called *negative* edges. See Figure [f2]{}. We would like to see how our edges are attached to each other due to the gluing of two squares. We will say an attachment $[a,b]-[c,d]$ takes place if the end point of $[a,b]$ is glued to the beginning point of $[c,d]$. So, say we have identified the two squares $a$ and $b$ according to the oriented chord $(a,b)$. As seen in Figure [f3]{}, the eight edges involved are attached to each other as follows: $[a-1,a]-[b,b+1]$,  $[b+1,b]-[a,a+1]$,  $[a+1,a]-[b,b-1]$,  $[b-1,b]-[a,a-1]$. This can be summarized by the following rule: \[r\] An attachment $[a,b]-[c,d]$ holds if the oriented chord $(b,c)$ exists, and the signs of $[a,b]$ and $[c,d]$ are the same, or if the oriented chord $(c,b)$ exists and the signs of $[a,b]$ and $[c,d]$ are opposite. Traveling along a boundary component of the surface $F(D)$ corresponds to a walk along the diagram $D$, which by Rule [r]{} proceeds as follows: When moving along the circle of the diagram, and arriving at an end of a chord, continue your motion along the chord to its other side. If your motion along the chord is in the direction (respectively, against the direction) of its orientation, then continue your motion along the circle in the same (respectively, opposite) direction as you have moved before entering the chord. We may thus read all boundary components directly from the diagram $D$, as follows: Choose an arbitrary edge, and start traveling along the diagram in the above way, alternatingly passing edges and chords, until you return to your initial edge. Then choose some unvisited edge, and similarly travel until you return to it. Continue until all edges have been visited. Notice that when this is done, each chord of the diagram has been visited four times, each visit corresponding to one of the four corners of the glued square. For $k \leq n$, a $k$-$n$-diagram is a choice of $2k$ out of the $2n$ dots, and a division of these $2k$ dots into $k$ oriented chords, i.e. ordered pairs. The remaining $2n-2k$ dots will be called *vacant* dots. So, an oriented chord diagram of order $n$ is an $n$-$n$-diagram. \[d0\] A *path* in a $k$-$n$-diagram $D$, is a sequence $[a_1,b_1]-[a_2,b_2]- \cdots - [a_r,b_r]$ of distinct edges, attached via oriented chords of $D$ according to Rule [r]{}. A path in $D$ is called a *loop* if the attachment $[a_r,b_r]-[a_1,b_1]$ also holds. A path in $D$ is called a *segment* if $a_1$ and $b_r$ are vacant dots (perhaps the same dot). Loops and segments are precisely those paths that cannot be further extended. We now specify our procedure for choosing a random $n$-$n$-diagram. Our procedure will choose the chords one by one. We first fix an ordering $e_1,\dots,e_{4n}$ of our edges, once and for all. Before the procedure begins, we announce $e_1$ as the “pointer” edge. Assume that after the $(j-1)$th step, we have already chosen $j-1$ oriented chords, and the pointer edge lies in a segment (rather than a loop) of the given $(j-1)$-$n$-diagram. The next chord is now chosen with one of its dots being the concluding dot of the segment in which the pointer lies, and its other dot is randomly chosen from the other $2n-2j+1$ vacant dots. The orientation of the new chord is also randomly chosen. If in the new $j$-$n$-diagram, the pointer’s segment continues to be a segment, i.e. it does not close into a loop, then the same edge remains the pointer. If on the other hand, after the $j$th chord is added, the pointer’s segment closes into a loop, then the edge with smallest index which lies in a segment in the new $j$-$n$-diagram, becomes the new pointer. This procedure indeed produces all $n$-$n$-diagrams with equal probability. Examples of two runs of our random procedure appear in Figures [f4]{} and [f5]{}, demonstrating some of the interesting features of the possible evolution of the pointer’s segment. In both figures the edge $e_1$ is the edge $[a,a+1]$ and is marked by 1. In Figure [f4]{} the chords are chosen in the following order: $(a+1,b),(c,b+1),(c-1,c+1),(a,a+2)$. The segment of $e_1$ after the four steps of this run is $[a+3,a+2]-[a,a+1]-[b,b+1]-[c,c-1]-[c+1,c]-[b+1,b]-[a+1,a+2]-[a,a-1]$ and these edges are marked in the figure by $0,\dots,7$, in this order. The evolution of $e_1$’s segment throughout the four steps of this run is $1-2$, $1-2-3$, $1-2-3-4-5-6$, $0-1-2-3-4-5-6-7$. Note that after the third step the segment revisits some chords that were chosen in previous steps. Note that after the fourth step the segment extends in both directions, and so $1$ is no longer the first edge in the segment. We point out the following difference between this run, and the run of Figure [f5]{}. In Figure [f4]{} the initial and final dots of the pointer’s segment are distinct, whereas in Figure [f5]{} it is the same dot. Upper bound for $d_n$ {#up} ===================== In this section we establish our upper bound for $d_n$. If $\ell$ is a loop in an $n$-$n$-diagram $D$, then the *size* of $\ell$ is the number of distinct chords visited by $\ell$. If $k$ is the size of the loop $\ell$ and $r$ is the number of edges it visits, then since $\ell$ alternatingly visits an edge and a chord, and since each chord is visited at most four times, we have $k \leq r \leq 4k$. For given $n$, let $L_k = L_k(n)$ denote the expected number of loops of size $k$ in a random $n$-$n$-diagram, then $d_n = \sum_{k=1}^n L_k$. We will show in Proposition [p4]{} below, that for $k \leq \frac{n}{{{100}}}$, $L_k \leq \frac{3}{k}$. On the other hand, since any chord is visited by at most four different loops, the total number of all loops of size $k > \frac{n}{{{100}}}$ is at most ${{400}}$, and so its expected value $\sum_{k > \frac{n}{{{100}}}} L_k$ is at most ${{400}}$. Together this gives $$d_n = \sum_{k \leq \frac{n}{{{100}}}} L_k + \sum_{k > \frac{n}{{{100}}}} L_k \leq \sum_{k \leq \frac{n}{{{100}}}} \frac{3}{k} + {{400}}\leq 3 \ln n + {{400}}.$$ In order to obtain our bound $L_k \leq \frac{3}{k}$ we will need to bound the probability that at a given $k$th step, the pointer’s segment closes into a loop. We now prove the following: \[p3\] For $k \leq \frac{n}{{{100}}}$, the probability that the pointer’s segment will close into a loop at the $k$th step is at most $\frac{3}{4n}$. Let $S$ be the pointer’s segment in our $(k-1)$-$n$-diagram after step $k-1$, and let $p$ be its concluding dot. We now need to choose the $k$th chord, with one end being $p$. We must determine how many choices will result in closing $S$ into a loop. Let $q$ be the initial dot of $S$. If $q \neq p$ (as occurs in the example in Figure [f4]{}, where $q=a+3$ and $p=a-1$), then for $S$ to be closed into a loop, we must choose $q$ as the second dot for the $k$th chord, and so the choice of unoriented chord is unique. Though usually only one of the two choices of orientation for this chord will indeed close the segment $S$ into a loop (as is the case in Figure [f4]{}), it may in fact occur that both orientations accomplish this. Since we are seeking an upper bound for the probability, we will always count both orientations as possible, or in other words we will ignore the choice of orientation in the computation of the probability. Since there are $2n-2k+1$ vacant dots from which we may choose the second dot for the new chord, the probability of choosing the correct dot $q$ is $\frac{1}{2n-2k+1}$, and since we assume $k \leq \frac{n}{100}$ we have $\frac{1}{2n-2k+1} \leq \frac{1.1}{2n}$. If $q=p$ (as in the example of Figure [f5]{}, where $q=p=a$), then at first sight it may seem that closing the pointer’s segment into a loop is impossible, and in most cases this is in fact true. But on the other hand, there are cases with $q=p$ where not only does there exist a choice of chord that closes the given segment into a loop, but there are in fact many such choices. The example in Figure [f5]{} is such case. The oriented chords $(a,a+4)$, $(a,a+7)$ or $(a,a+11)$ may each be added in the present step to close the pointer’s segment into a loop. For the analysis of this phenomenon, we define the following notion. \[d1\] A *plug* is a segment $[a_1,b_1]-\cdots-[a_r,b_r]$ for which $a_1=b_r$. The vacant dot $a_1=b_r$ is called the *entrance* to the plug. Examples of two plugs are depicted in Figure [f6]{}. In Figure [f6]{}a, the chord $(a-1,a+1)$ produces the plug $[a,a-1] - [a+1,a]$ with entrance $a$. In Figure [f6]{}b, the chords $(a+1,a+3)$ and $(a+4,a+2)$ produce the plug $[a,a+1] - [a+3,a+4] - [a+2,a+3] - [a+1,a]$ with entrance $a$. Note that the same vacant dot can be the entrance to two different plugs. In our present case, where $q=p$, the pointer’s segment is itself a plug, but this fact is not of interest to us. What enables us to close the pointer’s segment into a loop in Figure [f5]{}, is each one of the *additional* plugs that are present in the given $5$-$n$-diagram, namely the three plugs $[a+4,a+3]-[a+5,a+4]$, $[a+7,a+6]-[a+8,a+7]$, and $[a+11,a+10]-[a+12,a+11]$ (each of which is similar to the plug in Figure [f6]{}a). In fact, the next lemma shows that in order to close the pointer’s segment into a loop in the case $q=p$, it is necessary that the second dot of the new chord will be an entrance to a plug. \[l1\] Let $D$ be a $(j-1)$-$n$-diagram, and let $a$ be a vacant dot in $D$. Let $e$ be an edge entering $a$, (i.e. $e$ is $[a-1,a]$ or $[a+1,a]$). Assume $Q$ is an additional oriented chord with one end at $a$ and the other end at some other vacant dot $b$, such that in the $j$-$n$ diagram obtained by adding $Q$, the path beginning at $e$ leads to an edge $e'$ which is one of the two exiting edges at $a$ (i.e. $e'$ is $[a,a+1]$ or $[a,a-1]$). Then $b$ is the entrance to a plug in $D$. Assume $b$ is not an entrance to a plug. In Figure [f7]{}, edge $e$ is marked, and the two possibilities for $e'$ are marked. The path beginning at $e$ passes the point $x$, and in order for it to lead to $e'$, it must eventually arrive back into the region depicted in the figure. It does not arrive at $y_1$ or $y_3$ since we have assumed that $b$ is not an entrance to a plug. If it arrives at $y_4$ then it closes a loop without passing either possibility for $e'$. So, it must arrive at $y_2$ as depicted, and it then exits our region through point $z$. This time its only possibility for returning is through $y_4$, which as before prevents it from ever arriving at either possibility for the exiting edge $e'$. As we have seen, there may be many plugs available for completing our segment into a loop, but fortunately, the *expected* number of available plugs is small. The main technical effort of this work is the following proposition whose proof we defer to Section [plug]{}. \[pl\] For $k \leq \frac{n}{{{100}}}$, the expected number of plugs present after $k$ steps of the random procedure is at most $\frac{1}{4}$. Back to the proof of Proposition [p3]{} for the case $q=p$. By Lemma [l1]{}, in order for $S$ to close into a loop, the second dot we choose for the new chord must be the entrance to some plug. By Proposition [pl]{}, the expected number of plugs existing in the present stage of the random procedure (i.e. after $k-1$ steps), is at most $\frac{1}{4}$. Together with the case $q \neq p$ we have on average at most $1 + \frac{1}{4}$ choices for the new unoriented chord. (As before, we ignore the additional choice of orientation.) Note that we must take the sum and not the maximum of the bounds for the two possibilities $q \neq p$ and $q=p$, since the expectation for the number of plugs that we bound in Proposition [pl]{} is not conditional on $q = p$. We obtain that for $k \leq \frac{n}{100}$, the probability that the pointer’s segment closes into a loop at the $k$th step of the random procedure is at most $(1 + \frac{1}{4}) \cdot \frac{1.1}{2n} \leq \frac{3}{4n}$, which completes the proof of Proposition [p3]{}. \[p4\] Let $L_k$ denote the expected number of loops of size $k$ in a random $n$-$n$-diagram. Then for $k \leq \frac{n}{{{100}}}$ we have $L_k \leq \frac{3}{k}$. It is clear from the definition of our random procedure, that if the segment of $e_1$ closes into a loop at the $k$th step, then this loop is of size $k$. So the event that the edge $e_1$ lies in a loop of size $k$ is the same as the event that $e_1$ survives as pointer until step $k$, and then at step $k$ its segment closes into a loop. The probability for this event is at most the probability that at the $k$th step the pointer’s segment closes into a loop, and by Proposition [p3]{} this probability is at most $\frac{3}{4n}$. Now, our random procedure produces each $n$-$n$-diagram with equal probability, and so by the symmetry of our annulus, the probability for *any* given edge to lie in a loop of size $k$ is also at most $\frac{3}{4n}$, or alternatively, the probability $P_k$ that a *randomly* chosen edge will lie in a loop of size $k$ is at most $\frac{3}{4n}$. We obtain a lower bound for $P_k$ by noting that each loop of size $k$ includes at least $k$ edges, and the total number of edges is $4n$, and so $P_k \geq \frac{kL_k}{4n}$. Together we get $\frac{kL_k}{4n} \leq P_k \leq \frac{3}{4n}$ which proves our claim. As already explained above, the bound $L_k \leq \frac{3}{k}$ for $k \leq \frac{n}{{{100}}}$ implies the following upper bound for $d_n$: $$d_n \leq 3 \ln n + {{400}}.$$ Lower bound for $d_n$ {#low} ===================== We have asked in the proof of Proposition [p4]{}, what is the probability $P_k$ that a randomly chosen edge will lie in a loop of size $k$. We have noticed that this is precisely the probability that in our random procedure, the segment of $e_1$ survives until the $k$th step, and then at the $k$th step it closes into a loop. In this section we will find a lower bound for $P_k$, for $n \geq 50$ and $k \leq \sqrt{n}$, which in turn will provide a lower bound for $d_n$. A run of $j-1$ steps of the random procedure is called *good* if after these $j-1$ steps $e_1$ is still the pointer, and its segment is of the form $[a_1,b_1]-\cdots-[a_j,b_j]$ with all the dots $a_1,b_1,\dots,a_j,b_j$ being distinct. When $j-1=0$, i.e. before starting the random procedure, then the pointer’s segment is simply $e_1=[a_1,b_1]$, so the run is good. If the run is good after $j-1$ steps, and at the $j$th step the second dot chosen for the new chord is not adjacent to any of the dots $a_1,b_1,\dots,a_j,b_j$, then the run is still good after the $j$th step. This restriction for the choice of the $j$th chord means that if an edge of the segment is say $[a,a+1]$, then the four dots $a-1,a,a+1,a+2$ are not chosen. So, at the $j$th step we have at most $4j$ dots which we are forbidden to choose, so the number of allowed choices for a new dot at the $j$th step is at least $2n-4j$. Since the total number of dots from which we choose is $2n-2j+1$, the probability for such restricted choice at the $j$th step is at least $\frac{2n-4j}{2n-2j+1}$. If after $k-1$ steps of the random procedure the run is still good, then in particular, the initial and final dots of $e_1$’s segment are distinct. So, at the $k$th step there exists a choice of oriented chord that closes $e_1$’s segment into a loop, and the probability for this choice is $\frac{1}{2(2n-2k+1)}$. So together, for $n \geq 50$ and $k \leq \sqrt{n}$, the probability that the segment of $e_1$ survives until the $k$th step, and then at the $k$th step closes into a loop satisfies: $$\begin{gathered} P_k \geq \frac{1}{2(2n-2k+1)}\prod_{j=1}^{k-1} \frac{2n-4j}{2n-2j+1} \geq \frac{1}{4n}\prod_{j=1}^{k-1} (1-\frac{2j+1}{2n-2j+1}) \geq \frac{1}{4n}\prod_{j=1}^{k-1} (1-\frac{j+1}{n-k}) \\ \geq \frac{1}{4n}\prod_{j=1}^{k-1} e^{-\frac{6}{5}\cdot\frac{j+1}{n-k}} = \frac{1}{4n}e^{-\frac{6}{5}\sum_{j=1}^{k-1}\frac{j+1}{n-k}} \geq \frac{1}{4n}e^{-\frac{3}{5}\cdot\frac{k^2 + k}{n-k}} \geq \frac{1}{4n}e^{-\frac{3}{5}\cdot\frac{n + \sqrt{n}}{n-\sqrt{n}}} \geq \frac{1}{9n}.\end{gathered}$$ (We use the assumption $n \geq 50$ in the fourth and last inequalities.) As before, let $L_k$ be the expected number of loops of size $k$ in a random $n$-$n$-diagram, then since the number of edges in a loop of size $k$ is at most $4k$ we have $P_k \leq \frac{4k L_k}{4n}$. Together, for $n \geq 50$ and $k\leq \sqrt{n}$ we get $\frac{1}{9n} \leq P_k \leq \frac{4k L_k}{4n}$, so $L_k \geq \frac{1}{9k}$. We may now establish our lower bound for $d_n$, for $n \geq 50$: $$d_n = \sum_{k=1}^{n} L_k \geq \sum_{k=1}^{\sqrt{n}} \frac{1}{9k} \geq \frac{1}{9} \ln \sqrt{n} = \frac{1}{18} \ln n.$$ Together with the upper bound of Section [up]{} we obtain $d_n = \Theta(\ln n)$, which proves Theorem [main]{}, stating that the expected genus $g_n$ of a random diagram of order $n$ satisfies: $$g_n = \frac{n}{2} - \Theta(\ln n).$$ Upper bound for the expected number of plugs {#plug} ============================================ In this section we prove Proposition [pl]{}, stating that for $k \leq \frac{n}{100}$, the expected number of plugs present in our $k$-$n$-diagram after $k$ steps of the random procedure is at most $\frac{1}{4}$. \[d3\] Two vacant dots in a $k$-$n$-diagram $D$ are called *neighbors*, if they are the two end points of a segment in $D$. \[sgn\] A *positive* plug is a plug $[a_1,b_1]-\cdots-[a_r,b_r]$ for which the two edges $[a_1,b_1],[a_r,b_r]$ are of the same sign, that is, they are of the form $[a,a+1],[a-1,a]$ or $[a,a-1],[a+1,a])$, (as in Figure [f6]{}a). A *negative* plug is a plug for which these two edges are of opposite sign, that is, they are of the form $[a,a+1],[a+1,a]$ or $[a,a-1],[a-1,a]$, (as in Figure [f6]{}b). Note that if same vacant dot is the entrance to two different plugs, then these two plugs must be of the same sign. \[l2\] Under the assumptions of Lemma [l1]{}, if $e$ and $e'$ are of opposite sign, and if $b$ is not the entrance to a positive plug (and so by Lemma [l1]{} it is the entrance to one or two negative plugs), then either $a$ and $b$ are neighbors (Definition [d3]{}), or $a$ is also an entrance to a plug. Assume $a$ and $b$ are not neighbors. In order for us to arrive at $e'$, given that $b$ is not the entrance to a positive plug and $a$ and $b$ are not neighbors, our path must be as in Figure [f8]{}, which shows that $a$ is the entrance to a (negative) plug. Any chord is involved in at most four different segments, and so at each step, when adding a new chord, at most four new plugs can be created. But we will show that in fact the expected number of plugs created at each step $k \leq \frac{n}{{{100}}}$ is at most $\frac{25}{n}$. This implies that the expected number of plugs present after $k \leq \frac{n}{{{100}}}$ steps is at most $\frac{1}{4}$. To establish this bound we will in fact need to prove the following more detailed proposition, which distinguishes between positive and negative plugs. \[p2\] The following holds for $k \leq \frac{n}{{{100}}}$: 1. Let $G^+_k$ (respectively $G^-_k$) denote the expected number of positive (respectively negative) plugs completed at the $k$th step. Then $G^+_k \leq \frac{{{5}}}{n}$ and $G^-_k \leq \frac{{{20}}}{n}$. \[1\] 2. The expected number of plugs present after $k$ steps is at most $\frac{1}{4}$. \[2\] 3. Let $H^+_k$ (respectively $H^-_k$) denote the probability that after the $k$th step the concluding dot of the pointer’s segment is an entrance to a positive (respectively negative) plug. Then $H^+_k \leq \frac{{{6}}}{n}$ and $H^-_k \leq \frac{{{21}}}{n}$. \[3\] ([1]{}) Say at the $k$th step we have chosen a chord $Q$ between dots $a$ and $b$, and a plug has been completed, with dot $c$ being its entrance. This means that after adding $Q$ there is a segment with edges $e_{i_1} - e_{i_2} - \cdots - e_{i_r}$ beginning and ending at the vacant dot $c$, and before adding $Q$ this segment did not exist. This means that before adding $Q$, the segment $S_1$ beginning with $e_{i_1}$ ended at some vacant dot $a \neq c$, and the segment $S_2$ ending with $e_{i_r}$ began at some vacant dot $a' \neq c$. We now distinguish three cases as follows. If $a \neq a'$ then the new chord $Q$ must be between $a$ and $a'$. By definition of our random procedure, the concluding vacant dot $p$ of the pointer’s segment is one of the dots of the new chord $Q$, so must be either $a$ or $a'$. We will refer to this case as Case A. If on the other hand $a=a'$ then the new chord $Q$ must be between $a$ and some other vacant dot $b$. In this case either $p=a$ or $p=b$, and these two possibilities will be referred to as Case B and Case C, respectively. For Case A, we note that there are at most four different segments with one end being $p$. The other end of each such segment is a vacant dot that may be $c$ of the above description. For each such $c$ there is a unique second segment with which a configuration $S_1,S_2$ as described above may arise for a positive plug, and a unique such second segment for a negative plug. Our assumption is that $a \neq a'$ and so for each such configuration there is a unique choice of unoriented chord with which such a plug may be created. As discussed in Section [up]{}, it may be that both choices of orientation for this chord bring to the completion of the plug. So here and in all following cases, we do as we have done in Section [up]{}, and include both choices in our count by simply ignoring the choice of orientation. As before, the probability for the correct unoriented chord to be chosen in each case is $\frac{1}{2n-2k+1}$ since there are $2n-2k+1$ additional vacant dots, and for $k \leq \frac{n}{100}$ we have $\frac{1}{2n-2k+1} \leq \frac{1.1}{2n}$. So, the contribution of this case to $G^+_k$ and $G^-_k$ is at most $4 \cdot \frac{1.1}{2n}$. Note that it may be that different configurations in our count are completed into a plug by the same choice of chord, but by the additivity of expectation, the contributions of all configurations may be added regardless of the dependence between them. In Case B, $Q$ is between the dot $p=a$ and the dot $b$, and by Lemma [l1]{}, $b$ must be an entrance to an existing plug. We bound all possible contributions that may be from choosing the second dot of the new chord as the entrance to an existing plug. Any new chord may participate in at most four different segments, and so at most four new plugs may be completed. By induction, we may use ([2]{}) of the present proposition for $k-1$ to conclude that on average we have at most $\frac{1}{4}$ existing plugs available. So, the contribution is on average at most $4 \cdot \frac{1}{4} \cdot \frac{1.1}{2n} = \frac{1.1}{2n}$. We cannot determine how this contribution will divide between $G^+_k$ and $G^-_k$ and so we add it to both. In Case C, $p=b$, and our choice of the second dot $a$ for $Q$ is such that $a$ is part of a configuration of segments $S_1,S_2$ and dots $c,a$ as described above. The segments $S_1$ and $S_2$ may or may not pass chords, but there is just one special configuration for $S_1,S_2$ in which both $S_1$ and $S_2$ do not pass any chord, namely, the configuration where $a$ and $c$ are adjacent dots along the annulus, and $S_1,S_2$ are the two edges connecting them. If the configuration is not this special one, then necessarily the dot $c$ is adjacent along the annulus to a dot that is the end of one of the $k-1$ existing chords. So there are at most $4(k-1)$ possibilities for such dot. For each such dot $c$ there are two possibilities for a pair of segments $S_1,S_2$ that may give rise to a positive plug, and two possibilities for a negative plug. Together this gives at most $8k$ possible pairs of segments for positive plugs and for negative plugs. Now we note that in order for us to be in Case C, our dot $p$ must be at the entrance to an existing plug after step $k-1$. By induction we may use ([3]{}) of the present proposition for $k-1$ to conclude that the probability for us being in Case C is at most $\frac{{{6}}}{n} + \frac{{{21}}}{n}$. And so the contribution of the non-special configurations to $G^+_k$ and $G^-_k$ is at most $8k (\frac{{{6}}}{n} + \frac{{{21}}}{n})\frac{1.1}{2n} \leq 8 \cdot \frac{n}{{{100}}} \cdot \frac{{{27}}}{n} \cdot \frac{1.1}{2n} \leq 3 \cdot \frac{1.1}{2n}$. For the special configuration, if a plug is completed then it is necessarily a negative plug, so contributes only to $G^-_k$. If $p=b$ is the entrance to a positive plug, which happens, by induction on part ([3]{}), with probability at most $\frac{{{6}}}{n}$, then we take our bound to be simply the total number of choices $2n-2k+1$ for $a$. There may be a special configuration on each side of $a$, and so the contribution to $G^-_k$ is at most $2 \cdot (2n-2k+1) \cdot \frac{{{6}}}{n} \cdot \frac{1.1}{2n} \leq 24 \cdot \frac{1.1}{2n}$. If $p=b$ is the entrance to a negative plug, which happens by part ([3]{}), by induction, with probability at most $\frac{{{21}}}{n}$, then by Lemma [l2]{}, we must choose $a$ which is either a neighbor of $b$ or the entrance to a plug. The dot $b$ has at most 4 neighbors. For each such neighbor $a$ there is at most one special configuration that may be completed into a plug, since it may not be on the side of $a$ where the segment from $b$ arrives at $a$. So, the contribution of this case is at most $4 \cdot \frac{{{21}}}{n} \cdot \frac{1.1}{2n}$. The second possibility is that $a$ itself is an entrance to a plug, but in Case B above we have already counted all possible contributions from connecting $p$ to a dot which is the entrance to an existing plug, and so we need not count this again here. The contribution to $G^-_k$ is thus at most $4 \cdot \frac{{{21}}}{n} \cdot \frac{1.1}{2n} \leq \frac{1.1}{2n}$, since $n \geq {{100}}$ whenever the assumption $k \leq \frac{n}{{{100}}}$ is relevant. We add all contributions for $G^+_k$: $$G^+_k \leq (4 + 1 + 3) \cdot \frac{1.1}{2n} \leq \frac{{{5}}}{n},$$ and for $G^-_k$: $$G^-_k \leq (4 + 1 + 3 + 24 + 1) \cdot \frac{1.1}{2n} \leq \frac{{{20}}}{n}.$$ ([2]{}) In each step $j \leq k$ on average at most $\frac{{{5}}}{n} + \frac{{{20}}}{n}$ plugs are completed, by ([1]{}), and so after $k$ steps the expected number of plugs is at most $k (\frac{{{5}}}{n} + \frac{{{20}}}{n}) \leq \frac{n}{{{100}}}(\frac{{{5}}}{n} + \frac{{{20}}}{n}) = \frac{1}{4}$. ([3]{}) If after the $k$th step, the final dot of our spanning segment is the entrance to a positive plug, then this plug may either be one that has existed previously, or one that has just been completed. If it is a plug that has existed previously, then in the $(k-1)$-$n$-diagram we had before the $k$th step, there is a unique segment $S$ leading to its entrance (which is not the plug itself), and let $a$ denote the vacant dot at the beginning of $S$. In order for us to end up at the entrance to the given plug after adding the $k$th chord, this chord must include $a$. As before, let $p$ denote the concluding dot of the pointer’s segment. If $p \neq a$ then we have one choice for such unoriented chord. If $p=a$ then in order for us to continue into the segment $S$, then by Lemma [l1]{} the other dot $b$ of the new chord must be the entrance to an existing plug. Together we see that in order for us to land at the entrance of an existing plug, we must choose the second dot for the new chord either as a dot $a$ as described above, which is uniquely determined by a plug, or as a dot which is itself the entrance to a plug. By ([2]{}) we know that there are on average at most $\frac{1}{4}$ previously existing plugs, and so this contributes at most $2 \cdot \frac{1}{4} \cdot \frac{1.1}{2n}$ to the probability. On the other hand, the probability that after the $k$th step we have landed at the entrance of a positive plug that has just been completed, is at most the probability that such a plug has at all been completed at the $k$th step. By ([1]{}) this probability is at most $\frac{{{5}}}{n}$, since the expected number of plugs completed is a bound to the probability that at least one plug has been completed. Together we get $H^+_k \leq 2 \cdot \frac{1}{4} \cdot \frac{1.1}{2n} + \frac{{{5}}}{n} \leq \frac{{{6}}}{n}$. In the same way, using $G^-_k \leq \frac{{{20}}}{n}$ we get $H^-_k \leq \frac{{{21}}}{n}$ Recall that what we have actually used from Proposition [p2]{} is only part (2), which bounds the total number of plugs. The need for this more detailed analysis is due to the large contribution of existing positive plugs to the completion of new negative plugs in Case C with the special configuration. This required that we separate between positive and negative plugs in the inductive proof, with a larger bound for the negative plugs. [StoPC]{} E. Babson, C. Hoffman, M. Kahle: “The fundamental group of random 2-complexes.” arXiv:0711.2704 J. S. Carter: “Classifying immersed curves.” *Proc. Amer. Math. Soc.* 111 (1991), no. 1, 281–287. C. H. Dowker, M. B. Thistlethwaite: “Classification of knot projections.” *Topology Appl.* 16 (1983) 19–31. N. M. Dunfield, W. P. Thurston: “Finite covers of random 3-manifolds.” *Invent. Math.* 166 (2006), no. 3, 457–521. C. G. Gauss: *Werke* 8, 271-286. N. Linial, R. Meshulam: “Homological connectivity of random 2-complexes.” *Combinatorica* 26 (2006), no. 4, 475–487. L. Lovász, M. L. Marx: “A forbidden substructure characterization of Gauss codes.” *Acta Sci. Math.* (Szeged), 38 (1976), 115–119. R. Meshulam, N. Wallach" “Homological connectivity of random k-dimensional complexes” *Random Structures Algorithms* 34 (2009) 408–417. N. Pippenger, K. Schleich: “Topological characteristics of random triangulated surfaces.” *Random Structures Algorithms* 28 (2006), no. 3, 247–288. R. C. Read, P. Rosenstiehl: “On the Gauss crossing problem.” *Combinatorics* (Proc. Fifth Hungarian Colloq., Keszthely, 1976), Vol. II, pp. 843–876. P. Rosenstiehl: “Solution algébrique du problème de Gauss sur la permutation des points d’intersection d’une ou plusieurs courbes fermés du plan” *C.R. Acad. Sci. Paris Sér. A-B* 283 (1976), A551–A553. V. Turaev: “Curves on surfaces, charts, and words.” *Geom. Dedicata* 116 (2005), 203–236.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We show that the Lusin area integral or the square function on the unit ball of ${{\mathbb C}}^n$, regarded as an operator in weighted space $L^2(w)$ has a linear bound in terms of the invariant $A_2$ characteristic of the weight. We show a dimension-free estimate for the “area-integral” associated to the weighted $L^2(w)$ norm of the square function. We prove the equivalence of the classical and the invariant $A_2$ classes.' address: - | Stefanie Petermichl\ Department of Mathematics\ University of Texas at Austin\ Austin, TX 78712\ - | Brett D. Wick\ Department of Mathematics\ Vanderbilt University\ 1326 Stevenson Center\ Nashville, TN 37240-0001 author: - 'Stefanie Petermichl$^\dagger$' - 'Brett D. Wick$^\ddagger$' title: 'A Weighted Estimate for the Square Function on the Unit Ball in ${{\mathbb C}}^n$' --- [^1] [^2] Introduction ============ Weighted inequalities for singular integral operators appear naturally in many areas of analysis. The theory of weights is very well understood for the “real analysis” case. In a fundamental paper of Hunt, Muckenhoupt, and Wheeden, [@HMW], it is shown that the so-called Calderón–Zygmund operators in harmonic analysis are bounded on weighted $L^p$ spaces if and only if the weight satisfies the $A_p$ condition. Once this characterization was known, it then became of interest to determine exactly *how* the norms of the operators from harmonic analysis are bounded in terms of the characteristic of the weight. One seeks the smallest power $r=r(p)$ so that $\|Tf\|_{L^p({\omega})}\leq CQ_p(w)^r\|f\|_{L^p({\omega})}$, with $C$ an absolute constant. The best possible power for $p=2$ usually is conjectured to be $1$. Though this is only known for very specific operators. The first successful estimate is for a dyadic analog of the square function, see [@HTV] and the martingale transforms in [@W] shortly after. As for classical Calderón–Zygmund operators, the best bound is only known for specific operators with certain invariance properties, such as the Hilbert and Beurling transform and the Riesz transforms. See [@Phil], [@Priesz] and [@PV]. Such estimates have applications in PDE, see [@FKP] and [@PV], and have received considerable attention. Much of the theory of harmonic analysis can be extended to the boundary of the unit ball ${{\mathbb B}}$ since it is a domain of homogeneous type. Many of the notions make sense on the boundary of the unit ball, but some care is needed. In studying “complex analysis” questions, the choice of metric plays a distinguished role, and in this context one should use the non-isotropic metric. This provides some additional difficulties since the “balls” in this metric are actually elliptical, and this provides some difference in the geometry. So, it makes sense to ask about the boundedness properties of singular integral operators whose natural domain are functions on the boundary of the unit ball, ${{\mathbb S}}$, for example the Cauchy transform. In particular the paper [@LR] demonstrates that the Cauchy transform is bounded on $L^p(w)$ if and only if $w\in A_p$. One can then inquire about other operators from harmonic analysis, such as the square function A major motivation for this note was the paper [@HTV]. In this paper, the authors determined the exact dependence of the norm of the square function (Lusin’s Area Integral, g-functions, etc.) on $L^2(\mathbb{T};w)$ in terms of the invariant characteristic of the weight. We extend the work in [@HTV] to the case of the unit ball and its boundary ${{\mathbb S}}$. It is interesting to note that our proof has an underlying dyadic idea - similar to all other proofs of optimal bounds in weighted spaces. This is so, even though the unit sphere itself lacks a simple dyadic structure. One manages to utilize the dyadic model in one real variable to obtain a result on the unit sphere, despite the differences in geometry. Acknowledgment {#acknowledgment .unnumbered} -------------- The authors would like to thank Texas A & M University and the organizers of the Workshop in Analysis and Probability. Portions of this paper were completed while the authors were in attendance at the workshop during Summer 2006. The authors also thank an observant and skilled referee for many detailed observations and comments. The presentation of the paper benefited greatly. Definitions ----------- Let ${{\mathbb B}}$ denote the unit ball in ${{\mathbb C}}^n$, i.e. ${{\mathbb B}}=\{z \in {{\mathbb C}}^n : |z|<1\}$ and let ${{\mathbb S}}=\{z\in {{\mathbb C}}^n: |z|=1\}$. We write $z=(z_1,...,z_n)$ with $z_k=x_k+iy_k$ and recall that $$\partial_k = \frac{\partial}{\partial z_k}= \frac12 \left(\frac{\partial }{\partial x_k}-i\frac{\partial}{\partial y_k}\right)\text{ and } \bar \partial_k = \frac{\partial}{\partial \bar z_k}= \frac12 \left(\frac{\partial }{\partial x_k}+i\frac{\partial}{\partial y_k}\right).$$ The Bergman kernel on the unit ball is given by $$K(z,w)=\frac{n!}{\pi^n (1-\langle z,w \rangle)^{n+1}}.$$ Define $$g_{ij}(z)=\partial_i \bar\partial_j\log K(z,z)=\frac{n+1}{(1-|z|^2)^2} [(1-|z|^2)\delta_{ij} +\bar z_i z_j]$$ The Bergman metric on ${{\mathbb B}}$ is $$\beta^2(z,\xi)=\sum_{i,j}g_{ij}(z)\xi_i\bar \xi_j.$$ So the volume element associated to the Bergman metric is $$dg(z)=K(z,z)dV(z)= \frac{n!}{\pi^n}\frac{dV(z)}{(1-|z|^2)^{n+1}}=\frac{d\nu (z)}{(1-|z|^2)^{n+1}}.$$Here $V$ is Lebesgue measure on the ball and $\nu$ is normalized Lebesgue measure on ${{\mathbb B}}$. Let $g^{ij}$ be the inverse to the matrix $g_{ij}$, so the Laplace–Beltrami operator on the ball is given by $$\widetilde \Delta = 4\sum_{i,j}g^{ij}\partial_j \bar \partial_i=4\frac{(1-|z|^2)}{n+1}\sum_{i,j}[\delta_{ij}-\bar z_i z_j]\partial_j \bar \partial_i.$$ It also has a radial form, given by $$\widetilde \Delta f(z)=\frac{(1-r^2)}{n+1}[(1-r^2)f''(r)+\frac{2n-r^2-1}{r}f'(r)]$$ if $f(z)=f(|z|)$ for $f\in C^2({{\mathbb B}})$. The invariant gradient of a $C^1({{\mathbb B}})$ function is the vector field given by $$\widetilde \nabla u = 2\sum_{i,j}g^{ij}\left(\bar\partial_i u\ \partial_j+\partial_j u\ \bar\partial_i\right).$$ The Poisson kernel for $\widetilde\Delta$ is given by $P(z,\zeta)=\frac{(1-|z|^2)^n}{|1-\langle z, \zeta \rangle|^{2n}}$ and the Green’s function for $\widetilde \Delta$ is given by $$G(z)=\frac{n+1}{2n}\int_{|z|}^1(1-t^2)^{n-1}t^{-2n+1}dt.$$ For $n=1$ this of course becomes $\log \frac1{|z|}$. Let $f\in L^2({{\mathbb S}})$, define the Poisson–Szegö integral $$\widetilde f(z)=\int_{{{\mathbb S}}}P(z,\zeta)f(\zeta)d\sigma(\zeta)$$ where $\sigma$ is normalized surface measure. For any $\zeta \in {{\mathbb S}}$ we define the Korányi admissible approach region (the analogue of the non-tangential approach region or cone in the unit disc) with aperture $a>0$ as $$\Gamma_a(\zeta) = \{z\in{{\mathbb B}}: |1-\langle z, \zeta \rangle|<a (1-|z|^2)\}.$$ Note that we must assume $a>\frac{1}{2}$ otherwise $\Gamma_a(\zeta)$ is empty. It is well known that $F$ converges to $f$ admissibly almost everywhere. The generalized Lusin area integral or square function on the ball with respect to $\Gamma_a(\xi)$ is $$S_a(f)(\zeta)=\left(\int_{\Gamma_a(\zeta)}|\widetilde \nabla \widetilde f(z)|^2dg(z)\right)^{1/2}.$$ We will be concerned with weighted $L^2$ spaces on the ball. We say that a positive $w \in L^1_{loc}({{\mathbb S}})$ function is in the class $ A_2$ if $$Q_2(w)=\sup_B \left(\frac1{\sigma(B)}\int_B w(\zeta) d\sigma(\zeta) \,\cdot \frac1{\sigma(B)}\int_B w^{-1}(\zeta) d\sigma(\zeta)\right)<\infty ,$$ where the supremum runs over all non-isotropic balls $B$ on ${{\mathbb S}}$. Here, we will be more concerned with the invariant $A_2$ class, denoted by $\widetilde A_2$. A weight is in $\widetilde A_2$ if and only if $$\widetilde Q_2(w)=\sup_{z\in{{\mathbb B}}}\widetilde w(z)\, \widetilde{w^{-1}}(z) <\infty.$$ The quantity $\widetilde Q_2(w)$ is invariant under Möbius transforms and is therefore more suited for questions on the ball. We will see later that these two classes are the same, i.e. that $A_2=\widetilde A_2$ and that $$Q_2(w) \lesssim \widetilde Q_2(w) \lesssim Q_2(w)^2 ,$$ where the implied constants depend upon the dimension $n$. Here and throughout the paper $A\lesssim B$ means $A\leq CB$ for some absolute constant $C$. Let ${\mathcal{L}}^2(w)$ denote the space of measurable functions in the ball that are square integrable with respect to the measure $\widetilde w(z)G(z)dg(z)$. Also let $L^2(w)$ denote the space of measurable functions that are square integrable with respect to the measure $w(\xi)d\sigma(\xi)$. We define the operator $$\widetilde\nabla:L^2(w)\to\mathcal{L}^2(w)$$ by sending the function $f$ to $\widetilde\nabla\widetilde f$ where $\widetilde f$ is the Poisson–Szegö extension. It is an easy calculation that $$\|S_af\|_{L^2(w)}\le c(n)\|\widetilde \nabla \widetilde f \|_{{\mathcal{L}} ^2(w)}.$$ For more information on some of these concepts the reader can consult [@G], [@K], [@R], [@St1], [@St2] or [@S]. Main Results ------------ The main result in this paper is the following: Under the assumptions above, we have: $$\|S_a(f)\|_{L^2(w)} \lesssim \widetilde Q_2(w)\|f\|_{L^2(w)}$$ where the implied constant may depend upon the dimension $n$. Moreover, $$\|\widetilde \nabla(\widetilde f)\|_{{\mathcal{L}}^2(w)}\lesssim \widetilde Q_2(w)\|f\|_{L^2(w)}$$ where we have no dependence on the dimension. Moreover, the dependence upon $\widetilde Q_2(w)$ is sharp for both inequalities. This theorem is proved by showing that the following string of inequalities holds: $$\label{string} \|S_a(f)\|_{L^2(w)}\stackrel{\tiny(1)}{\leq} c(n)\|\widetilde \nabla \widetilde f \|_{{\mathcal{L}} ^2(w)}\stackrel{\tiny(2)}{\leq} c\,\widetilde Q_2\|f\|_{L^2(w)}.$$ Inequality (1) is shown by changing the order of integration, while inequality (2) is demonstrated by Bellman function techniques. The other result demonstrated in this note is the equivalence between the weight classes $A_2$ and $\widetilde A_2$. This is a notable result since the corresponding fact fails in ${{\mathbb R}}^n$. With the notation above, we have the following, $$Q_2(w)\lesssim \widetilde Q_2(w)\lesssim Q_2(w)^2.$$ In particular the classes $A_2$ and $\widetilde A_2$ define the same class of weights on the unit sphere ${{\mathbb S}}$. This result is demonstrated in the last section. The “Area-Integral" Estimate ============================ We prove part $(2)$ of inequality (\[string\]) using Bellman functions. The Bellman function and its properties --------------------------------------- Let us consider the following function $$B(X,x,w,v)=(1+\frac1{Q})X -\frac{x^2}{Qw} -\frac{Q^2x^2}{Q^2w + (4Q^2+1)w-w^2v-4Q^2/v)}$$ on the following domain $$\mathcal{O}=\{(X,x,w,v)>0:x^2<Xw,1\le wv \le Q\}.$$ Note that it is convenient to think of $B=Q^{-1}B_1+B_2$ where $$B_1=X-\frac{x^2}{w}$$ and $$B_2=X-\frac{Q^2x^2}{Q^2w + (4Q^2+1)w-w^2v-4Q^2/v)}.$$ On $\mathcal{O}$, it enjoys the following properties: $$\label{Bsize} 0\le B \le 2X,$$ $$\label{Bconcavity}-d^2B\ge C\frac1{Q^2}v(dx)^2.$$ It is of course very hard to guess such a function. It was taken from [@HTV] where a careful analysis was done to come up with this expression. The properties stated above are a direct calculation, we refer the reader to [@HTV] for detail and briefly sketch here how to get the estimates. The upper estimate on $B$ for (\[Bsize\]) is obvious. For the positivity of $B$, we split $B$ into $B_1$ and $B_2$ and note that $X-x^2/w \ge 0$ and hence it suffices to show that $$\frac{Q^2x^2}{Q^2w + (4Q^2+1)w-w^2v-4Q^2/v)}\le \frac{x^2}{w}.$$ To see the latter, note that it is equivalent to $w^2v^2+4Q^2 \le (4Q^2+1)vw$ which in turn is equivalent to $(vw-2Q)^2\le(2Q-1)^2vw.$ The last inequality is true for $1\le vw\le Q$. To establish (\[Bconcavity\]), observe that $-d^2B\ge \frac2{Q} vx^2(\frac{dx}{x}-\frac{dw}{w})^2$. Also notice that $B_2(X,x,w,v)=B_1(X,x,w')$ where $w'=w+\frac{(4Q^2+1)w-w^2v-4Q^2/v}{Q^2}$. Then use the chain rule to get the estimate $-d^2B_2\gtrsim \frac1{Q^2}vx^2(\frac{dw}{w})^2$. Combining the estimates on the Hessians of $B_1$ and $B_2$ gives $-d^2B\gtrsim \frac1{Q^2} v (dx)^2$. A Dimension-Free Littlewood–Paley Formula ----------------------------------------- On the unit disc one has a Littlewood–Paley formula that says $$\frac1{2\pi}\int_{-\pi}^{\pi}f(e^{i\theta})\bar h(e^{i \theta})d\theta = \frac1{\pi} \int_{{{\mathbb D}}} \langle \nabla \widetilde f(z), \nabla \widetilde h(z) \rangle \log \frac1{|z|}dA(z)$$ as long as $f(0)h(0)=0$. In [@Z] a Littlewood–Paley formula for the unit ball, using the Bergman metric was derived: $$\int_{{{\mathbb S}}}f(\xi)\bar h(\xi) d\sigma(\xi) = C_n \int_{{{\mathbb B}}}\langle \widetilde\nabla \widetilde f(z), \widetilde\nabla \widetilde h(z) \rangle G(z) dg(z)$$ where $f,h \in L^2({{\mathbb S}})$ with $f(0)h(0)=0$. Here, the author was mostly concerned with showing that there be a finite constant $C_n$. This result is not immediate, recall that the ball, equipped with the Bergman metric is not a compact manifold as the metric blows up on ${{\mathbb S}}$. With $u\in C^2({{\mathbb B}})\cap C(\overline{{{\mathbb B}}})$ and under the assumptions above, we have the dimension-free formula $$\int_{{{\mathbb B}}}\widetilde \Delta u(z) G(z)dg(z) =\int_{{{\mathbb S}}}u(\xi) d\sigma(\xi) - u(0),$$ here $\sigma$ is normalized surface measure. The constants in this lemma can be seen by testing Green’s formula for the unit ball ${{\mathbb B}}$ from [@Z] using the radial function $f(z)=f(|z|)=|z|^2$. The main inequality ------------------- Given ${\bf x} \in {{\mathbb R}}^k$ and smooth functions $B({\bf x})$ and $\widetilde{\bf v}(z)=(\widetilde v_1(z),...,\widetilde v_k(z))$ where the $\widetilde v_s$ are Poisson extensions, then we have the following formula for the invariant Laplacian for $b(z)=B(\widetilde{\bf v}(z))$: $$\widetilde \Delta b(z)= 4\sum_{i,j}g^{ij}(d^2B(\widetilde{\bf v}(z))\bar \partial_i \widetilde {\bf v}(z),\overline{\partial_j\widetilde{\bf v}(z)})$$ If $-d^2B({\bf x})\ge F({\bf x}) (dx_s)^2$ then we have the estimate $-\widetilde \Delta b(z)\ge F(\widetilde{\bf v}(z))|\widetilde \nabla \widetilde v_s(z)|^2$. Here $(dx_s)^2$ is the operator represented by the matrix with an entry $1$ in the corresponding to the second derivative in the s$^{th}$ variable and $0$ entries everywhere else. The proof is certainly a direct computation, using harmonicity of the entries of $\widetilde{\bf v}$. Applied to our situation we have $$-\widetilde \Delta b(z)\ge C\frac1{Q^2}\widetilde{w^{-1}} |\widetilde \nabla \widetilde f|^2.$$ Notice that with our choice of $Q=\widetilde Q(w)$ and choices for variables all entries are in our domain $\mathcal{O}$. Using equation (\[Bsize\]) and Green’s formula applied to $b$ we have: $$\begin{aligned} \lefteqn{\int_{{{\mathbb S}}}\frac{f^2}{w}d\sigma \ge b(0) - \int_{{{\mathbb S}}}bd\sigma}\\ &=& \int_{{{\mathbb B}}} -\widetilde \Delta b(z) G(z) dg(z)\\ &\ge& C\frac1{Q^2}\int_{{{\mathbb B}}}|\widetilde \nabla \widetilde f(z)|^2 \widetilde{w^{-1}}(z)G(z)dg(z). \end{aligned}$$ This then proves the second inequality. This estimate is valid for any weight $w\in A_2$. Replacing $w$ by $w^{-1}$ and recalling that $w^{-1}\in A_2$ as well, we arrive at the desired result. The Square Function Estimate ============================ We now turn to (1) of inequality (\[string\]). Let $\mathcal{L}^2(w)$ denote the space of measurable functions on the unit ball ${{\mathbb B}}$ that are square integrable with respect to $\widetilde w(z)G(z)dg(z)$. We have $$\|S_a(f)\|_{L^2(w)}\leq c(n)\|\widetilde\nabla \widetilde f\|_{\mathcal{L}^2(w)}.$$ This follows from the following observation, $\frac{1}{\sigma(Q)}\int_{Q}w(\eta)d\sigma(\eta)\leq c(n)\widetilde w(z)$. Indeed, use the observation that for any non-isotropic ball $Q=Q(\xi,\delta)$ of radius $\delta$ and center $\xi\in{{\mathbb S}}$ there exists a $z_\xi\in{{\mathbb B}}$ such that $\sigma(Q)=(1-{\ensuremath{\vertz_\xi\vert}}^2)^n$ with $\xi= \frac{z_{\xi}}{|z_{\xi}|}$. Additionally, by the triangle inequality for the non-isotropic metric (recall the triangle inequality also holds in ${{\mathbb B}}$), for any $\eta\in Q(\xi,\delta)$ $${\ensuremath{\vert1-\langle z_\xi,\eta\rangle\vert}}^{1/2}\leq{\ensuremath{\vert1-\langle \xi,\eta\rangle\vert}}^{1/2}+{\ensuremath{\vert1-\langle z_\xi,\xi\rangle\vert}}^{1/2}\leq \delta + \sigma(Q)^{\frac1{2n}}.$$ Then for any $Q=Q(\xi,\delta)$, $$\begin{aligned} \frac{1}{\sigma(Q)}\int_Q w(\eta)d\sigma(\eta) & = & \frac{1}{\sigma(Q)}\int_Q \frac{(1-{\ensuremath{\vertz_\xi\vert}}^2)^n}{{\ensuremath{\vert1-\langle z_\xi,\eta\rangle\vert}}^{2n}}\frac{{\ensuremath{\vert1-\langle z_\xi,\eta\rangle\vert}}^{2n}}{(1-{\ensuremath{\vertz_\xi\vert}}^2)^n}w(\eta)d\sigma(\eta)\\ & = & \frac{1}{\sigma(Q)}\int_Q\mathcal{P}_{z_\xi}(\eta)\frac{{\ensuremath{\vert1-\langle z_\xi,\eta\rangle\vert}}^{2n}}{(1-{\ensuremath{\vertz_\xi\vert}}^2)^n}w(\eta)d\sigma(\eta)\\ & \leq & \frac{(\delta + \sigma(Q)^{\frac1{2n}})^{4n}}{\sigma(Q)^2}\int_Q\mathcal{P}_{z_\xi}(\eta)w(\eta)d\sigma(\eta)\\ &=& c_1(n)\widetilde w(z_\xi).\end{aligned}$$ Since $c_l(n)\delta^{2n}\le \sigma(Q)\le c_u(n)\delta^{2n}$ (there is only a comparison, due to the fact the “balls” in this metric are elliptical) then notice that $$c_1(n)\le \frac{(1+c_u^{\frac1{2n}})^{4n}}{c_l(n)^2}.$$ $$\begin{aligned} \lefteqn{\int_{{{\mathbb S}}}S^2_a(f)(\zeta)w(\zeta)d\sigma(\zeta)}\\ & = & \int_{{{\mathbb S}}}\int_{\Gamma_a(\zeta)}|\widetilde \nabla f(z)|^2dg(z)w(\zeta)d\sigma(\zeta)\\ & = & \int_{{{\mathbb B}}}|\widetilde \nabla f(z)|^2\int_{{{\mathbb S}}}\textbf{1}_{\Gamma_a(\zeta)}(z)w(\zeta)d\sigma(\zeta)dg(z).\end{aligned}$$ For fixed $z\in{{\mathbb B}}$, let $$E(z):=\{\zeta\in{{\mathbb S}}:{\ensuremath{\vert1-\langle \frac{z}{|z|},\zeta \rangle\vert}}^{1/2}<(a^{1/2}+1)(1-{\ensuremath{\vertz\vert}}^2)^{1/2}\}.$$ Then one sees by the triangle inequality for the non-isotropic metric that for any $z\in\Gamma_a(\zeta) \Rightarrow \zeta \in E(z)$. Note that $E(z)$ is a non-isotropic ball with center $\frac{z}{|z|}$ and radius $(a^{1/2}+1)(1-{\ensuremath{\vertz\vert}}^2)^{1/2}$. Continuing our estimate we have, $$\begin{aligned} \lefteqn{\int_{{{\mathbb S}}}S^2_a(f)(\zeta)w(\zeta)d\sigma(\zeta)}\\ & = & \int_{{{\mathbb S}}}\int_{\Gamma_a(\zeta)}|\widetilde \nabla \widetilde f(z)|^2dg(z)w(\zeta)d\sigma(\zeta)\\ & = & \int_{{{\mathbb B}}}|\widetilde \nabla \widetilde f(z)|^2\int_{{{\mathbb S}}}\textbf{1}_{\Gamma_a(\zeta)}(z)w(\zeta)d\sigma(\zeta)dg(z)\\ & \leq & \int_{{{\mathbb B}}}|\widetilde \nabla \widetilde f(z)|^2\left(\frac{1}{\sigma(E(z))}\int_{E(z)}w(\zeta)d\sigma(\zeta)\right)\sigma(E(z))dg(z)\\ & \leq & c_1(n) \int_{{{\mathbb B}}}|\widetilde \nabla \widetilde f(z)|^2 \widetilde w(z)\sigma(E(z))dg(z)\\ & \leq & c_1(n)c_u(n)(a^{1/2}+1)^{2n}\int_{{{\mathbb B}}}|\widetilde \nabla \widetilde f(z)|^2 \widetilde w(z)d\sigma(\zeta)(1-{\ensuremath{\vertz\vert}}^2)^ndg(z). \end{aligned}$$ We now use the inequality $(1-{\ensuremath{\vertz\vert}}^2)^n\leq\frac{4n^2}{n+1}G(z)$ in the last integral from above, and continue the estimate: $$\begin{aligned} \int_{{{\mathbb S}}}S^2_a(f)(\zeta)w(\zeta)d\sigma(\zeta) & \leq & c_1(n)c_u(n)(a^{1/2}+1)^{2n} \frac{4n^2}{n+1}\int_{{{\mathbb B}}}|\widetilde \nabla \widetilde f(z)|^2 \widetilde w(z)G(z)dg(z)\\ & = & c(n)\int_{{{\mathbb B}}}|\widetilde \nabla \widetilde f(z)|^2 \widetilde w(z)G(z)dg(z). \end{aligned}$$ This proves inequality (1) in (\[string\]). Sharpness of the linear dependence on $\widetilde Q_2({\omega})$. ================================================================= To see that our estimate is sharp, we supply a family of examples. We utilize power weights adapted to the non-isotropic metric on the sphere. Let ${\omega}_{\alpha}(\xi)=|1-\langle \xi, \eta_0 \rangle|^{\alpha}$ for some fixed $\eta_0 \in {{\mathbb S}}$. Choose $f_{\alpha}(\xi)=|1-\langle \xi, \eta_0 \rangle|^{-\alpha}$. One can observe that ${\omega}_{\alpha} \in \widetilde A_2$ if and only if $|\alpha|<n-1$. To see this, one integrates using the following formula (valid for $n\geq2$) from [@S]: $$\int_{{{\mathbb S}}} g(\langle \xi,\eta\rangle)d\sigma(\xi) = \frac{n-1}{\pi}\int_{{{\mathbb T}}}\int_{0}^1(1-r^2)^{n-2}g(re^{i\theta}) rdr d\theta$$ with $g(z)=|1-z|^{\alpha}$. The largest contribution of the integrand in our case appears near $\eta_0$. Recall that $\widetilde Q_2({\omega}_{\alpha})=\sup_z \widetilde{\omega}_{\alpha}(z)\widetilde{{\omega}^{-1}_{\alpha}}(z)$. The supremum here is attained for $z$ on the ray from 0 to $\eta_0$. One sees that for $\alpha \to n-1$ we have $\widetilde Q_2({\omega}_{\alpha})\sim (n-1-\alpha)^{-1}$. Similarly one calculates that $\|f_{\alpha}\|_{L^2({\omega}_{\alpha})} \sim (n-1-\alpha)^{-1/2}$. It remains to see that $\|\widetilde\nabla \widetilde f_\alpha\|_{\mathcal{L}^2(w)}$, $\|S_af_{\alpha}\|_{L^2({\omega}_{\alpha})} \gtrsim (n-1-\alpha)^{-3/2}$. Letting $\alpha \to n-1$ then establishes sharpness. To see this, one computes the invariant gradient of the Poisson extension of the function $f_{\alpha}$ directly. One then establishes the weighted norm of the area integral similar to [@HTV] just more computationally involved. The Comparison of Classical and Invariant $A_2$ =============================================== In this section we establish the that the class $A_2$ and $\widetilde A_2$ are in fact the same. This is shown by demonstrating that $$Q_2(w)\lesssim \widetilde Q_2(w) \lesssim Q_2^2(w)$$ where the implied constants only depend upon the dimension. We begin with the following fact. First, some notation. Let $$f _Q=\frac{1}{\sigma(Q)}\int_Q f(\xi)d\sigma(\xi)\quad w(Q)=\int_Qw(\xi)d\sigma(\xi).$$ Now, observe that by Cauchy-Schwarz $$f_Q^2 w(Q) \le Q_2(w)(f^2w)(Q).$$ If $P\subset Q$, and applying the above to $\textbf{1}_{Q\setminus P}$ we get $$\left(1-\frac{\sigma(P)}{\sigma(Q)}\right)^2\leq Q_2(w)(w(Q)-w(P))$$ Upon rearrangement we arrive at: $$\label{betterthandoubling} w(P)\le \left(1-\frac{(1-\sigma(P)/\sigma(Q))^2}{Q_2(w)}\right)w(Q).$$ Given any non-isotropic ball $Q$ on the sphere, choose $z_Q\in {{\mathbb B}}$ so that the center of $Q$ is $z_Q/|z_Q|$ and its volume is $(1-|z_Q|)^n$ (hence its radius is $\sim (1-|z_Q|)^{1/2}$). There is a one-to-one correspondence between $Q$’s and $z_Q$’s. $\widetilde w(z_Q)$ will approximately correspond to $w_Q$. We estimate $$\begin{aligned} \lefteqn{\widetilde w(z_Q)\widetilde{w^{-1}}(z_Q)\ge \left(\int_{Q} P_{z_Q}wd\sigma \right) \left(\int_{Q} P_{z_Q}w^{-1}d\sigma \right) }\\ &\gtrsim& \left(\frac{(1-|z_Q|^2)^{n}}{(1-|z_Q|)^{2n}}\right)^2 \left(\int_{Q}wd\sigma \right) \left(\int_{Q} w^{-1}d\sigma \right)\\ &\gtrsim & \frac1{(1-|z_Q|)^{2n}}\left(\int_{Q} wd\sigma \right) \left(\int_{Q} w^{-1}d\sigma \right)\\ &\gtrsim& Q_2(w).\end{aligned}$$ Hence $Q_2(w)\lesssim \widetilde Q_2(w)$. We now turn to the other inequality. We have to estimate $\widetilde w(z)\widetilde{w^{-1}}(z)\lesssim Q_2^2(w)$ with implied constant independent of $w$ and $Q$. Using the Möbius invariance of $\widetilde Q_2$ it suffices to assume $z=(r,0,\ldots,0)$ with $0<r<1$ arbitrary. Now fix $r$ and let $Q$ be the non-isotropic ball with center $(1,0,\ldots,0)$ and $\sigma(Q)=(1-r)^n$. To obtain the upper estimate, we exhaust the sphere by enlarging $Q$. Let us denote by $cQ$ the ball with the same center and $c^n$-fold volume (hence $c^{1/2}$-fold radius). $$\int_{{{\mathbb S}}}P_{z}wd\sigma = \int_QP_{z}wd\sigma + \sum_{k\ge 1} \int_{2^kQ\setminus 2^{k-1}Q} P_{z}w d\sigma$$ Observe that for all $\xi \in {{\mathbb S}}$ $$P_{z}(\xi)\le\frac{(1+r)^n}{(1-r)^n}.$$ Moreover, if $\xi \in 2^kQ \setminus 2^{k-1}Q$ we have the better estimate $$P_{z}(\xi)\lesssim 2^{-2nk}\frac{(1+r)^n}{(1-r)^n}.$$ So we get $$\widetilde w(z) \lesssim \frac{(1+r)^n}{(1-r)^n}\sum_{k\ge 0}2^{-2kn}w(2^kQ)$$ and similarly for $\widetilde{w^{-1}}$. We start to estimate the product $$\begin{aligned} \lefteqn{\widetilde w(z)\widetilde{w^{-1}}(z) \lesssim \frac{(1+r)^{2n}}{(1-r)^{2n}} \sum_{k\ge 0}2^{-2nk}w(2^kQ) \cdot \sum_{l\ge 0}2^{-2nl}w^{-1}(2^lQ) }\\ &\le& \frac{(1+r)^{2n}}{(1-r)^{2n}} \sum_{k\ge 0} 2^{-2nk}w(2^kQ)\sum_{l=0}^k 2^{-2nl}w^{-1}(2^lQ)\\ &&+\frac{(1+r)^{2n}}{(1-r)^{2n}} \sum_{k\ge 0} 2^{-2nk}w(2^kQ)\sum_{l=k}^{\infty} 2^{-2nl}w^{-1}(2^lQ).\end{aligned}$$ The latter sums are equivalent (with roles of $w$ and $w^{-1}$ switched) so we estimate the second sum only. Iterating the estimate (\[betterthandoubling\]) we estimate for $l\ge k$ $$w(2^kQ)\le \left(1-\frac{1-4^{-n}}{Q_2(w)}\right)^{l-k}w(2^lQ),$$ and using that $\sigma (2^kQ) = 2^{nk}(1-r)^n$ we get $$\begin{aligned} \lefteqn{\frac{(1+r)^{2n}}{(1-r)^{2n}} \sum_{k\ge 0} 2^{-2nk}w(2^kQ)\sum_{l=k}^{\infty} 2^{-2nl}w^{-1}(2^lQ)}\\ &\lesssim& Q_2(w)\sum_{k\ge 0}2^{-2nk}\sum_{s\ge 0}\left(1-\frac{1-4^{-n}}{Q_2(w)}\right)^s\\ &\lesssim&Q_2^2(w).\end{aligned}$$ Concluding Remarks ================== The result in [@LR] establishes that the Cauchy transform on $L^p(w)$ is bounded if and only if the weight is in $A_p$. It is however not known how the norm of the Cauchy transform grows with respect to the characteristic of the weight. Thus, it is an interesting task to understand the dependence of the norm of the Cauchy transform on weighted $L^2$ spaces of the unit sphere by means of Bellman functions. On the unit disc, using Bellman function techniques, a very simple proof of the boundedness of the Cauchy transform on weighted $L^2$ spaces is given, see [@NT] for continuity and [@PW] for the sharp result in terms of the invariant characteristic. In the case of the disc the sharp weighted bound for the square function was a first step and a tool to obtain the sharp bound for the Cauchy transform, which was a major motivation for this note. As for the Cauchy transform $C$, the goal is to provide a linear and sharp estimate of the form $\|C f\|_{L^2({\omega})}\lesssim \widetilde Q_2({\omega})\|f\|_{L^2({\omega})}$ where the implied constant is independent of $f$, ${\omega}$ and the dimension $n$. The claim is that the invariant characteristic of the weight is the correct one, so that the dependence is both linear and no dependence on the dimension occurs. Our dimension-free estimate on the “area integral” illustrates that this guess was a good one. The missing ingredient for the estimate for the Cauchy transform is the following formula: $|\widetilde \nabla C f|\sim|\widetilde \nabla \widetilde f|$, which holds in one dimension, but is not true in several variables. The estimate for the square function avoids this deficiency. Let $w\in A_2$ and $C$ denote the Cauchy transform in ${{\mathbb C}}^n$. Then, is it true that $$\|C(f)\|_{L^2(w)}\lesssim \widetilde Q_2(w)\|f\|_{L^2(w)},$$ where the implied constant does not depend upon the dimension? [30]{} , The Theory of Weights and the Dirichlet Problem for Elliptic Equations, Ann. of Math. 134 (1991), 65–124. , Bounded Analytic Functions, Acad. Press, New York, 1981. , The Bellman Functions and Sharp Weighted Inequaities for Square Functions, in Oper. Theory Adv. Appl. 113, pp. 97-113, Birkhaeuser, Basel, 2000. , Weighted Norm Inequalities for the Conjugate Function and Hilbert Transform, Trans. Amer. Math. Soc. 176 (1973), 227–251. , Function Theory of Several Complex Variables, AMS Chelsea Press, Providence, 2001. , Weighted Norm Inequalities for Pluriharmonic Conjugate Functions, J. Math. Anal. Appl. 268 (2002), 707–717. , The Weighted Norm Inequalities for the Hilbert Transform are Now Trivial, C. R. Acad. Sci. Paris, Série I 323 (1996), 717–722. , The Sharp Bound for the Hilbert Transform on Weighted Lebesgue Spaces in terms of the Classical $A_p$ Characteristic, to appear in Amer. J. of Math.. , The Sharp Weighted Bound for the Riesz Transforms, to apprear in Proc. Amer. Math. Soc.. , Heating of the Beurling Operator: Weakly Quasiregular Maps on the plane are Quasiregular, Duke Math. J. Vol. 112, No 2 (2002), pp. 281-305. , A sharp weighted estimate on the norm of Hilbert transform via invariant $A_2$ characteristic of the weight, Mich. Math. J. 50 (2002), pp. 71-87. , Function Theory in the Unit Ball of ${{\mathbb C}}^n$, Springer-Verlag, New York, 1980. , Boundary Behavior of Holomorphic Functions of Several Complex Variables”, Princeton University Press, 1972. , Harmonic Analysis: Real-Variable Methods, Orthogonality, and Oscillatory Integrals, with the assistance of Timothy S. Murhpy, Princeton Math., 1993. , Invariant Potential Theory in the Unit Ball of ${{\mathbb C}}^n$, Cambridge University Press, 1994. , A Sharp Estimate on the Norm of the Martingale Transform, Math. Res. Lett. v.7 (2000) N1, pp. 1-12. , Toeplitz Operators and Hankel Operators on the Hardy Space of the Unit Sphere, J. Func. Anal. 149 (1997), 1–24. [^1]: $\dagger$Research supported in part by a National Science Foundation Grant. [^2]: $\ddagger$Research supported in part by a National Science Foundation RTG Grant to Vanderbilt University.
{ "pile_set_name": "ArXiv" }
--- abstract: 'This article gives solutions to the exercises in Bestvina and Feighn’s paper [@BF03] on Sela’s work on limit groups. We prove that all constructible limit groups are limit groups and give an account of the shortening argument of Rips and Sela.' author: - Henry Wilton bibliography: - 'exercises.bib' date: 16th March 2006 title: 'Solutions to Bestvina & Feighn’s Exercises on Limit Groups' --- Mladen Bestvina and Mark Feighn’s beautiful first set of notes [@BF03] on Zlil Sela’s work on the Tarski problems (see [@Se1] *et seq.*) provides a very useful introduction to the subject. It gives a clear description of the construction of Makanin–Razborov diagrams, and precisely codifies the structure theory for limit groups in terms of *constructible limit groups (CLGs)*. Furthermore, the reader is given a practical initiation in the subject with exercises that illustrate the key arguments. This article is intended as a supplement to [@BF03], to provide solutions to these exercises. Although we do give some definitions in order not to interrupt the flow, we refer the reader to [@BF03] for all the longer definitions and background ideas and references. Definitions and elementary properties {#Introduction} ===================================== In this section we present solutions to exercises 2, 3, 4, 5, 6 and 7, which give some of the simpler properties and the first examples and non-examples of limit groups. $\omega$-residually free groups ------------------------------- Fix ${\mathbb{F}}$ a free group of rank $r>1$. A finitely generated group $G$ is *$\omega$-residually free* if, for any finite subset $X\subset G$, there exists a homomorphism $h:G\rightarrow{\mathbb{F}}$ whose restriction to $X$ is injective. (Equivalently, whenever $1\notin X$ there exists a homomorphism $h:G\to{\mathbb{F}}$ so that $1\notin h(X)$.) Residually free groups inherit many of the properties of free groups; the first and most obvious property is being torsion-free. \[LGs are TF\] Any residually free group is torsion-free. Let $G$ be $\omega$-residually free (indeed $G$ can be thought of as merely residually free). Then for any $g\in G$, there exists a homomorphism $h:G\rightarrow{\mathbb{F}}$ with $h(g)\neq 1$; so $h(g^k)\neq 1$ for all integers $k$, and $g^k\neq 1$. It is immediate that any subgroup of an $\omega$-residually free group is $\omega$-residually free (exercise 6 of [@BF03]). That the choice of ${\mathbb{F}}$ does not matter follows from the observation that all finitely generated free groups are $\omega$-residually free. \[Free groups are LGs\] Let $F$ be a finitely generated free group. Realize ${\mathbb{F}}$ as the fundamental group of a rose $\Gamma$ with $r$ petals; that is, the wedge of $r$ circles. Then $\Gamma$ has an infinite-sheeted cover that corresponds to a subgroup $F'$ of ${\mathbb{F}}$ of countably infinite rank. The group $F$ can be realized as a free factor of $F'$; this exhibits an injection $F\hookrightarrow{\mathbb{F}}$. In particular, every free group is $\omega$-residually free. \[Free abelian groups are LGs\] Let $A$ be a finitely generated free abelian group, and let $a_1,\ldots,a_n\in A$ be non-trivial. Fix a basis for $A$, and consider the corresponding inner product. Let $z\in A$ be such that $\langle z,a_i\rangle\neq 0$ for all $i$. Then inner product with $z$ defines a homomorphism $A\to{\mathbb{Z}}$ so that the image of every $a_i$ is non-trivial, as required. Examples \[Free groups are LGs\] and \[Free abelian groups are LGs\] give exercise 3 of [@BF03]. Limit groups ------------ Groups that are $\omega$-residually free are natural examples of limit groups. Let ${\mathbb{F}}$ be as above, and $\Gamma$ a finitely generated group. A sequence of homomorphisms $(f_n:\Gamma\rightarrow{\mathbb{F}})$ is *stable* if, for every $g\in G$, $f_n(g)$ is either eventually $1$ or eventually not $1$. The *stable kernel* of a stable sequence of homomorphisms $(f_n)$ consists of all $g\in G$ with $f_n(g)$ eventually trivial; it is denoted $\underrightarrow{ker}f_n$. A *limit group* is a group arising as a quotient $\Gamma/\underrightarrow{ker}f_n$ for $(f_n)$ a stable sequence. \[ORFs are LGs\] Every $\omega$-residually free group is a limit group. Let $G$ be an $\omega$-residually free group. Fix a generating set, and let $X_n\subset G$ be the ball of radius $n$ about the identity in the word metric. Let $f_n:G\rightarrow{\mathbb{F}}$ be a homomorphism that is injective on $X_n$. Now $f_n$ is a stable sequence and the stable kernel is trivial, so $G$ is a limit group. In fact every limit group is $\omega$-residually free (lemma 1.11 of [@BF03]). Henceforth, we shall use the terms interchangeably. Negative examples ----------------- Let’s see some examples of groups that aren’t limit groups. The first three examples are surface groups that aren’t even residually free. It follows from lemma \[LGs are TF\] that the fundamental group of the real projective plane is not a limit group. A slightly finer analysis yields some other negative examples. \[2-generator LGs\] The only 2-generator residually free groups are the free group of rank 2 and the free abelian group of rank 2. Let $G$ be a residually free group generated by $x$ and $y$. If $G$ is non-abelian then $[x,y]\neq 1$ so there exists a homomorphism $f:G\rightarrow{\mathbb{F}}$ with $f([x,y])\neq 1$. So $f(x)$ and $f(y)$ generate a rank 2 free subgroup of ${\mathbb{F}}$. Therefore $G$ is free. In particular, the fundamental group of the Klein bottle is not a limit group. Only one other surface group fails to be $\omega$-residually free. This was first shown by R. S. Lyndon in [@L59]. \[Lyndon’s example\] Let $\Sigma$ be the closed surface of Euler characteristic -1. Then any homomorphism $f_*:\pi_1(\Sigma)\to{\mathbb{F}}$ has abelian image. In particular, since $\pi_1(\Sigma)$ is not abelian, it is not residually free. Let $\Gamma$ be a bouquet of circles so ${\mathbb{F}}=\pi_1(\Gamma)$. Realize the homomorphism $f_*$ as a map from $\Sigma$ to $\Gamma$, which we denote by $f$. Our first aim is to find an essential simple closed curve in the kernel of $f_*$. Consider $x$ the mid-point of an edge of $\Gamma$. Altering $f$ by a homotopy, it can be assumed that $f$ is transverse at $x$; in this case, $f^{-1}(x)$ is a collection of simple closed curves. Let $\gamma$ be such a curve. If $\gamma$ is null-homotopic in $\Sigma$ then $\gamma$ can be removed from $f^{-1}(x)$ by a homotopy. If all components of the pre-images of all midpoints $x$ can be removed in this way then $f_*$ was the trivial homomorphism. Otherwise, any remaining such component $\gamma$ lies in the kernel of $f$ as required. We proceed with a case-by-case analysis of the components of $\Sigma\smallsetminus\gamma$. 1. If $\gamma$ is 2-sided and separating then, by examining Euler characteristic, the components of $\Sigma\smallsetminus\gamma$ are a punctured torus or a Klein bottle, together with a Möbius band. So $f$ factors through the one-point union $T\vee {\mathbb{R}}P^2$ or $K\vee{\mathbb{R}}P^2$. In either case, it follows that the image is abelian. 2. If $\gamma$ is 2-sided and non-separating then $\Sigma-\gamma$ is the non-orientable surface with Euler characteristic -1 and two boundary components, so $f_*$ factors through ${\mathbb{Z}}*{\mathbb{Z}}/2{\mathbb{Z}}$ and hence through ${\mathbb{Z}}$. 3. If $\gamma$ is 1-sided then $\gamma^2$ is 2-sided and separating, and case 1 applies. This finishes the proof. Lemmas \[LGs are TF\], \[2-generator LGs\] and \[Lyndon’s example\] give exercise 4 of [@BF03]. Here is a more interesting obstruction to being a limit group. A group $G$ is *commutative transitive* if every non-trivial element has abelian centralizer; equivalently, if $[x,y]=[y,z]=1$ then $[x,z]=1$. Note that ${\mathbb{F}}$ is commutative transitive, since every non-trivial element has cyclic centralizer. \[Abelian centralizers\] Limit groups are commutative transitive. Let $G$ be $\omega$-residually free, let $g\in G$, and suppose $a,b\in G$ commute with $g$. Then there exists a homomorphism $$f:G\rightarrow{\mathbb{F}}$$ injective on the set $\{1,g,[a,b]\}$. Then $f([g,a])=f([g,b])=1$; since ${\mathbb{F}}$ is commutative transitive, it follows that $$f([a,b])=1.$$ So $[a,b]=1$, as required. A stronger property also holds. A subgroup $H\subset G$ is *malnormal* if, whenever $g\notin H$, $gHg^{-1}\cap H=1$. The group $G$ is *CSA* if every maximal abelian subgroup is malnormal. If $G$ is CSA then $G$ is commutative transitive. For, let $g\in G$ with centralizer $Z(g)$. Consider maximal abelian $A\subset Z(g)$ and $h\in Z(g)$. Then $g\in hAh^{-1}\cap A$, so $h\in A$. Therefore $Z(g)=A$. \[LGs are CSA\] Limit groups are CSA. Let $H\subset G$ be a maximal abelian subgroup, consider $g\in G$, and suppose there exists non-trivial $h\in gHg^{-1}\cap H$. Let $f:G\rightarrow{\mathbb{F}}$ be injective on the set $$\{1,g,h,[g,h]\}.$$ Then $f([h,ghg^{-1}])=1$, which implies that $f(h)$ and $f(ghg^{-1})$ lie in the same cyclic subgroup. But in a free group, this is only possible if $f(g)$ also lies in that cyclic subgroup; so $f([g,h])=1$, and hence $[g,h]=1$. By lemma \[Abelian centralizers\] it follows that $g$ commutes with every element of $H$, so $g\in H$. Embeddings in real algebraic groups =================================== In this section we provide solutions to exercises 8 and 9 of [@BF03], which show how to embed limit groups in real algebraic groups and also $PSL_2({\mathbb{R}})$, and furthermore give some control over the nature of the embeddings. First, we need a little real algebraic geometry. By, for example, proposition 3.3.13 of [@BCR], every real algebraic variety $V$ has an open dense subset $V_\mathrm{reg}\subset V$ with finitely many connected components, so that every component $V'\subset V_\mathrm{reg}$ is a manifold. \[Regular components\] Consider a countable collection $V_1,V_2,\ldots\subset V$ of closed subvarieties. Then for any component $V'$ of $V_\mathrm{reg}$ as above, either there exists $k$ so that $V'\subset V_k$, or $$V'\cap\bigcup_{i=1}^\infty V_i$$ has empty interior. Suppose $V'\cap\bigcup_i V_i$ doesn’t have empty interior. Then, by Baire’s Category Theorem, there exists $k$ such that $V'\cap V_k$ doesn’t have empty interior. Consider $x$ in the closure of the interior of $V'\cap V_k$ and let $f$ be an algebraic function on $V'$ that vanishes on $V_k$. Then $f$ has zero Taylor expansion at $x$, so $f$ vanishes on an open neighbourhood of $x$. In particular, $x$ lies in the interior of $V'\cap V_k$. So the interior of $V'\cap V_k$ is both open and closed, and $V'\subset V_k$ since $V'$ is connected. \[Embedding in PSL2\] Let ${\mathcal{G}}$ be an algebraic group over ${\mathbb{R}}$ in which ${\mathbb{F}}$ embeds. Then for any limit group $G$ there exists an embedding $$G\hookrightarrow{\mathcal{G}}.$$ In particular, $G$ embeds into $SL_2({\mathbb{R}})$ and $SO(3)$. Consider the variety $V={\mathrm{Hom}}(G,{\mathcal{G}})$. (If $G$ is of rank $r$, then $V$ is a subvariety of ${\mathcal{G}}^r$, cut out by the relations of $G$. By Hilbert’s Basis Theorem, finitely many relations suffice.) For each $g\in G$, consider the subvariety $$V_g=\{f\in V|f(g)=1\}.$$ If $G$ does not embed into ${\mathcal{G}}$ then $V$ is covered by the subvarieties $V_g$ for $g\neq 1$. By lemma \[Regular components\] every component of $V_\mathrm{reg}$ is contained in some $V_g$, so $$V=V_{g_1}\cup\ldots\cup V_{g_n}$$ for some non-trivial $g_1,\ldots,g_n\in G$. So every homomorphism from $G$ to ${\mathcal{G}}$ kills one of the $g_i$. But ${\mathbb{F}}$ embeds in ${\mathcal{G}}$, so this contradicts the assumption that $G$ is $\omega$-residually free. Given a limit group $G$ and an embedding $f:G\hookrightarrow SL_2({\mathbb{R}})$ we have a natural map $G\to PSL_2({\mathbb{R}})$. This is also an embedding since any element in its kernel satisfies $f(g)^2=1$ and $G$ is torsion-free. We can gain more control over embeddings into $PSL_2({\mathbb{R}})$ by considering the trace function. If $G$ is a limit group and $g_1,\ldots,g_n\in G$ are non-trivial then there is an embedding $G\hookrightarrow PSL_2({\mathbb{R}})$ whose image has no non-trivial parabolic elements, and so that the images of $g_1,\ldots,g_n$ are all hyperbolic. We abusively identify each element of the variety $V={\mathrm{Hom}}(G,SL_2({\mathbb{R}}))$ with the corresponding element of ${\mathrm{Hom}}(G,PSL_2({\mathbb{R}}))$, and call it elliptic, hyperbolic or parabolic accordingly. For each $g\in G$, consider the closed subvariety $U_g$ of homomorphisms that map $g$ to a parabolic, and the open set $W_g$ of homomorphisms that map $g$ to a hyperbolic. (Note that $\gamma\in SL_2({\mathbb{R}})$ is parabolic if $|\mathrm{tr}\gamma|=2$ and hyperbolic if $|\mathrm{tr}\gamma|<2$.) Fix an embedding ${\mathbb{F}}\hookrightarrow SL_2({\mathbb{R}})$ whose image in $PSL_2({\mathbb{R}})$ is the fundamental group of a sphere with open discs removed; such a subgroup is called a *Schottky group*, and every non-trivial element is hyperbolic. Call a component $V'$ of $V_\mathrm{reg}$ *essential* if its closure contains a homomorphism $G\to SL_2({\mathbb{R}})$ that factors through ${\mathbb{F}}\hookrightarrow SL_2({\mathbb{R}})$ and maps the $g_i$ non-trivially. Suppose every essential component $V'$ of $V_\mathrm{reg}$ is contained in some $U_{g'}$ for non-trivial $g'$. Then, since there are only finitely many components, for certain non-trivial $g'_1,\ldots,g'_m$, every homomorphism $G\to{\mathbb{F}}$ kills one of the $g_i$ or one of the $g'_j$, contradicting the assumption that $G$ is $\omega$-residually free. Therefore, by lemma \[Regular components\], there exists an essential component $V'$ so that $V'\cap\bigcup_{g\neq 1}U_g$ has empty interior. In particular, $$\bigcap_i W_{g_i}\smallsetminus \bigcup_{g\neq 1} U_g$$ is non-empty, as required. GADs for limit groups ===================== In this section we provide solutions to exercises 10 and ll, and also the related exercise 17. For the definitions of the modular group ${\mathrm{Mod}}(G)$, and of generalized Dehn twists, see definitions 1.6 and 1.17 respectively in [@BF03]. \[Generalized Dehn twists\] ${\mathrm{Mod}}(G)$ is generated by inner automorphisms and generalized Dehn twists. Since the mapping class group of a surface is generated by Dehn twists (see, for example, [@CB88]), it only remains to show that unimodular automorphisms of abelian vertices are generated by generalized Dehn twists. Given such a vertex $A$, we can write $$G=A*_{\bar{P}(A)} B$$ for some subgroup $B$ of $G$. Any unimodular automorphism of $A$ is a generalized Dehn twist of this splitting. \[Elliptic abelian subgroups\] Let $M$ be a non-cyclic maximal abelian subgroup of a limit group $G$. 1. If $G=A*_C B$ for $C$ abelian then $M$ is conjugate into either $A$ or $B$. 2. If $G=A*_C$ with $C$ abelian then either $M$ is conjugate into $A$ or there is a stable letter $t$ so that $M$ is conjugate to $M'=\langle C,t\rangle$ and $G=A*_C M'$. We first prove 1. Suppose $M$ is not elliptic in the splitting $G=A*_C B$. (Note that we don’t yet know that $M$ is finitely generated.) Non-cyclic abelian groups have no free splittings, so $C$ is non-trivial. Let $T$ be the Bass–Serre tree of the splitting. Either $M$ fixes an axis in $T$, or it fixes a point on the boundary. In the latter case, there is an increasing chain of edge groups $$C_1\subset C_2\subset\ldots M.$$ But every $C_i$ is a conjugate of $C_1$, and since $M$ is malnormal it follows that $C_i=C_1$. So $M=C_i$, contradicting the assumption that $M$ is not elliptic. If $M$ fixes a line in $T$ then $M$ can be conjugated to $M'$ fixing a line $L$ so that $A$ stabilizes a vertex $v$ of $L$ and $M'$ is of the form $M'=C\oplus{\mathbb{Z}}$; $C$ fixes $L$ pointwise, and ${\mathbb{Z}}$ acts as translations of $L$. Consider the edges of $L$ incident at $v$, corresponding to the cosets $C$ and $aC$, for some $a\in A\smallsetminus C$. Since $aCa^{-1}=C$ and $C$ is non-trivial, it follows from lemma \[LGs are CSA\] that $a\in M'$. But $a$ is elliptic so $a\in C$, a contradiction. In the HNN-extension case, assuming $M$ is not elliptic in the splitting we have as before that $M$ preserves a line in the Bass–Serre tree $T$. Conjugating $M$ to $M'$, we may assume $C$ fixes an edge in the preserved line $L$, so $C\subset M'$. The stabilizer of an adjacent edge is of the form $(ta)C(ta)^{-1}$, where $a\in A$ and $t$ is the stable letter of the HNN-extension. Therefore $C=(ta)C(ta)^{-1}$, so since $G$ is CSA it follows that $M'=C\oplus \langle ta\rangle $ and $$G=A*_C M'$$ as required. \[Remark on CSA assumptions\] Note that, in fact, the proof of lemma \[LGs are CSA\] only used that the vertex groups are CSA and that the edge group was maximal abelian on one side. A one-edge splitting of $G$ is said to satisfy *condition JSJ* if every non-cyclic abelian group is elliptic in it. Recall that a limit group is *generic* if it is freely indecomposable, non-abelian and not a surface group. If $G$ is a generic limit group then ${\mathrm{Mod}}(G)$ is generated by inner automorphisms and generalized Dehn twists in one-edge splittings satisfying JSJ. Furthermore, the only generalized Dehn twists that are not Dehn twists can be taken to be with respect to a splitting of the from $G=A*_C B$ with $A=C\oplus{\mathbb{Z}}$. By lemma \[Generalized Dehn twists\], ${\mathrm{Mod}}(G)$ is generated by inner automorphisms and generalized Dehn twists. By lemma \[Elliptic abelian subgroups\], any splitting of $G$ as an amalgamated product satisfies JSJ. Consider, therefore, the splitting $$G=A*_C.$$ A (generalized) Dehn twist $\delta_z$ in this splitting fixes $A$ and maps the stable letter $t\mapsto tz$, for some $z\in Z_G(C)$. Suppose that this HNN-extension doesn’t satisfy JSJ, so there exists some (without loss, maximal) abelian subgroup $M$ that is not elliptic in the splitting. By lemma \[Elliptic abelian subgroups\], after conjugating $M$ to $M'$, we have that $$G=A*_C M'$$ and $M'=C\oplus\langle t\rangle$ where $t$ is the stable letter. Since $M'=Z_G(C)$, a Dehn twist $\delta_z$ along $z$ (for $c\in C$ and $n\in{\mathbb{Z}}$) fixes $A$ and maps $$t\mapsto z+t.$$ But this is a generalized Dehn twist in the amalgamated product. So ${\mathrm{Mod}}(G)$ is, indeed, generated by generalized Dehn twists in one-edge splittings satisfying JSJ. Any generalized Dehn twist $\delta$ that is not a Dehn twist is in a splitting of the form $$G=A*_C B$$ with $A$ abelian, and acts as a unimodular automorphism on $A$ that preserves $\bar{P}(A)$. Recall that $A/\bar{P}(A)$ is finitely generated, by remark 1.15 of [@BF03]. To show that it is enough to use splittings in which $A=C\oplus{\mathbb{Z}}$, we work by induction on the rank of $A/\bar{P}(A)$. Write $A=A'\oplus{\mathbb{Z}}$, where $\bar{P}(A)\subset A'$. Then there is a modular automorphism $\alpha$, agreeing with $\delta$ on $A'$, generated by generalized Dehn twists of the required form, by induction. Now $\delta$ and $\alpha$ differ by a generalized Dehn twist in the splitting $$G=A*_{A'} (A'*_C B)$$ which is of the required form. Constructible Limit Groups ========================== For the definition of CLGs see [@BF03]. The definition lends itself to the technique of proving results by a nested induction, first on level and then on the number of edges of the GAD $\Delta$. To prove that CLGs have a certain property, this technique often reduces the proof to the cases where $G$ has a one-edge splitting over groups for which the property can be assumed. In this section we provide solutions to exercises 12, 13, 14 and 15, which give the first properties of CLGs, culminating in the result that all CLGs are $\omega$-residually free. CLGs are CSA ------------ We have seen that limit groups are CSA. This section is devoted to proving that CLGs are also CSA. Knowing this will prove extremely useful in deducing the other properties of CLGs. Note that the property of being CSA passes to subgroups. \[CLGs are CSA\] CLGs are CSA. By induction on the number of edges in the graph of groups $\Delta$, it suffices to consider a CLG $G$ such that $$G=A*_C B$$ or $$G=A*_C$$ where each vertex group is assumed to be CSA and the edge group is taken to be maximal abelian on one side. (In the first case, we will always assume that $C$ is maximal abelian in $A$.) First, we have an analogue of lemma \[Elliptic abelian subgroups\]. \[CLGs Elliptic abelian subgroups\] Let $G$ be as above. Then $G$ decomposes as an amalgamated product or HNN-extension in such a way that all non-cyclic maximal abelian subgroups are conjugate into a vertex group. Furthermore, in the HNN-extension case we have that $C\cap C^t=1$ where $C$ is the edge group and $t$ is the stable letter. By induction, we can assume that the vertex groups are CSA. Note that the proof of the first assertion of lemma \[Elliptic abelian subgroups\] only relies on the facts that the vertex groups are CSA and the edge group is maximal abelian in one vertex group. So the amalgamated product case follows. Now consider the case of an HNN-extension. Suppose that, for some stable letter $t$, $C\cap C^t$ is non-trivial. Then since the vertex group $A$ is commutative transitive, it follows that $C\subset C^t$ (or $C^t\subset C$, in which case replace $t$ by $t^{-1}$). If $\rho:G\to G'$ is the retraction to a lower level then, since $G'$ can be taken to be CSA, $\rho(C^t)=\rho(C)$. But $\rho$ is injective on edge groups so $C=C^t$ and, furthermore, $t$ commutes with $C$. Otherwise, for every choice of stable letter $t$, $C\cap C^t$ is trivial. In either case, the result now follows as in the proof of the second assertion of lemma \[Elliptic abelian subgroups\]. Recall that a simplicial $G$-tree is *$k$-acylindrical* if the fixed point set of every non-trivial element of $G$ has diameter at most $k$. \[Acylindricality\] In the graph-of-groups decomposition given by lemma \[CLGs Elliptic abelian subgroups\], the Bass–Serre tree is 2-acylindrical. In the amalgamated product case this is because, for any $a\in A\smallsetminus C$, $C\cap C^a=1$. Likewise, in the HNN-extension case this is because $C\cap C^t=1$ where $t$ is the stable letter. [lemma \[CLGs are CSA\]]{} Let $M\subset G$ be a maximal abelian subgroup and suppose $$1\neq m\subset M^g\cap M.$$ Let $T$ be the Bass–Serre tree of the splitting. If $M$ is cyclic then it might act as translations on a line $L$ in $T$. Then $g$ also maps $L$ to itself. But it follows from acylindricality that any element that preserves $L$ lies in $M$; so $g\in M$ as required. We can therefore assume that $M$ fixes a vertex $v$ of $T$. If $g$ also fixes $v$ then, since the vertex stabilizers are CSA, $g\in M$. Consider the case when $$G=A*_C B;$$ the case of an HNN-extension is similar. Then without loss of generality $M\subset A$ and $g=ba$ for some $a\in A$ and $b\in B$ so $M^g$ fixes a vertex stabilized by $A^b$ for some $b\in B$. Since $C$ is maximal abelian in $A$ we have $C=M$ and since $B$ is CSA and $m\in C\cap C^b$ it follows that $b$ commutes with $C$ so $b\in M$, hence $g\in M$. Abelian subgroups ----------------- It is no surprise that CLGs share the most elementary property of limit groups. \[CLGs are torsion free\] CLGs are torsion-free. The freely decomposable case is immediate by induction. Therefore assume $G$ is a freely indecomposable CLG of level $n$, with $\Delta$ and $\rho:G\rightarrow G'$ as in the definition. Suppose $g\in G$ is of finite order. Then $g$ acts elliptically on the Bass–Serre tree of $\Delta$, so $g$ lies in a vertex group. Clearly if the vertex is QH then $g=1$, and by induction if the vertex is rigid then $g=1$. Suppose therefore that the vertex is abelian. Then $g$ lies in the peripheral subgroup. But $\rho$ is assumed to inject on the peripheral subgroup, so by induction $g$ is trivial. However, it is far from obvious that limit groups have abelian subgroups of bounded rank; indeed, it is not obvious that all abelian subgroups of limit groups are finitely generated. But this is true of CLGs. \[Exercise 13 in [@BF03]\] \[Abelian subgroups of CLGs\] Abelian subgroups of CLGs are free, and there is a uniform (finite) bound on their rank. The proof starts by induction on the level of $G$. Let $G$ be a CLG. Since non-cyclic abelian subgroups have no free splittings we can assume $G$ is freely indecomposable. Let $\Delta$ be a generalized abelian decomposition. Let $T$ be the Bass–Serre tree of $\Delta$. If $A$ fixes a vertex of $T$ then the result follows by induction on level. Otherwise, $A$ fixes a line $T_A$ in $T$, on which it acts by translations. The quotient $\Delta'=T_A/A$ is topologically a circle; after some collapses $\Delta'$ is an HNN-extension; so the rank of $A$ is bounded by the maximum rank of abelian subgroups of the vertex groups plus 1. So by induction the rank of $A$ is uniformly bounded. That $A$ is free follows from lemma \[CLGs are torsion free\]. Heredity -------- Let $\Sigma$ be a (not necessarily compact) surface with boundary. Then a boundary component $\delta$ is a circle or a line, and defines up to conjugacy a cyclic subgroup $\pi_1(\delta)\subset\pi_1(\Sigma)$. These are called the *peripheral subgroups* of $\Sigma$. \[Free splittings of non-compact surfaces\] Let $\Sigma$ be a non-compact surface with non-abelian fundamental group. Then there exists a non-trivial free splitting of $\pi_1(\Sigma)$, with respect to which all peripheral subgroups are elliptic. \[Exercise 12 in [@BF03]\] \[Subgroups of CLGs\] Let $G$ be a CLG of level $n$ and $H$ a finitely generated subgroup. Then $H$ is a free product of finitely many CLGs of level at most $n$. The subgroup $H$ can be assumed to be freely indecomposable by Grushko’s theorem, so we can also assume that $G$ is freely indecomposable. Let $\Delta$ and $\rho:G\rightarrow G'$ be as in the definition. Then subgroup $H$ inherits a graph-of-groups decomposition from $\Delta$, namely the quotient of the Bass–Serre tree $T$ by $H$. Since $H$ is finitely generated, it is the fundamental group of some finite core $\Delta'\subset T/H$. Every vertex of $\Delta'$ covers a vertex of $\Delta$, from which it inherits its designation as QH, abelian or rigid. The edge groups of $\Delta'$ are subgroups of the edge groups of $\Delta$, so they are abelian and $\rho$ is injective on them. Furthermore, it follows from lemma \[CLGs are CSA\] that $H$ is commutative transitive, so each edge group of $\Delta'$ is maximal abelian on one side of the associated one-edge splitting. Let $V'$ be a vertex group of $\Delta'$, a subgroup of the vertex group $V$ of $\Delta$. There are three case to consider. 1. $V'\subset V$ are abelian. Since every map $f':V'\to{\mathbb{Z}}$ with $f'(P(V'))=0$ extends to a map $f:V\to{\mathbb{Z}}$ with $f(P(V))=0$ we have that $$\bar{P}(V')\subset\bar{P}(V)$$ so $\rho$ is injective on $\bar{P}(V')$. 2. $V'\subset V$ are QH. If $V'$ is of infinite index in $V$ then $V'$ is the fundamental group of a non-compact surface, so by remark \[Free splittings of non-compact surfaces\] $H$ is freely decomposable. Therefore it can be assumed that $V'$ is of finite degree $m$ in $V$. In particular, $V$ is the fundamental group of a compact surface that admits a pseudo-Anosov automorphism. Furthermore, let $g,h\in V$ be such that $\rho([g,h])\neq 1$. Then because CLGs are commutative transitive, $$\rho([g^m,h^m])\neq 1.$$ But $g^m,h^m\in V'$, so $\rho(V')$ is non-abelian. 3. $V'\subset V$ are rigid. Then $\tilde{V}'\subset \tilde{V}$ because CLGs are commutative transitive, so $\rho|_{V'}$ is injective. Therefore $\rho|_H:H\rightarrow G'$ and $\Delta'$ satisfy the properties for $H$ to be a CLG. Coherence --------- A group is *coherent* if every finitely generated subgroup is finitely presented. Note that free groups and free abelian groups are coherent. For limit groups, coherence is an instance of a more general phenomenon, as in the next lemma. Recall that a group is *slender* if every subgroup is finitely generated. Finitely generated abelian groups are slender. \[Graphs of coherent groups\] The fundamental group of a graph of groups with coherent vertex groups and slender edge groups is coherent. Let $\Delta$ be a graph of groups, with coherent vertex groups and slender edge groups. Let $G=\pi_1(\Delta)$ and $H\subset G$ a finitely generated subgroup. Then $H$ inherits a graph-of-groups decomposition from $\Delta$ given by taking the quotient of the Bass–Serre tree $T$ of $\Delta$ by the action of $H$. Since $H$ is finitely generated it is the fundamental group of some finite core $\Delta'\subset T/H$. But, by induction on the number of edges, $H=\pi_1(\Delta')$ is finitely presented. \[CLGs are coherent\] CLGs are coherent, in particular finitely presented. In the case of a free decomposition the result is immediate. In the other case, the (free abelian) edge groups of $\Delta$ are finitely generated, so slender and coherent, by lemma \[Abelian subgroups of CLGs\]. Therefore all vertex groups are finitely generated; in particular, abelian vertex groups are coherent. Finitely generated surface groups are also coherent. Rigid vertex groups embed into a CLG of lower level, so by lemma \[Subgroups of CLGs\] they are free products of coherent groups and hence coherent by induction. The result now follows by lemma \[Graphs of coherent groups\]. Finite $K(G,1)$ --------------- That CLGs have finite $K(G,1)$ follows from the fact that graphs of aspherical spaces are aspherical. \[Aspherical graphs of groups\] Let $\Delta$ be a graph of groups; suppose that for every vertex group $V$ there exists finite $K(V,1)$, and for every edge group $E$ there exists a finite $K(E,1)$. Then for $G=\pi_1(\Delta)$, there exists a finite $K(G,1)$. Surface groups and abelian groups have finite Eilenberg–Mac Lane spaces. Rigid vertices embed into a CLG of lower level, so by lemma \[Subgroups of CLGs\] and induction they also have finite Eilenberg–Mac Lane spaces. \[Exercise 13 in [@BF03]\] \[CLGs have finite K(G,1)\] If $G$ is a CLG then there exists a finite $K(G,1)$. Principal cyclic splittings --------------------------- A *principal cyclic splitting* of $G$ is a one-edge splitting of $G$ with cyclic edge group, such that the image of the edge group is maximal abelian in one of the vertex groups; further, if it is an HNN-extension then the edge group is required to be maximal abelian in the whole group. The key observation about principal cyclic splittings is that any non-cyclic abelian subgroup is elliptic with respect to them—in other words, they are precisely those cyclic splittings that feature in the conclusion of lemma \[CLGs Elliptic abelian subgroups\]. Applying lemma \[CLGs Elliptic abelian subgroups\], to prove that every freely indecomposable, non-abelian CLG has a principal cyclic splitting it will therefore suffice to produce any non-trivial cyclic splitting (since we now know that CLGs are CSA). \[Principal cyclic splittings of CLGs\] Every non-abelian, freely indecomposable CLG admits a principal cyclic splitting. Let $G$ be a CLG. As usual, by induction it suffices to consider the cases when $G$ splits as an amalgamated product or HNN-extension. It suffices to exhibit any cyclic splitting of $G$, as observed above. Suppose $$G=A*_C B.$$ If $C$ is cyclic the result is immediate, so assume $C$ is non-cyclic abelian. If either vertex group is freely decomposable then so is $G$, since $C$ has no free splittings; if both vertex groups are abelian then so is $G$. Therefore $A$, say, is freely indecomposable and non-abelian so has a principal cyclic splitting, which we shall take to be of the form $$A=A'*_{C'} B'.$$ (It might also be an HNN-extension, but this doesn’t affect the proof.) Because it is principal $C$ is conjugate into a vertex, say $B'$; so $G$ now decomposes as $$G=A'*_{C'} (B'*_C B).$$ which is a cyclic splitting as required. The proof when $G=A*_C$ is the same. A criterion in free groups -------------------------- To prove that a group $G$ is $\omega$-residually free, it suffices to show that for any finite $X\subset G\smallsetminus 1$ there exists a homomorphism $f:G\rightarrow{\mathbb{F}}$ with $1\notin f(X)$. So a criterion to show that an element of ${\mathbb{F}}$ is not the identity will be useful. \[Free group criterion\] Let $z\in{\mathbb{F}}\smallsetminus 1$, and consider an element $g$ of the form $$g=a_0z^{i_1} a_1z^{i_2}a_2\ldots a_{n-1}z^{i_n}a_n$$ where $n\geq 1$ and, whenever $0<k<n$, $[a_k,z]\neq 1$. Then $g\neq 1$ whenever the $|i_k|$ are sufficiently large. Choose a generating set for ${\mathbb{F}}$ so the corresponding Cayley graph is a tree $T$. An element $u\in{\mathbb{F}}$ specifies a geodesic $[1,u]\subset T$. Likewise, a string of elements $u_0,u_1,\ldots,u_n\in{\mathbb{F}}$ defines a path $$[1,u_0]\cdot u_0[1,u_1]\cdot\ldots\cdot (u_0\ldots u_{n-1})[1,u_n]$$ in $T$, where $\cdot$ denotes concatenation of paths. The key observation we will use is as follows. The length of a word $w\in{\mathbb{F}}$ is denoted by $|w|$. \[Key observation\] Suppose $z$ is cyclically reduced and has no proper roots. Let $a\in{\mathbb{F}}$ be such that $a$ and $az$ both lie in $L\subset T$ the axis of $z$. If $j$ is minimal such that $z^j$ lies in the geodesic $[a,az]$ then, setting $u=z^ja^{-1}$ and $v=az^{1-j}$, it follows that $uv=z=vu$; in particular, either $u$ or $v$ is trivial and $[a,z]=1$. [lemma \[Free group criterion\]]{} It can be assumed that $z$ is cyclically reduced and has no proper roots. Assume that, for each $k$, $|z^{i_k}|\geq |a_{k-1}|+|a_k|+|z|$. Let $L\subset T$ be the axis of $z$. Denote by $g_k$ the partial product $$g_k=a_0z^{i_1} a_1z^{i_2}a_2\ldots a_{k-1}z^{i_k}a_k.$$ The path $\gamma$ corresponding to $g$ is of the form $$[1,a_0]\cdot g_0[1,z^{i_1}]\cdot g_0z^{i_1}[1,a_1]\cdot\ldots\cdot g_{n-1}[1,z^{i_n}]\cdot g_{n-1}z^{i_n}[1,a_n].$$ Suppose that $g=1$ so this path is a loop. Each section of the form $g_k[1,z^{i_k}]$ lies in a translate of $L$, the axis of $z$. Since $T$ is a tree, for at least one such section $\gamma$ enters and leaves $g_kL$ at the same point—otherwise $\gamma$ is a non-trivial loop. Since $|z^{i_k}|>|a_{k-1}|+|a_k|+|z|$ it follows that both $g_kz^{i_k}a_k$ and $g_kz^{i_k}a_kz$ lie in $g_kL$ and so $[a_k,z]=1$ by remark \[Key observation\]. CLGs are limit groups --------------------- \[CLGs are LGs\] CLGs are $\omega$-residually free. Since the freely decomposable case is immediate, let $\Delta$, $G'$ and $\rho$ be as in the definition of a CLG in [@BF03]. By induction, $G'$ can be assumed $\omega$-residually free. As a warm up, and for use in the subsequent induction, we first prove the result in the case of abelian and surface vertices. \[Abelian case\] Let $A$ be a free abelian group and $\rho:A\rightarrow G'$ a homomorphism to a limit group. Suppose $P\subset A$ is a nontrivial subgroup of finite corank closed under taking roots, on which $\rho$ is injective. Then for any finite subset $X\subset A\smallsetminus1$ there exists an automorphism $\alpha$ of $A$, fixing $P$, so that $1\notin \rho\circ\alpha(X)$. Since $\ker\rho$ is a subgroup of $A$ of positive codimension, for given $x\in A\smallsetminus 0$ a generic automorphism $\alpha$ certainly satisfies $\alpha(x)\notin\ker\rho$. Since $X$ is finite, therefore, there exists $\alpha$ such that $\alpha(x)\notin\ker\rho$ for any $x\in X$. We now consider the surface vertex case. \[Surface case\] Let $S$ be the fundamental group of a surface $\Sigma$ with non-empty boundary, with $\chi(\Sigma)\leq -1$, and $\rho:S\rightarrow G'$ a homomorphism injective on each peripheral subgroup and with non-abelian image. Then for any finite subset $X\subset S\smallsetminus1$ there exists an automorphism $\alpha$ of $S$, induced by an automorphism of $\Sigma$ fixing the boundary components pointwise, such that $1\notin\rho\circ\alpha(X)$. Let the surface $\Sigma$ have $b>0$ boundary components and Euler characteristic $\chi<0$. When $\Sigma$ is cut along a two-sided simple closed curve $\gamma$, the resulting pieces either have lower genus (defined to be $1-\frac{1}{2}(\chi+b)$) or fundamental groups of strictly lower rank, depending on whether $\gamma$ was separating or not. The simplest cases all have fundamental groups that are free of rank 2. \[Simplest cases\] Suppose $S$ is free of rank 2. By lemma \[2-generator LGs\], $\rho(S)$ is free or free abelian; but $\rho(S)$ is assumed non-abelian, so $\rho(S)$ is free and $\rho$ is injective. For the more complicated cases, the idea is to find a suitable simple closed curve $\zeta$ along which to cut to make the surface simpler. In order to apply the proposition inductively, $\zeta$ needs the following properties: 1. $\rho(\zeta)\neq 1$; 2. the fundamental group $S'$ of any component of $\Sigma\smallsetminus\zeta$ must have $\rho(S')$ non-abelian. Let’s find this curve in some examples. ![A four-times punctured sphere](punctured){width="70.00000%"} \[Punctured spheres\] Suppose $\Sigma$ is a punctured sphere, so $$S=\langle d_1,\ldots,d_n|\prod_j d_j\rangle.$$ Assume $n\geq 4$ and $\rho(d_i)\neq 1$ for all $i$. Define a relation on $\{1,\ldots,n\}$ by $$i\sim j \Leftrightarrow \rho([d_i,d_j])= 1.$$ Since $G'$ is commutative transitive, $\sim$ is an equivalence relation. Because the image is non-abelian, there are at least two equivalence classes. Since $$\prod_i d_i=1$$ any equivalence class has at least two elements in its complement. Relabelling if necessary, it can now be assumed that $$\rho([d_1,d_2]),\rho([d_3,d_4])\neq 1.$$ Now if the boundary curves have been coherently oriented then $d_1 d_2$ has a representative that is a simple closed curve. Take $\zeta$ as this representative. The case when $\Sigma$ is non-orientable is closely related. \[Non-orientable surfaces\] Suppose $\Sigma$ is non-orientable so $$S=\langle c_1,\ldots,c_m,d_1,\ldots,d_n|\prod_i c_i^2\prod_jd_j\rangle.$$ Exactly the same argument as in the case of a punctured sphere would work if it could be guaranteed that $\rho(c_i)\neq 1$ for all $i$. Fix some $c_k$, therefore, and suppose $\rho(c_k)=1$. Let $\gamma$ be a simple closed curve representing $c_k$. Then $d_1c_k$ has a representative $\delta$ which is a simple closed curve, and $\rho(d_1c_k)\neq 1$. Furthermore, $\Sigma\smallsetminus\gamma$ and $\Sigma\smallsetminus\delta$ are homeomorphic surfaces, and a homeomorphism between them extends to an automorphism of $\Sigma$ mapping $\gamma$ to $\delta$. This homeomorphism can be chosen not to alter any of the other $c_i$ or the $d_j$. Therefore, after an automorphism of $\Sigma$, it can be assumed that $\rho(c_i)\neq 1$ for all $i$, so a suitable $\zeta$ can be found as in the previous example. ![The positive-genus case.](handle){width="70.00000%"} \[Positive genus\] Suppose $\Sigma$ is an orientable surface of positive genus, so $$S=\langle a_1,b_1,\ldots,a_g,b_g,d_1,\ldots,d_n|\prod_i[a_i,b_i]\prod_j d_j \rangle.$$ Assume that $g,n\geq 1$. If, for example, $\rho(a_1)\neq 1$ then $\zeta$ can be taken to be a simple closed curve representing $a_1$. Otherwise, $\rho(a_1d_1)\neq 1$ and $a_1d_1$ has a simple closed representative. It remains to show that the single component of $\Sigma\smallsetminus\zeta$ has non-abelian image. Cutting along $\zeta$ expresses $S$ as an HNN-extension: $$S=S'*_{\mathbb{Z}}.$$ Let $t$ be the stable letter, and suppose $\rho(S')$ is abelian. Then $\zeta\in tS't^{-1}\cap S'$ so in particular $\rho(tS't^{-1}\cap S')$ is non-trivial. But $G'$ is a limit group and hence CSA, so $\rho(t)$ commutes with $\rho(S')$, contradicting the assumption that $\rho(S)$ is non-abelian. Note that examples \[Punctured spheres\], \[Non-orientable surfaces\] and \[Positive genus\] cover all the more complicated surfaces with boundary. [proposition \[Surface case\]]{} Example \[Simplest cases\] covers all the simplest cases. Suppose therefore that $\Sigma$ is more complicated. To apply the inductive hypothesis, an essential simple closed curve $\zeta\in\Sigma$ is needed such that $\rho(\zeta)\neq 1$ and, for any component $S'$ of $\Sigma\smallsetminus\zeta$, $\rho(S')$ is non-abelian. This is provided by examples \[Punctured spheres\], \[Non-orientable surfaces\] and \[Positive genus\]. For simplicity, assume $\zeta$ is separating. The non-separating case is similar. Then $\Sigma\smallsetminus\zeta$ has two components, $\Sigma_1$ and $\Sigma_2$. Let $S_i=\pi_1(\Sigma_i)$, and denote by $X_i$ the *syllables* of $X$ in $S_i$—the elements of $S_i$ that occur in the normal form of some $x\in X$ with respect to the splitting over $\langle\zeta\rangle$. Because the pieces $\Sigma_i$ are simpler than $\Sigma$ there exists $\alpha\in{\mathrm{Aut}}_0(\Sigma)$ and $f:G'\rightarrow{\mathbb{F}}$ such that $$1\notin f\circ\rho\circ\alpha([\zeta,X_1\cup X_2]).$$ Consider $\xi\in X$. The proposition follows from the claim that, for all sufficiently large $k$, $$f\circ\rho\circ\delta_\zeta^k\circ\alpha(\xi)\neq 1$$ where $\delta_\zeta$ is a Dehn twist in $\zeta$. If $\xi$ is a power of $\zeta$ then the result is immediate. Otherwise, with respect to the one-edge splitting of $G$ over $\langle\zeta\rangle$, $\xi$ has reduced form $$\sigma_0\tau_0\sigma_1\tau_1\ldots\sigma_n\tau_n$$ where the $\sigma_i\in X_1$ and the $\tau_j\in X_2$. The image $x^{(k)}=f\circ\rho\circ\delta_\zeta^k\circ\alpha(\xi)$ is of the form $$z^ks_0z^{-k}t_0z^ks_1z^{-k}t_1\ldots z^ks_nz^{-k}t_n$$ where $z=f\circ\rho(\zeta)$, $s_i=f\circ\rho\circ\alpha(\sigma_i)$ and $t_i=f\circ\rho\circ\alpha(\tau_i)$. This expression for $x^{(k)}$ satisfies the hypotheses of lemma \[Free group criterion\], so $x^{(k)}\neq 1$ for all sufficiently large $k$. The proof of theorem \[CLGs are LGs\] is very similar to the proof of proposition \[Surface case\]. The theorem follows from the following proposition, by induction on level. \[Heart of the matter\] Let $G$ be a freely indecomposable CLG, let $G'$ be $\omega$-residually free, and let $\Delta$ and $\rho$ be as usual. For any finite subset $X\subset G\smallsetminus1$ there exists a modular automorphism $\alpha$ of $G$ such that $1\notin \rho\circ\alpha(X)$. As usual, the proposition is proved by induction on the number of edges of $\Delta$. The case of $\Delta$ having no edges follows from lemmas \[Abelian case\] and \[Surface case\], and the fact that $\rho$ is injective on rigid vertices. By induction on level, $G'$ is a limit group. Now suppose $\Delta$ has an edge group $E$. For simplicity, assume $E$ is separating. The non-separating case is similar. Then removing the edge corresponding to $E$ divides $\Delta$ into two subgraphs $\Delta_1$ and $\Delta_2$. Let $G_i=\pi_1(\Delta_i)$, and denote by $X_i$ the syllables of $X$ in $G_i$. Without loss assume $E$ is maximal abelian in $G_1$. Fix non-trivial $\zeta\in E$. By induction there exists $\alpha\in{\mathrm{Mod}}(\Delta)$ and $f:G'\rightarrow{\mathbb{F}}$ such that $$1\notin f\circ\rho\circ\alpha([\zeta,X_1]\cup X_2).$$ Consider $\xi\in X$. The proposition follows from the claim that, for all sufficiently large $k$, $$f\circ\rho\circ\delta_\zeta^k\circ\alpha(\xi)\neq 1.$$ If $\xi\in E$ then the result is immediate. Otherwise, with respect to the one-edge splitting of $G$ over $E$, $\xi$ has reduced form $$\sigma_0\tau_0\sigma_1\tau_1\ldots\sigma_n\tau_n$$ where the $\sigma_i\in X_1$ and the $\tau_j\in X_2$. The image $x^{(k)}=f\circ\rho\circ\delta_\zeta^k\circ\alpha(\xi)$ is of the form $$z^ks_0z^{-k}t_0z^ks_1z^{-k}t_1\ldots z^ks_nz^{-k}t_n$$ where $z=f\circ\rho(\zeta)$, $s_i=f\circ\rho\circ\alpha(\sigma_i)$ and $t_i=f\circ\rho\circ\alpha(\tau_i)$. In particular, canceling across those $t_i$ that commute with $z$, we have $$x^{(k)}=u_0z^{k\epsilon_1}u_1\ldots u_{n-1}z^{k\epsilon_n}u_n$$ where $\epsilon_i=\pm1$ and $u_i$ don’t commute with $z$ for $0<i<n$. This second expression for $x^{(k)}$ satisfies the hypotheses of lemma \[Free group criterion\], so $x^{(k)}\neq 1$ for all sufficiently large $k$. The Shortening Argument {#Shortening argument section} ======================= We consider a sequence of $G$-trees $T_i$, arising from homomorphisms $f_i:G\to{\mathbb{F}}$, that converge in the Gromov topology to a $G$-tree $T$. By the results of section 3 of [@BF03], if the action of $G$ on the limit tree $T$ is faithful then it gives rise to a generalized abelian decomposition for $G$. This section is entirely devoted to the solution of exercise 16, which is essentially Rips and Sela’s shortening argument—an ingenious means of using this generalized abelian decomposition to force the action on the limit tree to be unfaithful. Preliminary ideas {#Shortening argument subsection} ----------------- Once again, fix a generating set for ${\mathbb{F}}$ so that the corresponding Cayley graph is a tree, and let $|w|$ denote the length of a word $w\in{\mathbb{F}}$. Fix a generating set $S$ for $G$. For $f:G\rightarrow{\mathbb{F}}$, let $$|f|=\max_{g\in S}|f(g)|.$$ A homomorphism is *short* if $$|f|\leq |\iota\circ f\circ\alpha|$$ whenever $\alpha$ is a modular automorphism of $G$ and $\iota$ is an inner automorphism of ${\mathbb{F}}$. \[Shortening argument\] Suppose every $f_i$ is short. Then the action on $T$ is not faithful. The proof is by contradiction. We assume therefore, for the rest of section \[Shortening argument section\], that the action is faithful. By the results summarized in section 3 of [@BF03], the action of $G$ on $T$ gives a GAD $\Delta$ for $G$. The idea is, if $T_i$ are the limiting trees with basepoints $x_i$, to construct modular automorphisms $\phi_n$ so that $$d_i(x_i,f_i\circ\phi_i(g)x_i)<d_i(x_i,f_i(g)x_i)$$ for all sufficiently large $i$. Then apply these automorphisms to carefully chosen basepoints. All constructions of the limit tree $T$, such as the asymptotic cone [@vdDW], use some form of based convergence: basepoints $x_i\in T_i$ are fixed, and converge to a basepoint $[x_i]\in T$. Because the $f_i$ are short, $$\max_{g\in S}d_{\mathbb{F}}(1,f_i(g))\leq\max_{g\in S}d_{\mathbb{F}}(t,f_i(g)t)$$ for all $t\in T_{\mathbb{F}}$; otherwise, conjugation by the element of ${\mathbb{F}}$ nearest to $t$ leads to a shorter equivalent homomorphism. It follows that $1\in T_i$ is always a valid basepoint; we set $x=[1]\in T$ to be the basepoint for $T$. The proof of theorem \[Shortening argument\] goes on a case-by-case basis, depending on whether $[x,gx]$ intersects a simplicial part or a minimal part of $T$. The abelian part ---------------- The next proposition is a prototypical shortening result for a minimal vertex. \[Proposition 5.2\] Let $V$ be an abelian vertex group of $\Delta$. For $g\in G$, let $l(g)$ be the translation length of $g$ on $T$. Fix $\epsilon>0$. Then for any finite subset $S\subset V$ there exists a modular automorphism $\phi$ of $G$ such that $$\max_{g\in S}l(\phi(g))<\epsilon.$$ The minimal $V$-invariant subtree $T_V$ is a line in $T$, on which $V$ acts indiscretely. Since $S$ is finite, $V$ can be assumed finitely generated. It suffices to prove the theorem in the case where $S$ is a basis for $V$. Assume furthermore that each element of $S$ translates $T_V$ in the same direction. Suppose the action of $V$ on $T_V$ is free. Let $S=\{g_1,\ldots,g_n\}$, ordered so that $$l(g_1)>l(g_2)>\ldots>l(g_n)>0.$$ Since the action is indiscrete, there exists an integer $\lambda$ such that $$l(g_1)-\lambda l(g_2)<\frac{1}{2}l(g_2).$$ Applying the automorphism that maps $g_1\mapsto g_1-\lambda g_2$ and proceeding inductively, we can make $l(g_1)$ as short as we like. If the action of $V$ is not free then $V=V'\oplus V_0$ where $V'$ acts freely on $T_V$ and $V_0$ fixes $T_V$ pointwise. Applying the free case to $V'$ gives the result. The aim is to prove the following theorem. \[Abelian shortening\] Let $V$ be an abelian vertex. Then for any finite subset $S\subset G$ there exists a modular automorphism $\phi$ such that for any $g\in S$: 1. if $[x,gx]$ intersects a translate of $T_V$ in a segment of positive length then $$d(x,\phi(g)x)<d(x,gx);$$ 2. otherwise, $\phi(g)=g$. By a result of J. Morgan (claim 3.3 of [@Mo]—the article is phrased in terms of laminations), the path $[x,gx]$ intersects finitely many translates of $T_V$ in non-trivial segments. Let $\epsilon$ be the minimal length of all such segments across all $g\in S$. Assume that $g\in S$ is such that $[x,gx]$ intersects a translate of $T_V$ non-trivially. Suppose first that $x$ lies in a translate of $T_V$, so without loss of generality $x\in T_V$. Then $g$ has a non-trivial decomposition in the GAD provided by corollary 3.16 of [@BF03] of the form $$g=a_0b_1a_1\ldots a_n$$ where the $a_i$ lie in $V$ and the $b_i$ are products of elements of other vertices and loop elements. Write $g_i=a_0b_1\ldots b_{i-1}a_{i-1}$. The decomposition can be chosen so that each component of the geodesic $[x,gx]$ that lies in $g_iT_V$ is non-trivial. For each $i$, decompose $[x,b_ix]$ as $$[x,s_i]\cdot[s_i,t_i]\cdot[t_i,b_ix]$$ where $[x,s_i]$ and $[t_i,b_ix]$ are maximal segments in $T_V$ and $b_iT_V$ respectively. Then $$[x,gx]=[x,a_0s_1]\cdot g_1[s_1,t_1]\cdot g_1[t_1,b_1a_1s_2]\cdot\ldots\cdot g_n[s_n,t_n]\cdot g_n[t_n,b_na_nx]\\$$ where each $[s_i,t_i]$ and $[t_i,b_ia_is_{i+1}]$ is a non-trivial segment. Therefore $$\begin{aligned} d(x,gx) & = & d(x,a_0s_1)+\sum_{i=1}^n d(s_i,t_i)+\sum_{i=1}^{n-1}d(t_i,b_ia_is_{i+1})+d(t_n,b_na_nx)\\ & \geq & \sum_{i=1}^n d(s_i,t_i) + (n+1)\epsilon.\end{aligned}$$ Since $V$ acts indiscretely on the line $T_V$, by modifying the $b_i$ by elements of $V$ it can be assumed that $$d(x,s_i),d(t_i,b_ix)<\frac{1}{4}\epsilon.$$ By proposition \[Proposition 5.2\] there exists $\phi\in{\mathrm{Mod}}(G)$ such that $\phi(b_i)=b_i$ for all $b_i$ and $$d(x,\phi(a_i)x)<\frac{1}{2}\epsilon$$ for all $a_i$. Now as before $[x,\phi(g)x]$ decomposes as $$[x,\phi(a_0)s_1]\cdot \phi(g_1)[s_1,t_1]\cdot\ldots\cdot\phi(g_n)[t_n,b_n\phi(a_n)x]$$ so $$\begin{aligned} d(x,\phi(g)x) & = & d(x,\phi(a_0)s_1)+\sum_{i=1}^n d(s_i,t_i)+\sum_{i=1}^{n-1}d(t_i,b_i\phi(a_i)s_{i+1})\\& &+d(t_n,b_n\phi(a_n)x)\\ & < & d(x,\phi(a_0)x)+\sum_{i=1}^n d(s_i,t_i)+\sum_{i=1}^{n-1}d(b_ix,b_i\phi(a_i)x)\\& &+d(b_nx,b_n\phi(a_n)x)+\frac{n}{2}\epsilon\\ & < & \sum_{i=1}^n d(s_i,t_i)+ \big(n+\frac{1}{2}\big)\epsilon.\end{aligned}$$ Therefore, $d(x,\phi(g)x)<d(x,gx)-\frac{1}{2}\epsilon$ and in particular the result follows. Now suppose $x$ does not lie in a translate of $T_V$. Then $g$ has a non-trivial decomposition in the corresponding GAD of the form $$g=b_0a_1b_1\ldots b_n$$ where the $a_i$ lie in $V$ and the $b_i$ are products of elements of other vertices and loop elements. Let $g'=a_1b_1\ldots b_{n-1}a_n$. Let $x'$ be the first point on $[x,gx]$ in a translate of $T_V$, so $x'=b_0y\in b_0T_V$ for some $y\in T_V$. Likewise let $x''$ be the last point on $[x,gx]$, so $x''=b_0g'z\in b_0g'T_V$ for some $z\in T_V$. Since the action of $V$ on $T_V$ is indiscrete we can modify $b_n$ by an element of $V$ and assume that $d(y,z)<\frac{1}{4}\epsilon$. Then the geodesic $[x,gx]$ decomposes as $$[x,gx]=[x,b_0y]\cdot[b_0y,b_0g'z]\cdot[b_0g'z,gx]$$ so $$d(x,gx)> d(x,b_0y)+d(y,g'y)+d(z,b_nx)-\frac{1}{4}\epsilon.$$ Applying the first case to $g'$ and $y$ we obtain $\phi\in{\mathrm{Mod}}(G)$ such that $$d(y,\phi(g')y)<d(y,g'y)-\frac{1}{2}\epsilon$$ so $$\begin{aligned} d(x,\phi(g)x)&<&d(x,b_0y)+d(y,\phi(g')y)+d(z,b_nx)+\frac{1}{4}\epsilon\\ &<& d(x,b_0y)+d(y,g'y)-\frac{1}{2}\epsilon+d(z,b_nx)+\frac{1}{4}\epsilon\\ &<& d(x,gx)\end{aligned}$$ as required. The surface part ---------------- The surface part is dealt with by Rips and Sela, in [@RS94], in the following theorem. \[Surface shortening\] Let $V$ be a surface vertex. Then for any finite subset $S\subset G$ there exists a modular automorphism $\phi$ such that for any $g\in S$: 1. if $[x,gx]$ intersects a translate $T_V$ in a segment of positive length then $$d(x,\phi(g)x)<d(x,gx);$$ 2. otherwise, $\phi(g)=g$. Rips and Sela use the notion of groups of interval exchange transformations, which are equivalent to surface groups, and prove an analogous result to proposition \[Proposition 5.2\]. The rest of the proof is the same as that of theorem \[Abelian shortening\]. The simplicial part ------------------- It remains to consider the case where $[x,gx]$ is contained in the simplicial part of $T$. \[Simplicial shortening\] Let $S\subset G$ be finite and let $x\in T$. Then there exist $\phi_n\in{\mathrm{Mod}}(\Delta)$ such that, for all $g\in S$, $$d(x,\phi_n(g)x)=d(x,gx);$$ furthermore, for all $g\in S$ that do not fix $x$, and for all sufficiently large $n$, $$d_n(x_n,f_n\circ\phi_n(g)x_n)<d_n(x_n,f_n(g)x_n).$$ Let $e$ be a closed simplicial edge containing $x$. The proof of the theorem is divided into cases, depending on whether the image of $e$ is separating in $T/G$. In both cases, the following lemma will prove useful. \[Defining the mn\] Let $A$ be a vertex group of the splitting over $e$. Let $T_A$ be the minimal $A$-invariant subtree of $T$; conjugating $A$, we can assume that $T_A\cap e$ is precisely one point, $y$. Fix any non-trivial $c\in C={\mathrm{Stab}}(e)$. Then there exists a sequence of integers $m_n$ such that, for any $a\in A$, $$d_n(x_n,f_n(c^{-m_n}ac^{m_n})x_n)\rightarrow d(y,ay)$$ as $n\rightarrow \infty$. The key observation is that $$2d_n(x_n,{\mathrm{Axis}}(f_n(c)))<d_n(x_n,f_n(c)x_n)\rightarrow 0,$$ and the same holds for the $y_n$. Let $x'_n$ be the nearest point on ${\mathrm{Axis}}(f_n(c))$ to $x_n$; likewise, let $y'_n$ be the nearest point on ${\mathrm{Axis}}(f_n(c))$ to $y_n$. Then, for each $n$, there exists $m_n$ such that $$d_n(f_n(c^{m_n})x'_n,y'_n)<l(f_n(c))\rightarrow 0$$ as $n\rightarrow\infty$. Therefore $$d_n(f_n(c^{m_n})x_n,y_n)\rightarrow 0$$ as $n\rightarrow\infty$, and the result follows. The next lemma helps with the case where the image of $e$ is separating. \[Separating automorphism\] Assume the image of $e$ is separating, so the induced splitting is $$G=A*_CB.$$ Assume furthermore that, with the notation of the previous lemma, $x\neq y$. Then there exists $\alpha_n\in{\mathrm{Mod}}(\Delta)$ such that, for all $g\in S$: 1. if $g\in A$ then $\alpha_n(g)=g$; 2. if $g\notin A$ then $$d_n(x_n,f_n\circ\alpha_n(g)x_n)<d_n(x_n,f_n(g)x_n).$$ Fix a non-trivial $c\in C$, and let $\delta_c$ be the Dehn twist in $c$ that is the identity when restricted to $A$. Let $\alpha_n=\delta_c^{m_n}$ where $m_n$ are the integers given by lemma \[Defining the mn\]. Any $g\notin A$ has normal form $$g=a_0b_1a_1\ldots b_la_l$$ with the $a_i\in A\smallsetminus C$ and the $b_i\in B\smallsetminus C$, except for $a_0$ and $a_l$ which may be trivial. Therefore $d(x,gx)=\sum_i d(x,b_ix)+\sum_i d(x,a_ix)$. Fix $\epsilon>0$. If $a_i$ is non-trivial then $a_i\notin C$ and so $$d(x,a_ix)=2d(x,y)+d(y,a_iy).$$ Let $k$ be the number of $a_i$ that are non-trivial (so $l-1\leq k\leq l+1$). Therefore, for all sufficiently large $n$, $$d_n(x_n,f_n(g)x_n)>\sum_i d(x,b_ix) + \sum_i d(y,a_iy)+2kd(x,y)-\epsilon.$$ By contrast, for all sufficiently large $n$, $$d_n(x_n,f_n\circ\alpha_n(g)x_n)<\sum_i d(x,b_ix)+\sum_i d(y,a_iy)+\epsilon$$ by lemma \[Defining the mn\]. By assumption $x\neq y$, so $d(x,y)>0$. Therefore taking $\epsilon<kd(x,y)$ gives the result. We now turn to the non-separating case. \[Defining the pn\] Assume the image of $e$ is non-separating, so the splitting induced by $e$ is $$G=A*_C.$$ Let $t$ be a stable letter. As before, conjugate $A$ so that $T_A\cap e$ is precisely one point $y$. Fix any non-trivial $c\in C={\mathrm{Stab}}(e)$. Then there exists a sequence of integers $p_n$ such that $$d_n(y_n,f_n(tc^{p_n})y_n)\rightarrow 0$$ as $n\rightarrow \infty$. Therefore, for any fixed integer $j$, $$d_n(y_n,f_n(tc^{p_n})^j y_n)\rightarrow 0$$ as $n\rightarrow \infty$. As in the proof of lemma \[Defining the mn\], by the definition of Gromov convergence, $$2d_n(y_n,{\mathrm{Axis}}(f_n(c)))<d_n(y_n,f_n(c)y_n)\rightarrow 0$$ as $n\rightarrow \infty$, and similarly, $$2d_n(f_n(t^{-1})y_n,{\mathrm{Axis}}(f_n(c)))\rightarrow 0$$ as $n\rightarrow \infty$. Let $y'_n$ be the nearest point on ${\mathrm{Axis}}(f_n(c))$ to $y_n$, and let $y''_n$ be the nearest point on ${\mathrm{Axis}}(f_n(c))$ to $f_n(t^{-1})y_n$. Then there exist integers $p_n$ such that $$d_n(f_n(c^{p_n})y'_n,y''_n)\rightarrow 0$$ as $n\rightarrow\infty$. The result now follows. \[Non-separating automorphism\] Assume the situation is in lemma \[Defining the pn\]. Then there exists $\alpha_n\in{\mathrm{Mod}}(\Delta)$ such that, for all $g\in S$: 1. if $g\in C$ then $\alpha_n(g)=g$; 2. if $g\notin C$ then $$d_n(x_n,f_n\circ\alpha_n(g)x_n)<d_n(x_n,f_n(g)x_n).$$ Fix a stable letter $t$ that translates $x$ away from $y$. Fix a non-trivial $c\in C$ and let $i_c\in{\mathrm{Mod}}(G)$ be conjugation by $c$. Set $\alpha_n=i_c^{m_n}\circ\delta_c^{p_n}$, where $m_n$ are integers given by lemma \[Defining the mn\] and $p_n$ are given by lemma \[Defining the pn\]. Any $g$ is of the form $$g=a_0t^{j_1}a_1\ldots t^{j_l}a_l$$ with $j_i\neq 0$ and the $a_i\in A\smallsetminus C$ except for $a_0$ and $a_l$ which may be trivial. Unlike in the case of a separating edge, we have to be a little more careful in estimating $d(x,gx)$ because the natural path from $x$ to $gx$ given by the decomposition of $g$ may backtrack. To be precise, backtracking occurs when $a_i\neq 1$ and $j_{i+1}<0$ and also when $j_i>0$ and $a_{i+1}\neq 1$. Let $k$ be the number of $i$ for which backtracking does not occur, so $0\leq k\leq 2$. Then $$d(x,gx)=\sum_i d(y,a_iy)+\sum_i d(x,t^{j_i}x)+2kd(x,y).$$ Fix $\epsilon>0$. Then for all sufficiently large $n$, $$d_n(x_n,f_n(g)x_n)>\sum_i d(y,a_iy)+2kd(x,y)+\sum_i d(x,t^{j_i}x)-\epsilon.$$ Now for each $i$, $$d_n(f_n(c^{m_n})x_n,f_n((tc^{p_n})^{j_i})f_n(c^{m_n})x_n)\rightarrow 0$$ and $$d_n(f_n(c^{m_n})x_n,f_n(a_ic^{m_n})x_n)\rightarrow d(y,a_iy).$$ So for all sufficiently large $n$, $$d_n(x_n,f_n\circ\alpha_n(g)x_n)<\sum_i d(y,a_iy)+\epsilon.$$ Taking $2\epsilon<2kd(x,y)+\sum_i d(x,t^{j_i}x)$ gives the result. We are now ready to prove the theorem. [theorem \[Simplicial shortening\]]{} Suppose first that $x$ lies in the interior of an edge $e$. If $e$ has separating image in the quotient then lemma \[Separating automorphism\] can be applied both ways round, giving rise to modular automorphisms $\alpha_n$ and $\beta_n$. The theorem is then proved by taking $\phi_n=\alpha_n\circ\beta_n$. If $e$ is non-separating then applying lemma \[Non-separating automorphism\] and taking $\phi_n=\alpha_n$ gives the result. Suppose now that $x$ is a vertex. For each orbit of edges $[e]$ adjoining $x$, let $\alpha_n^e$ be the result of applying lemma \[Separating automorphism\] or lemma \[Non-separating automorphism\] as appropriate to $e$. Now taking $$\phi_n=\alpha_n^{e_1}\circ\ldots\circ\alpha_n^{e_p}$$ where $[e_1],\ldots,[e_p]$ are the orbits adjoining $x$ gives the required automorphism. This is the final piece of the shortening argument. [theorem \[Shortening argument\]]{} Fix a generating set $S$ for $G$. Let $f_i:G\to{\mathbb{F}}$ be a sequence of short homomorphisms corresponding to the convergent sequence of $G$-trees $T_i$. Let $T$ be the limiting $G$-tree and suppose that the action of $G$ on $T$ is faithful. By corollary 3.16 of [@BF03] this induces a GAD $\Delta$ for $G$. Let $x\in T$ be the basepoint fixed in subsection \[Shortening argument subsection\]. Composing the automorphisms given by theorems \[Abelian shortening\] and \[Surface shortening\] there exists $\alpha\in{\mathrm{Mod}}(\Delta)$ such that, for any $g\in G$, $$d(x,\phi(g)x)<d(x,gx)$$ if $[x,gx]$ intersects an abelian or surface component of $T$ and $\phi(g)=g$ otherwise. By theorem \[Simplicial shortening\], for all sufficiently large $i$ there exist $\beta_i\in{\mathrm{Mod}}(\Delta)$ such that $d(x,\beta_i(g)x)=d(x,gx)$ and, furthermore, $$d_i(1,f_i\circ\beta_i(g))<d_i(1,f_i(g))$$ whenever $[x,gx]$ is a non-trivial arc in the simplicial part of the tree. It follows that for $\phi_i=\beta_i\circ\alpha$, $$d_i(1,f_i\circ\phi_i(g))<d_i(1,f_i(g))$$ for all $g\in S$ and all sufficiently large $i$. This contradicts the assumption that the $f_i$ were short. Bestvina and Feighn’s geometric approach ======================================== In section 7 of [@BF03], Bestvina and Feighn provide a more geometric proof of their Main Proposition. In this section we provide proofs of the exercises needed in this argument. The space of laminations ------------------------ Recall that ${\mathcal{ML}}(K)$ is the space of measured laminations on $K$, and ${\mathbb{P}}{\mathcal{ML}}(K)$ is its quotient by the action of ${\mathbb{R}}_+$. Let $E$ be the set of edges of $K$. The space of measured laminations on $K$ can be identified with a closed cone in ${\mathbb{R}}_+^E-\{0\}$, given by the triangle inequality for each 2-cell of $K$. Hence, when ${\mathcal{ML}}(K)$ is endowed with the corresponding topology, ${\mathbb{P}}{\mathcal{ML}}(K)$ is compact. Recall that two laminations are considered equivalent if they assign the same measure to each edge. Therefore it suffices to show existence of a lamination with the prescribed values on the edges. First, for each edge $e$ with $\int_e \mu>0$, fix a closed proper subinterval $I_e$ contained in the interior of $e$. Now fix a Cantor function $c_e:I_e\rightarrow [0,\int_e\mu]$. This gives a measure $\mu$ on $e$, given by $$\int_J\mu=\int_{I_e\cap J} c_e d\lambda$$ where $d\lambda$ is Lebesgue measure on ${\mathbb{R}}$. Now suppose $e_1, e_2,e_3$ are the edges of a simplex in $K$. Divide $e_1$ into intervals $e_1^2$ and $e_1^3$ so that $$2\int_{e_1^2} d\mu=\int_{e_1}d\mu +\int_{e_2}d\mu - \int_{e_3}d\mu$$ and $e_1^2$ shares a vertex with $e_2$, and similarly for $e_1^3$. Divide $e_2$ and $e_3$ likewise. Fix a Cantor set in each $e_i^j$. Now for each distinct $i,j$ inscribe a lamination between $e_i^j$ and $e_j^i$. Since any path transverse to this lamination can be homotoped to an edge path respecting the lamination, the measure on the edges determines a transverse measure to the lamination. Matching resolutions in the limit --------------------------------- A measured lamination on $K$ defines a $G$-tree. The next exercise shows the close relation between the topology on the space of laminations and the topology on the space of trees. For the definition of a resolution, see [@BF03]. The solution is most easily phrased in terms of some explicit construction of the limiting tree. I shall use the asymptotic cone, $T_\omega$; $T$ can be realized as the minimal $G$-invariant subtree of $T_\omega$. For the definition of the asymptotic cone see, for example, [@vdDW]. To see how to choose basepoints and scaling to ensure that the action is non-trivial see, for example, [@P88]. Consider $f_i$-equivariant resolutions $$\phi_i:\tilde{K}\to T_{\mathbb{F}}.$$ Suppose $\lim T_{f_i}=T$, $\lim \Lambda_{\phi_i}=\Lambda$ and the sequences $(|f_i|)$ and $(||\phi_i||_K)$ are comparable. Then there is a resolution that sends lifts of leaves of $\Lambda$ to points of $T$ and is a Cantor function on edges of $\tilde{K}$. A resolution $\phi:\tilde{K}\rightarrow T$ is determined by a choice of $\phi(\tilde{v})$ for a lift $\tilde{v}$ of each vertex $v$ of $K$. First, define a resolution $\phi':\tilde{K}\to T_\omega$ by setting $\phi'(\tilde{v})=[\phi_i(\tilde{v})]$. Since $(|f_i|)$ and $(||\phi_i||_K)$ are comparable, $\phi'(\tilde{v})$ is a valid point of $T_\omega$. The resolution $\phi'$ maps leaves of $\Lambda$ to points, and is a Cantor function on edges. However, $T_\omega$ is far from minimal. Let $\pi:T_\omega\to T$ be closest-point projection to the minimal invariant subtree, which is equivariantly isomorphic to $T$. Now let $\phi=\pi\circ\phi'$’; this is a resolution that still maps leaves of $\Lambda$ to points, and is a Cantor function on edges, as required. Finding kernel elements carried by leaves ----------------------------------------- Exercise 20 of [@BF03] relies heavily on the results of [@BF95]. The most important result is a structure theorem for resolutions of stable actions on real trees, summarized in the following theorem. \[FP actions on trees\] Let $\Lambda$ be a lamination on a 2-complex $K$, resolving a stable action of $G=\pi_1(K)$ on a real tree $T$. Then $$\Lambda=\Lambda_1\sqcup\ldots\sqcup\Lambda_k.$$ Each component has a standard neighbourhood $N_i$ carrying a subgroup $H_i\subset G$. Let $T_i$ be the minimal $H_i$-invariant subtree of $T$. Each component is of one of the following types. 1. **Surface type**. $N_i$ is a cone-type 2-orbifold, with some annuli attached. $H_i$ fits into a short exact sequence $$1\rightarrow \ker T_i\rightarrow H_i\rightarrow\pi_1(O)\rightarrow 1$$ where $O$ is a cone-type 2-orbifold. 2. **Toral type**. $T_i$ is a line, and $H_i$ fits into a short exact sequence $$1\rightarrow \ker T_i\rightarrow H_i\rightarrow A\rightarrow 1$$ where $A\subset{\mathrm{Isom}}({\mathbb{R}})$. 3. **Thin type**. $H_i$ splits over an arc stabilizer, carried by a leaf of $\Lambda_i$. 4. **Simplicial type**. All the leaves of $\Lambda_i$ are compact, and $N_i$ is an interval bundle over a leaf. $H_i$ fits into a short exact sequence satisfying $$1\rightarrow\ker T_i\rightarrow H_i\rightarrow C\rightarrow 1$$ where $C$ is finite. Furthermore, if $E$ is a subgroup carried by a leaf, $E$ fits into a short exact sequence of the form $$1\rightarrow\kappa\rightarrow E\rightarrow C\rightarrow 1$$ where $\kappa$ fixes an arc of $T$ and $C$ is finite or cyclic. In particular, the standard neighbourhoods induce a graph-of-spaces decomposition for $K$, and a corresponding graph-of-groups decomposition for $G$. The vertex spaces are the $N_i$ and the closures of the components of $K-\cup_iN_i$. The edge spaces are boundary components of the $N_i$, and are all contained in a leaf. See theorem 5.13 of [@BF95]. The proof of this exercise will also make use of the following result. \[Stability\] If $h\in H_i$ fixes an arc of $T_i$ then $h\in\ker T_i$. We are now ready to prove the exercise. In the situation of the exercise, the lamination $\Lambda$ has a leaf carrying non-trivial elements of the kernel. Note that $G/\ker T$ is a limit group. Suppose no elements of $\ker T$ are carried by a leaf of $\Lambda$. Consider $\Gamma$ the graph of groups for $G$ induced by $\Lambda$. The aim is to show that $\Gamma$ really is a GAD. Since a GAD decomposition can be used to shorten, this contradicts the assumption that the $f_i$ are short. We deal with each sort of vertex in turn. 1. Suppose $\Lambda_i$ is of surface type. Then $N_i$ is a cone-type 2-orbifold, with some annuli attached. Suppose $g\in H_i$ is carried by an annulus. Then $g$ fixes an arc of $T_i$, so by proposition \[Stability\], $g\in\ker T_i$. But $T_i$ contains a tripod, and tripod stabilizers are trivial, so $g\in\ker T$ contradicting the assumption. Therefore $N_i$ can be assumed to have no attached annuli. Consider an element $g\in H_i$ carried by the leaf corresponding to a cone-point. Then $g$ has finite order, so $g\in\ker T$, since $G/\ker T$ is a limit group. This contradicts the assumption, so $N_i$ has no cone-points. Therefore $N_i$ is genuinely a surface. Moreover, $N_i$ carries a pseudo-Anosov homeomorphism, since it carries a minimal lamination. 2. If $\Lambda_i$ is toral, then $H_i$ is an extension $$1\rightarrow\ker T_i\rightarrow H_i\rightarrow A\rightarrow 1$$ for $A\subset{\mathrm{Isom}}{\mathbb{R}}$. The elements of $\ker T_i$ are carried by annuli in $N_i$. But $\ker T_i$ itself fits into an exact sequence $$1\rightarrow\kappa\rightarrow\ker T_i\rightarrow A'\rightarrow 1$$ where $\kappa\subset\ker T$ and $A'$ is abelian. In order not to contradict the assumption that no elements of the kernel are carried by a leaf, therefore, $\kappa$ must be trivial; so we have $$1\rightarrow A'\rightarrow H_i\rightarrow A\rightarrow 1.$$ and $H_i$ acts faithfully on $T$. In particular, $H_i$ embeds in the limit group $G/\ker T$, and so is a limit group. But $A'$ is normal; since limit groups are torsion-free and CSA, it follows that $H_i$ is free abelian. 3. If $\Lambda_i$ is thin, then $G$ splits over a subgroup $H$ fixing an arc of $T_i$. By proposition \[Stability\], $H\subset \ker{T_i}$. But $T_i$ contains a tripod, and tripod stabilizers are trivial, so $H\subset\ker T$; since $H$ is carried by a leaf, $H$ must be trivial by assumption. But this contradicts the assumption that $G$ is freely indecomposable. 4. If $\Lambda_i$ is simplicial, then $H_i$ fits into the short exact sequence $$1\rightarrow\ker T_i\rightarrow H_i\rightarrow C\rightarrow 1.$$ for finite $C$. As in the toral case, the assumption implies that $\ker T_i$ is abelian, and $H_i$ embeds in $G/\ker T$, and so is a limit group. But, again, $H_i$ is torsion-free and CSA; so $C$ is trivial, and $H_i$ is abelian and fixes an arc of $T$. Now consider an edge-group $E$ of $\Gamma$. Then $E$ is carried by a leaf, and satisfies $$1\rightarrow\kappa\rightarrow E\rightarrow C\rightarrow 1$$ where $C$ is cyclic and $\kappa$ fixes an arc of $T$. Then $\kappa$ fits into a short exact sequence $$1\rightarrow\kappa'\rightarrow\kappa\rightarrow A\rightarrow 1$$ where $\kappa'\subset\ker T$ and $A$ is abelian. By assumption, therefore, $\kappa'$ is trivial and $\kappa$ is abelian; furthermore, $E$ acts faithfully on $T$, so embeds in $G/\ker T$ and is a limit group. Therefore $E$ is free abelian. In conclusion, $\Gamma$ is a GAD. Just as in the proof of theorem \[Shortening argument\], this contradicts the assumption that the $f_i$ are all short. Examples of limit groups ------------------------ To complete their argument, Bestvina and Feighn need some elementary examples of limit groups. This theorem is required. Let $\Delta$ be a 1-edged GAD of a group $G$ with a homomorphism $q$ to a limit group $\Gamma$. Suppose: 1. the vertex groups of $\Delta$ are non-abelian, 2. the edge group of $\Delta$ is maximal abelian in each vertex group, and 3. $q$ is injective on vertex groups of $\Delta$. Then $G$ is a limit group. This theorem is just a special case of proposition \[Heart of the matter\].
{ "pile_set_name": "ArXiv" }
--- abstract: | In this article two methods to distinguish between polynomial and exponential tails are introduced. The methods are mainly based on the properties of the residual coefficient of variation for the exponential and non-exponential distributions. A graphical method, called CV-plot, shows departures from exponentiality in the tails. It is, in fact, the empirical coefficient of variation of the conditional excedance over a threshold. The plot is applied to the daily log-returns of exchange rates of US dollar and Japan yen. New statistics are introduced for testing the exponentiality of tails using multiple thresholds. Some simulation studies present the critical points and compare them with the corresponding asymptotic critical points. Moreover, the powers of new statistics have been compared with the powers of some others statistics for different sample size. author: - | Joan del Castillo\ Universitat Autonoma de Barcelona, Spain - | Jalila Daoudi\ Universitat Autonoma de Barcelona, Spain - | Richard Lockhart\ Simon Fraser University, Canada date: title: Methods to distinguish between polynomial and exponential tails --- #### Keywords: Residual coefficient of variation. Multiple testing problem. Heavy tailed distributions. Power distributions. Extreme value theory. Introduction ============ Since Balkema-DeHaan (1974) and Pickands (1975), it has been well known that the conditional distribution of any random variable over a high threshold — what is known in reliability as the residual life — has approximately a generalized Pareto distribution (GPD). The exponential distribution is a particular case that appears between compact support distributions and heavy-tailed distributions, in GPD. Applications of extreme value theory to risk management in finance and economics are now of increasing importance. The GPD has been used by many authors to model excedances in several fields such as hydrology, insurance, finance and environmental science, see McNeil et al. (2005), Finkenstadt and Rootzén (2003), Coles (2001) and Embrechts et al. (1997). It is especially important for applications to distinguish between polynomial and exponential tails. Often, the methodology is based on graphical methods to determine the threshold where the tail begins, see Embrechts et al. (1997) and Ghosh and Resnick (2010). In this cases, multiple testing problem occurs when one considers a wide set of thresholds. The main objective of this paper is providing ways to distinguish the behavior of tails, avoiding the multiple testing problems. The methods are mainly based on the properties of the residual coefficient of variation that is closely related to the likelihood functions of the exponential and Pareto distributions, see Castillo and Puig (1999) and Castillo and Daoudi (2009). The empirical coefficient of variation, or equivalent statistics (e.g., Greenwood’s statistic, Stephens $W_{s}$) are omnibus tests used for testing exponentiality against arbitrary increasing failure rate or decreasing failure rate alternatives. A good description of these tests has been given by D’Agostino and Stephens (1986). A large number of tests for exponentiality have been proposed in the literature. Montfort and Witter (1985) propose the maximum/median statistic for testing exponentiality against GPD. Smith (1975) and Gel, Miao and Gastwirth (2007) show that powerful tests of normality against heavy-tailed alternatives  are obtained using the average absolute deviation from the median. Lee *et al.* (1980) and Ascher (1990) discuss tests based on the equation $E\left(X^{p}\right)/E\left(X\right)^{p}=\Gamma\left(1+p\right)$, for some $p>0$, where $X$ is an exponential random variable. The limit case, when $p$ tends to $0$, is studied in Mimoto and Zitikis (2008), see also references therein. The case $p=2$ is equivalent to the coefficient of variation test. Lee *et al.* (1980) show that in this case the power is poor testing against distributions whose coefficient of variation is $1$ (the exponential case) as happens testing against the absolute values of the Student distribution $t_{4}$. Our methods based on a multivariate point of view are also useful in this situation, since the exponential distribution is the unique distribution with the residual coefficient of variation over any threshold equal to $1$; see Sullo and Rutherford (1977), Gupta (1987) and Gupta and Kirmani (2000). In Section 2 the asymptotic distribution of the residual coefficient of variation is studied as a random process in terms of the threshold. This provides a clear graphical method, called a CV-plot, for assessing departures from exponentiality in the tails. The qualitative behavior of the CV-plot is made more precise in Section 3. The plot is applied to the daily log-returns of exchange rates of US dollar and Japan yen. New statistics are introduced for testing the exponentiality of tails using multiple thresholds in Section 4. Some simulation studies present the critical points and compare them with the corresponding asymptotic critical points. In Section 5, the powers of new statistics have been compared with the powers of some others statistics against heavy-tailed alternatives, given by Pareto and absolute values of the Student distributions, for different sample size. The residual coefficient of variation ===================================== Let $X$ be a continuous non negative random variable with distribution function $F\left(x\right)$. For any threshold, $t>0$, the distribution function of threshold excedances, $\left(X-t\mid X>t\right)$, denoted $F_{t}\left(x\right)$, is defined by $$1-F_{t}\left(x\right)=\frac{1-F\left(x+t\right)}{1-F\left(t\right)}.$$ The coefficient of variation (CV) of the conditional excedance over a threshold, $t$, (the residual CV) is $$CV\left(t\right)=Var\left(X-t\mid X>t\right)^{1/2}/E\left(X-t\mid X>t\right)$$ where $E\left[\cdot\right]$ and $\ Var\left[\cdot\right]$ denote the expected value and the variance. The $CV\left(t\right)$ is independent of scale parameters. It will be useful find the distribution of the empirical $CV$ process for all values of $t$. It is well known that the mean residual lifetime determines the distribution for random variables. Gupta and Kirmani (2000) showed that mean residual life is a function of the residual coefficient of variation, hence it also characterize the distribution. In this context, generalized Pareto distributions appear as the simple case in which the residual coefficient of variation is a constant. Hence, from Pickands (1975) and Balkema-DeHaan (1974), it is almost constant for a sufficiently high threshold. Denote $X1_{\left(X>t\right)}$ the random variable $X$ if it is larger that $t$ and zero otherwise. Denote $\mu_{0}\left(t\right)=\Pr\left\{ X>t\right\} $ and $\mu_{k}\left(t\right)=E\left[X^{k}1_{\left(X>t\right)}\right]$, $k>0$. Throughout this paper $\mu_{0}\left(t\right)>0,$ foll all $t$, is assumed. Note that$$\mu_{k}\left(t\right)=\mu_{0}\left(t\right)E\left(X^{k}\mid X>t\right).$$ Given a sample $\left\{ X_{j}\right\} $ of size $n$, let $n\left(t\right)=\sum_{j=1}^{n}1_{\left(X_{j}>t\right)}$ the number of excedance over a threshold, $t$. By the law of large numbers, $n\left(t\right)/n$ converges to $\mu_{0}\left(t\right)$. *The empirical CV of the conditional excedance* is given by$$\begin{gathered} cv_{n}\left(t\right)=\frac{n\left(t\right)}{\sum_{j=1}^{n}\left(X_{j}-t\right)1_{\left(X_{j}>t\right)}}\times\\ \left[\frac{\sum_{j=1}^{n}X_{j}^{2}1_{\left(X_{j}>t\right)}}{n\left(t\right)}-\left(\frac{\sum_{j=1}^{n}X_{j}1_{\left(X_{j}>t\right)}}{n\left(t\right)}\right)^{2}\right]^{1/2}\label{cv1}\end{gathered}$$ The $cv_{n}\left(t\right)$ is also independent of scale parameters, since the mean and standard deviation have the same units. The $cv_{n}\left(t\right)$ is a consistent estimator of $CV\left(t\right)$, assuming finite second moment, since the limit in probability of $cv_{n}\left(t\right)$, as $n$ goes to infinity is$$m_{cv}\left(t\right)=\frac{\sqrt{\mu_{2}\left(t\right)\mu_{0}\left(t\right)-\mu_{1}\left(t\right)^{2}}}{\mu_{1}\left(t\right)-t\mu_{0}\left(t\right)}=CV\left(t\right)$$ Fixed $t$, as $n$ goes to infinity$$\frac{1}{n\left(t\right)}\sum_{j=1}^{n}X_{j}^{k}1_{\left(X_{j}>t\right)}=\frac{n}{n\left(t\right)}\frac{1}{n}\sum_{j=1}^{n}X_{j}^{k}1_{\left(X_{j}>t\right)}\rightarrow\mu_{k}\left(t\right)/\mu_{0}\left(t\right)=E\left[X^{k}\mid X>t\right],$$ by the law of large numbers. Hence, the limit in probability of $cv_{n}\left(t\right)$ is$$\frac{\sqrt{\mu_{2}\left(t\right)/\mu_{0}\left(t\right)-\left(\mu_{1}\left(t\right)/\mu_{0}\left(t\right)\right)^{2}}}{\mu_{1}\left(t\right)/\mu_{0}\left(t\right)-t}=\frac{\sqrt{Var\left(X-t\mid X>t\right)}}{E\left(X-t\mid X>t\right)}$$ Let us define the standardized $k$-th sampling moment of the conditional excedance by$$W_{k,n}\left(t\right)=\frac{1}{\sqrt{n}}\sum_{j=1}^{n}\left\{ X_{j}^{k}1_{\left(X_{j}>t\right)}-\mu_{k}\left(t\right)\right\} ,$$ hence,$$\sum_{j=1}^{n}X_{j}^{k}1_{\left(X_{j}>t\right)}=\sqrt{n}W_{k,n}\left(t\right)+n\mu_{k}\left(t\right).\label{as}$$ Note that normalizing constant $1/\sqrt{n}$ is used in order to have $W_{k,n}\left(t\right)=O_{p}\left(1\right)$, with orders of convergence in probability notation. The covariance of this random process is given by $$\begin{aligned} cov\left(W_{i,n}\left(s\right),W_{j,n}\left(t\right)\right) & =cov\left(X^{i}1_{\left(X>s\right)},X^{j}1_{\left(X>t\right)}\right)\notag\\ & =\mu_{i+j}\left(s\vee t\right)-\mu_{i}\left(s\right)\mu_{j}\left(t\right),\label{cov}\end{aligned}$$ Throughout this paper the quantities $cv$ and $W_{k}$ among others depend on $n$;  wherever possible the dependence of quantities on $n$ is suppressed for simplicity. Even the dependence on $t$ is dropped for $W_{k}=W_{k}\left(t\right)$ and $\mu_{k}=\mu_{k}\left(t\right)$, in many places. \[p2\]Let $X$ be a continuous non negative random variable with finite fourth moment. Then, the following expansion holds$$\begin{aligned} \sqrt{n}\left(cv\left(t\right)-m_{cv}\left(t\right)\right)\! &\!\! =\!\! &\! \frac{\mu_{0}W_{2}}{2\left(\mu_{1}-t\mu_{0}\right)\!\sqrt{\mu_{2}\mu_{0}-\mu_{1}^{2}}}+\frac{\mu_{0}\left(t\mu_{1}-\mu_{2}\right)W_{1}}{\left(\mu_{1}-t\mu_{0}\right)^{2}\!\sqrt{\mu_{2}\mu_{0}-\mu_{1}^{2}}}+\!\notag\label{cv2}\\ \\ \!&\! & +\frac{\left(-2t\mu_{1}^{2}+t\mu_{0}\mu_{2}+\mu_{1}\mu_{2}\right)W_{0}}{2\left(\mu_{1}-t\mu_{0}\right)^{2}\sqrt{\mu_{2}\mu_{0}-\mu_{1}^{2}}}+O_{p}\left(\frac{1}{\sqrt{n}}\right).\notag\end{aligned}$$ The expression $\left(\ref{cv1}\right)$ in terms of $W_{k}=W_{k,n}\left(t\right)$ is $$\begin{gathered} cv\left(t\right)=\frac{\mu_{0}\left(t\right)+W_{0}/\sqrt{n}}{\mu_{1}\left(t\right)+W_{1}/\sqrt{n}-t\left(\mu_{0}\left(t\right)+W_{0}/\sqrt{n}\right)}\times\\ \left[\frac{\mu_{2}\left(t\right)+W_{2}/\sqrt{n}}{\mu_{0}\left(t\right)+W_{0}/\sqrt{n}}-\left(\frac{\mu_{1}\left(t\right)+W_{1}/\sqrt{n}}{\mu_{0}\left(t\right)+W_{0}/\sqrt{n}}\right)^{2}\right]^{1/2}\label{cv3}\end{gathered}$$ Let $w_{k}=W_{k}/\sqrt{n}=O_{p}\left(1/\sqrt{n}\right)$, since $W_{k}=O_{p}\left(1\right).$ Then, let us replace $w_{k}$ in $\left(\ref{cv3}\right)$. Taking a Taylor expansion of $\sqrt{n}\left(cv\left(t\right)-m_{cv}\left(t\right)\right)$ with respect to $w_{k}$ near zero the result follows. \[pe\]Let $X$ be a random variable with an exponential distribution with mean $\mu$. Conditional moments of $X$, $\mu_{k}\left(t\right)$, can be obtained from the conditional moments of the exponential distribution of mean $1,\mu_{k}^{1}\left(t\right)$ by$$\mu_{k}\left(t\right)=\mu^{k}\mu_{k}^{1}\left(t/\mu\right)$$ where $$\begin{aligned} \mu_{0}^{1}\left(t\right) & \!\!\!\!=\!\!\!\! & e^{-t},\qquad\mu_{1}^{1}\left(t\right)=e^{-t}\left(1+t\right),\qquad\mu_{2}^{1}\left(t\right)=e^{-t}\left(2+t\left(2+t\right)\right)\\ \mu_{3}^{1}\left(t\right) &\!\!\!\!=\!\!\!\! & e^{-t}\left(6+t\left(6+t\left(3+t\right)\right)\right),\,\mu_{4}^{1}\left(t\right)=e^{-t}\left(24+t\left(24+t\left(12+t\left(4+t\right)\right)\right)\right).\end{aligned}$$ In particular $$m_{cv}\left(t\right)=1.$$ In this Section several results on the convergence of random processes are shown, in the sense of convergence of finite-dimensional distributions. These results are sufficient for the applications given in Section 4. If tightness is proved then weak convergence in the Skorokhod space follows, but this will not be considered here. \[c1\]Let $X$ be a random variable with exponential distribution of mean $\mu$; then $\sqrt{n}\left(cv\left(t\right)-1\right)$ converges to a Gaussian process with zero mean and covariance function given by$$\rho\left(s,t\right)=\exp\left(\frac{s\wedge t}{\mu}\right)\text{.}$$ In particular$$\sqrt{n}\left(cv\left(0\right)-1\right)\overset{d}{\rightarrow}N\left(0,1\right)\text{,}\label{Green}$$ that corresponds to the asymptotic distribution of Greenwood’s statistic. From Theorem $\ref{p2}$ and Example $\ref{pe}$ it follows that$$\sqrt{n}\left(cv\left(t\right)-1\right)=\left(W_{0},W_{1},W_{2}\right)\ a\left(t\right)+O_{p}\left(n^{-1/2}\right)$$ where$$a\left(t\right)^{\prime}=(e^{t/\mu}\,\left(t^{2}+4\, t\,\mu+2\,{\mu}^{2}\right)/\left(2\,{\mu}^{2}\right),-e^{t/\mu}\,\left(t+2\,\mu\right)/{\mu}^{2},e^{t/\mu}/\left(2\,{\mu}^{2}\right))$$ Then, the covariance matrix of $W=\left(W_{0},W_{1},W_{2}\right)^{\prime}$, from $\left(\ref{cov}\right)$ and Example \[pe\], assuming $s\leq t$, is $$\mathrm{cov}\left(W(s),W(t)\right)\equiv M\left(s,t\right)=\left(\mu_{i+j}\left(t\right)-\mu_{i}\left(s\right)\mu_{j}\left(t\right)\right)_{i,j=0,1,2}.$$ Some algebra shows$$a\left(s\right)^{\prime}M\left(s,t\right)\ a\left(t\right)=\exp\left(s/\mu\right).$$ Let $X$ be a random variable with exponential distribution of mean $\mu$; then using a new time scale, $\tau=\mu\log t$, for $t\geq1$, the random process of $\sqrt{n}\left(cv\left(\tau\right)-1\right)$ converges to standard Brownian Motion. From $\left(\ref{c1}\right)$, given $s,t\geq1$, $$\rho\left(\mu\log s,\mu\log t\right)=\exp\left(\log s\wedge\log t\right)=s\wedge t$$ Corollary $\ref{c1}$ uses the same $n$ in $\sqrt{n}\left(cv\left(t\right)-1\right)$ for all $t$. The next result uses the sample size adapted to the corresponding $t$. Let $X$ be a random variable with an exponential distribution, then $\sqrt{n\left(t\right)}\left(cv\left(t\right)-1\right)$ converges to a Gaussian process with zero mean and covariance function given by $$\exp\left(-\left\vert s-t\right\vert /\left(2\mu\right)\right).$$ This is the covariance function of the Ornstein-Uhlenbeck process, the continuous time version of an $AR(1)$ process. It is a stationary Markov Gaussian process. In particular, for any fixed $t$$$\sqrt{n\left(t\right)}\left(cv\left(t\right)-1\right)\overset{d}{\rightarrow}N\left(0,1\right)\text{.}\label{ex}$$ We remember that $n\left(t\right)/n$ converges to $\mu_{0}\left(t\right)=\Pr\left\{ X>t\right\} >0$. Hence, if $n$ tends to infinity $n(t)$ tends to infinity too$.$We can write $$\sqrt{n\left(t\right)}\left(cv\left(t\right)-1\right)=\sqrt{n\left(t\right)/n}\sqrt{n}\left(cv\left(t\right)-1\right).$$ From $\left(\ref{as}\right)$ and Example $\ref{pe}$, we have$$\frac{n\left(t\right)}{n}=\exp\left(-t/\mu\right)+\frac{W_{0}}{\sqrt{n}}.$$ Then $\sqrt{n\left(t\right)}\approx\sqrt{n}\exp\left(-t/2\mu\right)$ and we have that $$\exp\left(-s/2\mu\right)\exp\left(\frac{s\wedge t}{\mu}\right)\exp\left(-t/2\mu\right)=\exp\left(-\left\vert s-t\right\vert /\left(2\mu\right)\right).$$ CV-plot ======= Given a sample $\left\{ x_{k}\right\} $ of positive numbers of size $n$, we denote by $\left\{ x_{\left(k\right)}\right\} $ the ordered sample, so that $x_{\left(1\right)}\leq x_{\left(2\right)}\leq...\leq x_{\left(n\right)}$. We denote by *CV-plot* the representation of the empirical CV of the conditional excedance $\left(\ref{cv1}\right)$, given by$$k\rightarrow cv\left(x_{\left(k\right)}\right)\text{.}\label{plot1}$$ The CV-plot does not depend on scale parameters, since the $cv_{n}\left(t\right)$ does not. That is, the CV-plots for samples $\left\{ x_{k}\right\} $ and $\left\{ \lambda\ x_{k}\right\} $ are the same, for any $\lambda>0$.  In order to have a reference for the behavior of $\left(\ref{plot1}\right)$, pointwise error limits for these plots can be obtained for large samples using $\left(\ref{ex}\right)$, from the null hypothesis of exponentiality. In Section 4, pointwise error limits of the CV-plot are computed by simulation for samples of several sizes. Then, the points are joined by linear interpolation and plotted in the CV-plots. Under regularity conditions, the conditional distribution of any random variable over a high threshold is approximately GPD and this model is characterized as the family of distributions with constant residual CV, as has been said. Hence, the CV-plot can be a complement tool to the Hill-plot or the ME-plot, which are used as diagnostics in the extreme values theory, see Ghosh and Resnick (2010). In order to illustrate the usefulness of the residual coefficient of variation, we are going to examine the behavior of exchange rates between the US dollar and the Japanese yen (JPY), from January 1, 1979 to December 31, 2003. The data set is available from OANDA Corporation at http://www.oanda.com/ convert/fxhistory. The *daily returns* for the dollar price, $P_{k}$, are given by$$x_{k}=\log\left(P_{k}\right)-\log\left(P_{k-1}\right)$$ The daily returns are assumed to be independent here, as in the most basic financial models. However, the theory may be extended even for short-range correlations, see Coles (2001, chap. 5) The set of positive returns is called the *positive part* of returns* *and* * the set of minus the negative returns is called the *negative part*. Both cases are samples of positive random variables. From the 25 years considered we have 9131 daily returns, 3840 of which are positive, 3642 negative and 1649 are equal to zero. In Figure 1, the plots (a) and (b) are the CV-plots of the $n=2000$ largest values for the positive and negative part of dollar/yen returns, respectively. Pointwise 90% limits around the line $cv=1$ are included, the lowest sample size we consider is $20$, since not relevant information comes from smaller samples. Since the basic model for returns is the normal distribution, we will assume that the distribution has support in $(0,\infty)$.  Then, their threshold excedances, for large thresholds, are very nearly Pareto distributed with parameter $\xi>0$ (Pickands, 1975). Some remarks arise from Figure 1. The plot (a) shows that the process $\left(\ref{plot1}\right)$ for the positive part of dollar/yen returns is always inside the pointwise limits for the exponential distribution. Moreover, since we are only interested to test against Pareto alternatives, we have to consider only upper bounds; thus the pointwise level is $95\%$. Hence, the hypothesis that $CV=1$ can be accepted and we can say that the tails decrease at an exponential rate. Note that use of simultaneous confidence limits would make the bounds wider, reinforcing our conclusion. The plot (b) shows that the process $\left(\ref{plot1}\right)$ for the negative part of dollar/yen returns is clearly outside the error limits for the exponential distribution in most of the range. It seems clear that we have to reject the hypothesis of exponentiality. However, the coefficient of variations looks like a constant, approximately. Hence, a Pareto distribution might be accepted for the sample. ![The plots (a) and (b) are the CV-plots of the $n = 2000$ largest values for the positive and negative parts of dollar/yen returns, respectively, with pointwise 90[%]{} error limits under the exponential distribution hypothesis.](Plot.pdf) Testing exponentiality allowing multiple thresholds =================================================== The CV-plot, explained in the last subsection, provides a clear graphical method for assessing departures from exponentiality in the tails. This qualitative behavior shall be made more precise here by introducing new tests of exponential tails adapted to the present situation. The tests are more powerful than most tests against the absolute values of the Student distribution, as we will see in Section \[powerestimates\], including the empirical coefficient of variation, or equivalent statistics as Greenwood’s statistic or Stephens $W_{s}$ (D’Agostino and Stephens, 1986). Our approach is the following: Given a sample $\left\{ x_{j}\right\} $ from an exponential distribution, for any set of thresholds $t_{0}<t_{1}<...<t_{m}$, let $n\left(t_{k}\right)$ be the number of events in $\left\{ x_{j}:x_{j}>t_{k}\right\} $, and $cv\left(t_{k}\right)$ the empirical CV given by $\left(\ref{cv1}\right)$, where $0\leq k\leq m$. From $\left(\ref{ex}\right)$, asymptotically $n\left(t_{k}\right)\left(cv\left(t_{k}\right)-1\right)^{2}$ is distributed as a $\chi_{1}^{2}$ distribution. Let us consider the statistic$$T=\sum_{k=0}^{m}n\left(t_{k}\right)\left(cv\left(t_{k}\right)-1\right)^{2}.\label{cd0}$$ Clearly the asymptotic expectation of $T$ is $m+1$; however, its asymptotic distribution is not $\chi_{m+1}^{2}$, since the random variables $cv\left(t_{k}\right)$ are not independent. Its distribution does not depend on scale parameters and it is straightforward to simulate the distribution of $T$. It is important to note that lower values for $T$ are expected under the null hypothesis of exponentiality, when the expected values for $cv\left(t_{k}\right)$ are $1$. Hence, high values for $T$ show departure from exponential tails. The thresholds $\left\{ t_{k}\right\} $ can be arbitrary but some practical simplicity is obtained by taking thresholds approximately equally spaced, under the null hypothesis of exponentiality. The next result shows a way of doing this. \[p4\]If $X$ is a random variable with exponential distribution of mean $\mu$, then $$\Pr\left\{ X>\left(\mu\log2\right)k\right\} =1/2^{k}$$ Given a sample $\left\{ x_{j}\right\} $ of size $n$ with exponential distribution, the subsample of the last $n/2^{k}$ elements (assuming that $n/2^{k}$ is integer) corresponds to the elements greater than the order statistic $x_{\left(n-n/2^{k}\right)}$ and $x_{\left(0\right)}=0$, $x_{\left(n/2\right)}$, $x_{\left(3\ n/4\right)}$, $x_{\left(7\ n/8\right)}$, ... are approximately equally spaced, from Proposition $\ref{p4}$.   For a general sample, the quantiles $q_{k}$ corresponding to the last $n/2^{k}$ elements are considered $(q_{1}$ is the median, $q_{2}$ is the third quartile, ...$)$. From $\left(\ref{p4}\right)$, $q_{k}\approx\left(\mu\log2\right)k\approx x_{\left(n-n/2^{k}\right)}$. Taking the set of thresholds corresponding to these sampling quantiles, $\left(\ref{cd0}\right)$ became $$T_{m}=n\sum_{k=0}^{m}2^{-k}\left(cv\left(q_{k}\right)-1\right)^{2}\label{cd}$$ Asymptotic distribution ----------------------- It is possible to write $\left(\ref{cd}\right)$ in the form $T_{m}$ $=V^{\prime}V$, where $$V^{\prime}=\sqrt{n}\left[cv\left(q_{0}\right)-1,2^{-1/2}\left(cv\left(q_{1}\right)-1\right),...,2^{-m/2}(cv\left(q_{m}\right)-1)\right]$$ The asymptotic distribution of $T_{m}$ can be found from Corollary $\ref{c1}$ in the following way. From Proposition $\ref{p4},$ we have that $q_{k}\approx\left(\mu\log2\right)k$. Then, asymptotically, the covariance matrix for $V$ is$$\Sigma_{m}=\left(2^{-i/2}\rho\left(q_{i},q_{j}\right)2^{-j/2}\right)_{i,j=0,...,m}=\left(2^{-\left\vert i-j\right\vert /2}\right)_{i,j=0,...,m}$$ \[p3\]The asymptotic distribution of $T_{m}$ is $\sum_{0}^{m}\lambda_{i}Z_{i}^{2}$ with $Z_{i}$ distributed as independent $N(0,1)$ and $\lambda_{i}$ the eigenvalues of $\Sigma_{m}$. From the central limit theorem $V$ is asymptotically multivariate normal $N\left(0,\Sigma_{m}\right)$. Then, in a classical argument, $\Sigma_{m}=A\ \Lambda\ A^{\prime}$ with $A$ an orthogonal matrix and $\Lambda$ the diagonal matrix of the eigenvalues. It follows that $V=A\ \Lambda^{1/2}Z$ with $Z$ asymptotically multivariate normal with the identity as covariance matrix, $N\left(0,I\right)$. Then $T_{m}$ $=V^{\prime}V=Z^{\prime}\Lambda\ Z=\sum_{0}^{m}\lambda_{i}Z_{i}^{2}$, because $A$ is an orthogonal matrix. For instance, for $m=2$,$$\Sigma_{2}=\left(\begin{array}{ccc} 1 & 1/\sqrt{2} & 1/2\\ 1/\sqrt{2} & 1 & 1/\sqrt{2}\\ 1/2 & 1/\sqrt{2} & 1\end{array}\right)$$ and the eigenvalues are given by$$\lambda_{0}=\left(5+\sqrt{17}\right)/4,\lambda_{1}=1/2,\lambda_{2}=\left(5-\sqrt{17}\right)/4$$ Note also that for $m=0$, the asymptotic distribution of $T_{0}$ is simply a $\chi_{1}^{2}$ distribution. Numerical values of the eigenvalues $\lambda_{i}$ are given in Table 1 for other small values of $m$. Approximate critical points --------------------------- Simulation methods are now easily available to compute critical values and p-values of $T_{m}$. However, the asymptotic distribution of $T_{m}$, given by Theorem $\ref{p3}$, provides a way to compute such p-values for large sample sizes without heavy simulation. For instance, if the sample size is $n=2000$ and $m=3$, the direct method needs samples of $2000$ exponential random numbers and the asymptotic distribution only needs samples of $4$ normal random numbers. Moreover, the asymptotic distribution of $T_{m}$, given by Theorem $\ref{p3}$, can be approximated by $a+b\chi_{\nu}^{2}$, where $\chi_{\nu}^{2}$ has gamma distribution with parameters $\left(\nu/2,2\right)$, fitting the constants $a,b,\nu$ in order the three first moments of $\sum_{0}^{m}\lambda_{i}Z_{i}^{2}$ and $a+b\chi_{\nu}^{2}$ be equal. This leads us to solve:$$a+b\ \nu=\sum_{0}^{m}\lambda_{i},\quad b^{2}\nu=\sum_{0}^{m}\lambda_{i}^{2},\quad b^{3}\nu=\sum_{0}^{m}\lambda_{i}^{3}\label{app}$$ Table 1 shows the eigenvalues of the asymptotic covariance matrix of $T_{m}$ and the corresponding constants, $a$, $b$, and $\nu$ for $m=1,2,3$ and $4$. Table 2 shows the critical points, obtained by simulation, for the $T_{m}$ statistics ($m=0,1,2,3$ and $4$) for samples of size $50$, $100$, $200$, $500$, $1000$ and $2000$, corresponding to the $90$, $95$ and $99$ percentiles, as well as the values obtained by simulation of the asymptotic distribution $\left(\ref{p3}\right)$ and the approximation given from $\left(\ref{app}\right)$. The simulations are all run with $50,000$ samples. It can be seen that the asymptotic and approximate methods are useful for samples larger than $500$. These two methods are particularly useful for finding rough p-values. Note that for the approximate method,$$\Pr\left\{ T_{m}>t\right\} =\Pr\left\{ \chi_{\nu}^{2}>(t-a)/b\right\}$$ where $a,b,\nu$ are the solutions of $\left(\ref{app}\right)$. $$\begin{array}{|c|c|c|c|c|c|c|c|c|} \hhline{~--------} \omit & \multicolumn{5}{|c|}{\text{EigenValues}} & \multicolumn{3}{c|}{\text{Parameters}}\\ \hhline{~--------} \omit&\multicolumn{1}{|c}{\lambda 1} & \lambda 2 & \lambda 3 & \lambda 4 & \lambda 5 & a & b & \nu \\ \hhline{---------} T1 & 1.7071 & 0.2929 & - & - & - & 0.2000 & 1.6667 & 1.0800 \\ T2 & 2.2808 & 0.5000 & 0.2192 & - & - & 0.4792 & 2.1818 & 1.1554 \\ T3 & 2.7503 & 0.7420 & 0.3104 & 0.1974 & - & 0.7971 & 2.5758 & 1.2435 {\\}T4 & 3.1381 & 1.0000 & 0.4241 & 0.2500 & 0.1879 & 1.1323 & 2.8764 & 1.3446 {\\}\hhline{---------} \end{array}$$ An example ---------- This analysis is based on the $n=2000$ largest values for the positive and negative parts of dollar/yen returns, respectively, introduced in Section 3. The corresponding CV-plots are (a) and (b) in Figure 1. Looking at the CV plot it can be think that exponentiality is accepted for high order statistics, even in the negative part. In fact when the sample is small enough always the null hypothesis is accepted. But looking at the CV plot hundreds of test are done. Here, the statistic $T_{m}$, for $m=7$, is used; see $\left(\ref{cd}\right)$. The coefficients of variation over tresholds, $cv_{k}$ , for $k=0,...,7$, and samples size $n_{k}=n2^{-k}$ are the following: for the positive part$$\{0.978,0.959,1.008,1.002,1.018,0.919,1.015,0.968\}$$ and for the negative part$$\{1.088,1.135,1.141,1.111,1.088,1.138,1.16,1.585\}\label{cvs}$$ The $T_{m}$ statistics and their corresponding p-values are given by $T_{m}=3.15$ and $p=0.784$, for the positive part; $T_{m}=54.92$ and $p=0.002$, for the negative part. Hence, we accept exponential tails for the positive part and reject this hypothesis for the negative part. Note that in the first case we accept exponentiality for a really large sample, not only the high upper tail of the distribution, and that our test uses simultaneously eight thresholds. The CV-plot in Figure 1(b) suggests a constant coefficient of variation greater than $1$; thus a Pareto distribution can be assumed (Sullo and Rutherford, 1977). In our analysis we conclude that the tails for the positive part of the returns decrease exponentially fast. However, for the negative part we conclude that the tails decrease at a polynomial rate. These conclusions can be surprising, since by considering the yen denominated in dollars the positive and negative part change from one to the other. Note that in these 25 years the price of one dollar went down from 200 yen to 100 yen, more or less. Perhaps this fact and the different sizes of the two economies can explain the difference between positive and negative parts. Probably the traders use different strategies when these two currencies go up or go down. We do not know what the dollar will do in future years. We believe that if it goes down a polynomial rates would be correct to measure risks. Comparisons with other inference approaches ------------------------------------------- The CV-plot (b) in Figure 1 suggest to model the negative part of dollar/yen returns by a Pareto distribution. The *generalized Pareto family of distributions (GPD)* has probability distribution function, for $\beta>0$,$$F\left(x\right)=1-\left(1+\xi x/\beta\right)^{-1/\xi},\label{pareto}$$ defined on $x>0$ for $\xi>0$ and defined on $0<x<\beta/\left\vert \xi\right\vert $ for $\xi<0$. The limit case $\xi=0$ corresponds to the exponential distribution. When $\xi>0$, the GPD is simply the *Pareto distribution*. In this case the tail function decrease like a power law and the inverse of the shape parameter, $\xi^{-1}$, is called the *power of the tail*. Hence we can estimate the parameters of $\left(\ref{pareto}\right)$ by maximum likelihood (ML), using the sample of size $n=2000$ in the last Example. We find $\hat{\xi}^{-1}=13.473$ and $\hat{\beta}=$ $0.024\ $ and, the corresponding coefficient of variation is $\hat{c}_{\xi}=1.084$. Note that this result is not far from $c_{0}=1.088$ in $\left(\ref{cvs}\right)$. In the same way, estimating the Pareto parameters by ML, from samples of size $n_{k}$, we find coefficients of variation near $c_{k}$ in $\left(\ref{cvs}\right)$. The clasical approach from extreme values theory uses the *generalized extreme value distribution*. This distribution is defined by the cumulative distribution function$$G\left(x\right)=\exp\left[-\left(1+\xi\left(\frac{x-\mu}{\sigma}\right)\right)_{+}^{-1/\xi}\right].\label{gev}$$ For $\xi>0$ the model $\left(\ref{gev}\right)$ is the Frechet distribution, for $\xi=0$ the Gumbel distribution and for $\xi<0$ the Weibull distribution, see Embrechts et al. (1997). Using $\left(\ref{gev}\right)$, with the anual maximums gives the ML estimation $$\left(\hat{\mu},\hat{\sigma},\hat{\xi}^{-1}\right)=\left(0.023,0.005,5.485\right)$$ and leads to $\hat{c}_{\xi}=1.255$. The standard error for $\hat{\xi}^{-1}$ has been computed with the inverse of the observed information matrix, and gives $sd($ $\hat{\xi}^{-1})=5.326$. Hence, the $95\%$ confidence interval for $\xi^{-1}$ includes the estimation above. However, the range for $\xi^{-1}$ is really wide, including distributions with no finite mean and distributions with compact support. We conclude that the estimation done with Pareto distribution seems correct and it agrees with the hypotesis of a coefficient of variation over thresholds constant. However, the tail estimated with generalized extreme value distribution looks away of the coefficients of variation over threshold in $\left(\ref{cvs}\right)$. Power estimates {#powerestimates} =============== The $T_{m}$ statistics test simultaneously at several points whether $CV=1$, though at each new point only one half of the sample of the previous point is used. Hence, $T_{m}$ statistics are especially useful for testing exponentiality in the tails, when the exact point where the tail begins is unknown, avoiding the problem of multiple comparisons. However, in this Section $T_{m}$ is considered as a simple test of exponentiality. Two experiments are conducted. The first one considers as the alternative distribution the absolute value of the Student distribution (with degrees of freedom $\nu=1$ to $10)$. In the second case the alternative distribution is a Pareto distribution. In both cases the empirical powers of the $T_{m}$ statistics ($m=0,1,2,3$ and $4$) have been compared with the empirical powers of the empirical coefficient of variation (D’Agostino and Stephens, 1986) and the tests suggested by Montfort and Witter (1985) and Smith (1975) as tests against heavy-tailed alternatives. Every empirical power is estimated running $10,000$ samples and using the critical points of Tables 2 and 3. All the statistics considered are invariant to changes in scale parameters. Hence, the powers estimated do not depend on scale parameters under the null hypothesis of exponentiality or under the alternative distributions. Montfort and Witter (1985) propose the maximum/median statistic for testing exponentiality against the GPD. Given a sample $\left\{ X_{i}\right\} $, let us denote $$MW=Max\left(X_{i}\right)/X_{m}\label{HW}$$ where $X_{m}$ is the median of the sample. Smith (1975) and Gel, Miao and Gastwirth (2007) show that powerful tests of normality against heavy-tailed alternatives are obtained using the average absolute deviation from the median. The same statistic suggested by Smith (1975) is used here for testing exponentiality against heavy-tailed alternatives. Let us denote$$SU=\left[\sum_{i}\left(X_{i}-\bar{X}\right)^{2}/n\right]^{1/2}/\left[\sum_{i}\left\vert X_{i}-X_{m}\right\vert /n\right]\label{SU}$$ where $\bar{X}$ is the sample mean. The empirical coefficient of variation statistics is (D’Agostino and Stephens, 1986)$$cv=\left[\sum_{i}\left(X_{i}-\bar{X}\right)^{2}/n\right]^{1/2}/\bar{X}.$$ Table 3 shows the critical points for the empirical coefficient of variation and the statistics $MW$ and $SU$, for samples of size equal to $50$, $100$, $200$, $500$, $1000$ and $2000$, corresponding to several quantiles. The simulations are all run with $50,000$ samples. Note that here two-sided test are considered. This one is the unique difference between $cv$ and $T_{0}$. The cumulative distribution function of the Pareto distribution is $$F\left(x\right)=1-\left(1+\xi x/\psi\right)^{-1/\xi},\label{GPD}$$ where $\psi>0$ and $\xi>0$ are scale and shape parameters and $x>0$. The limit case $\xi=0$ corresponds to the exponential distribution. The parameter $\alpha=1/\xi$ is called the power of the tail. The probability density function of the Student distribution with $\nu$ degrees of freedom is$$t_{\nu}\left(x\right)=\frac{\Gamma\left(\left(\nu+1\right)/2\right)}{\sqrt{\nu\pi}\ \Gamma\left(\nu/2\right)}\left(1+\frac{x^{2}}{\nu}\right)^{-\left(\nu+1\right)/2}$$ Hence, a Student distribution is a distribution of regular variation with index $\alpha=\nu$. That is, the tails of the Student distribution are like the Pareto distribution for $\xi=1/\nu$. When $\nu$ tends to infinity the Student distribution tends to the standard normal distribution, hence it is a usual alternative when the tails are heavier than in the normal case. For $\nu=1$ the distribution is also called the Cauchy distribution. In order to test exponentiality only the positive part, or equivalently the absolute value, of the Student distribution is considered. Note that in finance often models with only three finite moments (infinite kurtosis) are considered; that corresponds to a Student distribution with $\nu=3$ or $\nu=4$. Table 4 reports the results for the eight statistics with sample sizes, $n$, of $50,100,200,500,1000$ and $2000$, at significance level $5\%$, testing exponentiality against the absolute value of the Student distribution with degrees of freedom from $\nu=1$ to $10$. Several overall observations can be made on the basis of these sampling experiments. First of all, the powers are high for $\nu=1$ (Cauchy distribution) or $\nu=2$ (unbounded variance) and clearly increase with sample size for $\nu\geq7$. In most cases $cv$ (or $T_{0}$) is superior to the other tests. However, its power is poor against some particular cases. Even for samples of size $2000$ the power is only $38\%$ against the absolute values of the Student distribution $t_{4}$. This is easily explained since the alternative has coefficient of variation $CV=1$, as in the null hypothesis of exponentiality. In this case the powers of $T_{1}$, $T_{2}$ and $T_{3}$ are $96\%$, $98\%$ and $97\%.$ In general the power of $cv$ is something higher than $T_{1}$ or $T_{2}$ but in some cases very much lower. Table 5 reports the results of the eight statistics with sample sizes, $n$, of $50,100,200,500$ and $1000$, at significance level $5\%$, testing exponentiality against a Pareto distribution with scale parameter $\psi=1$ and shape parameters $\xi$ from $0.05$ to $0.5$ with increments of $0.05$. The Pareto distribution has constant coefficient of variation, hence the $T_{m}$ statistics do not have any advantage testing for $CV=1$ at different points. Moreover, at each new point only one half of the sample of the previous point is used. The overall observation that can be made on the basis of these sampling experiments is that again $cv$ (or $T_{0}$) is superior to other tests; this agrees with the results Castillo and Daoudi (2009). Moreover, other $T_{m}$ statistics are not far away from $cv$. The main conclusion is that, though $cv$ is in general a good test, the $T_{m}$ statistics have a very similar power and clearly improve the poor power of $cv$ in testing against distributions with coefficient of variation near $1$, which often appear in finance. Bibliography ============ 1. Ascher, S. (1990). A survey of tests for exponentiality. *Communications in Statistics: Theory and Methods*, 19, 1811–1825. 2. Balkema, A. and DeHaan, L. (1974). Residual life time at great age. *Annals of Probability*. 2, 792-804. 3. Coles, S. (2001). *An Introduction to Statistical Modelling of Extreme Values*. Springer, London. 4. Castillo, J. and Daoudi, J. (2009). Estimation of the generalized Pareto distribution. *Statistics and Probability Letters*. 79, 684-688. 5. Castillo, J. and Puig, P. (1999). The Best Test of Exponentiality Against Singly Truncated Normal Alternatives. *Journal of the American Statistical Association*. 94, 529-532. 6. D’Agostino, R. and Stephens, M.A. (1986). *Goodness-of-Fit Techniques*. Marcel Dekker, New York. 7. Embrechts, P. Klüppelberg, C. and Mikosch, T. (1997). *Modeling Extremal Events for Insurance and Finance*. Springer, Berlin. 8. Finkenstadt, B.and Rootzén, H. (edit) (2003). *Extreme values in Finance, Telecommunications, and the Environment*. Chapman & Hall. 9. Lee, S., Locke, C. and Spurrier, J. (1980). On a class of tests of exponentiality. *Technometrics*, 22, 547–554. 10. McNeil, A. Frey, R. and Embrechts, P. (2005). *Quantitative Risk Management: Concepts, Techniques and Tools*. Princeton UP. New Jersey. 11. Mimoto, N. and Zitikis, R. (2008). The Atkinson index, the Moran statistic and testing exponentiality. *J. Japan Statist. Soc*. 38, 187–205. 12. Montfort, M. and Witter, J. (1985). Testing exponentiality against generalizad Pareto distribution. *Journal of Hydrology*, 78, 305-315. 13. Gel, Y., Miao, M. and Gastwirth, J. (2007). Robust directed tests of normality against heavy-tailed alternatives. *Computational Statistics & Data Analysis* 51, 2734-2746. 14. Ghosh, S. and Resnick, S. (2010). A discussion on mean excess plots. * Stochastic Processes and their Applications* 120, 1492-1517. 15. Gupta, R. (1987). On the monotonic properties of the residual variance and their applications in reliability. *Journal of Statistical Planning and Inference* 16, 329-335. 16. Gupta, R. and Kirmani, S. (2000). Residual coefficient of variation and some characterization results. *Journal of Statistical Planning and Inference* 91, 23-31. 17. Pickands, J. (1975). Statistical inference using extreme order statistics. *The Annals of Statistics* 3, 119-131. 18. Smith, V. (1975). A Simulation analysis of the Power of Several Test for Detecting Heavy-Tailed Distributions. *Journal of the American Statistical Association*, 70, 662-665. 19. Sullo, P. and Rutherford, D. (1977). Characterizations of the Power Distribution by Conditional Exceedance, in American Statistical Association, Proceedings of the Business and Economic Statistics Section, Washington, D.C.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Direct reaction experiments provide a powerful tool to probe the structure of neutron-rich nuclei like Beryllium-11. We use halo effective field theory to calculate the cross section of the deuteron-induced neutron transfer reaction [${{}^{10}\text{Be}}(\text{d},\,\text{p}){{}^{11}\text{Be}}$]{}. The effective theory contains dynamical fields for the Beryllium-10 core, the neutron, and the proton. In contrast, the deuteron and the Beryllium-11 halo nucleus are generated dynamically from contact interactions using experimental and ab-initio input. The reaction amplitude is constructed up to next-to-leading order in an expansion in the ratio of the length scales characterizing the core and the halo. The Coulomb repulsion between core and proton is treated perturbatively. Finally, we compare our results to cross section data and other calculations.' author: - 'M. Schmidt' - 'L. Platter' - 'H.-W. Hammer' bibliography: - 'HaloReactions.bib' title: Neutron transfer reactions in halo effective field theory --- Introduction {#Sec:Intro} ============ Nuclear processes such as capture and transfer reactions are one focus of ongoing research at existing and forthcoming experimental facilities with radioactive ion beams [@Fahlander:2013nfa]. However, the consistent theoretical description of such reactions in ab initio calculations poses significant challenges. Tremendous process has been made for lighter systems in calculating elastic nucleus-nucleon scattering processes by combining the variational approach of the resonating group model and the no-core shell model in the no-core shell model with continuum [@Navratil:2016ycn]. However, for larger systems it remains a challenging task to calculate reactions in a controlled way and with reliable uncertainty estimates; see for example Refs. [@Yoshida:2017pxa; @Capel:2018kss; @King:2018vzw; @Nunes:2018erz; @Lovell:2018bao]. One alternative approach is to reduce the number of dynamical degrees of freedom. A process can then be described as an effective two- or three-body problem using a Lippmann-Schwinger or Faddeev equation. The remaining challenge is to model the interaction between the degrees of freedom appropriately. A reduction to the minimal degrees of freedom required to obtain a certain observable is frequently the starting point of an effective field theory (EFT) treatment of a system. EFTs can be applied if a system displays two disparate scales that can be combined to form a small expansion parameter. The large scale can for example be the excitation energy of a degree of freedom or a heavy state not included in the approach. EFT is the theory in which these high energy modes are integrated out. Halo nuclei display such a separation of scales [@Zhukov:1993aw; @Hansen:1995pu; @Jonson:2004; @Jensen:2004zz]. They consist of a tightly bound core with large excitation energy ${E_\text{x}}$ and some weakly bound valence nucleons. The EFT that has been developed for these systems is called halo effective field theory (Halo EFT) [@Bertulani:2002sz; @Bedaque:2003wa]. It treats the core as a fundamental degree of freedom, which is a valid approximation as long as energies smaller than ${E_\text{x}}$ are considered. Halo EFT has been applied to a variety of processes including electromagnetic transitions and Coulomb dissociation of one-neutron halo nuclei. The formalism has been extended to one-proton and two-neutron halo nuclei. For a recent review, see Ref. [@Hammer:2017tjm]. In this work, we explore the potential of Halo EFT to describe the experimentally important process of a deuteron-induced transfer reaction. Such a calculation has not been carried out yet due to the challenging continuum structure of the reaction. As a test case, we consider [${{}^{10}\text{Be}}(\text{d},\,\text{p}){{}^{11}\text{Be}}$]{}. The effective three-body system is given by a ${{}^{10}\text{Be}}$ core, a neutron, and a proton. The one-neutron halo nucleus ${{}^{11}\text{Be}}$ represents a neutron-core state with a binding energy much smaller than the $2^+$ core excitation energy ; see Fig. \[Fig:Levels\]. This intrinsic scale separation reflects itself also in the small core radius and the large halo radius [@Nortershauser:2008vp]. Exploiting these length scales, we calculate the reaction cross section up to first order in ${R_\text{c}}/{R_\text{h}}$. We find that dynamical core excitations, excited ${{}^{11}\text{Be}}$ states and proton-core resonances in Fig. \[Fig:Levels\] can be neglected at leading order (LO). ![\[Fig:Levels\]Thresholds relative to ${{}^{10}\text{Be}}$$+$n$+$p. The center column shows the ground and first excited state of ${{}^{10}\text{Be}}$. Bound and resonance states of the core-neutron (${{}^{11}\text{Be}}$) and core-proton (${{}^{11}\text{B}}$) systems are depicted in the left and right columns, respectively. We only show ${{}^{11}\text{B}}$ levels, which have been seen in the experiment of Ref. [@Goosman:1970vn]. In this work, we explicitly include those states with thick lines.](Pictures/picture1.pdf) We expect that the expansion works best for center-of-mass energies $E$ well below ; see Fig. \[Fig:Levels\]. However, in the absence of approriate data, we compare our theory to data at , measured by Schmitt [*et al.*]{} at Oak Ridge National Laboratory [@Schmitt:2012bt]. In fact, previous works suggest that Halo EFT could still be appropriate for the lower experimental energies. For example, Deltuva [*et al.*]{} calculated the differential cross section in a Faddeev approach, using model interactions that reproduce elastic proton-core scattering data and optical potentials that account for loss channels [@Deltuva:2016his]. Their work suggests that core excitations barely influence the cross section for . More recently, Yang and Capel [@Yang:2018nzr] reanalyzed the reaction by combining the adiabatic distorted wave approximation (ADWA) reaction model with a Halo EFT description of ${{}^{11}\text{Be}}$. They found out that, for the lower beam energies and forward angles, the reaction is purely peripheral. I.e., it only depends on the asymptotic form of the ${{}^{11}\text{Be}}$ wave function, while being independent of short-range details. Indeed, we will be able to describe data for the lower beam energies. This manuscript is structured as follows. In section \[Sec:Lagrangian\], we present the EFT Lagrangian. Strong interactions between core, neutron, and proton are described by contact forces and the Coulomb interaction follows from photon couplings. Section \[Sec:2Body\] explains how the two-body states ${{}^{11}\text{Be}}$, ${{}^{11}\text{Be}}^\ast$ and the deuteron dynamically emerge from the given interactions. We then turn to the three-body system in section \[Sec:3Body\]. A Faddeev equation for the reaction will be constructed up to next-to-leading order (NLO) in the ${R_\text{c}}/{R_\text{h}}$ expansion. Following work carried out for the three-nucleon sector [@Rupak:2001ci; @Konig:2014ufa], it will include the dominant Coulomb contributions. After presenting results for the reaction cross section, we summarize our work and give an outlook in section \[Sec:SummaryOutlook\]. EFT Lagrangian {#Sec:Lagrangian} ============== The EFT Lagrangian $\mathcal{L}$ can be written as the sum $$\label{Eq:Lagrangian-sum} \mathcal{L}=\mathcal{L}_1+\mathcal{L}_2+\mathcal{L}_3+\mathcal{L}_\gamma$$ of one-, two-, and three-body interactions and a photon part. The kinetic part reads $$\label{Eq:Lagrangian-1b} \mathcal{L}_1= n_\alpha^\dagger \left( iD_0+\frac{{\boldsymbol{D}}^2}{2{m_\text{N}}} \right)n_\alpha +p_\alpha^\dagger \left( iD_0+\frac{{\boldsymbol{D}}^2}{2{m_\text{N}}} \right)p_\alpha +c^\dagger \left( iD_0+\frac{{\boldsymbol{D}}^2}{2{m_\text{c}}} \right)c\,.$$ It introduces fields $n_\alpha$, $p_\alpha$ and $c$ for neutron, proton, and the ${{}^{10}\text{Be}}$ core. They are treated as distinguishable particles. Sums over doubly appearing indices are implicit. Masses are taken to be and . The photon’s kinetic and gauge fixing terms are given by $$\mathcal{L}_\gamma=-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}-\frac{1}{2\xi} \left( \partial_\mu A^\mu -\eta_\mu \eta_\nu \partial^\nu A^\mu \right)^2$$ with time-like unit vector $\eta_\mu$. We only consider Coulomb photons, which induce a static potential. The covariant derivative in Eq.  with charge operator $\hat{Q}$ induces respective photon couplings $+ie\,Q_\text{p/c}$ with and . As done in Ref. [@Konig:2015aka], we introduce a screened Coulomb photon propagator $$iG_\gamma({\boldsymbol{p}})\equiv i\left[ {\boldsymbol{p}}^2+\lambda^2-i\epsilon \right]^{-1}.$$ The artificial photon mass $\lambda$ has to be taken to zero at the end of each calculation. The two-body part $\mathcal{L}_2$ involves the auxiliary fields $\sigma_\alpha$ and $d_i$ for the shallow bound states ${{}^{11}\text{Be}}$ and deuteron, respectively. It reads $$\begin{aligned} \nonumber \mathcal{L}_2 =&\ \sigma_\alpha^\dagger \left[ \Delta_\sigma^{(0)} -\left( i\partial_0+\frac{{\boldsymbol{\nabla}}^2}{2M_\text{Nc}} \right) \right]\sigma_\alpha -{g_\sigma}\left[ \sigma_\alpha^\dagger \left(n_\alpha c\right) + \text{H.c.} \right] \\ \nonumber \label{Eq:Lagrangian-2b} &+d_i^\dagger \left[ \Delta_\d^{(0)} -\left( i\partial_0+\frac{{\boldsymbol{\nabla}}^2}{4{m_\text{N}}} \right) \right]d_i -{g_\text{d}}\,{\text{C}_{1/2\,\alpha,1/2\,\beta}^{1i'}} \left[ d_{i'}^\dagger (p_\alpha n_\beta) + \text{H.c.} \right] \\ &+\mathcal{L}_{2,{{}^{11}\text{Be}}^\ast}+\cdots\end{aligned}$$ with and a Clebsch-Gordan coefficient ${\text{C}_{s_1m_1,s_2m_2}^{sm_s}}$. The expression "H.c.“ denotes the Hermitian conjugate. The regularization-dependent parameters will be matched to experiment. Derivatives in Eq.  induce range corrections at NLO. Higher-order terms in the ellipses are negligible at NLO. The part $\mathcal{L}_{2,{{}^{11}\text{Be}}^\ast}$ accounts for NLO contributions from the first excited state ${{}^{11}\text{Be}}^\ast$. It is discussed in appendix \[App:Be11Exc\]. The three-body part $\mathcal{L}_3$ contains an $s$-wave deuteron-core interaction $C_0$ which will be used to renormalize the reaction amplitude. We write $$\label{Eq:Lagrangian-3b} \mathcal{L}_3 =-{g_\text{d}}^2\,C_0\,(d_i c)^\dagger (d_i c) +\cdots\,.$$ Two-body states {#Sec:2Body} =============== Due to the small neutron separation energies of deuteron and ${{}^{11}\text{Be}}$, three-body processes like deuteron breakup are crucial for the transfer reaction; see for example Refs. [@Yilmaz:2001pi; @Gomez-Ramos:2017ihe]. In this section, we summarize how ${{}^{11}\text{Be}}$, the deuteron, and ${{}^{11}\text{Be}}^\ast$ emerge dynamically from contact interactions of the EFT Lagrangian. Moreover, we explain the effective treatment of core excitation effects in the ${{}^{11}\text{Be}}$ system. The Beryllium-11 ground state ----------------------------- In Halo EFT, the ${{}^{11}\text{Be}}$ ground state ($1/2^+$) is treated as a pure neutron-core $s$-wave state. Already at LO, its propagator $iG_\sigma$, depicted as a solid-dashed double line, contains iterations of the neutron-core self-energy loop to all orders; see Fig. \[Fig:Dyson\_sigma\]. As a consequence of the EFT’s Galilean invariance, $iG_\sigma$ is a function of the center-of-mass energy only, where $p^\mu$ denotes the total four-momentum. After resumming the self-energy, $iG_\sigma$ takes the well-known effective range expansion form [@Bethe:1949yr] $$\label{Eq:EffRangeExp} iG_\sigma({E_\text{cm}})= -i\,{g_\sigma}^{-2}\,\frac{2\pi}{{\mu_\text{Nc}}} \left[ -a_\sigma^{-1} +\frac{r_\sigma}{2}k^2+\cdots -ik \right]^{-1},$$ where is the on-shell relative momentum and the ellipses denote higher order terms.[^1] In the Power Divergence Subtraction (PDS) scheme with mass scale $\Lambda_\text{PDS}$ [@Kaplan:1998tg; @Kaplan:1998we], the scattering length $a_\sigma$ and effective range $r_\sigma$ are given by $$\begin{aligned} \label{Eq:Matching_sigma} a_\sigma^{-1}&=\frac{2\pi}{{\mu_\text{Nc}}}\frac{\Delta_\sigma^{(0)}(\Lambda_\text{PDS})}{{g_\sigma}^2}+\Lambda_\text{PDS}\,, \\ \label{Eq:Matching_sigma2} r_\sigma &= \frac{2\pi}{{\mu_\text{Nc}}^2}{g_\sigma}^{-2}\,.\end{aligned}$$ ![\[Fig:Dyson\_sigma\]The full ${{}^{11}\text{Be}}$ propagator $iG_\sigma$ (solid-dashed double line with filled circle) is given by the bare one (with empty circle) and the neutron-core self-energy loop, where the solid (dashed) line represents the neutron (core). A similar expression for the deuteron propagator can be obtained by replacing all dashed core lines by dotted proton lines.](Pictures/picture2.pdf) The propagator has a pole at , or equivalently at , where [@TUNL_Be11] and are the small binding energy and binding momentum. Thus, Eq.  can be rearranged by writing $$\label{Eq:EffRangeExpPole} iG_\sigma({E_\text{cm}})= i\,{g_\sigma}^{-2}\,\frac{2\pi}{{\mu_\text{Nc}}} \left[ \gamma_\sigma+ik -\frac{r_\sigma}{2}\left( k^2+\gamma_\sigma^2 \right) +\cdots \right]^{-1}.$$ Note that the coupling ${g_\sigma}$ is not an observable. We thus eliminate it using the redefined auxiliary fields ; see for example Ref. [@Griesshammer:2004pe]. Consequently, we multiply $G_\sigma$ by ${g_\sigma}^2$ and each (neutron-core)-${{}^{11}\text{Be}}$ vertex by ${g_\sigma}^{-1}$. ### Halo EFT counting & ANC In Halo EFT, all parameters in Eqs. – scale with certain powers of the large halo radius and the small core radius ${R_\text{c}}$, which represents the natural nuclear physics length scale [@Hammer:2011ye]. We may estimate from the core excitation energy . Thus, the EFT expansion parameter is . As one of the first applications of Halo EFT to electromagnetic processes, Hammer and Phillips used data of the low-energy E1 strength of ${{}^{11}\text{Be}}$ breakup, to determine a value for $r_\sigma$ [@Hammer:2011ye]. Their result $\unit[2.7]{fm}$ scales like ${R_\text{c}}$. In contrast, the binding momentum is as small as ${R_\text{h}}^{-1}\approx\unit[28]{MeV}$. It follows that for low momenta , the effective range term in Eq.  is of NLO compared to . Higher-order terms in the ellipses are of the order ${R_\text{c}}^3/{R_\text{h}}^4$ (N${}^3$LO) at most [@Hammer:2011ye]. Once physics in the pole region is reproduced at a desired accuracy, it becomes obsolete to scale the ${{}^{11}\text{Be}}$ ground state wave function with a spectroscopic factor. Such experimentally nonobservable quantities are not required in Halo EFT. Instead, Eq.  yields an asymptotic normalization coefficient (ANC) $$\label{Eq:ANC} A_\sigma = \sqrt{ \frac{ 2\gamma_\sigma}{ 1-\gamma_\sigma r_\sigma +{\mathcal{O}\left({R_\text{c}}^3/{R_\text{h}}^3\right)} } }$$ for the radial wave function , which is fully determined by low-energy observables [@Hammer:2011ye]. Recently, Calci [*et al.*]{} were able to calculate the ANC using the no-core shell model with continuum (NCSMC) [@Calci:2016dfb]. Their result was afterwards confirmed by Yang and Capel in Ref. [@Yang:2018nzr], who extracted the value from cross section data [@Schmitt:2012bt]. We will use the ANC of Calci [*et al.*]{} as an input parameter at NLO. Equation  can then be inverted to give a value for the effective range, which reads $$\label{Eq:rSigmaFromANC} r_\sigma \equiv \left( \gamma_\sigma^{-1}-\frac{2}{A_\sigma^2} \right)\left(1+{\mathcal{O}\left({R_\text{c}}^2/{R_\text{h}}^2\right)}\right) \approx \unit[3.5]{fm}\,.$$ This value is larger than the one obtained by Hammer and Phillips in Ref. [@Hammer:2011ye]. It will still be counted as ${R_\text{c}}$, since differs by only from . ### Propagator expansion From NLO, the propagator in Eq.  exhibits spurious deep poles in addition to the physical one representing ${{}^{11}\text{Be}}$ [@Ji:2011qg]. We solve this issue by expanding $iG_\sigma$ around in terms of ${R_\text{c}}/{R_\text{h}}$, yielding the series $$\begin{aligned} \nonumber iG_\sigma({E_\text{cm}}) =&\ i\frac{2\pi}{{\mu_\text{Nc}}} \left[ \gamma_\sigma-\sqrt{-2{\mu_\text{Nc}}(E_\text{cm}+i\epsilon)} \right]^{-1} \\ \label{Eq:PoleExp_sigma} &\times\left(1+ \frac{r_\sigma}{2} \left( \gamma_\sigma+\sqrt{-2{\mu_\text{Nc}}(E_\text{cm}+i\epsilon)} \right) +{\mathcal{O}\left({R_\text{c}}^2/{R_\text{h}}^2\right)} \right).\end{aligned}$$ The residue of $G_\sigma$ has an analogue expansion and reads $$\label{Eq:Residue_sigma} Z_\sigma=\Bigg[ \left. \frac{\partial G_\sigma^{-1}}{\partial E_\text{cm}} \right|_{E_\text{cm}=-B_\sigma} \Bigg]^{-1} =\frac{2\pi}{\mu_\text{Nc}^2}\,\gamma_\sigma \left( 1+\gamma_\sigma r_\sigma +{\mathcal{O}\left({R_\text{c}}^2/{R_\text{h}}^2\right)} \right).$$ In section \[Sec:3Body\], $G_\sigma$ will enter the three-body Faddeev equation and $Z_\sigma$ is needed to normalize the reaction amplitude. At LO, we will truncate Eqs. – after the leading term "$1$“, yielding expressions $G_\sigma^{(\text{LO})}$ and $Z_\sigma^{(\text{LO})}$. The NLO forms $G_\sigma^{(\text{NLO})}$ and $Z_\sigma^{(\text{NLO})}$ also include the terms linear in $r_\sigma$. We will follow Bedaque [*et al.*]{} by replacing in the Faddeev kernel at NLO [@Bedaque:2002yg]. This straightforward technique is often referred to as "partial resummation“, because it induces specific amplitude terms proportional to $r_\sigma^n$, $n\geq 2$. In principle, such terms only occur at higher orders. However, for natural cutoffs, they are smaller then NLO terms and do not undermine the validity of the NLO calculation [@Ji:2011qg; @Ji:2012nj]. ### Core excitation effects So far, we have treated ${{}^{11}\text{Be}}$ as a pure neutron-core state. However, in principle, it also couples to the configuration of a neutron and a core excitation ${{}^{10}\text{Be}}^\ast$. Note that this threshold resides far above the pole at an energy separation ; see Fig. \[Fig:Levels\]. Close to the pole, $G_\sigma$ is insensitive to nonanalyticities of this remote channel. Instead, it only receives residual modifications, which are automatically taken into account by renormalization onto low-energy observables like $\gamma_\sigma$, $r_\sigma$, etc. Indeed, Deltuva [*et al.*]{} confirmed that dynamical core excitations within the ${{}^{11}\text{Be}}$ bound state barely influence the reaction cross section [@Deltuva:2016his]. In other words, our *effective* single-channel description readily contains all the relevant core excitation information in the pole regime. For illustration, we show in appendix \[App:CoreExc\] that our approach is equivalent to a theory with an explicit ${{}^{10}\text{Be}}^\ast$ field. The deuteron ------------ The deuteron is treated as an $s$-wave neutron-proton bound state with binding energy [@arXiv:nucl-th/9509032]. The product of the small binding momentum and the effective range [@arXiv:nucl-th/9509032] is as small as ${R_\text{c}}/{R_\text{h}}$. Up to NLO , the deuteron can be solved in analogy to ${{}^{11}\text{Be}}$, including field redefinitions . Expressions for the propagator[^2] $G_\d$ around the pole, its residue $Z_\d$, and respective truncations, can be obtained from Eqs. – by replacing all subscripts "$\sigma$“ by "$\d$“, the total mass ${M_\text{Nc}}$ by $2{m_\text{N}}$, and the reduced mass ${\mu_\text{Nc}}$ by ${m_\text{N}}/2$. Relativistic effects and $s$-$d$ mixing are negligible up to NLO, as shown by Chen [*et al.*]{} [@Chen:1999tn]. The Beryllium-11 excited state ------------------------------ A second neutron-core state close to threshold is the first excited state ${{}^{11}\text{Be}}^\ast$ ($1/2^-$). In Halo EFT, it is treated as a $p$-wave bound state [@Hammer:2011ye] with binding energy [@TUNL_Be11], or binding momentum . The Lagrangian part $\mathcal{L}_{2,{{}^{11}\text{Be}}^\ast}$ is given in appendix \[App:Be11Exc\]. As shown in Ref. [@Bertulani:2002sz], shallow $p$-wave states require the inclusion of at least two low-energy parameters. Close to the pole, we choose and the $p$-wave effective range . The propagator expansion then reads $$\label{Eq:PoleExp_pi} iG_\pi({E_\text{cm}}) = i\frac{6\pi}{{\mu_\text{Nc}}}\frac{2}{-r_\pi}\left[ \gamma_\pi^2+2{\mu_\text{Nc}}({E_\text{cm}}+i\epsilon) \right]^{-1}\left( 1+{\mathcal{O}\left({R_\text{c}}/{R_\text{h}}\right)} \right).$$ Similar to the ground state, $r_\pi$ can be obtained from the respective ANC $A_\pi$ [@Hammer:2011ye]. Taking the value of Calci [*et al.*]{} [@Calci:2016dfb], we find $$\label{Eq:rPiFromANC} r_\pi = -\frac{2\gamma_\pi^2}{A_\pi^2}\left( 1+{\mathcal{O}\left({R_\text{c}}/{R_\text{h}}\right)} \right)\approx -\unit[0.95]{fm^{-1}}\,.$$ In the transfer reaction [${{}^{10}\text{Be}}(\text{d},\,\text{p}){{}^{11}\text{Be}}$]{}, intermediate ${{}^{11}\text{Be}}^\ast$ states represent NLO corrections to the reaction amplitude since , and higher orders in Eq.  are at most of N${}^2$LO. For the moment, we neglect the excited state. It will be subject to the NLO discussion in section \[Sec:NLO\]. We note that there are further two-body states, which are neglected in this work. They include the ${{}^{1}\!S}_0$ virtual neutron-proton state and strong proton-core resonances shown in Fig. \[Fig:Levels\]. Corrections from these states to the reaction amplitude are stronger suppressed than such from ${{}^{11}\text{Be}}^\ast$ in our scheme. Further details will be given at the end of section \[Sec:NLO\]. Three-body system {#Sec:3Body} ================= In this section, we derive an integral integration for the reaction cross section from interactions of the Lagrangian $\mathcal{L}$ up to NLO in the ${R_\text{c}}/{R_\text{h}}$ expansion. Firstly, we show which three-body diagrams are induced by strong and Coulomb interactions of the Lagrangian $\mathcal{L}$. Secondly, we construct the LO transfer amplitude and present results for the cross section. At the end of the section, we discuss NLO corrections. Power counting & LO diagrams ---------------------------- The transfer amplitude $T_{\sigma\d}$ connects the two states $${\left|\sigma\right\rangle}\equiv\ {\left|\text{p}+{{}^{11}\text{Be}}\right\rangle},\ {\left|\d\right\rangle}\equiv\ {\left|{{}^{10}\text{Be}}+\text{d}\right\rangle}$$ through strong and Coulomb interactions. In EFT calculations, these interactions have to be classified in a power counting, which exploits the typical momentum scales of the system. ### Momentum scales The typical momentum scales of the three-body system are given by the small binding momentum scale and the core radius ${R_\text{c}}$. The largest subleading corrections in the strong sector are suppressed by ; see above. Coulomb interactions additionally introduce the small "Coulomb momentum“ $$\label{Eq:CoulMom} p_\text{c}\equiv Q_\text{c}\,\alpha\,{\mu_\text{Nc}}\approx \unit[25]{MeV}\lesssim \gamma\,,$$ where is the fine structure constant. Moreover, Rupak and Kong pointed out that external momenta $p$ have to be counted separately from $\gamma$ in the presence of Coulomb photons [@Rupak:2001ci]. In this work, we calculate cross sections for center-of-mass energies . Thus, $p$ is of the order . The two scales $p_\text{c}$ and $p$ form a second expansion parameter , which we will count like $({R_\text{c}}/{R_\text{h}})^2$. ### Strong interaction In Fig. \[Fig:Vab\], we display the neutron exchange diagrams that form the elementary building blocks of the strong interaction part of the reaction amplitude. We denote them by $-iV_{\sigma\d}^{Sm,1m'}$ and $-iV_{\d\sigma}^{1m,S'm'}$, where and $m,m'$ represent total incoming and outgoing spins and their projections, respectively. Let ${\boldsymbol{p}}$ (${\boldsymbol{q}}$) be the incoming (outgoing) relative momentum[^3] and $E$ the center-of-mass energy. We then find $$\begin{aligned} \label{Eq:Vsd} V_{\sigma\d}^{Sm,1m'}\left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) =&\ -\delta^{S1}\delta^{mm'}\,{m_\text{N}}\left[ {\boldsymbol{p}}\cdot{\boldsymbol{q}} + p^2+\frac{1+y}{2}{q}^2-{m_\text{N}}(E+i\epsilon) \right]^{-1}, \\ \label{Eq:Vds} V_{\d\sigma}^{1m,S'm'}\left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) =&\ V_{\sigma\d}^{S'm',1m}\left({\boldsymbol{q}},\,{\boldsymbol{p}};\,E\right),\end{aligned}$$ where is the mass ratio. Due to the $s$-wave nature of the short-range interactions, only transitions between spin states $S=S'=1$ with projections $m=m'$ are possible. In the following, we will refer to the functions in Eqs. – as "neutron exchange potentials“. For neutron exchange iterations, we use the standard power counting of pionless EFT, which counts all momenta formally like $\gamma\sim {R_\text{h}}^{-1}$. Loops, one-body propagators, and $s$-wave two-body propagators then count like $\gamma^5/{m_\text{N}}$, ${m_\text{N}}/\gamma^2$, and $1/(\gamma\,{m_\text{N}})$, respectively. It follows that all neutron exchange iterations are of order ${m_\text{N}}{R_\text{h}}^2$ and have to be resummed at LO. ### Coulomb contributions Next, we consider the Coulomb force, whose repulsion is expected to lower the reaction probability. In calculations, it is usually included as a static two-body potential in addition to some nuclear model interaction. In a strict EFT approach, however, Coulomb interactions can be analyzed in a systematic power counting, which exploits the system’s momentum scales. This procedure reveals the relative importance of strong and Coulomb interactions. Photon couplings in $\mathcal{L}$ induce the diagrams $-i\Gamma_{ab}^{Sm,S'm'}$ in Fig. \[Fig:CoulombDiagrams\]. Their mathematical expressions are given in appendix \[App:CoulDiags\]. In the following, we analyze the diagrams using the Coulomb power counting developed by Rupak and Kong [@Rupak:2001ci]. Bubble diagrams : The one-loop diagrams (a) and (b) in Fig. \[Fig:CoulombDiagrams\] are proportional to the photon propagator ($\sim p^{-2}$) and to $p_\text{c}$; see Eq. . All momenta in the loop ("bubble“) may be counted like $\gamma$.[^4] I.e., we count one-body propagators like ${m_\text{N}}/\gamma^2$ and the loop integration by $\gamma^5/{m_\text{N}}$. The resulting scaling suggests that bubble diagrams are small compared to neutron exchanges since . Box diagrams : In the box diagrams of Fig. \[Fig:CoulombDiagrams\] (c) and (d), the photon is part of a loop. In this case, it is not straightforward to see if the corresponding integral is governed by powers of $p$ or $\gamma$. Since in our case, the safest option is to count the loop like ${m_\text{N}}/\gamma^3$. This scheme is in line with Ref. [@Konig:2014ufa]. The overall scaling ${m_\text{N}}\,p_\text{c}/\gamma^3$ implies that box diagrams are of the same order as neutron exchanges since . In summary, the Rupak and Kong counting suggests that box diagrams should be iterated at LO, while bubble diagrams are subleading . However, one important feature of the bubble diagrams is not captured by this counting. Their photon propagators exhibit infrared divergences at small momentum transfers in the limit of vanishing photon mass; see Eqs. –. In principle, this enhancement could compensate for the discussed suppression. We account for this possibility by including the bubble diagrams already in the LO calculation, as was also done in Ref. [@Konig:2014ufa]. We will then critically assess this choice by comparing the numerical influence of the box and bubble diagrams on the cross section. Note that we only consider diagrams with one photon exchange between two strong interactions. Corrections from two or more successive exchanges should be small since they involve further powers of the small Coulomb momentum $p_\text{c}$. In principle, they could be included by replacing each photon propagator with the full Coulomb $T$ matrix; see for example [@Konig:2014ufa]. We have checked that, for example, $-i\Gamma_{\d\d}$ would be modified by around in the on-shell case. Such effects are neglected in this work. Transfer amplitude at LO ------------------------ By iterating neutron exchanges, Coulomb bubble, and box diagrams to all orders we obtain the LO transfer amplitude $T^{\text{(LO)}}_{\sigma\d}$. The corresponding Faddeev equation (without three-body force) is shown diagrammatically in Fig. \[Fig:T\_LO\]. ![\[Fig:T\_LO\]Transfer and elastic amplitude at LO. Loop integrals one the right-hand side contain LO propagators $G_{a}^{(\text{LO})}$ ($a\in\{\d,\,\sigma\}$) (drawn without circles). The three-body force $C_0(\Lambda)$ is omitted.](Pictures/picture9.pdf) ### Partial wave channels It is beneficial for our purposes to perform a partial wave projection onto the total angular momentum with total spin $S$ and total orbital angular momentum $L$. The respective neutron exchange potentials $$\begin{aligned} \label{Eq:Vab_PW} V_{\sigma\d}^{{{}^{2S+1}\!L}_J,{{}^{3}\!L'}_J}\left(p,\,q;\,E\right) =&\ \delta^{S1}\delta^{LL'}\,\frac{{m_\text{N}}}{pq}\, Q_L\left( -\frac{p^2+\frac{1+y}{2}q^2-{m_\text{N}}(E+i\epsilon)}{pq} \right), \\ V_{\d\sigma}^{{{}^{3}\!L}_J,{{}^{2S'+1}\!L'}_J}\left(p,\,q;\,E\right) =&\ V_{\sigma\d}^{{{}^{2S'+1}\!L'}_J,{{}^{3}\!L}_J}\left(q,\,p;\,E\right),\end{aligned}$$ depend on Legendre functions of the 2^nd^ kind $$Q_L(x_0)\equiv \frac{1}{2}\int_{-1}^1\!\d x\ \frac{P_L(x)}{x+x_0}\,.$$ Unfortunately, partial wave expressions of the Coulomb interactions in Eqs. – are impractically lengthy. Instead, we calculate them numerically via $$\Gamma_{ab}^{{{}^{2S+1}\!L}_J,{{}^{2S'+1}\!L'}_J} \left(p,\,q;\,E\right) =\delta^{LL'}\,\frac{1}{2} \int_{-1}^1\!\d x\,P_L(x)\,\Gamma_{ab}^{S0,S'0} \left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right)\ (a\in\{\d,\,\sigma\})$$ with $x\equiv {\boldsymbol{p}}\cdot {\boldsymbol{q}}/(pq)$. As indicated in Fig. \[Fig:T\_LO\], the LO elastic and transfer amplitudes can be summarized into an amplitude vector $\vec{T}^{\text{(LO)}}$. Due to the fact that the total spins and orbital angular momenta are conserved at LO, we identify a specific partial wave system by the superscript "$[L,J]$“. For incoming (outgoing) relative momenta $p$ ($p'$), we finally obtain the scattering equations $$\begin{aligned} \nonumber \vec{T}^{\text{(LO)}\,[L,J]} &\!\left(p,\,p';\,E\right) =- \underline{\underline{K}}^{\text{(LO)}\,[L,J]} \!\left(p,\,p';\,E\right) \cdot \vec{e}_1 \\ \label{Eq:LSEq_PW} &+4\pi\!\int\!\frac{\d q\,q^2}{(2\pi)^3}\ \underline{\underline{K}}^{\text{(LO)}\,[L,J]} \!\left(p,\,q;\,E\right)\cdot \underline{\underline{\mathcal{G}}}^{\text{(LO)}}\left(q;\,E\right)\cdot \vec{T}^{\text{(LO)}\,[L,J]} \!\left(q,\,p';\,E\right)\end{aligned}$$ with LO kernel and propagator matrices $$\begin{aligned} \label{Eq:TVec_LO} \vec{T}^{\text{(LO)}\,[L,J]} \equiv&\ \begin{pmatrix} T_{\d\d}^{\text{(LO)}}\\ T_{\sigma\d}^{\text{(LO)}} \end{pmatrix}^{{{}^{3}\!{L}}_J,{{}^{3}\!{L}}_J}\,, \\ \label{Eq:KMat_LO} \underline{\underline{K}}^{\text{(LO)}\,[L,J]} \equiv&\ \begin{pmatrix} \Gamma_{\d\d} & V_{\d\sigma}+\Gamma_{\d\sigma}\\ V_{\sigma\d}+\Gamma_{\sigma\d} & \Gamma_{\sigma\sigma} \end{pmatrix}^{{{}^{3}\!{L}}_J,{{}^{3}\!{L}}_J}\,, \\ \label{Eq:GMat_LO} \underline{\underline{\mathcal{G}}}^{\text{(LO)}} \equiv&\ \text{diag}\left[ \mathcal{G}_\d^{\text{(LO)}},\, \mathcal{G}_\sigma^{\text{(LO)}} \right],\end{aligned}$$ and in channel space. For convenience, we introduced the new functions $$\begin{aligned} \label{Eq:NewPropFunctions} \mathcal{G}_a^{(\text{N${}^n$LO})}\left(q;\,E\right) \equiv&\ G_a^{(\text{N${}^n$LO})}\left(E-q^2/(2\mu_a)\right)\ (a\in\{\d,\,\sigma\},\,n\in\mathbb{N}_0)\,,\end{aligned}$$ where $\mu_\d\equiv 2{m_\text{N}}\,{m_\text{c}}/(2{m_\text{N}}+{m_\text{c}})$ and $\mu_\sigma\equiv ({m_\text{N}}+{m_\text{c}}){m_\text{N}}/(2{m_\text{N}}+{m_\text{c}})$. The full transfer amplitude is given as a sum over the partial wave amplitudes and respective projection operators as shown in appendix \[App:PWProj\]. In all calculations, we truncate the sum at some maximal orbital angular momentum $L_\text{max}$ and increase this value toward convergence. Similarly, whenever including Coulomb diagrams, we decrease the photon mass . We find that the cross section convergences at and . ### Unphysical deep bound states To see if Eq.  requires a three-body force for renormalization, we perform an asymptotic analysis for large incoming and loop momenta similar to Ref. [@Griesshammer:2005ga]. In this limit, the nucleon exchange potential dominates over the Coulomb interactions [@Konig:2014ufa]. Thus, we may neglect the Coulomb force for the moment. It turns out that for , the potentials in Eq.  fall of fast enough to produce unique amplitudes solutions. In the system, however, that is not the case. Instead, the amplitudes approach a power law behavior with . It follows that the system exhibits an Efimov effect, i.e., a geometric spectrum of three-body bound states at energies $E=-B_\d-B_3$ [@Efimov:1970zz; @Braaten:2004rn; @Naidon:2016dpf]. We note that reproduces the universal scaling factor of three distiguishable particles with mass ratio presented in Ref. [@Braaten:2004rn]. In the following, we equip Eq.  with a momentum cutoff . The resulting spectrum is shown in Fig. \[Fig:Spectra\] as dashed lines. Coulomb interactions do not influence the large momentum behavior of the system qualitatively. They only push the Efimov states to higher cutoffs (solid lines in Fig. \[Fig:Spectra\]). The system will be renormalized using the three-body coupling $C_0(\Lambda)$ of Eq. . It enters the kernel matrix of Eq.  as a constant $s$-wave potential like $$K_{\d\d}^{\text{(LO)}\,[0,1]}\rightarrow K_{\d\d}^{\text{(LO)}\,[0,1]}+C_0(\Lambda)\,.$$ Note that the choice of this specific three-body force is not unique. One could also introduce it in the transfer or the ${\left|\sigma\right\rangle}$ elastic channel. ![\[Fig:Spectra\]Unrenormalized three-body spectra at LO without (dashed lines) and with (solid lines) Coulomb interactions ($\lambda=\unit[0.1]{MeV}$, converged) for various cutoffs $\Lambda$.](Pictures/picture10.pdf) The quantum numbers of the Efimov states correspond to those of a level in ${{}^{12}\text{B}}$. Experimentally, three such states are known [@TUNL_B12]. In a deuteron-${{}^{10}\text{Be}}$ cluster picture, their binding energies correspond to spatial separations of the deuteron-${{}^{10}\text{Be}}$ pair. Being of the order ${R_\text{c}}$, they do not reflect a separation of scales in the three-body sector. Thus, the cluster picture is not justified and the Efimov states can be understood as artifacts of the short-range approach. However, although unphysical, they do not impose a problem as long as they lie outside the EFT’s region of applicability. Indeed, after renormalization onto cross section data, all three-body states will occur at binding energies and thus far away from the low-energy region; see Fig. \[Fig:Renorm\] (b). Cross section ------------- The differential cross section of the reaction [${{}^{10}\text{Be}}(\text{d},\,\text{p}){{}^{11}\text{Be}}$]{} at a deuteron beam energy $${E_\text{d}}= \frac{2{m_\text{N}}}{\mu_\d}\,\left(E+B_\d\right)$$ can be obtained by multiplying the transfer amplitude by the residue factor $(Z_\sigma Z_\d)^{1/2}$ and evaluating it at on-shell relative momenta $$\bar{p}_a\equiv \sqrt{2\mu_a(E+B_a+i\epsilon)}\,,\ (a\in\{\d,\,\sigma\})\,.$$ The cross section depends on the center-of-mass angle $\theta_\text{cm}$ with . In the ${\left|\d\right\rangle}$ channel, we set the relative momentum to and in the ${\left|\sigma\right\rangle}$ channel we take . The spin-averaged reaction cross section then reads $$\label{Eq:CrossSection} \left(\frac{\d\sigma}{\d\Omega}\right)\left(\theta_\text{cm};\,E\right) =\frac{1}{3}\,\sum_{m,S',m'} \frac{\mu_\d\mu_\sigma}{4\pi^2}\frac{\bar{p}_\sigma}{\bar{p}_\d} Z_d Z_\sigma \left| T_{\d\sigma}^{1m,S'm'}\left(\bar{{\boldsymbol{p}}}_\d,\,\bar{{\boldsymbol{p}}}_\sigma;\,E\right) \right|^2\,,$$ where . Table \[Tab:InputParameters\] summarizes the input parameters needed for the calculation of the reaction cross section up to NLO in the ${R_\text{c}}/{R_\text{h}}$ expansion. At LO, only the binding energies $B_\d$ and $B_\sigma$ are required. At NLO, also the effective range $r_\d$, the ANC $A_\sigma$ of ${{}^{11}\text{Be}}$, and the binding energy $B_\pi$ and ANC $A_\pi$ of ${{}^{11}\text{Be}}^\ast$ enter. Order deuteron ${{}^{11}\text{Be}}$ ${{}^{11}\text{Be}}^\ast$ ------------------------------------------------------------------------- -------------------------------------------------- ----------------------------------------------------- -------------------------------------------------- LO \[${\mathcal{O}\left(1\right)}$\] $B_\d=\unit[2.22]{MeV}$ [@arXiv:nucl-th/9509032] $B_\sigma=\unit[0.50]{MeV}$ [@TUNL_Be11] – \[0.5em\] NLO \[${\mathcal{O}\left({R_\text{c}}/{R_\text{h}}\right)}$\] $r_\d=\unit[1.75]{fm}$ [@arXiv:nucl-th/9509032] $A_\sigma=\unit[0.786]{fm^{-1/2}}$ [@Calci:2016dfb] $B_\pi=\unit[0.18]{MeV}$ [@TUNL_Be11], $A_\pi=\unit[0.129]{fm^{-1/2}}$ [@Calci:2016dfb] : \[Tab:InputParameters\]EFT inputs for the calculation of the reaction cross section up to NLO. ### Coulomb suppression & improved LO system Our first goal is to critically assess the Coulomb power counting performed above. In particular, we would like to validate the proposed LO nature of the Coulomb force in general and of the bubble diagrams specifically, for the experimental energies used by Schmitt [*et al.*]{} [@Schmitt:2012bt]. Given the cutoff-dependence of the $L=0$ channel, we vary $\Lambda$ in the large range $\Lambda\in[300,\,1500]\,\unit{MeV}$ in each calculation. This procedure reveals the potential impact of the $s$-wave three-body force $C_0(\Lambda)$ on the LO reaction cross sections. ![\[Fig:CS\_LO\]LO cross section of [${{}^{10}\text{Be}}(\text{d},\,\text{p}){{}^{11}\text{Be}}$]{} as function of the center-of mass angle ${\theta_\text{cm}}$. For different deuteron energies ${E_\text{d}}$ (lab frame), the results are compared to data (black points) from Ref. [@Schmitt:2012bt]. All bands show cutoff variations . Each single curve is converged at and . Hatched bands exclude Coulomb interactions. Light (dark) filled bands enclosed by dotted lines include the Coulomb box (and bubble) diagrams. Dash-dotted curves represent a $\chi^2$-fit of the full equation system in Fig. \[Fig:T\_LO\] onto the depicted data using the three-body force $C_0(\Lambda)$; see also Fig. \[Fig:Renorm\]. The fit is cutoff-independent for .](Pictures/picture11.pdf) In a first step, we switch off all Coulomb interactions, which yields the uppermost bands (hatched) in Fig. \[Fig:CS\_LO\]. Each curve is converged at percent level for . At all four deuteron beam energies (lab frame), the bands lie high above the experimental data by Schmitt [*et al.*]{} [@Schmitt:2012bt]. Apparently, the strong interaction alone does not produce enough repulsion between the scattering partners, even if $C_0(\Lambda)$ is included. In order to understand the relative importance of the Coulomb box and bubble diagrams, we add them successively to the Faddeev equation. The light bands surrounded by dotted lines in Fig. \[Fig:CS\_LO\] show that the box diagrams alone lower the cross sections drastically at all beam energies as expected. Indeed, it is important to include them at LO. Further repulsion comes from the bubble diagrams. Their inclusion yields the dark lowermost bands in Fig. \[Fig:CS\_LO\]. Apparently, the influence of the bubble diagrams on the cross section is smaller than the one of the box diagrams. Thus, it seems as if we have overestimated the enhancement due to the bubble diagrams’ infrared divergences by one order in ${R_\text{h}}/{R_\text{c}}$. A posteriori, the bubble diagrams are of NLO and could in principle be neglected at LO. The "pure LO“ system then only contains neutron transfer and box diagrams. Interestingly, however, the inclusion of the bubble diagrams as one specific NLO correction leads to a surprisingly good agreement with the cross section data at lower beam energies and forward angles. Thus, choosing the "improved LO“ system of Fig. \[Fig:T\_LO\] significantly accelerates the EFT convergence. This statement will be verified later by including the remaining NLO corrections. Moreover, the improved LO system, unlike the pure one, can be renormalized onto data at since the respective band comprises all data points. ### Peripherality regions Although subleading in a strict sense, the bubble diagrams do not introduce any new parameters like, for example, effective range coefficients. Thus, the improved LO system stays independent of short-range details. Cross sections are then only affected by the tail of the ${{}^{11}\text{Be}}$ wave function, i.e., the reaction is purely "peripheral“. Yang and Capel argued that such a description is sufficient to describe the reaction at lower beam energies and forward angles [@Yang:2018nzr]. Our results provide clear evidence for this claim since the improved LO band for perfectly describes the whole data region . Moreover, according to Yang and Capel, the peripherality region increases (decreases) in size for lower (higher) energies. Indeed, at , only forward scattering is captured by the improved LO band. Deviations at larger angles are of NLO size. At even higher energies , however, the bands deviate from data by . We conclude that the reaction is indeed only peripheral at forward angles and low energies. That implies that our short-range EFT may fail at energies above . ### Cutoff-dependence & renormalizability Out of all components , only the $L=0$ part is cutoff-dependent. Due to this circumstance, the band widths in Fig. \[Fig:CS\_LO\] are only $\unit[20]{\%}$ the size of the box diagram shift (LO). Such contributions are negligible up to NLO. Thus, in principle, each curve within the filled bands represents an LO result itself and renormalization is not required. Let us emphasize that the only inputs to our LO system are then given by the binding energies $B_\d$ and $B_\sigma$; see Tab. \[Tab:InputParameters\]. At astrophysical energies, however, the $L=0$ component is of much greater importance, leading to a much stronger cutoff dependence. We demonstrate the renormalizability of the improved LO system using the three-body force $C_0(\Lambda)$. For various cutoffs , we adjust it in a $\chi^2$-fit to the depicted ${E_\text{d}}=\unit[12]{MeV}$ data set. This procedure yields the two solutions for $C_0(\Lambda)$ shown in Fig. \[Fig:Renorm\] (a). Their fit values (solid curve) and (dot-dashed curve) are, within numerical uncertainties, equal in size and respectively constant for . For illustration, we show fit results for in Fig. \[Fig:CS\_LO\] as dot-dashed curves. The first three-body state occurs at (or ); see Fig. \[Fig:Renorm\] (b). It lies above (or ) and converges to even higher values as . Corrections at NLO and beyond {#Sec:NLO} ----------------------------- We now discuss NLO contributions to the reaction cross section in the ${R_\text{c}}/{R_\text{h}}$ expansion, stemming from range corrections in the two-body sectors and from the excited state ${{}^{11}\text{Be}}^\ast$. ### Effective range corrections A straightforward way to include effective range corrections in the deuteron and ${{}^{11}\text{Be}}$ is to replace the LO propagators $\mathcal{G}_a^{\text{(LO)}}$ by $\mathcal{G}_a^{\text{(NLO)}}$ in Eq.  [@Bedaque:2002yg].[^5] Note that this approach reintroduces a cutoff-dependence in the $L=0$ channel. In principle, it could be cured by readjusting the three-body force $C_0(\Lambda)$ [@Hammer:2001gh]. In order to see the impact of the additional cutoff-dependence, we include effective range corrections in the renormalized improved LO system for various .[^6] Figure \[Fig:CS\_NLO\] shows that the resulting red hatched bands lie well within the $\unit[\pm 40]{\%}$ LO uncertainty band (blue, enclosed by thin solid lines) of the improved LO estimates (blue dot-dashed curves). The band widths are comparably small, giving rise to a mild cutoff dependence. ![\[Fig:CS\_NLO\]Cross section of [${{}^{10}\text{Be}}(\text{d},\,\text{p}){{}^{11}\text{Be}}$]{} up to NLO as function of the center-of-mass angle ${\theta_\text{cm}}$. The outer blue LO bands (enclosed by thin solid lines) correspond to $\unit[\pm 40]{\%}$ variations of the improved LO results (dot-dashed). The red hatched bands result from effective range corrections and the red bands enclosed by dotted lines also include corrections from the excited state ${{}^{11}\text{Be}}^\ast$ at NLO. Both NLO calculations involve cutoff variations . The final red NLO bands (enclosed by thick solid lines) represent $\unit[\pm 16]{\%}$ variations of the averaged NLO results.](Pictures/picture14.pdf) It has to be mentioned that a small fraction of the band widths stems from an unexpected cutoff dependence in the $L=1$ sector. It can be understood as an artifact of the choice, not to perturb the amplitude itself to first order in ${R_\text{c}}/{R_\text{h}}$, but the integration kernel. That modifies the UV behavior of the partial wave amplitudes, leading to a divergence in the $L=1$ sector. This divergence would not be present in a strictly perturbative approach [@Griesshammer:2005ga]. Even though desirable, such a more involved NLO treatment lies beyond the scope of this work. In fact, we have checked that the influence of the cutoff on the $L=1$ amplitude is less than $\unit[2]{\%}$ over the range . Thus, this issue can be neglected at NLO. ### The Beryllium-11 excited state The excited state ${{}^{11}\text{Be}}^\ast$ introduces a third channel to the three-body system. It couples to ${\left|\d\right\rangle}$ via the diagrams $-iV_{\pi\d},\,-iV_{\d\pi}$ shown in Fig. \[Fig:VabExc\] (a,b). Their mathematical forms and partial wave projections are given in appendix \[App:Be11Exc\]. We note that ${\left|\pi\right\rangle}$ only occurs as an intermediate state in the reaction. Thus, the NLO nature of ${{}^{11}\text{Be}}^\ast$ follows from the propagator scaling ; see section \[Sec:2Body\]. A typical contribution to the reaction amplitude is given by Fig. \[Fig:VabExc\] (c). Again, we count all loop momenta like . Note that the two (neutron-core)-${{}^{11}\text{Be}}^\ast$ vertices contribute a factor $\gamma^2$. The overall scaling ${m_\text{N}}\,{R_\text{c}}{R_\text{h}}$ is then one order smaller than the LO scaling ${m_\text{N}}{R_\text{h}}^2$. We complete the NLO system by inserting both effective range corrections in $G_\d$ and $G_\sigma$, and the potentials $V_{\pi\d},V_{\d\pi}$ into the integration kernel. The resulting Faddeev equations are given in appendix \[App:NLOEquations\]. Similar to the previous calculation, we vary and include the LO three-body force $C_0(\Lambda)$. Figure \[Fig:CS\_NLO\] shows that the results of the previous calculation (hatched bands) get shifted back toward the improved LO results, ending up as red bands enclosed by dotted lines. Thus, the influence of ${{}^{11}\text{Be}}^\ast$ is indeed of NLO, in agreement with our power counting. The remaining cutoff dependences of the $L=0$ and $L=1$ sectors are negligible compared to N${}^2$LO corrections ($\pm \unit[16]{\%}$, red uncertainty bands enclosed by thick solid lines). Thus, no further renormalization is needed at NLO. Recall that the NLO parameters and were calculated in Eqs.  and from the ANCs of Calci [*et al.*]{} [@Calci:2016dfb]. Instead, one could directly use the Halo EFT values and of Hammer and Phillips [@Hammer:2011ye]. The relative differences $\unit[30]{\%}$ and $\unit[40]{\%}$ are of size ${R_\text{c}}/{R_\text{h}}$ and should thus be negligible at NLO. We have checked that the final NLO bands would only indeed only change by ca. $\unit[5]{\%}$. Thus, both choices for $r_\sigma,\,r_\pi$ are consistent with the proposed power counting. ### Higher-order channels At higher orders in Halo EFT, additional scattering channels enter the calculation. For example, the proton-neutron sector exhibits a shallow ${{}^{1}\!S}_0$ virtual state [@Kaplan:1998tg; @Kaplan:1998we]. It does not occur at LO, because the total neutron-proton spin $S=1$ is conserved if all interactions are of $s$-wave type. In the presence of the $p$-wave state ${{}^{11}\text{Be}}^\ast$, however, $S$ may change, and transitions become possible; see Fig. \[Fig:VirtualState\]. However, the virtual state is not only suppressed due to the intermediate ${\left|\pi\right\rangle}$ channel. Since multiple spin changes ( or smaller) are negligible at NLO, a virtual state leads to $S=0$ in the final state of [${{}^{10}\text{Be}}(\text{d},\,\text{p}){{}^{11}\text{Be}}$]{}. The corresponding phase space is $1/3$ the size of $S=1$, yielding a suppression of . We thus neglect virtual states at NLO. ![\[Fig:VirtualState\]The excited state ${{}^{11}\text{Be}}^\ast$ allows transitions from total spin $S=1$ to $S=1$ (${\left|\d\right\rangle}\rightarrow{\left|\pi\right\rangle}\rightarrow{\left|\d\right\rangle}$) or to $S=0$ (${\left|\d\right\rangle}\rightarrow{\left|\pi\right\rangle}\rightarrow{\left|\text{np}({{}^{1}\!S}_0)+{{}^{10}\text{Be}}\right\rangle}$). The thickened solid-dotted double line represents the neutron-proton ${{}^{1}\!S}_0$ virtual state. Multiple transitions via ${\left|\pi\right\rangle}$ are negligible at NLO.](Pictures/picture18.pdf) In Ref. [@Goosman:1970vn], several ${{}^{11}\text{B}}$ resonances have been observed in ; see Fig. \[Fig:Levels\]. The lowest one ($1/2^+$) occurs at a proton-core center-of-mass energy . It has a total width and the branching ratio for decay into is close to $1$ [@Goosman:1970vn]. The resonance represents a pole at in the Coulomb-modified resonance propagator; see for example Refs. [@Kok:1982dr; @Kong:1998sx]. This pole position implies effective range terms and , which scale like ${R_\text{c}}^{-1}$. Moreover, in three-body diagrams, the resonance propagator comes along with a Gamow-Sommerfeld factor [@Kong:1999mp]. It gives the probability of two charged particles to meet in one point. At resonance, it takes the small value . It follows that the influence of the resonance propagator on the reaction is suppressed by three orders in ${R_\text{c}}/{R_\text{h}}$ compared to $G_\sigma$. Note that there are more ${{}^{11}\text{B}}$ states around $E=0$, which could possibly couple strongly to the proton-core system. However, transitions to those states would involve even smaller Gamow factors . Thus, we neglect strong proton-core interactions at NLO. During the reaction process, the ${{}^{11}\text{Be}}$ state may break up into an excited core ${{}^{10}\text{Be}}^\ast$ and a neutron. Thus, ${\left|\sigma\right\rangle}$ in principle couples to the additional intermediate channel via neutron exchanges. However, each such channel comes along with two couplings of order ${R_\text{c}}^2$; see appendix \[App:CoreExc\] for details. Thus, dynamical core excitations can be neglected at NLO. Summary & outlook {#Sec:SummaryOutlook} ================= In this work, we carried out the first Halo EFT calculation of deuteron-induced transfer reactions. As a working example, we considered [${{}^{10}\text{Be}}(\text{d},\,\text{p}){{}^{11}\text{Be}}$]{}, involving the one-neutron halo nucleus ${{}^{11}\text{Be}}$. The degrees of freedom in this approach are the ${{}^{10}\text{Be}}$ core, the neutron, and the proton. Strong interactions are described by contact forces alone. To obtain the differential cross section, the reaction amplitude was constructed diagrammatically in an expansion in the ratio of core and halo radius. The corresponding Faddeev equation contains all dynamical features of a transfer reaction. A three-body force ensures internal consistency. We included the Coulomb force by considering the dominant photon exchange diagrams, which were iterated to all orders in the Faddeev equation. The differential cross section was compared to experimental data by Schmitt [*et al.*]{} [@Schmitt:2012bt]. In agreement with Yang and Capel [@Yang:2018nzr], who calculated the cross section in the adiabatic distorted wave approximation, we found that Halo EFT is able to describe scattering at low beam energies (center-of-mass energies ). In this regime, the reaction can be considered peripheral, i.e., it predominantly depends on the long-range tail of the ${{}^{11}\text{Be}}$ wave function. This part is systematically reproduced by the ${R_\text{c}}/{R_\text{h}}$ expansion. Our theory contains only few information on the spectra of the involved particles. We included, in particular, only two-body states with a binding momentum $\gamma$ clearly smaller than the respective momentum scale of short-range physics; see Fig. \[Fig:Levels\]. The influence of such states should be enhanced by powers of $\gamma^{-1}$ compared to those far away from the two-body threshold. As a consequence of this reduction, we were able to describe data using only a minimal amount of experimental input. At leading order (LO) in the ${R_\text{c}}/{R_\text{h}}$ expansion, only the binding energies of deuteron and ${{}^{11}\text{Be}}$ are needed; see Table \[Tab:InputParameters\]. Next-to-leading-order (NLO) corrections arise from respective effective ranges and the first exited state ${{}^{11}\text{Be}}^\ast$. The effective ranges of the ${{}^{11}\text{Be}}$ states were extracted from the ANCs of the ab-initio calculation by Calci [*et al.*]{} [@Calci:2016dfb]. Both NLO corrections modify the cross section at a $\unit[40]{\%}$ level, as predicted by the power counting. While our results describe data at fairly well, they strongly overestimate the cross section at higher beam energies. Apparently, the low-energy expansion of Halo EFT converges, if at all, slowly at these energies. In order to improve the expansion, it might be necessary to modify the three-body power counting, which, at the moment, counts loop momenta like small binding momenta. For example, dynamical core excitations might then already occur at lower orders. Further hints on missing ingredients in our approach can be inferred from previous theoretical analyses, e.g. by Schmitt [*et al.*]{} in Ref. [@Schmitt:2012bt], Deltuva [*et al.*]{} in Ref. [@Deltuva:2016his] or Yang and Capel in Ref. [@Yang:2018nzr], which were successfull in describing also scattering for . We note that the model used in Ref. [@Yang:2018nzr] contains the same amount of information on the ${{}^{11}\text{Be}}$ spectrum as our work. Thus, we do not expect the inclusion of ${{}^{11}\text{Be}}$ levels beyond the first excited state to be of prime importance. Instead, we might need to consider not explicitly measured loss channels, in particular due to deep ${{}^{11}\text{B}}$ states indicated in Fig. \[Fig:Levels\], at these energies. Usually, such effects are included using optical model potentials, adjusted to, for example, proton-core scattering data. In the future, we will instead introduce imaginary contact terms to the strong Lagrangian, a method called "Open EFT“ [@Braaten:2016sja]. It was applied successfully to a broad range of inelastic processes including quarkonium decays in nonrelativistic QCD [@Bodwin:1994jh] and three-body recombinations of ultracold atoms [@Braaten:2001hf]. Let us emphasize again, that Halo EFT is ideally suited for the description of strong interactions at low energies. In this sense, our long-term goal is to apply the developed framework to the astrophysical regime. Although Coulomb interactions become nonperturbative then, short-range effects like higher-order states should be even less important. Moreover, it will be interesting to calculate cross sections for other deuteron-induced reactions like . We thank D. R. Phillips for giving valuable feedback on the manuscript and S. König for providing information on the calculation of the Coulomb box diagrams. M. S. appreciates stimulating discussions with I. Thompson, D. Baye, and other participants of the INT Program INT-17-1a "Toward Predictive Theories of Nuclear Reactions Across the Isotopic Chart“. Moreover, M. S. sincerely thanks the Nuclear Theory groups of UT Knoxville and Oak Ridge National Laboratory for their kind hospitality and support during his research stay. This work has been funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Projektnummer 279384907 – SFB 1245, by the National Science Foundation under Grant No. PHY-1555030, and by the Office of Nuclear Physics, U.S. Department of Energy under Contract No. DE-AC05-00OR22725. Core excitation effects {#App:CoreExc} ======================= In this section, we show that core excitation effects in the pole region are taken care of in this work due to renormalization onto low-energy observables. For that, we consider a theory with an explicit ${{}^{10}\text{Be}}^\ast$ field $C_m$ by adding a piece $$\label{Eq:L1Be10Exc} \mathcal{L}_{1,{{}^{10}\text{Be}}^\ast} =C_m^\dagger\left( \partial_0 + \frac{\nabla^2}{2{m_\text{c}}}-{E_\text{x}}\right)C_m$$ to the Lagrangian. A similar approach has been chosen by Zhang [*et al.*]{} to analyze effects of the core excitation ${{}^{7}\text{Li}}^\ast$ on the reaction [@Zhang:2013kja]. Moreover, Zhang [*et al.*]{} and Ryberg [*et al.*]{} used a ${{}^{7}\text{Be}}^\ast$ core excitation field in their calculation of the $S$-factor of [@Zhang:2014zsa; @Ryberg:2014exa]. In both systems, the core excitation occurs at low energies. That, however, is not true in our case since . Together with a neutron, ${{}^{10}\text{Be}}^\ast$ couples to the ${{}^{11}\text{Be}}$ ground state in a $d$-wave. In terms of the redefined field $\tilde{\sigma}_{\alpha}$, we thus write $$\label{Eq:L2Be10Exc} \mathcal{L}_{2,{{}^{10}\text{Be}}^\ast} =-\!\!\!\!\sum_{s\in\{3/2,\,5/2\}}\ \frac{g_{\sigma,\text{x}}^{(s)}}{g_\sigma}\, {\text{C}_{1/2\alpha,2m}^{sm_s}}{\text{C}_{2m_l,sm_s}^{1/2 \alpha'}} \left[ \tilde{\sigma}_{\alpha'}^\dagger \left( n_\alpha \left[-i\overleftrightarrow{\nabla}\right]_{2m_l}C_m \right)+\text{H.c.} \right].$$ The vertex term contains a Galilei-invariant derivative . It is embedded in the tensor structure $$\label{Eq:LWaveTensor} \left[\mathcal{O}\right]_{lm_l} \equiv \sqrt{\frac{4\pi}{2l+1}}\,\left| {\boldsymbol{\mathcal{O}}} \right|^l\, Y_{l}^{m_l}\left(\hat{{\boldsymbol{\mathcal{O}}}}\right)$$ with $l=2$, where $Y_l^{m_l}(\hat{{\boldsymbol{\mathcal{O}}}})$ denotes a spherical harmonic, evaluated at . The mass difference in the transition is of natural size. Thus, we assume no fine-tuning in this scattering channel and count . It follows that the overall coupling is natural as well, since ; see Eq. . The core excitation modifies the ${{}^{11}\text{Be}}$ propagator through the ${{}^{10}\text{Be}}^\ast$-neutron self-energy loop . It resembles the neutron-core self-energy loop in Fig. \[Fig:Dyson\_sigma\], but all core lines have to be replaced by core excitation lines. Using the PDS scheme, we find $$\begin{aligned} \nonumber \Sigma_{\sigma,\text{x}}({E_\text{cm}}) =&\ -\sum_s\left(\frac{g_{\sigma,\text{x}}^{(s)}}{g_\sigma}\right)^2\frac{{\mu_\text{Nc}}}{10\pi}\, \left[ 2{\mu_\text{Nc}}({E_\text{cm}}-{E_\text{x}}+i\epsilon) \right]^2 \\ &\hspace{2cm}\times\left( \Lambda_\text{PDS}-\left[ -2{\mu_\text{Nc}}({E_\text{cm}}-{E_\text{x}}+i\epsilon) \right]^{1/2} \right) \\ \equiv&\ -g_\sigma^{-2}\sum_n \Delta_{\sigma,\text{x}}^{(n)}({E_\text{cm}}+i\epsilon)^n.\end{aligned}$$ Note that $\Sigma_{\sigma,\text{x}}$ is analytic for , i.e., it can be expanded at ${E_\text{cm}}=0$. The resulting coefficients $\Delta_{\sigma,\text{x}}^{(n)}$ then contribute to the unrenormalized parameters $\Delta_\sigma^{(n)}$ of the bare ${{}^{11}\text{Be}}$ propagator; see Eq. . Thus, renormalization onto observables $\gamma_\sigma$ (or $a_\sigma$), $r_\sigma$, etc. automatically takes care of core excitation effects at small ${E_\text{cm}}$, where the pole is located. In other words, $C_m$ does not introduce any new information to the two-body sector and can be integrated out. Partial wave expansion {#App:PWProj} ====================== Let us consider a general interaction $\mathcal{I}$, which could be an amplitude $T$, a neutron exchange potential $V$ or a Coulomb interaction $\Gamma$. We expand $\mathcal{I}$ in tensor spherical harmonics $$\label{Eq:VecSphHarm} \left({\boldsymbol{Y}}_{(L,S)Jm_J}(\hat{{\boldsymbol{p}}})\right)^m \equiv \sqrt{4\pi}\,\sum_{m_L} {\text{C}_{Lm_L,Sm}^{Jm_J}} Y_{L}^{m_L}(\hat{{\boldsymbol{p}}})$$ by writing $$\begin{aligned} \label{Eq:PartialWaveExpansion} \mathcal{I}^{Sm,S'm'}\left({\boldsymbol{p}},{\boldsymbol{q}};\,E\right) =&\ \sum_J \sum_{L,L'} \mathcal{I}^{{{}^{2S+1}\!L}_J,{{}^{2S'+1}\!L'}_J}\left(p,\,q;\,E\right) P_{\,{{}^{2S+1}\!L}_J,{{}^{2S'+1}\!L'}_J}^{\,m,m'}\left(\hat{{\boldsymbol{p}}},\,\hat{{\boldsymbol{q}}}\right), \\ P_{\,{{}^{2S+1}\!L}_J,{{}^{2S'+1}\!L'}_J}^{\,m,m'}\!\left(\hat{{\boldsymbol{p}}},\,\hat{{\boldsymbol{q}}}\right) \equiv&\ \sum_{m_J} \left( {\boldsymbol{Y}}_{(L,S)Jm_J}\left(\hat{{\boldsymbol{p}}}\right) \right)^m \left({\boldsymbol{Y}}_{(L',S')Jm_J} \left(\hat{{\boldsymbol{q}}}\right)\right)^{m'\,\ast}\,.\end{aligned}$$ Specific partial waves can be extracted via $$\begin{aligned} \mathcal{I}^{{{}^{2S+1}\!L}_J,{{}^{2S'+1}\!L'}_J}\left(p,\,q;\,E\right) = \frac{(4\pi)^{-2}}{2J+1}\sum_{m,m'}\int_{\Omega_{{\boldsymbol{p}}},\Omega_{{\boldsymbol{q}}}} P_{\,{{}^{2S'+1}\!L'}_J,{{}^{2S+1}\!L}_J}^{\,m',m}\!\left(\hat{{\boldsymbol{q}}},\,\hat{{\boldsymbol{p}}}\right) \, \mathcal{I}^{Sm,S'm'}\left({\boldsymbol{p}},{\boldsymbol{q}};\,E\right).\end{aligned}$$ Coulomb diagrams {#App:CoulDiags} ================ The Coulomb interactions in Fig. \[Fig:CoulombDiagrams\] resemble such considered by König [*et al.*]{} for the three-nucleon system [@Konig:2015aka]. However, they exhibit nontrivial dependencies on the mass ratio . The bubble interactions read $$\begin{aligned} \nonumber \Gamma_{\d\d}^{1m,1m'} \left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) =&\ \delta^{mm'}\,\frac{ Q_\text{c}\alpha\,{m_\text{N}}^2 }{ \left({\boldsymbol{p}}-{\boldsymbol{q}}\right)^2 +\lambda^2-i\epsilon } \\ \label{Eq:Gamma_dd} &\ \times f\Big( {\boldsymbol{p}}-{\boldsymbol{q}},\,A_\d(p;\,E),\,A_\d(q;\,E) \Big)\,, \\ \nonumber \Gamma_{\sigma\sigma}^{Sm,S'm'} \left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) =&\ \delta^{SS'}\delta^{mm'}\,\frac{ Q_\text{c}\alpha\,(2{\mu_\text{Nc}})^2 }{ \left({\boldsymbol{p}}-{\boldsymbol{q}}\right)^2 +\lambda^2-i\epsilon } \\ \label{Eq:Gamma_ss} &\ \underbrace{ \times f\Bigg( \frac{y}{\xi}\left({\boldsymbol{p}}-{\boldsymbol{q}}\right),\,A_\sigma(p;\,E),\,A_\sigma(q;\,E) \Bigg) \,, }_{=\big( \sqrt{A_\sigma(q;\,E)}-\sqrt{A_\sigma(p;\,E)} \big)/\big( A_\sigma(q;\,E)-A_\sigma(p;\,E) \big) +{\mathcal{O}\left(y^2\right)} }\end{aligned}$$ and the box interactions are given by $$\begin{aligned} \nonumber \Gamma_{\sigma\d}^{Sm,1m'} \left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) =&\ -Q_\text{c}\alpha\,{m_\text{N}}\, V_{\sigma\d}^{Sm,1m'}\left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) \\ \nonumber &\ \times\Bigg[ f\Bigg( {\boldsymbol{p}}-y{\boldsymbol{q}},\,\xi^2 A_\sigma(p;\,E),\,A_\d(q;\,E) \Bigg) \\ \label{Eq:Gamma_sd} &\hspace{4em} -\frac{\lambda}{ {\boldsymbol{p}}\cdot{\boldsymbol{q}} + p^2+\xi{q}^2 -{m_\text{N}}(E+i\epsilon) } +{\mathcal{O}\left(\lambda^2\right)} \Bigg]\,, \\ \label{Eq:Gamma_ds} \Gamma_{\d\sigma}^{1m,S'm'} \left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) =&\ \Gamma_{\sigma\d}^{S'm',1m} \left({\boldsymbol{q}},\,{\boldsymbol{p}};\,E\right),\end{aligned}$$ where we defined . Moreover, is the fine structure constant and is the core charge. All interactions involve the function $$f\left({\boldsymbol{\Delta}},\,A_1,\,A_2\right) \equiv \frac{1}{ \left|{\boldsymbol{\Delta}}\right| }\, \text{tan}^{-1}\!\left( \frac{ A_1-A_2 +{\boldsymbol{\Delta}}^2/4 }{ \left|{\boldsymbol{\Delta}}\right| \sqrt{A_2} } \right) +\left[A_1\leftrightarrow A_2\right],$$ whose arguments involve the expressions $$\begin{aligned} A_\d(p;\,E)\equiv&\ \frac{1+2y}{4}\,p^2-{m_\text{N}}(E+i\epsilon) \xrightarrow{\text{\ on-shell\ }}\gamma_\d^2\,, \\ A_\sigma(p;\,E)\equiv&\ \xi^{-2}\,\frac{1+2y}{4}\,p^2-\xi^{-1}{m_\text{N}}(E+i\epsilon) \xrightarrow{\text{\ on-shell\ }}\gamma_\sigma^2\,.\end{aligned}$$ The form of $\Gamma_{\sigma\sigma}$ can be simplified significantly by neglecting terms of order ${\mathcal{O}\left(y^2\right)}$; see Eq. . This approximation is justified since is a tiny number. The only angular dependence then comes from the photon propagator, which can be projected onto certain partial waves analytically. The bubble diagrams $-i\Gamma_{\d\d}^{1m,1m'}$ and $-i\Gamma_{\sigma\sigma}^{Sm,S'm'}$ are linear in the Coulomb propagator. Thus, their largest contributions to the transfer reaction comes from the region of small momentum transfers ${\boldsymbol{p}}-{\boldsymbol{q}}$. For ${\boldsymbol{p}}={\boldsymbol{q}}$, the values of the function $f$ in Eqs. – collapse to . Thus, the deuteron and halo loops of the LO bubble diagrams in Fig. \[Fig:CoulombDiagrams\] may be counted like ${m_\text{N}}/\gamma$. The Coulomb interactions can be connected to the $s$-wave projected functions $K_\text{bubble}$ and $K_\text{box}$ of Ref. [@Konig:2015aka] by taking the limits . We find $$\begin{aligned} \left. \int_{-1}^1\! \d x\ \Gamma_{aa}^{10,10} \left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) \right|_{y,Q_\text{c}\rightarrow 1} =&\ -\frac{{m_\text{N}}}{4\pi}\,K_\text{bubble}\left(E;\,p,\,q\right)\ (a\in\{\d,\,\sigma\}), \\ \left. \int_{-1}^1\! \d x\ \Gamma_{\sigma\d}^{10,10} \left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) \right|_{y,Q_\text{c}\rightarrow 1} =&\ -\frac{{m_\text{N}}}{2\pi}\,K_\text{box}\left(E;\,p,\,q\right),\end{aligned}$$ where . Excited state of Beryllium-11 {#App:Be11Exc} ============================= In this section, we discuss the inclusion of the excited state ${{}^{11}\text{Be}}^\ast$ at NLO in the reaction calculation. The Lagrangian part $$\begin{aligned} \nonumber \mathcal{L}_{{{}^{11}\text{Be}}^\ast} =&\ \pi_\alpha^\dagger \left[ \Delta_\pi^{(0)} +\left( i\partial_0+\frac{{\boldsymbol{\nabla}}^2}{2M_\text{Nc}} \right) \right]\pi_\alpha \\ \label{Eq:LBe11Exc} &\ -g_\pi\ {\text{C}_{1/2\,\alpha,1m_l}^{1/2\,\alpha'}} \left[ \pi_{\alpha'}^\dagger \left( n_\alpha \left[ -i\overleftrightarrow{\nabla} \right]_{1m_l}\, c \right) + \text{H.c.} \right]\end{aligned}$$ of Eq.  contains an auxiliary field $\pi_\alpha$ for ${{}^{11}\text{Be}}^\ast$ with renormalization-dependent parameters $\Delta_\pi^{(0)},\,g_\pi\in\mathbb{R}$. The Galilei-invariant derivative $\overleftrightarrow{\boldsymbol{\nabla}}$ and the $p$-wave tensor structure $[\mathcal{O}]_{1m_l}$ are defined in appendix \[App:CoreExc\]. Unlike in the $s$-wave case, both the constant and derivative part of the bare propagator term in Eq.  are needed to describe the shallow $p$-wave state [@Bertulani:2002sz; @Hammer:2011ye]. The full ${{}^{11}\text{Be}}^\ast$ propagator can be obtained by resumming all two-body loops, similarly to Fig. \[Fig:Dyson\_sigma\]. For more details, we refer to Ref. [@Hammer:2011ye]. After proper renormalization and field redefinitions , the propagator $G_\pi$ around the pole at ${E_\text{cm}}=-B_\pi$ is given by Eq. . In the NLO three-body system, the intermediate state couples to ${\left|\d\right\rangle}$ via neutron exchange potentials shown in Fig. \[Fig:VabExc\]. They read $$\begin{aligned} \nonumber V_{\pi\d}^{Sm,1m'}\left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) =&\ {m_\text{N}}\,\sqrt{6}\left\lbrace \begin{matrix} S&1&1\\ 1/2&1/2&1/2 \end{matrix} \right\rbrace \\ &\times \frac{ \sum_{m_l}{\text{C}_{1m_l,1m'}^{Sm}}\left[\frac{1}{1+y}\,{\boldsymbol{p}}+{\boldsymbol{q}}\right]_{1m_l}^\ast }{ {\boldsymbol{p}}\cdot{\boldsymbol{q}} + p^2+\frac{1+y}{2}{q}^2-{m_\text{N}}(E+i\epsilon) }\,, \\ V_{\d\pi}^{1m,S'm'}\left({\boldsymbol{p}},\,{\boldsymbol{q}};\,E\right) =\ &\left[ V_{\pi d}^{S'm',1m}\left({\boldsymbol{q}},\,{\boldsymbol{p}};\,E\right) \right]^\ast\end{aligned}$$ with in the ${\left|\pi\right\rangle}$ channel and involve a $6j$-symbol. Partial wave projections are given by $$\begin{aligned} \nonumber V_{\pi\d}^{{{}^{2S+1}\!L}_J\,{{}^{3}\!L'}_J}\left(p,\,q;\,E\right) =&\ (-1)^{J+1}\sqrt{2\,(2S+1)(2L+1)(2L'+1)} \\ \nonumber &\hspace{-1cm}\times{\text{C}_{L0,L'0}^{10}}\,\left\lbrace \begin{matrix} S&1&1\\ 1/2&1/2&1/2 \end{matrix} \right\rbrace \left\lbrace \begin{matrix} S&1&1\\ L'&L&J \end{matrix} \right\rbrace \\ \label{Eq:Vpid} &\hspace{-1cm}\times\frac{{m_\text{N}}}{pq}\left[ \frac{1}{1+y}\,p\,Q_{L'}+q\,Q_L \right]\left( -\frac{p^2+\frac{1+y}{2}q^2-{m_\text{N}}(E+i\epsilon)}{pq} \right), \\ \label{Eq:Vdpi} V_{\d\pi}^{{{}^{3}\!L}_J\,{{}^{2S'+1}\!L'}_J}\left(p,\,q;\,E\right) =&\ V_{\pi\d}^{{{}^{2S'+1}\!L'}_J\,{{}^{3}\!L}_J}\left(q,\,p;\,E\right).\end{aligned}$$ A direct transition potential between ${\left|\sigma\right\rangle}$ and ${\left|\pi\right\rangle}$ is not induced by the Lagrangian, i.e., these states can only be connected via an intermediate state ${\left|\d\right\rangle}$. The Clebsch-Gordan coefficient and the $6j$-symbols in Eq.  imply some selection rules. Firstly, only transitions with are allowed. It follows that for $J=0$, we have , and for fixed , the system decouples into the two subsystems (1) and (2) . Secondly, is fixed in subsystem (1), while both options are allowed in subsystem (2). Lastly, in subsystem (2), the two channels further decouple after defining rotated spin states $$\label{Eq:RotatedPiStates} \begin{pmatrix} {\left|\pi,\,{{}^{\bar{3}}\!J}_J\right\rangle}\\[1ex] {\left|\pi,\,{{}^{\bar{1}}\!J}_J\right\rangle} \end{pmatrix} \equiv \frac{1}{\sqrt{2J+1}} \begin{pmatrix} \sqrt{J+1} & \sqrt{J} \\ -\sqrt{J} & \sqrt{J+1} \end{pmatrix} \begin{pmatrix} {\left|\pi,\,{{}^{3}\!J}_J\right\rangle}\\[1ex] {\left|\pi,\,{{}^{1}\!J}_J\right\rangle} \end{pmatrix}\,.$$ Note that and for $J=0$. The corresponding partial wave potentials read $$\begin{aligned} \nonumber V_{\pi\d}^{{{}^{\overline{2\mp1}}\!J}_J,{{}^{3}\!\,(J\pm1)}_J}(p,\,q;\,E) =&\ \mp\frac{1}{\sqrt{3}} \\ &\hspace{-2cm}\times\frac{{m_\text{N}}}{pq}\left[ \frac{1}{1+y} p\, Q_{J\pm1} + q\, Q_J \right]\left( -\frac{p^2+\frac{1+y}{2}q^2-{m_\text{N}}(E+i\epsilon)}{pq} \right)\,, \\ V_{\d\pi}^{{{}^{3}\!\,(J\pm1)}_J,{{}^{\overline{2\mp1}}\!J}_J}(p,\,q;\,E) =&\ V_{\pi\d}^{{{}^{\overline{2\mp1}}\!J}_J,{{}^{3}\!\,(J\pm1)}_J}(q,\,p;\,E)\,.\end{aligned}$$ Subsystem $L_\d=L_\sigma$ $S_\d=S_\sigma$ $L_\pi$ $S_\pi$ ----------- ----------------- ----------------- --------- ------------------------------------------------------- (1) $J$ $1$ $J\pm1$ $1$ (2a) $J-1$ $1$ $J$ $\bar{3}\propto \sqrt{J+1}\times 3+\sqrt{J}\times 1$ (2b) $J+1$ $1$ $J$ $\bar{1}\propto -\sqrt{J}\times 3+\sqrt{J+1}\times 1$ : \[Tab:NLOSubsystems\]Subsystems of fixed $J$ after including the excited state channel ${\left|\pi\right\rangle}$. Subsystems (1) and (2a) require $J\geq 1$. The quantum numbers $\bar{3}$ and $\bar{1}$ in subsystems (2a) and (2b) refer to rotated spin states of ${\left|\pi\right\rangle}$; see Eq. . In summary, for fixed $J\geq 1$, we find the three decoupled subsystems (1), (2a), and (2b) presented in Tab. \[Tab:NLOSubsystems\]. Just as in the LO case, they can be identified by the conserved quantum number (1) , (2a) , and (2b) . In the case , only system (2b) is allowed. NLO equations {#App:NLOEquations} ============= As explained in appendix \[App:Be11Exc\], the introduction of the excited state ${{}^{11}\text{Be}}^\ast$ produces three decoupled scattering systems for fixed $J\geq 1$, corresponding to , and a single system for $J=0$ with . The respective NLO amplitude vectors $\vec{T}^{\text{(NLO)}\,[L_\d,J]}$ read $$\begin{aligned} \vec{T}^{\text{(NLO)}\,[J,J]} =&\ \begin{pmatrix} T_{\d\d}^{\text{(NLO)}\,{{}^{3}\!J}_J,{{}^{3}\!J}_J}\\[0.5ex] T_{\sigma\d}^{\text{(NLO)}\,{{}^{3}\!J}_J,{{}^{3}\!J}_J}\\[0.5ex] T_{\pi\d}^{{{}^{3}\!(J-1)}_J,{{}^{3}\!J}_J}\\[0.5ex] T_{\pi\d}^{{{}^{3}\!(J+1)}_J,{{}^{3}\!J}_J} \end{pmatrix}\ (J\geq 1)\,, \\ \vec{T}^{\text{(NLO)}\,[J\pm 1,J]} =&\ \begin{pmatrix} T_{\d\d}^{\text{(NLO)}\,{{}^{3}\!(J\pm1)}_J,{{}^{3}\!(J\pm1)}_J}\\[0.5ex] T_{\sigma\d}^{\text{(NLO)}\,{{}^{3}\!(J\pm1)}_J,{{}^{3}\!(J\pm1)}_J}\\[0.5ex] T_{\pi\d}^{{{}^{\overline{2\mp1}}\!J}_J,{{}^{3}\!(J\pm1)}_J} \end{pmatrix}\ (J\geq 0\ \text{and}\ J\geq 1)\,.\end{aligned}$$ They are determined by the kernel and propagator matrices $$\begin{aligned} \underline{\underline{K}}^{\text{(NLO)}\,[J,J]} =&\ \begin{pmatrix} \Gamma_{\d\d}^{{{}^{3}\!J}_J,{{}^{3}\!J}_J} & (V_{\d\sigma}+\Gamma_{\d\sigma})^{{{}^{3}\!J}_J,{{}^{3}\!J}_J} & V_{\d\pi}^{{{}^{3}\!J}_J,{{}^{3}\!(J-1)}_J} & V_{\d\pi}^{{{}^{3}\!J}_J,{{}^{3}\!(J+1)}_J} \\[0.5ex] (V_{\sigma\d}+\Gamma_{\sigma\d})^{{{}^{3}\!J}_J,{{}^{3}\!J}_J} & \Gamma_{\sigma\sigma}^{{{}^{3}\!J}_J,{{}^{3}\!J}_J} &0&0 \\[0.5ex] V_{\pi\d}^{{{}^{3}\!(J-1)}_J,{{}^{3}\!J}_J} &0&0&0 \\[0.5ex] V_{\pi\d}^{{{}^{3}\!(J+1)}_J,{{}^{3}\!J}_J} &0&0&0 \end{pmatrix}\,, \\ \underline{\underline{\mathcal{G}}}^{(\text{NLO})\,[J,J]} =&\ \text{diag}\left[ \mathcal{G}_\d^{\text{(NLO)}},\,\mathcal{G}_\sigma^{\text{(NLO)}},\,\mathcal{G}_\pi^{\text{(LO)}},\,\mathcal{G}_\pi^{\text{(LO)}} \right],\end{aligned}$$ and $$\begin{aligned} \underline{\underline{K}}^{\text{(NLO)}\,[J\pm1,J]} =&\ \begin{pmatrix} \Gamma_{\d\d}^{{{}^{3}\!(J\pm1)}_J,{{}^{3}\!(J\pm1)}_J} & \!\!\!\!(V_{\d\sigma}+\Gamma_{\d\sigma})^{{{}^{3}\!(J\pm1)}_J,{{}^{3}\!(J\pm1)}_J} & V_{\d\pi}^{{{}^{3}\!(J\pm1)}_J,{{}^{\overline{2\mp1}}\!J}_J} \\[0.5ex] (V_{\sigma\d}+\Gamma_{\sigma\d})^{{{}^{3}\!(J\pm1)}_J,{{}^{3}\!(J\pm1)}_J} & \Gamma_{\sigma\sigma}^{{{}^{3}\!(J\pm1)}_J,{{}^{3}\!(J\pm1)}_J} &0 \\[0.5ex] V_{\pi\d}^{{{}^{\overline{2\mp1}}\!J}_J,{{}^{3}\!(J\pm1)}_J} &0&0 \end{pmatrix}\,, \\ \underline{\underline{\mathcal{G}}}^{(\text{NLO})\,[J\pm1,J]} =&\ \text{diag}\left[ \mathcal{G}_\d^{\text{(NLO)}},\,\mathcal{G}_\sigma^{\text{(NLO)}},\,\mathcal{G}_\pi^{\text{(LO)}} \right],\end{aligned}$$ respectively, similar to Eq. . The propagator function $\mathcal{G}_\pi^\text{(LO)}$ is defined via Eq.  with $a=\pi$ and reduced mass . [^1]: The propagator is diagonal in spin space. Respective factors $\delta^{\alpha,\alpha'}$ will be omitted in the following. [^2]: The deuteron propagator is diagonal in spin space, i.e., it has to be multiplied by $\delta^{i,i'}$ in diagrams. [^3]: In this work, relative momenta in the three-body center-of-mass system are defined as the momentum of the respective spectator particle. I.e., they equal in ${\left|\d\right\rangle}$, or in ${\left|\sigma\right\rangle}$. [^4]: This statement can be verified by applying the so-called "bubble approximation“ [@Konig:2014ufa]; see appendix \[App:CoulDiags\]. [^5]: Correspondingly, one has to use the residues $Z_a^{\text{(NLO)}}$ in the calculation of the cross section in Eq. . [^6]: Below , the renormalized improved LO result is not yet converged. Note that the cutoff variation up to $\unit[1500]{MeV}$ is only used to estimate higher-order corrections. It does, however, not reveal the necessity of additional counter terms.
{ "pile_set_name": "ArXiv" }
--- abstract: 'A spectral method is described for solving coupled elliptic problems on an interior and an exterior domain. The method is formulated and tested on the two-dimensional interior Poisson and exterior Laplace problems, whose solutions and their normal derivatives are required to be continuous across the interface. A complete basis of homogeneous solutions for the interior and exterior regions, corresponding to all possible Dirichlet boundary values at the interface, are calculated in a preprocessing step. This basis is used to construct the influence matrix which serves to transform the coupled boundary conditions into conditions on the interior problem. Chebyshev approximations are used to represent both the interior solutions and the boundary values. A standard Chebyshev spectral method is used to calculate the interior solutions. The exterior harmonic solutions are calculated as the convolution of the free-space Green’s function with a surface density; this surface density is itself the solution to an integral equation which has an analytic solution when the boundary values are given as a Chebyshev expansion. Properties of Chebyshev approximations insure that the basis of exterior harmonic functions represents the external near-boundary solutions uniformly. The method is tested by calculating the electrostatic potential resulting from charge distributions in a rectangle. The resulting influence matrix is well-conditioned and solutions converge exponentially as the resolution is increased. The generalization of this approach to three-dimensional problems is discussed, in particular the magnetohydrodynamic equations in a finite cylindrical domain surrounded by a vacuum.' address: | Center for Turbulence Research\ Stanford University, Bldg. 500, Stanford, CA 94305-3035\ e-mail: author: - Piotr Boronski title: Spectral method for matching exterior and interior elliptic problems --- influence matrix, spectral method, Chebyshev polynomials, boundary integral method, magnetohydrodynamics, Green’s functions, harmonic functions, Laplace’s equation, exterior problem Motivation {#sec:motivation} ========== The search for a self-sustaining magnetohydrodyamic dynamo has taken on great momentum in recent years, as researchers have sought to produce dynamos in the laboratory [@Steglitz02; @Riga01; @VKS02; @Forest02; @Shew02] and in simulations [@Jepps67; @Dudley89; @Glatzmaier95; @Tilgner97; @Hollerbach00; @Willis02; @Matsui04; @Iskakov04; @Xu04; @Guermond06]. One of the fundamental problems in numerical magnetohydrodynamics is the formulation of boundary conditions. The governing equations describe the velocity and magnetic field in in a finite container of electrically conducting fluid. At the container boundaries, the velocity is specified, but the magnetic field is not. Instead, the magnetic field is required to satisfy continuity conditions with the exterior magnetic field in the domain surrounding the fluid. The nature of these conditions depends on the properties of the surrounding medium; a complete discussion can be found in [@Roberts67]. Under the quasi-static approximation [@Roberts67], for a given velocity field ${{\mathbf u}}$ and magnetic Reynolds number $Rm$, the equations describing the interior magnetic field are \[eq:Bint\] $$\begin{aligned} {\partial}_t {{\mathbf B}}&=& {{\nabla} \times}({{\mathbf u}}\times{{\mathbf B}})+ \frac{1}{Rm} {\Delta}{{\mathbf B}}\label{eq:induction}\\ {{\nabla} \cdot}{{\mathbf B}}&=& 0 \label{eq:nomonopoles}\end{aligned}$$ The case of a fluid of finite electric conductivity restricted to a finite volume and surrounded by vacuum is of special importance because it models a number of experimental, geophysical, and astrophysical configurations. Since there are no electrical currents in a vacuum, ${{\mathbf B}}^{vac}$ is curl-free, and is therefore the gradient of a potential if the exterior domain is simply connected. The exterior magnetic field then obeys: $$\begin{aligned} {{\mathbf B}}^{\rm vac}&=&{{\nabla}}\phi^{\rm vac}\\ {\Delta}\phi^{\rm vac} &=& 0 \label{eq:phiharm}\\ {{\mathbf B}}^{\rm vac} &\rightarrow& 0\qquad |{{\mathbf x}}|\rightarrow\infty\end{aligned}$$ \[eq:Bext\] but is otherwise not fixed. The magnetic field is required to be continuous at the boundary: $${{\mathbf B}}- {{\mathbf B}}^{\rm vac} = 0 \qquad {{\mathbf x}}\in{{\partial\Omega}}\label{eq:Bcont}$$ In this case, continuity of all three components of the magnetic field are sufficient to uniquely determine both the interior and exterior fields. Our ultimate goal is to transform - into boundary conditions that can be applied to without calculating ${{\mathbf B}}^{\rm vac}$. The general principle we will employ is to construct a complete basis of exterior solutions ${{\mathbf B}}^{\rm vac}$ of in a preprocessing step, and to calculate ${{\mathbf B}}|_{{\partial\Omega}}$ for each member of the basis. The matching conditions will then yield boundary conditions for ${{\mathbf B}}$. To explore this approach, we will apply it to the simpler analogous scalar problem of the Poisson problem in an interior domain. We will require the solution to match continuously to an exterior solution satisfying Laplace’s equation. In this case, both Dirichlet and Neumann matching conditions are necessary to specify a unique solution. Formally, we wish to solve the following problem: \[eq:two\_domains\] $$\begin{aligned} {\Delta}\Phi=\rho &&\quad\text{in }\Omega \\ {\Delta}\phi=0 && \quad \text{outside }\Omega\end{aligned}$$ with boundary conditions: \[eq:match\] $$\begin{aligned} \Phi({{\mathbf x}})-\phi({{\mathbf x}}) &=& 0 \qquad {{\mathbf x}}\in{{\partial\Omega}}\label{eq:matchD}\\ {\partial_n}\Phi({{\mathbf x}})-{\partial_n}\phi({{\mathbf x}}) &=& 0 \qquad {{\mathbf x}}\in{{\partial\Omega}}\label{eq:matchN}\\ \nabla\phi({{\mathbf x}})&\rightarrow& 0 \qquad |{{\mathbf x}}|\rightarrow\infty\end{aligned}$$ where $\Omega$ is a bounded domain with boundary ${{\partial\Omega}}$. A physical interpretation of - is that of an electrostatic potential $\Phi$ of a field generated by charges distributed in space with the density $-\rho$, where the electrical permeability of the vacuum is taken to be one by the choice of units. We wish to calculate the interior solution $\Phi$ without explicitly constructing $\phi$. There exists a vast literature on the numerical solution of the fundamental physical problems – and -. We will briefly survey a small portion of this literature here, and postpone a more detailed comparison between our method and others to a later section. The main tool by which exterior domains can be eliminated is Green’s theorem, which replaces elliptic differential equations over a domain with integrals over the bounding surfaces. The use of methods based on boundary integrals has grown explosively since the 1970s-1980s to solve engineering problems from fields such as acoustics, elasticity, electromagnetism and fluid mechanics [@Hsiao73; @Jaswon77; @Brebbia78; @Banerjee81; @Atkinson82; @Rokhlin83; @Brebbia83; @Pozrikidis92; @KirkupWeb]. Hybrid methods, coupling a differential equation formulation in a domain and a boundary integral formulation at the boundary via an influence matrix, were also developed at the same time to solve – and similar problems. The majority of these approaches have been based on finite elements and are hence applicable to complicated real-world geometries. The boundary integrals are discretized with techniques derived from finite element theory, leading to the term boundary element method, and the hybrid methods use finite elements to solve the equations in the domain. This situation contrasts with magnetohydrodynamics, which has been dominated by spectral methods. Spherical domains are standard, for geophysical and astrophysical reasons. Spectral methods can then be based on spherical harmonics and the poloidal-toroidal decomposition [@Dudley89; @Glatzmaier95; @Tilgner97; @Hollerbach00]. The solution to Laplace’s equation on the exterior is immediate and, moreover, solutions and associated boundary conditions for each spherical harmonic and toroidal or poloidal component are decoupled. Boundary conditions at the interface can then be formulated for each mode without the use of an influence matrix. The technique which we will describe is based on spectral methods, but the geometry is assumed to be somewhat more complicated. Our technique occupies the niche which spectral methods staked out in the 1970s-1980s when the use of Chebyshev polynomials became commonplace [@Orszag71; @Canuto88] to represent domains with one or more non-periodic directions. In keeping with this tradition, we expect its main application to be to tensor-product domains whose boundaries consist of a small number of piecewise-smooth surfaces, such as the finite three-dimensional cylinder which is our eventual goal. We also mention here some other recent approaches to solving the magnetohydrodynamic equations [@Matsui04; @Iskakov04; @Iskakov05; @Xu04; @Guermond06], with a view to generalizing the geometry and/or increasing parallelization. In [@Iskakov04; @Iskakov05], a finite volume method is used to discretize the solution in the interior, which is matched to that in the exterior vacuum via a boundary element method. [@Xu04] describes an integral equation formulation for the entire domain, and [@Guermond06] uses finite elements with a penalty method to apply boundary conditions. Influence matrix formulation {#sec:Influence matrix formulation} ============================ We formulate a two-stage method for solving –, consisting of an initial preprocessing step which depends only on the geometry, followed by a step whereby solutions for many different distributions $\rho$ can be generated at little incremental cost. This is the usual description of the decomposition of $\Phi$ into homogeneous and particular solutions, with the additional proviso that solutions in the exterior domain are to be taken into account in the preparation of the homogeneous solutions. We will construct the homogeneous solutions by generating harmonic bases $\{\Phi_j^h\}$ and $\{\phi_j\}$ of interior and exterior solutions, corresponding to Dirichlet boundary data $\{f_j\}$ to be specified later. We decompose – into the Poisson and Laplace problems: $$\begin{array}{lll} {\Delta}\Phi^p=\rho \;\;\text{in } \Omega\qquad & \qquad {\Delta}\Phi_j^h=0 \;\;\text{in } \Omega\qquad & \qquad {\Delta}\phi_j = 0 \;\;\text{outside }\Omega\\ \Phi^p|_{{\partial\Omega}}=0 \qquad & \qquad \Phi_j^h|_{{\partial\Omega}}= f_j \qquad & \qquad \phi_j|_{{\partial\Omega}}= f_j\\ &&\qquad\nabla\phi_j|_\infty = 0 \end{array} \label{eq:construct}$$ and then construct the linear superpositions: $$\begin{array}{ccc} \Phi=\Phi^p+\Phi^h \qquad & \qquad \Phi^h =\sum_j c_j\Phi_j^h \qquad & \qquad \phi=\sum_j c_j\phi_j \end{array} \label{eq:super}$$ Then $$\begin{aligned} {\Delta}\Phi = &{\Delta}\Phi^p+\sum_j c_j{\Delta}\Phi_j^h = \rho &\qquad\text{ in } \Omega\\ {\Delta}\phi = &\sum_j c_j{\Delta}\phi_j = 0 &\qquad\text{ outside }\Omega\\ \left.(\Phi-\phi)\right|_{{\partial\Omega}}= &\Phi^p|_{{\partial\Omega}}+ \sum_j c_j \left.(\Phi^h_j-\phi_j)\right|_{{\partial\Omega}}= 0 & \\ \nabla\phi|_\infty =& \sum_j c_j \:\nabla\phi_j|_\infty = 0 &\end{aligned}$$ are already satisfied by construction, while $$\left.{\partial_n}(\Phi-\phi)\right|_{{\partial\Omega}}= \left.{\partial_n}{\Phi^p}\right|_{{\partial\Omega}}+ \sum_j c_j\left.{\partial_n}(\Phi^h_j-\phi_j)\right|_{{\partial\Omega}}= 0 \label{eq:matchN2}$$ constitutes a system of equations to be solved for ${c_j}$, where the derivative with respect to the normal is taken in the direction from the interior to the exterior region for both $\Phi$ and $\phi$. $\Phi$ is then set equal to the sum in . If the interior harmonic functions are not stored, $\Phi$ can be obtained by solving: $$\begin{aligned} {\Delta}\Phi &=& \rho \;\;\text{in } \Omega \\ \Phi|_{{\partial\Omega}}&=& \sum_j c_j f_j\end{aligned}$$ Using ${{\mathbf x}}_i$ to index points on the boundary, can be discretized as: $$\sum_j \left[{\partial_n}(\phi_j-\Phi^h_j)({{\mathbf x}}_i) \right]c_j = \left.{\partial_n}{\Phi^p}\right({{\mathbf x}}_i) \label{eq:matchN3}$$ Equation shows that the goal of the preprocessing step is the construction and inversion of the influence or capacitance matrix: $$C_{ij} \equiv \left[{\partial_n}\left(\phi_j - \Phi^h_j\right)\left({{{\mathbf x}}_i}\right) \right] \label{eq:def_influ_matrix}$$ The functions $\{f_j\}$ are required to constitute a complete set for values along the discretized boundary ${{\partial\Omega}}$. Another way to describe the influence matrix is as a discrete representation of the difference between the Dirichlet-to-Neumann mappings in the exterior and in the interior regions. Equivalently, ${\partial_n}(\phi_j-\Phi^h_j)|_{{\partial\Omega}}$ can be represented as coefficients of a basis set $\{g_i\}$ (which may be identical with the set of boundary value functions $\{f_i\}$) along each boundary. Equation is then discretized as: $$\sum_j \langle{\partial_n}(\phi_j-\Phi^h_j),g_i\rangle \: c_j = \langle {\partial_n}{\Phi^p},g_i \rangle \label{eq:matchN4}$$ Although we will use ${{\mathbf x}}_i$ and the notation in in what follows, the method is easily reformulated using . Solution of Poisson and Laplace problems ======================================== Interior domain {#sec:Interior domain} --------------- We now turn to the solution of . For the interior problems listed in the first two columns, we assume that we dispose of a solver able to compute solutions to Poisson’s equation in $\Omega$ with any specified boundary values. In principle, any numerical method can be used. In our particular case, we use a spectral discretization [@Canuto88] $$\Phi(x,y) = \sum_{k,l=0}^{K,L}{\mathcal{T}}_k(x/H){\mathcal{T}}_l(y) \label{eq:ChebCheb}$$ for the rectangle $[-H,H]\times[-1,1]$. The spectral basis functions are the Chebyshev polynomials ${\mathcal{T}}_k(x)=\cos(k\arccos(x))$. Taking $H\geq 1$, we set $K\geq L$. We use a standard method [@Canuto88] to solve the Poisson equation with Dirichlet boundary conditions, diagonalizing the discretized second derivative operator in $y$, and using recursion relations to treat the second derivative in $x$. Exterior harmonic functions {#sec:Exterior harmonic functions} --------------------------- Our main focus is on the construction of the exterior harmonic solutions, specified in the third column of . In order to avoid truncating or spatially discretizing the exterior domain, we will construct $\{\phi_j\}$ using the fundamental solution of the Laplace equation: the Green’s function satisfying $$\begin{aligned} {\Delta}_{{\mathbf x'}}G({{\mathbf x}}; {{\mathbf x'}}) =& \delta({{\mathbf x}}-{{\mathbf x'}}) & \\ \nabla G({{\mathbf x}}; {{\mathbf x'}}) =& 0 & \qquad\text{for } {{\mathbf x'}}\rightarrow \infty \label{eq:fundamental_Green}\end{aligned}$$ For a specified boundary value distribution $f({{\mathbf x}})$, we first calculate an appropriate source distribution $\sigma({{\mathbf x}})$ on the boundary by solving the integral equation: $$\int_{{\partial\Omega}}G({{\mathbf x}}; {{\mathbf x'}}) \sigma ({{\mathbf x'}}) = f({{\mathbf x}}) \qquad\text{for }{{\mathbf x}}\in{{\partial\Omega}}\label{eq:chargedist}$$ The exterior harmonic function $\phi({{\mathbf x}})$ required is then: $$\phi({{\mathbf x}}) \equiv \int_{{\partial\Omega}}G({{\mathbf x}}; {{\mathbf x'}}) \sigma ({{\mathbf x'}}) \label{eq:phiconstruct}$$ where ${{\mathbf x}}$ takes values either on or off ${{\partial\Omega}}$. We now apply - to our particular test problem of a rectangle. We divide the set of boundary distributions into four sets, each taking non-zero values on only one side of the rectangle. In two dimensions, the fundamental Green’s function solving is $$\frac{-1\;}{2\pi}\ln|{{\mathbf x}}-{{\mathbf x'}}|$$ Equation thus reduces to: $$\int_a^b \frac{-1\;}{2\pi}\ln|x-{x^\prime}|\sigma({x^\prime})\,d{x^\prime}= f(x) \label{eq:Carleman_equation}$$ Equation is known as Symm’s or Carleman’s equation and has the following solution [@Jorgens70; @Polyanin98]: $$\begin{aligned} \sigma(x)=\frac{-2}{\pi\sqrt{(x-a)(b-x)}} \left[\int_a^b\frac{\sqrt{({x^\prime}-a)(b-{x^\prime})}f'({x^\prime})\,d{x^\prime}}{{x^\prime}-x}\right. &&\nonumber\\ +\frac{1}{\ln ((b-a)/4)}\left.\int_a^b\frac{f({x^\prime})\,d{x^\prime}}{\sqrt{({x^\prime}-a)(b-{x^\prime})}}\right]&& \label{eq:Carleman_sol}\end{aligned}$$ if $b-a\neq 4$. (If $b-a=4$ then the second integral in can be replaced by an arbitrary constant if $\int _a^b f(t)\left[(t-a)(b-t)\right]^{-\frac{1}{2}}\,dt=0$.) Up to now, we have not specified the Dirichlet boundary values $f$. The choice of boundary value distributions is restricted only by the requirement, stated in section \[sec:Influence matrix formulation\], that the set of distributions form a basis for functions defined on the boundary ${{\partial\Omega}}$. Because we use Chebyshev polynomials to represent the interior solutions, it is convenient to take as boundary values $f_k(x)$ each of the functions ${\mathcal{T}}_k(x/H)$ on the interval $[-H,H]$. The corresponding solutions $\sigma_k(x;H)$ obtained from evaluating are: $$\sigma_k(x;H)= A_k\frac{{\mathcal{T}}_k(x/H)} {\pi H\sqrt{1-\left(\frac{x}{H}\right)^2}} \quad;\quad A_k=\left\{\begin{array}{ll} 2\pi k\quad & k>0\\ -2\pi\left[\ln(H/2)\right]^{-1}\quad & k=0 \end{array}\right. \label{eq:Carleman_sol_Cheb}$$ This remarkable property – the fact that that weighted Chebyshev polynomials are also obtained as the source distributions $\sigma_k(x)$ when the boundary values $f_k(x)$ are Chebyshev polynomials – is related to the very reason that Chebyshev polynomials are optimal in approximating polynomials on the interval. The function $1/\pi\sqrt{1-x^2}$ in (for $H=1$) is the weight with respect to which Chebyshev polynomials are orthogonal on the interval and is the asymptotic density of the Chebyshev interpolation points $\cos(\pi j/J)$, the extrema of the Chebyshev polynomials. See [@Mason00; @Trefethen00] for further details. Note also that the orthogonality of the Chebyshev polynomials with respect to this weight causes the second integral in to vanish except for ${\mathcal{T}}_0$. The corresponding harmonic functions $\phi^x_k$ are constructed via $$\phi^x_k({{\mathbf x}})=\int_{-H}^{H} \frac{-1\;}{2\pi}\ln\,|{{\mathbf x}}-{x^\prime}{\hat{\mathbf e}_x}|\:\sigma_k({x^\prime};H)\,d{x^\prime}\label{eq:reconvolvex}$$ and are illustrated in figure \[fig:harmonic\_basis\]. Specifying values along along the segment $[-1,1]$ in the $y$ direction, we obtain: $$\phi^y_l({{\mathbf x}})=\int_{-1}^{1} \frac{-1\;}{2\pi}\ln\,|{{\mathbf x}}-y'{\hat{\mathbf e}_y}|\:\sigma_l(y';1)\,dy' \label{eq:reconvolvey}$$ Note that $\sigma_k(x;H)=\sigma_k(x/H;1)/H$ for $k>0$ and $\sigma_0(x;H)=\frac{\ln(1/2)}{H\ln(H/2)}\sigma_0(x/H;1)$. The harmonic functions corresponding to specified values along the lower or upper boundaries ($y=\pm 1$, $|x|< H$) are $\phi^x_k\left({{\mathbf x}}\mp 1{\hat{\mathbf e}_y}\right)$; those corresponding to the left or right boundaries ($x=\pm H$, $|y|<1$) are $\phi^y_l\left({{\mathbf x}}\mp H{\hat{\mathbf e}_x}\right)$. We do not require the functions $\phi^x_k$, $\phi^y_l$ either inside or outside $\Omega$, but only the values and normal derivatives on the boundary. Although, for example, the values of $\phi^x_k\left({{\mathbf x}}+ 1{\hat{\mathbf e}_y}\right)$ on the lower boundary are merely the specified values $f_k(x)$, its values on the other three boundaries must be calculated via . When evaluating the normal derivatives, the kernel $G({{\mathbf x}};x')$ is differentiated before integration: $$\begin{aligned} {\partial}_y\phi^x_k(x,y)&=&\partial_y\int_{-H}^{H} \frac{-1\;}{2\pi}\ln\,\sqrt{(x-x')^2+y^2}\;\sigma_k(x';H)\,dx'\nonumber\\ &=&\int_{-H}^{H}\frac{-1\;}{2\pi}\frac{y}{(x-x')^2+y^2}\;\sigma_k(x';H)\,dx' \label{eq:normalder}\end{aligned}$$ and similarly for ${\partial}_x\phi^y_l$. Any exterior harmonic function can be approximated by the truncated series: $$\begin{aligned} \phi^{K,L}({{\mathbf x}})= \sum_{k=0}^{K-1}\left[c^{x,-}_k\phi^x_k\left({{\mathbf x}}+H{\hat{\mathbf e}_x}\right) + c^{x,+}_k\phi^x_k\left({{\mathbf x}}-H{\hat{\mathbf e}_x}\right)\right] \nonumber\\ + \sum_{l=0}^{L-1}\left[c^{y,-}_l\phi^y_l\left({{\mathbf x}}+1{\hat{\mathbf e}_y}\right) + c^{y,+}_l\phi^y_l\left({{\mathbf x}}-1{\hat{\mathbf e}_y}\right)\right] \label{eq:truncser}\end{aligned}$$ The potential $\phi({{\mathbf x}})$ of is defined by the $2(K+L)$ coefficients $\{c^{x,-}_k,c^{x,+}_k,c^{y,-}_l,c^{y,+}_l\}$. A very important property of the harmonic basis $\{\phi^x_k({{\mathbf x}}),\phi^y_l({{\mathbf x}})\}$ is that it represents a near-boundary field uniformly. This means that the truncated series converges uniformly for any smooth boundary data and for all locations ${{\mathbf x}}$ near the boundary. This is a direct consequence of the excellent convergence properties of Chebyshev approximation applied to ; a proof can be found in [@Levesley]. This property does not necessarily hold for other harmonic bases, in particular spherical harmonics, for which near-boundary convergence cannot be achieved, leading to a strong Gibbs effect. In the taxonomy of boundary integral methods, equations and constitute an indirect method, in that the intermediate surface charge density $\sigma$ is constructed; this is done by solving the Fredholm integral equation of the first kind . The surface charge density is a single-layer rather than a double-layer (dipole) potential; equivalently $G$, rather than ${\partial}G/{\partial}n$, is used in the representation. Because the method determines only the boundary values and normal derivatives of the exterior solution, it is not the preferred approach when the exterior potential is itself required at each time step: although the exterior solution can be sampled at any location, this is computationally expensive, as is often the case for boundary integral methods. ![Potentials $\phi^x_k({{\mathbf x}})$ generated by line source distributions $\sigma_k(x;H=2)$. Values of the potentials on the line segment $x\in[-1,1]$ correspond to the Chebyshev polynomials ${\mathcal{T}}_k(x)$. From top to bottom: $\sigma_0(x)$, $\sigma_1(x)$, $\sigma_4(x)$.[]{data-label="fig:harmonic_basis"}](line_mono.eps){width="8cm"} ![Potentials $\phi^x_k({{\mathbf x}})$ generated by line source distributions $\sigma_k(x;H=2)$. Values of the potentials on the line segment $x\in[-1,1]$ correspond to the Chebyshev polynomials ${\mathcal{T}}_k(x)$. From top to bottom: $\sigma_0(x)$, $\sigma_1(x)$, $\sigma_4(x)$.[]{data-label="fig:harmonic_basis"}](line_dipol.eps){width="8cm"} ![Potentials $\phi^x_k({{\mathbf x}})$ generated by line source distributions $\sigma_k(x;H=2)$. Values of the potentials on the line segment $x\in[-1,1]$ correspond to the Chebyshev polynomials ${\mathcal{T}}_k(x)$. From top to bottom: $\sigma_0(x)$, $\sigma_1(x)$, $\sigma_4(x)$.[]{data-label="fig:harmonic_basis"}](line_quad.eps){width="8cm"} Electrostatic example ===================== We apply our method to a simple problem from electrostatics, the distribution of electric charges $-\rho_m$ confined in a rectangular domain but localized around the origin: $$\rho_m(r,\theta)=\left\{\begin{array}{ll} r^me^{-r^2/\delta^2}\cos(m\theta) &\text{for }|x|\leq 1\text{ and }|y|\leq 1 \\ 0 &\text{for } |x|>1 \text{ or } |y|>1 \end{array}\right. \label{eq:src_dist}$$ The $r^m$ factor in (\[eq:src\_dist\]) ensures regularity of $\rho_m$ at $r=0$. The potential $\tilde{\Phi}_m$ due to [*unbounded*]{} sources (not restricted to the interior domain) can be found analytically: $$\begin{aligned} \tilde{\Phi}_{m=0}(r,\theta) &=& \frac{\delta^2}{4}\left[Ei\left(1,\frac{r^2}{\delta^2}\right)+2\log(r)\right]\label{eq:pot_sols_a} \\ \tilde{\Phi}_{m=1}(r,\theta) &=& \frac{\delta^4}{4r}\left[e^{-\frac{r^2}{\delta^2}}-1\right]\cos\theta \label{eq:pot_sols_b}\\ \tilde{\Phi}_{m=2}(r,\theta) &=& \frac{\delta^4}{4r^2}\left[\left(\delta^2+r^2\right)e^{-\frac{r^2}{\delta^2}}-\delta^2\right]\cos2\theta \label{eq:pot_sols_c}\end{aligned}$$ \[eq:pot\_sols\] \ where $Ei(a,\lambda)$ is the error function $Ei(a,\lambda)\equiv \int_1^\infty{e^{-\lambda r}r^{-a}\,dr}$ and $\tilde{\Phi}_m$ are chosen to be finite at $r=0$. We seek the corresponding electric potential. The parameter $\delta$ is chosen to make $\rho_m$ very small near the boundaries. We expect the solution to be almost unaffected by the presence of boundaries. The source distribution $\rho_{m=0}(r,\theta)$ should therefore lead to a potential which is almost axisymmetric. Figure \[fig:axi\_sol\] shows the potential obtained numerically for $\delta^2=0.15$ using the spectral resolution $N=8$ in both directions. The domain boundary is represented by a bold square. The contours are almost perfectly circular, as should be the case for $\delta$ small, showing that the presence of the boundaries has minimal effect. \[l\]\[\]\[1.0\] \[l\]\[\]\[0.9\][$\log_{10}E_{0}(N)$]{} ![ Convergence test: Convergence test: $\log_{10}E_{m=0}(N)$ plotted for $N=[6,\ldots,16]$, $\delta^2=0.1$.[]{data-label="fig:conv"}](monopol8.eps){width="6.5cm"} ![ Convergence test: Convergence test: $\log_{10}E_{m=0}(N)$ plotted for $N=[6,\ldots,16]$, $\delta^2=0.1$.[]{data-label="fig:conv"}](convergence.eps){width="6.9cm"} To evaluate the error convergence of the method we computed the relative error $E_m(N)$ defined as $$E_m(N)=\sup_{r,\theta}\frac{|\tilde{\Phi}_m(r,\theta)-\Phi^N_m(r,\theta)|}{|\Phi_m(r,\theta)|} \label{eq:error}$$ where $\Phi^N_m(r,\theta)$ is the solution computed numerically with spectral resolution $N$ in both spatial directions of the bounding square and $\tilde{\Phi}_m(r,\theta)$ is the analytic solution in the absence of the bounding square. Figure \[fig:conv\] proves the exponential convergence of the method. Figures \[fig:dipol\]–\[fig:quad\] show the electric potentials $\Phi^{N=16}_{m=1}$ and $\Phi^{N=16}_{m=2}$ for $\delta=0.1$. Figure \[fig:dipole\] shows $\Phi_{m=1}^{N=16}$ with $\delta=2$. In figure \[fig:turned\_dipole\], the dipole source distribution has been rotated by $45^\circ$ about the origin. For this large value of $\delta$, charges are located near the boundary. In each case with $\delta \ll 1$, we observed exponential convergence toward solution . Convergence can only be confirmed up to a limited precision since the analytic solution does not correspond exactly to the problem we are solving numerically, in which sources are confined to the interior square. The best agreement can be achieved for small values of $\delta$. If the numerical solution with highest spectral resolution (here $N=64$) is instead taken as a reference, then the method converges to this solution spectrally up to machine precision for any value of $\delta$. ![$\Phi_{m=1}^{N=16}(r,\theta+\frac{\pi}{4})$ for $\delta=2$.[]{data-label="fig:turned_dipole"}](dipol16.eps){width="6.5cm"} ![$\Phi_{m=1}^{N=16}(r,\theta+\frac{\pi}{4})$ for $\delta=2$.[]{data-label="fig:turned_dipole"}](quad32.eps){width="6.3cm"} ![$\Phi_{m=1}^{N=16}(r,\theta+\frac{\pi}{4})$ for $\delta=2$.[]{data-label="fig:turned_dipole"}](dipol16a.eps){width="6.5cm"} ![$\Phi_{m=1}^{N=16}(r,\theta+\frac{\pi}{4})$ for $\delta=2$.[]{data-label="fig:turned_dipole"}](dipol16b.eps){width="6.5cm"} Implementation ============== Summary and computation cost ---------------------------- We describe the implementation of the method for our illustrative example of the rectangle $[-H,H]\times [-1,1]$ with double Chebyshev discretization $(K+1)\times (L+1)$. The total preprocessing step consists of:\ $\bullet$ Evaluation of the values and the normal derivatives of the exterior harmonic solutions on the boundary.\ $\bullet$ Calculation of the interior harmonic solutions.\ $\bullet$ Inversion or LU decomposition of the influence matrix, For each particular right-hand-side $\rho$, the operations consist of:\ $\bullet$ Solving for a single particular solution.\ $\bullet$ Acting with the inverse of the influence matrix.\ $\bullet$ Using the corrected Dirichlet boundary conditions to calculate the final solution. The total number $J$ of boundary points is $2(K+L)$. The inversion or LU-decomposition of the influence matrix $C$ in the preprocessing stage requires a time proportional to $J^3$, while each solution of the linear system determining the coefficients of the homogeneous solutions requires a time proportional to $J^2$. Each interior solution is calculated at a cost proportional to $KL^2$. Symmetry can be used to reduce the cost of each step. The symmetry of the rectangle divides all the independent harmonic solutions into four mutually orthogonal classes. Decoupling the Laplacian operator according to parity in $x$ and $y$ leads to four Poisson problems, each with resolution $K/2 \times L/2$, thus reducing the time by a factor of two. Decoupling by parity also reduces the influence matrix $C$ to four matrices, the dimensions of which are one fourth of that of the original matrix. Table \[tab:cost\] gives the operation count of each step, taking into account the reductions permitted by symmetry. [|l|ll|]{} Calculation & Result & Cost\ \ Exterior harmonic solutions & $\phi_j|_{{\partial\Omega}}$, ${\partial_n}\phi_j|_{{\partial\Omega}}$ & $\begin{array}{l} K\times(\mbox{Int}_x + \mbox{Int}^\prime_x)\\ +L\times(\mbox{Int}_y + \mbox{Int}^\prime_y) \end{array}$\ Interior harmonic solutions & ${\partial_n}\Phi^h_j$ & $(K+L)KL^2/4$\ Influence matrix inversion/decomposition & $C^{-1}$ & $(K+L)^3/2$\ \ Particular solution & $\Phi^p$ & $KL^2/2$\ Action with influence matrix & $c_j$ & $(K+L)^2$\ Corrected solution & $\Phi$ & $KL^2/2$\ \[tab:cost\] Singular integrals ------------------ The integrations in - are performed numerically. Special attention must be paid in doing so since both the kernel $G({{\mathbf x}};x')$ and the density $\sigma(x')$ have integrable singularities within the domain of integration. The singular points are $x'{\hat{\mathbf e}_x}={{\mathbf x}}$ and $x'=\pm H$ for $\phi^x_k$ and $y'{\hat{\mathbf e}_y}={{\mathbf x}}$ and $y'=\pm 1$ for $\phi^y_l$. Dedicated adaptive quadratures (see [@Press86]) can be used to compute these integrals accurately. It is also possible to evaluate the singular part of the integral analytically, reducing the numerical problem to the evaluation of integrals with non-singular integrands. The remaining integrand is piecewise $C^\infty$ and can be integrated with spectral precision over each of the regular subdomains. Singularity subtraction greatly decreases the variation in grid density needed to sample the integrand homogeneously, thereby significantly accelerating the numerical integrations in -. Specifically, an adaptive method requires a smaller number of iterations, or, alternatively, a non-adaptive method requires a coarser resolution. However, the convergence of our approximation with $K,L$ is exponential (spectral), regardless of whether the singular part of the integral is subtracted or included in the numerical evaluation. Conditioning of matrices ------------------------ \[r\]\[\]\[0.9\][$3.58N^2-18.16N+59$]{} \[r\]\[\]\[0.9\][$\mathcal{C}(N)$]{} ![Quadratic fit of the condition number $\mathcal{C}(N)$ of the influence matrix defined in . ](cond.eps "fig:"){width="8cm"} \[fig:cond\_fit\] The influence matrix is not immediately invertible. Because of the redundancy of information at the corners, this matrix has exactly four zero singular values or eigenvalues. The corresponding linear system can be solved after arbitrarily correcting singular values or eigenvalues of the influence matrix; see [@Tuckerman89; @Boronski05] for more details. The condition number $\mathcal{C}$ of the corrected matrix depends on the spatial resolution $N$ and the maximal order of derivatives used to express the boundary conditions. In our case of a Neumann boundary condition and a resolution $N$ in each direction, the condition number scales as $\mathcal{C}=O\left(N^2\right)$. Fitting the condition numbers computed for $N\in[2,32]$ with a parabola (see fig. \[fig:cond\_fit\]) yields a formula for predicting the condition number for an arbitrary resolution: $$\mathcal{C}(N)=3.58N^2-18.16N+59 \label{eq:cond_number}$$ It can then be deduced from that a reasonably conditioned matrix with $\mathcal{C}<10^7$ is obtained for a spatial resolution as high as $O(1000)$. The method can therefore be applied to problems where small-scale field features require use of high spatial resolution. Generalizations =============== We now discuss the applicability of this method to other geometries, problems, and spatial discretizations. The decomposition into interior particular and homogeneous functions and exterior homogeneous functions described in section \[sec:Influence matrix formulation\] is, of course, completely general and not related to any particular spatial discretization. The method described in section \[sec:Interior domain\] for constructing the exterior harmonic functions relies on the Chebyshev-Chebyshev discretization of the rectangle which is widely used since the Chebyshev polynomials are optimal approximants of smooth functions. This property, as well as the straightfoward correspondence between interior and exterior solutions, make the Chebyshev discretization especially suitable for the construction of the exterior harmonic solutions as well. However, the method is easily generalizable to other basis functions $f_k(x)$ for the potential values, which can be substituted into in order to calculate the corresponding charge densities $\sigma_k(x)$, if dictated by the geometry or numerical method used for the interior problem. In fact, since our real interest is in generating the complete set of $\sigma_k(x)$ necessary to generate the complete set of $\phi_k(x)$, rather than in calculating the specific $\sigma_k(x)$ corresponding to each particular $f_k(x)$, the only information really required in is the singularity $1/\pi\sqrt{1-x^2}$. One may then allow the set of $\sigma_k$’s to be the products of this singularity with the members of any appropriate basis set of analytic functions on the boundary in question. In three dimensions, the fundamental Green’s function is $$G({{\mathbf x}},{{\mathbf x'}}) = \frac{1\;}{4\pi}\frac{1}{|{{\mathbf x}}-{{\mathbf x'}}|}$$ In an axisymmetric geometry with a Fourier representation of the azimuthal direction, all of the problems to be solved decouple according to Fourier mode. The operation count would then scale linearly with the number of azimuthal points or Fourier modes. The elliptic problems in would remain two-dimensional, and the integral equations equivalent to would remain one-dimensional. This method can also be applied to other elliptic problems or to parabolic problems. As stated in section \[sec:motivation\], our motivation for developing this method is to apply it to the magnetohydrodynamic equations -, in which is a parabolic equation. A general parabolic problem may be written as: $${\partial}_t \Phi = {\Delta}\Phi + \mathcal{F}(\Phi,\rho) \label{eq:parabolic}$$ where $\mathcal{F}$ may include nonlinear and/or time-dependent source terms. First-order implicit temporal discretization of results in the inhomogeneous Helmholtz equation: $$(I-\delta t {\Delta}) \Phi(t+\delta t) = \mathcal{F} \label{eq:implicit}$$ where $\mathcal{F}$ may depend on previous values of $\Phi$. This Helmholtz operator $(I-\delta t {\Delta})$ can replace the Laplacian in and . It is known that replacement of the Helmholtz equation by a boundary integral equation can lead to singularities for certain values of the wavenumber (here $i/\sqrt{\delta t}$); a large body of work, e.g. [@Amini95; @Givoli98; @Gerdes96; @Hwang99; @TCLin04], addresses this problem. However, in the magnetohydrodynamic case of a conducting fluid surrounded by an exterior vacuum, no such difficulties would be introduced, since the exterior problem remains governed by Laplace’s equation. More complicated vectorial operators may appear, as occur in the Navier-Stokes or magnetohydrodynamic equations. Comparison with other approaches ================================ We mention here some other techniques that have been used to solve exterior problems or to match interior and exterior domains. Spectral methods can be combined with various transformations and mappings. The inner region can be surrounded by a sphere, and the outer domain decomposed into the region inside and outside the sphere. The exterior domain can be mapped into an interior domain via a $1/r$ mapping [@Jepps67; @Grandclement01; @Lai06]; spectral methods can then be used to treat either or both domains. The region exterior to one or more spheres has been mapped to the interior of a rectangle [@Ansorg04] or a pentangular [@Ansorg05] region rotated about an axis, and Chebyshev-Fourier expansions used to solve elliptic equations arising in the study of black holes in general relativity. A smooth boundary can be parameterized by angle, and the boundary values represented as a series of trigonometric functions or spherical harmonics [@Atkinson82; @Gerdes96; @Hwang99; @TCLin04; @Grandclement01; @Lai06; @Meddahi02; @Chen94; @Ganesh98]. Our method differs from these in that a Chebyshev approximation is used to represent the boundary values on each segment of a non-smooth boundary, and an analytic formula is used to calculate the surface density which exactly yields this Chebyshev approximation. Conformal mapping is another technique which can be used to calculate interior or exterior harmonic functions. The Riemann mapping theorem guarantees the existence of a conformal transformation from the interior or exterior of a simply connected domain into the interior or exterior of a unit disk; its proof is, however, non-constructive, and does not explicitly derive the transformation. For some geometries, including the exterior of a rectangle, an analytical formula can be derived [@Ivanov]. For polygon-bounded regions with piecewise-constant boundary conditions, the Schwarz-Christoffel [@TrefethenSC] mapping has proved to be a very robust tool, applied to problems arising in magneto- and electro- statics, potential flows, inverse problems and many other fields. Our influence matrix approach relies on calculating harmonic functions with arbitrary boundary data, for which conformal mapping is much more problematic. More general conformal mappings are often computed by solving Symm’s or Carleman’s equation numerically on the domain boundary, making this approach similar in terms of numerical cost and precision to boundary integral equation methods. It is interesting to note that, for domains including corners, the Chebyshev approximation is especially well suited, guaranteeing superconvergence of the mapping function [@Levesley]. Conclusion ========== As a test case for the magnetohydrodynamic equations, we have developed a method for solving the two-dimensional Poisson equation in a bounded domain, where the solution satisfies matching conditions with a harmonic potential outside the domain. The method solves only the interior problem and determines the boundary conditions ensuring smooth matching with the exterior solution. The essential element of this approach is construction of a basis of harmonic functions which represent the near-boundary exterior solutions uniformly. This basis is used to construct the influence matrix which serves to impose the coupled boundary conditions between the interior and exterior solutions. The method is numerically reasonably well conditioned and can be used for quite high spatial resolutions. For a spectral solver, this method guarantees exponential convergence. Instead of corresponding to point sources on the boundary, each exterior harmonic solution corresponds to a spectral basis function. The most costly process – the construction of a basis of exterior harmonic functions – depends only on geometry and spatial resolution. Once the basis is computed it can be stored and used for any computation using the same resolution and domain shape. When used as a preprocessing step for time-dependent simulations, the cost of constructing the exterior harmonic basis is negligible compared to that of thousands of time steps. Since cylindrical coordinates have one periodic direction, it should be possible to apply this method separately to each of the Fourier modes, treated individually as two-dimensional problems. The extension of this method to the MHD equations in a finite cylindrical geometry is currently under investigation. [99]{} , Experimental demonstration of a homogenous two-scale dynamo, Magnetohydrodynamics [**38**]{}, 27 (2002). Magnetic field saturation in the Riga dynamo experiment, Phys. Rev. Lett. [**86**]{}, 3024 (2001). Magnetohydrodynamics measurements in the von Kármán sodium experiment, Phys. Fluids [**14**]{}, 3046 (2002). Hydrodynamic and numerical modeling of a spherical homogeneous dynamo experiment, Magnetohydrodynamics [**38**]{}, 107 (2002). Mechanically forced and thermally driven flows in liquid sodium, Magnetohydrodynamics [**38**]{}, 121 (2002). , Numerical models of hydromagnetic dynamos, J. Fluid Mech. [**67**]{}, 625 (1967). , Time-dependent kinematic dynamos with stationary flows, Proc. Roy. Soc. London A [**425**]{}, 407–429 (1989). A three-dimensional self-consistent computer simulation of a geomagnetic field reversal, Nature [**337**]{}, 203 (1995). , A kinematic dynamo with a small scale velocity field, Phys. Rev. [**A 226**]{}, 75–79 (1997). , A spectral solution of the magneto-convection equations in spherical geometry, Int. J. for Num. Meth. in Fluids [**32**]{}, 773–797 (2000). A Taylor-Couette Dynamo, Astronomy & Astrophysics [**393**]{}, 339–343 (2002). , Development of a simulation code for MHD dynamo processes using the GeoFEM platform, Intl. J. of Comp. Fluid Dyn. [**18**]{}, 323–332 (2004). , An integro-differential formulation for magnetic induction in bounded domains: boundary element-finite volume method, J. Comput. Phys. [**7**]{}, 540–554 (2004). , On magnetic boundary conditions for non-spectral dynamo simulations, Geophys. Astrophys. Fluid Dyn. [**99**]{}, 481–492 (2005). , The integral equation method for a steady kinematic dynamo problem, J. Comput. Phys. [**196**]{}, 102–125 (2004). , An interior penalty Galerkin method for the MHD equations in heterogeneous domains, J. Comput. Phys. [**221**]{}, 349–369 (2007). , An Introduction to Magnetohydrodynamics, Longmans Green, London, 1967. Solution of Boundary Value Problems by Integral Equations of the First Kind SIAM Review [**15**]{}, 687–705 (1973). , Integral equation methods in potential theory and elastostatics, Academic Press, 1977. , The boundary element method for engineers, Pentech Press, 1978. , Boundary element methods in engineering science, McGraw-Hill, 1981. , The numerical solution of Laplace’s equation in three dimensions, SIAM J. Num. Anal. [**19**]{}, 263–274 (82). , Rapid solution of integral equations of classical potential theory, J. Comput. Phys. [**60**]{}, 187–207 (1983). , Boundary Elements: Proceedings of the 5th International Conference, Springer, 1983. , Boundary integral and singularity methods for linearized viscous flow, Cambridge Univ. Press, 1992. , [http://www.boundary-element-method.com]{} , Numerical simulation of incompressible flows within simple boundaries: I. Galerkin (spectral) representations, Stud. Appl. Math. [**50**]{}, 293–327 (1971). , Spectral Methods in Fluid Dynamics, Springer, 1988. , Integral operators, Pitman, Boston, 1982. Translation from German of Lineare Integraloperatoren, Teubner, Stuttgart, 1970. , Handbook of Integral Equations, CRC Press, Boca Raton, 1998. , Chebyshev Polynomials, Chapman & Hall/CRC Press, 2000. , Spectral Methods in Matlab (SIAM, Philadelphia, 2000). , A Chebyshev collocation method for solving Symm’s integral equation for conformal mapping: a partial error analysis, IMA J. Num. Anal. [**14**]{}, 57–79 (1993). , Numerical Recipes: The Art of Scientific Computing, Cambridge Univ. Press, 1986. , Divergence-free velocity fields in nonperiodic geometries, J. Comput. Phys. [**80**]{}, 403–441 (1989). , A Method Based on Poloidal-Toroidal Potentials Applied to the von Kármán Flow in a Finite Cylinder Geometry, Ph.D. Thesis, Ecole Polytechnique, 2005. , Solution of Helmholtz equation in the exterior domain by elementary boundary integral methods, J. Comput. Phys. [**118**]{}, 208–221 (1995). , eds., Special Issue on Exterior Problems of Wave Propagation, Comput. Methods Appl. Mech. Engrg. [**164**]{}, 1-266 (1998). , Solution of 3D-Laplace and Helmholtz equations in exterior domains using hp-infinite elements, Comput. Methods Appl. Mech. Engrg. [**137**]{}, 239–272 (1996). , A boundary spectral method for solving exterior acoustical problems with hypersingular integrals, Int. J. Numer. Meth. Engng. [**44**]{}, 1775–1783 (1999). , The numerical solution of exterior Neumann problem for Helmholtz’s equation via modified Green’s functions approach, Computers and Mathematics with Applications [**47**]{}, 593–609 (2004). , A multidomain spectral method for scalar and vectorial Poisson equations with noncompact sources, J. Comput. Phys. [**170**]{}, 231–260 (2001). , Fast solvers for 3D Poisson equations involving interfaces in a finite or the infinite domain, J. Comput. Appl. Math. [**181**]{}, 106–125 (2006). , A single-domain spectral method for black hole puncture data, Phys. Rev. D [**70**]{}, 064011 (2004). , Double-domain spectral method for black hole excision data, Phys. Rev. D [**72**]{}, 024018 (2005). , A combination of spectral and finite elements for an exterior problem on the plane, Appl. Numer. Math. [**43**]{}, 275–295 (2002). , Galerkin methods for solving single layer integral equations in three dimensions, Ph.D. Thesis, Univ. of Iowa, 1994. , A new spectral boundary integral collocation method for three-dimensional potential problems, SIAM J. Num. Anal. [**35**]{}, 778–805 (1998). , Handbook of Conformal Mapping with Computer-Aided Visualization, CRC Press, Boca Raton, 1995. , Schwarz-Christoffel Mapping, Cambridge Univ. Press, 2002.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We study odd-dimensional modular tensor categories and maximally non-self dual (MNSD) modular tensor categories of low rank. We give lower bounds for the ranks of modular tensor categories in terms of the rank of the adjoint subcategory and the order of the group of invertible objects. As an application of these results, we prove that MNSD modular tensor categories of ranks 13 and 15 are pointed. In addition, we show that MNSD tensor categories of ranks 17, 19, 21 and 23 are either pointed or perfect.' address: - 'Department of Mathematics, University of Oregon' - 'Department of Mathematics, Indiana University' author: - Agustina Czenky - Julia Plavnik title: 'On odd-dimensional Modular Tensor Categories' --- Introduction ============ Integral modular tensor categories have been deeply studied in the last decade, see for example [@BR], [@BGHKNNPR], [@DGNO1], [@DLD], [@DN], [@DT], [@EGO], [@ENO1], [@ENO2], [@NR]. One large class of examples is given by odd-dimensional modular tensor categories  [@GN]. Ng and Schauenburg proved that if the dimension of a modular tensor category is odd then the category is maximally non-self dual (MNSD), i.e. the only self-dual simple object is the unit object \[NS\]. In the other direction, Hong and Rowell showed in [@HR Theorem 2.2] that MNSD modular tensor categories are always integral, and as a consequence they must be odd-dimensional. In [@BR], the authors studied more in detail odd-dimensional modular tensor categories. In particular, they asked if odd-dimensional modular tensor categories are neccesarily group-theorerical. A negative answer to this question can be deduced from results of Larsen and Jordan in [@JL]. Roughly speaking, they proved that integral fusion categories of dimension $pq^2$, where $p$ and $q$ are distinct primes, need not be group-theoretical. Bruillard and Rowell showed in [@BR] that any MNSD modular tensor category of rank at most 11 is always pointed. Recall that a fusion category is called pointed if all its simple objects are invertible, which is equivalent to requiring that all its simple objects have Frobenius-Perron dimension equal to 1. They also found an example of a MNSD modular tensor category of rank $25$ that is not pointed (but it is group-theoretical). A natural follow-up question is if there exists a non-pointed MNSD modular tensor category of rank less than $25$ [@BR]. In this manuscript we continue the study of odd-dimensional modular tensor categories. We prove some useful relations between the ranks of the different components of a faithful grading of an odd-dimensional modular tensor category. Moreover, we give some bounds of the rank of the category in terms of the order of the group of invertibles, the rank of the adjoint subcategory, and certain prime dividing the order of the group of invertibles of the adjoint. We apply these general results to classify MNSD modular tensor categories of low rank. In particular, we give a partial answer to the question mentioned above: MNSD modular tensor categories of rank 13 and 15 are pointed, and MNSD modular tensor categories of rank at most 23 are either pointed or perfect. By a perfect fusion category, we mean a fusion category with trivial group of invertible objects. As a consequence of our results, a non-pointed MNSD modular tensor category of rank between 17 and 23 cannot be weakly group-theoretical. There is an important conjecture in fusion categories that states that every weakly integral fusion category is weakly group-theoretical [@ENO2 Question 2]. Hence the veracity of this conjecture would imply that MNSD modular tensor categories of ranks between 17 and 23 are pointed. Moreover, this leads us to the following conjecture: Odd-dimensional modular tensor categories have at least one non-trivial invertible object, i.e. they cannot be perfect. Notice that this is equivalent to: Odd-dimensional fusion categories are solvable. See Section \[Future directions\] for more details. The paper is organized as follows. In Section \[section: preliminaries\] we introduce the basic notions and we prove some results that we will use throughout this article. In Section \[section: cd\_p\] we study modular tensor categories whose simple objects have dimension a multiple of an odd prime number $p$ and modular tensor categories with dimension a power of $p$, in particular, we prove that a modular tensor category of dimension $p^5$, with $p$ an odd prime, must be pointed. In Section \[section: zeros S\] we investigate the entries of the $S$-matrix of a modular tensor category (with no restrictions on the dimension this time); specifically, we give conditions for having (or not) zeros on it. In Section \[section: perfect\] we show that odd-dimensional modular tensor categories with exactly one invertible object (the unit) have no non-trivial symmetric categories. This fact has strong consequences, such as that every fusion subcategory of such category is modular and therefore the category is split. In Section \[section: ranks\] we find bounds for the rank of a the category in terms of data associated to its universal grading. We also give conditions on the rank of the graded components in relation to the fixed points by the action of the group of invertible objects of the category. In Section \[section: mnsd low rank\] we apply the results on the previous sections to classify MNSD modular tensor categories of rank between $13$ and $23$. Lastly, in Section \[Future directions\] we give some equivalent statements to Conjectures \[conjecture: odd dim non perfect\] and \[Conjecture: odd fusion are solvable\]. Acknowledgements {#acknowledgements .unnumbered} ================ The authors specially thank A. Bruguières for enlightening discussions. They also thank P. Bruillard, E. Campagnolo, H. Peña Polastri, G. Sanmarco and L. Villagra for useful comments. The authors thank C. Galindo, M. , Müller, V. Ostrik, E. Rowell, and A. Schopieray for helpful remarks on an early draft. Substantial portions of this project were discussed at the CIMPA Research Schools on Quantum Symmetries (Bogotá, 2019) and Hopf algebras and Tensor Categories (Córdoba, 2019) and the authors would like to thank all the various organizers and granting agencies involved. Both authors were partially supported by NSF grant DMS-0932078, administered by the Mathematical Sciences Research Institute while the first author attended a workshop and the second author was in residence at MSRI during the Quantum Symmetries program in Spring 2020. JP was partially supported by NSF grants DMS-1802503 and DMS-1917319. JP gratefully acknowledges the support of Indiana University, Bloomington, through a Provost’s Travel Award for Women in Science. Preliminaries {#section: preliminaries} ============= In this paper, we will always work over an algebraically closed field $\textbf{k}$ of characteristic zero. We refer to [@BK] and [@EGNO] for the basic theory of fusion categories and braided fusion categories, and for unexplained terminology used throughout this paper. A fusion category $\mathcal C$ over $\textbf{k}$ is a $\textbf{k}$-linear semisimple rigid tensor category with a finite number of simple objects and finite dimensional spaces of morphisms, and such that the endomorphism algebra of the identity object (with respect to the tensor product) $\textbf{1}$ is $\textbf{k}$. Let $\mathcal C$ be a fusion category over $\textbf{k}$. We shall denote by $\mathcal{O(C)}$ the set of isomorphism classes of simple objects of $\mathcal C$. After fixing an enumeration of $\mathcal{O(C)}$, we denote the fusion coefficients by $N^k_{ij} := \dim {\operatorname{Hom}}(X_i \otimes X_j , X_k)$, where $X_i, X_i, X_k \in \mathcal{O(C)}$. The Frobenius-Perron dimension of an object $X$ is denoted by ${\operatorname{FPdim}}{X}$ (see [@ENO1]). We will use the notation ${\operatorname{c.d.}}(\mathcal C) = \{{\operatorname{FPdim}}(X) : X \in \mathcal{O(C)}\}$. The Frobenius-Perron dimension of $\mathcal C$ will be denoted by ${\operatorname{FPdim}}(\mathcal C)$. When $\mathcal{C}$ is endowed with a pivotal structure, ${\operatorname{d}}_X$ denotes the quantum dimension of an object $X$ and $\dim(\mathcal C)$ denotes the global dimension of $\mathcal C$. When the Frobenius-Perron dimension of a fusion category is an integer, it coincides with the global dimension (see [@ENO1 Proposition 8.24]). A fusion category is *pointed* if all simple objects are invertible, which is equivalent to having ${\operatorname{c.d.}}(\mathcal C)=\{1\}$. In this case $\mathcal C$ is equivalent to the category of finite dimensional $G$-graded vector spaces ${\operatorname{\textbf{Vec}_G^{\omega}}}$, where $G$ is a finite group and $\omega$ is a 3-cocycle on $G$ with coefficients in $\textbf{k}^{\times}$ codifying the associativity constraint. The group of isomorphism classes of invertible objects of $\mathcal C$ will be indicated by $\mathcal{G(C)}$. The largest pointed subcategory of $\mathcal C$ will be denoted by $ \mathcal{C}_\mathrm{pt}$, that is the fusion subcategory of $\mathcal{C}$ generated by $\mathcal{G(C)}$. Fusion categories ----------------- ### Nilpotent, solvable and weakly group-theoretical fusion categories Let $\mathcal{C}$ be a fusion category. The *adjoint subcategory*, indicated by $\mathcal{C}_{\mathrm{ad}}$, is the fusion subcategory of $\mathcal{C}$ generated by $\{ X\otimes X^* \colon X \in \mathcal{O(C)} \}$. The *upper central series* of $\mathcal{C}$ is the sequence of fusion subcategories of $\mathcal{C}$ defined recursively by $$\mathcal{C}^{(0)}=\mathcal{C} \text { and }\mathcal{C}^{(n)}=\big(\mathcal{C}^{(n-1)}\big)_{\mathrm{ad}} \ \text{ for all } n\geq 1.$$ The fusion category $\mathcal{C}$ is *nilpotent* if its upper central series converges to the trivial fusion subcategory $\operatorname{Vec}$ of finite dimensional vector spaces; that is, there exists $n\in \mathbb N_{\geq 0}$ such that $\mathcal{C}^{(n)}=\operatorname{Vec}$ [@GN], [@ENO2]. Fusion categories of Frobenius-Perron dimension a power of a prime integer are nilpotent [@GN]. We make repeated use of this result. More generally, $\mathcal{C}$ is *weakly group-theoretical* if it is Morita equivalent to a nilpotent fusion category [@ENO2]. The class of weakly group-theoretical fusion categories is closed under taking extensions and equivariantizations, Morita equivalence, tensor product, Drinfeld centers and subcategories [@ENO2 Proposition 4.1]. A fusion category $\mathcal{C}$ is *solvable* if it is Morita equivalent to a cyclically nilpotent fusion category. The class of solvable categories is closed under taking extensions and equivariantizations by solvable groups, Morita equivalent categories, tensor products, Drinfeld center, fusion subcategories and component categories of quotient categories [@ENO2 Proposition 4.5]. For a solvable braided fusion category $\mathcal C$, we have that either $\mathcal C=\text{Vec}$ or $\mathcal {G(C)}$ is not trivial [@ENO2 Proposition 4.5]. It is also known that braided nilpotent fusion categories are solvable [@ENO1 Proposition 4.5]. When $\mathcal C$ is nilpotent, then ${\operatorname{FPdim}}(X)^2$ divides ${\operatorname{FPdim}}(\mathcal C_{\mathrm{ad}})$ for all $X\in \mathcal{O(C)}$ [@GN]. ### The universal grading Let $G$ be a finite group. A $G$-grading on a fusion category $\mathcal C$ is a decomposition $\mathcal C = \oplus_{g\in G} \mathcal C_g$, such that $\otimes: \mathcal C_g\times \mathcal C_h \to \mathcal C_{gh}$, $\textbf{1}\in \mathcal C_{e}$, and $ ^*:\mathcal C_g \to \mathcal{C}_{g^{-1}}$. A $G$-grading is said to be *faithful* if $\mathcal C_g\ne 0$ for all $g \in G$. If $\mathcal C$ is a fusion category endowed with a faithful grading $\mathcal C = \oplus_{g\in G} \mathcal C_g$, then all the components $\mathcal C_g$ have the same Frobenius-Perron dimension [@ENO1 Proposition 8.20]. Hence, ${\operatorname{FPdim}}(\mathcal C)=|G|{\operatorname{FPdim}}(\mathcal{C}_e)$. By [@GN], any fusion category $\mathcal{C}$ admits a canonical faithful grading $\mathcal{C}= \oplus_{g \in \mathcal{U(C)}} \mathcal{C}_g$, called the *universal grading*; its trivial component is the adjoint subcategory $\mathcal{C}_{\mathrm{ad}}$ of $\mathcal{C}$. If $\mathcal{C}$ is equipped with a braiding, then $\mathcal{U(C)}$ is abelian. Moreover, if $\mathcal C$ is modular then $\mathcal{U(C)}$ is isomorphic to the group of (isomorphism classes of) invertibles $\mathcal{G(C)}$ [@GN Theorem 6.3]. ### Pseudo-unitary fusion categories A pivotal fusion category $\mathcal{C}$ is said to be *pseudo-unitary* if $\dim(\mathcal{C})={\operatorname{FPdim}}(\mathcal{C})$. This happens to be equivalent to ${\operatorname{d}}_X^2={\operatorname{FPdim}}(X)^2$ for all $X \in \mathcal{O(C)}$. By [@EGNO Proposition 9.6.5], weakly integral fusion categories are pseudo-unitary. We will sometimes use the Frobenius-Perron dimension and the global dimension of $\mathcal C$ indifferently when working with pseudo-unitary categories [@EGNO Corollary 9.6.6]. Modular tensor categories ------------------------- Let $\mathcal{C}$ be a braided fusion category endowed with a braiding $\sigma$. A *twist* in $\mathcal{C}$ is a natural isomorphism $\theta:{\operatorname{Id}}_\mathcal{C} \to {\operatorname{Id}}_\mathcal{C}$ such that $$\begin{aligned} \theta_{X\otimes Y}=(\theta_X \otimes \theta_Y) \circ \sigma_{Y,X} \circ \sigma_{X,Y}, \end{aligned}$$ for all $X,Y \in \mathcal{C}$. A twist is called a *ribbon structure* if $(\theta_X)^*=\theta_{X^*}$ for all $X \in \mathcal{C}$. A *pre-modular fusion category* is a fusion category endowed with a compatible ribbon structure. Equivalently, a pre-modular tensor category is a braided fusion category equipped with a spherical structure [@Br]. That is, ${\operatorname{d}}_X={\operatorname{d}}_{X^*}$ for all $X \in \mathcal{O(C)}$. In a spherical category, for an endomorphism $f\in {\operatorname{End}}_{\mathcal C}(X)$ we have a notion of *trace*, which we will denote by ${\operatorname{Tr}}(f)$ (see [@EGNO Definition 4.7.1]). Let $\mathcal C$ be a premodular tensor category, with braiding $\sigma_{X,Y}:X\otimes Y\xrightarrow{\sim}Y\otimes X$. The *S-matrix* $S$ of $\mathcal C$ is defined by $S:= \left(s_{X,Y}\right)_{X,Y \in \mathcal{O(\mathcal{C})}}$, where $s_{X,Y}={\operatorname{Tr}}(\sigma_{Y,X}\sigma_{X,Y})$. In a premodular tensor category $\mathcal C$, we can obtain the entries of the $S$-matrix in terms of the twists, fusion rules, and quantum dimensions via the so-called *balancing equation*\[balancing\] $$\begin{aligned} s_{X,Y}=\theta_X^{-1}\theta_Y^{-1} \sum\limits_{Z \in \mathcal{O(\mathcal{C})}} N_{XY}^Z \theta_Z {\operatorname{d}}_Z, \end{aligned}$$ for all $X,Y \in \mathcal{O(\mathcal{C})}$ [@EGNO Proposition 8.13.7]. A premodular tensor category $\mathcal C$ is said to be *modular* if the $S$-matrix $S$ is non-degenerate. ### Centralizers in braided fusion categories Let $\mathcal{C}$ be a braided fusion category and let $\mathcal{K}$ be a fusion subcategory of $\mathcal{C}$. The *centralizer* $\mathcal{K}'$ of $\mathcal{K}$ is the fusion subcategory of $\mathcal C$ with objects all those $Y$ in $\mathcal{C}$ such that $$\begin{aligned} \label{centralizador} \sigma_{Y,X}\sigma_{X,Y}={\operatorname{id}}_{X\otimes Y}, \ \ \text{for all}\ X \in \ \mathcal{K} \ \ \ \ \ \ \text{\cite{Mu}}.\end{aligned}$$ In particular, if $\mathcal{C}$ is modular then $\mathcal{K}=\mathcal{K}''$ and $\dim(\mathcal{K})\dim(\mathcal{K}')=\dim(\mathcal{C})$ [@Mu Theorem 3.2]; moreover, $\mathcal{C}_\mathrm{pt}=\mathcal{C}_{\mathrm{ad}}'$ and $\mathcal{C}_\mathrm{pt}'=\mathcal{C}_{\mathrm{ad}}$ [@GN Corollary 6.9]. A necessary and sufficient condition for $\mathcal{K}$ to be modular is $\mathcal{K} \cap \mathcal{K}' = \ \text{Vec}$. In this case $\mathcal{K}'$ is also modular and $\mathcal{C} \simeq \mathcal{K} \boxtimes \mathcal{K}'$ as braided fusion categories [@Mu]. A braided fusion category is *symmetric* if the square of the braiding is the identity. Hence $\mathcal{K}$ is symmetric if and only if $\mathcal{K} \subseteq \mathcal{K}'$. \[Cad\_pt\_symm\] If $\mathcal{C}$ is a braided fusion category, then $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is symmetric. First note that $(\mathcal{C}_{\mathrm{ad}})'=(\mathcal{C}')^{co}$ [@DGNO2 Proposition 3.25], i.e, $X\in (\mathcal{C}_{\mathrm{ad}})'$ if and only if $X\otimes X^* \in \mathcal{C}'$. In particular, $\mathcal{C}_\mathrm{pt}\subseteq (\mathcal{C}_{\mathrm{ad}})' $, and the claim follows since $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}} \subseteq \mathcal{C}_\mathrm{pt} \subseteq (\mathcal{C}_{\mathrm{ad}})' \subseteq ((\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}})'.$ Given a fusion category $\mathcal{C}$, we will identify the elements in $\mathcal{G}(\mathcal{C})$ with the invertible objects in $\mathcal{C}_\mathrm{pt}$. Let $\mathcal{C}$ be a modular tensor category and $\mathcal K$ be a fusion subcategory of $\mathcal C$. Then $\mathcal K \cap \mathcal K'$ is symmetric. From now on let $\mathcal{C}$ be a pre-modular fusion category. \[gX=X\] Let $g$ be an invertible object in $\mathcal{C}_{\mathrm{ad}}$ such that $\theta_g=1$. Suppose $g\otimes X=X$ for all non-invertible simple $X\not\in \mathcal{C}_{\mathrm{ad}}$. Then the rows of the S-matrix corresponding to (the isomorphism classes of) $g$ and **1** are equal. Let $g$ be as above. The balancing equation yields $$\label{G2} s_{g,X}= \theta_g^{-1}\theta_X^{-1}{\operatorname{d}}_X\theta_{X}= {\operatorname{d}}_X, \text{ for all non-invertible simple $X\not\in \mathcal C_{\mathrm{ad}}$},$$ that is, for all simple $X$ such that $X\not\in \mathcal {C}_{\mathrm{ad}}\cup \mathcal{C}_\mathrm{pt}.$ Now we compute $s_{g,X}$ for $X\in \mathcal{C}_\mathrm{pt}\cup \mathcal {C}_{\mathrm{ad}}$. Assume first that $X\in \mathcal C_{\mathrm{ad}}$. Since $g\in (\mathcal{C}_\mathrm{pt})\subseteq (\mathcal {C}_{\mathrm{ad}})'$, it follows from [@Mu Proposition 2.5] that $$\label{G1} s_{g,X} ={\operatorname{d}}_X .$$ Lastly, assume $X\in \mathcal{C}_\mathrm{pt}$. Since $g\in (\mathcal {C}_{\mathrm{ad}})_{\mathrm{pt}}$ and $\mathcal {C}_{\mathrm{pt}}\subseteq ((\mathcal {C}_{\mathrm{ad}})_{\mathrm{pt}})'$, again, by [@Mu Proposition 2.5], we get $$\label{G3} s_{g,X} ={\operatorname{d}}_X.$$ Now the result follows from equations , and . For every $X\in \mathcal {O(C)}$ let $G[X]=\{ g \in \mathcal C : g \text{ is invertible and } g \otimes X =X\}$. We define $$\textbf{G} := \bigcap\limits_{ \text{non-invertible simple} \ X \not\in \mathcal{C}_{\mathrm{ad}} } G[X].$$ Note that $G[X]$ is a subgroup of $\mathcal{G}(\mathcal{C}_{\mathrm{ad}})$ for all simple $X \in \mathcal C$, and thus so is $\textbf{G}$ . \[G\] If $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is isotropic, i.e. if $\theta_g=1$ for all $g\in (\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$, then the rows of the S-matrix that correspond to the elements of $\textbf{G}$ are all equal. In particular, the previous result holds if $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is odd-dimensional [@DGNO1 Corollary 2.7]. From Corollary \[G\] we get \[intersection\] Let $\mathcal{C}$ be a modular tensor category. If $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is isotropic, then $\textbf{G} =\{ 1\}$. Modular tensor categories with certain irreducible degrees {#section: cd_p} ========================================================== In this section, we study modular tensor categories whose irreducible degrees are a multiple of an odd prime number $p$. In particular, we study modular tensor categories that are integral and whose adjoint subcategory has dimension a power of $p$. We also need some more general auxiliary results that we prove in this section, (see Proposition  \[p2\] and Lemma  \[invertible\_square free\]). \[cyclic\] Let $\mathcal{C}$ be a not pointed modular tensor category and $p$ be an odd prime number such that ${\operatorname{c.d.}}(\mathcal C) \subset \{1\} \cup p \mathbb{Z}$. If $\mathcal{G}(\mathcal{C}_{\mathrm{ad}})$ is a non-trivial p-group, then it is not cyclic. In particular, $|\mathcal{G}(\mathcal{C}_{\mathrm{ad}})| \geq p^2$. Assume $\mathcal{G}(\mathcal{C}_{\mathrm{ad}})$ is a cyclic group of order $p^k$, for some $k\in \mathbb{N}$. For every $i \in \{0,\dots, k\}$ denote by $H_i$ the unique subgroup of $\mathcal{G}(\mathcal{C}_{\mathrm{ad}})$ of order $p^i$. Note that $$\label{cyclic 1} H_0=\{e\} \subset H_1 \subset \dots \subset H_k.$$ We claim that there exists $1\leq l\leq k$ such that $H_l \subseteq G[X]$ for every non-invertible simple object $X\in \mathcal C$. As $p$ is odd, this is a contradiction by Corollary \[intersection\]. Indeed, taking the Frobenius-Perron dimension of both sides of $$X\otimes X^*= \bigoplus\limits_{g \in G[X]} g \oplus \bigoplus\limits_{Y \in \mathcal{C}_{\mathrm{ad}}, {\operatorname{d}}_Y>1} N_{XX^*}^Y Y,$$ we obtain that $p$ divides $|G[X]|$. Thus $G[X]$ has order $p^{j_X}$, where $1\leq j_X \leq k$. Define $l := \min\limits_{X} j_X \geq 1$ and consider the subgroup $H_l$ of $\mathcal{G}(\mathcal{C}_{\mathrm{ad}})$. By Equation , we have that $H_l \subseteq G[X]$ for every simple non-invertible $X$ in $\mathcal{C}$ and then the claim follows. \[cyclic 2\] Let $\mathcal{C}$ be a not pointed modular tensor category and $p$ an odd prime integer such that ${\operatorname{c.d.}}(\mathcal C) \subset \{1\} \cup p \mathbb{Z}$. If $\mathcal{G}(\mathcal{C})$ is a cyclic p-group, then $\mathcal{G}(\mathcal{C}_{\mathrm{ad}})$ is trivial. \[A\] Let $\mathcal{C}$ be a not pointed modular tensor category and $p$ an odd prime number such that ${\operatorname{c.d.}}(\mathcal C) \subset \{1\} \cup p \mathbb{Z}$. If $\mathcal{C}$ is solvable, then $\mathcal{G(C)}$ is not a cyclic p-group. This is a direct consequence of Corollary \[cyclic 2\] since $\mathcal{C}_{\mathrm{ad}}$ contains a non-trivial invertible object because it is a non-trivial solvable category [@ENO1 Proposition 4.5]. In this work we are mostly interested in odd-dimensional categories, but some results hold also for $p = 2$. We do not assume that $p$ is odd unless otherwise stated. \[p2\] Let $\mathcal{C}$ be an integral modular tensor category and $p$ be a prime number that divides ${\operatorname{FPdim}}(\mathcal{C})$. Then ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}}) \ne p^2$. Assume ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}}) = p^2$. Then $\mathcal{C}$ is a non-pointed nilpotent category. Furthermore, $\mathcal{C}_{\mathrm{ad}}$ is solvable and thus it contains a non-trivial invertible object [@ENO2 Proposition 4.5]. This together with the fact that every simple object has Frobenius-Perron dimension 1 or $p$ [@GN Corollary 5.3] implies that $\mathcal{C}_{\mathrm{ad}}$ is necessarily pointed. Note that by [@DGNO1 Corollary 2.7] there is a non-trivial object $h \in \mathcal{C}_{\mathrm{ad}}$ such that $\theta_h=1$. We show that the rows of the S-matrix corresponding to (the isomorphism classes of) $h$ and $\textbf{1}$ are equal. Given $g \in \mathcal{U(C)}$, let $a_g$ and $b_g$ denote the number of isomorphism classes of simple objects of dimension 1 and $p$ in the component $\mathcal{C}_g$, respectively. Since $p^2 = {\operatorname{FPdim}}(\mathcal{C}_g) = a_g + b_g p^2$, the following holds: - Either $\mathcal{C}_g$ has exactly $p^2$ simple objects, all of which are invertible, or - $\mathcal{C}_g$ has exactly one simple object, which has dimension $p$. Let $X$ be a simple object of Frobenius-Perron dimension $p$ and let $g\in \mathcal U(C)$ such that $X \in \mathcal{C}_g$. As $X$ is the unique simple object in $\mathcal{C}_g$ we have $h \otimes X =X$. Using the balancing equation we obtain $$\label{6} s_{h,X}= \theta_h^{-1}\theta_X^{-1} {\operatorname{d}}_X \theta_X= {\operatorname{d}}_X.$$ On the other hand, since $h\in \mathcal{C}_{\mathrm{ad}} =\mathcal{C}'_{\mathrm{pt}}$ it follows from [@Mu Proposition 2.5] that for all invertible object $g \in \mathcal{C}$ we have $$\label{5} s_{h,g}= {\operatorname{d}}_h{\operatorname{d}}_g= {\operatorname{d}}_g.$$ Equations and prove our claim. This is a contradiction since $S$ is invertible. Note that this yields an alternate proof of Lemma 4.11 of [@DN]. \[p4\] Let $p$ be a prime number. Modular tensor categories of Frobenius-Perron dimension $p^4$ are pointed. It is easy to see that $p^2$ divides ${\operatorname{FPdim}}(\mathcal{C}_\mathrm{pt})$. If $\mathcal{C}$ is not pointed we have ${\operatorname{FPdim}}({\mathcal C}_{\mathrm{pt}})=p^2$ [@DLD Lemma 3.2]. Thus the Frobenius-Perron dimension of $\mathcal C_{\mathrm{ad}}$ is also $p^2$, which cannot happen by Proposition \[p2\]. \[p5\]Let $p$ be an odd prime. Modular tensor categories of Frobenius-Perron dimension $p^5$ are pointed. Let $\mathcal{C}$ be an integral modular tensor category of Frobenius-Perron dimension $p^5$, and suppose that $\mathcal C$ is not pointed. Then by [@DLD Lemma 3.2] and Proposition \[p2\] we get that ${\operatorname{FPdim}}(\mathcal{C}_{pt})=p^2$, and so $\mathcal C_{pt}\subseteq \mathcal C_{ad}$ [@DLD Lemma 3.4]. Hence $\mathcal C_{pt}$ is an odd-dimensional symmetric subcategory of $\mathcal C$ and thus $\mathcal C_{pt}\simeq {\operatorname{Rep}}(G)$ for some group $G$ of order $p^2$. Consider the de-equivariantization $\mathcal C_G$ of $\mathcal C$ by the Tannakinan subcategory ${\operatorname{Rep}}(G).$ Let $\mathcal C_G^0$ denote the neutral component with respect to the associated $G$-grading. Since $\mathcal C$ is non-degenerate then $C_G^0$ is non-degenerate [@DGNO2 Proposition 4.6]. Moreover, we have that ${\operatorname{FPdim}}(\mathcal C_G^0)={\operatorname{FPdim}}(\mathcal C)/|G|^2=p$. Hence $\mathcal C$ must be the gauging of a modular tensor category $\mathcal D$ of dimension $p$ by the group $G$. Since $G$ is odd-dimensional, it follows from [@BGPR Proposition 2.8] that the action of $G$ on $\mathcal D$ is the trivial action. Then, the gauging of $\mathcal D$ by $G$ is $\mathcal C= \mathcal D\boxtimes \mathcal Z(\text{Vec}_H^{\omega})$. Note that the Frobenius-Perron dimensions of $\mathcal D$ and $Z(\text{Vec}_G^{\omega})$ are $p$ and $p^4$, respectively, so $\mathcal D$ and $Z(\text{Vec}_G^{\omega})$ are both pointed, see Corollary \[p4\]. Consequently $\mathcal C$ is also pointed, and we arrive at a contradiction. \[Cad\] Let $\mathcal C$ be a fusion category and $p$ a prime number such that ${\operatorname{FPdim}}(\mathcal{C})= p^3$. Then $\mathcal{C}_{\mathrm{ad}}$ is pointed. As $\mathcal{C}$ is nilpotent, ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})=1, p$ or $p^2$. If $X$ is a simple object in $\mathcal{C}_{\mathrm{ad}}$, then ${\operatorname{FPdim}}(X)$ can be either 1 or $p$ [@GN Corollary 5.3]. By a dimension argument, $\mathcal{C}_{\mathrm{ad}}$ must be pointed. \[Cad2\] Let $\mathcal C$ be an integral modular tensor category and $p$ an odd prime number such that ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})= p^3$. Then one of the following is true: 1. $\mathcal{C}_{\mathrm{ad}}$ is pointed and $\mathcal{G}(\mathcal{C}_{\mathrm{ad}})\simeq \mathbb{Z}_p \times \mathbb{Z}_p \times \mathbb{Z}_p.$ 2. $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{ad}}=(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ and $\mathcal{G}(\mathcal{C}_{\mathrm{ad}})\simeq \mathbb{Z}_{p}\times \mathbb{Z}_p$. First note that the dimensions of simple objects of $\mathcal{C}$ are either $1$ or $p$ by [@GN Corollary 5.3]. Assume $\mathcal{C}_{\mathrm{ad}}$ is not pointed. Then $\mathcal{C}_{\mathrm{ad}}$ has at least one simple object of dimension $p$. On the other hand, Lemma \[Cad\] implies that $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{ad}}$ must be pointed, and ${\operatorname{FPdim}}(X)^2$ divides ${\operatorname{FPdim}}((\mathcal{C}_{\mathrm{ad}})_{\mathrm{ad}})$ for all simple $X \in \mathcal{C}_{\mathrm{ad}}$ [@GN Corollary 5.3]. Therefore ${\operatorname{FPdim}}((\mathcal{C}_{\mathrm{ad}})_{\mathrm{ad}})=p^2$. Hence, by Proposition \[cyclic\] we get $\mathcal{G}(\mathcal{C}_{\mathrm{ad}})\simeq\mathbb{Z}_{p}\times \mathbb{Z}_p$. Now if $\mathcal{C}_{\mathrm{ad}}$ is pointed then $\mathcal{G}(\mathcal{C}_{\mathrm{ad}}) \simeq \mathbb{Z}_p \times \mathbb{Z}_p \times \mathbb{Z}_p$ or $\mathbb{Z}_{p^2} \times \mathbb{Z}_p$ by Proposition \[cyclic\]. Assume the latter holds. Let $X$ be a simple object in $\mathcal{C}$ of dimension $p$. Since $X\otimes X^* \in \mathcal{C}_{\mathrm{ad}}$, then $X\otimes X^*= \bigoplus\limits_{g\in G[X]} \ g$. Hence, $|G[X]|=p^2$ for every non-invertible simple $X$. As the intersection of all subgroups of order $p^2$ of $\mathbb{Z}_{p^2}\times \mathbb{Z}_p$ is non-trivial, this is a contradiction by Proposition \[intersection\]. \[invertible\_square free\] Let $\mathcal{C}$ be a weakly-integral modular tensor category such that $|\mathcal{G(C)}|$ is square-free. Then $\gcd\{{\operatorname{FPdim}}(X) \ / \text{ non-invertible simple} \ X \}=1.$ Assume there exists a prime $p$ such that $p$ divides $\dim(X)$ for all non-invertible simple $X$ in $\mathcal{C}$. By [@ENO2 Theorem 2.11] we have that $p^2$ divides $\dim(\mathcal{C})$. Note that $${\operatorname{FPdim}}(\mathcal{C})=|\mathcal{G(C)}| + \sum\limits_{X \in \mathcal{O(C)}\setminus \mathcal{G(C)}} {\operatorname{FPdim}}(X)^2,$$ and thus $p^2$ divides $|\mathcal{G(C)}|$ which is a square free number, a contradiction. The next proposition mimics the argument in the proof of Theorem 8.2 in [@N1]. Let $\mathcal C$ be a modular tensor category of dimension $cp^2q^2r^2$, where $p, q$ and $r$ are odd prime numbers and $c$ is a square-free odd integer such that $\gcd(c, pqr) = 1.$ Then $\mathcal C$ is weakly group-theoretical. By the proof of [@ENO2 Lemma 9.3] $\mathcal C$ contains a nontrivial symmetric subcategory $\mathcal D$. Since $\mathcal C$ is odd-dimensional then $\mathcal D$ is Tannakian and thus $\mathcal D \simeq \text{Rep}(G)$ for some finite group $G$. Since $\mathcal C$ is non-degenerate, so is its core $\mathcal C_G^0$ [@ENO1]. Moreover, $\dim(\mathcal C_G^0)=\dim(\mathcal C)/|G|^2$. Thus by [@N2 Theorem 7.4] $\mathcal C_G^0$ is weakly group theoretical, and hence so is $\mathcal C.$ Zeros of the $S$-matrix {#section: zeros S} ======================= Let $\mathcal{C}$ be a modular tensor category and $g$ be an invertible object in $\mathcal{C}$. Note that for any simple object $X$ in $\mathcal{C}$ we have that $\sigma_{g,X}^2 \in {\operatorname{Hom}}(g\otimes X, g\otimes X)\simeq \textbf{k}$. Define $\xi_g(X) \in \textbf{k}^\times$ by $$\sigma_{g,X}^2=\xi_g(X){\operatorname{id}}_{g\otimes X}.$$ \[ceros\] Let $\mathcal{C}$ be a modular tensor category. For all $X \in \mathcal{O(C)}$ such that $G[X]$ is non-trivial there exists $Y\in \mathcal{O(C)}$ such that $s_{X,Y}=0.$ For such $X$, let $g \in G[X]/\{\textbf{1}\}$. Note that for all $Y \in \mathcal{O(C)}$, $$\label{gX} {\operatorname{Tr}}(\sigma_{g\otimes X,Y}^2)=s_{g\otimes X,Y}=s_{X,Y}= {\operatorname{Tr}}(\sigma_{X,Y}^2).$$ By the hexagon axioms, $$\begin{aligned} \sigma_{g\otimes X,Y}^2=({\operatorname{id}}_g\otimes \sigma_{Y,X})(\sigma_{Y,g} \otimes {\operatorname{id}}_X)(\sigma_{g,Y}\otimes {\operatorname{id}}_X)({\operatorname{id}}_g \otimes \sigma_{X,Y})=\xi_g(Y)({\operatorname{id}}_g\otimes \sigma_{X,Y}^2). \end{aligned}$$ Taking trace on both sides of the previous equation we get ${\operatorname{Tr}}(\sigma_{g\otimes X,Y}^2)=\xi_g(Y)s_{X,Y}$. This together with equation implies $(1-\xi_g(Y))s_{X,Y}=0$. As $g\ne \textbf{1}$, $\xi_g$ is non-trivial, there exists a simple $Y$ such that $\xi_g(Y)\ne 1$. Lastly, note that for all simple $Y$ such that $\xi_g(Y) \ne 1$, we have that $s_{X,Y}=0$. Let $\mathcal{C}$ be a modular tensor category. If its $S$-matrix has no null entries, $\mathcal{G(C)}$ acts freely over $\mathcal{O(C)}$. Suppose $g\otimes X= h\otimes X $ for some $g,h \in \mathcal{G(C)}$ and $X \in \mathcal{O(C)}$. Then $h^{-1}g\in G[X]$ which is trivial by Theorem \[ceros\]. Hence, $h=g$. For all $X\in \mathcal{O(C)}$ let $$\mathcal{G(C)} \cdot X:=\{ Y \in \mathcal{O(C)} \ / \ h\otimes X= Y \ \text{for some} \ h\in \mathcal{G(C)} \}$$ Let $\mathcal C$ be a modular tensor category and $X\in \mathcal{O(C)}$. If $s_{X,Z}=0$ for some $Z\in \mathcal{O(C)},$ then $$\begin{aligned} s_{YZ}=0 \ \ \text{for all} \ Y \in \mathcal{G(C)} \cdot X.\end{aligned}$$ In particular, if $s_{X,Z}=0$ for $Z \in \mathcal{G(C)} \cdot X$, then $s_{X,Y}=0$ for all $Y \in \mathcal{G(C)} \cdot X$. Fix $Z\in \mathcal{O(C)}$ such that $s_{X,Z}=0$. For $Y \in \mathcal{G(C)} \cdot X$ and $h \in \mathcal{G(C)}$ such that $h\otimes X \simeq Y$ we have that $$0=s_{X,Z}s_{h,Z}={\operatorname{d}}_Z \sum\limits_{W \in \mathcal{O(C)}} N_{Xh}^W s_{Z,W}= {\operatorname{d}}_Z s_{Z,Y}.$$ Hence, $s_{Z,Y}=0$. Perfect modular tensor categories {#section: perfect} ================================= In analogy with group theory, we call a fusion category *perfect* if the only $1$-dimensional simple object is the unit, that is, if $\mathcal{G(C)}= \{\textbf{1}\}$. Perfect fusion categories are sometimes called unpointed. \[nosymmetric\] Let $\mathcal{C}$ be a perfect odd-dimensional braided fusion category. Then $\mathcal{C}$ has no non-trivial symmetric subcategories. In particular, $\mathcal C$ is modular. Let $\mathcal{E}$ be a symmetric subcategory of $\mathcal{C}$. As $\mathcal{C}$ is odd-dimensional, $\mathcal{E}$ is Tannakian and thus $\mathcal{E} \simeq {\operatorname{Rep}}(G)$ for some finite group $G$. Note that the unit is the only invertible object in $\mathcal{C}$, hence $G$ must be a perfect group. Thus, $G$ must be trivial, as the order of every non-trivial finite perfect group is divisible by four and $\mathcal{C}$ is odd-dimensional. This implies that $\mathcal{E}$ is trivial. \[subcategoriesaremodular\] Let $\mathcal{C}$ be a perfect odd-dimensional modular tensor category. Then every fusion subcategory of $\mathcal{C}$ is modular. In particular, $$\mathcal C \simeq \mathcal C[X] \boxtimes \mathcal C[X]'$$ for all $X\in \mathcal{O(C)}.$ Let $\mathcal{D}$ be a non-trivial fusion subcategory of $\mathcal{C}$, and consider its Müger centralizer $\mathcal{D}'$. Note that $\mathcal{D} \cap \mathcal{D}'$ is a symmetric subcategory of $\mathcal{C}$, and thus by Lemma \[nosymmetric\] it must be trivial. That is, $\mathcal{D} \cap \mathcal{D}' \simeq \text{Vec}$, and therefore $\mathcal{D}$ is modular by [@DGNO1 Section 2.5]. \[perfect: nosubcategories\] Let $\mathcal{C}$ be a perfect odd-dimensional modular tensor category of prime rank. Then $\mathcal{C}$ has no non-trivial fusion subcategories. Let $\mathcal{C}$ be an odd-dimensional perfect modular tensor category If $X$ and $Y$ are simple objects with coprime Frobenius-Perron dimensions then $s_{X, Y} = 0$ or $s_{X, Y} = d_Xd_Y$. Moreover, $s_{X, Y} \not = 0$ if and only if $X$ and $Y$ centralize each other. Fix $X$ a simple object in $\mathcal C$. Consider the full subcategory $\mathcal{D}_{\lambda}^X$ of objects $Z \in \mathcal C$ such that $\sigma_{Z,X}\sigma_{X,Z} = \lambda \cdot {\operatorname{id}}_{X\otimes Z}$ as in [@DGNO1 Lemma 3.15]. Moreover, the category $\mathcal{D}^X = \oplus_{\lambda\in \textbf{k}^*} \mathcal{D}_{\lambda}^X$ is a fusion subcategory of $\mathcal{C}$. By [@DGNO1 Proposition 3.22], $\mathcal{D}^X = \langle Y\in \mathcal{O(C)} | Y \text{centralizes } X\otimes X^*\rangle = C_{\mathcal C}((\mathcal{C}[X])_{\text{ad}})$. By Corollary \[subcategoriesaremodular\], $\mathcal{C}[X]$ is modular and, since $\mathcal{C}$ is perfect, so is $\mathcal{C}[X]$. Therefore, $\mathcal{C}[X])_{\text{ad}} = \mathcal{C}[X]$ and $\mathcal{D}^X = C_{\mathcal C}((\mathcal{C}[X])_{\text{ad}}) = ((\mathcal{C}[X]))'$, by [@GN Corollary 6.9]. It follows from [@ENO2 Lemma 7.1] that, since $\mathcal{D}^X = ((\mathcal{C}[X]))'$, for all $X$, $Y\in \mathcal{O(C)}$ of coprime Frobenius-Perron dimensions we have that $s_{X, Y} = 0$ or $s_{X, Y} = d_Xd_Y$ as desired. The following Corollary extends Lemma 10.2 in [@NPa] when the category is prime to the case $\gcd({\operatorname{FPdim}}(X), {\operatorname{FPdim}}(Y)) = 2$. Let $\mathcal{C}$ be a prime odd-dimensional perfect modular tensor category. If $X$ and $Y$ are simple objects with coprime Frobenius-Perron dimensions then $s_{X, Y} = 0$. Bounds on the ranks of graded components {#section: ranks} ======================================== \[rank1\] Let $\mathcal{C}$ be a modular tensor category and consider the universal grading $\mathcal{C}=\bigoplus\limits_{g\in \mathcal{U(C)}} \mathcal{C}_g.$ Then for each prime $p>2$ that divides $| \mathcal{G}(\mathcal{C}_{\mathrm{ad}})|$ there exists $h\in \mathcal{U(C)}$ such that $\mathcal{C}_h$ has at least $p$ non-invertible simple objects of the same dimension. In particular, if $\mathcal{C}$ is not pointed then ${\operatorname{rank}}(\mathcal{C}) \geq {\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}}) + |\mathcal{G(C)}| + p -2$ for all odd prime $p$ that divides $|\mathcal{G}(\mathcal{C}_{\mathrm{ad}})|$. Let $p>2$ that divides $|\mathcal{G}(\mathcal{C}_{\mathrm{ad}})|$ and let $g \in \mathcal{G}(\mathcal{C}_{\mathrm{ad}})$ of order $p$. Consider the fusion subcategory $\mathcal{C}[g]$ generated by $g$. Note that $\mathcal {C}[g]\subseteq (\mathcal C_{\mathrm{ad}})_{\mathrm{pt}}$ and thus $\mathcal {C}[g]$ is a symmetric subcategory of $\mathcal{C}$ (see Remark \[Cad\_pt\_symm\]). Since ${\operatorname{FPdim}}(\mathcal C[g])=p$ is odd then $\theta_g=1$ [@DGNO1 Corollary 2.7]. For all $h\in \mathcal{G(C)}\setminus\{e\}$ consider the action of $g$ on the non-invertible simple elements of $\mathcal C_h$ given by left multiplication. As the order of $g$ is $p$ it follows that for all $h\in \mathcal{G(C)}\setminus\{e\}$ this action is given either by the identity or by a cycle of length $p$. If the former holds for all $h\in \mathcal{G(C)}\setminus\{e\}$, by Lemma \[gX=X\] the rows of the $S$- matrix corresponding to $g$ and $\textbf{1}$ are equal, which is a contradiction as S is invertible. Thus, there must exist $h\ne 1$ such that $g$ acts as a cycle of length $p$ on the non-invertible simple elements of $\mathcal C_h$. Therefore there are at least $p$ different non-invertible simple objects in $\mathcal{C}_h$ of the same dimension. Finally recall that all the components of the universal grading have at least one simple element. Hence ${\operatorname{rank}}(\mathcal{C}) \geq {\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}}) + |\mathcal{U(C)}| + p -2 = {\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}}) + |\mathcal{G(C)}| + p -2$. \[rank2\] Let $\mathcal{C}$ be an odd-dimensional modular tensor category and $p$ an odd prime that divides $|\mathcal{G}(\mathcal{C}_{\mathrm{ad}})|$. Then $${\operatorname{rank}}(\mathcal{C}) \geq {\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}}) + |\mathcal{G}(\mathcal C)| + 2p - 3 .$$ Let $p$ be an odd prime that divides $|\mathcal{G}(\mathcal{C}_{\mathrm{ad}})|$. By Lemma \[rank1\] there exists $h\ne \textbf 1$ such that $\mathcal{C}_h$ has at least $p$ non invertible simple objects. As $\mathcal{C}$ is odd-dimensional, by [@NS Corollary 8.2(ii)], it is maximally non-self-dual. Thus, $\mathcal{C}_{h^{-1}}=\mathcal C_h^*$ also has at least $p$ non invertible simple objects, and the result follows. \[rank 4\] Let $\mathcal{C}$ be a modular tensor category such that $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is trivial. Then ${\operatorname{rank}}(\mathcal{C})= {\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}}) {\operatorname{rank}}(\mathcal{C}_\mathrm{pt})$. Note that $\mathcal{C}_{\mathrm{ad}} \cap \mathcal{C}_{\mathrm{ad}}' \simeq \mathcal{C}_{\mathrm{ad}} \cap (\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}} \simeq \text{Vec}$. Thus $\mathcal{C}_{\mathrm{ad}}$ is modular and we have a braided equivalence $\mathcal{C} \simeq \mathcal{C}_{\mathrm{ad}} \boxtimes \mathcal{C}_\mathrm{pt}$. Therefore ${\operatorname{rank}}(\mathcal{C})= {\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}}) {\operatorname{rank}}(\mathcal{C}_\mathrm{pt})$. \[primerank\] Let $\mathcal{C}$ be a modular tensor category of prime rank such that $ (\mathcal C_{\mathrm{ad}})_{\mathrm{pt}}$ is trivial. Then either $\mathcal{C}$ is pointed or $\mathcal{C}_\mathrm{pt}$ is trivial. \[dim cong rank\] Let $\mathcal{C}$ be an odd-dimensional fusion category. Then ${\operatorname{rank}}(\mathcal{D})\equiv \dim(\mathcal{D})$ ${\operatorname{mod}}{8}$ for every fusion subcategory $\mathcal{D}$ of $\mathcal{C}$. Any odd integer $n$ satisfies $n^2\equiv 1 \mod 8$. As the dimension of every simple element in $\mathcal{C}$ is an odd integer we get $$\dim(\mathcal{D})= \sum\limits_{X \in \mathcal{O(D)}} \dim(X)^2 \equiv \sum\limits_{X \in \mathcal{O(D)}} 1 = {\operatorname{rank}}(\mathcal{D}) \ \ \ \ {\operatorname{mod}}{8}.$$ \[cong 8\] Let $\mathcal{C}$ be an odd-dimensional modular tensor category and consider the universal (faithful) grading $\mathcal{C}=\bigoplus\limits_{g \in \mathcal{U(C)}} \mathcal{C}_g$. As a direct consequence of Lemma  \[dim cong rank\], We have $$\begin{aligned} {\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}}) \equiv {\operatorname{rank}}(\mathcal{C}_g) {\operatorname{mod}}{8},\end{aligned}$$ for all $g \in \mathcal{U(C)}.$ \[rank Cg\] Let $\mathcal C$ be a non-pointed modular tensor category such that $\mathcal{C}_\mathrm{pt}\subseteq \mathcal C_{\mathrm{ad}}$ and ${\operatorname{FPdim}}(\mathcal{C}_\mathrm{pt})=p^k$ for some odd prime $p$ and $k\in \mathbb N$. Consider the universal grading $\mathcal C=\bigoplus_{g\in \mathcal{G(C)}} \mathcal C_g$. Then $$\begin{aligned} {\operatorname{rank}}(\mathcal C_g)\equiv 0 {\operatorname{mod}}{p},\end{aligned}$$ for all $g\in \mathcal{G(C)}$ such that $g\ne 1$. Let $g\in \mathcal{G(C)}$ such that $g\ne 1$. Since $\mathcal{G(C)}$ acts on $\mathcal O(\mathcal C_g)$ by left multiplication and $\mathcal{G(C)}$ is a $p$-group, we have that the number of fixed elements by the action must be congruent to $|\mathcal O(\mathcal C_g)|$ modulo $p$. We show that there can be no fixed elements, and so the statement follows. Suppose there exists an element $X\in \mathcal O(\mathcal C_g)$ that is fixed by the action. Since $\mathcal{C}_\mathrm{pt}\subseteq \mathcal C_{\mathrm{ad}}$ and $\mathcal{C}_\mathrm{pt}'=\mathcal C_{\mathrm{ad}}$ we have that $\mathcal{C}_\mathrm{pt}$ is symmetric and odd-dimensional, and thus by [@DGNO1 Corollary 2.7] we get $\theta_h=1$ for all $h\in \mathcal{C}_\mathrm{pt}.$ Hence by the balancing equation $$\begin{aligned} s_{h,X}=\theta_h^{-1}\theta_X^{-1}\theta_X {\operatorname{d}}_X={\operatorname{d}}_X={\operatorname{d}}_h{\operatorname{d}}_X, \text{ for all } h\in \mathcal{C}_\mathrm{pt}.\end{aligned}$$ Therefore $X\in \mathcal{C}_\mathrm{pt}'=\mathcal C_{\mathrm{ad}}$ [@Mu Proposition 2.5], a contradiction. \[rank a\] Let $\mathcal C$ be a non-pointed modular tensor category such that $\mathcal{C}_\mathrm{pt}\subseteq \mathcal C_{\mathrm{ad}}$ and ${\operatorname{FPdim}}(\mathcal{C}_\mathrm{pt})=p^k$ for some odd prime $p$ and $k\in \mathbb N$. Then for all $g\in \mathcal{G(C)}$ such that $g\ne 1$ and $a\in \mathbb C$ the number of simple objects in $\mathcal C_g$ of dimension $a$ is congruent to 0 modulo $p$. Let $g\in \mathcal{G(C)}$ such that $g\ne 1$ and let $a\in \mathbb C$. Let $\mathcal C_g^a$ be the set of simple objects in $\mathcal C_g$ of dimension $a$. If $\mathcal C_g^a$ is empty the statement is clear. Assume $\mathcal C_g^a$ is not empty. Since $\mathcal{G(C)}$ acts on $\mathcal C_g^a$ by left multiplication and $\mathcal{G(C)}$ is a $p$-group, we have that the number of fixed elements by the action must be congruent to $|\mathcal C_g^a|$ modulo $p$. By the same argument given in the proof of Proposition \[rank Cg\] there can be no fixed elements, and so the statement follows. Application: low rank MNSD modular tensor categories {#section: mnsd low rank} ==================================================== In this section we prove that MNSD modular tensor categories of rank 13 and 15 are pointed. Moreover, we show that MNSD modular tensor categories of rank 17, 19, 21 and 23 are either pointed or perfect, that is, they have exactly one invertible object which is the unit object in the category. Let $\mathcal C$ be a MNSD modular tensor category such that ${\operatorname{rank}}(\mathcal{C})\in\{13, 15, 17, 19, 21, 23 \}$. Note that our claim is equivalent to showing that 1. if ${\operatorname{rank}}(\mathcal C)=13$ or $15$, then $|\mathcal{G(C)}|={\operatorname{rank}}(\mathcal{C})$, 2. if ${\operatorname{rank}}(\mathcal C)=17, 19, 21$ or $23$, then $|\mathcal{G(C)}|={\operatorname{rank}}(\mathcal{C})$ or $1$. We will prove our statement discarding the different possibilities for $|\mathcal{G(C)}|$ until we are left with the cases stated above. We start by proving the following useful Lemma. \[Cadpt no trivial\] Assume $\mathcal{C}$ is not pointed. Then $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is trivial if and only if $\mathcal{C}_\mathrm{pt}$ is trivial. It follows from Corollary \[primerank\] that $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ trivial implies $\mathcal{C}_\mathrm{pt}$ trivial for ranks $13, 17, 19$ or $23$. Let ${\operatorname{rank}}(\mathcal{C})=15$ and assume $(\mathcal C_{\mathrm{ad}})_{\mathrm{pt}}$ is trivial. If $\mathcal{C}_\mathrm{pt}$ is not trivial, then by Corollary \[primerank\] we have that $\mathcal{C}_{\mathrm{ad}}$ is a MNSD modular tensor category of rank $3$ or $5$, and thus it is pointed [@HR; @RSW; @BR], which is a contradiction. Similarly, if ${\operatorname{rank}}(\mathcal{C})=21$ and we assume that $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is trivial but $\mathcal{C}_\mathrm{pt}$ is not trivial, then $\mathcal{C}_{\mathrm{ad}}$ is a MNSD modular of rank $3$ or $7$ [@HR; @RSW; @BR], which is a contradiction. \[casos descartados\] Note that $|\mathcal{G(C)}|$ must be an odd integer smaller or equal to ${\operatorname{rank}}(\mathcal{C})$. By Corollary \[rank2\] we must have that ${\operatorname{rank}}(\mathcal{C}) \geq {\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}}) + |\mathcal{G}(\mathcal{C})| + 2p - 3 $ for all odd prime $p$ that divides $|\mathcal{G}(\mathcal{C}_{\mathrm{ad}})|$. From this and Lemma \[Cadpt no trivial\] we conclude that the following are all the possible options for $|\mathcal{G(C)}|$: 1. If ${\operatorname{rank}}(\mathcal{C})=13$, then $|\mathcal{G(C)}|= 3$ or $1$. 2. If ${\operatorname{rank}}(\mathcal{C})=15, 17, 19, 21$ or $23$, then $|\mathcal{G(C)}|= 9, 5, 3$ or $1$. We proceed discarding, case by case, all the possibilities stated above for ${\operatorname{rank}}(\mathcal{C})=13, 15$ and all possibilities stated above, besides $|\mathcal{G(C)}| = 1$, for ${\operatorname{rank}}(\mathcal{C})=17, 19, 21, 23$. Let $\mathcal{C}$ be a MNSD modular tensor category of rank 13. Then $\mathcal{C}$ is pointed. By Remark , it is enough to discard the possibilities $|\mathcal{G(C)}|=3$ and $1$. Recall that by Lemma \[Cadpt no trivial\] $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is not trivial if $|\mathcal {G(C)}|\ne 1$. Assume $|\mathcal{G(C)}|= 3$. Note that $\mathcal{C}_\mathrm{pt} \subseteq \mathcal{C}_{\mathrm{ad}}$ and ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to 3. Hence, there must exist a simple non-invertible element in $\mathcal{C}_{\mathrm{ad}}$, and as $\mathcal{C}$ is MNSD the rank of $\mathcal{C}_{\mathrm{ad}}$ is at least five. Moreover, by Lemma $\ref{rank1}$ there exists $g \in \mathcal{G(C)}\simeq \mathbb{Z}_3$ such that $ 3 \leq {\operatorname{rank}}(\mathcal{C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Thus, $\mathcal{C}_{\mathrm{ad}}$ has rank either 5 or 7, and both cases are discarded by Remark \[cong 8\]. Assume now that $|\mathcal{G(C)}|=1$. We will denote the non-invertible simple objects in $\mathcal{C}$ by $X_1, X_1^*, \cdots, X_6, X_6^*$, and their respective Frobenius-Perron dimensions by ${\operatorname{d}}_1, \cdots, {\operatorname{d}}_6$. Up to relabeling the simple objects, we have that ${\operatorname{d}}_1 \geq {\operatorname{d}}_2 \geq \cdots \geq {\operatorname{d}}_6$. Hence, $$\begin{aligned} \label{h1} \dim(\mathcal{C})= 1 + 2 {\operatorname{d}}_1^2 + \cdots + 2 {\operatorname{d}}_6^2 \leq 1 + 12 {\operatorname{d}}_1^2.\end{aligned}$$ On the other hand, by [@ENO2 Theorem 2.11] there exists an odd integer $l$ such that $\dim(\mathcal{C}) = l {\operatorname{d}}_1^2$. Equation implies that $l\leq 12$, and therefore $l=5$ (see Lemma \[dim cong rank\]). Consequently, $$\begin{aligned} \label{h2} 3{\operatorname{d}}_1^2= 1 + 2 {\operatorname{d}}_2^2 + \cdots + 2 {\operatorname{d}}_6^2 \leq 1 + 10 {\operatorname{d}}_2^2.\end{aligned}$$ Again, by [@ENO2 Theorem 2.11], we know that ${\operatorname{d}}_2^2$ divides $\dim(\mathcal{C})= 5{\operatorname{d}}_1^2,$ and so there exists an odd integer $m$ such that $ {\operatorname{d}}_1^2 = m^2 {\operatorname{d}}_2^2$. Equation implies that $m=1$, that is, ${\operatorname{d}}_1= {\operatorname{d}}_2$ and $$\begin{aligned} \label{h3} {\operatorname{d}}_2^2= 1 + 2 {\operatorname{d}}_3^2 + \cdots + 2 {\operatorname{d}}_6^2 \leq 1 + 8 {\operatorname{d}}_3^2.\end{aligned}$$ By the same argument as before, there exists an odd integer $n$ such that ${\operatorname{d}}_2^2= n^2 {\operatorname{d}}_3^2$, and Equation implies that $n=1$. Hence, $$\begin{aligned} {\operatorname{d}}_3^2= 1 + 2 {\operatorname{d}}_3^2 + \cdots + 2 {\operatorname{d}}_6^2 ,\end{aligned}$$ which is a contradiction. Let $\mathcal{C}$ be a MNSD modular tensor category of rank 15. Then $\mathcal{C}$ is pointed. By Remark it is enough to discard the possibilities $|\mathcal{G(C)}|= 9, 5, 3$ and $1$. Recall that, by Lemma \[Cadpt no trivial\], $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is not trivial if $|\mathcal {G(C)}|\ne 1$. Case $|\mathcal{G(C)}|= 9$: since $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is not trivial its Frobenius-Perron dimension must be at least 3. As ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to 3, the rank of $\mathcal{C}_{\mathrm{ad}}$ is at least five. Thus, this case is discarded by Corollary \[rank2\], taking $p=3.$ Case $|\mathcal{G(C)}|= 5$: here $\mathcal{C}_\mathrm{pt} \subset \mathcal{C}_{\mathrm{ad}}$. As ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to five, the rank of $\mathcal{C}_{\mathrm{ad}}$ is at least seven. Thus, this case is discarded by Corollary \[rank2\], taking $p=5.$ Case $|\mathcal{G(C)}|= 3$: again, $\mathcal{C}_\mathrm{pt} \subset \mathcal{C}_{\mathrm{ad}}$ and the rank of $\mathcal{C}_{\mathrm{ad}}$ must be at least five. Moreover, by Lemma $\ref{rank1}$ there exists $g \in \mathcal{G(C)}$ such that $ 3 \leq {\operatorname{rank}}({\mathcal C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Thus, $\mathcal{C}_{\mathrm{ad}}$ has rank either $5, 7$ or $9$, and the last two are discarded by Remark \[cong 8\]. Consider now the case where ${\operatorname{rank}}(\mathcal C_{\mathrm{ad}})=5$. By Remark \[cong 8\], we get that ${\operatorname{rank}}(\mathcal C_g)=5$ for all $g\ne 1$, which is a contradiction by Proposition \[rank Cg\]. Lastly, assume $|\mathcal {G(C)}|=1$. We will denote the simple non-invertible objects in $\mathcal{C}$ by $X_1, X_1^*, \cdots, X_7, X_7^*$ and their respective Frobenius-Perron dimensions by ${\operatorname{d}}_1, \cdots, {\operatorname{d}}_7$. Relabel the simple objects so that ${\operatorname{d}}_1 \geq {\operatorname{d}}_2 \geq \cdots \geq {\operatorname{d}}_7$. Hence, $$\begin{aligned} \label{h4} \dim(\mathcal{C})= 1 + 2 {\operatorname{d}}_1^2 + \cdots + 2 {\operatorname{d}}_7^2 \leq 1 + 14 {\operatorname{d}}_1^2.\end{aligned}$$ On the other hand, by [@ENO2 Theorem 2.11], there exists an odd integer $l$ such that $\dim(\mathcal{C}) = l {\operatorname{d}}_1^2$. Equation implies that $l\leq 14$, and therefore $l=7$ (see Lemma \[dim cong rank\]). Consequently, $$\begin{aligned} \label{h5} 5{\operatorname{d}}_1^2= 1 + 2 {\operatorname{d}}_2^2 + \cdots + 2 {\operatorname{d}}_6^2 + 2 {\operatorname{d}}_7^2 \leq 1 + 12 {\operatorname{d}}_2^2.\end{aligned}$$ Again, by [@ENO2 Theorem 2.11], we know that ${\operatorname{d}}_2^2$ divides $\dim(\mathcal{C})= 7{\operatorname{d}}_1^2$, and so there exists an odd integer $m$ such that $ {\operatorname{d}}_1^2 = m^2 {\operatorname{d}}_2^2$. Equation \[h5\] implies that $m=1$, that is, ${\operatorname{d}}_1= {\operatorname{d}}_2$ and $$\begin{aligned} \label{h6} 3 {\operatorname{d}}_2^2= 1 + 2 {\operatorname{d}}_3^2 + \cdots + 2 {\operatorname{d}}_6^2 + 2 {\operatorname{d}}_7^2\leq 1 + 10 {\operatorname{d}}_3^2.\end{aligned}$$ By the same argument as before, there exists an odd integer $n$ such that ${\operatorname{d}}_2^2= n^2 {\operatorname{d}}_3^2$, and equation implies that $n=1$. Hence, $$\begin{aligned} \label{h7} {\operatorname{d}}_3^2= 1 + 2 {\operatorname{d}}_4^2 + \cdots + 2 {\operatorname{d}}_6^2 + 2 {\operatorname{d}}_7^2\leq 1 + 8 {\operatorname{d}}_4^2,\end{aligned}$$ Once again, there exists an odd integer $q$ such that ${\operatorname{d}}_3^2 = q^2 {\operatorname{d}}_4^2$, and equation implies $q=1$. Therefore, $$\begin{aligned} {\operatorname{d}}_4^2= 1 + 2 {\operatorname{d}}_4^2 + \cdots + 2 {\operatorname{d}}_6^2 + 2 {\operatorname{d}}_7^2\end{aligned}$$ which is a contradiction Let $\mathcal{C}$ be a MNSD modular tensor category of rank 17. Then $\mathcal{C}$ is pointed or $\mathcal{C}_\mathrm{pt}\simeq \text{Vec}$. By Remark , it is enough to discard the cases $|\mathcal{G(C)}|= 9,5$ and $3$. Recall that, by Lemma \[Cadpt no trivial\], $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is not trivial if $|\mathcal {G(C)}|\ne 1$. Case $|\mathcal{G(C)}|= 9$: Since $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is not trivial its Frobenius-Perron dimension must be at least 3. As ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to 3, the rank of $\mathcal{C}_{\mathrm{ad}}$ is at least five. Thus, this case is discarded by Corollary \[rank2\] taking $p=3$. Case $|\mathcal{G(C)}|= 5$: here $\mathcal{C}_\mathrm{pt} \subset \mathcal{C}_{\mathrm{ad}}$. As ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to five, the rank of $\mathcal{C}_{\mathrm{ad}}$ is at least seven. Thus, this case is discarded by Corollary \[rank2\] taking $p=5$. Case $|\mathcal{G(C)}|= 3$: as $\mathcal{C}_\mathrm{pt} \subset \mathcal{C}_{\mathrm{ad}}$, the rank of $\mathcal{C}_{\mathrm{ad}}$ is at least five. Moreover, by Lemma \[rank1\] there exists $g \in \mathcal{G(C)}\simeq \mathbb{Z}_3$ such that $ 3 \leq {\operatorname{rank}}({\mathcal C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Thus, $\mathcal{C}_{\mathrm{ad}}$ has rank either 5, 7 , 9 or 11. The first three cases are discarded by Remark \[cong 8\]. Assume ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})=11$. We denote the non-invertible objects in $\mathcal{C}_{\mathrm{ad}}$ by $X_1, X_1^*, \cdots, X_4, X_4^*, $ and the invertible ones by $\textbf{1}, g, g^2$. Note that since $|\mathcal{G(C)}| = 3$, the action of $\mathcal{G(C)}$ by left multiplication on $\{X_1, X_1^*, \cdots, X_4, X_4^*\}$ has $2$ or $8$ fixed elements. Lets consider first the case in which there are exactly $2$ fixed elements by the action. That is, we have that (up to relabeling) the simple objects $X_1$ and $X_1^*$ are the only simple objects fixed by the action. Denote by ${\operatorname{d}}_i$ the Frobenius-Perron dimensions of the objects $X_i$ and $X_i^*$ for all $i$. It is easy to see that since $X_2, X_2^*, X_3, X_3^*, X_4, X_4^*$ are not fixed by the action, we have that ${\operatorname{d}}:={\operatorname{d}}_{2}= {\operatorname{d}}_{3} = {\operatorname{d}}_{4}$. Thus, $$\dim(\mathcal{C})= \dim(\mathcal{C}_\mathrm{pt}) \dim(\mathcal{C}_{\mathrm{ad}}) = 9 + 6 {\operatorname{d}}_{1}^2 + 18 {\operatorname{d}}^2.$$ Hence, $\gcd({\operatorname{d}}_{1}, {\operatorname{d}})=1,3$. Assume $\gcd({\operatorname{d}}_{1}, {\operatorname{d}})=3$, and consider the decomposition $$X_2 \otimes X_2^* = \textbf{1} \oplus N_{X_2 X_2^*}^{X_1} X_1 \oplus \cdots \oplus N_{X_2X_2^*}^{X_4^*} X_4^*.$$ Taking dimensions on both sides, we get $${\operatorname{d}}^2= 1 + {\operatorname{d}}_{1} (N_{X_2 X_2^*}^{X_1} + N_{X_2 X_2^*}^{X_1^*}) + {\operatorname{d}}(N_{X_2 X_2^*}^{X_2}+ \cdots+ N_{X_2 X_2^*}^{X_4^*}),$$ and thus 3 divides 1, which is a contradiction. Consequently, $\gcd({\operatorname{d}}_{1}, {\operatorname{d}}) =1$. Let $ Y \in \{X_2, X_2^*, \cdots, X_4^*\}$. Consider the decomposition $$X_1 \otimes Y= N_{X_1 Y}^{X_1} X_1 \oplus \cdots \oplus N_{X_1 Y}^{X_4^*} X_4^*.$$ Notice that neither $g$ nor $g^2$ are subobjects of $X_1\otimes Y$ since $g$ fixes $X_1$ and $Y \not\simeq X_1^*$. Taking dimensions on both sides on the previous equation we get $${\operatorname{d}}_{1}{\operatorname{d}}= {\operatorname{d}}_{1} (N_{X_1 Y}^{X_1} + N_{X_1 Y}^{X_1^*}) + {\operatorname{d}}(N_{X_1 Y}^{X_2}+ \cdots N_{X_1 Y}^{X_4^*}).$$ Thus, ${\operatorname{d}}_{1} $ divides $N_{X_1 Y}^{X_2}+ \cdots +N_{X_1 Y}^{X_4^*}$ and ${\operatorname{d}}$ divides $N_{X_1 Y}^{X_1} + N_{X_1 Y}^{X_1^*} $. Since $\gcd({\operatorname{d}}_1, {\operatorname{d}})=1$, either ${\operatorname{d}}_{1} = N_{X_1 Y}^{X_2}+ \cdots + N_{X_1 Y}^{X_4^*}$ and $N_{X_1 Y}^{X_1} + N_{X_1 Y}^{X_1^*}=0$ or ${\operatorname{d}}= N_{X_1 Y}^{X_1} + N_{X_1 Y}^{X_1^*}$ and $N_{X_1 Y}^{X_2}+ \cdots + N_{X_1 Y}^{X_4^*}=0$. Assume the latter is true for some $Y\in \{X_2, \dots, X_4^*\}$. Then $N_{X_1 Y}^{X_2}= \cdots = N_{X_1 Y}^{X_4^*}=0$. In particular, $N_{X_1Y}^Y=0$ and so by the fusion rules we have that $N_{Y Y^*}^{X_1}=0$. Thus $$Y \otimes Y^* = \textbf{1} \oplus N_{Y Y^*}^{X_2} X_2 \oplus\dots \oplus N_{Y Y^*}^{X_4^*} X_4^*,$$ and taking dimensions on both sides, we get $${\operatorname{d}}^2=1 + {\operatorname{d}}( N_{Y Y^*}^{X_2} +\dots + N_{Y Y^*}^{X_4^*}),$$ and so ${\operatorname{d}}$ divides 1, a contradiction. Therefore ${\operatorname{d}}_{1} = N_{X_1 Y}^{X_2}+ \cdots + N_{X_1 Y}^{X_4^*}$ and $N_{X_1 Y}^{X_1} + N_{X_1 Y}^{X_1^*}=0$ for all $Y\in \{X_2, \dots, X_4^*\}$. Consequently, $N_{X_1 Y}^{X_1} = N_{X_1 Y}^{X_1^*}=0$ for all $Y\in \{X_2, \dots, X_4^*\}$, and so by the fusion rules we get that $N_{X_1 X_1^*}^{Y} = N_{X_1 X_1^*}^{Y^*}=0$ for all $Y\in \{X_2, \dots, X_4^*\}$. Thus, $$X_1 \otimes X_1^* = \textbf{1} \oplus g \oplus g^2 \oplus N_{X_1 X_1^*}^{X_1} X_1 \oplus N_{X_1 X_1^*}^{X_1^*} X_1^*,$$ which implies that ${\operatorname{d}}_{1}=3$. Hence, $\dim(\mathcal{C})= 9 + 3^26 + 18 {\operatorname{d}}^2. $ By [@ENO2 Theorem 2.11], we get that ${\operatorname{d}}^2$ divides $3^27$. Thus, ${\operatorname{d}}^2= 9$, which is a contradiction as $\gcd({\operatorname{d}}_{1}, {\operatorname{d}})=1$ and ${\operatorname{d}}_1 = 3$. Lastly, we consider the case in which all simple non-invertible elements in $\mathcal{C}_{\mathrm{ad}}$ are fixed by $\mathbb{Z}_3$. Recall that $\mathcal{C}= \mathcal{C}_{\mathrm{ad}} \oplus \mathcal{C}_g \oplus \mathcal{C}_{g^2}$, where ${\operatorname{rank}}(\mathcal{C}_{g})={\operatorname{rank}}(\mathcal{C}_{g^2}) =3$. We denote the simple elements of $\mathcal{C}_{g}$ by $Y_1, Y_2, Y_3$ and their respective Frobenius-Perron dimensions by ${\operatorname{d}}_{Y_1}, {\operatorname{d}}_{Y_2}, {\operatorname{d}}_{Y_3}$. Note that the simple elements of $\mathcal{C}_{g}$ are exactly $Y_1^*, Y_2^*, Y_3^*$, and thus by Corollary \[intersection\] the action of $\mathcal{G(C)}\simeq \mathbb{Z}_3$ by left multiplication on $\{Y_1, Y_2, Y_3\}$ must be non-trivial. We may relabel the simples so that $g \otimes Y_1 = Y_2$ and $g^2\otimes Y_1 = Y_3$. So ${\operatorname{d}}_{Y_1}={\operatorname{d}}_{Y_2}={\operatorname{d}}_{Y_3}=:{\operatorname{d}}$. Now, for all $i=1, \cdots, 4$, we have that $X_i\otimes Y_1 \in \mathcal{C}_{g}$, so $$\begin{aligned} \label{e1} X_i \otimes Y_1 = N_{X_i Y_1}^{Y_1} Y_1 \oplus N_{X_i Y_1}^{Y_2} Y_2 \oplus N_{X_i Y_1}^{Y_3} Y_3.\end{aligned}$$ On the other hand, $$\begin{aligned} &X_i \otimes Y_1 = g \otimes X_i \otimes Y_1= N_{X_i Y_1}^{Y_1} Y_2 \oplus N_{X_i Y_1}^{Y_2} Y_3 \oplus N_{X_i Y_1}^{Y_3} Y_1,\label{e2}\\ &X_i \otimes Y_1 = g^2 \otimes X_i \otimes Y_1= N_{X_i Y_1}^{Y_1} Y_3 \oplus N_{X_i Y_1}^{Y_2} Y_1 \oplus N_{X_i Y_1}^{Y_3} Y_2.\label{e3}\end{aligned}$$ From equations , , we get that $N_{X_i Y_1}^{Y_1} =N_{X_i Y_1}^{Y_2}= N_{X_i Y_1}^{Y_3} $, hence $$\begin{aligned} X_i \otimes Y_1 = N_{X_i Y_1}^{Y_1} (Y_1 \oplus Y_2 \oplus Y_3).\end{aligned}$$ Consequently, ${\operatorname{d}}_{X_i} = 3 N_{X_i Y_1}^{Y_1} $. So, 3 divides ${\operatorname{d}}_{X_i}$ for all $i=1,\cdots, 4$. Let $c_{X_i}= {\operatorname{d}}_{X_i}/3.$ Note that $ \dim(\mathcal{C})= 3 \dim(\mathcal{C}_{\mathrm{ad}})= 3 \dim(\mathcal{C}_g) = 9 {\operatorname{d}}^2.$ As ${\operatorname{d}}_{X_i}^2 $ divides $\dim(\mathcal{C}),$ we get $c_{X_i}^2$ divides ${\operatorname{d}}^2.$ Reordering the indices so that $c_{X_1} \geq c_{X_2} \geq c_{X_3} \geq c_{X_4},$ and letting $l$ be an odd integer such that ${\operatorname{d}}^2= l^2 c_{X_1}^2$, we get that $$\begin{aligned} \label{e4} 3 + 2 {\operatorname{d}}_{X_1}^2 + \cdots + 2 {\operatorname{d}}_{X_4}^2= \dim(\mathcal{C}_{\mathrm{ad}})= \dim(\mathcal{C}_g) = 3 {\operatorname{d}}^2. \end{aligned}$$ Dividing each side of equation by 3, we get $$l^2 c_{X_1}^2 ={\operatorname{d}}^2= 1+ 6 c_{X_1}^2 + \cdots + 6 c_{X_4}^2 \leq 1 + 24 c_{X_1}^2.$$ Hence, $l^2 \leq 24$, and so $l^2=1$ or $9$. If $l^2=9,$ then 9 divides ${\operatorname{d}}^2$, and as $9$ also divides ${\operatorname{d}}_{X_1}^2, \cdots, {\operatorname{d}}_{X_4}^2$, by equation $\eqref{e4}$ we have that $9$ divides 3. Consequently, $l^2=1$, i.e, ${\operatorname{d}}^2 = c_{X_1}^2, $ and ${\operatorname{d}}_{X_1}^2 = 9 {\operatorname{d}}^2= \dim(\mathcal{C})= 3 \dim(\mathcal{C}_{\mathrm{ad}})= 9 + 6 {\operatorname{d}}_{X_1}^2 + \cdots+ 6 {\operatorname{d}}_{X_4}^2$, which is again a contradiction. Let $\mathcal{C}$ be a MNSD modular tensor category of rank 19. Then $\mathcal{C}$ is either pointed or $\mathcal{C}_\mathrm{pt}\simeq \text{Vec}$. By Remark \[casos descartados\], it is enough to discard the cases $|\mathcal{G(C)}|= 9,5,$ and $3$. Recall that, by Lemma \[Cadpt no trivial\], $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is not trivial if $|\mathcal {G(C)}|\ne 1$. Case $|\mathcal{G(C)}|= 5$: here, $\mathcal{C}_\mathrm{pt} \subset \mathcal{C}_{\mathrm{ad}}$. As ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to five, the rank of $\mathcal{C}_{\mathrm{ad}}$ is at least seven. Thus, this case is discarded by Remark \[cong 8\]. Case $|\mathcal{G(C)}|= 3$ or $9$: as ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to 3, the rank of $\mathcal{C}_{\mathrm{ad}}$ is at least 5. By Lemma \[rank1\] there exists $g \in \mathcal{G(C)}$ such that $ 3 \leq {\operatorname{rank}}(\mathcal{C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Hence, ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 5, 7, 9, 11 $ or $13$, and all cases are discarded by Remark \[cong 8\]. Let $\mathcal{C}$ be a MNSD modular tensor category of rank 21. Then $\mathcal{C}$ is either pointed or $\mathcal{C}_\mathrm{pt}\simeq \text{Vec}$. By Remark \[casos descartados\], it is enough to discard the cases $|\mathcal{G(C)}|= 15, 9, 5,$ and $3$. Recall that $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is not trivial if $|\mathcal {G(C)}|\ne 1$ by Lemma \[Cadpt no trivial\]. Case $|\mathcal{G(C)}|= 5$: here, $\mathcal{C}_\mathrm{pt}\subseteq \mathcal C_{\mathrm{ad}}$, and since ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to 5 we get that ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})\geq 7$. By Lemma \[rank1\] there exists $g \in \mathcal{G(C)}$ such that $ 5 \leq {\operatorname{rank}}(\mathcal{C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Therefore ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 7$ or $9$, and both cases are discarded by Remark \[cong 8\]. Case $|\mathcal{G(C)}|= 9$: since ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to 3 we get that ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})\geq 5$. By Lemma \[rank1\] there exists $g \in \mathcal{G(C)}$ such that $ 3 \leq {\operatorname{rank}}(\mathcal{C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Therefore ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 5, 7$ or $9$, and all cases are discarded by Remark \[cong 8\]. Case $|\mathcal{G(C)}|= 3$: as ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to 3 we have that ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})\geq 5$. By Lemma \[rank1\] there exists $g \in \mathcal{G(C)}$ such that $ 3 \leq {\operatorname{rank}}(\mathcal{C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Hence, ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 5, 7, 9, 11, 13$ or $15$, and all cases but ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 7$ are discarded by Remark \[cong 8\]. If ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 7$ then ${\operatorname{rank}}(\mathcal C_g)=7$ for all $g\in \mathcal {G(C)}$, which is a contradiction by Proposition \[rank Cg\]. Let $\mathcal{C}$ be a MNSD modular tensor category of rank 23. Then $\mathcal{C}$ is either pointed or $\mathcal{C}_\mathrm{pt}\simeq \text{Vec}$. By Remark \[casos descartados\], it is enough to discard the cases $|\mathcal{G(C)}|= 15, 9, 5$, and $3$. Recall that by Lemma \[Cadpt no trivial\] $(\mathcal{C}_{\mathrm{ad}})_{\mathrm{pt}}$ is not trivial if $|\mathcal {G(C)}|\ne 1$. Case $|\mathcal{G(C)}|= 9$: since ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to 3 we have that ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})\geq 5.$ By Lemma \[rank1\] there exists $g \in \mathcal{G(C)}$ such that $ 3 \leq {\operatorname{rank}}(\mathcal{C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Hence, ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 5, 7, 9$ or $11$, and all cases are discarded by Remark \[cong 8\]. Case $|\mathcal{G(C)}|= 5$: since ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to five we get that $\mathcal{C}_{\mathrm{ad}}\geq 7.$. By Lemma \[rank1\] there exists $g \in \mathcal{G(C)}$ such that $ 5 \leq {\operatorname{rank}}(\mathcal{C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Hence, ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 7, 9$ or $11$, and all cases are discarded by Remark \[cong 8\]. Case $|\mathcal{G(C)}|= 3$: since ${\operatorname{FPdim}}(\mathcal{C}_{\mathrm{ad}})$ cannot be equal to three, the rank of $\mathcal{C}_{\mathrm{ad}}$ is at least five. By Lemma \[rank1\] there exists $g \in \mathcal{G(C)}$ such that $ 3 \leq {\operatorname{rank}}(\mathcal{C}_g) = {\operatorname{rank}}(\mathcal{C}_{g^{-1}})$. Hence, ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 5, 7, 9, 11, 13, 15$, and all cases but ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 13$ are discarded by Remark \[cong 8\]. Now, if ${\operatorname{rank}}(\mathcal{C}_{\mathrm{ad}})= 13$ then ${\operatorname{rank}}(\mathcal{C}_g)=5$ for $g\in \mathcal {G(C)}$ such that $g\ne 1$, which is a contradiction by Proposition \[rank Cg\]. Future Directions {#Future directions} ================= As a consequence of our results in Section \[section: mnsd low rank\], a non-pointed MNSD modular tensor category of rank between 17 and 23 cannot be weakly group-theoretical . This follows since for weakly group-theoretical fusion categories there is a version of the Feit-Thompson theorem: if a weakly group-theoretical fusion category is odd-dimensional then it is solvable [@NP Proposition 7.1]. It is known that solvable fusion categories contain non-trivial invertible objects [@ENO2 Proposition 4.5]. This result is relevant because if there exists a non-pointed MNSD modular tensor category of odd rank between 17 and 23 then it must be non-weakly group-theoretical but also odd-dimensional (in particular, weakly integral). This would be a counter-example for an important conjecture in fusion categories that states that every weakly integral fusion category is weakly group-theoretical [@ENO1]. In fact, any perfect odd-dimensional modular tensor category would yield a counter-example for said conjecture. We conjecture: \[conjecture: odd dim non perfect\] Odd-dimensional modular tensor categories have at least one non-trivial invertible object, i.e. they cannot be perfect. By the argument explained above using Feit-Thompson for weakly group-theoretical fusion categories, this conjecture would be an immediate consequence of the weakly integral conjecture. Notice that if Conjecture \[conjecture: odd dim non perfect\] is false, then we can construct an odd-dimensional modular tensor category that is neither perfect nor weakly group-theoretical via the Deligne product of an odd-dimensional perfect modular tensor category and an odd-dimensional pointed modular tensor category. Another question is if it is possible to remove the hypothesis of weakly group-theoretical in the fusion categorical version of Feit-Thompson’s Theorem. Equivalent statements to Conjecture \[conjecture: odd dim non perfect\] are the following[^1]. \[Conjecture: odd fusion are solvable\] Odd-dimensional fusion categories are solvable. \[Conjecture:odd modular are solvable\] Odd-dimensional modular tensor categories are solvable. Note that Conjectures \[Conjecture: odd fusion are solvable\] and \[Conjecture:odd modular are solvable\] are equivalent by [@ENO2 Proposition 4.5]. In fact, if $\mathcal C$ is an odd-dimensional fusion category, then its Drinfeld center $\mathcal {Z(C)}$ is an odd-dimensional modular tensor category. Assuming Conjecture \[Conjecture:odd modular are solvable\] this would imply that $\mathcal{Z(C)}$ is solvable, and since it is Morita equivalent to $\mathcal C \boxtimes \mathcal C^{\operatorname{op}}$ we conclude that $\mathcal C$ is also solvable [@ENO2 Proposition 4.5]. On the other hand, if we assume Conjecture \[conjecture: odd dim non perfect\] then Conjecture \[Conjecture:odd modular are solvable\] is also true as follows. Let $\mathcal C$ be a modular tensor category of odd dimension. We prove our implication by induction on $\dim(\mathcal C).$ Let $g$ be a non-trivial invertible object in $\mathcal C$; we may assume that the order of $g$ is prime. Consider the fusion subcategory $\mathcal C[g]$. If $\mathcal C[g]$ is not Tannakian then it must be modular and $\mathcal C= \mathcal C \boxtimes \mathcal C[g]'$. Since $\mathcal C[g]'$ is a modular tensor category of dimension strictly less than $\dim(\mathcal C)$ then by induction it is solvable and thus $\mathcal C$ is solvable [@ENO2 Proposition 4.5]. On the other hand, if $\mathcal C[g]$ is Tannakian then $\mathcal C[g]$ is equivalent to ${\operatorname{Rep}}(G)$ for $G=\langle g\rangle$, and so the trivial component $(\mathcal C_G)_0$ of the de-equivariantization $\mathcal C_G$ of $\mathcal C$ is a modular tensor category of dimension strictly less than $\dim(\mathcal C)$ and hence solvable by induction. It follows that $\mathcal C$ is also solvable by [@ENO2 Proposition 4.5]. Lastly, Conjecture \[conjecture: odd dim non perfect\] follows from Conjecture \[Conjecture:odd modular are solvable\] and [@ENO2 Proposition 4.5]. By an analogous reasoning, we have that Conjecture $\ref{conjecture: odd dim non perfect}$ is equivalent to [@ENO2 Question 2] for the odd-dimensional case, namely, odd-dimensional weakly integral fusion categories are weakly group theoretical. Notice that given the previous equivalences, the assumption of modularity in Conjecture \[conjecture: odd dim non perfect\] can be dropped. [AAAAAAAAA]{} *Lectures on tensor categories and modular functors.* University Series Lectures, vol. **21**, AMS (2001). . *Catègories prèmodulaires, modularisations et invariants des variétés de dimension 3*. Mathematische Annalen **316** (2000), no. 2, 215-236. . *Classification of integral modular categories of Frobenius-Perron dimension $pq^4$ and $p^2q^2$*. Canad. Math. Bull. **57** (2014), no. 4, 721-734. . *Dimension as a quantum statistic and the classification of metaplectic categories.* Preprint `arXiv 1909.09843`, to appear in J. Algebra Appl. . *Modular categories, integrality and egyptian fractions*. Proc. Amer. Math. Soc. **140** (2012), no. 4, 1141-1150. . *Integral almost square-free modular categories*. J. Algebra Appl. **16** (2017), no. 5, 1750104. . *On the classification of almost square free modular categories*. Alg. and Rep. Th. **21** (2018), no. 6, 1353-1368. . *Integral modular categories of Frobenius-Perron dimension $pq^n$*. Alg. and Rep. Th. **19** (2016), 33-46. . *Group-theoretical properties of nilpotent modular categories*. Preprint `arXiv 0704.0195` (2007). . *On braided fusion categories I*. Selecta Math. **16** (2010), no. 1, 1-119. . *Tensor categories*. Math. Surv. Monog. **205** (2015), AMS. . *Classification of fusion categories of dimension $pq$*. Int. Math. Res. Not. **57** (2004), 3041-3056. . *On fusion categories*. Annals of Mathematics **162** (2005), 581-642. . *Weakly group-theoretical and solvable fusion categories*. Advances in Mathematics **226** (2011), no. 1, 176-205. . *Nilpotent fusion categories*. Advances in Mathematics **217** (2008), no. 3, 1053-1071. . *On the classification of the Grothendieck semirings of non-self-dual modular categories*. J. Algebra **324** (2010), 1000-1015. . *On the classification of certain fusion categories*. J. Noncommut. Geom. **3** (2009), no. 3, 481-499. . *On the Structure of Modular Categories*. Proc. London Math. Soc. **87** (2003), 291-308. . *A finiteness property for braided fusion categories.* Algebr. Represent. Theory. **14** (2011), no. 5 837-855. . *On weakly group-theoretical non-degenerate braided fusion categories.* J. Noncommut. Geom. **8** (2014), 1043-1060. . *The core of a weakly group-theoretical braided fusion category*. Int. J. Math. **29** (2018), no. 2, Article ID 1850012, 23 p. . *Graphs attached to simple Frobenius-Perron dimensions of an integral fusion category.* Monatsh. Math. **179** (2016), no. 4, 615-649. . *On fusion categories with few irreducible degrees.* Algebra Number Theory **6** (2012), 1171-1197. . *Frobenius-Schur Indicators and Exponents of Spherical Categories*. Adv. Math. **211** (2007), no. 1, 34–71. , *On classification of modular tensor categories*, Comm. Math. Phys. **292** (2009), no. 2, 343-389. [^1]: The equivalence of these statements was pointed out to us by C. Galindo
{ "pile_set_name": "ArXiv" }
--- abstract: 'The eigenvalue problem for $3\times3$ octonionic Hermitian matrices contains some surprises, which we have reported elsewhere [@Eigen]. In particular, the eigenvalues need not be real, there are 6 rather than 3 real eigenvalues, and the corresponding eigenvectors are not orthogonal in the usual sense. The nonassociativity of the octonions makes computations tricky, and all of these results were first obtained via brute force (but exact) [[*Mathematica*]{}]{} computations. Some of them, such as the computation of real eigenvalues, have subsequently been implemented more elegantly; others have not. We describe here the use of [[*Mathematica*]{}]{} in analyzing this problem, and in particular its use in proving a generalized orthogonality property for which no other proof is known.' address: - | Department of Mathematics, Oregon State University, Corvallis, OR 97331, USA\ [tevian[@]{}math.orst.edu]{}\ - | Department of Physics, Oregon State University, Corvallis, OR 97331, USA\ [corinne[@]{}physics.orst.edu]{}\ author: - Tevian Dray - 'Corinne A. Manogue' date: 17 June 1998 title: | **Finding Octonionic Eigenvectors\ Using [[*Mathematica*]{}]{}** --- octonions, eigenvectors, Mathematica 15A33, 15A18, 17A35, 17C90 02.70.Rw INTRODUCTION ============ Finding the eigenvalues and eigenvectors of a given matrix is one of the basic techniques in linear algebra, with countless applications. The familiar case of Hermitian (complex) matrices is very important, for instance in quantum mechanics, where the fact that such matrices have real eigenvalues allows them to represent physically observable quantities. The eigenvalue problem is usually formulated over the complex numbers ${{\mathbb C}}$, including the reals ${{\mathbb R}}$ as a special case. In recent work [@Eigen], we considered the generalization to the other normed division algebras, namely the quaternions ${{\mathbb H}}$ and the octonions ${{\mathbb O}}$. Most of the basic properties are retained, provided they are reinterpreted to take into account the lack of commutativity of ${{\mathbb H}}$ and ${{\mathbb O}}$, and the lack of associativity of ${{\mathbb O}}$. However, there are a number of surprises, including the fact that such matrices admit non-real eigenvalues. Our most important results concern the eigenvalue problem for $3\times3$ octonionic Hermitian matrices, the [*Jordan matrices*]{}. It turns out [@Eigen] that such matrices admit more than the expected 3 real eigenvalues, and that eigenvectors corresponding to different eigenvalues fail in general to be orthogonal in the usual sense, although they do seem to be orthogonal in a generalized sense. Because of the lack of both commutativity and associativity, working with octonionic matrices is rather tricky. All of the above results were initially discovered using a [[*Mathematica*]{}]{} package we have developed over the years for just this purpose. While we have subsequently been able to derive more elegant derivations by hand for some of these results, we have not succeeded in doing this for all. In particular, the only current proof of the generalized orthogonality property consists of a lengthy, brute force [[*Mathematica*]{}]{} computation, which used 6 hours of CPU time on a Sparc20 with 224 Mb of memory. This article describes both the mathematics behind this result, and the [[*Mathematica*]{}]{}computation used to obtain it. We set the stage in Section \[Background\] by reviewing some basic properties about both the standard eigenvalue problem and the octonions, as well as introducing our [[*Mathematica*]{}]{} package. We then summarize the theory of $3\times3$ octonionic Hermitian matrices in Section \[JordanMatrices\], pointing out that the only known proof of some of the results in this section involves direct computation using [[*Mathematica*]{}]{}. In Section \[Example\] we give an explicit example, and we discuss our results in Section \[Discussion\]. BACKGROUND {#Background} ========== The Standard Eigenvalue Problem ------------------------------- We begin by collecting some standard results about the standard eigenvalue problem. We give the details of some of the proofs in order to emphasize the use of both the commutativity and associativity of ${{\mathbb C}}$. The eigenvalue problem as usually stated is to find solutions $\lambda,v$ to the equation $$\label{Orig} A v = \lambda v$$ for a given square matrix $A$. The basic properties of the eigenvalue problem for [$n \times n$]{} complex Hermitian matrices are well-understood. An [$n \times n$]{} complex Hermitian matrix $A$ has $n$ eigenvalues (counting multiplicity), all of which are real. We give here only the proof that the eigenvalues are real. Let $A$, $v$, $\lambda$ satisfy (\[Orig\]), with $A^\dagger=A$. Then $${\overline}\lambda v^\dagger v = (A v)^\dagger v = v^\dagger A v = \lambda v^\dagger v$$ so that if $v\ne0$ we have $v^\dagger v\ne0$, which forces ${\overline}\lambda=\lambda$. Eigenvectors of an [$n \times n$]{} complex Hermitian matrix $A$ corresponding to different eigenvalues are orthogonal. For $m=1,2$, let $v_m$ be an eigenvector of $A=A^\dagger$ with eigenvalue $\lambda_m$. By the previous lemma, $\lambda_m\in{{\mathbb R}}$. Then $$\lambda_1 v_1^\dagger v_2 = (A v_1)^\dagger v_2 = v_1^\dagger A v_2 = \lambda_2 v_1^\dagger v_2$$ Then either $\lambda_1=\lambda_2$ or $v_1^\dagger v_2=0$. For any [$n \times n$]{} complex Hermitian matrix $A$, there exists an orthonormal basis of ${{\mathbb C}}^n$ consisting of eigenvectors of $A$. If all eigenvalues have multiplicity one, the result follows from the previous lemma. But the Gram-Schmidt orthogonalization process can be used on any eigenspace corresponding to an eigenvalue with multiplicity greater than one. These lemmas are equivalent to the standard result that a complex Hermitian matrix can always be diagonalized by a unitary transformation. It is important for what follows to realize that the form of the proofs given above relies on both the commutativity and the associativity of ${{\mathbb C}}$. Combining the above results, it is easy to see that any (complex) Hermitian matrix $A$ admits a decomposition in terms of an orthonormal basis of eigenvectors. Let $A$ be an [$n \times n$]{} complex Hermitian matrix. Then $A$ can be expanded as $$A = \sum_{m=1}^n \lambda_m v_m v_m^\dagger \label{Decomp}$$ where $\{v_m; ~m=1,...,n\}$ is an orthonormal basis of eigenvectors corresponding to eigenvalues $\lambda_m$. By the previous lemma, there exists an orthonormal basis $\{v_m\}$ of eigenvectors. It then suffices to check that $$\sum_{m=1}^n \lambda_m v_m v_m^\dagger v_k = \lambda_k v_k$$ But this follows by direct computation using orthonormality. Furthermore, the set of eigenvalues $\{\lambda_m\}$ is unique, and the (unit) eigenvectors are unique up to unitary transformations in the separate eigenspaces (which reduce to multiplication by a complex phase for eigenvalues of multiplicity one). Octonions --------- The quaternions ${{\mathbb H}}$ double the dimension of the complex numbers by adding two additional square roots of $-1$, usually denoted $j$ and $k$. The multiplication table follows from $$i^2=j^2=k^2=-1 \qquad ij=k=-ji$$ and associativity; note that ${{\mathbb H}}$ is not commutative. Equivalently, ${{\mathbb H}}$ can be viewed as the sum of 2 copies of the complex numbers $${{\mathbb H}}= {{\mathbb C}}+ k{{\mathbb C}}$$ with $j$ being defined by $j=ki$. The octonions ${{\mathbb O}}$ in turn can be viewed as the direct sum of two copies of the quaternions [^1] $${{\mathbb O}}= {{\mathbb H}}+ {{\mathbb H}}\ell = ({{\mathbb C}}+ k{{\mathbb C}}) + ({{\mathbb C}}+ k{{\mathbb C}})\ell$$ where $\ell$ is yet another square root of $-1$. The octonions are thus spanned by the identity element $1$ and the 7 imaginary units $\{i,j,k,k\ell,j\ell,i\ell,\ell\}$. These units can be grouped into (the imaginary parts of) quaternionic subspaces in 7 different ways; these will be referred to as “triples”. Any three of these imaginary units which do not lie in a such a triple anti-associate. The multiplication table can be neatly summarized by appropriately labeling the 7-point projective plane, as shown in Figure \[Omult\]. [=3.5in]{} Even though the octonions are not associative, since any 2 octonions lie in a quaternionic subspace, products involving only 2 different octonions (and their octonionic conjugates) do associate. For example, $$p(pq) = p^2 q$$ which is a weak form of associativity known as [*alternativity*]{}. We use the notation ${\overline}{a}$ to denote the (octonionic) conjugate of the octonion $a$, $$|a|^2 := a{\overline}{a}$$ to denote the squared norm of $a$, $A^\dagger$ to denote the (octonionic) Hermitian conjugate of the matrix $A$, $$[a,b] := ab - ba$$ to denote the commutator of $a$ and $b$, and $$[a,b,c] := (ab)c - a(bc)$$ to denote the associator of $a$, $b$, $c$. Both the commutator and the associator are purely imaginary, totally antisymmetric, and change sign if any one of their arguments is replaced by its conjugate. Another octonionic product with these properties is given by the [*associative 3-form*]{} [@ReeseHarvey; @Gureirch] $$\Phi(a,b,c) = {1\over2} {{\rm Re}}\Big( a({\overline}{b}c) - c({\overline}{b}a) \Big)$$ which reduces to the vector triple product when $a$, $b$, $c$ are imaginary quaternions. $$\Phi(a,b,c) = {1\over2} \, {{\rm Re}}( [a,{\overline}{b}] c) \label{PhiEq}$$ [[*Mathematica*]{}]{} Package {#Package} ----------------------------- We needed a way to easily manipulate octonions and octonionic matrices — it is quite difficult to unlearn associativity! There are 2 complementary approaches, depending on whether it is desired to manipulate abstract octonions or whether an explicit basis can be used. For our purposes, it was initially quite sufficient to work with an abstract basis: We define an octonion to be a list with 8 elements $$\begin{aligned} a &=& a_1 +a_2 {\,\hbox{\boldmath $i$}}+a_3 {\,\hbox{\boldmath $j$}}+a_4 {\,\hbox{\boldmath $k$}}-a_5 {\,\hbox{\boldmath $k\ell$}}-a_6 {\,\hbox{\boldmath $j\ell$}}-a_7 {\,\hbox{\boldmath $i\ell$}}+a_8 {\,\hbox{\boldmath $\ell$}}\nonumber\\ &=& \{a_1,a_2,a_3,a_4,a_5,a_6,a_7,a_8\}\end{aligned}$$ where the signs are conventional, and where we have emphasized the role of the imaginary units as “basis vectors” by writing them in boldface. Octonionic multiplication can then be expressed as a series of rules of the form $${\,\hbox{\boldmath $i$}}* {\,\hbox{\boldmath $j$}}= {\,\hbox{\boldmath $k$}}$$ where we have introduced the symbol $*$ to represent octonionic multiplication. Octonionic conjugation can be defined as a series of rules of the form $${\overline}{\hbox{\boldmath $i$}} = -{\,\hbox{\boldmath $i$}}$$ and everything else can be defined in terms of these 2 basic operations. [=2.5in]{} 1 Furthermore, as already stated, computations involving small numbers of octonions can be dramatically simplified. For instance, 3 arbitrary octonions can be assumed to take the form $$\begin{aligned} O1 &=& a_1 + a_2 {\,\hbox{\boldmath $i$}}\nonumber\\ O2 &=& b_1 + b_2 {\,\hbox{\boldmath $i$}}+ b_3 {\,\hbox{\boldmath $j$}}\label{Generic}\\ O3 &=& c_1 + c_2 {\,\hbox{\boldmath $i$}}+ c_3 {\,\hbox{\boldmath $j$}}+ c_4 {\,\hbox{\boldmath $k$}}+ c_8 \nonumber{\,\hbox{\boldmath $\ell$}}\end{aligned}$$ When implementing these ideas using [[*Mathematica*]{}]{}, it turned out to be more efficient to define the 2 fundamental operations directly on lists, rather than building them up in terms of rules. For instance, conjugation is more easily defined by $$\begin{aligned} \tt bar[\{x1\_,x2\_,x3\_,x4\_,x5\_,x6\_,x7\_,x8\_\}]:= \nonumber\\ \qquad\qquad\tt\{x1,-x2,-x3,-x4,-x5,-x6,-x7,-x8\}\end{aligned}$$ and there is an analogous definition of octonionic multiplication, called [Omult]{}. A illustration of computation in an explicit basis appears in Figure \[BasicsII\]. Matrices can now be constructed as nested lists, and operations such as matrix multiplication can easily be defined. However, matrix expressions involving several octonions quickly become unwieldy. We therefore supplemented the above basic definitions in terms of an explicit basis with an alternative set of definitions using abstract octonions. In the process, we took full advantage of the formatting capabilities in [[*Mathematica*]{}]{} 3.0. An illustration of computation using the abstract approach appears in Figure \[BasicsI\]. 1 Even though we chose not to implement the concrete version of the basic octonionic operations in terms of rules, [[*Mathematica*]{}]{}’s ability to manipulate such rules was crucial in constructing the abstract version. Especially when teaching [[*Mathematica*]{}]{} to both manipulate and print matrices of lists (i.e.octonions), the ability to easily modify the code to recognize special cases proved extremely helpful. (Octonions are closely related to [*Clifford algebras*]{}; see for instance [@Jorg], in which the representation theory of Clifford algebras is extended so as to employ octonions. There are a number of packages for manipulating Clifford algebras, some of which are described in [@CliffordBook]. Several of these can handle octonions, such as the program CLICAL by Pertti Lounesto [@CLICAL] and the Maple package CLIFFORD by Rafał Abłamowicz [@CLIFFORD], both of which introduce the octonions as paravectors over an appropriate Clifford algebra [@Lounesto2]. There is also an older Maple package Octonion (and the companion package Clifford) by Jörg Schray [@Schray].) $3\times3$ OCTONIONIC HERMITIAN MATRICES {#JordanMatrices} ======================================== It is not immediately obvious that $3\times3$ octonionic Hermitian matrices have a well-defined determinant, let alone a characteristic equation. We therefore first discuss some of the properties of these matrices before turning to the eigenvalue problem. Jordan Matrices --------------- The $3\times3$ octonionic Hermitian matrices, henceforth referred to as the [*Jordan matrices*]{}, form the [*exceptional Jordan algebra*]{} under the [*Jordan product*]{} $$A \circ B := {1\over2} (AB + BA)$$ which is commutative, but not associative. A special case of this is $$A^2 \equiv A \circ A$$ and we [*define*]{} $$A^3 := A^2 \circ A = A \circ A^2$$ Remarkably, with these definitions, Jordan matrices satisfy the usual characteristic equation (see e.g. [@ReeseHarvey]) $$A^3 - ({{\rm tr\,}}A) \, A^2 + \sigma(A) \, A - (\det A) \, I = 0 \label{Char}$$ where $\sigma(A)$ is defined by $$\sigma(A) := {1\over2} \left( ({{\rm tr\,}}A)^2 - {{\rm tr\,}}(A^2) \right)$$ and where the determinant $\det A$ of $A$ is defined abstractly in terms of the Freudenthal product [@Jordan; @Freudenthal] $$\begin{aligned} \nonumber A*B = A \circ B - {1\over2} \Big(A\,{{\rm tr\,}}(B)+B\,{{\rm tr\,}}(A)\Big) + {1\over2} \Big({{\rm tr\,}}(A)\,{{\rm tr\,}}(B)-{{\rm tr\,}}(A\circ B)\Big) \\\end{aligned}$$ which leads to $$\det(A) = {1\over3} \, {{\rm tr\,}}\Big( (A*A) \circ A \Big)$$ Concretely, if $$A = \pmatrix{p& a& {\overline}{b}\cr {\overline}{a}& m& c\cr b& {\overline}{c}& n\cr} \label{Three}$$ with $p,m,n\in{{\mathbb R}}$ and $a,b,c\in{{\mathbb O}}$, then $$\begin{aligned} {{\rm tr\,}}A &=& p + m + n \\ \sigma(A) &=& pm + pn + mn - |a|^2 - |b|^2 - |c|^2 \\ \det A &=& pmn + b(ac) + {\overline}{b(ac)} - n|a|^2 - m|b|^2 - p|c|^2 \label{ThreeEq}\end{aligned}$$ The determinant calculation is illustrated in Figure \[MathMatrices\]. 1 [=5.55in]{} [=5.4in]{} 1 The Real Eigenvalue Problem --------------------------- Each division algebra can be rewritten as a real matrix algebra of appropriate dimension (see e.g. [@Eigen]). Under this identification, a Hermitian matrix over any of the division algebras becomes a real symmetric matrix. It is therefore clear that a $3\times3$ octonionic Hermitian matrix must have $8\times3=24$ real eigenvalues [@Horwitz]. However, as we now discuss, instead of having (a maximum of) 3 distinct real eigenvalues, each with multiplicity 8, there appear to be (a maximum of) 6 distinct real eigenvalues, each with multiplicity 4. The reason for this is that, somewhat surprisingly, a (real) eigenvalue $\lambda$ of a Jordan matrix $A$ does [*not*]{} in general satisfy the characteristic equation. To see this, consider the eigenvalue equation (\[Orig\]), with $A$ as in (\[Three\]), $\lambda\in{{\mathbb R}}$, and where $v=\pmatrix{x\cr y\cr z\cr}$. Assuming without loss of generality that $z\ne0$, explicit computation yields [@Eigen] $$\begin{aligned} \Big[ \det(\lambda I - A) \Big] z &\equiv& \left[ \lambda^3 - ({{\rm tr\,}}A) \, \lambda^2 + \sigma(A) \, \lambda - \det A \right] z \nonumber\\ &=& b \Big( a (cz) \Big) + {\overline}{c} \left( {\overline}{a} ({\overline}{b}z) \right) - \left[ b(ac) + ({\overline}{c}\,{\overline}{a}){\overline}{b} \right] z \label{CharLam}\end{aligned}$$ If $a$, $b$, $c$, and $z$ associate, the RHS of (\[CharLam\]) vanishes, and $\lambda$ does indeed satisfy the characteristic equation (\[Char\]); this will not happen in general. However, since the LHS of (\[CharLam\]) is a real multiple of $z$, this must also be true of the RHS, so that $$b \Big( a (cz) \Big) + {\overline}{c} \left( {\overline}{a} ({\overline}{b}z) \right) - \left[ b(ac) + ({\overline}{c}\,{\overline}{a}){\overline}{b} \right] z = rz \qquad\qquad r\in{{\mathbb R}}\label{RReal}$$ which can be solved to yield a quadratic equation for $r$ as well as constraints on $z$. The real eigenvalues of the $3\times3$ octonionic Hermitian matrix $A$ satisfy the modified characteristic equation $$\det(\lambda I - A) = \lambda^3 - ({{\rm tr\,}}A) \, \lambda^2 + \sigma(A) \, \lambda - \det A = r \label{Lameq}$$ where $r$ is either of the two roots of $$r^2 + 4\Phi(a,b,c) \, r - \Big| [a,b,c] \Big|^2 = 0 \label{Req}$$ with $a,b,c$ as defined by (\[Three\]) and where $\Phi$ was defined in (\[PhiEq\]). \[CharThm\] Furthermore, provided that $[a,b,c]\ne0$, each of $x$, $y$, and $z$ can be shown to admit an expansion of the form (given for $z$ only) With $A$ and $r$ as above, and assuming $[a,b,c]\ne0$, $$z = (\alpha a + \beta b + \gamma c + \delta) \left( 1 + {[a,b,c] \, r \over \Big| [a,b,c] \Big|^2} \right) \label{Zeq}$$ with $\alpha,\beta,\gamma,\delta\in{{\mathbb R}}$. Both of these results were obtained using [[*Mathematica*]{}]{} to solve (\[RReal\]) by brute force for real $r$ and octonionic $z$ given generic octonions $a$, $b$, $c$. An outline of the computation appears in Figure \[MathREq\]. 1 [=5.45in]{} 1 1 The real parameters $\alpha$,$\beta$,$\gamma$,$\delta$ may be freely specified for one (nonzero) component, say $z$; the remaining components $x$,$y$ have a similar form. The solutions of (\[RReal\]) are real, since the corresponding $24\times24$ real symmetric matrix has 24 real eigenvalues. We will refer to the 3 real solutions of (\[RReal\]) corresponding to a single value of $r$ as a [*family*]{} of eigenvalues of $A$. There are thus 2 families of real eigenvalues, each corresponding to a 4 independent (over ${{\mathbb R}}$) eigenvectors. We note several intriguing properties of these results. If $A$ is in fact complex, then the only solution of (\[Req\]) is $r=0$, and we recover the usual characteristic equation with a unique set of 3 (real) eigenvalues. If $A$ is quaternionic, then one solution of (\[Req\]) is $r=0$, leading to the standard set of 3 real eigenvalues and their corresponding quaternionic eigenvectors. However, unless $a$, $b$, $c$ involve only two independent imaginary quaternionic directions (in which case $\Phi(a,b,c)=0=[a,b,c]$), there will also be a nonzero solution for $r$, leading to a second set of 3 real eigenvalues. Finally, if $A$ is octonionic (so that in particular $[a,b,c]\ne0$), then there are two distinct solutions for $r$, and hence two different sets of real eigenvalues, with corresponding eigenvectors. Note that if $\det{A}=0\ne[a,b,c]$ then all of the eigenvalues of $A$ will be nonzero! Orthogonality ------------- The final surprise lies with the orthogonality condition for eigenvectors $v,w$ corresponding to different eigenvalues. It is [*not*]{} true that $v^\dagger w=0$, although the real part of this expression does vanish [@Eigen]. But, at least in the $2\times2$ case [@Eigen], it is straightforward to show that what is needed to ensure a decomposition of the form (\[Decomp\]) is the following generalized notion of orthogonality, which does in fact hold. Let $v$ and $w$ be two octonionic vectors. We will say that $w$ is [*orthogonal*]{} to $v$ if $$\label{Ortho} (v v^\dagger) \, w = 0$$ In the $3\times3$ case, a lengthy, direct computation verifies that eigenvectors with different real eigenvalues satisfy (\[Ortho\]) [*provided that*]{} the same value of $r$ is used for both eigenvectors. If $v$ and $w$ are eigenvectors of the $3\times3$ octonionic Hermitian matrix $A$ corresponding to different real eigenvalues in the same family (same $r$ value), then $v$ and $w$ are mutually orthogonal in the sense of (\[Ortho\]). \[OrthoThm\] The modified characteristic equation (\[Lameq\]) can be used to eliminate cubic and higher powers of $\lambda$ from any expression. Furthermore, given two distinct eigenvalues $\lambda_1\ne\lambda_2$, subtracting the two versions of (\[Lameq\]) and factoring the result leads to the equation $$(\lambda_1^2+\lambda_1\lambda_2+\lambda_2^2) - ({{\rm tr\,}}{A}) (\lambda_1+\lambda_2) + \sigma(A) = 0 \label{TwoLamEq}$$ which can be used to eliminate quadratic terms in one of the eigenvalues. We used [[*Mathematica*]{}]{} to implement these simplifications in a brute force verification of (\[Ortho\]) in this context, which used 6 hours of CPU time on a SUN Sparc20 with 224 Mb of RAM. A summary of the computation appears in Figures \[MathOrthoI\] and \[MathOrthoII\]. For Jordan matrices, we thus obtain [*two*]{} decompositions of the form (\[Decomp\]), corresponding to the two sets of real eigenvalues. For each, the eigenvectors are fixed up to orthogonal transformations which preserve the form (\[Zeq\]) of $z$. 1 1 1 Let $A$ be a $3\times3$ octonionic Hermitian matrix. Then $A$ can be expanded as in (\[Decomp\]), where $\{v_1, v_2, v_3\}$ is an orthonormal basis, as per (\[Ortho\]), of eigenvectors of $A$ corresponding to the real eigenvalues $\lambda_m$, which belong to the same family (same $r$ value). \[DecompThm\] Fix a family of real eigenvalues of $A$ by fixing $r$. If the eigenvalues are distinct, then the previous theorem guarantees the existence of an orthonormal basis of eigenvectors, and the result follows. If the eigenvalues are the same, the family consists of a single real eigenvalue $\lambda$ with multiplicity 3. Then ${{\rm tr\,}}(A)=3\lambda$ and $\sigma(A)=3\lambda^2$. Writing out these two equations in terms of the components (\[Three\]) of $A$, and inserting the first into the second, results in a quadratic equation for $\lambda$; the discriminant $D$ of this equation satisfies $D\le0$. But $\lambda$ is assumed to be real, which forces $D=0$, which in turn forces $A$ to be a multiple of the identity matrix, for which the result holds. The remaining case is when one eigenvalue, say $\mu$, has multiplicity 2 and one has multiplicity 1. Letting $v$ be a (normalized) eigenvector with eigenvalue $\mu$, consider the matrix $$X = A - \alpha \> v v^\dagger \label{Two}$$ with $\alpha\in{{\mathbb R}}$. For most values of $\alpha$, $X$ will have 3 distinct real eigenvalues, whose eigenvectors will be orthogonal by the previous theorem. But this means that eigenvectors of $X$ are also eigenvectors of $A$; the required decomposition of $A$ is obtained from that of $X$ simply by solving (\[Two\]) for $A$. Note in particular that for some quaternionic matrices with determinant equal to zero, one and only one of these two decompositions will contain the eigenvalue zero. Example {#Example} ======= Let $s$ be given by $$s=\cos\theta+k\ell\,\sin\theta$$ and consider the matrix $$B = \pmatrix{~~p & ~~iq & kqs \cr -iq & ~~p & jq\cr -kqs & -jq & p \cr} \label{BDef}$$ noting that $B$ is quaternionic if $\theta=0$. Turning first to the equation for $r$, (\[RReal\]) becomes $$r^2 + 4 q^3 r \cos\theta - 4 q^6 \sin^2\theta = 0$$ with solutions $$r_\pm = - 2 q^3 ( \cos\theta \pm 1 )$$ Since $$\begin{aligned} {{\rm tr\,}}B &=& 3p \\ \sigma(B) &=& 3 (p^2-q^2) \\ \det B &=& p^3 - 3 p q^2 + 2 q^3 \cos\theta\end{aligned}$$ the eigenvalue equation (\[Lameq\]) becomes $$\begin{aligned} 0 &=& \lambda^3 - 3p \, \lambda^2 + 3(p^2-q^2) \, \lambda - (p^3 - 3 p q^2 \mp 2 q^3) \\ &=& (\lambda - p \mp q)^2 \, (\lambda - p \pm 2q)\end{aligned}$$ [=4.75in]{} 1 An orthonormal basis of eigenvectors associated with these eigenvalues is $$\begin{aligned} \label{EigenI} \lambda_u &= p \pm q: \quad u_\pm &= \pmatrix{i\cr 0\cr j\cr} (f_u R_\pm) \\ \noalign{\smallskip} \lambda_v &= p \pm q: \quad v_\pm &= \pmatrix{j\cr 2ks\cr i\cr} (f_v R_\pm) \\ \noalign{\smallskip} \label{EigenIII} \lambda_w &= p \mp 2q: \quad w_\pm &= \pmatrix{~j\cr -ks\cr ~i\cr} (f_w R_\pm)\end{aligned}$$ where $f_u$, $f_v$, $f_w$ are arbitrary linear combinations of $a=iq$, $b=-kqs$, $c=jq$, i.e.$$f_u,f_v,f_w \in \langle 1,a,b,c \rangle$$ and where $R_\pm$ is given by $$R_\pm = \cases {\sin{\theta\over2}+k\ell\,\cos{\theta\over2}\cr \noalign{\smallskip} \cos{\theta\over2}-k\ell\,\sin{\theta\over2}\cr }$$ Note that in the limiting case $\theta\to0$, each $f$ is quaternionic, and $R_-$ reduces to $1$ while $R_+$ becomes $k\ell$. As expected, [*provided*]{} one fixes a family of eigenvectors and eigenvalues arising from a given choice of $r$, these eigenvectors satisfy the orthogonality property (\[Ortho\]) and thus lead to a decomposition of the form (\[Decomp\]). A partial verification of this using [[*Mathematica*]{}]{} is given in Figure \[MathEx\]. Discussion {#Discussion} ========== There are 2 quite different surprising aspects of our work: the mathematical changes needed to extend the eigenvalue problem to the octonions, and the fact that we have only been able to prove one of our key results using computer algebra. We discuss each of these in turn. It is of course intriguing that the eigenvalue problem over the octonions changes so much, for instance in that there are unexpectedly many real eigenvalues. But we find it remarkable that so much of the standard structure remains, provided it is reinterpreted appropriately. The most striking example of this is the need to generalize what is meant by orthogonality. We can relate our notion of orthonormality to the usual one by noting that a basis of ${{\mathbb O}}^n$ which is orthonormal in the sense (\[Ortho\]) satisfies $$vv^\dagger + ... + ww^\dagger=I$$ which follows directly from the definition. If we define a matrix $Q$ whose columns are just $v,...,w$, then this statement is equivalent to $$Q Q^\dagger = I$$ Over the quaternions, left matrix inverses are the same as right matrix inverses, and we would also have $$Q^\dagger Q = I$$ or equivalently $$v^\dagger v = 1 = ... = w^\dagger w; \qquad v^\dagger w = 0 = ...$$ which is just the standard notion of orthogonality. These two notions of orthogonality fail to be equivalent over the octonions; we have been led to view the former as more fundamental. Turning to our proof-by-computer, we reiterate that the only proof we currently have of our main orthogonality result, namely Theorem \[OrthoThm\], uses [[*Mathematica*]{}]{} to explicitly perform a horrendous, but exact, algebraic computation. While one could hope for a more elegant mathematical proof of this result, the [[*Mathematica*]{}]{} computation nevertheless establishes a result which would otherwise remain for the moment merely a conjecture. This is a good example of being able to use the computer to verify one’s intuition when it may not be possible to do so otherwise. But even more is true: Throughout our work with the octonions, the ability to manipulate octonionic expressions quickly and accurately has been crucial in [*developing*]{} our intuition in the first place. We do not feel that we would have been able to reach anything like our current understanding of the applications of the octonions to physics, on which we continue to be working actively, without the availability of a package such as the one described here. Finally, this computation was initially done in 1996 using [[*Mathematica*]{}]{} 2.2. While preparing this paper, we attempted to reproduce the computation using [[*Mathematica*]{}]{}3.0 — and couldn’t! Even for [[*Mathematica*]{}]{} 2.2, it was necessary to massage the computation by hand in order to succeed. One way this was done (between Figure \[MathOrthoI\] and Figure \[MathOrthoII\]) was by saving some intermediate steps to files and then restarting the kernel. Another technique was not to simplify all the components of an expression at the same time. For instance, in Figure \[MathOrthoII\], the vector ${\tt VVW}=(V V^\dagger)W$ contains 3 octonions, each of which requires roughly 8 Mb. Comparing the computations both versions of [[*Mathematica*]{}]{} could handle, [[*Mathematica*]{}]{} 3.0 appears to require nearly 4 times as much CPU time for the same computation; this was for identical inputs, with a minimum of special formatting. It is unfortunate that the many nice features of [[*Mathematica*]{}]{} 3.0 appear to require such a high price. [9]{} Tevian Dray and Corinne A. Manogue, [*The Octonionic Eigenvalue Problem*]{}, (in preparation). F. Reese Harvey, [**Spinors and Calibrations**]{}, Academic Press, Boston, 1990. I. L. Kantor and A. S. Solodovnikov, [**Hypercomplex Numbers, An Elementary Introduction to Algebras**]{}, Springer Verlag, Berlin and Heidelberg, 1989. Per-Erik Hagmark and P. Lounesto, [*Walsh functions, Clifford algebras, and the Cayley-Dickson process*]{}, in [**Clifford Algebras and their Applications in Mathematical Physics**]{}, eds. J. S. R. Chisholm and A. K. Common, D. Reidel Publishing, 1986, pp. 531–540. G. P. Wene, [*A construction relating Clifford algebras and Cayley-Dickson algebras*]{}, J. Math. Phys. [**25**]{}, 2351–2353 (1984). R. D. Schafer, [*On the algebras formed by the Cayley-Dickson process*]{}, Amer. J. Math. [**76**]{}, 435–446 (1954). G. B. Gureirch, [**Foundations of the Theory of Algebraic Invariants**]{}, P. Noordhoff, Groningen, 1964. Jörg Schray & Corinne A. Manogue, [*Octonionic Representations of Clifford Algebras and Triality*]{}, Found. Phys. [**26**]{}, 17–70 (1996). , eds. Rafał Abłamowicz, Pertti Lounesto, Josep M. Parra, Birkhäuser, Boston, 1996. , Pertti Lounesto, [http://www.hit.fi/[’176]{}lounesto/CLICAL.htm]{}, 1987; , P. Lounesto, R. Mikkola, V. Vierros, Institute of Mathematics, Helsinki University of Technology, 1987. , Rafał Abłamowicz, [http://math.gannon.edu/rafal/cliff3/]{}, 1996; Rafał Abłamowicz, [*Clifford algebra computations with Maple*]{}, in [**Geometric (Clifford) Algebras with Applicatino to Physics, Mathematics, and Engineering**]{}, Birkhäuser, Boston, 1996, pp. 463–501. P. Lounesto, [**Clifford Algebras and Spinors**]{}, Cambridge University Press, Cambridge, 1997. Jörg Schray, private communication, 1994. P. Jordan, J. von Neumann, and E. Wigner, Ann. Math. [**36**]{}, 29 (1934). H. Freudenthal, Adv. Math. [**1**]{}, 145 (1964). H. H. Goldstine & L. P. Horwitz, [*On a Hilbert Space with Nonassociative Scalars*]{}, Proc. Nat. Aca. [**48**]{}, 1134 (1962). [^1]: This construction of a new division algebra from 2 copies of another is a special case of the Cayley-Dickson process; for modern treatments, see [@ReeseHarvey; @Kantor; @LounestoI; @Wene; @Schafer].
{ "pile_set_name": "ArXiv" }
--- abstract: 'Let $K$ be a number field with ring of integers $\operatorname{\mathcal{O}}_{K}$. We prove that if $3$ does not divide $ [K:\operatorname{\mathbb{Q}}]$ and $3$ splits completely in $K$, then the unit equation has no solutions in $K$. In other words, there are no $x, y \in \operatorname{\mathcal{O}}_{K}^{\times}$ with $x + y = 1$. Our elementary $p$-adic proof is inspired by the Skolem-Chabauty-Coleman method applied to the restriction of scalars of the projective line minus three points.' address: 'N. Triantafillou, Department of Mathematics, University of Georgia, Athens, GA 30602, USA' author: - Nicholas George Triantafillou bibliography: - 'biblio.bib' title: The unit equation has no solutions in number fields of degree prime to $3$ where $3$ splits completely --- Let $K$ be a number field of degree $d$ over $\operatorname{\mathbb{Q}}$ and let $\operatorname{\mathcal{O}}_{K}$ be the ring of integers of $K$. The set $E_{K} \colonequals \{x \in \operatorname{\mathcal{O}}_{K}^{\times}: 1-x \in \operatorname{\mathcal{O}}_{K}^{\times}\}$ of *exceptional units* in $K$ is well-known to be finite, dating back to Siegel [@siegel-21]. Each $x \in E_{K}$ corresponds to a solution in $\operatorname{\mathcal{O}}_{K}^{\times}$ to the *unit equation*, $x + y = 1$. Solutions to the unit equation and the $S$-unit equation (which allows $x$ and $1-x$ to be units up to a fixed-in-advance finite set $S$ of prime ideals) remain of substantial practical interest because of a wide variety of applications to number theory and other fields. These include: enumerating elliptic curves over $K$ with good reduction outside a fixed set of primes [@smart-97]; understanding finitely generated groups, arithmetic graphs, and recurrence sequences [@evertse1988s]; and many Diophantine problems [@gyory-92]. Some work on exceptional units focuses on general upper bounds. Building on work of Baker and Györy on of linear forms in (complex/$p$-adic) logarithms, Evertse proved an explicit upper bound on $\# E_{K}$ which is exponential in the degree of $K$ [@evertse-84]. More recent work (e.g. [@gyory-19]) has refined these bounds somewhat, but the best known bounds remain exponential, while the ‘true’ upper bound is conjectured by Stewart to be sub-exponential (see p. 120 of [@evertse1988s].) Other work focuses on low-degree number fields and/or computation. For instance, [@nagell-70] and [@niklasch-smart-98] study the number of exceptional units in fields of degree $3$ and $4$. There has also been recent progress *computing* the set of of solutions to $S$-unit equations over low-degree number fields [@akmrvw18] both in practice and as a test-case for computations by variants of Chabauty’s method [@dan-cohen-wewers-15; @triantafillou-19]. Instead of studying low-degree $K$ or general upper bounds, we impose a local condition on $K$, showing: \[thm:3-splits\] Let $K$ be a number field. Suppose that $3 \nmid [K:\operatorname{\mathbb{Q}}]$ and $3$ splits completely in $K$. Then there is no solution to the unit equation in $K$. In other words, there is no pair $x,y \in \operatorname{\mathcal{O}}_{K}^\times$ such that $x + y = 1$. The set of degree $d$ polynomials in $\operatorname{\mathbb{Z}}[x]$ which generate number fields where $3$ splits completely have positive density (ordered by height). Indeed, if $g(x) = \sum_{j=0}^{d} a_{j} x^j$ satisfies $v_{3}(a_{d- i}) = i(i-1)/2$, a Newton polygon computation shows that the roots of $g$ have distinct $3$-adic valuations. If $g$ is also irreducible then $\operatorname{\mathbb{Q}}[x]/(g(x))$ is a field where $3$ splits completely. The set of number fields $K$ where $3$ splits completely is expected to have positive density in the set of degree $d$ number fields ordered by discriminant (for any $d$); there are precise conjectures of what this density should be [@bhargava-07]. Theorem \[thm:3-splits\] does not give the *first*-known infinite family of number fields of high degree without exceptional units. Indeed, if *any* prime $\operatorname{\mathfrak{p}}$ above $2$ in $K$ has residue field $\operatorname{\mathbb{F}}_{\operatorname{\mathfrak{p}}} \cong \operatorname{\mathbb{F}}_{2}$ then there are no exceptional units in $K$ for a trivial reason. The values $x$ and $1-x$ cannot simultaneously be non-zero modulo$\operatorname{\mathfrak{p}}$. To our knowledge, Theorem \[thm:3-splits\] yields the first-known infinite family of number fields of high degree without exceptional units outside of these trivial examples. The hypothesis that $3 \nmid [K:\operatorname{\mathbb{Q}}]$ in Theorem \[thm:3-splits\] is necessary. The set of degree $3$ number fields containing exceptional units has been well-understood since at least [@nagell-70]. One can construct infinitely many degree $3$ number fields with an exceptional unit and where $3$ splits completely as follows: Choose an integer $c \equiv 40 \pmod{81}$. Let $g(x) = (x - c) x (x-1) - 2x + 1$, which is irreducible over $\operatorname{\mathbb{Q}}$ be the rational root theorem. Let $\alpha$ be a root of $g$. Let $K = \operatorname{\mathbb{Q}}(\alpha)$. Since $\operatorname{\mathrm{Nm}}_{K/\operatorname{\mathbb{Q}}}(\alpha) = -g(0) = -1$ and $\operatorname{\mathrm{Nm}}_{K/\operatorname{\mathbb{Q}}}(1-\alpha) = g(1) = -1$, we see that $\alpha$ is an exceptional unit. Since the minimal polynomial of $(\alpha -2)/3$ — namely $\frac{1}{27} g(3x + 2)$ — has integer coefficients and is congruent to $x(x-1)(x+1)$ modulo $3$, we see that $3$ splits completely in $K$. If we replace the hypotheses ‘$3 \nmid [K:\operatorname{\mathbb{Q}}]$ and $3$ splits completely in $K$’ with ‘$5 \nmid [K:\operatorname{\mathbb{Q}}]$ and $5$ splits completely in $K$’ then Theorem \[thm:3-splits\] becomes false. Let $g(x) = x^3 - 4x^2 + x + 1$, let $\alpha$ be any root of $g$, and let $K = \operatorname{\mathbb{Q}}(\alpha)$. Then $5$ splits completely in $K$. Moreover, $\operatorname{\mathrm{Nm}}_{K/\operatorname{\mathbb{Q}}}(\alpha) = -g(0) = -1$ and $\operatorname{\mathrm{Nm}}_{K/\operatorname{\mathbb{Q}}}(1-\alpha) = g(1) = -1$, so $\alpha$ and $1-\alpha$ are both units, i.e. $\alpha$ is an exceptional unit. Suppose that $u, v \in \operatorname{\mathcal{O}}_{K}^{\times}$ satisfy $-u-v = 1$, so that $-u$ and $-v$ are solutions to the unit equation. Since $3$ splits completely in $K$, there are $d$ embeddings $\operatorname{\mathcal{O}}_{K} \hookrightarrow \operatorname{\mathbb{Z}}_{3}$. Let $u_{1}, \dots, u_{d}$ be the images of $u$ in $\operatorname{\mathbb{Z}}_{3}$ under these embeddings. Since $u$ and $v$ are units, $u_{i} \in 1 + 3 \operatorname{\mathbb{Z}}_{3}$ for all $i \in \{1, \dots, d\}$. Also, $\operatorname{\mathrm{Nm}}_{K/\operatorname{\mathbb{Q}}}(u), \operatorname{\mathrm{Nm}}_{K/\operatorname{\mathbb{Q}}}(v) \in \operatorname{\mathbb{Z}}^{\times} = \{\pm 1\}$. We have $$\begin{aligned} \prod_{i=1}^{d} u_{i} = \operatorname{\mathrm{Nm}}_{K/\operatorname{\mathbb{Q}}}(u) = 1\, \qquad \text{and} \qquad \prod_{i=1}^{d}(1 + u_{i}) = \operatorname{\mathrm{Nm}}_{K/\operatorname{\mathbb{Q}}}(-v) = (-1)^{d}\,. \end{aligned}$$ We see that $n = 1$ is a zero of the $3$-adic analytic function $$f(n) \colonequals (1 + u_1^n) \cdots (1 + u_{d}^{n}) - (-1)^{d} \,$$ and $$\begin{aligned} f(-n) & = \prod_{i=1}^{d} (1 + u_i^{-n}) - (-1)^{d} = \prod_{i=1}^{d} u_{i}^{-n} \prod_{i=1}^{d} (1 + u_i^{n}) - (-1)^{d} = \prod_{i=1}^{d} (1 + u_i^{n}) - (-1)^{d} = f(n)\,. \end{aligned}$$ In particular, expanding $f$ as a $p$-adic power series, all coefficients in odd degrees are zero. Now, $$f(n) = -(-1)^{d} + \prod_{i=1}^{d}(1 + \exp(n \log u_{i})),.$$ Let $v_{3}$ be the $3$-adic valuation normalized so that $v_3(3) = 1$. Since $v_{3}(\log u_i) \geq 1$ and $\exp$ converges when $v_{3}(n \log u_i) > 1/2$ (see [@gouvea-97]), this expression converges for all $n \in \operatorname{\mathbb{Z}}_{3}$. Expanding $f$ as a power series, $$f(n) = -(-1)^d + \prod_{i=1}^{d}(2 + n \log u_i + \frac{n^2}{2} (\log u_i)^2 + \frac{n^3}{3!} (\log u_i)^3 + \cdots) \equalscolon \sum_{j =0}^{\infty} a_{j} n^j\,.$$ Now, $$\begin{aligned} a_{0} = 2^d - (-1)^d\,, \quad a_{1} = 0\,, \quad a_{2} = 2^{d-3} \sum_{i=1}^{d} (\log u_i)^2\,, \quad a_{3} = 0\,, \quad \text{and} \quad v_{3}(a_{j}) \geq 3\,\, \text{for} \,\, j \geq 4\,.\\ \end{aligned}$$ Since $v_3(a_{2}) \geq 2$ and $f(1) = 0$ we have $v_{3}(a_{0}) \geq 2$. But $v_{3}(2^d - (-1)^d) \geq 2$ if and only if $3|d$. The inspiration for the proof of Theorem \[thm:3-splits\] is a variant of the method of Skolem-Chabauty-Coleman applied to the *restriction of scalars* of $\operatorname{\mathbb{P}}^{1}_{\operatorname{\mathcal{O}}_{K}} \smallsetminus \{0,1,\infty\}$ from $\operatorname{\mathcal{O}}_{K}$ to $\operatorname{\mathbb{Z}}$. In this setting, $\operatorname{\mathbb{P}}^{1}_{\operatorname{\mathcal{O}}_{K}} \smallsetminus \{0,1,\infty\}$ embeds into its the *generalized* Jacobian $\operatorname{\mathbb{G}}_{m, \operatorname{\mathcal{O}}_{K}} \times \operatorname{\mathbb{G}}_{m, \operatorname{\mathcal{O}}_{K}}$ via the Abel-Jacobi map $x \mapsto (x, x-1)$. To prove that $\operatorname{\mathbb{P}}^{1}\smallsetminus \{0,1,\infty\} = \emptyset$, we consider the restriction of scalars of the Abel-Jacobi map. In this language, the proof of Theorem \[thm:3-splits\] amounts to showing that for any unit $u \in \operatorname{\mathcal{O}}_{K}^{\times}$ the intersection $$E_{u} \colonequals (\operatorname{Res}_{\operatorname{\mathcal{O}}_{K}/\operatorname{\mathbb{Z}}} \operatorname{\mathbb{P}}^1\smallsetminus\{0,1,\infty\})(\operatorname{\mathbb{Z}}_{3}) \cap \overline{\{u^n: n \in \operatorname{\mathbb{Z}}\} \times \operatorname{\mathcal{O}}_{K}^{\times}}$$ inside $(\operatorname{Res}_{\operatorname{\mathcal{O}}_{K}/\operatorname{\mathbb{Z}}} (\operatorname{\mathbb{G}}_m \times \operatorname{\mathbb{G}}_{m}))(\operatorname{\mathbb{Z}}_{3})$ is empty. Here, the closure on the right is respect to the $3$-adic topology. To conclude, $\bigcup_{u \in \operatorname{\mathcal{O}}_{K}^{\times}} E_{u} = \emptyset$ is the set of solutions to the unit equation in $K$. See [@triantafillou-19] for a more general discussion of using Skolem-Chabauty-Coleman applied to the *restriction of scalars* to compute solutions to the $S$-unit equation. Acknowledgements {#acknowledgements .unnumbered} ---------------- Thank you to Joe Rabinoff and Bjorn Poonen for comments which simplified the proof, to Pete Clark, Dino Lorenzini, and Paul Pollack for helpful feedback on an early draft of this manuscript and to Vishal Arul, Jack Petok, and Padmavathi Srinivasan for helpful conversations. Thank you to the NSF Graduate Research Fellowship under grant \#1122374, Simons Foundation grant \#550033, and the RTG grant DMS-1344994 in Algebra, Algebraic Geometry, and Number Theory at UGA for funding this work.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Given a pedestrian image as a query, the purpose of person re-identification is to identify the correct match from a large collection of gallery images depicting the same person captured by disjoint camera views. The critical challenge is how to construct a robust yet discriminative feature representation to capture the compounded variations in pedestrian appearance. To this end, deep learning methods have been proposed to extract hierarchical features against extreme variability of appearance. However, existing methods in this category generally neglect the efficiency in the matching stage whereas the searching speed of a re-identification system is crucial in real-world applications. In this paper, we present a novel deep hashing framework with Convolutional Neural Networks (CNNs) for fast person re-identification. Technically, we simultaneously learn both CNN features and hash functions/codes to get robust yet discriminative features and similarity-preserving hash codes. Thereby, person re-identification can be resolved by efficiently computing and ranking the Hamming distances between images. A structured loss function defined over positive pairs and hard negatives is proposed to formulate a novel optimization problem so that fast convergence and more stable optimized solution can be obtained. Extensive experiments on two benchmarks CUHK03 [@FPNN] and Market-1501 [@Market1501] show that the proposed deep architecture is efficacy over state-of-the-arts.' author: - | Lin Wu$^{\ddag}$, Yang Wang$^{\dag}$\ $^{\ddag}$ISSR, ITEE, The University of Queensland, Brisbane, QLD, 4072, Australia\ $^{\dag}$The University of New South Wales, Kensington, Sydney, Australia\ [email protected] [email protected] bibliography: - 'ijcai16.bib' title: 'Structured Deep Hashing with Convolutional Neural Networks for Fast Person Re-identification' --- Introduction {#sec:intro} ============ re-identification is task of matching persons observed from non-overlapping camera views based on visual appearance. It has gained considerable popularity in video surveillance, multimedia, and security system by its prospect of searching a person of interest from a large amount of video sequences [@Zero-shot-person; @Person-ranking; @Person-context]. The major challenge arises from the variations in human appearances, poses, viewpoints and background cluster across camera views. Some examples are shown in Fig.\[fig:example\]. Towards this end, many approaches [@Farenzena2010Person; @Yang:color-person; @MidLevelFilter; @Pedagadi2013Local; @LADF; @paul2015ensemble] have been proposed by developing a combination of low-level features (including color histogram [@Gray2008Viewpoint], spatial co-occurrence representation [@WangShape2007], LBP [@Xiong2014Person] and color SIFT [@eSDC]) against variations ([*e.g., *]{}poses and illumination) in pedestrian images. However, these hand-crafted features are still not discriminative and reliable to such severe variations and misalignment across camera views. Recently, deep learning methods [@FPNN; @JointRe-id; @DeepReID; @Ding:DeepRank; @PersonNet; @DeepRanking; @DomainDropout] have been proposed to address the problem of person re-identification by learning deeply discriminative Convolutional Neural Network (CNN) features in a *feed-forward* and *back-propagation* manner. It extracts hierarchical CNN features from pedestrian images; the subsequent metric-cost part compares the CNN features with a chosen metric encoded by specific loss functions, [*e.g., *]{}contrastive (pair-wise) [@FPNN; @JointRe-id; @PersonNet] or triplet [@DeepReID; @DeepRanking] loss functions. However, such typical deep learning methods are not efficient in real-time scenario, due to the less-efficiency of matching two pedestrian images by extracting and comparing hierarchical CNN features. In fact, the excellent recognition accuracy in neural network-based architectures comes at expense of high computational cost both at training and testing time. The main computational expense for these deep models comes from convolving filter maps with the entire input image, making their computational complexity at least linear in the number of pixels. And matching these CNN features to obtain similarity values is not fast enough to be applicable in real-world applications. In this paper, we aim to reduce the computational burden of person re-identification by developing a fast re-identification framework. Motivation ---------- To cope with ever-growing amounts of visual data, deep learning based hashing methods have been proposed to simultaneously learn similarity-preserved hashing functions and discriminative image representation via a deep architecture [@Lai:hash-deep; @DeepSemanticHash; @DeepRegularlizedHash]. Simply delving existing deep hashing approaches into a person re-identification system is not trivial due to the difficulty of generalizing these pre-trained models to match pedestrian images in disjoint views. Fine-tuning is a plausible way to make pre-trained models suitable to re-identification, however, to suit their models, training images are commonly divided into mini-batches, where each mini-batch contains a set of *randomly* sampled positive/negative pairs or triplets. Thus, a contrastive or triplet loss is computed from each mini-batch, and the networks try to minimize the loss function feed-forwardly and update the parameters through back-propagation by using Stochastic Gradient Decent (SGD) [@SGD]. We remark that randomly sampled pairs/triplets carry little helpful information to SGD. For instance, many triplet units can easily satisfy the relative comparison constraint in a triplet loss function (Eq ), resulting into a slow convergence rate in the training stage. Worse still, mini-batches with random samples may fail to obtain a stable solution or collapsed into a local optimum if a contrastive/triplet loss function is optimized [@LiftEmbed]. To this end, a suitable loss function is highly demanded to work well with SGD over mini-batches. In this paper, we propose a deep hashing based on CNNs to efficiently address the problem of person re-identification. To mitigate the undesirable effects caused by contrastive/triplet loss function, we propose a structured loss function by *actively* adding hard negative samples into mini-batches, leading to a structured deep hashing framework. The proposed structured loss can guide sub-gradient computing in SGD to have correct directions, and thus achieves a fast convergence in training. Our Approach ------------ One may easily generate a straightforward two-stage deep hashing strategy by firstly extracting CNN features from a pre-trained model [*e.g., *]{}AlexNet [@Krizhevsky2012Imagenet], followed by performing the learned hash functions (separate projection and quantization step) to convert such CNN features into binary codes. However, as demonstrated in section \[sec:exp\], such a strategy cannot obtain optimal binary codes. As such binary codes may not well characterize the supervised information from training data [*i.e., *]{}intra-personal variation and inter-personal difference, due to the *independence* of two stages. In fact, such two stages can boost each other to achieve much better performance, that is, the learned binary codes can guide the learning of useful CNN features, while CNN features can in turn help learn semantically similarity-preserving hash function/codes. Motivated by this, we present a structured deep hashing architecture to *jointly* learn feature representations and hash codes for person re-identification. The overall framework is illustrated in Fig.\[fig:framework\]. In our architecture, mini-batches contain all positive pairs for a particular pedestrian, meanwhile each positive pair (has a query image and its correct match image from a different camera view) is augmented by actively selected hard negatives for its query and match image, respectively. Such mini-batches are taken into the inputs of deep network with a structured loss function optimized to learn CNN features and hash functions jointly. The major contributions are summarized below: - To the best of our knowledge, we are the first to solve person re-identification efficiently by presenting a structured deep hashing model. This makes our paper distinct from existing studies [@Zero-shot-person; @Person-ranking; @Person-context] where the matching efficiency is not addressed. - By simultaneously learning CNN features and hash functions/codes, we are able to get robust yet discriminative features against complex pedestrian appearance and boosted hash codes, so that every two deep hashing codes learned from the same identities are close to each other while those from different identities are kept away. - To combat the drawbacks of the contrastive/triplet loss, we propose a structured loss function where mini-batches are augmented by considering hard negatives. Also, the proposed structured loss function that is imposed at the top layer of the network can achieve fast convergence and a stable optimized solution. ![Overview of our deep hashing framework for person re-identification. Our deep neural network takes a feed-forward, back-propagation strategy to learn features and hash codes simultaneously. During the feed-forward stage, the proposed network performs inference from a mini-batch. The mini-batch is put through a stack of convolutional layers to generate nonlinear yet discriminative features, which are subsequently mapped to output feature vector by fully-connected layers (FC). Meanwhile, a hash function layer is introduced at the top of FC layer to learn hash codes that are optimized by a structured loss function to preserve their similarities/dissimilarities. In back-propagation, parameters are updated by computing their Stochastic Gradient Decent (SGD) w.r.t. the mini-batch.[]{data-label="fig:framework"}]({hash_framework}){width="3.5in" height="2in"} Related Work {#sec:related} ============ In this section, we briefly review deep learning based on CNNs for person re-identification and several typical hashing methods, as they are closely related to our proposed technique. In literature of person re-identification, many studies try to address this challenging problem by either seeking a robust feature representation [@Farenzena2010Person; @MidLevelFilter; @YangCIKM13; @LinMM13; @YangKAIS16; @YangMM15; @YangTIP15; @TNNLS17; @PAKDD14; @YangIJCAI16; @Gray2008Viewpoint; @eSDC] or casting it as a metric learning problem where more discriminative distance metrics are learned to handle features extracted from person images across camera views [@LADF; @KISSME; @Pedagadi2013Local; @Xiong2014Person; @LOMOMetric; @NullSpace-Reid]. The first aspect considers to find features that are robust to challenging factors while preserving identity information. The second stream generally tries to minimize the intra-class distance while maximize the inter-class distance. Also, person re-identification can be approached by a pipeline of image search where a Bag-of-words [@Market1501] model is constructed to represent each pedestrian image and visual matching refinement strategies can be applied to improve the matching precision. Readers are kindly referred to [@PersonReid-book] to have more reviews. A notable improvement on person re-identification is achieved by using Convolutional Neural Networks (CNNs) [@FPNN; @JointRe-id; @DeepReID; @Ding:DeepRank; @PersonNet; @Lin-PR2016; @DeepRanking; @Yang-TIP2017; @DomainDropout], which can jointly learn robust yet discriminative feature representation and its corresponding similarity value in an end-to-end fashion. However, existing deep learning methods in person re-identification are facing a major challenge of efficiency, where computational time required to process an input image is very high due to the convolution operations with the entire input through deep nets. Thus, from a pragmatical perspective, an advanced yet fast neural network-based architecture is highly demanded. This motivated us to develop an efficient deep learning model to alleviate the computational burden in person re-identification. Hashing is an efficient technology in approximate nearest neighbor search with low storage cost of loading hash codes. Learning-based hash methods can be roughly divided into two categories: unsupervised methods and supervised methods. Unsupervised methods including Spectral Hashing [@Weiss:SH; @Yang-SIGIR15] and Iterative Quantization [@Gong:cvpr2011] only use the training data to learn hash functions. Supervised methods try to leverage supervised information to learn compact binary codes. Some representative methods are Binary Reconstruction Embedding (BRE) [@Kulis:nips2009], Minimal Loss Hashing (MLH) [@Norouzi:loss-hash], and Supervised Hashing with Kernels (KSH) [@Liu:kernel-hash]. Nonetheless, these hashing methods often cope with images represented by hand-crafted features ([*e.g., *]{}SIFT [@eSDC]), which are extracted before projection and quantization steps. Moreover, they usually seek a linear projection which cannot capture the nonlinear relationship of pedestrian image samples[^1]. Even though some kernel-based hashing approaches [@Liu:kernel-hash; @Lin-IVC2017] have been proposed, they are stricken with the efficiency issue. To capture the non-linear relationship between data samples while keeping efficient, Liong [*et al. *]{}[@liong:deep] present a Deep Hashing to learn multiple hierarchical nonlinear transformation which maps original images to compact binary code and thus supports large-scale image retrieval. A supervised version named Semantic Deep Hashing is also presented in [@liong:deep] where a discriminative item is introduced into the objective function. However, the above methods did not include a pre-training stage in their networks, which may make the generated hash codes less semantic. To keep the hash codes semantic, Xia [*et al. *]{}[@Xia:hash2014] proposed a deep hashing architecture based on CNNs, where the learning process is decomposed into a stage of learning approximate hash codes from supervised priors, which are used to guide a stage of simultaneously learning hash functions and image representations. More recently, to generate the binary hash codes directly from raw images, deep CNNs are utilized to train the model in an end-to-end manner where discriminative features and hash functions are simultaneously optimized [@Lai:hash-deep; @DeepSemanticHash; @DeepRegularlizedHash]. However, in training stage, they commonly take mini-batches with randomly sampled triplets as inputs, which may lead to local optimum or unstable optimized solution. By contrast, in this paper we deliver the first efforts in proposing a structured deep hashing model for person re-identification, which allows us to jointly learn deep feature representations and binary codes faithfully. The proposed structured loss function benefits us from achieving fast convergence and more stable optimized solutions, compared with pairwise/triplet ranking loss. ![Illustrations on different loss functions. (a) Contrastive loss; (b) Triplet ranking loss; (c) Our structured loss. Here, $\bx$’s and $\by$’s indicate hash codes of pedestrian images captured by query and gallery camera view, respectively. For a specific pedestrian’s hash codes $\bx_i$, its correct match’s code is $\by_i$ from a different view. Green edges and red edges represent similar and dissimilar examples, respectively. Our method explicitly adds hard negatives ([*e.g., *]{}$y_m$, $y_n$) for all positive pairs ([*e.g., *]{}($x_1, y_1$) and ($x_2$, $y_2$) ) into mini-batches.[]{data-label="fig:loss"}]({revisit_loss}){width="3.5in" height="1.2in"} Structured Deep Hashing for Person Re-identification {#sec:method} ==================================================== Our major contribution is to jointly learn feature representation from raw person images and their mappings to hash codes by presenting an improved deep neural network. The proposed network takes a mini-batch as its input which contains images in a form of positive/negative pairs. The architecture consists of three components: 1) a stack of convolution layers followed by max pooling to learn non-linear feature mappings from raw pedestrian images; 2) a hash layer connected to the first and the second fully connected layers; 3) a structured loss function is designed to optimize the whole mini-batch. The architecture overview is illustrated in Fig.\[fig:framework\]. Learning Deep Hashing Functions ------------------------------- Assuming $\mathcal{I}$ to be the original image space, a hash function $f: \mathcal{I} \rightarrow \{0,1\}^r$ is treated as a mapping that projects an input image $\mathcal{I}$ into a $r$-bit binary code $f(\mathcal{I})$ while preserving the similarities of person images across camera views. Learning based hashing methods aim to seek a set of hash functions to map and quantize each sample into a compact binary code vector. Assuming we have $r$ hash functions to be learned, which map an image $\mathcal{I}$ into a $r$-bit binary code vector $f(\mathcal{I})=[f_1(\mathcal{I}), f_2(\mathcal{I}), \ldots, f_r(\mathcal{I})]$. Although many learning-based hashing methods have been proposed [@Gong:angular; @Gong:cvpr2011; @He:kmeans-hash; @Norouzi:loss-hash; @Kulis:nips2009], most of them essentially learn a single linear projection matrix, which can not well capture the nonlinear relationship of samples. Admittedly, some kernel-based hashing methods are available [@Liu:kernel-hash; @He:kdd2010], they instead suffer from the efficiency issue because kernel-based methods cannot have explicit nonlinear mapping. In this work, we propose to learn deep hash functions with CNNs to jointly learn feature representation from raw pixels of pedestrian images and their mappings to hash codes. In this way, feature representations for person images can be learned more optimally compatible with the coding process, thus producing optimal hash codes. During training, the input to our network is a mini-batch containing pairs of fixed-size 160$\times$60 RGB images. The images are passed through four convolutional layers, where we use filters with a very small receptive filed: 3$\times$3. The convolution stride is fixed to 1 pixel. Spatial pooling is carried out by three max-pooling layers. Max-pooling is performed over a 2$\times$2 pixel window, with stride 2. After a stack of convolution layers, we have two fully-connected layers where the first one has 4096 dimension and the second is 512-dim, which are then fed into the hash layer to generate a compact binary code. We show details of layers in CNNs in Table \[tab:layer\]. Name Type Output Dim FS Sride ------- ------ ------------------------ ------------ ------- Conv0 C $157\times57\times32$ 3$\times3$ 1 Pool0 MP $79\times 29\times 32$ 2$\times2$ 2 Conv1 C $76\times 26\times 32$ 3$\times$3 1 Pool1 MP $38\times 13\times 32$ 2$\times$2 2 Conv2 C $35\times 10\times 32$ 3$\times$3 1 Pool2 C $18\times 5\times 32$ 3$\times$3 2 Conv3 C $15\times 2\times 32$ 3$\times$3 1 Pool4 MP $15\times 2\times 32$ 1$\times$1 1 FC1 FC - 4096 - FC2 FC - 512 - : Layer parameters of convolutional neural networks. The output dimension is given by height$\times$width$\times$width. FS: filter size for convolutions. Layer types: C: convolution, MP: max-pooling, FC: fully-connected. All convolution and FC layers use hyperbolic tangent as activation function.[]{data-label="tab:layer"} Inspired by [@Sun:deep-face], we add a bypass connection between the first fully connected layer and the hash layer to reduce the possible information loss. Another reason is features from the second fully connected layer is very semantic and invariant, which is unable to capture the subtle difference between person images. Thus, we rewrite the deep hash function as: $$f(\mathcal{I},\bw_i)=sigmoid\left(\bw_i^T [g_1(\mathcal{I}); g_2(\mathcal{I})]\right),$$ where $sigmoid(t)=1/(1+\exp^{(-\bw^T t)})$, $\bw_i$ denotes the weights in the $i$-th hash function, $g_1(\cdot)$ and $g_2(\cdot)$ represent feature vectors from the outputs of the two fully connected layers, respectively. Then, we have $f(\mathcal{I},\bW)=[f(\mathcal{I},\bw_1),\dots,f(\mathcal{I}, \bw_r)]$. After the deep architecture is trained, the hashing code for a new image $\mathcal{I}$ can be done by a simple quantization $b=sign(f(\mathcal{I}, \bW))$, where $sign(v)$ is a sign function on vectors that for $i=1,2,\dots,r$, $sign(v_i)=1$ if $v_i>0$, otherwise $sign(v_i)=0$. Structured Loss Optimization ---------------------------- In deep metric learning for person re-identification, the network is often trained on data in the form of pairs [@FPNN; @JointRe-id; @DeepReID] or triplet ranking [@Ding:DeepRank]. Thus, there are two commonly used cost functions, contrastive/pairwise loss and triplet ranking loss, which can be used in hash code optimization. We briefly revisit the two loss functions and then introduce the proposed structured loss function. #### Contrastive/Pairwise Loss Function Given a person’s binary codes $\bx_i$ and its correct match’s codes $\by_i$ from a different camera view, the contrastive training tries to minimize the Hamming distance between a positive pair of $(\bx_i, \by_i)$ and penalize the negative pairs $(\bx_i,\by_j)$ ($i\neq j$) with a Hamming distance smaller than a margin. The contrastive cost function can be defined as $$F = \sum_{(i,j)} a_{i,j} ||\bx_i - \by_j||_H + (1-a_{ij}) \left(\max\left(0,1 - ||\bx_i - \by_j||_H\right)\right)$$ where $\bx_i$, $\by_j$ $\in \{0,1\}^r$ and $||\cdot||_H$ represents the Hamming distance. The label $a_{ij}\in \{0,1\}$ indicates whether a pair of binary codes $(\bx_i,\by_j)$ depicting the same person. #### Triplet Ranking Loss Function Some recent studies have been made to learn hash functions that preserve relative similarities of the form of a triplet data $\left(\mathcal{I}, \mathcal{I}^+, \mathcal{I}^-\right)$ where image $\mathcal{I}$ (*anchor*) of a specific person is more similar to all other images $\mathcal{I}^+$(*positive*) of the same person than it is to any image $\mathcal{I}^-$ (*negative*) of any other person (images $\mathcal{I}^+$ and $\mathcal{I}^-$ are from a different camera view from $\mathcal{I}$). Specifically, in hash function learning, the goad is to find a mapping $f(\cdot)$ such that the binary code $f(\mathcal{I})=\bx_i$ is closer to $f(\mathcal{I}^+)=\by_i$ than to $f(\mathcal{I}^-)=\by_j$ ($j\neq i$). Thus, we want $$\label{eq:triplet_constraint} ||\bx_i-\by_i||_H + 1 < ||\bx_i-\by_j||_H, \forall (\bx_i, \by_i, \by_j) \in \mathbb{T},$$ where $\mathbb{T}$ is the set of all possible triplets in the training set and has cardinality $N$. Accordingly, the triplet ranking hinge loss is defined by $$\begin{aligned} &F=\sum_i \max\left( 0, 1-\left( ||\bx_i -\by_j||_H - ||\bx_i-\by_i||_H \right)\right)\\ &s.t. ~~~ \bx_i, \by_i, \by_j \in \{0,1\}^r. \end{aligned}$$ A noticeable difference between a contrastive embedding and a triplet embedding is that a triplet unit with similar and dissimilar inputs provide some distance comparison context for the optimization process, as opposed to the contrastive loss that the network minimizes (same class) or maximizes (different classes) as much as possible for each pair independently [@TripletNet]. In triplet embedding, however, generating all possible triplets would result in many triplets that easily fulfill the constraint in Eq , which is known as over-sampling. These triplets would not contribute to the training whereas resulting in slow convergence. An alternative strategy is to perform a smart sampling where one must be careful to avoid too much focus on hard training exemplars due to the possibility of over-fitting. Thus, it is crucial to *actively* select informative hard exemplars in order to improve the model. Below, we introduce our structured loss which can avoid aforementioned over or under-sampling dilemmas by virtue of actively adding difficult neighbors to positive pairs into training batches. #### The Proposed Structured Loss Function Previous works on person re-identification implement a Stochastic Gradient Decent (SGD) [@SGD] by drawing pairs or triplets of images uniformly at random. They didn’t fully makes use of the information of the mini-batch that is sampled at a time and not only individual pairs or triplets. By contrast, we propose a structured loss over a mini-batch in order to take fully advantage of the training batches used in SGD. Meanwhile, the proposed structured loss can ensure fast convergence and stableness in training. As shown in Fig.\[fig:loss\] (c), the structured loss is conducted on all positive pairs and corresponding close (“difficult") negative pairs across camera views. Specifically, it can be formulated as $$\label{eq:structured_loss} \begin{aligned} &F = \frac{1}{|\bar{P}|} \sum_{\bx_i,\by_i \in \bar{P}} \max\left( 0, F_{\bx_i,\by_i}\right),\\ &F_{\bx_i,\by_i } = \max\left( \max\left(0, 1 - ||\bx_i-\by_k||_H\right), \max\left( 0, 1-||\by_i -\by_l||_H\right)\right)\\ & + ||\bx_i-\by_i||_H, \\ &s.t. ~~ \bx_i,\by_i,\by_k,\by_l \in \{0,1\}^r, (\bx_i,\by_k)\in \bar{N}, (\by_i,\by_l)\in \bar{N}, \end{aligned}$$ where $\bar{P}$ and $\bar{N}$ denote the set of positive and negative pairs in each mini-batch. The process of selecting positive and negative samples is elaborated in Section \[ssec:hard\_negative\]. **Difference to contrastive and triplet ranking loss:** - In pairwise training with $O(m)$ separate pairs in the batch, a total of $O(m^2)$ pairs can be generated accordingly. However, these negative edges induced between randomly sampled pairs carry very limited information [@LiftEmbed]. By contrast, selected difficult exemplars are sharper cases that a full sub-gradient method would more likely focus on; - Compared with triplet embedding containing randomly sampled triplets, our training batch is augmented by adding negative neighbors bilaterally for each positive pairs. By doing this, the optimization process is conducted on most violate constraints, leading to fast convergence. Fig.\[fig:loss\] (a) and (b) illustrates a batch of positive/negative pairs and triplets with corresponding contrastive loss and triplet ranking loss. Green edges represent positive pairs (the same person) and red edges represent negative pairs (different individuals). Please note that these pairs and triplets are sampled completely random into a mini-batch. Fig.\[fig:loss\] (c) illustrates the mining process for two positive pairs in the batch where for each image in a positive pair we seek its close (hard) negative images. We can see that our method allows mining the hard negatives from both the query image ([*e.g., *]{}$\bx_1$ ) and its correct match ([*e.g., *]{}$\by_1$) of a pair against gallery images ([*e.g., *]{}$\by_m, m \neq 1$). #### Optimization For ease of optimization, we relax Eq. by replacing the Hamming norm with the $\ell_2$-norm and replacing the integer constraints on $\bx$’s and $\by$’s with the range constraints. The modified loss function is $$\label{eq:new_loss} \begin{aligned} &F = \frac{1}{|\bar{P}|} \sum_{\bx_i,\by_i \in \bar{P}} \max\left( 0, F_{\bx_i,\by_i}\right),\\ &F_{\bx_i,\by_i } = \max\left( \max\left(0, 1 - ||\bx_i-\by_k||_2^2\right), \max\left( 0, 1-||\by_i -\by_l||_2^2\right)\right)\\ & + ||\bx_i-\by_i||_2^2, \\ & s.t. ~~ \bx_i,\by_i,\by_k,\by_l \in [0,1]^r, (\bx_i,\by_k)\in \bar{N}, (\by_i,\by_l)\in \bar{N}. \end{aligned}$$ The variant of structured loss is convex. Its sub-gradients with respect to $\bx_i$, $\by_i$, $\by_k$, and $\by_l$ are $$\begin{aligned} &\frac{\partial F }{\partial \bx_i}= (2\by_k-2\by_i) \times \mathbb{I}[2+ ||\bx_i-\by_i||_2^2 > ||\bx_i-\by_k||_2^2+||\by_i -\by_l||_2^2 ]\\ &\frac{\partial F}{\partial \by_i}=(2\by_l-2\bx_i) \times \mathbb{I}[2+ ||\bx_i-\by_i||_2^2 > ||\bx_i-\by_k||_2^2+||\by_i -\by_l||_2^2]\\ &\frac{\partial F}{\partial \by_k}=2\bx_i \times \mathbb{I}[2+ ||\bx_i-\by_i||_2^2 > ||\bx_i-\by_k||_2^2+||\by_i -\by_l||_2^2]\\ & \frac{\partial F}{\partial \by_l}=2\by_i \times \mathbb{I}[2+ ||\bx_i-\by_i||_2^2 > ||\bx_i-\by_k||_2^2+||\by_i -\by_l||_2^2] \end{aligned}$$ The indicator function $\mathbb{I}[\cdot]$ is the indicator function which outputs 1 if the expression evaluates to true and outputs 0 otherwise. Thus, the loss function in Eq. can be easily integrated into back propagation of neural networks. We can see that our structured loss provides informative gradients signals for all negative pairs which are within the margin of any positive pairs. In contrast to existing networks like [@FPNN; @JointRe-id] where only hardest negative gradients are updated, making the training easily over-fit, the proposed structured loss makes the optimization much more stable. Hard Negative Mining for Mini-batches {#ssec:hard_negative} ------------------------------------- As mentioned before, our approach differs from existing deep methods by making full information of the mini-batch that is sampled at a time, including positive pairs and their difficult neighbors. Please note that difficult neighbors are defined only with respect to the gallery camera view. The motivation of doing this is to enhance the mini-batch optimization in network training because the sub-gradient of $F_{\bx_i,\by_i }$ would use the close negative pairs. Thus, our approach biases the sample towards including “difficult" pairs. In this paper, we particularly select a few positive pairs at random, and then **actively** add their difficult (hard) neighbors into the training mini-batch. This augmentation adds relevant information that a sub-gradient would use. Specifically, we determine the elements in mini-batches by online generation where all anchor-positive pairs in any identity are kept while selecting the hard negatives for both the anchor and its positive correspondence. In fact, this procedure of mining hard negative edges amounts to computing the loss augmented inference in structured prediction setting [@Ioannis:icml2004; @cutting-plane; @LiftEmbed]. Intuitively, the loss from hard negative pairs should be penalized more heavily than a loss involving other pairs. In this end, our structured loss function contains enough negative examples within the margin bound, which can push the positive examples towards the correct direction and thus making the optimization much more stable. ![ Illustration on degeneration of contrastive and triplet ranking embedding with randomly sampled training pairs. Pink circles, green squares, and purple triangles indicate three different classes. Dotted blue circles regulate the margin bound where the loss becomes zero out of the bound. Magenta arrows denote the negative sub-gradient direction for positive samples. Left: Contrastive embedding. Middle: Triplet embedding. Right: Proposed structured embedding.[]{data-label="fig:failure_mode"}]({failure_mode}){width="4in" height="1.6in"} Fig.\[fig:failure\_mode\] shows failure cases in 2D profile with samples from three different classes, visualized by pink circles, green squares, and magenta triangles, respectively. The contrastive embedding has failure conditioned that randomly sampled negative $y_j$ is collinear [^2] with examples from a third class (purple triangles). For triplet embedding, the degenerated case happens when a negative $y_j$ is within the margin bound with respect to the anchor $x_i$ and its positive $y_i$. In this situation, both contrastive and triplet embedding incorrectly enforce the gradient direction of positives towards examples from the third class. By contrast, through explicitly mining hard negatives within the margin w.r.t. the positive $x_i$, the proposed structured embedding can push the positives towards the correct direction. Margin maximization. Hard negative mining on mini-batches is equivalent to computing the loss augmented inference, which promotes margin maximization in pairwise/triplet units. *Proof*. Following the definitions in Eq., the condition of zero training error can be compactly written as a set of non-linear constraints $$\label{eq:nonlinear} \forall i: \max_{\by\in \mathcal{Y} \char`\\ \by_i} \{\langle\bw, H(\bx_i,\by)\rangle\} < \langle\bw, H(\bx_i,\by_i)\rangle.$$ where $\mathcal{Y}$ contains training samples from cross-camera view against $\bx_i$. $H(.)$ denotes Hamming distance. Each non-linear inequality in Eq. can be equivalently replaced by $|\mathcal{Y}|-1$ linear inequalities, and thus we have $$\label{eq:margin} \begin{split} &\forall i, \forall \by \in \mathcal{Y} \char`\\ \by_i: \langle \bw, \delta H_i(\by)\rangle <0;\\ &\delta H_i(\by) \equiv H(\bx_i,\by)-H(\bx_i,\by_i). \end{split}$$ Recall Eq. that the hard negative mining is equivalent to augmenting the loss as $\bar{H}_i (\by)=H(\bx_i,\by)-H(\bx_i,\by_i)+H(\by_i,\by)$. Thus, the linear constraint in Eq. is updated as $$\label{eq:augmented_margin} \begin{split} &\forall i, \forall \by \in \mathcal{Y} \char`\\ \by_i: \langle \bw, \delta \bar{H}_i(\by)\rangle <0;\\ & \Leftrightarrow \langle \bw, \delta H_i(\by)\rangle + \langle \bw, H(\by_i,\by) \rangle<0. \end{split}$$ In Eq., since the term $\langle \bw, H(\by_i,\by) \rangle \geq 1-\epsilon_i$, $\epsilon_i\geq 0$ is a small slack variable, the term $\langle \bw, \delta H_i(\by)\rangle$ is imposed a more tight constraint on its margin maximization. $\Box$ ![image](precision_bits_cuhk03){width="1.8in" height="1.5in"} ![image](precision_recall_cuhk03){width="1.8in" height="1.5in"} ![image](precision_top_cuhk03){width="1.8in" height="1.5in"}\ (a) (b) (c)\ ![image](precision_bits_market){width="1.8in" height="1.5in"} ![image](precision_recall_market){width="1.8in" height="1.5in"} ![image](precision_top_market){width="1.8in" height="1.5in"}\ (a) (b) (c)\ ---------------------------------- ----------- ----------- ----------- ----------- ----------- ----------- ----------- ----------- Method 24 bits 32 bits 48 bits 128 bits 24 bits 32 bits 48 bits 128 bits Ours **0.579** **0.594** **0.602** **0.601** **0.452** **0.466** **0.481** **0.482** SFLHC [@Lai:hash-deep] 0.428 0.468 0.472 0.476 0.365 0.372 0.377 0.378 DSRH [@DeepSemanticHash] 0.494 0.486 0.482 0.484 0.403 0.421 0.426 0.423 DRSCH [@DeepRegularlizedHash] 0.509 0.517 0.523 0.521 0.421 0.443 0.452 0.447 CNNH [@Xia:hash2014] 0.403 0.411 0.417 0.414 0.305 0.337 0.346 0.345 KSH+AlexNet [@Liu:kernel-hash] 0.301 0.339 0.356 0.357 0.264 0.282 0.288 0.281 MLH+AlexNet [@Norouzi:loss-hash] 0.262 0.295 0.299 0.302 0.224 0.257 0.269 0.273 BRE+AlexNet [@Kulis:nips2009] 0.206 0.215 0.237 0.239 0.185 0.196 0.211 0.210 ---------------------------------- ----------- ----------- ----------- ----------- ----------- ----------- ----------- ----------- Experiments {#sec:exp} =========== In this section, we conduct extensive evaluations of the proposed architecture on two largest datasets in person re-identification: **CUHK03** [@FPNN] and **Market-1501** [@Market1501]. Experimental Settings --------------------- #### Datasets Person re-identification comes with a number of benchmark datasets such as VIPeR [@Gray2007Evaluating], PRID2011 [@Hirzer2011Person], and iLIDS [@Zheng2009Associating]. However, these datasets are moderately small/medium-sized, rendering them not suitable to be the test bed for our fast hashing learning framework. More recently, to facilitate deep learning in person re-identification, two large datasets [*i.e., *]{}CUHK03 and Market1501 are contributed with more identities captured by multiple cameras in more realistic conditions. - The **CUHK03** dataset [@FPNN] includes 13,164 images of 1360 pedestrians. The whole dataset is captured with six surveillance camera. Each identity is observed by two disjoint camera views, yielding an average 4.8 images in each view. This dataset provides both manually labeled pedestrian bounding boxes and bounding boxes automatically obtained by running a pedestrian detector [@DetectionPAMI]. In our experiment, we report results on labeled data set. - The **Market-1501** dataset [@Market1501] contains 32,643 fully annotated boxes of 1501 pedestrians, making it the largest person re-id dataset to date. Each identity is captured by at most six cameras and boxes of person are obtained by running a state-of-the-art detector, the Deformable Part Model (DPM) [@MarketDetector]. The dataset is randomly divided into training and testing sets, containing 750 and 751 identities, respectively. #### Competitors We present quantitative evaluations in terms of searching accuracies and compare our method with seven state-of-the-art methods: - Kernel-based Supervised Hashing (KSH) [@Liu:kernel-hash]: KSH is a kernel based method that maps the data to binary hash codes by maximizing the separability of code inner products between similar and dissimilar pairs. In particular, KSH adopts the kernel trick to learn nonlinear hash functions on the feature space. - Minimal Loss Hashing (MLH) [@Norouzi:loss-hash]: MLS is working by treating the hash codes ad latent variables, and employs the structured prediction formulation for hash learning. - Binary Reconstructive Embedding (BRE) [@Kulis:nips2009]: Without requiring any assumptions on data distributions, BRE directly learns the hash functions by minimizing the reconstruction error between the distances in the original feature space and the Hamming distances in the embedded binary space. - CNNH [@Xia:hash2014]: is a supervised hashing method in which the learning process is decomposed into a stage of learning approximate hash codes, followed by a second stage of learning hashing functions and image representations from approximate ones. - Simulaneous Feature Learning and Hash Coding based on CNNs (SFLHC) [@Lai:hash-deep]: SFLHC is a deep architecture consisting of stacked convolution layers and hashing code learning module. It adopts a triplet ranking loss to preserve relative similarities. - Deep Semantic Ranking Hashing (DSRH) [@DeepSemanticHash]: DSRH is a recently developed method that incorporates deep feature learning into hash framework in order to preserve multi-level semantic similarity between multi-label images. Also, their network is optimized on a triplet ranking embedding. - Deep Regularized Similarity Comparison Hashing (DRSCH) [@DeepRegularlizedHash]: DRSCH is a deep framework which aims to generate bit-scalabel hash codes directly from raw images. Their network is optimized by triplet ranking loss, and hash codes are regularized by adjacency consistency. The first three methods are conventional supervised methods and the last three are based on deep learning framework. The results of these competitors are obtained by the implementations provided by their authors. For fair comparison on three supervised methods [*i.e., *]{}KSH, MLH, and BRE, we extract CNN features for person images using AlexNet [@Krizhevsky2012Imagenet], and feed the feature vectors from the last fully-connected layer (4096-dim) to MLH and BRE, denoted as KSH+AlexNet, MLH+AlexNet, BRE+AlexNet, respectively. ![The difference between Average Precision (AP) and CMC measurements. The green and red boxes represent the position of true and false matches in rank lists. For all three rank lists, CMC curve at rank 1 remains 1 whilst AP=1 (rank list A), 1 (rank list B), and 0.7 (list C), respectively.[]{data-label="fig:AP_CMC"}](AP_CMC){width="3in" height="1.2in"} #### Evaluation Protocol We adopt four evaluation metrics in the experiments: Mean Average Precision (MAP), Precision curves with Hamming distance within 2, Precision-Recall curves, and Precision curves with respect to varied number of top returned samples. In person re-identification, a standard evaluation metric is Cumulated Matching Characteristics (CMC) curve, which shows the probability that a correct match to the query identity appears in different-sized candidate lists. This measurement is, however, is valid only in the single-shot setting where there is only one ground truth match for a given query (see an example in Fig.\[fig:AP\_CMC\]). In the case of one-shot, precision and recall are degenerated to be the same manner. Nonetheless, given multiple ground truths regarding to a query identity, the CMC curve is biased due to the fact that the recall issue is not considered. For instance, two rank lists A and B in Fig.\[fig:AP\_CMC\] can yield their CMC value equal to 1 at rank=1, respectively, whereas CMC curves fail to provide a fair comparison of the quality between the two rank lists. By contrast, Average Precision (AP) can quantitatively evaluate the quality of rank list for the case of multi-ground-truth. For Market-1501 (CUHK03) dataset, there are on average 14.8 (4.8) cross-camera ground truths for each query. Thus, we employ Mean Average Precision (MAP) to evaluate the overall performance. For each query, we calculate the area under the Precision-Recall curve, which is known as Average Precision (AP). Then, MAP is calculated as the mean value of APs over all queries. We have the definition of MAP in the following $$\label{eq:MAP} MAP(Q)=\frac{1}{Q}\sum_{j=1}^{|Q|}\frac{1}{m_j}\sum_{k=1}^{m_j}Precision(R_{jk}),$$ where $Q$ denotes a set of queries, and $\{d_1,\dots,d_{mj}\}$ are a set of relevant items with respect to a given query $q_j\in Q$. $R_{jk}$ is the set of ranked retrieval results from the top results until item $d_k$ is retrieved. Given a query, the precision with hamming distance within 2 (@ $r$-bits) w.r.t. the returned top $N$ nearest neighbors is defined as $$\small Precision (||\cdot||_H<=2) @ N =\frac{\sharp \left(imgs \cap ||imgs-query||_H<=2\right)}{N}$$ where $imgs$ denote similar images to the query, the hamming distance between two binary vectors is the number of coefficients where they differ. The four types of metrics are widely used to evaluate hashing models [@Liu:kernel-hash; @Lai:hash-deep]. #### Implementation Details We implemented our architecture using the Theano [@Theano] deep learning framework with contrastive, triplet, and the proposed structured loss. The batch size is set to 128 for contrastive and our method and to 120 for triplet. Network training converges in roughly 22-24 hours on NVIDIA GTX980. All training and test images are normalized to 160 by 60. We augment the training data by performing random 2D translation, as also done in [@FPNN; @JointRe-id]. For an original image of size $W\times H$, we sample 5 images around the image center, with translation drawn from a uniform distribution in the range $[-0.05H,0.05H]\times [-0.05W,0.05W]$. In training, we exhaustively use all the positive pairs of examples and randomly generate approximately equal number of negative pairs as positives. In Market-1501, there are 12,936 images for training and 19,732 images for test, corresponding to 750 and 751 identities, respectively. In CUHK03 dataset, we randomly partition the dataset into training, validation, and test with 1160, 100, and 100 identities, respectively. During testing, for each identity, we select one query image in each camera. The search process is performed in a *cross-camera* mode, that is, relevant images captured in the same camera as the query are regarded as “junk" [@ObjRetrieval], which means that this image has no influence to re-identification accuracy. In this scenario, for Market-1501 dataset, each identity has at most 6 queries, and there are 3,363 query images in total. For CUHK03 dataset, each identity has at most 2 queries, and there are 200 query images in total. In our implementation, we use all positive anchor positive pairs regarding to each identity. In pairwise training, anchor negative pairs are generated by randomly selecting a sample from a different identity with respect to the anchor’s identity. The same sampling scheme is applied on triplet selection. To add meaningful hard negatives into mini-batch in our model, we select hard neighbors from gallery view for each training image in a positive pair. Specifically, for an anchor $\mathcal{I}$ and its positive $\mathcal{I}^+$, their hard negatives $\mathcal{I}^-$s are selected such that $||s_{\mathcal{I} } - s_{\mathcal{I}^+}||_2^2 < ||s_{\mathcal{I} } - s_{\mathcal{I}^-} ||_2^2$, where $s_{ (\cdot) }$ is a visual descriptor and in our experiment we use SIFT features at the beginning of training [^3]. Since features are updated continuously as network is on training, $s_{ (\cdot) }$ corresponds to feature extracted after each 50 epochs. [|c|c|c|c|c|]{} Method & 24 bits & 32 bits & 48 bits & 128 bits\ \ FC2 & 0.529 & 0.546 & 0.571 & 0.584\ FC1+FC2 & **0.579** & **0.594** & **0.602** &**0.601**\ \ FC2 & 0.417 & 0.420 & 0.439 & 0.437\ FC1+FC2 & **0.452** & **0.466** & **0.481** & **0.482**\ Results on Benchmark Datasets ----------------------------- We test and compare the search accuracies of all methods against two datasets. Comparison results are reported in Table \[tab:MAP\] and Figs. \[fig:results\_cuhk03\]–\[fig:results\_market\]. We can see that - On the two benchmark datasets, the proposed method outperforms all supervised learning baselines using CNN features in terms of MAP, precision with Hamming distance 2, precision-recall, and precision with varying size of top returned images. For instance, compared with KSH + AlexNet, the MAP results of the proposed method achives a gain from 35.6% 58.5%, 28.8% 48.1% with 48 bits on CUHK03 and Market-1501, respectively. - Comparing with CNNH [@Xia:hash2014], which is a two-stage deep network based hashing method, our method indicates a better searching accuracies. Specifically, the MAP results achieve a relative increase by a margin of 16% and 13% on two datasets, respectively. This observation can verify that jointly learning features and hashing codes are beneficial to each other. - Comparing with the most related competitors DSRH [@DeepSemanticHash] and DRSCH [@DeepRegularlizedHash], our structured prediction suits well to SGD and thus achieves superior performance. For example, in terms of MAP on CUHK03 dataset, a notable improvement can be seen from 49.4% (50.9%) to 54.7%, compared with DSRH [@DeepSemanticHash] (DRSCH [@DeepRegularlizedHash]). We also conduct self-evaluation of our architecture with skip layer connected to hash layers and its alternative with only the second fully connected layer. As can be seen in Table \[tab:FC\_layer\], the results of the proposed architecture outperforms its alternative with only the second fully connected layer as input to the hash layer. One possible reason is the hash layer can see multi-scale features by connecting to the first and second fully connected layers (features in the FC2 is more global than those in FC1). And adding this bypass connections can reduce the possible information loss in the network. Comparison with State-of-the-art Approaches ------------------------------------------- In this section, we evaluate our method by comparing with state-of-the-art approaches in person re-identification. Apart from the above hashing methods, seven competitors are included in our experiment, which are FPNN [@FPNN], JointRe-id [@JointRe-id], KISSSME [@KISSME], SDALF [@Farenzena2010Person], eSDC [@eSDC], kLFDA [@Xiong2014Person], XQDA [@LOMOMetric], DomainDropout [@DomainDropout], NullSpace [@NullSpace-Reid] and BoW [@Market1501]. For KISSME [@KISSME], SDALF [@Farenzena2010Person], eSDC [@eSDC], kLFDA [@Xiong2014Person] and BoW model [@Market1501], the experimental results are generated by their suggested features and parameter settings. For XQDA [@LOMOMetric] and NullSpace [@NullSpace-Reid], the Local Maximal Occurrence (LOMO) features are used for person representation. The descriptor has 26,960 dimensions. FPNN [@FPNN] is a deep learning method with the validation set adopted to select parameters of the network. JointRe-id [@JointRe-id] is an improved deep learning architecture in an attempt to simultaneously learn features and a corresponding similarity metric for person re-identification. DomainDropout [@DomainDropout] presents a framework for learning deep feature representations from multiple domains with CNNs. We also extract the intermediate features from the last fully-connected layer, denoted as Ours (FC), to evaluate the performance without hash layer. To have fair comparison with DomainDropout [@DomainDropout], we particularly leverage training data from CUHK03, CUHK01 [@CUHK01] with domain-aware dropout, and Market-1501, denoted as Ours (DomainDropout). Table \[tab:compare\_state\_art\] displays comparison results with state-of-the-art approaches, where all of the Cumulative Matching Characteristics (CMC) Curves are single-shot results on CUHK03 dataset whilst multiple-shot on Market1501 dataset. All hashing methods perform using 128 bits hashing codes, and the ranking list is based on the Hamming distance. We can see that on Market-1501 dataset our method outperforms all baselines on rank 1 recognition rate except NullSpace [@NullSpace-Reid]. The superiority of NullSpace [@NullSpace-Reid] on Market-1501 comes from enough samples in each identity, which allows it to learn a discriminative subspace. Our result (48.06%) is very comparative to NullSpace [@NullSpace-Reid] (55.43%) while the time cost is tremendously reduced, as shown in Table \[tab:test\_time\]. Besides, the performance of our model without hash layer (Ours (FC)) is consistently better than that with hashing projection. This is mainly because the dimension reduction in hashing layer and quantization bring about certain information loss. On CUHK03 dataset, DomainDropout [@DomainDropout] achieves the best performance in recognition rate at rank from 1 to 10. This is mainly because DomainDropout [@DomainDropout] introduces a method to jointly utilize all datasets in person re-identification to produce generic feature representation. However, this action renders their model extremely expensive in training given a variety of datasets varied in size and distributions. To this end, we test the average testing time of our model and competing deep learning methods, and report results in Table \[tab:test\_time\]. The testing time aggregates computational cost in feature extraction, hash code generation, and image search. For all the experiments, we assume that every image in the database has been represented by its binary hash codes. In this manner, the time consumption of feature extraction and hash code generation is mainly caused by the query image. It is obvious that our model achieves comparable performance in terms of efficiency in matching pedestrian images. Our framework runs slightly slower than DRSCH and SFLHC due to the computation of structured loss on each mini-batch. -------------------------------------------------- ----------- ----------- ----------- ----------- ----------- ----------- ----------- ----------- Method r=1 r=5 r=10 r=20 r=1 r=5 r=10 r=20 Ours 37.41 61.28 77.46 88.42 48.06 61.23 75.67 87.06 Ours (FC2) 43.20 69.07 84.24 90.92 50.12 **63.50** **76.82** **89.24** Ours (DomainDropout) **74.20** **92.27** 94.24 95.92 **58.12** **68.50** **80.82** **92.24** SFLHC [@Lai:hash-deep] 12.38 30.52 49.34 71.55 37.74 59.09 74.25 86.52 DSRH [@DeepSemanticHash] 9.75 28.10 47.82 67.95 34.33 59.82 71.27 86.09 DRSCH [@DeepRegularlizedHash] 20.84 49.39 72.66 83.03 41.25 58.98 76.04 85.33 CNNH [@Xia:hash2014] 8.27 22.53 45.09 59.74 16.46 39.95 51.24 71.23 KSH+AlexNet [@Liu:kernel-hash] 5.65 15.71 22.75 34.68 12.57 31.22 48.66 66.72 MLH+AlexNet [@Norouzi:loss-hash] 5.75 15.62 27.61 42.68 10.89 29.93 46.78 66.32 BRE+AlexNet [@Kulis:nips2009] 4.91 11.24 17.83 29.20 12.65 32.68 49.08 67.70 FPNN [@FPNN] 20.65 50.09 66.42 80.02 - - - - JointRe-id [@JointRe-id] 54.74 86.71 91.10 **97.21** - - - - KISSME [@KISSME] 14.17 41.12 54.89 70.09 40.47 59.35 75.26 83.58 SDALF [@Farenzena2010Person] 5.60 23.45 36.09 51.96 20.53 47.82 62.45 80.21 eSDC [@eSDC] 8.76 27.03 38.32 55.06 33.54 58.25 74.33 84.57 kLFDA [@Xiong2014Person] 47.25 64.58 82.36 89.17 - - - - BoW [@Market1501] 24.33 58.42 71.28 84.91 47.25 62.71 75.33 86.42 DomainDropout [@DomainDropout] 72.60 92.22 **94.50** 95.01 - - - - NullSpace (LOMO [@LOMOMetric]) [@NullSpace-Reid] 58.90 85.60 92.45 96.30 55.43 - - - XQDA (LOMO) [@LOMOMetric] 52.20 82.23 92.14 96.25 43.79 - - - -------------------------------------------------- ----------- ----------- ----------- ----------- ----------- ----------- ----------- ----------- Method CUHK03 (ms) Market-1501 (ms) ---------------------------------- --------------- ------------------- Ours 4.617 (4.982) 7.374 (7.902) SFLHC [@Lai:hash-deep] 4.241 (4.782) 5.892 (6.417) DSRH [@DeepSemanticHash] 5.765 (6.019) 7.887 (8.445) DRSCH [@DeepRegularlizedHash] 2.332 (2.816) 3.609 (3.973) CNNH [@Xia:hash2014] 5.359 (5.743) 6.943 (7.410) KSH+AlexNet [@Liu:kernel-hash] 7.279 (7.805) 9.046 (9.537) MLH+AlexNet [@Norouzi:loss-hash] 6.727 (7.198) 8.092 (8.545) BRE+AlexNet [@Kulis:nips2009] 6.765 (7.214) 9.072 (9.613) FPNN [@FPNN] $\infty$ $\infty$ JointRe-id [@JointRe-id] $\infty$ $\infty$ DomainDropout [@DomainDropout] $\infty$ $\infty$ kLFDA [@Xiong2014Person] - $43.4\times 10^3$ XQDA [@LOMOMetric] - $1.6\times 10^3$ MFA [@Xiong2014Person] - $43.2\times 10^3$ NullSpace [@NullSpace-Reid] - $31.3\times 10^3$ : Comparison on the average testing time (millisecond per image) by fixing the code length to be 48 (128) bits. $\infty$ indicates that the computational cost is too high to be estimated and “-" indicates that the result is not available.[]{data-label="tab:test_time"} Convergence Study ----------------- ![Convergence study on two benchmark datasets. It is obvious that our structured embedding has fast convergence compared with contrastive and triplet embeddings.[]{data-label="fig:convergence"}](convergence_cuhk03 "fig:"){width="1.7in" height="1.3in"} ![Convergence study on two benchmark datasets. It is obvious that our structured embedding has fast convergence compared with contrastive and triplet embeddings.[]{data-label="fig:convergence"}](convergence_market "fig:"){width="1.7in" height="1.3in"}\ (a) (b)\ In this experiment, we study the convergence speed of optimizing contrastive, triplet, and structured embedding, respectively. The average loss values over all mini-batches are computed on three kinds of embeddings, as shown in Fig.\[fig:convergence\]. We can see that the proposed structured embedding is able to converge faster than other two embeddings. This can be regarded as the response to the augment from hard negatives which provide informative gradient direction for positives. Conclusion {#sec:con} ========== In this paper, we developed a structured deep hashing architecture for efficient person re-identification, which jointly learn both CNN features and hash functions/codes. As a result, person re-identification can be resolved by efficiently computing and ranking the Hamming distances between images. A structured loss function is proposed to achieve fast convergence and more stable optimization solutions. Empirical studies on two larger benchmark data sets demonstrate the efficacy of our method. In our future work, we would explore more efficient training strategies to reduce training complexity, and possible solutions include an improved loss function based on local distributions. [^1]: Pedestrian images typically undergo compounded variations in the form of human appearance, view angles, and human poses. [^2]: Three or more points are said to be collinear if they lie on a single straight line. [^3]: To extract SIFT/LAB features, we first rescale pedestrian images to a resolution of $48\times 128$ in order to remove some background regions. SIFT and color features are extracted over a set of 14 dense overlapping $32\times 32$-pixels regions with a step stride of 16 pixels in both directions. Thus, we have the feature descriptor vector with length 9408.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Active galactic nuclei present continuum and line emission. The former is produced by the accretion disk and the jets, whereas the latter is originated by gas located close to the super-massive black hole. The small region where the broad lines are emitted is called the broad-line region. The structure of this region is not well known, although it has been proposed that it may be formed by small and dense ionized clouds surrounding the supermassive black-hole. In this work, we study the interaction of one cloud from the broad line region with the jet of the active galactic nuclei. We explore the high-energy emission produced by this interaction close to the base of the jet. The resulting radiation may be detectable for nearby non-blazar sources as well as for powerful quasars, and its detection could give important information on the broad line region and the jet itself.' address: - | Instituto Argentino de Radioastronomía (CCT La Plata, CONICET), C.C.5, (1894)\ Villa Elisa, Buenos Aires, Argentina\ Facultad de Ciencias Astronómicas y Geofísicas, UNLP, Argentina\ [email protected] - | Max Planck Institut für Kernphysik, Saupfercheckweg 1\ Heidelberg 69117, Germany [email protected] - | Instituto Argentino de Radioastronomía (CCT La Plata, CONICET), C.C.5, (1894)\ Villa Elisa, Buenos Aires, UNLP, Argentina\ Facultad de Ciencias Astronómicas y Geofísicas, UNLP, Argentina\ [email protected] author: - 'Anabella T. Araudo' - 'Valentí Bosch-Ramon' - 'Gustavo E. Romero' title: 'Jet-Cloud Interactions in AGNs' --- Introduction ============ Active galactic nucleus (AGNs) are systems mainly composed by a supermassive black-hole (SMBH), an accretion disk, and bipolar relativistic jets. AGNs produce non-thermal continuum emission along the whole spectrum, from radio to $\gamma$-rays. The high-energy non-thermal radiation is expected to come from the jets, which are formed by a magnetized plasma moving relativistically. Besides the emission in the continuum, AGNs present optic and UV lines. Some of these lines are observed with a broad FWHM, with an associated velocity for the emitting gas of $\sim 10^9$ cm s$^{-1}$. The region where these lines are formed is called the broad line region (BLR), and surrounds the SMBH. The size of the BLR is related to its luminosity, $L_{\rm BLR}$ \[\]. For instance, for $L_{\rm BLR}\sim 10^{44}$ erg s$^{-1}$ (Faranoff-Riley (FR) I case), $R_{\rm BLR}\sim 6\times 10^{16}$ cm, and for $L_{\rm BLR}\sim 10^{46}$ erg s$^{-1}$ (FR II case) $R_{\rm BLR}\sim 5\times 10^{17}$ cm. The BLR is thought to be filled with a clumpy medium composed by cold clouds ($T\sim 10^4-10^5$ K) of radius $R_{\rm c}\sim 10^{13}$ cm \[\]. In this work we study the interaction between a cloud of the BLR with the relativistic jet close to the SMBH in an AGN. Assuming standard values of the cloud parameters, and adopting a hydrodynamical supersonic jet, we estimate the high-energy emission produced by this interaction for two different kinds of AGN: a non-blazar source of FR I type, and a powerful flat-spectrum radio quasar of FR II type. The physical scenario ===================== A cloud with a radius $R_{\rm c}$ that moves with a velocity $v_{\rm c}=10^9$ cm s$^{-1}$, completely enters into the jet in a time $t_{\rm c}\sim 2\,R_{\rm c}/v_{\rm c}\sim 2\times 10^4$ s. For an effective penetration of the cloud into the jet, a huge contrast between the cloud and the jet densities ($\chi=n_{\rm c}/n_{\rm j}\gg 1$) is necessary. The cloud density $n_{\rm c}$ is fixed to $\sim 10^{10}$ cm$^{-3}$ \[\] and the density of the jet, $n_{\rm j}$, is determined through the equation: $L_{\rm j}=\sigma_{\rm j}(\Gamma-1)n_{\rm j} m_{\rm p} v_{\rm j} c^2$, where $v_{\rm j}\sim c$ is the velocity of the jet, $\Gamma\sim 10$ the jet Lorentz factor, and $\sigma_{\rm j}=\pi R_{\rm j}^2$ is the section of the jet at the interaction height $z_{\rm int}$, where $R_{\rm j}\sim 0.1 z_{\rm int}$. In order to obtain $n_{\rm j}$ we need to fix $z_{\rm int}$ at which the cloud penetrates into the jet. When the cloud penetrates into the jet, a shock is formed and propagates through the cloud at a velocity $v_{\rm sh}=v_{\rm j}((\Gamma-1)/\chi)^{1/2}$. In a time $t_{\rm cc}\sim 2\,R_{\rm c}/v_{\rm sh}$ the shock crosses the whole cloud. We focus on the stage when the cloud is inside the jet ($t_{\rm cc}>t_{\rm c}$) at the zero order approximation, implying that the interaction should take place at least at $z_{\rm int}=2.5\times 10^{15}$ and $2.5\times 10^{16}$ cm for an FR I ($L_{\rm j}\sim 10^{44}$ erg s$^{-1}$) and an FR II ($L_{\rm j}\sim 10^{46}$ erg s$^{-1}$), respectively. At such $z_{\rm int}$, we obtain for all cases $n_{\rm j}=1.2\times 10^6$ cm$^{-3}$ and $\chi\sim 10^4$. The shock heats the cloud material up to a temperature $T\sim 2\times 10^9$ K. The hot plasma cools via thermal Bremsstrahlung radiation with a thermal luminosity $\sim 10^{38}$ erg s$^{-1}$, peaking at soft $\gamma$-rays. A bow shock is also formed in the jet, reaching the steady state at a distance $\sim R_{\rm c}$ from the cloud in a time $t_{\rm bs}\sim R_{\rm c}/v_{\rm j}\ll t_{\rm cc}$. The cloud could escape from the jet in a time $t_{\rm j}\sim 2\,R_{\rm j}/v_{\rm c}\sim 5\times 10^5$ (FR I case) and $5\times 10^6$ s (FR II case), being $t_{\rm j}\gg t_{\rm cc}$, although Kelvin-Helmholtz and Rayleigh-Taylor instabilities can destroy the cloud in a time $t_{\rm KH/RT}\sim$ few times $t_{\rm cc}$, shorter than $t_{\rm j}$. We note that the shocked cloud is also accelerated by the jet and might reach a velocity $\sim v_{\rm j}$, but it is likely that before this happens the cloud escapes from the jet. Non-thermal emission ==================== In the bow shock, particles can be accelerated via relativistic Fermi-I mechanism. Given the much lower velocity of the cloud shock, we will neglect at this stage its role to accelerate particles. The acceleration rate of particles, for which we adopt here a phenomenological prescription: $\dot{E}=0.1\,qBc$, depends on the magnetic field $B$ in the post-shock region of the bow shock \[\]. We will consider three different cases, varying the value of $B$ and the luminosity of the jet: i) case 0, with $L_{\rm j}\sim 10^{44}$ erg s$^{-1}$ and $B=2.8\times 10^{-2}$ G (FR I case; dominant BLR photon energy density); ii) case I, with $L_{\rm j}\sim 10^{46}$ erg s$^{-1}$ and $B=4.1\times 10^{-3}$ G (FR II case; dominant BLR photon energy density); and iii) case II, with $L_{\rm j}\sim 10^{46}$ erg s$^{-1}$ and $B=1.1\times 10^3$ G (FR II case; dominant magnetic energy density, 10% of equipartition with the post-shock matter). The magnetic fields for cases 0 and I are very much below equipartition. These cases are considered to explore the situation when external Compton dominates the radiation output, where the external (BLR) energy density is $u_{\rm BLR}\sim L_{\rm BLR}/\pi R_{\rm BLR}^2 c$. In case II, for $B$-values close to equipartition, synchrotron emission will dominate. Work on the case when synchrotron self-Compton is the dominant radiation channel is on-going. The maximum energy ($E_{\rm max}$) achieved by electrons is constrained by the escape of these particles from the bow-shock region, advected by the shocked material of the jet on a time $t_{\rm esc}\sim 3\,R_{\rm c}/c$ (case 0: $2\times 10^{13}$ eV and case I: $4\times 10^{12}$ eV), and by synchrotron losses (case II: $5\times 10^{11}$ eV). For protons the maximum energy is determined by the size of the acceleration region, $\sim R_{\rm c}$ (case 0: $8\times 10^{13}$ eV, case I: $10^{13}$ eV, and case II: $3\times 10^{18}$ eV). We assume here that the 20% of the $L_{\rm j}$ fraction transferred to the cloud is converted into (non-thermal) luminosity of the accelerated particles, i.e. $L_{\rm NT}=0.2\,(\sigma_{\rm c}/\sigma_{\rm j})\,L_{\rm j}$. We determine the constant $K_{e,p}$ of the injection function $Q$, assuming a power-law with an index $p=-2.2$ and a cut-off at higher energies: $Q_{e,p}=K_{e,p}\,E^{-p}\,e^{-E/E_{\rm max}}$. The electron energy distribution, $N_e$, is determined by the escape of particles and radiation losses (synchrotron and inverse Compton (IC)), reaching the steady state on a time $\ll t_{\rm cc}$. The steady distribution of these relativistic leptons is $N_e=Q_e\,t_{\rm esc}\propto E^{-p}$ (in cases 0 and I; escape dominance) and $N_e=Q_e\,t_{\rm synch}\propto E^{-p-1}$ (in case II; synchrotron cooling). On the other hand, protons escape from the bow-shock region after losing a negligible fraction of their energy by $pp$ interactions. As noted above, the most important emission process is IC scattering in cases 0 and I. In the former, the achieved luminosities are $L_{\rm synch}\sim 10^{36}$ and $L_{\rm IC}\sim 10^{38}$ erg s$^{-1}$ (peaking in hard X-rays and in $\gamma$-rays, respectively), and in the latter, $L_{\rm synch}\sim 10^{34}$ and $L_{\rm IC}\sim 10^{37}$ erg s$^{-1}$ (peaking in soft X-rays and in $\gamma$-rays, respectively). In case II, corresponding to the larger value of $B$, the synchrotron emission is the most important channel of energy loss, peaking in the the infrared and decreasing smoothly afterwards up to soft $\gamma$-rays. In this case, synchrotron and IC luminosities reach values $\sim 10^{39}$ and $10^{32}$ erg s$^{-1}$, respectively. The IC maximum is always in the sub-TeV range, given the strong photon-photon absorption in the BLR photon field, producing pairs that will also generate synchrotron and IC radiation (e.g. \[\]). Regarding relativistic protons, since the density of particles in the bow-shock region is much lower than that of the cloud density, $pp$ interactions are not an effective radiative process. For this reason, the most energetic protons ($E_p\sim E_{\rm max}$) can reach the cloud via diffusion before being advected away from the bow-shock region. Inside the cloud, the relativistic protons lose energy via $pp$ collisions, yielding (absorbed) $\gamma$-ray luminosities $L_{pp}\sim 10^{36}$, $10^{38}$ and $10^{39}$ erg s$^{-1}$ for the cases 0, I and II, respectively. Case II is an interesting one since radiation above 100 TeV may be detectable. We note that secondary leptons and neutrinos with luminosities similar to those of $\gamma$-rays, would be also produced inside the cloud due to $pp$ interactions, and their study will be carried out in future work. Final remarks ============= The relation flux/luminosity/distance: $F\sim 10^{-12} (L_{38}/d_{\rm Mpc}^2)$ erg s$^{-1}$ cm$^2$, and the luminosity values given above, show that one-cloud/jet interaction fluxes predicted in the present contribution could be detectable at X- or $\gamma$-rays up to distances of $\sim 10$ Mpc. The interaction of a cloud at the adopted $z_{\rm int}$ could lead either to persistent but variable emission if there were many clouds interacting with the jet \[\], or to sporadic emission with a certain duty cycle if clouds interact with the jet from time to time. For the parameters adopted here, the jet-cloud interaction would be a persistent activity if the number of clouds is $N_{\rm c}>10^8$ (for an FR II and assuming a life time of a cloud in the jet of $\sim t_{\rm j}$), and sporadic if $N_{\rm c}$ is smaller. In the former case, the actual luminosities should be obtained here multiplying by $(N_{\rm c}/10^8)$, and in the latter, the interaction duty cycle would be proportional to $100\times (N_{\rm c}/10^8)$%. Notice that at larger $z_{\rm int}$, the emission will be reduced even if a larger number of clouds could penetrate the jet, given the smaller cloud to jet section ratio. Finally, we remain that given the properties of the emitter, the jet-cloud interaction radiation is not beamed and almost isotropic. Acknowledgments {#acknowledgments .unnumbered} =============== A.T.A. thanks the Max Planck Institut für Kernphysik for its kind support and hospitality. G.E.R. and V.B-R. acknowledge support by DGI of MEC under grant AYA2007-6803407171-C03-01, as well as partial support by the European Regional Development Fund (ERDF/FEDER). V.B-R. gratefully acknowledges support from the Alexander von Humboldt Foundation. [0]{} S. Kaspi, D. Maoz, Dan and H. Netzer [*ApJ*]{} [ **629**]{} (2005) 61. M. Rees, H. Netzer and G. Ferland, 1989, [*ApJ*]{} [ **347**]{} (1989) 640. D. Kazanas, [*ApJ*]{} [**374**]{} (1989) 74. F. Tavecchio and G. Ghisellini, [*MNRAS*]{} [ **385**]{} (2008) L98. A. Asterberg, Y.A. Gallant, J.G. Kirk, A.W. Guthmann [*MNRAS*]{} [**328**]{} (2001) 393. F. Aharonian, D. Khangulyan and L. Costamante, [*MNRAS*]{} [**387**]{} (2008) 1206. S. Owocki, G.E. Romero, R. Townsend and A. Araudo, [*ApJ*]{} in press (2009) \[arXiv:0902.2278\].
{ "pile_set_name": "ArXiv" }
--- abstract: 'Optical stellar intensity interferometry with air Cherenkov telescope arrays, composed of nearly 100 telescopes, will provide means to measure fundamental stellar parameters and also open the possibility of model-independent imaging. In addition to sensitivity issues, a main limitation of image recovery in intensity interferometry is the loss of phase of the complex degree of coherence during the measurement process. Nevertheless, several model-independent phase reconstruction techniques have been developed. Here we implement a Cauchy-Riemann based algorithm to recover images from simulated data. For bright stars ($m_v\sim 6$) and exposure times of a few hours, we find that scale features such as diameters, oblateness and overall shapes are reconstructed with uncertainties of a few percent. More complex images are also well reconstructed with high degrees of correlation with the pristine image. Results are further improved by using a forward algorithm.' author: - | Paul D. Nuñez$^{1}$[^1], Richard Holmes$^{2}$, David Kieda$^{3}$, Stephan LeBohec$^{4}$\ $^{1,3,4}$University of Utah, Dept. of Physics & Astronomy, 115 South 1400 East, Salt Lake City, UT 84112-0830, USA\ $^{2}$Nutronics Inc., 3357 Chasen Drive, Cameron Park, CA 95682, USA\ title: High angular resolution imaging with stellar intensity interferometry using air Cherenkov telescope arrays --- techniques: high angular resolution - instrumentation: interferometers - stars: imaging - stars: fundamental parameters Introduction ============ Even though Stellar Intensity Interferometry (SII) was abandoned in the 1970’s, there has been a recent interest in reviving this technique, mainly due to the unprecedented $(u,v)$ plane coverage that future imaging air Cherenkov telescope (IACT) arrays will provide [@cta; @agis]. The possibility of probing stars at the sub-milliarcsecond scale and visible wavelengths has motivated new developments in instrumentation and simulations, the latter being the focus of this paper.\ Recent results obtained with amplitude (Michelson) interferometry have started to reveal stars as extended objects (e.g. @coast [@chara_review]), and with non-uniform light intensity distributions in the milliarcsecond scale. Such interesting results can be further investigated with SII taking advantage of the longer (km) baselines and relative ease of observing at shorter (blue) wavelengths. For example, measuring stellar diameters at different wavelengths, will make it possible to further investigate the wavelength dependence of limb darkening, [@Mozurkewich] and thus constrain stellar atmosphere models. Radii measurements with uncertainties of a few percent, along with spectroscopic measurements are necessary to constrain the position of stars in the HR diagram (e.g. @Kervella). With the methods described in this paper, we show that diameters can in principle be measured with accuracies better than $1\%$ when using realistic array configurations for future experiments such as CTA (Cherenkov Telescope Array). As another example we can consider fast rotating B stars, which are ideal candidates for imaging oblateness, pole brightening [@chara_review; @von_zeipel], radiatively driven mass loss [@Friend], and perhaps even pulsation modes [@pulsation]. The impact of rotation on stellar evolution is non-trivial, and several studies have been made in the subject (e.g. @martin [@maeder]). Images of rotating stars have become available in the past few years (e.g. @altair [@vega]), and measurements of oblateness with accuracies of a few percent have been made. We will show that this is comparable to what can be achieved with SII using large arrays of Cherenkov telescopes. There is also the case of interacting binaries, for which we can measure angular separation, diameters, and relative brightness. It may even be possible to measure mass transfer [@verhoelst]. Measurements of the angular separation in binaries is crucial for determining the masses of stars. These masses must be found to within $\sim 2\%$ [@Andersen] in order to test main sequence models. With the methods described in this paper, we show that angular separations can be found to within a few percent from reconstructed images. A more convenient and accurate analysis (beyond the scope of this paper) does not require imaging, and should yield sub-percent uncertainties, so that testing main sequence models will become possible with SII.\ In preparation for a large-scale SII observatory deployment, several laboratory experiments are in progress [@stephan.spie]. Their main goal is to measure light intensity correlation between two receivers. It is also worth mentioning the *Star Base* observatory (located in Grantsville, Utah) which consists of two $3\mathrm{m}$ light receivers separated by $24\,\mathrm{m}$ and which will be used to test high time resolution digital correlators, and to measure the second order degree of coherence for a few stars. Various analog and digital correlator technologies [@Dravins.timescale] are being implemented, and cross correlation of streams of photons with nanosecond-scale resolution has already been achieved.\ Intensity interferometry, unlike amplitude interferometry, relies on the correlation between intensity fluctuations averaged over the spectral band at electronic (nanosecond) time resolution. These averaged fluctuations are much slower than the (femtosecond) light wave period. This correlation is directly related to the complex degree of coherence $\gamma_{ij}$ as [@lipson] $$|\gamma_{ij}|^2=\frac{<\Delta I_i \,\Delta I_j>}{<I_i> <I_j>}.$$ Here, $<I_i>$ is the time average of the intensity received at a particular telescope $i$, and $\Delta I_i$ is the intensity fluctuation. Measuring a second-order effect results in lower signal-to-noise ratio when compared to amplitude interferometry [@holder2]. This sensitivity issue can be dealt with by using large light collection areas (such as those available with air Cherenkov telescopes), longer exposure times and baseline redundancy. The low frequency fluctuation can be interpreted classically as the beat formed by neighboring Fourier components. Since SII relies on low frequency fluctuations, which are typically several orders of magnitude smaller than the frequency of optical light, it does not rely on the phase difference between light waves, but rather in the difference between the relative phases of the two components at the detectors [@Brown3]. The main advantage is the relative insensitivity to atmospheric turbulence and the absence of a requirement for sub-wavelength precision in the optics and delay lines [@Brown3]. The complex mutual degree of coherence $\gamma$ is proportional to the Fourier transform of the radiance distribution of the object in the sky (Van Cittert-Zernike theorem). However, since with SII, the squared-modulus of $\gamma$ is the measurable quantity, the main disadvantage is that the phase of the Fourier transform is lost in the measurement process. The loss of phase information poses a severe difficulty, and images have in the past been reconstructed from the bispectrum technique, using monolithic apertures (e.g. @lawrence). The imaging limitations can be overcome using a model-independent phase recovery technique. Even though several phase reconstruction techniques exist [@fienup], we concentrate on a two dimensional version of the one dimensional analysis introduced by @Holmes, which is based on the theory of analytic functions, and in particular the Cauchy-Riemann equations.\ Following recent successes in Gamma ray astronomy, a next generation Cherenkov telescope array is in a preparatory stage. This project is currently known as CTA (Cherenkov telescope array) [@cta], and will contain between 50 and 100 telescopes with apertures ranging between $5\,\mathrm{m}$ and $25\,\mathrm{m}$. In this paper we investigate the sensitivity and imaging capabilities of SII implemented on such an atmospheric Cherenkov telescope array. We start with a discussion of sensitivity (section \[sensitivity\]), followed by a discussion of simulating noisy data as would be realistically obtained with such an array (section \[simulation\]). Since data have a finite sampling in the $(u,v)$ plane, we discuss our method of fitting an analytic function to the data in order to estimate derivatives which are needed for phase reconstruction (section \[fit\]). We then briefly discuss phase recovery (section \[phase\_rec\_sec\]) and proceed to quantify the reconstruction quality using several criteria. We start with the simple case of uniform disks (section \[disks\]) and progressively increase the degree of image complexity by including oblateness (section \[oblate\_sec\]), binary stars (section \[binary\_sec\]), and obscuring disks and spots (section \[complex\]). Sensitivity =========== The signal to noise ratio (SNR) for an intensity correlation measurement depends on the degree of correlation $\gamma$, the area $A$ of each of the light receivers, the spectral density $n$ (number of photons per unit area per unit time, per frequency), the quantum efficiency $\alpha$, the electronic bandwidth $\Delta f$, and the observation time $t$. The SNR can be expressed as [@Brown3] $$SNR=n(\lambda, T, m_v)\,A\;\alpha\;\gamma^2\sqrt{\Delta f t/2}. \label{snr}$$ The spectral density $n$ is related to the visual magnitude $m_v$ of the star as well as its temperature $T$ and observing wavelength $\lambda$. The spectral density $n(\lambda, T, m_v)$ is the number of black body photons per unit area, per unit frequency and per unit time. The dependence of the visual magnitude $m_v$ is found by recalling that the flux for a $0^{th}$ magnitude star with a temperature of $9550^{\circ}\mathrm{K}$ observed at $550\,\mathrm{nm}$ is $3.64\times 10^{-23}\mathrm{W m^{-2}Hz^{-1}}$ [@flux]. This in turn corresponds to a spectral density of $10^{-4}\,\mathrm{m^{-2} s^{-1} Hz^{-1}}$. The spectral density as a function of temperature (for different visual magnitudes and observing wavelengths) is shown in Fig. \[n\_vs\_T\], and we see that at constant visual magnitude and observing wavelength, higher temperatures correspond to higher spectral densities. We find that the increase in temperature $\Delta T(\lambda, T, \Delta m_v)$ is approximately $\frac{\lambda k T^2}{h c}\Delta m_v$ for the range of temperatures and wavelengths considered in Fig. \[n\_vs\_T\]. For example, at $400\,\mathrm{nm}$, a decrease of 1 visual magnitude is equivalent to increasing the temperature of the star from $5000\,\mathrm{K}$ to $5700\,\mathrm{K}$. For this reason high temperature objects are easier targets for SII.\ We use a preliminary design of the CTA project as an array configuration [@cta_simulation], which is shown in Fig. \[array\]. This array contains $N=97$ telescopes and $N(N-1)/2=4646$ baselines (many of which are redundant) which are shown in Fig. \[baselines\]. Each detector is assumed to have a light collecting area of $100\,\mathrm{m}^2$ and a light detection quantum efficiency of $\alpha=0.3$. Using a $\lambda/D$ criterion, we find that the largest baselines of $1.5\,\mathrm{km}$ resolve angular scales of $\sim 0.05\,\mathrm{mas}$ at $400\,\mathrm{nm}$. The smallest $48$ baselines of $35\,\mathrm{m}$ resolve angular scales of $\sim 2\,\mathrm{mas}$ . However, we show in section \[disks\], that the largest angular scales that can be realistically *imaged* with our analysis, in a model independent way, are more determined by baselines of $\sim 70\,\mathrm{m}$. This is because the estimation of derivatives of the phase (needed for phase recovery) degrades as the number of baselines is reduced. Baselines of $70\,\mathrm{m}$ resolve angular scales of $\sim 1.2\,\mathrm{mas}$ at $400\,\mathrm{nm}$.\ These order-of-magnitude considerations are taken into account when performing simulations and image reconstructions, i.e. the minimum and maximum size of pristine images that can be reconstructed by data analysis, do not go far beyond these limits. More precise array resolution limits are presented in section \[disks\] (diameters ranging between $0.06\,\mathrm{mas}\,-\,1.2\,\mathrm{mas}$) . By combining these angular scales with the SNR (eq. \[snr\]), we obtain Fig. \[MvsT\]. This figure shows the highest visual magnitude, for which photon correlations (with $|\gamma|=0.5$) can be detected (5 standard deviations), as a function of the temperature, and for several different exposure times. Also shown in Fig. \[MvsT\], is the shaded region corresponding to angular diameters between $0.03\,\mathrm{mas}$ and $0.6\,\mathrm{mas}$ [^2], and observable within $100\,\mathrm{hrs}$. From the Figure we can see how correlations of photons from faint stars can be more easily detected if they are hot. To quantify the number of stars for which photon correlations can be detected with the IACT array, we perform use the JMMC stellar diameters catalog [@catalog]. We find that $\sim 1000$ (out of $\sim33000$) stars from the JMMC catalog can be detected within $1\,\mathrm{hr}$, correlations from $\sim 2500$ stars can be detected within $10\,\mathrm{hrs}$, and $\sim 4300$ can be detected within $50\,\mathrm{hrs}$. In Fig. \[array\], we show a random sample of 2000 stars (out of $\sim33000$) from the JMMC catalog. Interstellar reddening may play a role in reducing the number of measurable targets Simulation of realistic data {#simulation} ============================ The original “pristine” image consists of $2048\times2048$ pixels corresponding to $\sim10\,\mathrm{mas}\times10\,\mathrm{mas}$ of angular extension and a wavelength of $\lambda= 400\,\mathrm{nm}$. The Fourier transform of the image is then performed via an FFT algorithm and normalized so that its value is one at zero baseline. This results in a Fourier transform sampled every $\mathrm{\sim(8\,\mathrm{m})/\lambda}$, i.e. $2\times 10^7$ cycles per radian field-of-view at a wavelength $\lambda$ of $400\,\mathrm{nm}$. We then find the squared-modulus of the degree of coherence between the members of each pair of telescopes. This is obtained from a linear interpolation of the squared Fourier magnitude in the finely sampled FFT. Diurnal motion is not taken into account in the simulations. Diurnal motion plays a significant role in increasing the $(u,v)$ coverage when exposure times are long. As a consequence there is less $(u,v)$ coverage in the simulations since projected baselines do not drift with time. The smaller $(u,v)$ coverage is however compensated by smaller statistical error in the correlation measurements.\ The final step in the simulation phase is the addition of noise to the correlation at each baseline. This noise was found to be Gaussian by performing the time integrated product of two random streams of simulated photons as detected by a pair of photo-multiplier tubes. The standard deviation of the noise added to each pair of telescopes is calculated from eq. \[snr\]. In this paper we take the signal bandwidth to be $\Delta f=200\,\mathrm{MHz}$. An example of simulated data as a function of telescope separation is shown in Fig. \[example\_simulation\]. This corresponds to a $3^{rd}$ magnitude uniform disk star ($T=6000^{\circ}\mathrm{K}$) of radius $0.1 \,\mathrm{mas}$ and $10\,\mathrm{hrs}$ of observation time. The software used for the simulations, as well as the analysis[^3], was developed in **C**.\ Fitting an Analytic function to the data {#fit} ======================================== The estimation of derivatives of the Fourier log-magnitude is at the heart of the Cauchy-Riemann phase recovery algorithm (section \[phase\_rec\_sec\]), and is thus greatly simplified when data is known on a square grid rather than in a ‘randomly’ sampled way as is directly available from observations. Once simulated data are available (or observations in the future), an analytic function is fitted to the data.\ An analytic function can be expressed as a linear combination of basis functions. When data $f(x_i)\equiv|\gamma(x_i)|^2$ are available at baselines $x_i$, with uncertainty $\delta f(x_i)$, the coefficients of the basis functions can be found by minimizing the following $\chi^2$: $$\chi^2=\sum_i\left[\frac{\left(f(x_i)-\sum_ka_kg_k(\alpha R(x_i))\right)}{\delta f(x_i)}\right]^2. \label{chi2}$$ Each $a_k$ is the coefficient of a basis function $g_k$. The constant $\alpha$ is a scaling factor, and $R$ is a rotation operator. The scaling factor and rotation angle are found by first performing a two-dimensional Gaussian fit. Finding the appropriate scale and rotation angle has the advantage of reducing the number of basis elements needed to fit the data.\ Basis functions which tend to zero at very large baselines, where data are scarce (see Fig. \[baselines\]), are ideal. For this reason, we use Hermite functions. There are situations where data are more easily fit with a different set of basis functions, e.g. a binary with unresolved members. In such a situation, data do not rapidly tend to zero at large baselines, so the Hermite function fit may contain a large number of elements and result in high frequency noise where data are scarce[^4]. The best choice of basis functions may therefore depend on the structure of the object. However, for consistency, we use the Hermite fit for all the objects that we analyze, and find that it gives reasonably good results.\ The $\chi^2$ minimization problem can be turned into a linear system by setting the set of partial derivatives $\frac{\partial \chi^2}{\partial a_k}$ to zero. We typically start with a small number of basis elements, say eight, and only increase the number of basis elements if the optimized reduced $\chi^2$ is greater than some predefined value. Cauchy-Riemann phase reconstruction {#phase_rec_sec} =================================== In order to perform a model-independent image reconstruction, the phase of the Fourier transform needs to be recovered from magnitude information only [@lipson]. Since the image is real, the Fourier magnitude is symmetric with respect to the origin in the $(u,v)$ plane. Lack of phase information results in the reconstructed image being arbitrary up to a translation and reflection. We discuss the one-dimensional phase reconstruction prior to developing the two-dimensional analysis. We can start by first approximating the continuous Fourier transform $I(x)$ by a discrete one, i.e. $I(m\Delta x)=\sum_j \mathcal{O}(j \Delta \theta) e^{ijmk_0 \Delta x \Delta \theta}$, where $\mathcal{O}(\theta)$ is the image in the sky and $k_0$ is the usual wave vector. Then it becomes convenient to express the discrete Fourier transform in phasor representation , i.e. $I(z)=R(z) e^{i\Phi(z)}$ where $z\equiv e^{imk_0 \Delta x \Delta \theta}$ is complex. Since the discrete Fourier transform is a polynomial in $z$, the theory of analytic functions can be applied. As a consequence, we obtain the Cauchy-Riemann equations in polar coordinates, which relate the log-magnitude and the phase along the purely real and imaginary axis. If data were available along the purely real or purely imaginary axis, we could directly solve for the phase by integrating along these directions. However, since $z$, the independent variable of the Fourier transform ($z\equiv e^{imk_0 \Delta x \Delta \theta}$), has modulus equal to 1, the phase differences that we seek lie along the unit circle in the complex plane. On the other hand, one can show that by using the Cauchy-Riemann equations, the phase differences along the radial direction in the complex plane[^5] are directly related to the differences in the logarithm of the magnitude which is available from the data (see appendix A for more details).\ The procedure to find the phase consists of first assuming a general solution form, then taking differences in the radial direction of the complex z-plane, and finally fitting the data to the radial differences of the assumed solution. A general form of the phase can be postulated by noting that the phase is a solution of the Laplace equation in the complex plane (applying the Laplacian operator on the phase and using the Cauchy-Riemann equations yields zero). Since the phase differences are known along the radial direction in the complex plane we can take radial differences of the general solution and then fit the log-magnitude differences (available from the data) to the radial differences of the general solution.\ We can think of this one-dimensional reconstruction as a phase estimation along a single slice in the Fourier plane. A generalization to two dimensions can be made by doing the same procedure for several slices as described in Fig. \[slices\]. In fact, the requirement that a two-dimensional complex function $(z_x, z_y)$ be analytic, is equivalent to satisfying the Cauchy-Riemann equations in both $z_x$ and $z_y$ [@2-d_complex]. The direction of the slices is arbitrary, however for simplicity we reconstruct the phase along an arbitrary set of perpendicular directions in the Fourier plane, and noting that one can relate all slices through a single orthogonal slice, i.e. once the phase at the origin is set to zero, the single orthogonal slice sets the initial values for the rest of the slices.\ One can also require that the phase at a particular point in the complex plane be exactly equal when reconstructed along $z_x$ or $z_y$ since each reconstruction is arbitrary up to a constant (piston) and a linear term (tip/tilt). However, imposing this requirement results in a severely over-determined linear system. More precisely, by imposing equality in $n^2$ points in the complex plane, and having $2n$ slices (each with an unknown constant and linear term), results in a linear system of $n^2$ equations and $4n$ unknowns. Alternative methods of requiring slice consistency are a possible way of improving phase reconstruction, but are beyond the scope of this paper.\ The Cauchy-Riemman approach, with horizontal or vertical slices, and a single orthogonal slice, gives reasonably good results, however, it is not the only possible approach. We have also investigated Gerchberg-Saxton phase retrieval, Generalized Expectation Maximization, and other variants of the Cauchy-Riemann approach. It is premature to conclude which of these approaches is best at this time, given the limited imagery and SNR levels that have been explored. However, the Cauchy-Riemman approach has shown to give better results in a number of cases [@Holmes.spie]. ![\[slices\] Schematic representation of two-dimensional phase reconstruction approach. Several parallel slices are related to a single orthogonal slice.](slices.eps) Imaging capabilities ==================== We investigated the imaging capabilities for simple objects[^6], namely uniform disk-like stars, oblate rotating stars, binaries, and more complex images. For most of the objects that we consider, image reconstruction is not necessary, i.e. from the Fourier magnitude alone, one can extract radii, oblateness, relative brightness in binaries, etc. Estimation of these parameters is probably more accurate when extracted directly from Fourier magnitude data only, especially if some a-priori knowledge of the original image is available. However, measuring simple parameters from reconstructed images is the first step in quantifying reconstruction capabilities with IACT arrays. We assume no a-priori knowledge of the images that are being reconstructed, and then do a statistical study of the uncertainties of the reconstructed parameters. Uniform disks \[disks\] ----------------------- In order to quantify the uncertainty in the reconstructed radius, we simulate data corresponding to $6^{th}$ magnitude stars ($T=6000^{\circ}\,\mathrm{K}$) with disk radii up to $1\,\mathrm{mas}$ for 50 hours of exposure time[^7]. An example of such a reconstruction is shown in Fig. \[radius\_a\], where the radiance is shown in arbitrary units between 0 and 1. For a uniform disk, the reconstructed phase is null in the first lobe, and we find that the rms deviations from the true phase are approximately $0.19\,\mathrm{rad}$ in the null zone. A first look at the reconstruction in Fig. \[radius\_a\] reveals that the edge of the reconstructed disk is not sharp, so a threshold in the radiance was applied for measuring the radius. The radius is measured by counting pixels above a threshold and noting that the area of the disk is proportional to the number of pixels passing the threshold. After experimenting with different radii, we chose the threshold for measuring the radius to be 0.2. We can now compare the simulated and reconstructed radii as is shown in Fig. \[radius\_b\], where each point in the Fig. corresponds to an individual simulation (including noise) and reconstruction. Further optimization in the threshold for measuring the radius should yield a slope even closer to unity in Fig. \[radius\_b\].\ Fig. \[radius\_b\] clearly shows that stellar radii ranging from $0.03\,\mathrm{mas}$ to $0.6\,\mathrm{mas}$ can be measured with uncertainties ranging between sub-percent and a few percent (Figures \[uncertainty\_a\] and \[uncertainty\_b\]). It can be seen from Fig. \[radius\_b\], that the uncertainty increases linearly as a function of the pristine (simulated) radius. This is due to a decrease in the number of baselines that measure a high degree of correlation when the angular diameter increases. As the pristine radius $\theta$ decreases, the distance to the first zero in the correlation increases as $\sim \theta^{-1}$, so the number of telescopes contained within the airy disk increases as $\sim \theta^{-2}$. Consequently, decreasing the pristine radius is equivalent to increasing the number of independent measurements by a factor of $\sim \theta^{-2}$. Since the uncertainty decreases as the square root of the number of independent measurements, the error decreases linearly with the radius. For radii above $0.6\,\mathrm{mas}$, there are simply not enough baselines to constrain the Fourier plane information for image reconstruction. For radii greater than $0.6\,\mathrm{mas}$, the distance to the first zero in the degree of correlation is of the order of $100\,\mathrm{m}$, but only baselines at $35\,\mathrm{m}$ and $50\,\mathrm{m}$ are capable of measuring the Fourier magnitude with more than 3 standard deviations (see eq. \[snr\]). In Fig. \[uncertainty\_b\] we show the relative percent error (RMS of a statistic) as a function of time for two different radii, where it can be seen that a relative error of a few percent is achieved after only a few hours. Oblate stars\[oblate\_sec\] --------------------------- For oblate stars we use the same magnitude and exposure parameters that are used for disk-like stars. Uniform oblate stars can be described by three parameters: the semi-major axis $a$, the semi-minor axis $b$, and the inclination angle $\theta$. Judging from the limitations obtained from reconstructing disks, we produce pristine images whose values for $a$ and $b$ are random numbers less than $1\,\mathrm{mas}$, and choose $1\leq a/b \leq2$ . The value of the inclination angle $\theta$ also varies randomly between $0^\circ$ and $90^\circ$. A typical image reconstruction can be seen in Fig. \[oblate\_example\_a\].\ After applying a threshold on pixel values as was done for disk shaped stars, the reconstructed parameters are found by calculating the inertia tensor of the reconstructed image. The eigenvalues and eigenvectors of the inertia tensor provide information for the reconstructed values of $a$, $b$ and $\theta$. To do this, we make use of the relation between the matrix eigenvalue and the semi-major/minor axes $I_{xx}=\frac{1}{4}a^2M$, where $M$ is the integrated brightness. A similar relation for $I_{yy}$ holds for the semi-minor axis $b$.\ The resulting reconstructed semi-major/minor axes as a function of their real values have a similar structure as the scatter plot for reconstructed radii shown in Fig. \[radius\_b\], and are well reconstructed up to $0.5\,\mathrm{mas}$ within a few percent. In Fig. \[oblate\_example\_b\], it can be seen how the uncertainty in the reconstructed oblateness $a/b$ increases with increasing oblateness. As with disk shaped stars (section \[disks\]), the uncertainty in the reconstructed semi-major/minor axes decreases as the square root of the number of baselines measuring a high degree of correlation. Therefore, the uncertainty in the reconstructed semi-major/minor axes is proportional to $\sim \sqrt{ab}$, and the error in the reconstructed oblateness is proportional to $\sim \sqrt{a/b+a^3/b^3}$.\ The reconstructed inclination angle as a function of the pristine angle is shown in Fig. \[theta\_oblate\], and several factors play a role in the uncertainty of the reconstructed value. For a fixed value of $a$ and $b$, the orientation of the telescope array with respect to the main lobe of the Fourier magnitude determines the number of baselines that measure a high degree of correlation. The number of baselines that measure a high degree of correlation is greater when the main lobe of the Fourier magnitude is aligned with the $x$ or $y$ direction of the array (see Fig. \[array\]), and is smaller by a factor of $\sim \sqrt{2}$ (assuming a uniform grid of telescopes) when its main axis is at $45^{\circ}$ with respect to the array. However, the uncertainty (proportional to spread of points) in Fig. \[theta\_oblate\] does not appear to be symmetric at $0^{\circ}$ and $90^{\circ}$, and is smaller at $90^{\circ}$. This due to the way the phase is reconstructed, i.e. due to the slicing of the Fourier plane along the $u$ or $v$ directions (see section \[phase\_rec\_sec\]). In the case of Fig. \[theta\_oblate\], the $(u,v)$ plane is sliced along the $u$ direction, with a single orthogonal reference slice along the $v$ direction. The main lobe of the Fourier magnitude of an oblate star has more slices passing through it when it is elongated along the $v$ direction (corresponding to an inclination angle of $90^{\circ}$ in image space), yielding a better reconstruction. This is in contrast to the orthogonal case of $0^{\circ}$, where the main lobe of the Fourier magnitude has a smaller number of slices passing through it. Binary stars\[binary\_sec\] --------------------------- Simulated data are generated for $5^{th}$ magnitude binary stars ($T=6000^{\circ}\,\mathrm{K}$), and an exposure of 50 hours after noting that the uncertainty in the degree of correlation is of the order of a few percent (eq. \[snr\]). Binaries stars are parameterized by the radii $r_1$ and $r_2$ of each star, their separation $d$, position angle $\theta$, and relative brightness in arbitrary units between 0 and 1. We generate pristine images with random parameters within the following ranges: radii are less than $0.3\,\mathrm{mas}$, angular separations are less than $1.5\,\mathrm{mas}$, the relative brightness per unit area is less than or equal to 1, and the orientation angle is less than $90^\circ$. A typical reconstruction can be seen in Fig. \[binary\_example\_a\].\ To measure the reconstructed parameters we identify the two brightest spots whose pixel values exceed a threshold of 0.2. We then find the radius for each bright spot and its centroid position. Identifying spots is a non trivial task in noisy reconstructed images and our analysis sometimes fails to identify the “correct” reconstructed spots. For example, a common issue is that close reconstructed spots that are faintly connected by artifacts, are sometimes interpreted as a single spot. It should be again stressed that image reconstruction may not be the best way to measure reconstructed parameters. For example, the data can just as well be fit by the general form of the Fourier magnitude of a resolved binary system (containing only a few parameters).\ In Fig. \[binary\_example\_b\], we show reconstructed angular separations as a function of their real values. The reconstructed values of the angular separation are found to within $\sim 5\%$ of their real values and cannot be much larger than what is allowed by the smallest baselines. We find that stars separated by more than $d_{max}\approx 0.75\,\mathrm{mas}$ are not well reconstructed since the variations in the Fourier magnitude start to become comparable to the shortest baseline.\ In Fig. \[relative\] we show the reconstructed values of the radii as a function of their pristine values. We find $\sim 10\%$ uncertainties in each of the reconstructed radii. Aside from the angular separation, a variable that plays a role in successfully reconstructing pristine radii is the ratio of absolute brightness[^8] of both binary members. When one of the two members is more than $\sim 3$ times brighter than the other, the fainter star is found to be smaller than its pristine value, and sometimes not found at all when one of the members is more than $\sim 10$ times brighter than the other. This is in part because the sinusoidal variations in the Fourier magnitude start to become comparable to the uncertainty. For example: a non-resolved binary star with one component 20 times brighter that the other, has relative variations of $\sim 10\%$. With all the redundant baselines, a few percent uncertainty in the measured degree of correlation is sufficient to accurately measure these variations. However, when the binary components are resolved, the relative variations decrease with increasing baseline and baseline redundancy is not sufficient to reduce the uncertainty in the measurement of the Fourier magnitude. This signal to noise issue can of course be improved by increasing exposure time.\ There are also issues related to algorithm performance. One such problem has to do with the fit of the data to an analytic function (see section \[fit\]). When the scale of the fit (found by an initial Gauss fit) is found to be too small, too many basis elements are used to reconstruct the data, and high frequency artifacts appear in reconstructions. Small initial scales are typically related to the binary separation as opposed to the size of individual components, and it is the latter which correctly sets the scale of the fit. Artifacts may be then mistaken for binary components, and incorrect reconstructed parameters may be found. Results improve significantly when either the correct scale is set or (model-independent) image post-processing is performed (see section \[mira\]). Featured images\[complex\] -------------------------- We now present two examples of more complex images, and show that the capabilities can, to a large extent, be understood from results of less complex images, such as uniform disks and binaries. In Fig. \[horizontal\_disk\] we show the reconstruction of a star with a dark band (obscuring disk), corresponding to a $4^{th}$ magnitude star and 10 hrs of observation time. The metric used to quantify the agreement with the pristine image (bottom left corner of Fig. \[horizontal\_disk\]) is a normalized correlation[^9] whose absolute value ranges between 0 (no correlation) and 1 (perfect correlation/anti-correlation). To quantify the uncertainty in the correlation, we perform the noisy simulation and reconstruction several times, and find the standard deviation of the degree of correlation. In the case of Fig. \[horizontal\_disk\], the correlation $c$ is $c=0.947\pm 0.001$. Note that the uncertainty in the correlation is apparently small, and the image reconstruction is not perfect, which implies that the reconstruction is not only affected by the SNR level, but also by the reconstruction algorithm performance.\ In order to determine the confidence with which we can detect the feature (dark region within the disk), this correlation is compared to the correlation of the reconstruction, and a featureless image, i.e. a uniform disk whose radius matches the radius of the pristine image. This comparison allows us to quantify the confidence with which we actually detect the feature (dark spot). Consequently, the correlation of the reconstructed image with a uniform disk is $c=0.880\pm 0.001$, which is lower by 61 standard deviations.\ Another example of a complex image reconstruction can be seen in Fig. \[dark\_spot\]. Here we show the reconstruction of a star with a dark spot, whose correlation with the pristine image is $c=0.940\pm0.001$. We can compare this correlation with the correlation of the reconstructed image and a uniform disk, which is $c=0.904\pm0.001$ (lower by 30 standard deviations).\ For both examples, we also calculate the correlation with the pristine image as a function of the angular scale (in $\mathrm{mas}$) of the pristine image. We then find the degree of correlation of each reconstruction and its pristine image, and also the degree of correlation of the reconstruction with a uniform disk. By comparing these two correlation values, we can find the smallest feature that can be reconstructed. Below some point we no longer distinguish between the reconstruction of the featured image and that of a uniform disk. We find that the smallest feature that can be reconstructed is close to $0.05\,\mathrm{mas}$. This can already be seen from the order of magnitude estimate made in section \[sensitivity\] and a comparable value of $0.03\,\mathrm{mas}$ is found in section \[disks\]. When pristine images have angular sizes greater than $0.8\, \mathrm{mas}$, the degree of correlation drops significantly due to lack of short baselines. Note that this value is higher than what is found in section \[disks\]. This is due to higher signal to noise ratio in the simulated data corresponding to Figures \[horizontal\_disk\] and \[dark\_spot\]. The resolution limits discussed in turn correspond to $\sim 16\times 16$ effective resolution elements (pixels).\ Post-processing\[mira\] ----------------------- Image post-processing is performed in order to improve the raw images presented above. The type of post-processing that is currently being investigated is analogous to the data analysis performed in conventional amplitude interferometry. In this analysis, the image is slightly modified so as to maximize the agreement between the data and the magnitude of the Fourier transform of the reconstructed image. See @Thibeaut for details. Also, one can introduce additional and very general constraints to this maximization procedure, but this is beyond the scope of this paper. Such an analysis depends strongly on the starting image, which can be provided by the analysis presented in this paper. An example of image post-processing performed on the oblate star of Fig. \[oblate\_example\_c\] is shown in Fig. \[improved\_oblate\].\ We have performed this type of post-processing on the images in Figures \[horizontal\_disk\] and \[dark\_spot\] using the MiRA software [@Thibeaut], and we obtain the ones shown in Figures \[mira\_image\_a\] and \[mira\_image\_b\]. Aside from optimizing agreement with the data, no additional constraints are imposed. These preliminary results show the overall reduction in noise and improvement in the sharpness of the reconstructed images. A systematic study of the improvements with image post-processing is currently being investigated [@Nunez]. Conclusions =========== We perform a simulation study of the imaging capabilities at $400\,\mathrm{nm}$ of an IACT array consisting of 97 telescopes separated up to a $\mathrm{km}$. This is a preliminary design for the CTA project, expected to be operational in 2018. Our method uses a model-independent algorithm to recover the phase from intensity interferometric data. We test the method on images of increasing degrees of complexity, parameterizing the pristine image, and comparing the reconstructed parameter with the pristine parameter. We now summarize our results and briefly discuss how fundamental stellar parameters can be constrained with the methods described above.\ We first find that for bright disk-like stars ($M_v=6$), radii are well reconstructed from $0.03\,\mathrm{mas}$ to $0.6\,\mathrm{mas}$. Even though using a Cauchy-Riemann based approach to recover images might not be the most efficient way to measure stellar radii, such a study starts to quantify the abilities of measuring other scale parameters in more complicated images (oblateness, distance between binary components, etc.). The range of angular radii that can be measured with a CTA-like array ($0.03\,-\,0.6\,\mathrm{mas}$) will complement existing measurements ($2\,-50\,\mathrm{mas}$) [@Haniff]. With the aid of photometry, the effective temperature[^10] scale of stars within $0.05\,-\,0.5\,\mathrm{mas}$ can be extended. Multi-wavelength angular diameter measurements will also reveal the wavelength dependence of limb-darkening [@Aufdenberg], and such measurements can be used to constrain energy transport models as is done in amplitude interferometry [@procyon].\ For oblate stars, similar results for extracting geometrical parameters are found. Due to the relative ease of SII to observe at short ($\sim 400\,\mathrm{nm}$) wavelengths, measuring fundamental parameters of hot B type stars is possible. B stars are particularly interesting since rapid rotation, oblateness, and mass loss are a common feature. We show that oblateness can be accurately measured, and a next step will be to quantify the capabilities of imaging realistic surface brightness distributions in hot stars. By imaging brightness (temperature) distributions, we will be able to study effects such as limb darkening and mass loss in hot massive stars [@ridgway], as well as gravity darkening [@von_zeipel].\ Binary stars are well reconstructed when one of the members is not much brighter (three times as bright) than the other, and when they are not too far apart ($\leq 0.75\,\mathrm{mas}$). As with amplitude interferometry, SII, along with spectroscopy, will allow the determination masses and orbital parameters in binary stars. If measured with enough precision ($\leq 2\%$) [@Andersen], the determination of the mass can be used to test main sequence stellar models. An advantage of using an array such as the one used in this study, is that individual radii can be resolved. An interesting phenomena to be studied with interacting binary stars is mass transfer (e.g. @zhao), and capabilities for imaging this phenomena can be further investigated.\ Two examples of reconstructions of more complex images are presented, and demonstrate the capability to resolve features in the sub-milliarcsecond scale. Results show improvement when post-processing is performed. An optimization procedure improves the image reconstructions found by the methods presented above by maximizing their agreement with the data. A more comprehensive study of reconstructions of complex images is currently being performed. Appendix A. Cauchy-Riemann phase recovery {#appendix-a.-cauchy-riemann-phase-recovery .unnumbered} ========================================= Recall that since SII can only allow to measure $|\gamma|^2$, the phase information is lost, and we would like to recover it for imaging purposes, with only the magnitude information. If we denote $I(z)=R(z) e^{i\Phi(z)}$, where $z\equiv\xi+i\psi$, we obtain the following relations from the Cauchy-Riemann equations[^11]: $$\begin{aligned} \frac{\partial\Phi}{\partial\psi}&=&\frac{\partial\ln{R}}{\partial\xi}\equiv \frac{\partial s}{\partial \xi}\\ \frac{\partial\Phi}{\partial\xi}&=&-\frac{\partial\ln{R}}{\partial\psi}\equiv -\frac{\partial s}{\partial \psi}, \label{cauchy}\end{aligned}$$ where we have defined $s$ as the log-magnitude. Notice the relation between the magnitude and the phase. By using the Cauchy-Riemann equations we can write the log-magnitude differences along the real and imaginary axes as: $$\begin{aligned} \Delta s_{\xi} &=&\frac{\partial s}{\partial \xi}\Delta\xi=\frac{\partial\Phi}{\partial\psi}\Delta\xi\\ \Delta s_{\psi}&=&\frac{\partial s}{\partial \psi}\Delta\psi=-\frac{\partial\Phi}{\partial\xi}\Delta\psi\\\end{aligned}$$ If the log-magnitude were available along purely the $\xi$ or the $\psi$ axes, we could solve the previous two equations for the phase.\ However, notice that because $|z|=1$, we can only measure the log-magnitude on the unit circle in the complex space ($\xi,\psi$).\ In general, we can write the log-magnitude differences along the unit circle as $$\begin{aligned} \Delta s_{||}&=&\frac{\partial s}{\partial \xi}\Delta\xi+\frac{\partial s}{\partial \psi}\Delta\psi \label{path0}\\ &=&\frac{\partial\Phi}{\partial\psi}\Delta\xi-\frac{\partial\Phi}{\partial\xi}\Delta\psi \label{path}\\ &=&\Delta \Phi_{\bot}. \nonumber\end{aligned}$$ Here $\Delta \Phi_{\bot}$ corresponds to phase differences along a direction perpendicular to $\Delta s_{||}$, that is, perpendicular to the unit circle in the $\xi-\psi$ plane. We are however interested in obtaining $\Delta\Phi_{||}$, so that we can integrate along the unit circle.\ The general form of $\Phi$ can be found by taking second derivatives in eq. (\[cauchy\]) and thus noting that $\Phi$ is a solution of the Laplace equation in the complex plane. $$\frac{\partial^2\Phi}{\partial\xi^2}+\frac{\partial^2\Phi}{\partial\psi^2}=0.$$ The general solution of $\Phi(z)$ in polar coordinates ($\rho, \phi$) is [@Jackson] $$\Phi(\rho,\phi)=a_0+b_0\phi+\sum_{j} \rho^j \left( a_j \cos{j\phi}+b_j\sin{j\phi}\right) \label{solution},$$ where terms singular at the origin ($\rho^{-j}$) have been omitted. These singular terms lead to ambiguous reconstructions including flipped images and have not been found to be essential for most reconstructions.\ Now taking the difference of $\Phi$ along the radial direction we obtain $$\Delta\Phi_{\bot}(\rho, \phi)=\sum_j \rho^j((1+\frac{\Delta\rho}{\rho})^j-1)\left( a_j \cos{j\phi}+b_j\sin{j\phi}\right).$$ Note from eq. (\[path\]) that the length in the complex plane associated with $\Delta s_{||}$ is $\Delta\phi=|\Delta\xi+i\Delta\psi|$, and that the length associated with $\Delta\Phi_{\bot}$ is $\Delta\rho=|\Delta\xi+i\Delta\psi|$. Now setting $\rho=1$, $\Delta\rho=\Delta\phi$, and for simplicity of presentation, expanding for small $\Delta\phi$, we obtain $$\Delta\Phi_{\bot}(\phi)=\sum_j j\Delta\phi\left( a_j \cos{j\phi}+b_j\sin{j\phi}\right).$$ So now the coefficients $a_j$ and $b_j$ can be found using equations (\[path0\]-\[path\]) from the measured $\Delta s_{||}$, and thus $\Phi$ can be found in the complex plane, with an uncertainty in $a_0$ and $b_0$. The coefficients $a_j$ and $b_j$ can be calculated by performing the following integrals: $$a_j=\frac{1}{2\pi\,j}\int_0^{2\pi} \frac{d\Phi_{\bot}}{d\phi} \cos{j\phi}\,d\phi \label{a}$$ $$b_j=\frac{1}{2\pi\,j}\int_0^{2 \pi} \frac{d\Phi_{\bot}}{d\phi} \sin{j\phi}\,d\phi \label{b}$$ Note however that the previous expressions must exist, which is not the general case. More explicitly, if the magnitude is zero, the log-magnitude is singular. When imaging finite objects in image space, there will always be zeros in the magnitude of the Fourier transform. In practice we are always sample limited and nothing prevents us from calculating $a_j$ and $b_j$ approximately. [100]{} Andersen J., 1991, AAPR, 3, 91 Aufdenberg J. P., Ludwig H. G., Kervella P., 2005, ApJ, 633, 424 Aufdenberg J. P. et al., 2006, ApJ, 645, 664 Aufdenberg J. P., Ridgway S., White R., 2009, astro2010: The Astronomy and Astrophysics Decadal Survey, 2010, 8 Baldwin J. E. et al., 1996, A&A, 306, L13 Bernl[ö]{}hr K., 2008, American Institute of Physics Conference Series, 1085, 874 Bessell M. S., 1979, PASP, 91, 589 CTA Consortium, 2010, arXiv:1008.3703 Dravins D. et al., 1994, Proc. SPIE 2198, 289 Fienup J. R., 1981, JOSA, 71, 1641 Friend D. B., Abbott D. C., 1986, ApJ, 311, 701 Hanbury Brown R., 1974, The Intensity Interferometer, Taylor & Francis, London. Haniff C. A., 2001, Galaxies and their Constituents at the Highest Angular Resolutions, 205, 288 Holder J., LeBohec S., 2006, ApJ 649, 399 Holmes R. B., Belen’kii M. S., 2004, JOSA 21, 697 Holmes R. B., Nuñez P. D., LeBohec S., 2010, Proc. SPIE 7818 Hörmander L., 1990, An introduction to complex analysis in several variables. Elsevier Science Jackson J. D., 1998, Classical Electrodynamics, 3rd Edition, by John David Jackson, Wiley-VCH. Kervella P., 2008, IAU Symposium, 252, 405 Labeyrie A., Lipson S. G., Nisenson P., 2006, An Introduction to Optical Stellar Interferometry, by A. Labeyrie, S. G. Lipson and P. Nisenson, Cambridge University Press, Cambridge Lafrasse S., Mella G., Bonneau D., Duvert G., Delfosse X., Chelli A., 2010, VizieR Online Data Catalog, 2300, 0 Lawrence T. W., Fitch J. P., Goodman D. M., Massie N. A., Sherwood R. J., 1990, Proc. SPIE, 1237, 522 LeBohec S. et al., 2010, Proc. SPIE, 7734, 7734 Maeder A., 1997, A&A, 321, 134 Maier G., preprint(arXiv:0907.5118) Martin E. L., Claret A., 1996, A&A, 306, 408 Monnier J. D., 2007, Bulletin of the American Astronomical Society, 38, 854 Monnier J. D. et al., 2007, Science, 317, 342 Mozurkewich D. et al., 2003, ApJ 126, 2502-2520 Nu[ñ]{}ez P. D., Lebohec S., Kieda D., Holmes R., Jensen H., Dravins, D., 2010, Proc. SPIE 7734 Nuñez P. D. et al., 2011, In prep. Ridgway S. et al., 2009, astro2010: The Astronomy and Astrophysics Decadal Survey, 2010, 247 Saio H. et al., 2006, ApJ, 650, 1111 Thiebaut E., 2009, New Astron. Rev. 53, 312 Verhoelst T., van Aarle E., Acke B., 2007, A&A, 470, L21 von Zeipel H., 1924, MNRAS, 84, 665 Zhao M. et al., 2008, ApJ, 684, L95 [^1]: E-mail: [email protected] [^2]: These curves of constant angular size can be found approximately by recalling that the visual magnitude $m_v$ is related to a calibrator star of visual magnitude $m_0$ by: $(m_v-m_0)=-2.5\log{F/F_0}$. Here, $F$ and $F_0$ refer to the flux in the visual band . To express $m_v-m_0$ as a function of the angular size, note that flux is proportional to $\theta^2T^4$, where $\theta$ is the angular size and $T$ is the temperature of the star. [^3]: See sections \[fit\] and \[phase\_rec\_sec\] for details on the analysis. Some variants of the analysis software were developed in **MATLAB**. All software is available upon request. [^4]: A basis set consisting of products of sines and cosines is more suitable in this situation [^5]: If $\xi$ is the real axis and $\psi$ is the imaginary axis, then a difference along the radial direction is $\Delta\xi+i\Delta\psi$ (assuming $\Delta\xi=a\xi$ and $\Delta\psi=a\psi$, where $a$ is a proportionality constant). [^6]: For preliminary study see @nunez.spie. [^7]: \[long\_exposures\] This brightness and exposure time correspond to uncertainties in the simulated data of a few percent. Such long exposure times can be reduced to a few hours as is shown in Fig. \[uncertainty\_b\] [^8]: Product of relative brightness per unit area (in arbitrary linear units between 0 and 1) and relative area of both stars [^9]: For two images $A_{i,j}$ and $B_{i,j}$, the normalized correlation $C_{i,j}$ is $C_{i,j}=Max_{k,l}\left\{N^{-2}(\sigma_A\sigma_B)^{-1} \sum_{i,j}^{N, N}A_{i,j}B_{i+k, j+l}\right\}$, where $\sigma_A$ and $\sigma_B$ are the standard deviations of images $A$ and $B$. [^10]: Defined as $T_{eff}=\left(\frac{L}{4\pi R^2\sigma}\right)^{1/4}$, where $R$ is the radius, $L$ is the luminosity, and $\sigma$ is the Stephan-Boltzmann constant. [^11]: The C-R equations can be applied because “I” is a polynomial in z.
{ "pile_set_name": "ArXiv" }
--- author: - 'Teresa Ruiz-Herrero' - Enrique Velasco - 'Michael F. Hagan' title: Mechanisms of budding of nanoscale particles through lipid bilayers --- Introduction ============ The mechanisms by which nanoscale particles cross cell membranes and the factors that control their uptake are essential questions for cellular physiology and modern biomedicine. Regulating the uptake (endocytosis) of nanoparticles is important for nanomedicine applications and for predicting nanoparticle toxicity [@Nel2009; @Mitragotri2009; @Poland2008]. Similarly, during the replication of many viruses an assembled nucleocapsid buds through the cell membrane, simultaneously exiting the cell and acquiring a membrane coating of host origin. Although endocytosis [@Liu2009], viral budding [@Welsch2007; @Gladnikoff2009], and scission of budded viruses [@Baumgartel2011] can be actively driven or assisted by cell machinery, both nanoparticle uptake and at least some aspects of budding of viruses or viral proteins [@Helenius1977; @Bonsdorff1978] can occur passively (without cell machinery or ATP hydrolysis) [@Bihan2009; @Hurley2010; @Solon2005; @Popova2010b]. Furthermore, evidence suggests that some viruses do or can undergo passive budding in vivo (e.g. [@Hurley2010; @Rossman2010]). It is therefore important to establish the aspects of particle budding which are generic to passive transport and thus underlie all forms of particle uptake or egress. In this paper we use elastic theory, moleculary dynamics (MD) simulations, and free energy calculations to characterize the dynamics and thermodynamics of the process by which a particle adheres to a membrane, is passively engulfed, and then spontaneously separates. Previous works first studied the equilibrium configurations of budding through a vesicle or infinite membrane as a function of membrane rigidity, particle size, and membrane-particle adhesion energy using elasticity theory [@Deserno2002; @Gao2005; @Zhang2008]. Subsequent studies used simulations to address further aspects of the problem, including Monte Carlo simulations on a randomly triangulated surface representation of a vesicle [@Fosnaric2009] and molecular dynamics (MD) on a coarse-grained lipid model [@Li2010] to investigate wrapping of charged particles, density functional theory to study the relationship between particle hydrophobicity and wrapping [@Ginzburg2007], and dissipative particle dynamics (DPD) to study wrapping of a particle by a inhomogeneous bilayer [@Smith2007] and wrapping behavior of ligand-coated nanoparticles [@Yue2011]. Recently, the wrapping behavior of ellipsoidal particles has been studied via DPD [@Yang2011b] and MD simulations [@Vacha2011]. While all of these treatments show that the adhesion energy required for wrapping depends on particle properties and membrane composition, there has not been a thorough comparison of predictions of elasticity theory with the results of more sophisticated computational models. In this work our primary objective is to understand the extent to which simplified elastic models can describe the thermodynamics and/or dynamics of particle uptake. To focus on aspects generic to all forms of exo- or endocytosis, we consider a minimal model in which the membrane is treated as a bilayer of homogeneous composition and the particle is pre-assembled (as in the case of nanoparticles or, e.g. type-D retroviruses [@Garoff1998; @Demirov2004]), and is spherically symmetric. Thus, in this work we do not consider the effects of membrane inhomogeneity (i.e. lipid rafts) or the association of viral membrane proteins [@Welsch2007; @Chan2010]. We compare the predictions of the elastic model [@Deserno2002] to results of dynamical simulations and free energy calculations. We find that the phase behavior predicted by the two descriptions agrees nearly quantitatively in most regions of parameter space, but there are important dynamical differences at parameter values near the transition between no uptake and particle budding. In particular, we identify a partially wrapped state which we show to be metastable. Methods {#methods .unnumbered} ======= ### The Membrane Model {#the-membrane-model .unnumbered} We model the amphiphilic lipids comprising the membrane with a coarse grained implicit solvent model from Cooke et al [@Cooke2005], in which each amphiphile is represented by one head bead and two tail beads that interact via repulsive WCA potentials [@Weeks1971], Eq.(1) $$V_\text{rep}(r)\!=\!\left\{\begin{array}{cr} \!4 \epsilon_{0} \! \left[ \left( \frac{b}{r}\right) ^{12}-\left( \frac{b}{r}\right) ^{6}\!+\!\frac{1}{4}\right]\! &\!;r\leq r_\text{c} \\ 0 &\! ;r>r_\text{c} \end{array} \right. \label{eq:wca}$$ with $r_\text{c}=2^{1/6}b$ and $b$ is chosen to ensure an effective cylindrical lipid shape: $b_\text{head-head}=b_\text{head-tail}=0.95\sigma$ and $b_\text{head-tail}=\sigma$, where $\sigma$ will turn out to be the typical distance between beads within a model lipid molecule. The beads belonging to a given lipid are connected through FENE bonds (Eq.(2)) [@Grest1986] and the linearity of the molecule is achieved via a harmonic spring with rest length $4\sigma$ between the first and the third bead, Eq.(3) $$V_\text{bond}(r)=-\frac{1}{2}\kappa_\text{bond} r_\infty^{2} \ln \left[ 1-\left( r/r_{\infty}\right) ^{2}\right] \label{eq:bond}$$ where $r_{\infty}=1.5\sigma$ $$V_\text{bend}(r)=\frac{1}{2}\kappa_\text{bend}\left( r-4\sigma \right)^{2} \label{eq:bend}$$ Since this is an implicit solvent model, the hydrophobicity is represented by an attractive interaction, Eq.(4), between all tail beads. $$V_\text{attr}(r)\!=\!\left\{ \!\begin{array}{ccc}\! \!-\epsilon_{0} &;& r<r_\text{c} \\ \!-\!\epsilon_{0}\! \cos^{2} \!\frac{\pi (r-r_\text{c})}{2\omega_\text{c}} \! &\! ;\!&\! r_\text{c}\! \leq \! r \! \leq \! r_\text{c} \!+\! \omega_\text{c} \\ \!0 & ; & r>r_\text{c}+\omega_\text{c} \end{array} \right. \label{eq:attr}$$ This model allows the formation of bilayers with physical properties such as fluidity, area per molecule and bending rigidity that are easily tuned via $\omega_{c}$. Moreover, diffusivity within the membrane, density, and bending rigidity are in good agreement with values of these parameters measured for biological membranes[@Cooke2005] (Figure 1) ![The properties of the membrane are easily tuned by ${\omega_\text{c}}$. [**(a)**]{} Bending rigidity $\kappa$ (in log scale) and [**(b)**]{} areal density $\eta$ of lipids as functions of ${\omega_\text{c}}$. The values in [**(a)**]{} are from Ref. [@Cooke2005] and the values in [**(b)**]{} were calculated from our simulations.[]{data-label="fig:properties"}](./properties.pdf){width="6.0cm"} ### Membrane-particle interaction {#membrane-particle-interaction .unnumbered} As noted in the introduction, the systems we have in mind include synthetic nanoparticles or viral particles which bud through attractive interactions with lipid membranes. These interactions can arise in part from electrostatic interactions between charged lipid head groups and charges on the nanoparticle surface or capsid exterior (e.g. basic residues on the matrix protein in retroviruses [@Welsch2007]). A second source of interaction can be protein mediated, including binding of nanoparticle-functionalized ligands to membrane receptors or insertion of hydrophobic tails on capsid proteins into the membrane [@Garoff2004].Finally, transmembrane viral ‘spike’ proteins can drive or facilitate budding. Importantly, each of these forms of interactions is short ranged. Receptor-ligand and spike protein-virus interactions operate on length scales of Å to nm; similarly, at physiological conditions of 100 mM salt electrostatic interactions have a Debye screening length of 1 nm. Thus, to keep our analysis general, we consider a short range attractive interaction between our model particle and head groups. In particular, we represent the combination of excluded volume and attractive interactions between the particle and head groups with a shifted Lennard Jones potential: $$\begin{aligned} V_\text{particle-head}\!=\!\left\{\!\begin{array}{cr} \! 4 \!\epsilon \!\left[ \!\left( \frac{\sigma}{r\!-\!s}\!\right) ^{12}\!-\!\left( \frac{\sigma}{\!r-\!s}\!\right) ^{6}\!\right]\! - \!V_\text{cut}\! \!&\!;\!r\!<\! r_\text{ph}\!+\!s \\ 0 \!& \!;\! r\!\geq\! r_\text{ph}\!+\!s \end{array} \right. \label{eq:vph}\end{aligned}$$ with $\epsilon$ a free parameter that controls the membrane-particle interaction strength, $s=R-\sigma /2$, $R$ the particle radius, $V_\text{cut}=4 \epsilon \left[ \left( \frac{\sigma}{r_\text{ph}}\right) ^{12}-\left( \frac{\sigma}{r_\text{ph}}\right) ^{6}\right]$, $r_\text{ph}=3.5\sigma$, and $\epsilon-V_\text{cut}$ the depth of the attractive interaction between the particle and the membrane. The particle experiences only excluded volume interactions with the tail groups, which are modeled with a shifted WCA potential [@Weeks1971], Eq.(6) $$V_\text{particle-tail}\!=\!4 \epsilon_{0} \!\left[ \!\left( \!\frac{\sigma}{r\!-\!s}\right) ^{12}\!-\!\left( \frac{\sigma}{r\!-\!s}\!\right) ^{6}\!+\!\frac{1}{4}\right] \label{eq:vpt}.$$ [**Parameters.**]{} From the phase diagram in [@Cooke2005] we set the temperature of our simulations to $k_\text{B}T/\epsilon_{0} =1.1$, which allows for a broad range of $\omega_\text{c}$ (between $1.3\sigma$ and $1.7\sigma$) within which the membrane is in the fluid state. Furthermore, the bending rigidity was calculated as a function of $\omega_\text{c}$ over this range of values for $k_{B}T/\epsilon_{0} =1.1$ in the same work; results from that reference are shown in Figure 1a. Values of the bilayer density calculated in our simulations over a similar range of $\omega_\text{c}$ are shown in Figure 1b. The units of energy, length, and time in our simulations are respectively $\epsilon_{0}$, $\sigma$ and $\tau_{0}$. The remaining parameters can be assigned physical values by setting the system to room temperature, $T=300K$, and noting that the typical width of a lipid bilayer is around 5 nm, and the mass of a typical phospholipid is about 660 g/mol. The units of our system can then be assigned as follows: $\sigma=0.9$ nm, $m_{0}=220$ g/mol, $\epsilon_{0}=3.77 \times 10^{-21} \text{J}=227\text{g}\text{\AA{}}^{2}/\text{ps}^{2}\text{mol}$, and $\tau_{0}=\sigma\sqrt{m_{0}/\epsilon}=8.86$ ps. Simulations {#simulations .unnumbered} ----------- Molecular dynamics (MD) simulations of budding were performed at constant temperature and pressure using the velocity Verlet algorithm, with a Langevin thermostat [@Frenkel2002a] to maintain constant temperature and a modified Andersen barostat [@Kolb1999] to maintain constant membrane tension to represent wrapping by an infinite membrane. The time step was $\Delta t=0.01\tau_{0}$, the friction constant was $\gamma = \tau_{0} ^{-1}$, the box friction for the Andersen barostat was $\gamma_\text{box}=2\cdot 10^{-4}$ and the box mass $Q=10^{-5}$ in the system units. The reference pressure, $P_{0}$, is set to 0, to simulate a tensionless membrane. The tension equals the pressure because the normal component to the membrane, the $z$-axis in our case, is free to fluctuate and does not contribute to the pressure. The $x$ and $y$ components of velocities and positions are rescaled according to the changes in the volume. In order to simulate an infinite membrane, periodic boundary conditions were employed. For most simulations the membrane was comprised of $n=21,492$ beads. An initial bilayer configuration was relaxed by MD and then placed normal to the $z$-axis in a cubic box of side-length $L=63.5 \sigma$. The particle was introduced in the center of the box with its pole located about 5$\sigma$ below the membrane surface with zero initial velocity. Since the membrane was kept tensionless by the barostat, the size of the box decreased during simulations as the particle was wrapped. To ensure that there were no finite size effects, additional sets of simulations were performed, following the same protocol, for membranes with $n=48,600$ beads and initial box size of 100x100x60 $\sigma ^{3}$ and with $n=86,400$ beads and initial box size 130x130x60 $\sigma ^{3}$. Except where mentioned otherwise, results are shown for the system with $n=21,492$ beads. [**Free energy calculations.**]{} In addition to performing dynamical simulations of budding, we calculated the potential of mean force as a function of particle penetration using umbrella sampling [@Torrie1977]. Simulations were performed in which the system was biased toward particular values of the penetration $p$ by introducing a biasing function $U_\text{bias}(\{\mathbf{r}\})= \frac{1}{2}\kappa_\text{umb}(p(\{\mathbf{r}\})-p_{0})^{2}$. Here $p(\{\mathbf{r}\})$ is the penetration for a configuration $\{\mathbf{r}\}$ and is defined as the distance between the top of the particle and the center of mass of the membrane. A series of windows were performed at different values of $p_0$; for all windows $\kappa_\text{umb}=200\epsilon_0/\sigma^2$. The simulations were started for an unwrapped particle ($p=-6\sigma$), and initial coordinates for each subsequent window were obtained from simulations in the previous one. Statistics from each window were stitched together and re-weighted to obtain the unbiased free energy using the weighted histogram analysis method [@Kumar1992; @WHAM]. Elastic model {#elastic-model .unnumbered} ------------- To evaluate the results of the dynamical simulations and free energy calculations, we compare the simulation results to a simplified elastic model for invagination of a particle in a membrane. Our elastic model closely follows that of Ref. [@Deserno2002] but we consider an infinite tensionless membrane rather than a vesicle. The total energy of the particle-membrane system arises from the energy of adhesion between the particle and the membrane ($e_\text{ad}$) and the elastic energy of the membrane ($e_\text{m}$). Following the simulation model, we assume that adhesion is mediated by short-range interactions with energy per area $-{\epsilon^{*}}$ so that the total energy of adhesion is ${e_\text{ad}}=-{\epsilon^{*}}{a_\text{wrap}}$ with ${a_\text{wrap}}$ the area of the membrane in contact with the particle. Note that ${\epsilon^{*}}$ actually describes a free energy since it includes the effects of counterion dissociation and other entropic factors involved in particle associations, but following Ref.  [@Deserno2002] we refer to it and the elastic terms described next as energies to emphasize that we are neglecting the (small) contribution to the free energy associated with fluctuations around the lowest free energy membrane configuration. To calculate the elastic contributions to the energy, we consider the Helfrich Hamiltonian for an infinitesimally thin membrane [@Helfrich1973] $$e_{m}=\int da \left( \sigma_{s} + \frac{\kappa}{2}(2H-C_{0})^{2}+\kappa_\text{G}K \right) \label{eq:helfrich}$$ where $\sigma_{s}$ is the surface tension, $\kappa$ and $\kappa _{G}$ are the bending rigidity and the Gaussian curvature modulus respectively, H and K are the mean and Gaussian curvatures, and $C_{0}$ is the spontaneous curvature. Our model membrane is symmetric and tensionless, so $C_{0}$ and $\sigma_{s}$ are 0. We will use this elastic model to describe the budding process up until the point of scission at the neck, and thus the topology of the membrane remains constant. Assuming that the Gaussian curvature modulus $\kappa_\text{G}$ is invariant throughout the membrane, the last term in Eq.(7) is constant under the Gauss-Bonnet theorem [@DoCarmo1976]. In practice, the properties of the membrane and thus $\kappa_\text{G}$ could change in the vicinity of the adsorbed particle, but this effect contributes a factor proportional to the adhesion area ${a_\text{wrap}}$ and thus only renormalizes the adhesion free energy ${\epsilon^{*}}$. The elastic energy for a general configuration of the membrane is then given by $$e_\text{m}=\int \frac{\kappa}{2}\left( \frac{1}{r_{1}}+\frac{1}{r_{2}}\right) ^{2}da \label{eq:adhesion}$$ with $r_{1}$ and $r_{2}$ the principal radii of curvature. ![Cross-section of the 3D geometry used for the elastic model of a membrane wrapping a particle of radius $R$. The particle, depicted as a green sphere, sticks to a section of the membrane in red with area ${a_\text{wrap}}$. The surrounding membrane, with area $a_\text{rim}$, drawn in blue, decays toward the flat configuration. The shape of this surrounding membrane is taken to be a section of a torus for simplicity. $\delta$ and $\rho$ stand for the outer and inner radius of the section of the torus formed by the rim region, and $\alpha$ and $\phi$ represent the polar and azimuthal angle of spherical coordinates. For a given penetration, $p$, there is a wrapping degree, $\theta$, that minimizes the elastic energy.[]{data-label="fig:scheme"}](./cartoon.pdf){width="7.5cm"} We follow Deserno et al. [@Deserno2002] to assume that the following geometry closely corresponds to the lowest free energy configuration for a partially wrapped particle (Figure 2). There is an area $a_\text{wrap}$ of the membrane tightly adhered to the particle with radius of curvature approximately equal to the particle radius $R$, and a rim area, $a_\text{rim}$, between the point at which the membrane separates from the particle and where it recovers a flat configuration. Because the particle is a featureless sphere, we assume that the lowest free energy configuration is axisymmetric, to give the energy $$\begin{array}{lc} e={e_\text{ad}}+e_\text{wrap}+e_\text{rim}= \\ =\!-\! {a_\text{wrap}}\!\epsilon^{*} \! +\!\kappa \!\left[\!\frac{2{a_\text{wrap}}}{R^{2}}\!+\!\int \!\frac{da_\text{rim}}{2} \!\left( \!\frac{1}{r_{1}}\!+\!\frac{1}{r_{2}}\right)^{2} \right] \end{array} \label{eq:energy1}$$ where $a_\text{rim}$ is the area of the rim surrounding the particle, and $r_{1}$ and $r_{2}$ are the principal radii of curvature in the rim area. We now recast Eq.(9) in terms of two new variables, the latitudinal degree of wrapping $\theta$ and the penetration $p$, which is the distance the particle travels along the direction normal to the flat membrane, measured from the point at which the surface of the particle first touches the flat membrane. Note that the theoretical penetration and the one from simulations have different definitions, although qualitatively both describe the system in a similar way. In Figure 2, a schematic of the system is depicted as a 2-D cross-sectional cut, in which the red line represents the section of the membrane bound to the particle and the blue line represents the rim region. As shown in the schematic, we assume that the rim corresponds to a section of a torus (appearing as a circular arc in the 2-D cross-section). Although this is only one of the multiple shapes the rim can form, it was shown to closely correspond to solutions from a full variational calculation in Ref. [@Deserno2002] and allows us to write the geometric properties of the system as explicit functions of our parameters. In particular, the radius of the torus depends uniquely on the particle size $R$, the wrapping degree $\theta$ and the penetration $p$; the area element on a torus and the two principal radii of curvature are [@Deserno2002] $\mathrm{d}a_\text{rim}=\rho(\delta-\rho \sin\alpha)d\alpha d\phi$, $r_{1}={\rho}$ and $r_{2}=-\frac{\delta-\rho \sin\alpha}{\sin\alpha}$, where $\alpha$ and $\phi$ are the polar and azimuthal angle in spherical coordinates. With the new parametrization, the area of the membrane in contact with the particle for a wrapping degree $\theta$, turns out to be ${a_\text{wrap}}(\theta)=2\pi R^{2}(1-\cos(\theta))$. Therefore, the energy of the system can be written in the following way: $$\begin{array}{lc} e=(-\frac{A\epsilon ^{*}}{2}+4\pi \kappa)(1-\cos\theta) +\\ + \!\pi \kappa\int_{0}^{\theta} \! \rho|\delta\!-\!\rho \sin \alpha | \left(\! \frac{1}{\rho}\!-\!\frac{\sin\alpha}{\delta \!-\!\rho \sin\alpha} \right) ^{2} d\alpha \end{array} \label{eq:energy2}$$ where $A=4\pi R^{2}$ is the surface area of the particle. For a given bending rigidity, particle size and membrane-particle interaction, the energy only depends on the penetration $p$ and the wrapping degree $\theta$. Finally, for each value of $p$ we minimize the energy Eq.(10) with respect to $\theta$ to obtain the membrane configuration and corresponding energy as a function of penetration alone. The results of the minimization are described below in section **Phase Diagrams** Results {#results .unnumbered} ======= System behavior {#system-behavior .unnumbered} --------------- To understand the influence of membrane and particle properties on budding, we began by performing dynamical simulations for a range of particle-membrane interaction strengths, $\epsilon$, particle radius $R$, and ${\omega_\text{c}}$, which controls the areal density of lipids, the bending rigidity $\kappa$, and diffusion rates within the membrane, as described in section [**The membrane model**]{}. Different values of these parameters lead to dramatically different behaviors, as shown in phase diagrams presented below (Figure 12). First we note that the behaviors can be grouped into four classes, which we illustrate by describing trajectories observed for various values of $\epsilon$ and constant values of $\kappa=13.9 {k_\text{B}T}$ and $R=12\sigma$ (Figure 3). ![Particle penetration into the membrane, $p$, as a function of time for molecular dynamics trajectories with different values of the adhesion strength for a system with particle radius $R=12\sigma$ and membrane bending rigidity $\kappa=13.9{k_\text{B}T}$. For $\epsilon=0.5\epsilon_{0}$ (red line) no wrapping occurs. For $\epsilon=1.0\epsilon_{0}$ (green line) and $\epsilon=1.2\epsilon_{0}$ (orange line), budding becomes stalled at a partially wrapped state whose value increases with $\epsilon$. For $\epsilon=2.0\epsilon_{0}$ (blue line), 3.0 (pink line), and 4.0 (cyan line) the particle undergoes complete encapsulation. On the right, slices of the system for corresponding values of $p$ are shown. Images were generated using VMD [@Humphrey1996][]{data-label="fig:md"}](./R12mdf.pdf){width="7.5cm"} For weak adhesion strengths $\epsilon$, no wrapping occurs; the membrane continues to exhibit only the usual spectrum of thermal fluctuations (Figure 4) after the particle adheres to it, and the penetration oscillates around negative values (Figure 3, case for $\epsilon=0.5\epsilon_{0}$). ![**Adhesion without wrapping.** Slices of configurations extracted from MD simulations with $\epsilon=1\epsilon_{0}$ ,$R=10\sigma$, and $\kappa=13.9{k_\text{B}T}$. Times shown are (a) $t=5\cdot10^{3}\tau_{0}$, (b) $t=4\cdot10^{4}\tau_{0}$,(c) $t=6\cdot10^{4}\tau_{0}$, (d) $t=1\cdot10^{5}\tau_{0}$. []{data-label="fig:confnwrap"}](./confnwrap45.pdf){width="7.5cm"} For a narrow intermediate range of $\epsilon$, the particle adheres to the membrane, but wrapping ceases at a partially wrapped state (Figure 5), after which the degree of particle penetration into the membrane fluctuates around a steady value (Figure 3, case for $\epsilon=1.0\epsilon_{0}$). The average value of the penetration remained unchanged for as long as we simulated (up to $4\cdot 10^{4}\tau_{0}$). The final degree of penetration in this arrested state increases with $\epsilon$, until approximately the point at which the particle is half wrapped (Figure 3, case for $\epsilon=1.2\epsilon_{0}$). A further increase in $\epsilon$ results in the next class of trajectories (Figure 3, case for $\epsilon=2.0\epsilon_{0}$ and $\epsilon=3.0\epsilon_{0}$), in which the particle is completely encapsulated (Figure 6). In this case, wrapping proceeds steadily until the particle is completely surrounded by membrane except for a narrow neck region (Figure 6d). Wrapping is then completed when a thermal fluctuation causes the neck to break and have its sides fused (Figure 6e), after which the fully wrapped particle diffuses away from the membrane (Figure 6f). Since fusion is a stochastic event, the budding time can be variable and we have observed neck configurations lasting between 500 and 5000 $\tau_{0}$. The elastic theory predicts that the shape and length of the neck depend on the balance between the adhesion energy and the bending energy with strong adhesion favoring a short neck and large bending energies favoring a long neck. The simulation results are consistent with this prediction; example configurations are shown in Figure 7. The figure shows snapshots from simulations with particle radius $R=6\sigma$, bending rigidity $\kappa=13.9{k_\text{B}T}$ and different values of $\epsilon$. A small particle size was chosen for the figure because the relationship between neck configuration and adhesion energy is most easily visualized when high membrane curvature is required for wrapping. The fact that fusion is accessible within the course of a typical simulation is an interesting contrast between the model studied here and that studied by Smith et al. [@Smith2007], where fusion was observed only for inhomogeneous membranes. ![**Long-lived partial wrapping.** Slices of configurations extracted from MD simulations with $\epsilon=1.3\epsilon_{0}$, $R=10\sigma$, and $\kappa=13.9{k_\text{B}T}$. The particle remains partially wrapped for the length of the simulation ($t=4\cdot 10^{4}\tau_0$). (a) $t=5\cdot10^{3}\tau_{0}$, (b) $t=1.5\cdot10^{4}\tau_{0}$,(c) $t=2.5\cdot10^{4}\tau_{0}$, (d) $t=3\cdot10^{4}\tau_{0}$.[]{data-label="fig:confhwrap"}](./confhwrap57.pdf){width="7.5cm"} ![**Wrapping.** Slices of configurations at different times extracted from MD simulations with $\epsilon=3\epsilon_{0},R=10\sigma$ and $\kappa=13.9{k_\text{B}T}$. The membrane wraps the particle (a-d) until a neck or channel connecting the flat bilayer and the membrane surrounding the particle forms (e). Thermal fluctuations break this narrow neck, resulting in the encapsulated particle escaping from the membrane (f). Configurations are shown for times (a) $t=5\cdot10^{3}\tau_{0}$, (b) $t=1\cdot10^{4}\tau_{0}$,(c) $t=1.5\cdot10^{4}\tau_{0}$, (d) $t=1.95\cdot10^{4}\tau_{0}$, (e) $t=2\cdot10^{4}\tau_{0}$,(f) $t=2.5\cdot10^{4}\tau_{0}$.[]{data-label="fig:confwrap"}](./confwrap70.pdf){width="7.5cm"} For higher values of $\epsilon$ wrapping proceeds extremely rapidly (Figure 3, case for $\epsilon=4.0\epsilon_{0}$). as there is a strong driving force to increase the number of head-particle interactions (Figure 8). As the curvature of the membrane in the vicinity of the wrapping front increases, the membrane structure undergoes ruptures in that region (Figure 8d), and a pore forms in the membrane (Figure 8e). The fully encapsulated particle then diffuses away and the pore heals through thermal motions of the lipids. The formation of a pore during these budding trajectories resembles the process by which a hydrophobic nanoparticle passes through membranes in the simulations described in [@Li2008; @Yue2011], but the physical driving forces are different in this case and the pore arises for kinetic reasons. Namely, the collective wrapping process proceeds more slowly than ruptures form in the membrane due to the large driving force to increase particle-head group contacts. ![**The neck profile depends on the adhesion strength.** Membrane configurations are shown shortly before the completion of budding for $R=6\sigma$, $\kappa=13.9{k_\text{B}T}$ and different adhesion strengths. (a) A relatively small adhesion strength, $\epsilon=3\epsilon_{0}$, leads to a long neck. (b) For $\epsilon=4\epsilon_{0}$ the neck is shorter. (c) For $\epsilon=5\epsilon_{0}$, close to the adhesion strength that leads to membrane rupture, the neck length is comparable to the height of typical membrane fluctuations. The top row of images shows a side view of system configurations and the bottom row of images gives the corresponding side-view slices.[]{data-label="fig:neck"}](./neckR6b.pdf){width="8.5cm"} ![**Wrapping via membrane rupture.** Slices of configurations at different times extracted from MD simulations with $\epsilon=5\epsilon_{0}$ ,$R=10\sigma$, and $\kappa=13.9{k_\text{B}T}$. Particle wrapping (a, b) leads to the formation of a pore (c, d). Eventually, the enveloped particle leaves the membrane (e) and the pore closes (f). Configurations shown occured at times (a) $t=2\cdot10^{3}\tau_{0}$, (b) $t=5\cdot10^{3}\tau_{0}$,(c) $t=5.4\cdot10^{3}\tau_{0}$, (d) $t=5.6\cdot10^{3}\tau_{0}$,(e) $t=8\cdot10^{3}\tau_{0}$, (f) $t=9.5\cdot10^{3}\tau_{0}$.[]{data-label="fig:confbr"}](./confbr31.pdf){width="7.5cm"} [**Free energy calculations.**]{} We were particularly interested in the partially wrapped states seen in the dynamical simulations described in the previous section, as the elastic model predicts only fully wrapped or non-wrapped states. To determine whether or not these observations corresponded to equilibrium configurations, the free energy was calculated as a function of the penetration using umbrella sampling (section **System Model**). Calculated free energy projections are shown for three values of $\epsilon$ in Figure 9, for which the finite-time dynamical simulations respectively ended in no wrapping ($\epsilon=0.5\epsilon_0$, Figure 9a), partial wrapping ($\epsilon=\epsilon_0$, Figure 9b), and complete wrapping ($\epsilon=1.25\epsilon_0$, Figure 9c). For the cases of full wrapping and no wrapping, the calculated free energy projections are consistent with the dynamics results. Namely, for $\epsilon=0.5\epsilon_0$ the minimum free energy value corresponds to no wrapping with a steep penalty for increasing penetration, while for $\epsilon=1.25\epsilon_0$ the free energy decreases monotonically with increasing penetration until the particle is completely wrapped. ![Free energy profiles as a function of the penetration, *p*, calculated from MD simulations using umbrella sampling are shown for $R=12\sigma$, $\kappa=13.9{k_\text{B}T}$, and indicated values of $\epsilon$. At the bottom, slices of the system as a function of the penetration are shown.[]{data-label="fig:freenR12f"}](./freenR12f2.pdf){width="7.5cm"} In contrast, the minimum value in the free energy profile for $\epsilon=1.0\epsilon_0$ does not correspond to the partially wrapped state observed in the dynamical simulations, but rather corresponds to complete wrapping. Further comparison of umbrella sampling results to dynamical trajectories suggests that, while the penetration coordinate $p$ is capable of describing the free energy basins corresponding to the unwrapped and wrapped states, $p$ alone is not sufficient to completely describe the transition dynamics. I.e., $p$ is a suitable order parameter for determining the free energies of the stable states, but not a complete reaction coordinate [@Dellago2002]. To see this, we chose a set of configurations from the umbrella sampling trajectories with different values of $p$. For each such configuration we performed several unbiased MD trajectories initialized with velocities using different random number seeds to obtain a crude estimate of the commitment probability [@Dellago2002]. We found that trajectories initiated from configurations with small values of $p\lesssim5 \sigma$ fluctuate around that value and configurations with $p\gtrsim18.8\sigma$ progressed steadily to complete wrapping. However, configurations with $5\sigma\leq p \lesssim 18.8\sigma$ tended to fluctuate around the value of $p$ corresponding to their initial configuration, which is inconsistent with the free energy profile for $p\ge15\sigma$ and indicates the presence of an additional slow degree of freedom at moderate $\epsilon$. It is not necessary to identify a perfect reaction coordinate to fulfill our primary objective of understanding the phase behavior, but we did attempt to identify the second relevant dynamical degree of freedom. Analysis of umbrella sampling configurations during the equilibration phase of the calculation indicates that, when the particle is held at a fixed penetration, the membrane configuration gradually relaxes to a state of increased adhesion and bending (Figure 10). Thus we expect that a reaction coordinate capable of describing the transition dynamics needs to include an additional collective variable that describes adhesion and/or membrane bending. Investigating this possibility, however, is beyond the scope of the present work focused on the phase behavior. ![Evolution of the membrane area in contact with the particle (a) and bending energy (b) as a function of time for MD simulations with a biasing potential holding the penetration near $p=16.9 \sigma$, $U_{\text{bias}}=0.5 \kappa_\text{umb}(p-16.875)^2$, and $\kappa_\text{umb}=200\epsilon_{0}/\sigma^2$. The time courses are averaged over 20 independent trajectories[]{data-label="fig:bendadhevolution"}](./propevolution.pdf){width="6.0cm"} Phase diagrams {#ssec:phasediagrams .unnumbered} -------------- Based on the results of dynamical simulations over a wide range of parameters, as well as umbrella sampling at parameter sets near the transition between no wrapping and wrapping, we determined phase diagrams as functions of adhesion energy $\epsilon$, membrane bending modulus $\kappa$, and particle radius $R$ (Figure 12). To enable comparison with the elastic theory Eq.(9), it is essential to note the the theoretical parameter $\epsilon^{*}$ corresponds to the adhesion free energy rather than simply the depth of the head group-particle attractive potential well $\epsilon$. Therefore, we plot the data as a function of the adhesion free energy per area, calculated as $\epsilon^*/{k_\text{B}T}=-\eta\log \left[ 1+\int_{s+\sigma}^{\infty}dr \left( e^{-V_\text{particle-head}(r)}-1 \right) \right]$, Figure 11. Here we have neglected the tiny contribution from cutting off the potential, we assume that each lipid head group in contact with the particle approaches roughly along a radial coordinate, and we neglect configurational entropy losses endured by the lipid molecules during adhesion. The parameter sets for which non-wrapping is the equilibrium configuration are shown with \* symbols, while the parameter sets which lead to equilibrium wrapping are separated into those which involve long-lived partially wrapped structures (x symbols), complete wrapping (+ symbols), and those for which the membrane undergoes rupture prior to budding ($\Box$ symbols). The wrapping binodal predicted by the elastic theory is shown as a dashed line on each plot. We see that while the theory and simulations agree to within about $0.2{k_\text{B}T}$, the theoretical binodal is below the computational results. This discrepancy could occur because we have not accounted for the configurational entropy contributed by the lipids during adhesion or due to the fact that the theory assumes an infinitesimally thin membrane. ![Relationship between the adhesion energy $\epsilon$ and the adhesion free energy per area $\epsilon*$ for $\kappa=13.9 {k_\text{B}T}$.[]{data-label="fig:epsrel"}](./epsrel.pdf){width="6.0cm"} The complete phase diagram predicted by the elastic theory is shown in Figure 13. Here, the dashed line is the binodal, given by $\epsilon*=\frac{2\kappa}{R^{2}}$, below which wrapping is energetically unfavorable and the solid line denotes the spinodal, above which wrapping proceeds without any energetic barrier. In between the lines there is a barrier to wrapping which begins at $p=0$, meaning that there are no long-lived partially wrapped states consistent with this theory for infinite membranes. ![Phase diagrams obtained from MD simulations and free energy calculations (a) as a function of $\epsilon$ and $\kappa$ for constant $R=10\sigma$ and (b) as a function of the particle size $R$ and adhesion free energy $\epsilon*$ for constant bending rigidity, $\kappa=13.9 {k_\text{B}T}$. Parameter sets are identified as those which lead to no wrapping ($*$ symbols), long-lived partially wrapped structures (x symbols), complete wrapping (+ symbols), and those for which the membrane undergoes rupture prior to budding ($\Box$ symbols). The wrapping binodal predicted by the elastic theory is shown as a dashed line on each plot. The relationship between the adhesion free energy $\epsilon*$ and the head group-particle attractive well depth $\epsilon$ is given in the text and in Figure 11.[]{data-label="fig:phasediagsim"}](./phasediagsimbind.pdf){width="6.5cm"} ![Phases diagrams obtained from the energy minimization in Eq.(10) (a) as a function of $\epsilon^{*}$ and $\kappa$ for constant $R=10\sigma$ and (b) as a function of the particle size, R, and the adhesion free energy per area, $\epsilon^{*}$, for constant bending rigidity $\kappa=13.9 {k_\text{B}T}$. The binodal ($\epsilon*=\frac{2\kappa}{R^{2}}$) above which wrapping is energetically favorable is indicated by a dashed line and the spinodal, above which wrapping proceeds without an energy barrier, is shown by a solid line. []{data-label="fig:phasediagtheo"}](./phasediagtheo.pdf){width="6.5cm"} Discussion {#discussion .unnumbered} ========== Our dynamical simulations of a minimal molecular model for the process of passive endo- or exocytosis identified four classes of behaviors resulting from the interaction of a particle with a membrane, no wrapping or adhesion resulting in a minimal perturbation of membrane configurations, partial wrapping, complete wrapping, and wrapping via rupture of the membrane. Equilibrium calculations showed that there are only two equilibrium configurations, corresponding to no wrapping or complete wrapping, and this equilibrium phase behavior for the molecular model shows strong agreement with the predictions of a simplified elastic theory [@Deserno2002]. The primary difference between the elastic model and the finite-time dynamical simulation results is the appearance of long-lived partial wrapping states. Since the long-lived partially wrapped states seen in this study could be significant for dynamical, time-sensitive particle uptake processes such as endocytosis or viral budding in living organisms, it is worth comparing them to observations of other models. Most closely related to our results, Yue and Zhang [@Yue2011] study a model comparable to ours except that the particle is coated with ligands that experience attractive interactions with membrane lipids. It appears that some of the configurations which they denote as ‘adhesion’ correspond to our long-lived partially wrapped states; however, it would be necessary to perform free energy calculations to determine whether they are metastable as we find here. In contrast to their model, we do not identify any parameter sets for which the particle partially penetrates into the hydrophobic interior of the membrane. Partially wrapped states have also been predicted from equilibrium theories in the context of finite-sized systems. Deserno et al. [@Deserno2002] examined the budding of colloid from the interior of a spherical vesicle using the same Hamiltonian as introduced in elastic theory described above. They found that partial wrapping corresponds to the equilibrium state when the vesicle size is on the order of the colloid diameter due to the increase in curvature energy of the finite-sized vesicle. We similarly obtain partial wrapping configurations as equilibrium solutions if we introduce finite size into the elastic theory studied here (Eq.(9)) by minimizing the energy of the elastic theory (Eq.(10)) with the total membrane surface area constrained to $16R^2$, so that the wrapping area plus the rim area cannot exceed the total area. Zhang and Nguyen [@Zhang2008] also identify partially wrapped states as equilibrium solutions to an elastic theory, but they observed that the catenoid configuration is the only solution to the full variational problem for a tensionless infinite membrane. This solution implies that the elastic energy of the rim is always zero, and thus wrapping is only determined by the balance between the bending energy in the wrapped region and the adhesion energy, which does not lead to partial wrapping states. Because the toroid approximation for membrane configurations assumed in our elastic model is more restrictive than the full variational problem, the wrapping binodal shown in Figure 13 is shifted to slightly higher values of the adhesion energy $\epsilon$ than obtained for their theory [@Zhang2008], but the behavior is qualitatively unchanged. Importantly, neither theory predicts the partially wrapped state as a metastable configuration in an infinite membrane. The membrane size in our simulations was chosen large enough to ensure that the theoretically predicted finite-size effects would not affect our results. To confirm that this was the case, we ran additional dynamical simulations and umbrella sampling calculations with membranes which were $50 \%$ and $100\%$ larger (16200 and 28800 lipids respectively). The simulation results were the same for all three membrane sizes. Finally, we consider our minimal model for passive endo- or exocytosis in the context of physical systems. Based on the length scales discussed in section **Parameters**, the particle diameters in our simulations range from 9 to 36 nm. Nanoparticles are available in a wide range of sizes, with particles smaller than 50 nm undergoing the most efficient uptake [@Jiang2008], and our simulation results indicate a range of possible uptake pathways. Our simulated particles are somewhat smaller than the size of viral capsids that undergo budding, which range from about 40 nm (e.g. hepadnavirus [@Seeger2000]) to hundreds of nanometers, but the results can be extrapolated into that range. As shown in Figure 13 the adhesion-wrapping transition decreases with radius as $1/R$ for constant bending rigidity. Our simulation results indicate that the existence of attractive interactions between a particle and lipid head groups, which has been proposed as the minimal requirement for viral budding [@Garoff1998; @Garoff2004], is indeed sufficient to drive efficient wrapping. However, to avoid stalled partially-wrapping states and membrane rupture, the system would be confined to a relatively narrow range of adhesion strengths spanning about $2 {k_\text{B}T}/\sigma^2$ (the + symbols in Figure 12). While this result is qualitative, since the range increases in width with the particle size and the three-bead representation of the lipid molecule may lead to model membranes which are more susceptible to rupture than those comprised of a more realistic lipid, it does establish important constraints on viral evolution if budding is limited to these ingredients. However, depending on the viral system, a number of additional phenomena contribute to budding, including membrane-associated viral envelope or spike proteins, preferential budding from lipid rafts [@Ono2001], the use of cytoskeletal machinery to actively drive or assist budding [@Gladnikoff2009; @Taylor2011] or scission [@Demirov2004; @Baumgartel2011], and the ability of the virus to remodel cell membrane properties[@Chan2010]. These effects can broaden the range of functional adhesion energies; e.g. using actin to drive assembly and budding [@Gladnikoff2009] could enable efficient viral egress even for adhesion energies at which spontaneous dynamics become stalled. In this case the presence of a barrier to budding could serve as a regulatory feature. The agreement between our simulation results and the elastic theory over some ranges of parameter space indicates that some or all of these effects could in principle be captured by extending existing elastic theories along the lines of Liu et. al.’s description of active endocytosis [@Liu2009], but care would be required to include all relevant slow degrees of freedom near transitions between wrapping and no wrapping. @ifundefined [46]{} Nel, A. E.; Mädler, L.; Velegol, D.; Xia, T.; Hoek, E. M. V.; Somasundaran, P.; Klaessig, F.; Castranova, V.; Thompson, M. *Nat Mater* **2009**, *8*, 543–557 Mitragotri, S.; Lahann, J. *Nat Mater* **2009**, *8*, 15–23 Poland, C. A.; Duffin, R.; Kinloch, I.; Maynard, A.; Wallace, W. A. H.; Seaton, A.; Stone, V.; Brown, S.; Macnee, W.; Donaldson, K. *Nat Nanotechnol* **2008**, *3*, 423–428 Liu, J.; Sun, Y.; Drubin, D. G.; Oster, G. F. *PLoS Biol* **2009**, *7*, e1000204 Welsch, S.; Müller, B.; Kräusslich, H.-G. *FEBS Letters* **2007**, *581*, 2089 – 2097, &lt;ce:title&gt;Membrane Trafficking&lt;/ce:title&gt; Gladnikoff, M.; Shimoni, E.; Gov, N. S.; Rousso, I. *Biophysical Journal* **2009**, *97*, 2419–2428 Baumgartel, V.; Ivanchenko, S.; Dupont, A.; Sergeev, M.; Wiseman, P. W.; Krausslich, H.-G.; Brauchle, C.; Muller, B.; Lamb, D. C. *Nat Cell Biol* **2011**, *13*, 469–474 Helenius, A.; Fries, E.; Kartenbeck, J. *J Cell Biol* **1977**, *75*, 866–880 von Bonsdorff, C. H.; Harrison, S. C. *Journal of Virology* **1978**, *28*, 578–583 Bihan, O. L.; Bonnafous, P.; Marak, L.; Bickel, T.; Trépout, S.; Mornet, S.; Haas, F. D.; Talbot, H.; Taveau, J.-C.; Lambert, O. *J Struct Biol* **2009**, *168*, 419–425 Hurley, J. H.; Boura, E.; Carlson, L.-A.; Rózycki, B. *Cell* **2010**, *143*, 875 – 887 Solon, J.; Gareil, O.; Bassereau, P.; Gaudin, Y. *J Gen Virol* **2005**, *86*, 3357–3363 Popova, E.; Popov, S.; Göttlinger, H. G. *J Virol* **2010**, *84*, 6590–6597 Rossman, J. S.; Jing, X.; Leser, G. P.; Lamb, R. A. *Cell* **2010**, *142*, 902–913 Deserno, M.; Gelbart, W. M. *The Journal of Physical Chemistry B* **2002**, *106*, 5543–5552 Gao, H. J.; Shi, W. D.; Freund, L. B. *Proceedings of the National Academy of Sciences of the United States of America* **2005**, *102*, 9469–9474 Zhang, R.; Nguyen, T. T. *Phys. Rev. E* **2008**, *78*, 051903 Fošnarič, M.; Iglič, A.; Kroll, D. M.; May, S. *The Journal of Chemical Physics* **2009**, *131*, 105103 Li, Y.; Gu, N. *Journal of Physical Chemistry B* **2010**, *114*, 2749–2754 Ginzburg, V. V.; Balijepalli, S. *Nano Letters* **2007**, *7*, 3716–3722 Smith, K. A.; Jasnow, D.; Balazs, A. C. *The Journal of Chemical Physics* **2007**, *127*, 084703 Yue, T. T.; Zhang, X. R. *Soft Matter* **2011**, *7*, 9104–9112 Yang, K.; Ma, Y. Q. *Australian Journal of Chemistry* **2011**, *64*, 894–899 Vacha, R.; Martinez-Veracoechea, F. J.; Frenkel, D. *Nano Letters* **2011**, *11*, 5391–5395 Garoff, H.; Hewson, R.; Opstelten, D. J. *Microbiol Mol Biol Rev* **1998**, *62*, 1171–1190 Demirov, D. G.; Freed, E. O. *Virus Res* **2004**, *106*, 87–102 Chan, R. B.; Tanner, L.; Wenk, M. R. *Chem Phys Lipids* **2010**, *163*, 449–459 Cooke, I. R.; Kremer, K.; Deserno, M. *Phys. Rev. E* **2005**, *72*, 011506 Weeks, J. D.; Chandler, D.; Andersen, H. C. *J. Chem. Phys.* **1971**, *54*, 5237+ Grest, G. S.; Kremer, K. *Phys. Rev. A* **1986**, *33*, 3628–3631 Garoff, H.; Sjöberg, M.; Cheng, R. H. *Virus Res* **2004**, *106*, 103–116 Frenkel, D.; Smit, B. *Understanding molecular simulation: from algorithms to applications*, 2nd ed.; Academic: San Diego, Calif. ; London, 2002 Kolb, A.; Dünweg, B. *The Journal of Chemical Physics* **1999**, *111*, 4453–4459 Torrie, G. M.; Valleau, J. P. *J Comput Phys* **1977**, *23*, 187–199 Kumar, S.; Bouzida, D.; Swendsen, R. H.; Kollman, P. A.; Rosenberg, J. M. *J. Comput. Chem.* **1992**, *13*, 1011–1021 Grossfield, A. WHAM: the weighted histogram analysis method, version 2.0.6. <http://membrane.urmc.rochester.edu/content/wham> Helfrich, W. *Zeitschrift Fur Naturforschung C-a Journal of Biosciences* **1973**, *C 28*, 693–703 do Carmo, M. P. *Differential Geometry of Curves and Surfaces*; Prentice Hall, 1976 Li, Y.; Chen, X.; Gu, N. *The Journal of Physical Chemistry B* **2008**, *112*, 16647–16653 Humphrey, W.; Dalke, A.; Schulten, K. *J. Mol. Graph.* **1996**, *14*, 33–38 Dellago, C.; Bolhuis, P. G.; Geissler, P. L. *Advances in Chemical Physics, Vol 123*; Advances in Chemical Physics; 2002; Vol. 123; pp 1–78 Jiang, W.; Kim, B. Y. S.; Rutka, J. T.; Chan, W. C. W. *Nat Nanotechnol* **2008**, *3*, 145–150 Seeger, C.; Mason, W. S. *Microbiol Mol Biol Rev* **2000**, *64*, 51–68 Ono, A.; Freed, E. O. *Proc Natl Acad Sci U S A* **2001**, *98*, 13925–13930 Taylor, M. P.; Koyuncu, O. O.; Enquist, L. W. *Nat Rev Microbiol* **2011**, *9*, 427–439
{ "pile_set_name": "ArXiv" }
--- abstract: 'Competing density waves play an important role in the mystery of high-temperature cuprates superconductors. In spite of the large amount of experimental evidence, the fundamental question of whether these modulations represent charge or pairing density waves (CDWs or PDWs) is still debated. Here we present a method to answer this question using both momentum and energy-resolved resonant X-ray scattering maps. Starting from a minimal model of density waves in superconductors, we identify distinctive signatures of incipient CDWs and PDWs. The generality of our approach is confirmed by a self-consistent solution of an extended Hubbard model with attractive interaction. By considering the available experimental data, we claim that the spatial modulations in cuprates have a predominant PDW character. Our work paves the way for using X-ray to identify competing and intertwined orders in superconducting materials.' author: - David Dentelski - 'Emanuele G. Dalla Torre' title: 'Minimal model of charge and pairing density waves in X-ray scattering experiments ' --- [*Introduction –*]{} Strongly correlated materials often exhibit competing phases with distinct charge and spin orders. A famous example is copper-oxide high-temperature superconductors, or cuprates, whose rich phase diagram poses many theoretical challenges. Since the discovery of unidirectional spin density waves in LSCO [@tranquada1995evidence], it has become increasingly accepted that in cuprates superconductivity is intertwined with other orders [@doiron2007quantum; @wu2011magnetic; @fradkin2012high; @leboeuf2013thermodynamic; @wu2013emergence]. In particular, in 2002, scanning tunneling experiments found incommensurate density waves on the surface of BSCCO [@hoffman2002four; @howald2003periodic; @vershinin2004local; @hanaguri2004checkerboard]. Ten years later, resonant X-ray scattering experiments detected a similar incommensurate order in the bulk of YBCO [@ghiringhelli2012long]. The same order was later found in a large number of cuprates, demonstrating that this effect is ubiquitous [@chang2012direct; @torchinsky2013fluctuating; @blackburn2013x; @comin2014charge; @da2014ubiquitous; @le2014inelastic; @hashimoto2014direct; @tabis2014charge; @huecker2014competing; @achkar2014impact; @gerber2015three; @hamidian2015magnetic; @peng2016direct; @chaix2017dispersive; @peng2018re; @jang2017superconductivity; @da2018coupling; @bluschke2019adiabatic; @kang2019evolution]. In spite of the large number of experimental studies, the physical interpretation of these periodic modulations is still debated. A common approach, also based on earlier theoretical predictions [@castellani1995singular; @castellani1997charge], claims that these modulations are due to a charge density wave (CDW) order that competes with superconductivity. While this approach is widely accepted in the literature, it is inconsistent with some experimental details. In particular, angle-resolved photoemission spectroscopy (ARPES) shows that these density waves are associated with a spectral gap that closes from below the Fermi energy [@he2011single], while CDWs’ gaps are expected to close from above. Accordingly, it was argued that the competing order is intimately related to superconductivity [@loret2019intimate] and thus interpreted as a pair density wave (PDW) [@chen2004pair; @lee2014amperean], or a CDW/PDW mixed order [@pepin2014pseudogap; @freire2015renormalization; @wang2015interplay; @wang2015coexistence]. This claim is also supported by recent scanning measurements in the halos of magnetic vortices [@edkins2019magnetic] and with superconducting tips [@hamidian2016detection; @du2020imaging] (see Ref. [@agterberg2019physics] for a recent review). Here we address the question of how to distinguish between CDW and PDW modulations in available X-ray scattering experiments. Our approach departs from earlier studies that focused on the normal state of cuprates [@dalla2016friedel; @arpaia2019dynamical] and included the effects of strong antiferromagnetic fluctuations [@sachdev2013bond] and Fermi arcs with hot spots [@efetov2013pseudogap; @allais2014density; @pepin2014pseudogap; @wang2015interplay; @freire2015renormalization; @wang2015interplay; @wang2015coexistence]. Instead, we base our analysis on the well-established description of the superconducting state of cuprates in terms of a Bardeen-Cooper-Schrieffer (BCS) Hamiltonian with a $d$-wave gap. In this phase, superconductivity suppresses competing orders, justifying a weak-coupling approach where the density waves are induced by weak pinning centers [@caplan2015long; @caplan2017dimensional]. By considering a minimal model of isotropic scatterers, we develop a method to distinguish between incipient CDW and PDW fluctuations, which become long-ranged at high magnetic fields. The validity of this approach is confirmed by the solution of an extended Hubbard model with attractive interactions in the presence of local impurities, which enables us to study the interplay between CDWs and PDWs. \[t\] ------------------------------------------------------- -------------------------------------------------- ------------------------------------------------- \(a) NCCO \(b) Hg1201 \(c) BSCCO ![image](ElectrondopedFitNew1.pdf){width="31.00000%"} ![image](HgNatureFitNew1.pdf){width="31.00000%"} ![image](CominBiFitNew1.pdf){width="31.00000%"} ------------------------------------------------------- -------------------------------------------------- ------------------------------------------------- . \[Fig1\] [*Weak coupling approach –*]{} Resonant X-ray scattering experiments probe density fluctuations at a fixed wavevector $\textbf{q}$ and frequency $\Omega$. In our weak-coupling approach we assume that incipient CDWs and PDWs can be modeled by a homogeneous state perturbed by a local pinning center (impurity). Under this approximation, the intensity of the X-ray signal is given by density response to the impurity, $\chi(\textbf{q},\Omega)$. In a BCS superconductor, one has (see, for example, Ref.[@altland2010condensed]) $$\begin{aligned} \label{eq:rhoq} \chi(\textbf{q},\Omega) = \int d\omega \int d^{d}k~\rm Tr \left[G_{0}(\textbf{k},\omega) V_{\bf k} G_{0}(\textbf{k}+\textbf{q}, \omega+\Omega)\sigma^z\right]. \end{aligned}$$ Here $\rm V$ models a static (time independent) and local ($q$ independent) impurity, $\sigma^z$ is a Pauli matrix, and $\rm G_0$ is the Green’s function $$\begin{aligned} \rm G^{-1}_{0}({\bf k},\omega)= \begin{pmatrix} - \omega +\varepsilon_{\textbf{k}}-\mu& \Delta_{\textbf{k}} \\ \Delta^{\star}_{\textbf{k}} &-\omega-\varepsilon_{\textbf{k}}+\mu\end{pmatrix}, \end{aligned}$$ where $\Delta_{\textbf{k}}=\dfrac{\Delta_0}{2} (\cos(k_{x})-\cos(k_{y}))$ is the pairing gap, $\varepsilon_{\textbf{k}}$ is the band structure of the material, and $\mu$ the chemical potential [@Note1]. We now introduce a minimal model for CDW modulations, by considering the scattering from a momentum-independent charge impurity, $\rm V_\textbf{k} = V_0\sigma^z$. In this case, the integral over $\omega$ in Eq. (\[eq:rhoq\]) delivers $$\begin{aligned} \chi(\textbf{q},\Omega) =&~ 2\pi V_{0}\int d^2 k~ \dfrac{E_{\textbf{k}}^2+\varepsilon_{\textbf{k}}\varepsilon_{\textbf{k}+\textbf{q}}-\Delta_{\textbf{k}}\Delta_{\textbf{k}+\textbf{q}}-E_\textbf{k}\Omega}% {E_{\textbf{k}}[E_{\textbf{k}+\textbf{q}}^2-(E_{\textbf{k}}-\Omega)^2]} \nonumber\\% & +\dfrac{E_{\textbf{k}+\textbf{q}}^2+\varepsilon_{\textbf{k}}\varepsilon_{\textbf{k}+\textbf{q}}-\Delta_{\textbf{k}}\Delta_{\textbf{k}+\textbf{q}}+E_{\textbf{k}+\textbf{q}}\Omega}{E_{\textbf{k}+\textbf{q}}[E_{\textbf{k}}^2-(E_{\textbf{k}+\textbf{q}}+\Omega)^2]},\label{eq:rhoq2} \end{aligned}$$ where $E_{\textbf{k}} = \sqrt{\varepsilon_{\textbf{k}}^2+\Delta_{\textbf{k}}^2}$. In the limit of $\Delta_\textbf{k}\to0$, one has $E_\textbf{k}=|\varepsilon_\textbf{k}|$ and Eq. (\[eq:rhoq2\]) recovers the Lindhard response function of free fermions used in Ref. [@dalla2016friedel]. To describe X-ray scattering experiments of cuprates, we use a tight-binding model, $\varepsilon_{\textbf{k}}=-2t[\cos(k_{x})+\cos(k_{y})]-4t'\cos(k_{x})\cos(k_{y})$, where $t$ and $t'$ are nearest neighbor (NN) and next-nearest neighbor (NNN) hopping coefficients. The parameter $t'$ strongly affects the shape of the Fermi surface: superconducting cuprates are close to half-filling and, for $t'=0$, their Fermi surface has a diamond shape. A negative $t'$ leads to a Fermi surface with parallel segments (nesting) at the antinodal wavevectors ${\bf k} = (\pm\pi/a,0)$ and $(0,\pm\pi/a)$. As pointed out long ago [@massidda1989electronic], these parallel segments are prone to induce finite wavevector instabilities, such as CDWs and PDWs. This approach matches the experimentally observed doping dependence of the wavevector (see the Supplemental Materials)[@Note2]. Let us first consider the elastic component ($\Omega=0$), by comparing Eq. (\[eq:rhoq2\]) with resonant X-ray scattering experiments of three different cuprates: electron-doped NCCO [@jang2017superconductivity] , underdoped BSCCO [@comin2014charge] and underdoped Hg1201 [@tabis2014charge]. The corresponding plots are shown as blue curves in Fig. \[Fig1\], where we select $\mu$ to match the experimental doping $x$. The superconducting gap $\Delta_0$ has a minor influence on these plots and is set to physically relevant values. The fitting parameter $t'$ is obtained by minimizing the difference between the theoretical curves and the actual experiments [@Note3]. The values of $t'/t$ obtained by this procedure are consistent with the Fermi surfaces determined by ARPES [@norman1995phenomenological; @markiewicz2005one]. For all three materials, we obtain an excellent agreement between the theoretical curves and the experiments: Eq. (\[eq:rhoq2\]) describes well both the period of the modulation and the width of the peak. As mentioned above, an alternative explanation for the observed signal are PDW fluctuations. Specifically, we consider short-ranged PDWs that coexist with a static and uniform (d-wave) pairing gap $\Delta_0$. Our analysis does not apply to materials where the PDWs are long ranged and give rise to a state, analogous to the FFLO state [@fulde1964superconductivity; @larkin1965nonuniform], where the pairing gap is periodically modulated in space (such as the striped superconductor LBCO near 1/8 doping [@berg2009theory; @berg2009striped]). Following our weak-coupling approach, we consider PDW fluctuations induced by a local modulation of the pairing gap, $\rm {V_\textbf{k}} = \Delta_\textbf{k}\sigma^{x}$ in Eq. (\[eq:rhoq\]), where $\rm \sigma^{x}$ is a Pauli matrix. By performing the integral over $\omega$ in Eq. (\[eq:rhoq\]), we obtain $$\begin{aligned} \label{eq:rhoq3} \chi(\textbf{q},\Omega) =2\pi \int d^2 k~ \Delta_{\bf k}\frac{\varepsilon_{\bf k}\Delta_{\bf k+q}+\varepsilon_{\bf k+q}\Delta_{\bf k}}{(E_{\bf k}-E_{\bf k+q})^2-\Omega^2}\left(\dfrac1{E_\textbf{k}}-\dfrac1{E_{\textbf{k}+\textbf{q}}}\right). \end{aligned}$$ The resulting plots are shown as red curves in Fig. \[Fig1\]. We find that the PDW signal shows pronounced peaks at approximately the same wavevector as the CDW one. The precise shape of the peaks depends on the details of the band structure and cannot be used to identify the type of modulation. As a result, one-dimensional scans of the X-ray scattering are not sufficient to distinguish unequivocally between CDW and PDW fluctuations. \[t\] ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------ ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------ \(a) CDW \(b) PDW ![Two dimensional maps of the elastic response $\chi(\textbf{q},\Omega=0)$ in the weak-coupling approach: (a) CDW, Eq. (\[eq:rhoq2\]), (b) PDW, Eq. (\[eq:rhoq3\]). Numerical parameters: $ x=0.14, ~t'/t=-0.6,~ \Delta_{0}/t = 0.2$. []{data-label="fig:CPDW"}](DpairSCDW3.pdf "fig:"){width="25.00000%"} ![Two dimensional maps of the elastic response $\chi(\textbf{q},\Omega=0)$ in the weak-coupling approach: (a) CDW, Eq. (\[eq:rhoq2\]), (b) PDW, Eq. (\[eq:rhoq3\]). Numerical parameters: $ x=0.14, ~t'/t=-0.6,~ \Delta_{0}/t = 0.2$. []{data-label="fig:CPDW"}](DpairDPDW3.pdf "fig:"){width="25.00000%"} ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------ ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------ [*Identifying CDW and PDW –*]{} We now present two distinct methods to distinguish between these two types of modulations, based respectively on the momentum and energy dependence of the X-ray scattering signal. The first method uses the full two dimensional map of $\chi(\textbf{q},\Omega=0)$. Two representative theoretical maps are shown in Fig. \[fig:CPDW\]. Although both maps have pronounced peaks at the same wave-vector ($q \approx \pm0.25$), their two-dimensional structure is very different: The CDW signal has four peaks at $\textbf{q} = (\pm q,\pm q)$ and four saddle points at $\textbf{q} = (0,\pm q)$ and $(\pm q,0)$. In contrast, the PDW signal has four strong peaks at $\textbf{q} = (0,\pm q)$ and $(\pm q,0)$ and four weaker peaks at $\textbf{q} = (\pm q,\pm q)$. We claim that the ratio between the intensity of the signal at these two wavevectors can be used to identify the type of modulation. For the parameters used in Fig. \[fig:CPDW\], we find that $R\equiv \chi(\textbf{q}=(q, 0), \Omega=0)/\chi(\textbf{q}=(q, q), \Omega=0) \approx 0.7$ for CDW and $R\approx 1.4$ for PDW. This result is very robust: Although the precise value of $R$ depends on the microscopic parameters of the model, we find generically that $R<1$ for CDWs and $R>1$ for PDWs. To understand this result, we recall that in our weak-coupling model of CDW and PDW, the scatterers are local and isotropic and, hence, their scattering matrices are momentum independent. In this minimal model, the intensity of the response function only depends on the density of states, i.e. on the shape of the Fermi surface and on the symmetry of the pairing gap. As mentioned before, the Fermi surface of cuprates has 4 pairs of parallel segments, leading to a doping-dependent nesting wavevector $q$ . Interestingly, the nesting at wavevector $(q,q)$ is more effective than at wavevector $(q,0)$: In the former case one obtains an overlap between all 4 pairs of segments of the Fermi surface, while in the latter only 2 pairs are involved. This observation explains why the CDW is more pronounced at wavevector $(q,q)$ than at $(q,0)$, i.e. $R<1$ [@Note4]. In the case of the PDW signal, Eq. (\[eq:rhoq3\]), each segment of the Fermi surface is weighted by the corresponding value of $\Delta_\textbf{k}$. This factor strongly favors the wavevector $(q,0)$, which connects antinodes to antinodes, with respect to $(q,q)$, which connects antinodes to nodes. Hence, for PDWs $R>1$ in agreement with the numerical result mentioned above. --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- -------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- \(a) CDW \(b) PDW ![Energy-momentum dependence of the inelastic response $\chi(\textbf{q},\Omega)$ along the line $(q, 0)$ in the weak coupling approach for (a) CDWs, Eq. (\[eq:rhoq2\]), and (b) PDWs, Eq. (\[eq:rhoq3\]). Same numerical values as in Fig. \[fig:CPDW\]. The black dashed lines in (a) are located at $\Omega/t = \pm 2\Delta_{0}/t$.[]{data-label="Fig:inelastic"}](InelasticCDWq0.pdf "fig:"){width="25.00000%"} ![Energy-momentum dependence of the inelastic response $\chi(\textbf{q},\Omega)$ along the line $(q, 0)$ in the weak coupling approach for (a) CDWs, Eq. (\[eq:rhoq2\]), and (b) PDWs, Eq. (\[eq:rhoq3\]). Same numerical values as in Fig. \[fig:CPDW\]. The black dashed lines in (a) are located at $\Omega/t = \pm 2\Delta_{0}/t$.[]{data-label="Fig:inelastic"}](InelasticPDW1.pdf "fig:"){width="25.00000%"} --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- -------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- The experimental data strongly supports the PDW scenario: (i) Transverse and longitudinal one-dimensional scans in the vicinity of $(q,0)$ show that the scattering amplitude peaks in both directions [@peng2018re]. This experimental finding is in stark contradiction with the CDW case, where a saddle-point is expected, and agrees with the PDW case (see Fig. \[fig:CPDW\]). (ii) The peak at $(q,q)$ is small [@GhiringhelliPrivate] or absent [@arpaia2019dynamical; @kang2019evolution], indicating that $R>1$, whereas for CDW this should be the dominant peak. Both observations are consistent with the PDW scenario only. [*Energy dependence –* ]{}Let us now turn to the energy dependence of the response function $\chi(\textbf{q},\Omega)$. Fig. \[Fig:inelastic\] shows our theoretical predictions for incipient CDWs, Eq. (\[eq:rhoq2\]), and PDWs, Eq. (\[eq:rhoq3\]). The energy dependence of the two graphs is very different: $\chi(\textbf{q},\Omega)$ is peaked at $\Omega \approx 2\Delta_{0}$ for CDWs and at $\Omega=0$ for PDWs. This discrepancy can be rationalized by noting that charge impurities create particle-hole pairs and, hence, need to overcome the energy gap $\Delta$. In contrast, local modulations of the pairing gap can create two holes (or two particles) at the same energy, below (or above) the gap. As a consequence, the response to CDWs is peaked at $2\Delta_0$, while the response to PDWs is peaked at zero energy. Recent energy-resolved inelastic X-ray scattering (RIXS) experiments [@ghiringhelli2012long; @chaix2017dispersive; @peng2018re; @da2018coupling; @arpaia2019dynamical] show that the signal is peaked at (or close to) zero energy and, again, favor the PDW scenario. Furthermore, our theoretical model accounts for the experimental observation of a dispersive peak that departs from the zero-energy peak towards higher energies [@chaix2017dispersive]. [*Hubbard model –* ]{}The weak coupling approach considered above does not take into account the interactions between quasiparticles, which can enhance the CDW and PDW fluctuations and lead to a competition between them. To capture these effects, we now consider a two-dimensional extended Hubbard model with on-site repulsion $U$ and NN attraction $V$. This model shows several competing phases, such as the Mott insulator and the d-wave superconductor, that are generic to cuprates [@micnas1988extended; @micnas1990superconductivity; @monthoux1994self; @newns1995van; @husslein1996quantum; @takigawa2004quasiparticle]. Under the usual mean-field approximation $n_{j}=\sum_{\sigma=\pm}\langle c^{\dagger}_{j, \sigma}c_{j, \sigma} \rangle/2$, $\Delta_{\hat{e}, j}=\sum_{\sigma=\pm}\sigma\langle c_{j, -\sigma}c_{j+\hat{e}, \sigma} \rangle/2$, the Hamiltonian reads $$\begin{aligned} \label{eq:hubbard} H&=-t\sum_{\langle i, j \rangle,\sigma} c^{\dagger}_{i, \sigma} c^{\phantom{\dagger}}_{j, \sigma} -t'\sum_{\langle\langle i, j \rangle \rangle,\sigma} \left(c^{\dagger}_{i, \sigma} c^{\phantom{\dagger}}_{j, \sigma} +\mathrm{H.c.}\right)\\ \nonumber &+ U \sum_{j, \sigma} n^{\phantom{\dagger}}_{j} c^{\dagger}_{j, \sigma} c^{\phantom{\dagger}}_{j, \sigma} + V \sum_{\hat{e}, j,\sigma}\left( \Delta^{\phantom{\dagger}}_{\hat{e}, j} c^{\dagger}_{j, \uparrow} c^{\dagger}_{j+\hat{e}, \downarrow} +\mathrm{H.c.} \right) \end{aligned}$$ where $\langle\cdot,\cdot\rangle$ denotes NN, $\langle\langle\cdot,\cdot\rangle\rangle$ denotes NNN, $\hat{e}$ connects NN sites, and H.c. stands for Hermitian conjugate. For $V<0$, the self-consistent solution of Eq. (\[eq:hubbard\]) delivers a superconductor with d-wave order parameter $\Delta_j = \frac{1}{4} (\Delta_{\hat{x}, j}+\Delta_{-\hat{x}, j} - \Delta_{\hat{y}, j}-\Delta_{-\hat{y}, j})$. \[t\] --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- \(a) site impurity ($\delta U$) \(b) bond impurity ($\delta V$) ![Two dimensional maps of the density fluctuations $n_\textbf{q}$ in the Hubbard model with a local modulation of the interaction: (a) on a single site, (b) on a single bond. For clarity, subfigure (b) has been symmetrized by 90 degrees. Numerical parameters: $x=0.16$, $t'/t=-0.6$ and $U/t=1.5, V/t=-0.5, ~\delta U/t = \delta V/t = 0.2$.[]{data-label="fig:hubbard"}](CDWdensFT1.pdf "fig:"){width="25.00000%"} ![Two dimensional maps of the density fluctuations $n_\textbf{q}$ in the Hubbard model with a local modulation of the interaction: (a) on a single site, (b) on a single bond. For clarity, subfigure (b) has been symmetrized by 90 degrees. Numerical parameters: $x=0.16$, $t'/t=-0.6$ and $U/t=1.5, V/t=-0.5, ~\delta U/t = \delta V/t = 0.2$.[]{data-label="fig:hubbard"}](PDWdensFT1.pdf "fig:"){width="25.00000%"} --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- To study the interplay between CDW and PDW we add a local impurity, which generically leads to spatial modulations of both the charge $n_{j}$ and the pairing gap $\Delta_{j}$. We classify the resulting density wave as CDW or PDW depending on which modulation is dominant, by comparing the relative standard deviations ${\delta n}$ and ${\delta \Delta}$ [@Note5]. For simplicity, we focus here on two representative impurities (see the Supplemental Materials for additional examples): a single site with $U \to U+\delta U$ and a single bond with $V \to V+\delta V$. In the former case, we find that the density wave has mixed CDW/PDW character ($\delta\Delta/\delta n \approx 1 $), while in the latter case it has a predominant PDW character ($\delta \Delta/\delta n \approx 140$). To model the response to X-ray scattering, we now focus on the Fourier transformed density $n_{\bf q}$, Fig. \[fig:hubbard\]. As expected, we find that interactions enhance density wave instabilities and lead to narrower peaks, with longer range correlations. In our calculations, the width of these peaks is limited by the system size ($L=26$ unit cells), suggesting that the Hubbard model is consistent with long ranged CDW/PDW modulations. By comparing the two subplots, we observe that the on-site impurity leads to peaks at both wavevectors $(q,q)$ and $(q,0)$, while the bond impurity leads to pronounced peaks in at the wavevector $(q,0)$ only. This is consistent with our proposal to compare the intensities of the two peaks to distinguish between CDW and PDW modulations. [*Discussion –* ]{}In this paper we described a method to distinguish between incipient CDW and PDW fluctuations, based on the analysis of X-ray scattering experiments. When these spatial modulations coexist with a homogeneous superconducting gap $\Delta_0$, both oscillations couple directly to any physical observable, such as the charge density and the tunneling density of states. Hence, both CDWs and PDWs can be detected using different experimental technique, including X-ray scattering, scanning tunneling spectroscopy (STS), and scanning Josephson probes. Here we focused on X-ray scattering and showed that one-dimensional cuts are not unequivocal, because they can be adequately fitted by both types of modulations. In contrast, two-dimensional maps of CDWs and PDWs are very different: the former are peaked at wavevector $(q,q)$, while the latter has stronger peaks at $(q,0)$. These two types of density waves can be further distinguished by energy-resolved RIXS measurements: CDWs are peaked at $\Omega=2\Delta_ {0}$, while PDWs are peaked at $\Omega=0$. These results strongly rely on our simplifying assumption of noninteracting quasiparticles, scattered by local and isotropic pinning centers. In principle, other sources of inhomogeneity, as well as strong electronic correlations, can yield different results and enhance CDW and PDW signals in other directions. To address this point, we considered an attractive Hubbard model, which demonstrated that our method to identify fluctuations with a dominant CDW or PDW character remains valid in the presence of strong interactions. In particular, we showed that local modulations of the pairing mechanism (in our case, a nearest-neighbor attractive interaction) give rise to density waves peaked at $(q,0)$ and with a predominant PDW character. Our theoretical model of PDW reproduces the main features of recent X-ray scattering experiments of superconducting NCCO, Hg1201, and BSCCO. Specifically, our theory explains why (i) the X-ray scattering signal is peaked at wavevector $(q,0)$, rather than at $(q,q)$, as expected for CDWs; (ii) the RIXS signal is peaked at frequency $\Omega=0$ and is accompanied by weaker dispersive inelastic peaks. Our findings also agree with earlier STS experiments of BSCCO [@howald2003periodic; @vershinin2004local; @hanaguri2004checkerboard], which found that the incommensurate checkerboard order has a dominant PDW character [@pereg2003theory; @nowadnick2012quasiparticle; @dalla2015exploring; @dai2018pair]. Attributing the X-ray signal to fluctuations of the pairing order parameter (PDW) explains its temperature dependence: these fluctuations are strongest at the critical temperature of superconducting order parameter, $T_c$, in agreement with the experimental observations ([@chang2012direct; @da2014ubiquitous]). Finally, The proposed PDW scenario explains why the signal detected in X-ray scattering of cuprates is orders of magnitude smaller than the one observed in ordinary CDW materials ($\delta n\ll \delta \Delta$). Our method can be further extended to include the effects of magnetic fields by considering a Hubbard model with complex hopping elements (Peierls substitution). In type-II superconductors, external magnetic field generate isolated vortices in whose core the pairing gap is locally suppressed. Hence, a magnetic vortex acts as a pinning site for a PDW modulation, in analogy to the bond impurity considered in this paper. Numerical studies of the Hubbard model in the presence of magnetic fields have indeed found that spatial modulations of the pairing gap develop in the proximity of the vortex core [@zhu2002spin; @takigawa2004quasiparticle; @simonucci2013temperature]. This finding is consistent with STS experiments demonstrating that the periodic modulations are mostly pronounced in the vicinity of the vortex core [@hoffman2002four; @matsuba2007anti; @yoshizawa2013high; @edkins2019magnetic], as well as with evidence that the density waves become long ranged at high magnetic fields [@doiron2007quantum; @wu2011magnetic; @wu2013emergence; @gerber2015three]. By locally suppressing superconductivity, large densities of magnetic vortices can lead to a long ranged PDW order. We thank Peter Abbamonte, Lucio Braicovich, Debanjan Chowdhury, Riccardo Comin, J. C. Seamus Davis, Andrea Damascelli, Eugene Demler, Giacomo Ghiringhelli, Marco Grilli, Jenny Hoffman, Amit Keren, Steve Kivelson, and Dror Orgad for useful discussions. This work is supported by the Israel Science Foundation grants No. 151/19 and 967/19. [10]{} Tranquada, J. M., Sternlieb, B. J., Axe, J. D., Nakamura, Y. & Uchida, S. Evidence for stripe correlations of spins and holes in copper oxide superconductors. , 375(6532):561, 1995. Doiron-Leyraud, N. et al. Quantum oscillations and the Fermi surface in an underdoped high-Tc superconductor. , 447(7144):565, 2007. Wu, T. et al. Magnetic-field-induced charge-stripe order in the high-temperature superconductor YBa$_2$Cu$_3$O$_y$. , 477(7363):191, 2011. Fradkin, E. & Kivelson, S. A. High-temperature superconductivity: Ineluctable complexity. , 8(12):864, 2012. LeBoeuf, D. et al. Thermodynamic phase diagram of static charge order in underdoped YBa$_2$Cu$_3$O$_y$. , 9(2):79, 2013. Wu, T. et al. Emergence of charge order from the vortex state of a high-temperature superconductor. , 4:2113, 2013. Hoffman, J. E. et al. A four unit cell periodic pattern of quasi-particle states surrounding vortex cores in Bi$_2$Sr$_2$CaCu$_2$O$_{8+\delta}$. , 295(5554):466–469, 2002. Howald, C., Eisaki, H., Kaneko, N., Greven, M. & Kapitulnik, A. Periodic density-of-states modulations in superconducting Bi$_2$Sr$_2$CaCu$_2$O$_{8+\delta}$. , 67(1):014533, 2003. Vershinin, M. et al. Local ordering in the pseudogap state of the high-tc superconductor Bi$_2$Sr$_2$CaCu$_2$O$_{8+ \delta}$. , 303(5666):1995–1998, 2004. Hanaguri, T. et al. A checkerboard electronic crystal state in lightly hole-doped Ca$_{2-x}$Na$_x$CuO$_2$Cl$_2$. , 430(7003):1001, 2004. Ghiringhelli, G. et al. Long-range incommensurate charge fluctuations in (Y, Nd)Ba$_2$Cu$_3$O$_{6+x}$. , 337(6096):821–825, 2012. Chang, J. et al. Direct observation of competition between superconductivity and charge density wave order in YBa$_2$Cu$_3$O$_{6.67}$. , 8(12):871, 2012. Torchinsky, D. H., Mahmood, F., Bollinger, A. T., Bo[ž]{}ovi[ć]{}, I. & and Gedik, N. Fluctuating charge-density waves in a cuprate superconductor. , 12(5):387, 2013. Blackburn, E. et al. X-ray diffraction observations of a charge-density-wave order in superconducting ortho-II YBa$_2$Cu$_3$O$_{6.54}$ single crystals in zero magnetic field. , 110(13):137004, 2013. Comin, E. et al. Charge order driven by fermi-arc instability in Bi$_2$Sr$_{2-x}$La$_x$CuO$_{6+\delta}$. , 343(6169):390–392, 2014. da Silva Neto, E. H. et al. Ubiquitous interplay between charge ordering and high-temperature superconductivity in cuprates. , 343(6169):393–396, 2014. Le Tacon, M. et al. Inelastic x-ray scattering in YBa$_2$Cu$_3$O$_{6.6}$ reveals giant phonon anomalies and elastic central peak due to charge-density-wave formation. , 10(1):52, 2014. Hashimoto, M. et al. Direct observation of bulk charge modulations in optimally doped Bi$_{1.5}$Pb$_{0.6}$Sr$_{1.54}$CaCu$_2$O$_{8+ \delta}$. , 89(22):220511, 2014. Tabis, W. et al. Charge order and its connection with fermi-liquid charge transport in a pristine high-Tc cuprate. , 5:5875, 2014. Huecker, M. et al. Competing charge, spin, and superconducting orders in underdoped Yba$_2$Cu$_3$O$_y$. , 90(5):054514, 2014. Achkar, A. J. et al. Impact of quenched oxygen disorder on charge density wave order in YBa$_2$Cu$_3$O$_{6+ x}$. , 113(10):107002, 2014. Gerber, S. et al. Three-dimensional charge density wave order in YBa$_2$Cu$_3$O$_{6.67}$ at high magnetic fields. , 350(6263):949–952, 2015. Hamidian, M. H. et al. Magnetic-field induced interconversion of cooper pairs and density wave states within cuprate composite order. , 2015. Peng, Y. Y. et al. Direct observation of charge order in underdoped and optimally doped Bi$_2$(Sr, La)$_2$CuO$_{6+\delta}$ by resonant inelastic x-ray scattering. , 94(18):184511, 2016. Chaix, L. et al. Dispersive charge density wave excitations in Bi$_2$Sr$_2$CaCu$_2$O$_{8+\delta}$. , 13(10):952, 2017. Peng, Y. Y. et al. Re-entrant charge order in overdoped (Bi, Pb)$_{2.12}$Sr$_{1.88}$CuO$_{6+\delta}$ outside the pseudogap regime. , 17(8):697, 2018. Jang, H. et al. Superconductivity-insensitive order at $q \sim 1/4$ in electron-doped cuprates. , 7(4):041066, 2017. da Silva Neto, E. H. et al. Coupling between dynamic magnetic and charge-order correlations in the cuprate superconductor Nd$_{2-x}$Ce$_x$CuO$_4$. , 98(16):161114, 2018. Bluschke, M. et al. Adiabatic variation of the charge-density-wave phase diagram in the 123 cuprate (Ca$_x$La$_{1-x}$)(Ba$_{1.75-x}$La$_{0.25+x}$)Cu$_3$O$_y$. , 100(3):035129, 2019. Kang, M. et al. Evolution of charge order topology across a magnetic phase transition in cuprate superconductors. , 15(4):335, 2019. Castellani, C., Di Castro, C. & Grilli, M. Singular quasiparticle scattering in the proximity of charge instabilities. , 75(25):4650, 1995. Castellani, C., Di Castro, C. & Grilli, M. The charge-density-wave quantum-critical-point scenario. , 282:260–263, 1997. He, R. H. et al. From a single-band metal to a high-temperature superconductor via two thermal phase transitions. , 331(6024):1579–1583, 2011. Loret, B. et al. Intimate link between charge density wave, pseudogap and superconducting energy scales in cuprates. , 15(8):771, 2019. Chen, H. D., Vafek, O., Yazdani, A. & Zhang, S. C. Pair density wave in the pseudogap state of high temperature superconductors. , 93(18):187002, 2004. Lee, P. A. Amperean pairing and the pseudogap phase of cuprate superconductors. , 4(3):031017, 2014. P[é]{}pin, C., De Carvalho, V. S., Kloss, T. & Montiel, X. Pseudogap, charge order, and pairing density wave at the hot spots in cuprate superconductors. , 90(19):195207, 2014. Freire, H., De Carvalho, V. S. & P[é]{}pin, C. Renormalization group analysis of the pair-density-wave and charge order within the fermionic hot-spot model for cuprate superconductors. , 92(4):045132, 2015. Wang, Y., Agterberg, D. F. & Chubukov, A. Interplay between pair-and charge-density-wave orders in underdoped cuprates. , 91(11):115103, 2015. Wang, Y., Agterberg, D. F. & Chubukov, A. Coexistence of charge-density-wave and pair-density-wave orders in underdoped cuprates. , 114(19):197001, 2015. Edkins, S. D. et al. Magnetic field–induced pair density wave state in the cuprate vortex halo. , 364(6444):976–980, 2019. Hamidian, M. H. et al. Detection of a cooper-pair density wave in Bi$_2$Sr$_2$CaCu$_2$O$_{8+ x}$. , 532(7599):343, 2016. Du, Z. et al. Imaging the energy gap modulations of the cuprate pair-density-wave state , 580(7801):65, 2020. Agterberg, D. F. et al. The physics of pair density waves. , 11:231, 2020. Dalla Torre, E. G., Benjamin, D., He, Y., Dentelski, D. & Demler, E. Friedel oscillations as a probe of fermionic quasiparticles. , 93(20):205117, 2016. Arpaia, R. et al. Dynamical charge density fluctuations pervading the phase diagram of a cu-based high-Tc superconductor. , 365(6456):906–910, 2019. Sachdev, S. & La Placa, R. Bond order in two-dimensional metals with antiferromagnetic exchange interactions. , 111(2):027202, 2013. Efetov, K. B., Meier, H. & P[é]{}pin, C. Pseudogap state near a quantum critical point. , 9(7):442, 2013. Allais, A., Bauer, j. & Sachdev, S. Density wave instabilities in a correlated two-dimensional metal. , 90(15):155114, 2014. Caplan, Y., Wachtel, G. & Orgad, D. Long-range order and pinning of charge-density waves in competition with superconductivity. , 92(22):224504, 2015. Caplan, Y. & Orgad, D. Dimensional crossover of charge-density wave correlations in the cuprates. , 119(10):107002, 2017. Altland, A. & Simons, B.D. (Cambridge university press, 2010). According to the interpretation of the pseudogap as a precursor of the superconducting pairing gap [@emery1995importance], this approach applies to the pseudogap region as well. Massidda, S., Hamada, N., Yu, J. & Freeman, A. J. Electronic structure of Nd-Ce-Cu-O, a fermi liquid superconductor. , 157(3):571–574, 1989. Increasing the hole doping reduces the antinodal distance and, accordingly, reduces the CDW/PDW wavevector observed in BSCCO and YBCO . Some authors disputed this interpretation due to the mismatch between the wavevector observed in X-ray experiments and the antinodal distance observed in ARPES. We point out that this mismatch is rather small (of the order of 10%) and is of the same order as the mismatch between the antinodal distances obtained by fitting the ARPES data with different phenomenological band structures . Furthermore X-ray scattering is sensitive to the bulk of the material, while ARPES is a surface probe. Hence, the mismatch between these two probes may be due to the common bulk-surface dichotomy of the chemical potential. Note that the integrals in Eq. (\[eq:rhoq2\]) involve diverging functions. To normalize these expressions, we substituted $\Omega \to \Omega +i\Gamma $ with $\Gamma /t=0.1$ and summed over $N=301$ equally spaced points, see the Supplemental Materials. We checked that the qualitative features of the resulting plots do not depend on the normalization procedure. Norman, M. R., Randeria, M., Ding, H. & Campuzano, J. C. Phenomenological models for the gap anisotropy of Bi$_2$Sr$_2$CaCu$_2$O$_8$ as measured by angle-resolved photoemission spectroscopy. , 52(1):615, 1995. Markiewicz, R. S., Sahrakorpi, S., Lindroos, M., Lin, H. & Bansil, A. One-band tight-binding model parametrization of the high-Tc cuprates including the effect of $k_z$ dispersion. , 72(5):054519, 2005. Fulde, P. & Ferrell, R. A. Superconductivity in a strong spin-exchange field. , 135(3A):A550, 1964. Larkin, A. I. & Ovchinnikov, Y. N. Nonuniform state of superconductors. , 20(3):762–762, 1965. Berg, E., Fradkin, E. & Kivelson, S. A. Theory of the striped superconductor. , 79(6):064515, 2009. Berg, E., Fradkin, E., Kivelson, S. A. & Tranquada, J. M. Striped superconductors: how spin, charge and superconducting orders intertwine in the cuprates. , 11(11):115004, 2009. See also the Supplemental Materials for a simplified model of the Fermi surface of cuprates predicting that for CDWs with $q/2\pi =0.25$, $R=3/4$. Ghiringhelli, G. private communication (unpublished). Micnas, R., Ranninger, J. & Robaszkiewicz, S. An extended hubbard model with inter-site attraction in two dimensions and high-tc superconductivity. , 21(6):L145, 1988. Micnas, R., Ranninger, J. & Robaszkiewicz, S. Superconductivity in narrow-band systems with local nonretarded attractive interactions. , 62(1):113, 1990. Monthoux, P. H. & Scalapino, D. J. Self-consistent $d_{x^2-y^2}$ pairing in a two-dimensional Hubbard model. , 72(12):1874, 1994. Newns, D. M., Tsuei, C. C. & Pattnaik, P.C. Van hove scenario for d-wave superconductivity in cuprates. , 52(18):13611, 1995. Husslein, T. et al. Quantum monte carlo evidence for d-wave pairing in the two-dimensional hubbard model at a van hove singularity. , 54(22):16179, 1996. Takigawa, M., Ichioka, M. & Machida, K. Quasiparticle structure in antiferromagnetism around the vortex and nuclear magnetic relaxation time. , 73(2):450–458, 2004. The relative standard deviation of $X$ is defined as $\delta X \equiv (\Sigma _{j} X^{2}_{j})^{1/2}/\Sigma _{j}X_{j}$. Pereg-Barnea, T. & Franz, M. Theory of quasiparticle interference patterns in the pseudogap phase of the cuprate superconductors. , 68(18):180506, 2003. Nowadnick, E. A., Moritz, B. & Devereaux, T. P. Quasiparticle interference and the interplay between superconductivity and density wave order in the cuprates. , 86(13):134509, 2012. Dalla Torre, E. G., He, Y., Benjamin, D. & Demler, E. Exploring quasiparticles in high-Tc cuprates through photoemission, tunneling, and x-ray scattering experiments. , 17(2):022001, 2015. Dai, Z., Zhang, Y. H., Senthil, T. & Lee, P. A. Pair-density waves, charge-density waves, and vortices in high-t c cuprates. , 97(17):174511, 2018. Zhu, J. X., Martin, I. & Bishop, A. R. Spin and charge order around vortices and impurities in high-t c superconductors. , 89(6):067003, 2002. Simonucci, S., Pieri, P. & Strinati, G. C. Temperature dependence of a vortex in a superfluid fermi gas. , 87(21):214507, 2013. Matsuba, K. et al. Anti-phase modulation of electron-and hole-like states in vortex core of Bi$_2$Sr$_2$CaCu$_2$O$_x$ probed by scanning tunneling spectroscopy. , 76(6):063704, 2007. Yoshizawa, S. et al. High-resolution scanning tunneling spectroscopy of vortex cores in inhomogeneous electronic states of Bi$_2$Sr$_2$CaCu$_2$O$_x$. , 82(8):083706, 2013. Emery, V. J. & Kivelson, S. A. Importance of phase fluctuations in superconductors with small superfluid density. , 374(6521):434, 1995. Wise, W. D. et al. Charge-density-wave origin of cuprate checkerboard visualized by scanning tunnelling microscopy. , 4(9):696, 2008. King, P. D. C. et al. Structural origin of apparent fermi surface pockets in angle-resolved photoemission of Bi$_2$Sr$_{2-x}$La$_x$CuO$_{6+\delta}$. , 106(12):127005, 2011.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We employ a self-consistent Green’s function approach to investigate the spin-wave relaxation $\Gamma(p)$ in diluted magnetic semiconductors. We find the trend of the spin-wave relaxation strongly depends on the ratio of the itinerant and impurity spin densities. For density ratios in the Ruderman-Kittel-Kasuya-Yosida phase, $\Gamma(p)$ decreases even though thermal fluctuations increase. On the other hand, in the strong coupling phase, an interesting peak structure appears. We discuss the implications of our numerical results for experiments.' author: - 'J. E. Bunder' - 'Shih-Jye Sun' - 'Hsiu-Hau Lin' title: 'Spin-Wave Relaxation in Diluted Magnetic Semiconductors within the Self-Consistent Green’s Function Approach' --- Carrier-mediated ferromagnetism, found in diluted magnetic semiconductors (DMS) such as III-V semiconductors doped with transition metals, opens up the possibility of electric manipulations of magnetic and optical properties because of the exchange coupling between the localized moments of the transition metal and the itinerant spins in the semiconducting bands [@MacDonald05; @Wolf01]. In recent years, experimental progress has stimulated intense theoretical investigations of DMS [@Akai98; @Dietl00; @Konig00; @Schliemann01; @Litvinov01; @Priour04], which mainly focus on estimates of the critical temperature. Although some experimental data is available [@Crooker96; @Akimoto97], little theoretical attention has been paid to spin dynamics, which is crucially important in understanding the spin coherence. To analyze how spin waves relax in DMS, both the itinerant and impurity spins need to be treated on an equal footing. In this Letter, we employ the self-consistent Green’s function approach[@Yang01; @Konig01; @Bouzerar02; @Sun04b; @Bouzerar05; @Sun06], which retains both spatial and thermal fluctuations, to investigate the spin-wave relaxation at finite temperatures. The Green’s functions allow us to numerically compute the spin spectral function and therefore the spin-wave relaxation. One may naively expect a smooth increase in the spin-wave relaxation as thermal fluctuations grow. However, our numerical results go against common intuition. For instance, in the Ruderman-Kittel-Kasuya-Yosida (RKKY) phase, the spin relaxation, shown in Fig. \[Fig1\], actually decreases with increasing temperatures. To study the temperature dependence of the spin-wave relaxation, we begin with the simple Zener model, $$H=H_0+J\int d^3 r\: \bm{S}(r) \cdot \bm{\sigma}(r),$$ where $\bm{S}(r)$ and $\bm{\sigma}(r)$ are the impurity and itinerant spin densities, coupled by the antiferromagnetic ($J>0$) exchange coupling. For simplicity, the kinetic energy of the itinerant carriers $H_0$ is assumed to be parabolic, $\epsilon_p = p^2/2m^*$. We take the typical values[@Ohno98] for the exchange coupling $J=150$ meV nm$^{3}$ and the impurity spin density $n_{I}=1$ nm$^{-3}$ but vary the ratio of itinerant to impurity spin densities $n_{h}/n_{I}$ in the following. We now introduce the self-consistent Green’s function approach. Note that, instead of treating spin waves as bosons, the Green’s function approach respects the spin kinematics which is crucially important when thermal fluctuations are not small. The thermal Green’s function, describing the (impurity) spin-wave propagation, is defined as $ iD(r_1, r_2;t)\equiv \big\langle \big\langle T\big[S^+(r_1,t) S^-(r_2,0)\big] \big\rangle \big\rangle, $ where $T$ is the time ordering operator and the double bracket implies both quantum and thermal averages. To complete the self-consistency, it is necessary to introduce another Green’s function which describes the correlations between the impurity and the itinerant spins, $ iF(r_1, r'_1, r_2;t) \equiv \big\langle\big\langle T\big[ \psi^{\dag}_{\uparrow}(r_1,t) \psi^{}_{\downarrow}(r_1', t) S^{-}(r_2, 0) \big] \big\rangle \big\rangle. $ To simplify the calculations, we assume that the disorder is not strong, i.e. not close to the percolation threshold, and the virtual crystal approximation is valid. Therefore, after a coarse-graining procedure, the translational invariance is approximately restored and the Green’s functions only depend on the relative distance. This approximation is justified by recent numerical calculations [@Schliemann01a] which found that magnetization curves with different disorder configurations are almost identical. The dynamical equations for the Green’s functions can be derived by the standard approach and higher-order Green’s functions may be decoupled within the random phase approximation. Making use of the translational invariance and the decoupling scheme, the dynamical equation for $D(r;t)$ in the Fourier space is $$\begin{aligned} \Omega D(p,\Omega) &=& 2 \langle S^z \rangle - J \langle \sigma^z \rangle D(p,\Omega) \nonumber\\ &+& J \langle S^z \rangle \int \frac{d^3k}{(2\pi)^3} F(k,k+p;\Omega), \label{D}\end{aligned}$$ where $F(k,k+p;\Omega)$ is the Fourier transformed Green’s function $F(r,r';t)$. Similarly, the other dynamical equation is $F(k, k+p;\Omega)= G (k, k+p;\Omega) D(p,\Omega)$ with $$G (k, k+p;\Omega)=\frac{J}{2}\frac{ f_{\uparrow}(\epsilon_k)- f_{\downarrow}(\epsilon_{k+p})}{\Omega+\epsilon_k- \epsilon_{k+p}+\Delta+i\gamma_{h}}, \label{eq:F-D}$$ where $f_{\alpha}(\epsilon_k)= [e^{\beta(\epsilon_k+\alpha\Delta/2-\mu)}-1]^{-1}$ is the Fermi distribution function for itinerant carriers with the Zeeman gap $\Delta=J\langle S_z\rangle$. On substituting the expression for $F(k,k+p;\Omega)$ into Eq. (\[D\]), the spin-wave propagator is solved, $$\begin{aligned} D(p,\Omega)&= & \frac{2\langle S_z\rangle}{\Omega-\Sigma(p,\Omega)+i \gamma_{I}},\label{eq:D2}\end{aligned}$$ where the prefactor $2\langle S^z \rangle$ comes from the exact treatment of spin-wave kinematics. The self-energy $\Sigma(p,\Omega)$ arises from interactions between the itinerant and the impurity spins and contains two terms, $$\begin{aligned} \Sigma(p,\Omega) = -J \langle\sigma^z\rangle +\Delta \int \frac{d^3k}{(2\pi)^3}G(k,k+p;\Omega). \label{eq:gf}\end{aligned}$$ If only the first term is retained, it coincides with the Weiss mean-field approximation. However, since the first term is always real, dropping the second term would remove all information about the Landau damping of the spin waves, described by the imaginary part of the self-energy $\Sigma_{I}(p,\Omega)$, inside the Stoner continuum. Once the spin-wave dispersion is obtained from $\omega_p -\Sigma_{R}(p,\omega_p)=0$, the polarization of the impurity spins can be computed by Callen’s formula [@Callen63], $$\frac{\langle S^{z}\rangle}{n_{I}}=S-\langle n_{sw}\rangle +\frac{(2S+1)\langle n_{sw}\rangle^{2S+1}}{(1+\langle n_{sw}\rangle)^{2S+1}-\langle n_{sw}\rangle^{2S+1}}. \label{eq:sz}$$ where the average number of spin waves is $ \langle n_{sw}\rangle=(1/n_I)\int d^3 k/(2\pi)^3 \:\:[e^{\beta\omega_k}-1]^{-1}$. The difference between the independent spin-wave theory and the self-consistent Green’s function method lies in the third term, which correctly accounts for the spin kinematics. The magnetization curves for different density ratios $n_{h}/n_{I}$ are shown in Fig. \[Fig2\]. While this is not the main focus of our results, we emphasize the sensitivity in the shape of the magnetization curve to the density ratio. From previous studies in the diffusive regime, there are three phases for the Zener model – mean-field, RKKY and strong coupling. For $n_h/n_I=0.3$, it belongs to the RKKY phase and the magnetization curve is rather conventional. For $n_h/n_I=0.1$, it lies in the strong coupling phase and the magnetization curve decreases linearly with a very small gradient and, only in the narrow regime near the critical temperature, the magnetization dives to zero and the phase transition occurs. Similar magnetization curves, quite different to what we expect for a Heisenberg-like model, have been observed in some experiments [@Edmonds04]. Finally, we choose $n_h/n_I=0.01$ (also in the strong coupling phase) to demonstrate that the magnetization curve can turn [*concave*]{}, though this is not necessarily related to percolation near the metal-insulator transition. The unique feature of the Green’s function approach is that we can also study the temperature evolution of the spin spectral function, $A(p,\omega) = -(1/\pi) \mbox{Im} D(p,\Omega)$, as shown in Fig. \[Fig1\]. From Eq. (\[eq:D2\]), the imaginary part of the spin-wave propagator takes the Lorentzian form, $$\begin{aligned} A(p,\omega) = 2\langle S^z \rangle \left(\frac{Z}{\pi}\right) \frac{\Gamma(p)}{(\Omega-\omega_p)^2 + \Gamma(p)^2}, \label{eq:imagD}\end{aligned}$$ where $Z \approx 1$ is the spectral weight for the gapless spin wave. After Taylor expansion of $\Sigma_I(p,\Omega)$ in the vicinity of the spin-wave dispersion $\Omega = \omega_p$, the relaxation rate of the spin waves $\Gamma(p) \approx \Sigma_I(\omega_p)$. The numerical results for the spin-wave relaxation rate versus temperature with $n_{h}/n_{I}=0.3, 0.1$ are shown in Fig. \[Fig3\]. One may expect that increasing thermal fluctuations lead to a increasing relaxation rate, but our numerical results disagree. In the RKKY phase with $n_{h}/n_{I}=0.3$, the spin relaxation rate $\Gamma(p)$ decreases as the temperature increases. This reduction can be understood as the weakening of the effective coupling strength between the itinerant and impurity spins in proportion to the spin polarization. However, since the decoupling scheme in the self-consistent approach only includes decaying channels in the diagonal parts (as in all random phase approximations), the computation of $\Gamma(p)$ breaks down in the vicinity of the critical temperature where off-diagonal channels dominate. Therefore, one should not take the vanishing $\Gamma(p)$ at the Curie temperature seriously. However, in the low temperature regime, the diagonal channels do dominate and our numerical result of decreasing $\Gamma(p)$ with increasing temperature is a real effect. In the strong coupling phase with $n_{h}/n_{I}=0.1$, the trend of the spin relaxation is completely different. The spin relaxation $\Gamma(p)$ is small and does not change much at low temperatures, then a peak structure appears just before hitting the critical temperature. This non-monotonic trend can be explained by the energy shift of the Stoner continuum. In the strong coupling phase, the Stoner continuum is separated from the gapless spin waves at low temperatures by a finite gap. The gap suppresses $\Gamma(p)$ and makes it insensitive to temperature changes. The gap disappears at the temperature where the polarization of the itinerant spins $\langle \sigma^z \rangle$ is no longer fully polarized. Due to the intersection of the spin-wave dispersion and the Stoner continuum, enhanced spin relaxation is expected. An analysis of our numerical results reveal that the partial polarization of $\langle \sigma^z \rangle$ and the peak in $\Gamma(p)$ occur at the same temperature. One may notice the close relationship between the magnetization curve and the spin relaxation rate and their sensitive dependence on the density ratio $n_h/n_I$. This is reminiscent of the interesting differences between as-grown samples and annealed ones[@MacDonald05]. The annealing process removes interstitial impurities and reduces the disorder strength. However, if the as-grown sample is already in the diffusive regime, the annealed one can also be described by the virtual crystal approximation. We believe the major difference arises from the increase in carrier density and thus the density ratio $n_h/n_I$. The annealing process increases $n_h/n_I$ from the strong coupling to the RKKY phase which changes the shape of the magnetization curve. Therefore, it would be exciting to study the spin dynamics for both as-grown and annealed samples and compare the experimental outcomes with our numerical predictions. Note that we did not include the realistic six-band Luttinger model here. In principle, our approach can be generalized to include more bands by introducing more Green’s functions but the self-consistent equations will be rather complicated, not fatal though. However, it is important to emphasize that the ratio $n_h/n_I$ to enter the RKKY regime is expected to be larger, as compared to the estimate from the two-band model. We thank Sankar Das Sarma and Allan MacDonald for fruitful discussions. HHL acknowledges supports from National Science Council in Taiwan through grants NSC-93-2120-M-007-005 and NSC-94-2112-M-007-031 and also supports from KITP by the National Science Foundation under Grant No. PHY99-07949. [99]{} For a recent review, see A. H. MacDonald, P. Schiffer and N. Samarth, Nat. Mat. **4**, 195 (2005). S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S. von Molnar, M. L. Roukes, A. Y. Chtchelkanova, and D. M. Treger, Science **294**, 1488 (2001). H. Akai, Phys. Rev. Lett. [**81**]{}, 3002 (1998). T. Dietl, H. Ohno, F. Matsukura, J. Cibert and D. Ferrand, Science [**287**]{}, 1019 (2000). J. König, H.-H. Lin, and A. H. MacDonald, Phys. Rev. Lett. [**84**]{}, 5628 (2000). J. Schliemann, J. König, H.-H. Lin and A. H. MacDonald, Appl. Phys. Lett. [**78**]{}, 1550 (2001). V. I. Litvinov and V. K. Dugaev, Phys. Rev. Lett. [**86**]{}, 5593 (2001). D. J. Priour, Jr., E. H. Hwang and S. Das Sarma, Phys. Rev. Lett. **92**, 117201 (2004). S. A. Crooker, J. J. Baumberg, F. Flack, N. Samarth and D. D. Awschalom, Phys. Rev. Lett. **77**, 2814 (1996). R. Akimoto, K. Ando, F. Sasaki, S. Kobayashi and T. Tani Phys. Rev. B **56**, 9726 (1997). M.-F. Yang, S.-J. Sun and M.-C. Chang, Phys. Rev. Lett. [**86**]{}, 5636 (2001). J. Konig, H.-H. Lin and A. H. MacDonald, Phys. Rev. Lett. [**86**]{}, 5637 (2001). G. Bouzerar and T. P. Pareek, Phys. Rev. B [**65**]{}, 153203 (2002). S.-J. Sun and H.-H. Lin, Phys. Lett. A [**327**]{}, 73 (2004). G. Bouzerar, T. Ziman and J. Kudrnovsky, Europhys. Lett. **69**, 812 (2005). S.-J. Sun and H.-H. Lin, Eur. Phys. J. B [**49**]{}, 403 (2006). H. Ohno, Science **281**, 951 (1998). J. Schliemann, J. König and A. H. MacDonald, Phys. Rev. B **64**, 165201 (2001). Herbert B. Callen, Phys. Rev. [**[130]{}**]{}, 890 (1963). K. W. Edmonds, P. Bogusawski, K. Y. Wang, R. P. Campion, S. N. Novikov, N. R. S. Farley, B. L. Gallagher, C. T. Foxon, M. Sawicki, T. Dietl, M. Buongiorno Nardelli and J. Bernholc, Phys. Rev. Lett. **92**, 037201 (2004).
{ "pile_set_name": "ArXiv" }
--- abstract: 'The widths of charge balance function in high energy hadron-hadron and relativistic heavy ion collisions are studied using the Monte Carlo generators PYTHIA and AMPT, respectively. The narrowing of balance function as the increase of multiplicity is found in both cases. The mean parton-freeze-out time of a heavy-ion-collision event is used as the characteristic hadronization time of the event. It turns out that for a fixed multiplicity interval the width of balance function is consistent with being independent of hadronization time.' author: - 'Du Jiaxin,Li Na  and   Liu Lianshou[^1]' title: | The Narrowing of Charge Balance Function\ and Hadronization Time in Relativistic Heavy Ion Collisions --- Introduction ============ The relativistic heavy ion collision experiments at CERN-SPS and especially at the relativistic heavy ion collider RHIC in Brookhaven National Lab provide clear evidence for the production of a dense matter in the collision processes [@evidence]. The central question is whether this matter is purely hadronic or has been going through a quark-parton phase. There exist experimental evidences in favor of the existence of a quark-parton phase at the early stage of collision processes [@qgp], but in view of the importance of this question, further confirmation is needed. Recently, the rapidity correlation between oppositely charged particles, which has been used in  [@ee] and hadron-hadron collisions [@hh] to study the hadronization in these processes, is proposed [@clock] as a measure of the hadronization time in relativistic heavy ion collisions. It is argued that if the system produced in heavy ion collisions has undergone a quark-parton phase, the hadronization will occur at a later time and, therefore, the temperature will be lower and the diffusive interaction with other particles will be lesser than those in the direct hadronization without going through a quark-parton phase. These will result in a narrower charge balance function for a system with quark-parton phase than that without such a phase. Two heavy ion experiments [@star; @na49] have measured the balance function at various centralities and for different colliding nuclei. A narrowing of the balance function is indeed observed with increasing centrality of the collision and with increasing size of the colliding nuclei. These observations are consistent with the assumption that the narrowing of balance function is correlated with late hadronization. On the other hand, recently it is reported [@na22] that in hadron-hadron collisions at $\sqrt s=22$ GeV the balance function also becomes narrower as the increasing of multiplicity. Therefore, whether the observed narrowing of balance function in relativistic heavy ion collisions is due to late hadronization or is simply due to the multiplicity effect is an open question. In this letter this question is examined using the Monte Carlo generators PYTHIA [@pythia] and AMPT [@ampt]. The former is a standard Monte Carlo generator with string fragmentation as hadronization scheme. There is not any quark-parton phase in this model and the hadronization is almost instantaneous. On the other hand, the latter is a “multi-phase” model, with a transport of quark-parton before hadronization. The results from PYTHIA will be presented in section II. The hadronization time in AMPT model is discussed in section III, and its connection with the width of balance function is presented in section IV. Section V is summary and discussion. The width of balance function in PYTHIA model ============================================= The balance function is defined as [@na49] $$\begin{aligned} B(\dt y|Y_{\rm w}) &=& \frac{1}{2}\left[ \frac{\la n_{+-}(\dt y)\ra-\la n_{++}(\dt y)\ra}{\la n_+\ra}\right. \nonumber \\ & &\hs15mm \left.+\frac{\la n_{-+}(\dt y)\ra-\la n_{--}(\dt y)\ra}{\la n_-\ra}\right],\end{aligned}$$ where, $n_{+-}(\dt y)$, $n_{-+}(\dt y)$ and $n_{++}(\dt y)$, $n_{--}(\dt y)$ are the numbers of pairs of opposite- and like-charged particles satisfying the criteria that they fall into the rapidity window $Y_{\rm w}$ and that their relative rapidity equals $\dt y$; $n_+$ and $n_-$ are the numbers of positively and negatively charged particles, respectively, in the interval $Y_{\rm w}$. The balance function $B(\dt y|Y_{\rm w})$ represents the probability that the balancing charges are separated by $\dt y$ [@clock]. The mean of $\dt y$ [@na49] y\_= is defined as the [*width of balance function*]{}. Proton-proton collision events are generated at four c.m. energies —— 22, 64, 130 and 200 GeV using PYTHIA5.720 generator. The event number for each sample is 100,000. The widths $\la\dt y\ra_{\infty}$ of balance function in the full phase space are calculated for different (charged) multiplicity bins and plotted in Fig. 1. ![\[Fig. 1\] The width of full-phase-space balance function for different multiplicities in p-p collisions at $\sqrt{s}=22,\; 64,\;130,\;200$ GeV.](figures/fig1.epsi){width="2.5in"} It can be seen from the figure that in this model even for p-p collision, where no quark-parton phase is expected and the hadronization is almost instantaneous, the width of balance function decreases with the increase of multiplicity,  the width of balance function is narrower for higher multiplicity. This effect has nothing to do with hadronization time. On the other hand, by definition balance function measures the correlation length between oppositely charged particles. For comparison we have calculated the standard 2-particle correlation function [@Foa] of oppositely charged particles &&-12mm R\^[+-]{}(y\_1,y\_2) = ( .\ && + . )-1for different multiplicities in p-p collision at c.m. energy $\sqrt s =200$ GeV, for $y_1=0$, $y_2=y$. The results plotted in Fig.2 show that the width of $R$ is consistent with being independent of multiplicity. A possible explanation of the width of $R$ is cluster decay. Comparing with the definition of balance function, Eq.(1), we see that it is the difference between the correlations of opposite- and like-charged particles that shows a clear multiplicity dependence, which is unrelated with cluster decay, and is mainly due to the string fragmentation mechanism implemented in PYTHIA model. ![\[Fig. 2\] The 2-particle correlation function $R^{+-}(0,y)$ of oppositely charged particles as function of $y$ for different multiplicities in p-p collision at c.m. energy $\sqrt s = 200$ GeV.](figures/fig2.epsi){width="2.5in"} It can also be seen from Fig.1 that the width of balance function depends on collision energy. For the same multiplicity, the higher the collision energy is, the wider the width of balance function. ![\[Fig. 4\] The width of balance function in the rapidity region $[-3,\, 3]$ for different multiplicities in p-p collision at $\sqrt{s}=22,\; 64,\;130,\;200$ GeV.](figures/fig3.epsi){width="2.5in"} ![\[Fig. 4\] The width of balance function in the rapidity region $[-3,\, 3]$ for different multiplicities in p-p collision at $\sqrt{s}=22,\; 64,\;130,\;200$ GeV.](figures/fig4.epsi){width="2.5in"} However, it should be noticed that the full rapidity region is wider for higher energy, Fig3. In order to get rid of the influence of the width of rapidity region we calculate the width of balance function in the region $-3 \leq y\leq 3$ for all four energies. The results, presented in Fig. 4, show that, when the (average) rapidity density $\fc{\Dt n}{\Dt y}$ is the same, the width of balance function is almost independent of energy, especially for high $\fc{\Dt n}{\Dt y}$. That is, in hadron-hadron collisions the width of balance function depends essentially [*only*]{} on multiplicity and is consistent with being independent on energy. How does the width of balance function behave in nucleus-nucleus collisions will be studied in the next two sections. Hadronization time in AMPT model ================================ The Monte Carlo generator AMPT1.11 is a multi-phase transport model, which contains a quark-parton transport phase before hadronization. The initial spatial and momentum distributions of hard partons and soft string excitations are obtained from the HIJING [@hijing] model. The parton cascade follows Zhang’s parton-cascade (ZPC) model [@zpc], which is based on two-body pQCD scattering with screening masses. When interaction ceases, the partons are recombined with their parent strings to form hadrons according to LUND string fragmentation mechanism [@pythia]. Then the scatterings among the resulting hadrons are described by a relativistic transport (ART) model [@art] which includes baryon-baryon, baryon-meson and meson-meson elastic and inelastic scatterings. In our calculation the parton cross section is chosen to be 10 mb. If the colliding nuclei are large and the energy is high, then the parton cascade will last long enough to make the parton distribution arrive at an equilibrium distribution [@ko]. However, no equilibrium thermodynamics has been included in the model. In particular, there is not any equilibrium phase transition from parton phase to hadron phase. Therefore, there is no unique hadronisation time for the whole system. Each parton has its own hadronisation time, or freeze out time $\ft$. In order to study the correlation, if any, between the width of balance function and the hadronization time, we use the event mean of parton freeze out time $\ft$ = \_[i=1]{}\^ \_i as the characteristic hadronization time for an event, where $\npt$ is the number of partons in the event, ${\ft}_i$ is the freeze out time of the $i$th parton. ![\[Fig. 5\] Scattering plots of $\bft$ vs. $n_{\rm ch}$ in Au-Au collision at $\sqrt{s_{NN}}=200$ GeV for two different centralities.](figures/fig5.eps){width="3.2in"} The AMPT1.11 (default) generator is utilized to generate Au-Au collision events at $\sqrt{s_{NN}}=200$GeV. Two event samples with 250,000 and 770,000 events, respectively, are generated for two centralities $b\leq 7$fm and $b>7$fm. In Fig’s.5 are shown the scattering plots of $\bft$ vs. $n_{\rm ch}$ in Au-Au collisions at $\sqrt{s_{NN}}=200$ GeV for the two different centralities —— $b\leq 7$fm and $b>7$fm, respectively. It can be seen that in central collisions ($b\leq 7$fm) $\bft$ is larger than 5fm, while in peripheral collisions ($b > 7$fm) $\bft$ is concentrated at $\bft<5$fm. That is, central collision events hadronize later than peripheral ones. The distributions of event-mean parton freeze out time $\bft$ for the two different centralities in Au-Au collision at $\sqrt{s_{NN}}=200$ GeV are shown in Fig.6. ![\[Fig. 6\] The distributions of event-mean parton freeze out time $\bft$ for two different centralities in Au-Au collision at $\sqrt{s_{NN}}=200$ GeV.](figures/fig6.epsi){width="2.5in"} Relation between the width of Balance function and hadronization time ===================================================================== In order to study the correlation between the width of balance function and the characteristics of single event — event-mean parton freeze out time $\bft$ and/or multiplicity $n_{\rm ch}$, each centrality sample is divided into sub-samples according to the intervals of mean parton freeze-out time $\bft$ and the resulting sub-samples are further divided into sub-samples by different multiplicity intervals. The widths of balance function in the rapidity region $Y_{\rm W}=[-3,3]$ for different intervals of mean parton freeze-out time $\bft$ and multiplicity $n_{\rm ch}$ are shown in Fig.7. The abscissa of the figure is the average multiplicity in the corresponding multiplicity interval. It can be seen from the figure that in all cases the width of balance function decreases with the increasing of multiplicity, while in the same multiplicity interval, the width of balance function is consistent of being constant, independent of the hadronization time. Therefore, to use the narrowing of balance function in relativistic heavy ion collision as a measure of hadronization time and as a signal of QGP is doubtful. ![\[Fig. 7\] The widths of balance function in the rapidity region $Y_{\rm W}=[-3,3]$ for different mean parton freeze-out time $\bft$ intervals versus multiplicity $n_{\rm ch}$ in Au-Au collision at $\sqrt{s_{NN}}=200$ GeV.](figures/fig7.epsi){width="3.0in"} Summary and discussion ====================== It is found using PYTHIA Monte Carlo that the width of charge balance function decreases with the increasing of multiplicity in high energy hadron-hadron collisions, where the hadronization is almost instantaneous. The relation between the hadronization time and the width of charge balance function in relativistic heavy ion collisions is examined using the default AMPT1.11 Monte Carlo generator. The mean parton freeze out time of an event is used as the characteristic hadronization time of the event. The narrowing of balance function as the increase of multiplicity is found to exist also for relativistic heavy ion collisions, while for a fixed multiplicity interval the width of balance function is consistent with being independent of hadronization time. Therefore, to use the narrowing of balance function in relativistic heavy ion collisions as a measure of hadronization time and as a signal of QGP is doubtful. It should be noticed that AMPT model is a multi-phase [*transport*]{} model. There is no equilibrium thermodynamics included in the model. In particular, there is no equilibrium phase transition, and consequently no unique hadronization time for an event. To use the average of parton freeze out time in an event as the characteristic hadronization time of the event is a crude approximation. Therefore, the observation made in the present work on the independence of the width of balance function on hadronization time is only a first step. Further investigation along this line is needed. [**Acknowledgement**]{}  This work is supported by NSFC under project 10375025 and by the Cultivation Fund of the Key Scientific and Technical Innovation Project of Ministry of Education of China NO CFKSTIP-704035. [99]{} J. Adams (STAR Collab.), Nucl. Phys. A 757 (2005) 102; K. Adcox (PHENIX Collab), Nuclear Physics A 757 (2005) 184. M. Gyulassy, e-Print-arXiv nucl-th/0403032. R. Brandelik [*et al.*]{}, Phys. Lett. B [**100**]{}, 357 (1981); M. Althoff [*et al.*]{}, Z. Phys. C [**17**]{}, 5 (1983); H. Aihara [*et al.*]{}, Phys. Rev. Lett. [**53**]{}, 2199 (1984) and [**57**]{}, 3140 (1986); P.D. Acton et al., Phys. Lett. B [**305**]{}, 415 (1993). D. Drijard [*et al.*]{}, Nucl. Phys. B [**155**]{}, 269 (1979) and [**166**]{}, 233 (1980); M. Barth [*et al.*]{}, Z. Phys. C [**16**]{}, 291 (1983); D.H. Brick [*et al.*]{}, Z. Phys. C [**31**]{}, 59 (1986); I.V. Ajinenko [*et al.*]{}, Z. Phys. C [**43**]{}, 37 (1989). S.A. Bass, P. Danielewicz, and S. Pratt, Phys. Rev. Lett. [**85**]{}, 2689 (2000). J. Adams [*et al.*]{} (STAR Coll.), Phys. Rev. Lett. [**90**]{}, 172301 (2003). C. Alt [*et al.*]{} (NA49 Coll.), Phys. Rev. C [**71**]{}, 034903 (2005). M.R. Atayan [*et al.*]{} (NA22 Coll.), hep-ex/0504027, to be published in Phys. Lett. B. L. FoÀ, Phys. Rep. [**22**]{}, 26 (1975). T. Sjöstrand, [Comput. Phys. Commun.]{} [**82**]{}, 74 (1994). Zi-Wei Lin, Che Ming Ko, Bao-An Li and Bin Zhang, nucl-th/0411110 v2 (2004). S. Jeon, and S. Pratt, Phys. Rev. C [**65**]{}, 044902(2002). M. Gyulassy and X. N. Wang, Comput. Phys. Commun. [**83**]{}, 307 (1994). B. Zhang, Comput. Phys. Commun. [**109**]{}, 193 (1998). B.A. LI and C.M. KO, Phys. Rev. C52 (1995) 2037. C. M. Ko, private communication. [^1]: Email: [email protected]
{ "pile_set_name": "ArXiv" }
--- address: | Institut für Theoretische Physik, Universität Frankfurt, Germany\ and\ Bogolyubov Institute for Theoretical Physics, Kyiv, Ukraine author: - 'A.P. Kostyuk' title: Statistical hadronization of charm in heavy ion collisions --- The thermal hadron gas (HG) model explains successfully the chemical composition of light hadrons produced by colliding nuclei in a wide range center-of-mass energies[@HG]. A large body of experimental data can be well fitted with only three[^1] free parameters: the temperature $T$, baryonic chemical potential $\mu_b$ and volume $V$ at the point of chemical freeze-out. This suggested the idea[@GG] to see, whether this model can be applied also to charm particles. A straightforward generalization, when particles containing heavy (anti)quarks are treated in the same way as light hadrons, could hardly be successful. The estimated relaxation time for charm production and annihilation in HG or even in a deconfined medium appeared to be much larger than the total lifetime of the thermal system. Therefore, it has been generally believed that the production mechanism of heavy-flavored hadrons is completely different from that of light ones. For instance, the standard picture of charmonium production in nucleus-nucleus collisions assumes that charmonia are created exclusively at the initial stage of the reaction in primary nucleon-nucleon collisions. During the subsequent evolution of the system, the number of hidden charm mesons is reduced because of absorption of pre-resonance charmonium states in the nuclei (the normal nuclear suppression), interactions of charmonia with secondary hadrons (comovers), dissociation of $c\bar{c}$ bound states in the deconfined medium. The last mechanism was expected to be especially strong and charmonia were proposed to be used as a probe of the state of matter created at the early stage of the collision. It was found that the $J/\psi$ suppression with respect to Drell-Yan muon pairs measured in proton-nucleus and nucleus-nucleus collisions with light projectiles can be explained by the normal nuclear suppression alone[@NA38]. In contrast, the NA50 experiment with a heavy projectile and target (lead-lead) revealed essentially stronger $J/\psi$ suppression for central collisions[@anomalous]. This [*anomalous*]{} $J/\psi$ suppression was attributed[@evidence] to formation of quark-gluon plasma (QGP). In the same time, purely hadronic scenarios still cannot be excluded[@comover]. Despite of quite successful agreement with the $J/\psi$ data, the standard scenario seems to be in trouble explaining the $\psi'$ yield. The recent lattice simulations[@Karsch] suggest that the temperature of $\psi'$ dissociation $T_d(\psi')$ lies far below the deconfinement point[@Satz01] $T_c$: $T_d(\psi') \approx 0.1$–$0.2 T_c$. Therefore, not only the quark-gluon plasma, but also a hadronic co-mover medium should completely eliminate $\psi'$ charmonia in central Pb+Pb collisions at SPS. However, the experiment revealed a sizable $\psi'$ yield (see, for instance[@Bordalo]). It was observed[@Shuryak] that $\psi'$ to $J/\psi$ ratio decreases with centrality only in peripheral lead-lead collisions, but saturates at sufficiently large number of participants $N_p \ge 100$. The value of the ratio in (semi)central collisions is approximately constant and equal to the ratio of the densities of these charmonium states in an equilibrium HG. Hence, on one hand, the production of charm cannot be thermal, because of the large relaxation time, on the other hand, the multiplicity ratio of different charmonium states is thermal. This paradox is resolved in the statistical coalescence model (SCM)[@BMS]. In this model, the charm quarks $c$ and antiquarks $\bar{c}$ are created at the initial stage of A+A reaction in the hard parton collisions. This is similar to the standard approach. But like in the thermal model[@GG], the formation of observed hadrons with open and hidden charm takes place later at the hadronization stage near the point of chemical freeze-out. Production and annihilation of charm quark-antiquark pairs at all stages after the initial are neglected. Therefore the number $c$ and $\bar{c}$ in the system may be very far from chemical equilibrium, but their distribution over different species of hadrons with open and hidden charm is controlled by the lows of statistical mechanics. In the ‘pure’ SCM, which will be a subject of the present discussion, it is assumed that the hot strongly-interacting medium destroys all initially formed (‘primordial’) charmonia or even prevents their formation. A combined scenario, assuming that some of primordial charmonia do survive suppression in the medium, but additional hidden charm meson can be as well formed at hadronization, has been also considered[@GR]. Within the grand canonical approach, the deviation of the amount of (anti)charm in the system from its equilibrium value can be taken into account simply by multiplying the multiplicities of all open single (anti)charm species in equilibrium hadron gas by some factor $\gamma_c$: $$\label{X} \langle X \rangle = \gamma_c N_X(T,\mu_b,V) = \gamma_c \exp\left(c_X \frac{\mu_c}{T} \right) \tilde{N}_X(T,\mu_b,V).$$ Here $c_X = \pm 1$ is the charm of the species $X$. The charm chemical potential $\mu_c$ is needed to keep the total balance of open charm and anticharm in the thermal system. At zero $\mu_c$ but nonzero baryonic chemical potential $\mu_b$, the total number of open charm in the equilibrium HG $$\label{N1} \tilde{N}_1 = \sum_X \tilde{N}_X(T,\mu_b,V)$$ is [*not*]{} equal to that of anticharm $$\label{N1b} \tilde{N}_{\bar{1}} = \sum_{\bar{X}} \tilde{N}_{\bar{X}}(T,\mu_b,V).$$ This is because, for instance, a positive value of $\mu_b$ enhances the number of open charm baryons, that contribute to (\[N1\]), and suppresses the number of antibaryons in (\[N1b\]). (The sums in (\[N1\]) and (\[N1b\]) run over all species of single open charm and anticharm hadrons, respectively.) The value of $\mu_c$ should be chosen to restore the balance: $$\label{balance} \exp(\mu_c/T) \tilde{N}_1 = \exp(-\mu_c/T) \tilde{N}_{\bar{1}}.$$ A hidden charm meson contains a charm quark as well as an antiquark. Therefore, their thermal numbers are not influenced by the charm chemical potential. To take into account the deviation from the chemical equilibrium, one should multiply the thermal numbers of charmonia by $\gamma_c^2$: $$\label{Y} \langle Y \rangle = \gamma_c^2 \tilde{N}_Y(T,\mu_b,V).$$ Provided that the average number of charm quarks and antiquarks $\langle C \rangle = \langle \bar{C} \rangle$ in the system is known, the factor $\gamma_c$ can be found from the equation: $$\label{eqgam} \langle C \rangle = \gamma_c \exp(\mu_c/T) \tilde{N}_1 + \gamma_c^2 \tilde{N}_H \ .$$ Here $\tilde{N}_H$ is the total number of hidden charm particles in the system at chemical equilibrium: $$\label{NH} \tilde{N}_H = \sum_Y \tilde{N}_Y(T,\mu_b,V).$$ (The sum runs over all charmonium states.) The thermal number of hidden charm is much smaller than that of open charm. Therefore, if the number of $c\bar{c}$ pairs in the system is not extremely large (does not exceed the chemical equilibrium value by a factor of several hundreds), the hidden charm term in (\[eqgam\]) can be neglected and $\gamma_c$ is simply given by $$\label{gamc} \gamma_c = \frac{\langle C \rangle}{\exp(\mu_c/T) \tilde{N}_1}.$$ The grand canonical approach does not exactly correspond to the physical reality. It keeps balance only between the [*average*]{} quantities of charm and anticharm, while the partition function includes also configurations with nonequal numbers of charm quarks and antiquarks in the system. This does not influence the result, if the average number of $c\bar{c}$ pairs in the system is large $\langle C \rangle,\langle \bar{C} \rangle \gg 1$ . In the opposite case[@We], however, it should be taken into account that strong interactions conserve charm (weak interactions can be neglected), therefore the number $C$ of charm quarks in the system is always [*exactly*]{} equal to the number $\bar{C}$ of antiquarks. Let us first consider the situation, when the numbers of charm quarks and antiquarks are fixed (not necessary equal to each other). In this case, the partition function for particles with charm can be written as $$Z_{C\bar{C}}(T,\mu_b,V) = \frac{\tilde{N}_1^{C}}{C !} \frac{\tilde{N}_{\bar{1}}^{\bar{C}}}{\bar{C}!} + \tilde{N}_H \frac{\tilde{N}_1^{C-1}}{(C-1) !} \frac{\tilde{N}_{\bar{1}}^{\bar{C}-1}}{(\bar{C}-1)!} + \dots$$ The dots stand for terms corresponding to more than one hidden charm meson in the system and also for terms including double and triple charm baryons. These terms, as it was already mentioned above, can be safely neglected, unless $C$ and $\bar{C}$ are extremely large. As far as $\tilde{N}_H \ll \tilde{N}_1 \tilde{N}_{\bar{1}}$ (which is true in all practically interesting situations), the numbers of open charm and anticharm hadrons are approximately equal to $C$ and $\bar{C}$, respectively, i.e. almost all charm quarks and antiquarks hadronize into open charm particles. Only a very small fraction form charmonia. The total number of hidden charm can be easily calculated: $$\label{HCC} \langle H \rangle_{C\bar{C}} = \tilde{N}_H \frac{\partial \log Z_{C\bar{C}}}{\partial \tilde{N}_H} \approx C\bar{C} \frac{\tilde{N}_H}{\tilde{N}_1 \tilde{N}_{\bar{1}}}$$ In reality, however, the number $K$ of $c\bar{c}$ pairs is not fixed. It fluctuates from one event to another. The pairs are produced in independent collisions of nucleons whose number is large and the probability to produce a $c\bar{c}$ pair in a single collision is small. Therefore, the fluctuations should follow the Poisson law: $$\label{pois} P(C=\bar{C}=K) = \exp(- \langle c\bar{c} \rangle_{AB(b)}) \frac{\left( \langle c\bar{c} \rangle_{AB(b)}\right)^{K}}{K!},$$ where $\langle c\bar{c} \rangle_{AB(b)}$ stands for average number of $c\bar{c}$ pairs produced by nuclei A and B colliding at impact parameter $b$. The number of hidden charm averaged over all events at fixed centrality is given by the convolution of Eq.(\[HCC\]) with the probability (\[pois\]). The result reads: $$\label{HAB} \langle H \rangle_{AB(b)} = \langle c\bar{c} \rangle_{AB(b)} (\langle c\bar{c} \rangle_{AB(b)}+1) \frac{\tilde{N}_H}{\tilde{N}_1 \tilde{N}_{\bar{1}}}\ .$$ If one interested in the multiplicity of a particular charmonium species, one should replace the $\tilde{N}_H$ in the numerator of (\[HAB\]) by the thermal multiplicity of this species. It must be also taken into account that an additional contribution to multiplicities of low-lying charmonium states comes from decays of excited ones. Note that the ratios of the numbers of different charmonium states in SCM are exactly the same as in equilibrium HG: $$\label{chrat} \frac{\langle \psi' \rangle_{AB(b)}}{\langle J/\psi \rangle_{AB(b)}} = \frac{\tilde{N}_{\psi'}}{\tilde{N}_{J/\psi}}.$$ It agrees with the behavior of $\psi'$ to $J/\psi$ ratio in central Pb+Pb collisions at CERN SPS. This is not the case for peripheral events ($N_p < 100$). It can be explained in the following way: SCM is valid if the momenta of charm (anti)quarks (not their number!) are thermalized. This is possible only in a sufficiently large system. The equation (\[HAB\]) gives the total ($4\pi$) multiplicity of hidden charm. In real experimental situation, however, measurements are made in a limited rapidity window. In the most simple case, when the fraction of charmonia that fall into the relevant rapidity window does not depend on the centrality, one can merely use Eq.(\[HAB\]) multiplied by some factor $\xi < 1$. This situation is likely to be relevant to charmonium production at CERN SPS, where the multiplicity of light hadrons, that determine the freeze-out volume of the system, are approximately proportional to the number of nucleon participants $N_p$ at all rapidities. Indeed, SCM fits well the measured[@anomalous; @evidence; @Ramelo] $J/\psi$ to Drell-Yan ratio at CERN SPS (see Fig. \[Jpsi\_PbPb\], the details of the fit can be found in Ref.[@SPS]). =3.9in At BNL RHIC, the situation is different: the total ($4\pi$) multiplicity of light hadrons are still approximately proportional to the number of participants, while at midrapidity, it grows faster[@PHOBOS]. The centrality dependence of charmonium production at different rapidities should in this case be also different. The formula (\[HAB\]) should be generalized, in such a way that the charmonium multiplicity not only in the whole system, but in a part of it, a subsystem (like a limited rapidity interval), could be calculated. Let, if $c$ (or $\bar{c}$) is present in the system, the probability to find it in the subsystem is $\xi < 1$. Then, if the number of $c\bar{c}$ pairs in the whole system is $K$, the probability to find $C$ charm quarks in the subsystem is given by the binomial law: $$\label{binom} w(C | K) = \frac{K!}{C! (K - C)!} \xi^C (1-\xi)^{K-C}.$$ We assume that the distributions of quarks and antiquarks are uncorrelated,[^2] and the probability distribution of the number of antiquarks $\bar{C}$ is given by the same binomial law. If the number of pairs $K$ is distributed in accordance with the Poisson law (\[pois\]), the average multiplicity of charmonia in the subsystem is $$\begin{aligned} \label{Jpsi1} \langle H \rangle_{AB(b)}^{(\xi)} &\approx & \sum_{K=0}^{\infty} P(K) \sum_{C=0}^{K} w(C | K) \sum_{\bar{C}=0}^{K} w(\bar{C} | K) \ C \bar{C} \frac{\tilde{N}_H}{\tilde{N}_1 \tilde{N}_{\bar{1}}} \\ &=& \xi^2 \ \langle c\bar{c} \rangle_{AB(b)} \ (\langle c\bar{c} \rangle_{AB(b)}+1) \ \frac{\tilde{N}_H}{\tilde{N}_1 \tilde{N}_{\bar{1}}}. \nonumber\end{aligned}$$ Here the meaning of $\langle c\bar{c} \rangle_{AB(b)}$ is the same as in Eq.(\[HAB\]): the average [*total*]{} (4$\pi$) number of $c\bar{c}$ pairs in an event with given centrality. In contrast, $\tilde{N}_H$, $\tilde{N}_1$ and $\tilde{N}_{\bar{1}}$ are related to the subsystem. These are the thermal multiplicities of charmonia, open charm and anticharm calculated with the thermodynamic parameters fitted within the HG model to the chemical composition and multiplicity of light hadrons [*in the rapidity window under interest*]{}. In Fig.\[Jpsi\_RHIC\], the SCM result is compared to the preliminary PHENIX data[@data] for $J/\psi$ production in the central rapidity interval at the top RHIC energy ($\sqrt{s}=200$ GeV). The values of freeze-out parameters were taken from Ref.[@BMSR]: $T=177$ MeV and $\mu_{B}=29$ MeV. The dependence of volume on the centrality was chosen to reproduce the measured multiplicity of charged hadrons.[@PHOBOS] The value of $\langle c\bar{c} \rangle_{AB(b)}$ was calculated in the Glauber approach. The charm production cross section in nucleon-nucleon collisions was fixed at[@Averbeck] $\sigma^{NN}_{c\bar{c}} = 650\ \mu$b, while $\xi$ was considered as a free parameter. The minimum value of $\chi^2/dof=0.75$ is reached at $\xi \approx 0.18$. =3.9in As is seen, the SCM dependence of the $J/\psi$ multiplicity [*at midrapidity*]{} per binary collision on the centrality is almost flat for $N_p \gtrsim 100$ in contrast to the [*total*]{} $J/\psi$ yield , which is expected to grow in the same centrality region.[@We_RHIC] This is because the multiplicity of light hadrons[@PHOBOS] and, consequently, the hadronization volume [*at midrapidity*]{} grows faster with the centrality than the [*total*]{} volume. If the influence of the in-nuclear modification of the parton distribution functions (shadowing) on charm production at RHIC is essential, a decrease should be observed instead of saturation (see Fig.\[Jpsi\_RHIC\]). The centrality dependence of $J/\psi$ multiplicity in the QGP suppression scenario[@Blaizot] is also shown in Fig.\[Jpsi\_RHIC\]. The parameters of the model were fixed from SPS data[@anomalous; @evidence; @Ramelo] and then extrapollated to RHIC. As is seen, the observed yield can hardly be explained by survived primordial charmonia only. It seems that $c\bar{c}$ coalescence at late stages of the reaction is here a dominant if not the only charmonium production mechanism. In conclusion, the statistical coalescence model is consistent with the data on $J/\psi$ production in heavy nucleus collisions at SPS and RHIC energies, while the standard scenario of $J/\psi$ dissociation in QGP predicts too strong suppression at RHIC that is not favored by the experimental data. Acknowledgment {#acknowledgment .unnumbered} ============== The author is thankful to the Alexander von Humboldt Foundation (Germany) for financial support. [99]{} P. Braun-Munzinger, I. Heppe and J. Stachel, Phys. Lett. B [**465**]{} (1999) 15;\ G. D. Yen and M. I. Gorenstein, Phys. Rev. C [**59**]{} (1999) 2788;\ F. Becattini [*et al.*]{}, Phys. Rev. C [**64**]{} (2001) 024901;\ W. Florkowski, W. Broniowski and M. Michalec, Acta Phys. Polon. B [**33**]{} (2002) 761. M. Gaździcki and M. I. Gorenstein, Phys. Rev. Lett.  [**83**]{} (1999) 4009. T. Matsui and H. Satz, Phys. Lett. B [**178**]{} (1986) 416;\ H. Satz, Rept. Prog. Phys.  [**63**]{} (2000) 1511. M. C. Abreu [*et al.*]{}, Phys. Lett. B [**466**]{} (1999) 408. M. C. Abreu [*et al.*]{} \[NA50 Collaboration\], Phys. Lett. B [**410**]{} (1997) 337;\ M. C. Abreu [*et al.*]{} \[NA50 Collaboration\], Phys. Lett. B [**450**]{} (1999) 456. M. C. Abreu [*et al.*]{} \[NA50 Collaboration\], Phys. Lett. B [**477**]{} (2000) 28. N. Armesto [*et al.*]{}, Nucl. Phys. A [**698**]{} (2002) 583. F. Karsch, E. Laermann and A. Peikert, Nucl. Phys. B [**605**]{} (2001) 579. S. Digal, P. Petreczky and H. Satz, Phys. Lett. B [**514**]{} (2001) 57. M. C. Abreu [*et al.*]{} \[NA50 Collaboration\], LIP-96-04 [*Invited talk at 26th International Symposium on Multiparticle Dynamics (ISMD 96), Faro, Portugal, 1-5 Sep 1996*]{}. H. Sorge, E. Shuryak and I. Zahed, Phys. Rev. Lett.  [**79**]{} (1997) 2775. P. Braun-Munzinger and J. Stachel, Phys. Lett. B [**490**]{} (2000) 196. L. Grandchamp and R. Rapp, Phys. Lett. B [**523**]{} (2001) 60; Nucl. Phys. A [**709**]{} (2002) 415. M. I. Gorenstein, A. P. Kostyuk, H. Stöcker and W. Greiner, Phys. Lett. B [**509**]{} (2001) 277; J. Phys. G [**27**]{} (2001) L47. L. Ramelo [*et al.*]{} \[NA50 Collaboration\], “Results on leptonic probes from NA50,” Presented at the Quark Matter 2002 conference, Nantes, France, July 18-24, 2002 A. P. Kostyuk, M. I. Gorenstein, H. Stocker and W. Greiner, Phys. Lett. B [**531**]{} (2002) 195; J. Phys. G [**28**]{} (2002) 2297. B. B. Back [*et al.*]{} \[PHOBOS Collaboration\], arXiv:nucl-ex/0301017; Phys. Rev. C [**65**]{} (2002) 061901. A. Andronic, P. Braun-Munzinger, K. Redlich and J. Stachel, arXiv:nucl-th/0209035; arXiv:nucl-th/0303036. A. D. Frawley \[PHENIX Collaboration\], arXiv:nucl-ex/0210013. R. Averbeck \[PHENIX Collaboration\], arXiv:nucl-ex/0209016. P. Braun-Munzinger, K. Redlich and J. Stachel, arXiv:nucl-th/0304013. M. I. Gorenstein [*et al.*]{}, J. Phys. G [**28**]{} (2002) 2151; Phys. Lett. B [**524**]{} (2002) 265. J. P. Blaizot, M. Dinh and J. Y. Ollitrault, Phys. Rev. Lett.  [**85**]{} (2000) 4012. [^1]: Sometimes one more parameter, the strangeness suppression factor $\gamma_s$ is introduced. [^2]: At this point, our approach differs from that of Ref.[@andr], where a strong correlation $C=\bar{C}$ between the numbers of $c$-quarks and $\bar{c}$-antiquarks is assumed.
{ "pile_set_name": "ArXiv" }
--- abstract: | In the present work we calculate the phonon-limited mobility in intrinsic $n$-type single-layer MoS$_2$ as a function of carrier density and temperature for $T>100$ K. Using a first-principles approach for the calculation of the electron-phonon interaction, the deformation potentials and Fr[ö]{}hlich interaction in the isolated MoS$_2$ layer are determined. We find that the calculated room-temperature mobility of $\sim 410$ cm$^2$ V$^{-1}$ s$^{-1}$ is dominated by optical phonon scattering via deformation potential couplings and the Fr[ö]{}hlich interaction with the deformation potentials to the intravalley homopolar and intervalley longitudinal optical phonons given by $4.1\times10^8$ eV/cm and $2.6\times10^8$ eV/cm, respectively. The mobility is weakly dependent on the carrier density and follows a $\mu \sim T^{-\gamma}$ temperature dependence with $\gamma=1.69$ at room temperature. It is shown that a quenching of the characteristic homopolar mode which is likely to occur in top-gated samples, boosts the mobility with $\sim 70$ cm$^2$ V$^{-1}$ s$^{-1}$ and can be observed as a decrease in the exponent to $\gamma=1.52$. Our findings indicate that the intrinsic phonon-limited mobility is approached in samples where a high-$\kappa$ dielectric that effectively screens charge impurities is used as gate oxide. author: - Kristen Kaasbjerg - 'Kristian S. Thygesen' - 'Karsten W. Jacobsen' title: 'First-principles study of the phonon-limited mobility in $n$-type single-layer MoS$_2$' --- Introduction ============ Alongside with the rise of graphene and the exploration of its unique electronic properties [@Geim:Graphene; @RMP:Graphene; @Sarma:RMP], the search for other two-dimensional (2D) materials with promising electronic properties has gained increased interest [@Novoselov:RPP]. Metal dichalcogenides which are layered materials similar to graphite, provide interesting candidates. Their layered structure with weak interlayer van der Waals bonds allows for fabrication of single- to few-layer samples using mechanical peeling/cleavage or chemical exfoliation techniques similar to the fabrication of graphene [@Geim:2D; @Fuhrer:Ultrathin; @Pati:Analogues; @Kis:MoS2Transistor]. However, in contrast to graphene they are semiconductors and hence come with a naturally occurring band gap—a property essential for electronic applications. The vibrational and optical properties of single- to few-layer samples of MoS$_2$ have recently been studied extensively using various experimental techniques [@Ryu:Anomalous; @Heinz:ThinMoS2; @Schuller:Photocarrier]. In contrast to the bulk material which has an indirect-gap, it has been demonstrated that single-layer MoS$_2$ is a direct-gap semiconductor with a gap of $1.8$ eV [@Heinz:ThinMoS2; @Wang:Emerging]. Together with the excellent electrostatic control inherent of two-dimensional materials, the large band gap makes it well-suited for low power applications [@Salahuddin:HowGood]. So far, electrical characterizations of single-layer MoS$_2$ have shown $n$-type conductivity with room-temperature mobilities in the range $0.5-3$ cm$^2$ V$^{-1}$ s$^{-1}$ [@Geim:2D; @Fuhrer:Ultrathin; @Kis:MoS2Transistor]. Compared to early studies of the intralayer mobility of bulk MoS$_2$ where mobilities in the range $100-260$ cm$^2$ V$^{-1}$ s$^{-1}$ were reported [@Mooser:Mobility], this is rather low. In a recent experiment the use of a high-$\kappa$ gate dielectric in a top-gated device was shown to boost the carrier mobility to a value of 200 cm$^2$ V$^{-1}$ s$^{-1}$ [@Kis:MoS2Transistor]. The observed increase in the mobility was attributed to screening of impurities by the high-$\kappa$ dielectric and/or modifications of MoS$_2$ phonons in the top-gated sample. ![(Color online) Atomic structure and conduction band of single-layer MoS$_2$. Left: Primitive unit cell and structure of an MoS$_2$ layer with the molybdenum and sulfur atoms positioned in displaced hexagonal layers. Right: Contour plot showing the lowest lying conduction band valleys as obtained with DFT-LDA in the hexagonal Brillouin zone of single-layer MoS$_2$. For $n$-type MoS$_2$, the low-field mobility is determined by the carrier properties in the $K,K'$-valleys which are separated in energy from the satellite valleys inside the Brillouin zone.[]{data-label="fig:mos2"}](mos2){width="0.99\linewidth"} ![(Color online) Atomic structure and conduction band of single-layer MoS$_2$. Left: Primitive unit cell and structure of an MoS$_2$ layer with the molybdenum and sulfur atoms positioned in displaced hexagonal layers. Right: Contour plot showing the lowest lying conduction band valleys as obtained with DFT-LDA in the hexagonal Brillouin zone of single-layer MoS$_2$. For $n$-type MoS$_2$, the low-field mobility is determined by the carrier properties in the $K,K'$-valleys which are separated in energy from the satellite valleys inside the Brillouin zone.[]{data-label="fig:mos2"}](conductionband_2d){width="0.99\linewidth"} In order to shed further light on the measured mobilities and the possible role of phonon damping due to a top-gate dielectric [@Kis:MoS2Transistor], we have carried out a detailed study of the phonon-limited mobility in single-layer MoS$_2$. Since phonon scattering is an intrinsic scattering mechanism that often dominates at room temperature, the theoretically predicted phonon-limited mobility sets an upper limit for the experimentally achievable mobilities in single-layer MoS$_2$ [@Sarma:100]. In experimental situations, the temperature dependence of the mobility can be useful for the identification of dominating scattering mechanisms and also help to establish the value of the intrinsic electron-phonon couplings [@Sarma:Hetero1]. However, the existence of additional extrinsic scattering mechanisms often complicates the interpretation. For example, in graphene samples impurity scattering and scattering on surface polar optical phonons of the gate oxide are important mobility-limiting factors [@Sarma:2DGraphene; @MacDonald:Massless; @Fuhrer:GrapheneSiO2; @Jena:HighKappa; @Kim:SO; @Sarma:Nonmonotonic]. It can therefore be difficult to establish the value and nature of the intrinsic electron-phonon interaction from experiments alone. Theoretical studies therefore provide useful information for the interpretation of experimentally measured mobilities. In the present work, the electron-phonon interaction in single-layer MoS$_2$ is calculated from first-principles using a density-functional based approach. From the calculated electron-phonon couplings, the acoustic (ADP) and optical deformation potentials (ODP) and the Fr[ö]{}hlich interaction in single-layer MoS$_2$ are inferred. Using these as inputs in the Boltzmann equation, the phonon-limited mobility is calculated in the high-temperature regime ($T > 100$ K) where phonon scattering often dominate the mobility. Our findings demonstrate that the calculated phonon-limited room-temperature mobility of $\sim 410$ cm$^2$ V$^{-1}$ s$^{-1}$ is dominated by optical deformation potential and polar optical scattering via the Fr[ö]{}hlich interaction. The dominating deformation potentials of $D^0_\text{HP} = 4.1\times10^8$ eV/cm and $D^0_\text{LO} = 2.6\times10^8$ eV/cm originate from the couplings to the homopolar and the intervalley polar LO phonon. This is in contrast to previous studies for bulk MoS$_2$, where the mobility has been assumed to be dominated entirely by scattering on the homopolar mode [@Mooser:Mobility; @Fivaz:Dimensionality; @Schmid:Layered]. Furthermore, we show that a quenching of the characteristic homopolar mode which is polarized in the direction normal to the MoS$_2$ layer can be observed as a change in the exponent $\gamma$ of the generic temperature dependence $\mu \sim T^{-\gamma}$ of the mobility. Such a quenching can be expected to occur in top-gated samples where the MoS$_2$ layer is sandwiched between a substrate and a gate oxide. The paper is organized as follows. In Section \[sec:II\] the band structure and phonon dispersion of single-layer MoS$_2$ as obtained with DFT-LDA are presented. The latter is required for the calculation of the electron-phonon coupling presented in Section \[sec:III\]. From the calculated electron-phonon couplings, we extract zero- and first-order deformation potentials for both intra and intervalley phonons. In addition, the coupling constant for the Fr[ö]{}hlich interaction with the polar LO phonon is determined. In Section \[sec:IV\], the calculation of the phonon-limited mobility within the Boltzmann equation is outlined. This includes a detailed treatment of the phonon collision integral from which the scattering rates for the different coupling mechanisms can be extracted. In Section \[sec:V\] we present the calculated mobilities as a function of carrier density and temperature. Finally, in Section \[sec:VI\] we summarize and discuss our findings. Electronic structure and phonon dispersion {#sec:II} ========================================== The electronic structure and phonon dispersion of single-layer MoS$_2$ have been calculated within DFT in the LDA approximation using a real-space projector-augmented wave (PAW) method [@GPAW; @GPAW1; @GPAW2]. The equilibrium lattice constant of the hexagonal unit cell shown in Fig. \[fig:mos2\] was found to be $a=3.14$ [Å]{} using a $11\times 11$ ${\mathbf{k}}$-point sampling of the Brillouin zone. In order to eliminate interactions between the layers due to periodic boundary conditions, a large interlayer distance of 10 [Å]{} has been used in the calculations. Band structure -------------- In the present study, the band structure of single-layer MoS$_2$ is calculated within DFT-LDA [@Eriksson:2D; @Hong:Interlayer]. While DFT-LDA in general underestimates band gaps, the resulting dispersion of individual bands, i.e. effective masses and energy differences between valleys, is less problematic. We note, however, that recent GW quasi-particle calculations which in general give a better description of the band structure in standard semiconductors [@GW], suggest that the distance between the $K,K'$-valleys and the satellite valleys located on the $\Gamma$-$K$ path inside the Brillouin zone not as large as predicted by DFT (see below) and that the valley ordering may even be inverted [@Ciraci:Functionalization; @Olsen:MoS2]. This, however, seems to contradict with the experimental consensus that single-layer MoS$_2$ is a direct-gap semiconductor [@Heinz:ThinMoS2; @Wang:Emerging] and further studies are needed to clarify this issue. In the following we therefore use the DFT-LDA band structure and comment on the possible consequences of a valley inversion in Section \[sec:VI\]. ![(Color online) Band structure of single-layer MoS$_2$. The inset shows a zoom of the conduction band along the path $\Gamma$-K-M. The parabolic band (red dashed line) with effective mass $m^*=0.48\;m_e$ is seen to give a perfect description of the conduction band valley in the $K$-point for the range of energies relevant for the low-field mobility.[]{data-label="fig:bandstructure"}](bandstructure){width="0.9\linewidth"} The calculated DFT-LDA band structure is shown in Fig. \[fig:bandstructure\]. It predicts a direct band gap of $1.8$ eV at the $K$-poing in the Brillouin zone. The inset shows a zoom of the bottom of the conduction band along the $\Gamma$-K-M path in the Brillouin zone. As demonstrated by the dashed line, the conduction band is perfectly parabolic in the $K$-valley. Furthermore, the satellite valley positioned on the path between the $\Gamma$-point and the $K$-point lies on the order of $\sim 200$ meV above the conduction band edge and is therefore not relevant for the low-field transport. The right plot in Fig. \[fig:mos2\] shows a contour plot of the lowest lying conduction band valleys in the two-dimensional Brillouin zone. Here, the $K,K'$-valleys that are populated in $n$-type MoS$_2$, are positioned at the corners of the hexagonal Brillouin zone. Due to their isotropic and parabolic nature, the part of the conduction band relevant for the low-field mobility can to a good approximation be described by simple parabolic bands $$\label{eq:parabolic} \varepsilon_{\mathbf{k}}= \frac{\hbar^2 k^2}{2m^*}$$ with an effective electron mass of $m^*= 0.48\;m_e$ and where $k$ is measured with respect to the $K,K'$-points in the Brillouin zone. The two-dimensional nature of the carriers is reflected in the constant density of states given by $\rho_0 = g_s g_v m^* / 2\pi\hbar^2$ where $g_s=2$ and $g_v = 2$ are the spin and $K,K'$-valley degeneracy, respectively. The large density of states which follows from the high effective mass of the conduction band in the $K,K'$-valleys, results in non-degenerate carrier distributions except for very high carrier densities. Phonon dispersion ----------------- The phonon dispersion has been obtained with the supercell-based small-displacement method [@Alfe:SDM] using a $9\times 9$ supercell. The resulting phonon dispersion shown in Fig. \[fig:phonon\_dispersion\] is in excellent agreement with recent calculations of the lattice dynamics in two-dimensional MoS$_2$ [@Ciraci:Lattice]. With three atoms in the unit cell, single-layer MoS$_2$ has nine phonon branches—three acoustic and six optical branches. Of the three acoustic branches, the frequency of the out-of-plane flexural mode is quadratic in $q$ for $q\rightarrow 0$. In the long-wavelength limit, the frequency of the remaining transverse acoustic (TA) and longitudinal acoustic (LA) modes are given by the in-plane sound velocity $c_\lambda$ as $$\label{eq:omega_acoustic} \omega_{{\mathbf{q}}\lambda} = c_\lambda q .$$ Here, the sound velocity is found to be $4.4\times 10^3$ and $6.5\times 10^3$ m/s for the TA and LA mode, respectively. The gap in the phonon dispersion completely separates the acoustic and optical branches even at the high-symmetry points at the zone-boundary where the acoustic and optical modes become similar. The two lowest optical branches belong to the non-polar optical modes. Due to an insignificant coupling to the charge carriers, they are not relevant for the present study. The next two branches with a phonon energy of $\sim 48$ meV at the $\Gamma$-point are the transverse (TO) and longitudinal (LO) polar optical modes where the Mo and S atoms vibrate in counterphase. In bulk polar materials, the coupling of the lattice to the macroscopic polarization setup by the lattice vibration of the polar LO mode results in the so-called LO-TO splitting between the two modes in the long-wavelength limit. The inclusion of this effect from first-principles requires knowledge of the Born effective charges [@Baroni:PhononSemi; @Gonze:DynamicalMatrices; @Wang:MixedSpace]. In two-dimensional materials, however, the lack of periodicity in the direction perpendicular to the layer removes the LO-TO splitting [@Hernandez:MonolayerBN]. The coupling to the macroscopic polarization of the LO mode will therefore be neglected here. ![Phonon dispersion of single-layer MoS$_2$ calculated with the small displacement method (see e.g. Ref. ) using a $9\times 9$ supercell. The frequencies of the two optical Raman active E$_{2g}$ and A$_{1g}$ modes at 48 meV and 50 meV, respectively, are in excellent agreement with recent experimental measurements [@Ryu:Anomalous].[]{data-label="fig:phonon_dispersion"}](phonon_dispersion){width="0.9\linewidth"} The almost dispersionless phonon at $\sim50$ meV is the so-called homopolar mode which is characteristic for layered structures. The lattice vibration of this mode corresponds to a change in the layer thickness and has the sulfur layers vibrating in counterphase in the direction normal to the layer plane while the Mo layer remains stationary. The change in the potential associated with this lattice vibration has previously been demonstrated to result in a large deformation potential in bulk MoS$_2$ [@Mooser:Mobility]. Electron-phonon coupling {#sec:III} ======================== In this section, we present the electron-phonon coupling in single-layer MoS$_2$ obtained with a first-principles based DFT approach. The method is based on a supercell approach analogous to that used for the calculation of the phonon dispersion and is outlined in App. \[app:e-ph\]. The calculated electron-phonon couplings are discussed in the context of deformation potential couplings and the Fr[ö]{}hlich interaction well-known from the semiconductor literature. Within the adiabatic approximation for the electron-phonon interaction, the coupling strength for the phonon mode with wave vector ${\mathbf{q}}$ and branch index $\lambda$ is given by $$\label{eq:g} g_{{\mathbf{k}}{\mathbf{q}}}^{\lambda} = \sqrt{\frac{\hbar}{2 MN\omega_{{\mathbf{q}}\lambda}}} M_{{\mathbf{k}}{\mathbf{q}}}^\lambda ,$$ where $\omega_{{\mathbf{q}}\lambda}$ is the phonon frequency, $M$ is an appropriately defined effective mass, $N$ is the number of unit cells in the crystal, and $$\label{eq:M_elph} M_{{\mathbf{k}}{\mathbf{q}}}^\lambda = {\langle\mathbf{k+q} \rvert} \delta V_{{\mathbf{q}}\lambda}({\mathbf{r}}) {\lvert {\mathbf{k}}\rangle}$$ is the coupling matrix element where ${\mathbf{k}}$ is the wave vector of the carrier being scattered and $\delta V_{{\mathbf{q}}\lambda}$ is the change in the effective potential per unit displacement along the vibrational normal mode. Due to the valley degeneracy in the conduction band, both intra and intervalley phonon scattering of the carriers in the $K,K'$-valleys need to be considered. Here, the coupling constants for these scattering processes are approximated by the electron-phonon coupling at the bottom of the valleys, i.e. with ${\mathbf{k}}=\mathbf{K},\mathbf{K}'$. With this approach, the intra and intervalley scattering for the $K,K'$-valleys are thus assumed independent on the wave vector of the carriers. In the following two sections, the calculated electron-phonon couplings are presented [@footnote1]. The different deformation potential couplings are discussed and the functional form of the Fr[ö]{}hlich interaction in 2D materials is established. Piezoelectric coupling to the acoustic phonons which occurs in materials without inversion symmetry, is most important at low temperatures [@Sarma:Hetero2] and will not be considered here. As deformation potentials are often extracted as empirical parameters from experimental mobilities, it can be difficult to disentangle contributions from different phonons. Here, the first-principles calculation of the electron-phonon coupling allows for a detailed analysis of the couplings and to assign deformation potentials to the individual intra and intervalley phonons. Deformation potentials ---------------------- The deformation potential interaction describes how carriers interact with the local changes in the crystal potential associated with a lattice vibration. Within the deformation potential approximation, the electron-phonon coupling is expressed as [@Ferry] $$\label{eq:g_longwavelength} g_{{\mathbf{q}}\lambda} = \sqrt{\frac{\hbar}{2A\rho\omega_{{\mathbf{q}}\lambda}}} M_{{\mathbf{q}}\lambda} ,$$ where $A$ is the area of the sample, $\rho$ is the atomic mass density per area and $M_{{\mathbf{q}}\lambda}$ is the coupling matrix element for a given valley which is assumed independent on the ${\mathbf{k}}$-vector of the carriers. This expression follows from the general definition of the electron-phonon coupling in Eq. \[eq:g\] by setting the effective mass $M$ equal to the sum of the atomic masses in the unit cell. With this convention for the effective mass, $MN=A\rho$, and the expression in Eq.  is obtained. For scattering on acoustic phonons, the coupling matrix element is linear in $q$ in the long-wavelength limit, $$\label{eq:M_acoustic} M_{{\mathbf{q}}\lambda} = \Xi_\lambda q ,$$ where $\Xi_\lambda$ is the acoustic deformation potential. In the case of optical phonon scattering, both coupling to zero- and first-order in $q$ must be considered. The interaction via the constant zero-order optical deformation potential $D_\lambda^0$ is given by $$\label{eq:M_optical} M_{{\mathbf{q}}\lambda} = D_\lambda^0 .$$ The coupling via the zero-order deformation potential is dictated by selection rules for the coupling matrix elements. Therefore, only symmetry-allowed phonons can couple to the carriers via the zero-order interaction. The coupling via the first-order interaction is given by Eq.  with the acoustic deformation potential replaced by the first-order optical deformation potential $D_\lambda^1$. Both the zero- and first-order deformation potential coupling can give rise to intra and intervalley scattering. ![(Color online) Deformation potential couplings in single-layer MoS$_2$. Only phonon modes with significant coupling strength are shown. The contour plots show the absolute value of the coupling matrix elements ${\lvert M_{{\mathbf{q}}\lambda} \rvert}$ in Eq.  with ${\mathbf{k}}=\mathbf{K}$ and ${\mathbf{k}}' = \mathbf{K} + {\mathbf{q}}$ as a function of the two-dimensional phonon wave vector ${\mathbf{q}}$ (note the different color scales in the plots). The upper four plots correspond to intravalley scattering and the two bottom plots to intervalley scattering on phonons with wave vector ${\mathbf{q}}= 2\mathbf{K}$. The two zero-order deformation potential couplings to the intravalley homopolar and intervalley LO phonon play an important role for the room-temperature mobility.[]{data-label="fig:couplings"}](intravalley){width="25.00000%"} ![(Color online) Deformation potential couplings in single-layer MoS$_2$. Only phonon modes with significant coupling strength are shown. The contour plots show the absolute value of the coupling matrix elements ${\lvert M_{{\mathbf{q}}\lambda} \rvert}$ in Eq.  with ${\mathbf{k}}=\mathbf{K}$ and ${\mathbf{k}}' = \mathbf{K} + {\mathbf{q}}$ as a function of the two-dimensional phonon wave vector ${\mathbf{q}}$ (note the different color scales in the plots). The upper four plots correspond to intravalley scattering and the two bottom plots to intervalley scattering on phonons with wave vector ${\mathbf{q}}= 2\mathbf{K}$. The two zero-order deformation potential couplings to the intravalley homopolar and intervalley LO phonon play an important role for the room-temperature mobility.[]{data-label="fig:couplings"}](g_TA "fig:"){width="49.00000%"} ![(Color online) Deformation potential couplings in single-layer MoS$_2$. Only phonon modes with significant coupling strength are shown. The contour plots show the absolute value of the coupling matrix elements ${\lvert M_{{\mathbf{q}}\lambda} \rvert}$ in Eq.  with ${\mathbf{k}}=\mathbf{K}$ and ${\mathbf{k}}' = \mathbf{K} + {\mathbf{q}}$ as a function of the two-dimensional phonon wave vector ${\mathbf{q}}$ (note the different color scales in the plots). The upper four plots correspond to intravalley scattering and the two bottom plots to intervalley scattering on phonons with wave vector ${\mathbf{q}}= 2\mathbf{K}$. The two zero-order deformation potential couplings to the intravalley homopolar and intervalley LO phonon play an important role for the room-temperature mobility.[]{data-label="fig:couplings"}](g_LA "fig:"){width="49.00000%"} ![(Color online) Deformation potential couplings in single-layer MoS$_2$. Only phonon modes with significant coupling strength are shown. The contour plots show the absolute value of the coupling matrix elements ${\lvert M_{{\mathbf{q}}\lambda} \rvert}$ in Eq.  with ${\mathbf{k}}=\mathbf{K}$ and ${\mathbf{k}}' = \mathbf{K} + {\mathbf{q}}$ as a function of the two-dimensional phonon wave vector ${\mathbf{q}}$ (note the different color scales in the plots). The upper four plots correspond to intravalley scattering and the two bottom plots to intervalley scattering on phonons with wave vector ${\mathbf{q}}= 2\mathbf{K}$. The two zero-order deformation potential couplings to the intravalley homopolar and intervalley LO phonon play an important role for the room-temperature mobility.[]{data-label="fig:couplings"}](g_TO "fig:"){width="49.00000%"} ![(Color online) Deformation potential couplings in single-layer MoS$_2$. Only phonon modes with significant coupling strength are shown. The contour plots show the absolute value of the coupling matrix elements ${\lvert M_{{\mathbf{q}}\lambda} \rvert}$ in Eq.  with ${\mathbf{k}}=\mathbf{K}$ and ${\mathbf{k}}' = \mathbf{K} + {\mathbf{q}}$ as a function of the two-dimensional phonon wave vector ${\mathbf{q}}$ (note the different color scales in the plots). The upper four plots correspond to intravalley scattering and the two bottom plots to intervalley scattering on phonons with wave vector ${\mathbf{q}}= 2\mathbf{K}$. The two zero-order deformation potential couplings to the intravalley homopolar and intervalley LO phonon play an important role for the room-temperature mobility.[]{data-label="fig:couplings"}](g_HP "fig:"){width="49.00000%"} ![(Color online) Deformation potential couplings in single-layer MoS$_2$. Only phonon modes with significant coupling strength are shown. The contour plots show the absolute value of the coupling matrix elements ${\lvert M_{{\mathbf{q}}\lambda} \rvert}$ in Eq.  with ${\mathbf{k}}=\mathbf{K}$ and ${\mathbf{k}}' = \mathbf{K} + {\mathbf{q}}$ as a function of the two-dimensional phonon wave vector ${\mathbf{q}}$ (note the different color scales in the plots). The upper four plots correspond to intravalley scattering and the two bottom plots to intervalley scattering on phonons with wave vector ${\mathbf{q}}= 2\mathbf{K}$. The two zero-order deformation potential couplings to the intravalley homopolar and intervalley LO phonon play an important role for the room-temperature mobility.[]{data-label="fig:couplings"}](intervalley){width="25.00000%"} ![(Color online) Deformation potential couplings in single-layer MoS$_2$. Only phonon modes with significant coupling strength are shown. The contour plots show the absolute value of the coupling matrix elements ${\lvert M_{{\mathbf{q}}\lambda} \rvert}$ in Eq.  with ${\mathbf{k}}=\mathbf{K}$ and ${\mathbf{k}}' = \mathbf{K} + {\mathbf{q}}$ as a function of the two-dimensional phonon wave vector ${\mathbf{q}}$ (note the different color scales in the plots). The upper four plots correspond to intravalley scattering and the two bottom plots to intervalley scattering on phonons with wave vector ${\mathbf{q}}= 2\mathbf{K}$. The two zero-order deformation potential couplings to the intravalley homopolar and intervalley LO phonon play an important role for the room-temperature mobility.[]{data-label="fig:couplings"}](g_TA_K "fig:"){width="49.00000%"} ![(Color online) Deformation potential couplings in single-layer MoS$_2$. Only phonon modes with significant coupling strength are shown. The contour plots show the absolute value of the coupling matrix elements ${\lvert M_{{\mathbf{q}}\lambda} \rvert}$ in Eq.  with ${\mathbf{k}}=\mathbf{K}$ and ${\mathbf{k}}' = \mathbf{K} + {\mathbf{q}}$ as a function of the two-dimensional phonon wave vector ${\mathbf{q}}$ (note the different color scales in the plots). The upper four plots correspond to intravalley scattering and the two bottom plots to intervalley scattering on phonons with wave vector ${\mathbf{q}}= 2\mathbf{K}$. The two zero-order deformation potential couplings to the intravalley homopolar and intervalley LO phonon play an important role for the room-temperature mobility.[]{data-label="fig:couplings"}](g_LO_K "fig:"){width="49.00000%"} The absolute value of the calculated coupling matrix elements ${\lvert M_{{\mathbf{q}}\lambda} \rvert}$ are shown in Fig. \[fig:couplings\] for phonons which couple to the carriers via deformation potential interactions. They are plotted in the range of phonon wave vectors relevant for intra and intervalley scattering processes [@footnote2], and only phonon modes with significant coupling strengths have been included. Although the coupling matrix elements are shown only for ${\mathbf{k}}= \mathbf{K}$, the matrix elements for ${\mathbf{k}}= \mathbf{K'}$ are related through time-reversal symmetry as $M_{{\mathbf{q}}\lambda}^\mathbf{K}=M_{-{\mathbf{q}}\lambda}^\mathbf{K'}$. The three-fold rotational symmetry of the coupling matrix elements in ${\mathbf{q}}$-space stems from the symmetry of the conduction band in the vicinity of the $K,K'$-valleys (see Fig. \[fig:mos2\]). Since the symmetry of the matrix elements has not been imposed by hand, slight deviations from the three-fold rotational symmetry can be observed. The acoustic deformation potential couplings for the TA and LA modes are shown in the two top plots of Fig. \[fig:couplings\]. Due to the inclusion of Umklapp processes here, the coupling to the TA mode does not vanish as is most often assumed [@Madelung]. Only along high-symmetry directions in the Brillouin zone is this the case. This results in a highly anisotropic coupling to the TA mode. On the other hand, the deformation potential coupling for the LA mode is perfectly isotropic in the long-wavelength limit but also becomes anisotropic at shorter wavelengths. In agreement with Eq. , both the TA and LA coupling matrix elements are linear in $q$ in the long-wavelength limit. The next two plots show the couplings to the intravalley polar TO and homopolar optical modes. While the interaction with the polar TO phonon corresponds to a first-order optical deformation potential, the interaction with the homopolar mode acquires a finite value of $\sim 4.8$ eV/[Å]{} in the $\Gamma$-point and corresponds to a strong zero-order deformation potential coupling. The large deformation potential coupling to the homopolar mode stems from its characteristic lattice vibration polarized in the direction perpendicular to the layer corresponding to a change in the layer thickness. The associated change in the effective potential from the counterphase oscillation of the two negatively charged sulfur layers results in a significant change of the potential towards the center of the MoS$_2$ layer. As the electronic Bloch functions have significant weight here, the homopolar lattice vibration gives rise to a significant shift of the $K,K'$-valley states. The large deformation potential associated with the coupling to the homopolar mode is also present in bulk MoS$_2$ [@Mooser:Mobility]. The two bottom plots in Fig. \[fig:couplings\] show the couplings for the intervalley TA and polar LO phonons. Due to the nearly constant frequency of the intervalley acoustic phonons, their couplings are classified as optical deformation potentials in the following. Both the shown coupling to the intervalley TA phonon and the coupling to the intervalley LA phonon result in first-order deformation potentials. The intervalley coupling for the polar LO phonon gives rise to a zero-order deformation potential which results from an identical in-plane motion of the two sulfur layers in the lattice vibration of the intervalley phonon. In general, the deformation potential approximation, i.e. the assumption of isotropic and constant/linear coupling matrix elements, in Eqs.  and  is seen to hold in the long-wavelength limit only. At shorter wavelengths, the first-principles couplings become anisotropic and have a more complicated $q$-dependence. When determined experimentally from e.g. the temperature dependence of the mobility [@Sarma:Hetero1], the deformation potentials in Eqs.  and  implicitly account for the more complex ${\mathbf{q}}$-dependence of the *true* coupling matrix element. In Section \[sec:V\], the theoretical deformation potentials for single-layer MoS$_2$ are determined from the first-principles electron-phonon couplings. This is done by fitting their associated scattering rates to the scattering rates obtained with the first-principles coupling matrix elements. The resulting the deformation potentials, can be used in practical transport calculations based on e.g. the Boltzmann equation or Monte Carlo simulations. It should be noted that similar routes for first-principles calculations of deformation potentials have been given in the literature [@Sjakste:GaAsInter; @Sjakste:Silicon; @Kim:ElPhGraphene]. Fr[ö]{}hlich interaction ------------------------ The lattice vibration of the polar LO phonon gives rise to a macroscopic electric field that couples to the charge carriers. For bulk three-dimensional systems, the coupling to the field is given by the Fr[ö]{}hlich interaction which diverges as $1/q$ in the long-wavelength limit [@Madelung], $$\label{eq:frohlich} g_\text{LO}(q) = \frac{1}{q} \sqrt{\frac{e^2 \hbar\omega_\text{LO}}{2\epsilon_0 V }} \left( \frac{1}{\varepsilon^\infty} - \frac{1}{\varepsilon^0} \right)^{1/2} ,$$ where $\epsilon_0$ is the vacuum permittivity, $V$ is the volume of the sample , and $\varepsilon^\infty$ and $\varepsilon^0$ are the high-frequency optical and static dielectric constant, respectively. In atomically thin materials, the two-dimensional nature of both the LO phonon and the charge carriers leads to a qualitatively different $q$-dependence of the Fr[ö]{}hlich interaction. The situation is similar to 2D semiconductor heterostructures where the Fr[ö]{}hlich interaction has been studied using dielectric continuum models and microscopic based approaches [@Ando:ElphHetero; @Lugli:Microscopic]. From the microscopic considerations presented in App. \[app:frohlich\], we derive the following functional form for the Fr[ö]{}hlich interaction to the polar LO phonon in 2D materials, $$\label{eq:g_Fr} g_\text{LO}(q) = g_\text{Fr} \times \text{erfc}(q \sigma / 2) .$$ Here, the coupling constant $g_\text{Fr}$ is the equivalent of the square root factors in Eq. , $\sigma$ is the effective width of the electronic Bloch states and $\text{erfc}$ is the complementary error function. Figure \[fig:frohlich\] shows the first-principles coupling $g_\text{LO}$ to the polar LO phonon in single-layer MoS$_2$ (dots) as a function of the phonon wave vector along the $\Gamma$-$K$ path. The dashed line shows a fit of the coupling in Eq.  to the long-wavelength limit of the calculated coupling. With a coupling constant of $g_\text{Fr} = 98$ meV and an effective width of $\sigma=4.41$ [Å]{}, the analytic form gives a perfect description of the calculated coupling in the long-wavelength limit. For shorter wavelengths, the zero-order deformation potential coupling to the intervalley LO phonon from Fig. \[fig:couplings\] becomes dominant, and a deviation away from the behavior in Eq.  is observed. ![(Color online) Fr[ö]{}hlich interaction for the polar optical LO mode. The dashed line shows the analytical coupling in Eq.  with the coupling constant $g_\text{Fr} = 98$ meV and the effective layer thickness $\sigma = 4.41$ [Å]{} fitted to the long-wavelength limit of the calculated coupling (dots).[]{data-label="fig:frohlich"}](frohlich){width="0.8\linewidth"} Boltzmann equation {#sec:IV} ================== In the regime of diffuse transport dominated by phonon scattering, the mobility can be obtained with semiclassical Boltzmann transport theory. In the absence of spatial gradients the Boltzmann equation for the out-of-equilibrium distribution function $f_{\mathbf{k}}$ of the charge carriers reads [@SmithJensen] $$\label{eq:boltzmann} \frac{\partial f_{\mathbf{k}}}{\partial t} + \dot{{\mathbf{k}}} \cdot \nabla_{\mathbf{k}}f_{\mathbf{k}}= \left. \frac{\partial f_{\mathbf{k}}}{\partial t} \right\vert_{\text{coll}} ,$$ where the time evolution of the crystal momentum is governed by $\hbar\dot{{\mathbf{k}}} = q \mathbf{E}$ and $q$ is the charge of the carriers. In the case of a time-independent uniform electric field, the steady-state version of the linearized Boltzmann equation takes the form $$\begin{aligned} \label{eq:linearized} \frac{q}{\hbar} \mathbf{E} \cdot \nabla_{{\mathbf{k}}} f_{\mathbf{k}}^0 & = - \frac{q}{k_\text{B}T} \mathbf{E} \cdot \mathbf{v}_{\mathbf{k}}f_{\mathbf{k}}^0(1 - f_{\mathbf{k}}^0) \nonumber \\ & = \left. \frac{\partial f_{\mathbf{k}}}{\partial t} \right\vert_{\text{coll}} ,\end{aligned}$$ where $f_{\mathbf{k}}^0=f^0(\varepsilon_{\mathbf{k}})$ is the equilibrium Fermi-Dirac distribution function, $\partial f_{\mathbf{k}}^0/\partial \varepsilon_{\mathbf{k}}= - f_{\mathbf{k}}^0(1 - f_{\mathbf{k}}^0)/k_\text{B}T$, and $\mathbf{v}_{\mathbf{k}}= \nabla_{{\mathbf{k}}} \varepsilon_{\mathbf{k}}/ \hbar$ is the band velocity. In linear response, the out-of-equilibrium distribution function can be written as the equilibrium function plus a small deviation $\delta f$ away from its equilibrium value [@SmithJensen], i.e. $$\label{eq:deviation} \delta f_{\mathbf{k}}= f_{\mathbf{k}}- f_{\mathbf{k}}^0 = f_{\mathbf{k}}^0 (1 - f_{\mathbf{k}}^0 ) \psi_{\mathbf{k}}.$$ Here, the last equality has defined the deviation function $\psi_{\mathbf{k}}$. Furthermore, following the spirit of the iterative method of Rode [@Rode], the angular dependence of the deviation function is separated out as $$\label{eq:Rode} \psi_{\mathbf{k}}= \phi_k \cos\theta_{\mathbf{k}},$$ where $\theta_{\mathbf{k}}$ is the angle between the ${\mathbf{k}}$-vector and the force exerted on the carriers by the applied field $\mathbf{E}$. In general, $\phi_{\mathbf{k}}$ is a function of the ${\mathbf{k}}$-vector and not only its magnitude $k$. However, for isotropic bands as considered here, the angular dependence of the deviation function is entirely accounted for by the cosine factor in Eq.  [@Ferry]. Considering the phonon scattering, the collision integral describes how accelerated carriers are driven back towards their equilibrium distribution by scattering on acoustic and optical phonons. In the high-temperature regime of interest here, the collision integral can be split up in two contributions. The first, accounting for the quasielastic scattering on acoustic phonons, can be expressed in the form of a relaxation time $\tau_{\text{el}}$. The second, describing the inelastic scattering on optical phonons, will in general be an integral operator $I_{\text{inel}}$. The collision integral can thus be written $$\label{eq:collision} \left. \frac{\partial f_{\mathbf{k}}}{\partial t} \right\vert_{\text{coll}} = - \frac{\delta f_{\mathbf{k}}} {\tau_{\mathbf{k}}^\text{el}} + I_{\text{inel}} \left[ \psi_{\mathbf{k}}\right] ,$$ where the explicit forms of the relaxation time and the integral operator will be considered in detail below. With the above form of the collision integral, the Boltzmann equation can be solved by iterating the equation $$\begin{aligned} \label{eq:Rode_solution} \phi_k = & \left[ \frac{eE v_{\mathbf{k}}}{k_\text{B}T} + I'_{\text{inel}} \left[ \phi_k \right] \right] \times \left( \frac{1}{\tau_{\mathbf{k}}^{\text{el}}} \right)^{-1} \end{aligned}$$ for the deviation function $\phi_k$. Here, the angular dependence $\cos\theta_{\mathbf{k}}$ and a $f_{\mathbf{k}}^0 (1 - f_{\mathbf{k}}^0 )$ factor have been divided out and the new integral operator $I'_{\text{inel}}$ is defined by $$\label{eq:I'} I_{\text{inel}} = f_{\mathbf{k}}^0 (1 - f_{\mathbf{k}}^0 ) \cos\theta_{\mathbf{k}}I'_{\text{inel}} .$$ From the solution of the Boltzmann equation the current and drift mobility of the carriers can be obtained. Taking the electric field to be oriented along the $x$-direction, the current density is given by $$\label{eq:current_density} j_x \equiv \sigma_{xx} E_x = -\frac{4 e}{A} \sum_{\mathbf{k}}v_{x} \delta f_{\mathbf{k}},$$ where $\sigma_{xx}$ is the conductivity, $A$ is the area of the sample, and $v_{i} = \hbar k_i/m^*$ is the band velocity for parabolic bands with $i= x, y, z$. From the definition $\mu_{xx} = \sigma_{xx} / ne$ of the drift mobility, it follows that $$\label{eq:mobility} \mu_{xx} = \frac{e {\langle \tilde{\phi}_k\rangle}}{m^*}$$ where the modified deviation function $\tilde{\phi}_k$ having units of time is defined by $$\label{eq:phi_tilde} \tilde{\phi}_k = \frac{m^* k_\text{B}T}{e E_x \hbar k} \phi_k ,$$ and the energy-weighted average ${\langle \cdot\rangle}$ is defined by $$\label{eq:tau_averaged} {\langle A\rangle} = \frac{1}{n} \int \! d\varepsilon_{\mathbf{k}}\; \rho(\varepsilon_{\mathbf{k}}) \varepsilon_{\mathbf{k}}A_{\mathbf{k}}\left(- \frac{\partial f^0}{\partial \varepsilon_{\mathbf{k}}} \right) .$$ Here, $n$ is the two-dimensional carrier density. In the relaxation time approximation $\tilde{\phi}_k = \tau_k$ and the well-known Drude expression for the mobility $\mu = e{\langle \tau_k\rangle}/m^*$ is recovered. Phonon collision integral ------------------------- The phonon collision integral has been considered in great detail in the literature for two-dimensional electron gases in semiconductor heterostructure (see e.g. Ref. ). In general, these treatments consider scattering on three-dimensional or quasi two-dimensional phonons. In atomically thin materials the phonons are strictly two-dimensional which results in a slightly different treatment. In the following, a full account of the two-dimensional phonon collision integral is given for scattering on acoustic and optical phonons. With the distribution function written on the form in Eq. , the linearized collision integral for electron-phonon scattering takes the form [@SmithJensen] $$\begin{aligned} \label{eq:collision_linearized} \left. \frac{\partial f_{\mathbf{k}}}{\partial t} \right\vert_{\text{coll}} = -\frac{2\pi}{\hbar} \sum_{{\mathbf{q}}\lambda} \frac{\left\vert g_{{\mathbf{q}}\lambda} \right\vert^2} {\epsilon^2(q, T)} \times & \bigg[ f_{\mathbf{k}}^0 \left( 1 - f_{{\mathbf{k}}+ {\mathbf{q}}}^0 \right) N_{{\mathbf{q}}\lambda}^0 \left( \psi_{\mathbf{k}}- \psi_{{\mathbf{k}}+ {\mathbf{q}}} \right) \delta(\varepsilon_{{\mathbf{k}}+ {\mathbf{q}}} - \varepsilon_{\mathbf{k}}- \hbar\omega_{{\mathbf{q}}\lambda}) \nonumber \\ & + f_{\mathbf{k}}^0 \left( 1 - f_{{\mathbf{k}}- {\mathbf{q}}}^0 \right) \left(1 + N_{{\mathbf{q}}\lambda}^0 \right) \left( \psi_{\mathbf{k}}- \psi_{{\mathbf{k}}- {\mathbf{q}}} \right) \delta(\varepsilon_{{\mathbf{k}}- {\mathbf{q}}} - \varepsilon_{\mathbf{k}}+ \hbar\omega_{{\mathbf{q}}\lambda}) \bigg] \end{aligned}$$ where $N_{{\mathbf{q}}\lambda}^0=N^0(\hbar\omega_{{\mathbf{q}}\lambda})$ is the equilibrium distribution of the phonons given by the Bose-Einstein distribution function and $f_{{\mathbf{k}}\pm {\mathbf{q}}}^0 = f^0(\varepsilon_{{\mathbf{k}}} \pm \hbar\omega_{{\mathbf{q}}})$ is understood. The different terms inside the square brackets account for scattering out of ($\psi_{\mathbf{k}}$) and into ($\psi_{{\mathbf{k}}\pm{\mathbf{q}}}$) the state with wave vector ${\mathbf{k}}$ via absorption and emission of phonons. Screening of the electron-phonon interaction by the carriers themselves is accounted for by the static dielectric function $\epsilon$, which is both wave vector $q$ and temperature $T$ dependent. Due to the large effective mass of the conduction band, the charge carriers in single-layer MoS$_2$ are non-degenerate except at very high carrier concentrations. Semiclassical screening where the screening length is given by the inverse of the Debye-H[ü]{}ckel wave vector $k_\text{D} = e^2 n / 2k_\text{B}T$ therefore applies [@FerryGoodnick]. For the considered values of carrier densities and temperatures, this corresponds to a small fraction of the size of the Brillouin zone, i.e. $k_\text{D} \ll 2\pi / a$. As scattering on phonons in general involves larger wave vectors, screening by the carriers can to a good approximation be neglected here. The calculated mobilities therefore provide a lower limit for the proper screened mobility. ### Quasielastic scattering on acoustic phonons For quasielastic scattering on acoustic phonons at high temperatures, the energy of the acoustic phonon can be neglected in the collision integral in Eq. . As a consequence, the collision integral can be recast in the form of the following relaxation time [@Sarma:Hetero2] $$\label{eq:tau_elastic} \frac{1}{\tau_{\mathbf{k}}^{\text{el}}} = \sum_{{\mathbf{k}}'} \left( 1 - \cos{\theta_{{\mathbf{k}}{\mathbf{k}}'}} \right) P_{{\mathbf{k}}{\mathbf{k}}'}$$ where the summation variable has been changed to ${\mathbf{k}}'={\mathbf{k}}\pm {\mathbf{q}}$ and the transition matrix element for quasielastic scattering on acoustic phonons is given by $$\begin{aligned} \label{eq:P_acoustic} P_{{\mathbf{k}}{\mathbf{k}}'} = \frac{2\pi}{\hbar} \sum_{\mathbf{q}}\left\vert g_{{\mathbf{q}}\lambda} \right\vert^2 \bigg[ & N_{{\mathbf{q}}\lambda}^0 \delta(\varepsilon_{{\mathbf{k}}'} - \varepsilon_{\mathbf{k}}) \bigg. \nonumber \\ & + \bigg. \left(1 + N_{{\mathbf{q}}\lambda}^0 \right) \delta(\varepsilon_{{\mathbf{k}}'} - \varepsilon_{{\mathbf{k}}}) \bigg] .\end{aligned}$$ For isotropic scattering as assumed here in Eq. , the square of the electron-phonon coupling can be expressed as $$\label{eq:g2_elastic} \left\vert g_{{\mathbf{q}}\lambda} \right\vert^2 = \frac{\Xi_\lambda^2\hbar q}{2A\rho c_\lambda},$$ where the acoustic phonon frequency has been expressed in terms of the sound velocity $c_\lambda$. Except at very low temperatures $\hbar\omega_{\mathbf{q}}\ll k_\text{B}T$ implying that the equipartition approximation $N^0_{\mathbf{q}}\sim k_\text{B}T / \hbar\omega_{\mathbf{q}}\gg 1$ for the Bose-Einstein distribution applies. With the resulting ${\mathbf{q}}$-factors in the transition matrix element in Eq.  canceling, the $\cos{\theta_{{\mathbf{k}}{\mathbf{k}}'}}$ in the ${\mathbf{k}}'$-sum in Eq.  vanishes and the first term yields a factor density of states divided by the spin and valley degeneracies $\rho_0/g_sg_v$. The relaxation time for acoustic phonon scattering becomes $$\label{eq:tau_acoustic} \frac{1}{\tau_{\mathbf{k}}^\text{el}} = \frac{m^*\Xi^2k_\text{B}T}{\hbar^3\rho c_\lambda^2} .$$ The independence on the carrier energy and the $\tau^{-1} \sim T$ temperature dependence of the acoustic scattering rate are characteristic for charge carriers in two-dimensional heterostructures and layered materials [@Mooser:Mobility; @Sarma:Hetero2]. ### Inelastic scattering on dispersionless optical phonons For inelastic scattering on optical phonons the phonon energy can no longer be neglected and the collision integral in Eq.  must be considered in full detail. Under the reasonable assumption of dispersionless optical phonons $\omega_{{\mathbf{q}}\lambda}=\omega_\lambda$, the inelastic collision integral can be treated semianalytically. The overall procedure for the evaluation of the collision integral is given below, while the calculational details have been collected in App. \[app:collision\]. The resulting expressions for the collision integral apply to both intravalley and intervalley optical and intervalley acoustic phonons. In the following, the integral operator for inelastic scattering in Eq.  is split up in separate out- and in-scattering contributions, $$\label{eq:I_inel_out_in} {I'}_\text{inel}[\phi_k] = {I'}_\text{inel}^\text{out}[\phi_k] + {I'}_\text{inel}^\text{in}[\phi_k] ,$$ which include the terms in Eq.  involving $\psi_{\mathbf{k}}$ and $\psi_{{\mathbf{k}}\pm {\mathbf{q}}}$, respectively. With the contributions from different phonon branches $\lambda$ adding up, scattering on a single phonon with branch index $\lambda$ is considered in the following. From Eq. , the out-scattering part of the collision integral follows directly as $$\label{eq:I_out} {I'}_\text{inel}^\text{out}[\phi_k] = - \phi_k \sum_{{\mathbf{k}}'} P_{{\mathbf{k}}{\mathbf{k}}'} \frac{1 - f_{{\mathbf{k}}'}^0}{1 - f_{\mathbf{k}}^0} ,$$ where the transition matrix element for optical phonon scattering is given by $$\begin{aligned} \label{eq:P_optical} P_{{\mathbf{k}}{\mathbf{k}}'} = \frac{2\pi}{\hbar} \sum_{\mathbf{q}}\left\vert g_{{\mathbf{q}}\lambda} \right\vert^2 \times \bigg[ & N_\lambda^0 \delta(\varepsilon_{{\mathbf{k}}'} - \varepsilon_{\mathbf{k}}- \hbar\omega_\lambda) \bigg. \nonumber \\ & + \bigg. \left(1 + N_\lambda^0 \right) \delta(\varepsilon_{{\mathbf{k}}'} - \varepsilon_{{\mathbf{k}}} + \hbar\omega_\lambda) \bigg] .\end{aligned}$$ For the in-scattering part, the desired $\cos{\theta_{\mathbf{k}}}$ factor in Eq.  can be extracted from $\psi_{{\mathbf{k}}'}$ using the relation $\cos{\theta_{{\mathbf{k}}'}} = \cos{\theta_{\mathbf{k}}}\cos{\theta_{{\mathbf{k}}{\mathbf{k}}'}} - \sin{\theta_{\mathbf{k}}}\sin{\theta_{{\mathbf{k}}{\mathbf{k}}'}}$. Since the sine term vanishes from symmetry consideration, the in-scattering part of the inelastic collision integral reduces to $$\label{eq:I_in} {I'}_\text{inel}^\text{in}[\phi_k] = \sum_{{\mathbf{k}}'} \phi_{k'} \cos \theta_{{\mathbf{k}}{\mathbf{k}}'} P_{{\mathbf{k}}{\mathbf{k}}'} \frac{1 - f_{{\mathbf{k}}'}^0}{1 - f_{\mathbf{k}}^0} .$$ The evaluation of the ${\mathbf{k}}'$-sum in Eqs.  and  is outlined in App. \[app:collision\]. Here, the assumption of dispersionless optical phonons allows for an semianalytical treatment. For zero- and first-order coupling within the deformation potential approximation, the sum can be carried out analytically. The resulting expressions for the collision integral are given in Eqs. , , ,  and . In the case of the Fr[ö]{}hlich interaction, the angular part of the ${\mathbf{q}}$-integral must be done numerically. ### Optical deformation potential scattering rates In spite of the fact that the collision integral for inelastic scattering on optical phonons cannot be recast in the form of a (momentum) relaxation time [@footnote3], a scattering rate related to the inverse carrier lifetime can still be defined from the out-scattering part of the inelastic collision integral alone. The scattering rate so defined is given by $$\label{eq:tau_optical} \frac{1}{\tau_{\mathbf{k}}^{\text{inel}}} = \sum_{{\mathbf{k}}'} P_{{\mathbf{k}}{\mathbf{k}}'} \frac{1 - f_{{\mathbf{k}}'}^0}{1 - f_{\mathbf{k}}^0}$$ and corresponds to the imaginary part of the electronic self-energy in the Born approximation [@SmithJensen]. Below, the resulting scattering rates for zero-order and first-order deformation potential scattering are given for non-degenerate carriers, i.e. with the Fermi factors in Eq.  neglected. They follow straight-forwardly from the expressions for the out-scattering part of the collision integral derived in App. \[app:collision\]. It should be noted that the scattering rate for the zero-order deformation potential interaction given below in Eq. , in fact defines a proper momentum relaxation time because the in-scattering part of the collision integral vanishes in this case (see App. \[app:collision\]). For zero-order deformation potential scattering, the scattering rate is independent of the carrier energy and given by $$\begin{aligned} \label{eq:tau_zero_optical} \frac{1}{\tau_{\mathbf{k}}^\text{inel}} = \frac{m^* (D_\lambda^0)^2}{2\hbar^2\rho\omega_\lambda} \bigg[ 1 + e^{\hbar\omega_\lambda / k_\text{B}T} \Theta(\varepsilon_{{\mathbf{k}}} - \hbar\omega_{\lambda} ) \bigg] N_{\lambda}^0 .\end{aligned}$$ Here, $\Theta(x)$ is the Heavyside step function which assures that only electrons with sufficient energy can emit a phonon. The scattering rate for coupling via the first-order deformation potential is found to be $$\begin{aligned} \label{eq:tau_first_optical} \frac{1}{\tau_{\mathbf{k}}^\text{inel}} = \frac{m^{*2} (D_\lambda^1)^2}{\hbar^4\rho\omega_\lambda} \bigg[ & \left( 2\varepsilon_{\mathbf{k}}+ \hbar\omega_\lambda \right) + e^{\hbar\omega_\lambda / k_\text{B}T} \nonumber \\ & \times \Theta(\varepsilon_{{\mathbf{k}}} - \hbar\omega_{\lambda} ) \left( 2\varepsilon_{\mathbf{k}}- \hbar\omega_\lambda \right) \bigg] N_\lambda^0 .\end{aligned}$$ Due to the linear dependence on the carrier energy, zero-order scattering processes dominate first-order processes at low energies. Only under high-field conditions where the carriers are accelerated to high velocities will first-order scattering become significant. The expressions for the scattering rates in Eqs.  and  above apply to scattering on dispersionless intervalley acoustic phonons and intra/intervalley optical phonons. Except for a factor of $\sqrt{\varepsilon_{\mathbf{k}}\pm \hbar\omega_\lambda}$ originating from the density of states, the energy dependence of the scattering rates is identical to that of their three-dimensional analogs [@Ferry]. Results {#sec:V} ======= In the following, the scattering rate and phonon-limited mobility in single-layer MoS$_2$ are studied as a function of carrier energy, temperature $T$ and carrier density $n$ using the material parameters collected in Tab. \[tab:parameters\]. Here, the reported deformation potentials represent effective coupling parameters for the deformation potential approximation in Eqs.  and  (see below). Parameter Symbol Value --------------------------------- ----------------------------------- ------------------------------ Lattice constant $a$ 3.14 Å Ion mass density $\rho$ $3.1\times 10^{-7}$ g/cm$^2$ Effective electron mass $m^*$ 0.48 $m_e$ Transverse sound velocity $c_\text{TA}$ $4.2 \times 10^3$ m/s Longitudinal sound velocity $c_\text{LA}$ $6.7 \times 10^3$ m/s Acoustic deformation potentials TA $\Xi_\text{TA}$ $1.6$ eV LA $\Xi_\text{LA}$ $2.8$ eV Optical deformation potentials TA $D_{\mathbf{K},\text{TA}}^1$ $5.9$ eV LA $D_{\mathbf{K},\text{LA}}^1$ $3.9$ eV TO $D_{\mathbf{\Gamma},\text{TO}}^1$ $4.0$ eV TO $D_{\mathbf{K},\text{TO}}^1$ $1.9$ eV LO $D_{\mathbf{K},\text{LO}}^0$ $2.6 \times 10^8$ eV/cm Homopolar $D_{\mathbf{\Gamma},\text{HP}}^0$ $4.1 \times 10^8$ eV/cm Fr[ö]{}hlich interaction (LO) Effective layer thickness $\sigma$ 4.41 Å Coupling constant $g_\text{Fr}$ $98$ meV Optical phonon energies Polar LO $\hbar\omega_\text{LO}$ 48 meV Homopolar $\hbar\omega_\text{HP}$ 50 meV : Material parameters for single-layer MoS$_2$. Unless otherwise stated, the parameters have been calculated from first-principles as described in the Secs. II and III. The $\mathbf{\Gamma}/\mathbf{K}$-subscript on the optical deformation potentials, indicate couplings to the intra/intervalley phonons.[]{data-label="tab:parameters"} Scattering rates ---------------- With access to the first-principles electron-phonon couplings, the scattering rate taking into account the anisotropy and more complex $q$-dependence of the first-principles coupling matrix elements can be evaluated. They are obtained using the expression for the scattering rate in Eq.  with the respective transition matrix elements for acoustic and optical phonon scattering given in Eqs.  and  [@footnote4]. In order to account for the different coupling matrix elements in the $K,K'$-valleys, the scattering rates have been averaged over different high-symmetry directions of the carrier wave vector ${\mathbf{k}}$. The resulting scattering rates are shown in Fig. \[fig:scatteringrates\] as a function of the carrier energy for non-degenerate carriers at $T=300$ K. The different lines show the contributions to the total scattering rate from the various electron-phonon couplings to the intra and intervalley phonons which have been grouped according to their coupling type, i.e. acoustic deformation potentials (ADPs), zero/first-order optical deformation potentials (ODPs) and the Fr[ö]{}hlich interaction. The acoustic deformation potential scattering includes the quasielastic intravalley scattering on the TA and LA phonons with linear dispersions. Scattering on intervalley acoustic phonons is considered as optical deformation potential scattering. Both the total scattering rate and the contributions from the different coupling types have been obtained using Matthiessen’s rule by summing the scattering rates from the individual phonons, i.e. $$\label{eq:matthiessen} \tau_\text{tot}^{-1} = \sum_\lambda \tau_\lambda^{-1} .$$ ![(Color online) Scattering rates for the different electron-phonon couplings as a function of carrier energy at $T=300$ K. The scattering rates have been calculated from the first-principles electron-phonon couplings in Figs. \[fig:couplings\] and \[fig:frohlich\]. The (black) dashed line shows the scattering rate obtained using the fitted deformation potential parameters defined in Eqs.  and . The kinks in the curves for the optical scattering rates mark the onset of optical phonon emissions.[]{data-label="fig:scatteringrates"}](scattering_rates_T_300){width="0.75\linewidth"} For carrier energies below the optical phonon frequencies, the total scattering rate is dominated by acoustic deformation potential scattering. At higher energies zero-order deformation potential scattering and polar optical scattering via the Fr[ö]{}hlich interaction become dominant. Due to the linear dependence on the carrier energy, the first-order deformation potential scattering on the intervalley acoustic phonons and the optical phonons is in general only of minor importance for the low-field mobility. This is also the case here, where it is an order of magnitude smaller than the other scattering rates for almost the entire plotted energy range. The jumps in the curves for the optical scattering rates at the optical phonon energies $\hbar\omega_\lambda$ are associated with the threshold for optical phonon emission where the carriers have sufficient energy to emit optical phonons Deformation potentials ---------------------- In this section, we determine the deformation potential parameters in single-layer MoS$_2$. They can be used in the following study of the low-field mobility within the Boltzmann approach outlined in Section \[sec:IV\]. The energy dependence of the first-principles based scattering rates in Fig. \[fig:scatteringrates\] to a high degree resembles that of the analytic expressions for the deformation potential scattering rates in Eqs. ,  and . For example, the acoustic and zero-order deformation potential scattering rates are almost constant in the plotted energy range. The first-principles electron-phonon couplings can therefore to a good approximation be described by the simpler isotropic deformation potentials in Eqs.  and . The deformation potentials are obtained by fitting the associated scattering rates for each of the intra and intervalley phonons separately to the first-principles scattering rates. The resulting deformation potential values are summarized in Tab. \[tab:parameters\]. In analogy with deformation potentials extracted from experimental mobilities, the theoretical deformation potentials represent effective coupling parameters that implicitly account for the anisotropy and the full $q$-dependence of the first-principles electron-phonon couplings. However, as momentum and energy conservation limit phonon scattering to involve phonons in the vicinity of the $\Gamma/K$-point [@footnote2], the fitting procedure yields deformation potentials close to the direction averaged ${\mathbf{q}}\rightarrow \mathbf{\Gamma}$/$\mathbf{K}$ limiting behavior of the first-principles coupling matrix elements. For the zero-order deformation potentials, the sampling of the coupling matrix elements away from the $\Gamma$-point in the ${\mathbf{k}}'$ sum in Eq. , leads to a deformation potentials which is slightly smaller than the $\Gamma$-point value of the coupling matrix element. The total scattering rate resulting from the fitted deformation potentials is shown in Fig. \[fig:scatteringrates\] to give an almost excellent description of the first-principles scattering rate. With the mobility in the relaxation time approximation given by the energy-weighted average of the relaxation time (see Eq. ), the difference between the associated mobilities will be negligible. Hence, the deformation potentials provide well-founded electron-phonon coupling parameters for low-field studies of the mobility. Mobility -------- The high density of states in the $K,K'$-valleys of the conduction band in general results in non-degenerate carrier distributions in single-layer MoS$_2$. This is illustrated in the left plot of Fig. \[fig:mu\_vs\_n\] which shows the carrier density versus the position of the Fermi level for different temperatures. At room temperature, carrier densities in excess of $\sim 8 \times 10^{12}$ cm$^{-2}$ are needed to introduce the Fermi level into the conduction band and probe the Fermi-Dirac statistics of the carriers. Thus, only at the highest reported carrier densities of $n \sim 10^{13}$ cm$^{-2}$ [@Geim:2D], are the carriers degenerate at room temperature. For the lowest temperature $T=100$ K, the transition to degenerate carriers occurs at a carrier density of $\sim 2\times10^{12}$ cm$^{-2}$. The transition from non-degenerate to degenerate distributions is also illustrated by the discrepancy between the full and dashed lines which shows the Fermi level obtained with Boltzmann statistics. ![(Color online) Left: Carrier density as a function of the Fermi energy for different temperatures. The Fermi energy is measured relative to the conduction band edge $E_c$ Right: Mobility as a function of carrier density for the same temperatures. In both plots, the (black) dashed lines show the results obtained with Boltzmann statistics and the relaxation time approximation (only right plot).[]{data-label="fig:mu_vs_n"}](fermilevel){width="0.98\linewidth"} ![(Color online) Left: Carrier density as a function of the Fermi energy for different temperatures. The Fermi energy is measured relative to the conduction band edge $E_c$ Right: Mobility as a function of carrier density for the same temperatures. In both plots, the (black) dashed lines show the results obtained with Boltzmann statistics and the relaxation time approximation (only right plot).[]{data-label="fig:mu_vs_n"}](mobility_vs_n_T){width="0.98\linewidth"} The right plot of Fig. \[fig:mu\_vs\_n\] shows the phonon-limited drift mobility calculated with the full collision integral as a function of carrier density for the same set of temperatures. The dashed lines represent the results obtained with Boltzmann statistics and the relaxation time approximation using the expressions in Section \[sec:IV\] for optical phonon scattering and Matthiessen’s rule for the total relaxation time. The strong drop in the mobility from $\sim 2450$ cm$^2$ V$^{-1}$ s$^{-1}$ at $T = 100$ K to $\sim 400$ cm$^2$ V$^{-1}$ s$^{-1}$ at $T = 300$ K, is a consequence of the increased phonon scattering at higher temperatures due to larger phonon population of, in particular, optical phonons. The relatively low intrinsic room-temperature mobility of single-layer MoS$_2$ can be attributed to both the significant phonon scattering and the large effective mass of $0.48$ $m_e$ in the conduction band. While the mobility decreases strongly with increasing temperature, it is relatively independent on the carrier density. The weak density dependence of the mobility originates from the energy-weighted average in Eq.  where the derivative of the Fermi-Dirac distribution changes slowly as a function of carrier density for non-degenerate carriers. As the Fermi energy is introduced into the band with increasing carrier density, the derivative of the Fermi-Dirac distribution to a larger extent probes the scattering rate at higher energies leading to a decrease of the mobility. This effect is most prominent at $T=100$ K where the level of degeneracy is larger compared to higher temperatures. Surprisingly, the relaxation time approximation is seen to work extremely well. The deviation from the full treatment at high carrier densities stems from the assumption of non-degenerate carriers and not the relaxation time approximation. The reason for the good performance of the relaxation time approximation shall be found in the in-scattering part of the collision integral in the full treatment. As the in-scattering part of the collision integral vanishes for zero-order deformation potential coupling and is small compared to the out-scattering part otherwise, the full collision integral does not differ significantly from the corresponding scattering rate as defined in Eq.  of Section \[sec:IV\], i.e. with the in-scattering part neglected. The good performance of the relaxation time approximation therefore seems to be of general validity for the phonon collision integral even in the presence inelastic scattering on optical phonons. Finally, we study the temperature dependence of the mobility in more detail. In general, room temperature mobilities are to a large extent dominated by optical phonon scattering. This is manifested in the temperature dependence of the mobility which follows a $\mu \sim T^{-\gamma}$ law where the exponent $\gamma$ depends on the dominating scattering mechanism. For acoustic phonon scattering above the Bloch-Gr[ü]{}neisen temperature, the temperature dependence of the scattering rate in Eq.  results in $\gamma=1$. At higher temperatures where optical phonon scattering starts to dominate, the mobility acquires a stronger temperature dependence with $\gamma > 1$. In this regime, the exponent depends on the optical phonon frequencies and the electron-phonon coupling strength. In an early study of the in-plane mobility of bulk MoS$_2$ [@Mooser:Mobility], the measured room-temperature exponent of $\gamma\sim 2.6$ was found to be consistent with scattering on the homopolar mode via a zero-order deformation potential. ![(Color online) Mobility vs temperature. For comparison, the mobility in the presence of only acoustic deformation potential scattering with the $\mu\sim T^{-1}$ temperature dependence is shown. The (gray) shaded area shows the variation in the mobility associated with a $10\%$ uncertainty in the calculated deformation potentials.[]{data-label="fig:mu_vs_T"}](mobility_vs_T_quench_n_11){width="0.75\linewidth"} In Fig. \[fig:mu\_vs\_T\], we show the temperature dependence of the mobility at $n=10^{11}$ cm$^{-2}$ calculated with the full collision integral. For comparison, the mobility limited by acoustic phonon scattering with the $\mu \sim T^{-1}$ temperature dependence is also included. To illustrate the effect of an uncertainty in the calculated deformation potentials, the shaded area shows the variation of the mobility with a change in the deformation potentials of $\pm 10 \%$. The calculated room-temperature mobility of $\sim 410$ cm$^2$ V$^{-1}$ s$^{-1}$ is in fair agreement with the recently reported experimental value of $\sim 200$ cm$^2$ V$^{-1}$ s$^{-1}$ in the top-gated sample of Ref.  where additional scattering mechanisms as e.g. impurity and surface-optical phonon scattering must be expected to exist. Over the plotted temperature range, the mobility undergoes a transition from being dominated by acoustic phonon scattering at $T=100$ K to being dominated by optical phonon scattering at higher temperatures with a characteristic exponent of $\gamma > 1$. At room temperature, the mobility follows a temperature dependence with $\gamma=1.69$. The increase in the exponent is almost exclusively due to the optical zero-order deformation potential couplings and the Fr[ö]{}hlich interaction while the first-order deformation potential couplings contribute only marginally. The exponent found here is considerably lower than the above-mentioned exponent of $\gamma\sim 2.6$ for bulk MoS$_2$, indicating that the electron-phonon coupling in bulk and single-layer MoS$_2$ differ. Indeed, the transition from an indirect band gap in bulk MoS$_2$ to a direct band gap in single-layer MoS$_2$ which shifts the bottom of the conduction band from the valley located along the $\Gamma$-K path to the $K,K'$-valleys, could result in a change in the electron-phonon coupling. In top-gated samples as the one studied in Ref. , the sandwiched structure with the MoS$_2$ layer located between substrate and gate dielectric, is likely to result in a quenching of the characteristic homopolar mode which is polarized in the direction normal to the layer. To address the consequence of such a quenching, the mobility in the absence of the zero-order deformation potential originating from the coupling to homopolar mode is also shown in Fig. \[fig:mu\_vs\_T\]. Here, the curve for the quenched case shows a decrease in the characteristic exponent to $\gamma=1.52$ and an increase in the mobility of $\sim 70$ cm$^2$ V$^{-1}$ s$^{-1}$ at room temperature. Despite the significant deformation potential of the homopolar mode, the effect of the quenching on the mobility is minor. Discussions and conclusions {#sec:VI} =========================== Based on our finding for the phonon-limited mobility in single-layer MoS$_2$, it seems likely that the low experimental mobilities of $\sim 1$ cm$^2$ V$^{-1}$ s$^{-1}$ reported in Refs.  are dominated by other scattering mechanisms as e.g. charged impurity scattering. The main reason for the increase in the mobility to 200 cm$^2$ V$^{-1}$ s$^{-1}$ observed when depositing a high-$\kappa$ dielectric on top of the MoS$_2$ layer in Ref.  is therefore likely to be impurity screening. The quenching of the out-of-plane homopolar phonon only led to a minor increase in the mobility and cannot alone explain the above-mentioned increase in the mobility. It may, however, also contribute. A comparison of the temperature dependence of the mobility in samples with and without the top-gate structure can help to clarify the extent to which a quenching of optical phonons contributes to the experimentally observed mobility increase. With our theoretically predicted room-temperature mobility of 410 cm$^2$ V$^{-1}$ s$^{-1}$, the observed enhancement in the mobility to 200 cm$^2$ V$^{-1}$ s$^{-1}$ induced by the high-$\kappa$ dielectric, suggests that dielectric engineering [@Konar:Engineering] is an effective route towards phonon-limited mobilities in 2D materials via efficient screening of charge impurities. A rough estimate of the impurity concentrations required to dominate phonon scattering can be inferred from the phonon scattering rate of $\tau^{-1} \sim 10^{13}$ s$^{-1}$ in Fig. \[fig:scatteringrates\]. The scattering rate is related to the mean-free path $\lambda$ of the carriers via $\lambda = \tau {\langle v\rangle}$ where ${\langle v\rangle}$ is their mean velocity. Using the velocity of the mean energy carriers for a non-degenerate distribution where ${\langle \varepsilon_{\mathbf{k}}\rangle} = k_\text{B} T$, we find a mean-free path of $\lambda \sim 14$ nm at $T=300$ K. In order for impurity scattering to dominate, the impurity spacing must be on the order of the phonon mean-free path or smaller. This results in a minimum impurity concentration of $\sim 5\times10^{11}$ cm$^{-2}$ in order to dominate phonon scattering. The high value of the estimated impurity concentration needed to dominate phonon scattering is in agreement with the experimental observation that low-mobility single-layer MoS$_2$ samples are heavily doped semiconductors [@Geim:2D]. As discussed in Section \[sec:II\], independent GW quasi-particle calculations [@Ciraci:Functionalization; @Olsen:MoS2] suggest that the ordering of the valleys at the bottom of the conduction band might not be as clear as predicted by DFT. If this is the case, the satellite valleys inside the Brillouin zone must also be taken into account in the solution of Boltzmann equation. This gives rise to additional intervalley scattering channels that together with the larger average effective electron mass $m^* \sim 0.6$ $m_e$ of the satellite valley will result in mobilities below the values predicted here. In conclusion, we have used a first-principles based approach to establish the strength and nature of the electron-phonon coupling and calculate the intrinsic phonon-limited mobility in single-layer MoS$_2$. The calculated room-temperature mobility of 410 cm$^2$ V$^{-1}$ s$^{-1}$ is to a large extent dominated by optical deformation potential scattering on the intravalley homopolar and intervalley LO phonons as well as polar optical scattering on the intravalley LO phonon via the Fr[ö]{}hlich interaction. The mobility follows a $\mu \sim T^{-1.69}$ temperature dependence at room temperature characteristic of optical phonon scattering. A quenching of the homopolar mode likely to occur in top-gated samples, results in a change of the exponent in the temperature dependence of the mobility to 1.52. With effective masses, phonons and measured mobilities in other semiconducting metal dichalcogenides being similar to those of MoS$_2$ [@Mooser:Mobility; @Fivaz:Dimensionality], room-temperature mobilities of the same order of magnitude can be expected in their single-layer forms. The authors would like to thank O. Hansen and J. J. Mortensen for illuminating discussions. KK has been partially supported by the Center on Nanostructuring for Efficient Energy Conversion (CNEEC) at Stanford University, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Award Number DE-SC0001060. CAMD is supported by the Lundbeck Foundation. First-principles calculation of the electron-phonon interaction {#app:e-ph} =============================================================== The first-principles scheme for the calculation of the electron-phonon interaction used in this work is outlined in the following. Contrary to other first-principles approaches for the calculation of the electron-phonon interaction which are based on pseudo potentials [@Sjakste:GaAsInter; @Louie:e-ph; @Sjakste:Silicon; @Kim:ElPhGraphene], the present approach is based on the PAW method [@Blochl:PAW] and is implemented in the GPAW DFT package [@GPAW; @GPAW1; @GPAW2]. Within the adiabatic approximation for the electron-phonon interaction, the electron-phonon coupling matrix elements for the phonon ${\mathbf{q}}\lambda$ and Bloch state ${\mathbf{k}}$ can be expressed as $$\label{eq:g_bloch} g^{\lambda}_{{\mathbf{k}}{\mathbf{q}}} = \sqrt{\frac{\hbar}{2MN\omega_{{\mathbf{q}}\lambda}}} \sum_{\alpha} {\langle\mathbf{k+q} \rvert} \hat{\mathbf{e}}_{\alpha}^{{\mathbf{q}}\lambda} \cdot \nabla_{\alpha}^{{\mathbf{q}}} V({\mathbf{r}}) {\lvert {\mathbf{k}}\rangle} ,$$ where $N$ is the number of unit cells, $M$ is an appropriately defined effective mass, $\omega_{{\mathbf{q}}\lambda}$ is the phonon frequency, and the ${\mathbf{q}}$-dependent derivative of the effective potential for a given atom $\alpha$ is defined by $$\label{eq:V_q} \nabla_\alpha^{{\mathbf{q}}} V({\mathbf{r}}) = \sum_l e^{i {\mathbf{q}}\cdot \mathbf{R}_l } \nabla_{\alpha l} V({\mathbf{r}}) ,$$ where the sum is over unit cells in the system and $\nabla_{\alpha l} V$ is the gradient of the potential with respect to the position of atom $\alpha$ in cell $\mathbf{R}_l$. The polarization vectors $\hat{\mathbf{e}}_{{\mathbf{q}}\lambda}$ of the phonons appearing in Eq.  are normalized according to $\sum_{\alpha} (M_{\alpha}/M)\vert \hat{\mathbf{e}}_{\alpha}^{{\mathbf{q}}\lambda} \vert^2 = 1$. In order to evaluate the matrix element in Eq. , the Bloch states are expanded in LCAO basis orbitals ${\lvert i \mathbf{R}_n\rangle}$ where $i=(\alpha,\mu)$ is a composite index for atomic site and orbital index and $\mathbf{R}_n$ is the lattice vector to the $n$’th unit cell. In the LCAO basis, the Bloch states are expanded as ${\lvert {\mathbf{k}}\rangle} = \sum_i c_i^{\mathbf{k}}{\lvert i {\mathbf{k}}\rangle}$ where $$\label{eq:bloch_sum} {\lvert i {\mathbf{k}}\rangle} = \frac{1}{N} \sum_n e^{i {\mathbf{k}}\cdot \mathbf{R}_n} {\lvert i \mathbf{R}_n\rangle} .$$ are Bloch sums of the localized orbitals. Inserting the Bloch sum expansion of the Bloch states in the matrix element in Eq. , the matrix element can be written $$\begin{aligned} \label{eq:g_blochsum} {\langle\mathbf{k+q} \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot & \nabla_{\mathbf{q}}V({\mathbf{r}}) {\lvert {\mathbf{k}}\rangle} = \sum_{ij} c_i^{*} c_j {\langlei \mathbf{k+q} \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot \nabla_{\mathbf{q}}V({\mathbf{r}}) {\lvert j {\mathbf{k}}\rangle} \nonumber \\ = & \frac{1}{N^2} \sum_{ij} c_i^{*} c_j \sum_{lmn} e^{i ({\mathbf{k}}+ \mathbf{G}) \cdot (\mathbf{R}_n - \mathbf{R}_m) - i {\mathbf{q}}\cdot (\mathbf{R}_m - \mathbf{R}_l ) } \nonumber \\ & \quad\quad \times {\langlei \mathbf{R}_m \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot \nabla_l V({\mathbf{r}}) {\lvert j \mathbf{R}_n\rangle}\end{aligned}$$ where the ${\mathbf{k}}$-labels on the expansion coefficients have been discarded for brevity. The phase factor from the reciprocal lattice vector $\mathbf{G}$ can be neglected since $\mathbf{R}_n \cdot \mathbf{G} = 2\pi N$. By exploiting the translational invariance of the crystal the matrix elements can be obtained from the gradient in the primitive cell as $$\begin{aligned} & {\langlei \mathbf{R}_m \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot \nabla_l V({\mathbf{r}}) {\lvert j \mathbf{R}_n\rangle} \nonumber \\ & = {\langlei \mathbf{R}_m - \mathbf{R}_l \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot \nabla_0 V({\mathbf{r}}) {\lvert j \mathbf{R}_n - \mathbf{R}_l\rangle}\end{aligned}$$ by performing the change of variables ${\mathbf{r}}' = {\mathbf{r}}- \mathbf{R}_l$ in the integration. Inserting in Eq.  and changing the summing variables to $\mathbf{R}_m - \mathbf{R}_l$ and $\mathbf{R}_n - \mathbf{R}_l$ we find for the matrix element $$\begin{aligned} \label{eq:g_blochsum_final} {\langle\mathbf{k+q} \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot \nabla_{\mathbf{q}}V({\mathbf{r}}) {\lvert {\mathbf{k}}\rangle} & = \frac{1}{N^2} \sum_{ij} c_i^{*} c_j \sum_{lmn} e^{i {\mathbf{k}}\cdot (\mathbf{R}_n - \mathbf{R}_m) - i {\mathbf{q}}\cdot \mathbf{R}_m} {\langlei \mathbf{R}_m \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot \nabla_0 V({\mathbf{r}}) {\lvert j \mathbf{R}_n\rangle} \nonumber \\ & = \frac{1}{N} \sum_{ij} c_i^{*} c_j \sum_{mn} e^{i {\mathbf{k}}\cdot (\mathbf{R}_n - \mathbf{R}_m) - i {\mathbf{q}}\cdot \mathbf{R}_m} {\langlei \mathbf{R}_m \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot \nabla_0 V({\mathbf{r}}) {\lvert j \mathbf{R}_n\rangle}\end{aligned}$$ Here, the sum over $k$ in the first equality produces a factor of $N$. The result for the matrix element in Eq.  is similar to the matrix element reported in Eq. (22) of Ref.  where a Wannier basis was used instead of the LCAO basis. From the last equality in Eq. , the procedure for a supercell-based evaluation of the matrix element has emerged. The matrix element in the last equality involves the gradients of the effective potential $\nabla_0 V$ with respect to atomic displacement in the reference cell at $\mathbf{R}_0$. These can be obtained using a finite-difference approximation for the gradient where the individual components are obtained as $$\label{eq:fd} \frac{\partial V}{\partial R_{\alpha i}} = \frac{V^+_{\alpha i}({\mathbf{r}}) - V^-_{\alpha i}({\mathbf{r}})}{2\delta} .$$ Here, $V^{\pm}_{\alpha i}$ denotes the effective potential with atom $\alpha$ located at $\mathbf{R}_\alpha$ in the reference cell displaced by $\pm \delta$ in direction $i=x,y,z$. The calculation of the gradient thus amounts to carrying out self-consistent calculations for six displacements of each atom in the primitive unit cell. Having obtained the gradients of the effective potential, the matrix elements in the LCAO basis of the supercell must be calculated and the sums over unit cells and atomic orbitals in Eq.  can be evaluated. In general, the matrix elements of the electron-phonon interaction must be converged with respect to the supercell size and the LCAO basis. In particular, since the supercell approach relies on the gradient of the effective potential going to zero at the supercell boundaries, the supercell must be chosen large enough that the potential at the boundaries is negligible. For polar materials where the coupling to the polar LO phonon is long-ranged in nature, a correct description of the coupling can only be obtained for phonon wave vectors corresponding to wavelengths smaller than the supercell size. Large supercells are therefore required to capture the long-wavelength limit of the coupling constant (see e.g. main text). PAW details ----------- In the PAW method, the effective single-particle DFT Hamiltonian is given by $$\label{eq:H_eff_PAW} H = -\frac{1}{2} \nabla^2 + v_\text{eff}({\mathbf{r}}) + \sum_\alpha \sum_{i_1 i_2} \; {\lvert \tilde{p}_{i_1}^\alpha\rangle} \Delta H_{i_1 i_2}^\alpha {\langle\tilde{p}_{i_2}^\alpha \rvert} ,$$ where the first term is the kinetic energy, $v_\text{eff}$ is the effective potential containing contributions from the atomic potentials and the Hartree and exchange-correlation potentials and the last term is the non-local part including the atom-centered projector functions $\tilde{p}_i^\alpha({\mathbf{r}})$ and the atomic coefficients $\Delta H_{i_1 i_2}^\alpha$. In contrast to pseudo-potential methods where the atomic coefficients are constants, they depend on the density and thereby also on the atomic positions in the PAW method. The diagonal matrix elements (i.e. ${\mathbf{q}}= \mathbf{0}$) in Eq.  correspond to the first-order variation in the band energy $\varepsilon_{\mathbf{k}}$ upon an atomic displacement in the normal mode direction $\hat{\mathbf{e}}_{{\mathbf{q}}\lambda}$. Together with the matrix elements for ${\mathbf{q}}\neq \mathbf{0}$, these can be obtained from the gradient of the PAW Hamiltonian with respect to atomic displacements. The derivative of  with respect to a displacement of atom $\alpha$ in direction $i$ results in the following four terms $$\begin{aligned} \frac{\partial H}{\partial R_{\alpha i}} = \frac{\partial v_\text{eff}}{\partial R_{\alpha i}} + \sum_{i_1,i_2} \bigg[ & {\lvert \frac{\partial\tilde{p}_{i_1}^\alpha}{\partial R_{\alpha i}}\rangle} \Delta H_{i_1 i_2}^\alpha {\langle\tilde{p}_{i_2}^\alpha \rvert} \nonumber \\ & + {\lvert \tilde{p}_{i_1}^\alpha\rangle} \Delta H_{i_1 i_2}^\alpha {\langle\frac{\partial\tilde{p}_{i_2}^\alpha}{\partial R_{\alpha i}} \rvert} \nonumber \\ & + {\lvert \tilde{p}_{i_1}^\alpha\rangle} \frac{\partial \Delta H_{i_1 i_2}^\alpha}{\partial R_{\alpha i}} {\langle\tilde{p}_{i_2}^\alpha \rvert} \bigg] .\end{aligned}$$ While the gradient of the projector functions in the first and second terms inside the square brackets can be evaluated analytically, the gradients of the effective potential and the projector coefficients in the last term are obtained using the finite-difference approximation in Eq. . Once the gradient of the PAW Hamiltonian has been obtained, the matrix elements in Eq.  can be obtained. Under the assumption that the smooth pseudo Bloch wave functions $\tilde{\psi}_{\mathbf{k}}$ from the PAW formalism is a good approximation to the true quasi particle wave function, the matrix element follows as $${\langle\mathbf{k+q} \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot \nabla_{\mathbf{q}}V({\mathbf{r}}) {\lvert {\mathbf{k}}\rangle} = {\langle\tilde{\psi}_\mathbf{k+q} \rvert} \hat{\mathbf{e}}_{{\mathbf{q}}\lambda} \cdot \nabla_{\mathbf{q}}H {\lvert \tilde{\psi}_{\mathbf{k}}\rangle} ,$$ where $\nabla_{\mathbf{q}}H$ is given in Eq. . With the pseudo wave function expanded in an LCAO basis, the matrix element is evaluated following Eq. . In order to verify the calculated matrix elements, we have carried out self-consistent calculations of the variation in the band energies with respect to atomic displacements along the phonon normal mode. For the coupling to the $\Gamma$-point homopolar phonon in single-layer MoS$_2$, we find that the calculated matrix element of 4.8 eV/[Å]{} agrees with the self-consistently calculated value to within $0.1$ eV/[Å]{}. As the matrix elements of the electron-phonon interaction are sensitive to the behavior of the potential and wave functions in the vicinity of the atomic cores, variations in electron-phonon couplings obtained with different pseudo-potential approximations can be expected. We have confirmed this via self-consistent calculations of the band energy variations, which showed that the coupling to the homopolar $\Gamma$-point phonon increases with $0.3$ eV/[Å]{} compared to the PAW value when using norm-conserving HGH pseudo potentials. Fr[ö]{}hlich interaction in 2D materials {#app:frohlich} ======================================== In the long-wavelength limit, the lattice vibration of the polar LO mode gives rise to a macroscopic polarization that couples to the charge carriers. In three-dimensional bulk systems, the coupling strength is given by Eq. \[eq:frohlich\] which diverges as $1/q$. Using dielectric continuum models, the coupling has been studied for confined carriers and LO phonons in semiconductor heterostructures [@Ando:ElphHetero; @Lugli:Microscopic] where a non-diverging interaction is found. For atomically thin materials, however, macroscopic dielectric models are inappropriate. Using an approach based on the atomic Born effective charges $Z^*$, a functional form that fits the calculated values of the Fr[ö]{}hlich interaction in Fig. \[fig:frohlich\] is here derived. Polarization field and potential -------------------------------- In two-dimensional materials, the polarization from the lattice vibration of the polar LO phonon is oriented along the plane of the layer. It can be expressed in terms of the relative displacement $\mathbf{u}_{\mathbf{q}}$ of the unit cell atoms as $$\mathbf{P}_{\mathbf{q}}(z) = \frac{Z^*}{\varepsilon_\infty A} \mathbf{u}_{\mathbf{q}}f_{\mathbf{q}}(z) ,$$ where ${\mathbf{q}}$ is the two-dimensional phonon wave vector, $\varepsilon_\infty$ is the optical dielectric constant, $Z^*$ is the Born effective charge of the atoms (here assumed to be the same for all atoms) and $f_{\mathbf{q}}$ describes the profile of the polarization in the direction perpendicular (here the $z$-direction) to the layer. The associated polarization charge $\rho = - \nabla \cdot \mathbf{P}$ is given by $\rho_{\mathbf{q}}= -i {\mathbf{q}}\cdot \mathbf{P}_{\mathbf{q}}$. The resulting scalar potential which couples to the carriers follows from Poisson’s equation. Fourier transforming in all three directions, Poisson’s equation takes the form $$\label{eq:poisson_ft} (q^2 + k^2) \phi_{\mathbf{q}}(k) = -i \frac{Z^*}{\varepsilon_\infty A} {\mathbf{q}}\cdot \mathbf{u}_{\mathbf{q}}f_{\mathbf{q}}(k) ,$$ where $k$ is the Fourier variable in the direction perpendicular to the plane of the layer and ${\mathbf{q}}\parallel \mathbf{u}_{\mathbf{q}}$ for the LO phonon. In 2D materials, the $z$-profile of the polarization field can to a good approximation be described by a $\delta$-function, i.e. $$f_{\mathbf{q}}(z) = \delta(z) .$$ Inserting the Fourier transform $f_{\mathbf{q}}(k) = 1$ in Eq. , we find for the potential $$\begin{aligned} \phi_{\mathbf{q}}(z) & = -i \frac{Z^*u_{\mathbf{q}}}{\varepsilon_\infty A} \int \! dk \; e^{ikz} f_{\mathbf{q}}(k) \frac{q}{q^2 + k^2} \nonumber \\ & = -i \frac{Z^*u_{\mathbf{q}}}{\varepsilon_\infty A} e^{-q {\lvert z \rvert}} \end{aligned}$$ which in agreement with the findings of Refs.  does not diverge in the long-wavelength limit. Electron-phonon interaction --------------------------- In three-dimensional bulk systems, the $q$-dependence of the Fr[ö]{}hlich interaction is given entirely by the $1/q$-divergence of the potential associated with the lattice vibration of the polar LO phonon. However, in two dimensions, the interaction follows by integrating the potential with the square of the envelope function $\chi(z)$ of the electronic Bloch state. Hence, the Fr[ö]{}hlich interaction is given by the matrix element $$g_\text{LO} (q) = \int \! dz \; \chi^*(z) \phi_{\mathbf{q}}(z) \chi (z) .$$ For simplicity, we here assume a Gaussian profile for the envelope function $$\chi(z) = \frac{1}{\pi^{1/4} \sqrt{\sigma}} e^{-z^2/2\sigma^2} ,$$ where $\sigma$ denotes the effective width of the electronic Bloch function. With this approximation for the envelope function, the Fr[ö]{}hlich interaction becomes $$\begin{aligned} g_\text{LO}(q) & = g_\text{Fr} \times e^{q^2 \sigma^2 / 4} \text{erfc}(q \sigma / 2) \nonumber \\ & \approx g_\text{Fr} \times \text{erfc}(q \sigma / 2) \end{aligned}$$ where $g_\text{Fr}$ is the Fr[ö]{}hlich coupling constant, $\text{erfc}$ is the complementary error function and the last equality holds in the long-wavelength limit where $q^{-1} \gg \sigma$. As shown in Fig. \[fig:frohlich\] in the main text, this functional form for the Fr[ö]{}hlich interaction gives a perfect fit to the calculated electron-phonon coupling for the polar LO mode. Evaluation of the inelastic collision integral {#app:collision} ============================================== Following Eq.  in the main text, the inelastic collision integral for scattering on optical phonons is split up in separate out- and in-scattering contributions, $${I'}_\text{inel}[\phi_k] = {I'}_\text{inel}^\text{out}[\phi_k] + {I'}_\text{inel}^\text{in}[\phi_k] ,$$ which are given by Eq.  and , respectively. With the assumption of dispersionless optical phonons, the Fermi-Dirac and Bose-Einstein distribution functions do not depend on the phonon wave vector and can thus be taken outside the ${\mathbf{q}}$-sum. The out-scattering part of the collision integral then takes the form $$\begin{aligned} \label{eq:I_out_1} \quad & {I'}_\text{inel}^\text{out}[\phi_k] = - \frac{2\pi}{\hbar} \frac{\phi_k}{{1 - f_{\mathbf{k}}^0}} \nonumber \\ & \times \Bigg[ N_\lambda^0 (1 - f_{{\mathbf{k}}+ {\mathbf{q}}}^0) \sum_{\mathbf{q}}\left\vert g_{\mathbf{q}}\right\vert^2 \delta(\varepsilon_{{\mathbf{k}}+ {\mathbf{q}}} - \varepsilon_{\mathbf{k}}- \hbar\omega_\lambda) \nonumber \\ & + (1 + N_\lambda^0 ) (1 - f_{{\mathbf{k}}- {\mathbf{q}}}^0) \sum_{\mathbf{q}}\left\vert g_{\mathbf{q}}\right\vert^2 \delta(\varepsilon_{{\mathbf{k}}- {\mathbf{q}}} - \varepsilon_{{\mathbf{k}}} + \hbar\omega_\lambda) \Bigg] .\end{aligned}$$ For the in-scattering part, the additional $\cos\theta_{{\mathbf{k}},{\mathbf{k}}\pm{\mathbf{q}}}$ factor is rewritten as $$\label{eq:cos_k_kplusq} \cos\theta_{{\mathbf{k}},{\mathbf{k}}\pm{\mathbf{q}}} = \frac{k \pm q \cos{\theta_{{\mathbf{k}},{\mathbf{q}}}}} {k\sqrt{1 \pm \frac{\hbar\omega_\lambda} {\varepsilon_{\mathbf{k}}}}} ,$$ leading to the following general form of the in-scattering part of the collision integral $$\begin{aligned} \label{eq:I_in_1} {I'}_\text{inel}^\text{in}& [\phi_k] = \frac{2\pi}{\hbar} \frac{1}{{1 - f_{\mathbf{k}}^0}} \nonumber \\ & \times \Bigg[ N_\lambda^0 (1 - f_{{\mathbf{k}}+ {\mathbf{q}}}^0) \frac{\phi_{k+q}}{k \sqrt{1 + \frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}} \nonumber \\ & \quad \times \sum_{\mathbf{q}}\left\vert g_{\mathbf{q}}\right\vert^2 (k + q \cos{\theta_{{\mathbf{k}},{\mathbf{q}}}} ) \delta(\varepsilon_{{\mathbf{k}}+ {\mathbf{q}}} - \varepsilon_{\mathbf{k}}- \hbar\omega_\lambda) \nonumber \\ & \quad + (1 + N_\lambda^0 ) (1 - f_{{\mathbf{k}}- {\mathbf{q}}}^0) \frac{\phi_{k-q}}{k \sqrt{1 - \frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}} \nonumber \\ & \quad \times \sum_{\mathbf{q}}\left\vert g_{\mathbf{q}}\right\vert^2 (k - q \cos{\theta_{{\mathbf{k}},{\mathbf{q}}}} ) \delta(\varepsilon_{{\mathbf{k}}- {\mathbf{q}}} - \varepsilon_{{\mathbf{k}}} + \hbar\omega_\lambda) \Bigg] .\end{aligned}$$ This depends on the deviation function in $\phi_{k\pm q} \equiv \phi(\varepsilon_{\mathbf{k}}\pm \hbar\omega_\lambda)$ and thus couples the deviation function at the initial energy $\varepsilon_{\mathbf{k}}$ of the carrier to that at $\varepsilon_{\mathbf{k}}\pm \hbar\omega_\lambda$. The two terms inside the square brackets account for emission and absorption out of the state ${\mathbf{k}}\pm {\mathbf{q}}$ and into the state ${\mathbf{k}}$, respectively. Integration over $q$ -------------------- The in- and out-scattering contributions to the inelastic part of the collision integral in Eqs.  and  can be written on the general form $$\label{eq:collision_general} {I'}_\text{inel}^\text{in/out}({\mathbf{k}}) = \sum_{\mathbf{q}}f(q) \delta(\varepsilon_{{\mathbf{k}}\pm {\mathbf{q}}} - \varepsilon_{{\mathbf{k}}} \mp \hbar\omega_{{\mathbf{q}}\lambda}) ,$$ where the function $f$ accounts for the $q$-dependent functions inside the ${\mathbf{q}}$-sum. The conservation of energy and momentum in a scattering event is secured by the $\delta$-function entering the collision integral. Following the procedure outlined in Ref. , the integral over $q$ resulting from the ${\mathbf{q}}$-sum in the collision integral is evaluated using the following property of the $\delta$-function, $$\int \! dq \; F(q) \delta\left[ g(q) \right] = \sum_n \frac{F(q_n)}{{\lvert g'(q_n) \rvert}} .$$ Here, $F(q)=qf(q)$ and $q_n$ are the roots of $g$. Rewriting the argument of the $\delta$-function in Eq.  as $$\begin{aligned} \label{eq:conservation} \varepsilon_{{\mathbf{k}}\pm {\mathbf{q}}} - \varepsilon_{{\mathbf{k}}} \mp \hbar\omega_{{\mathbf{q}}\lambda} & = \frac{\hbar^2 ({\mathbf{k}}\pm {\mathbf{q}})^2}{2m^*} - \frac{\hbar^2k^2}{2m^*} \mp \hbar\omega_{{\mathbf{q}}\lambda} \nonumber \\ & = \frac{\hbar^2}{2 m^*} \left(q^2 \pm 2 kq \cos{\theta_{{\mathbf{k}},{\mathbf{q}}}} \mp \frac{2 m^*\omega_{{\mathbf{q}}\lambda}}{\hbar} \right) \nonumber \\ & \equiv \frac{\hbar^2}{2 m^*} g(q)\end{aligned}$$ we get $$\label{eq:collision_delta} {I'}({\mathbf{k}}) = \frac{A}{(2\pi)^2} \int \! d\theta_{{\mathbf{k}},{\mathbf{q}}} \; \frac{2 m^*}{\hbar^2} \sum_n \frac{q_n f(q_n)}{{\lvert g'(q_n) \rvert}} .$$ Depending on the $q$-dependence of the phonon frequency, different solutions for the roots $q_n$ result. ### Dispersionless optical phonons With the assumption of dispersionless optical phonons, i.e. $\omega_{{\mathbf{q}}\lambda} = \omega_\lambda$, the roots of $g$ in Eq.  become $$\label{eq:roots} q / k = \mp \cos{\theta_{{\mathbf{k}},{\mathbf{q}}}} \pm \sqrt{\cos^2{\theta_{{\mathbf{k}},{\mathbf{q}}}} \pm \frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}$$ for absorption (upper sign in the first and last term) and emission (lower sign in the first and last term), respectively. For absorption there is only one positive root $q_+^a$ given by the plus sign in front of the square root, $$\label{eq:q^abs} q_+^a = -k \cos{\theta_{{\mathbf{k}},{\mathbf{q}}}} + k \sqrt{\cos^2{\theta_{{\mathbf{k}},{\mathbf{q}}}} + \frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}$$ where all values of $\theta_{{\mathbf{k}},{\mathbf{q}}}$ are allowed. The absorption terms in the collision integral then becomes $${I'}_a({\mathbf{k}}) = \frac{A}{(2\pi)^2} \int_0^{2\pi} \! d\theta_{{\mathbf{k}},{\mathbf{q}}} \; \frac{m^*}{k\hbar^2} \frac{q_+^a f(q_+^a)} {\sqrt{\cos^2{\theta_{{\mathbf{k}},{\mathbf{q}}}} + \frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}}.$$ In the case of phonon emission, both possible roots in Eq.  $$\label{eq:q^emis} q_\pm^e = k \cos{\theta_{{\mathbf{k}},{\mathbf{q}}}} \pm k \sqrt{\cos^2{\theta_{{\mathbf{k}},{\mathbf{q}}}} - \frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}$$ can take on positive values. However, the minus sign inside the square root restricts the allowed values of the integration angle to the range $\theta_{{\mathbf{k}},{\mathbf{q}}} \le \theta_0 = \arccos{(\hbar\omega_\lambda/\varepsilon_{\mathbf{k}})^{1/2}}$. Furthermore, in order to secure a positive argument to the square root, the electron energy must be larger than the phonon energy, $\varepsilon_{\mathbf{k}}> \hbar\omega_\lambda$. Hence, the emission terms can be obtained as $$\begin{aligned} {I'}_e({\mathbf{k}}) = & \frac{A}{(2\pi)^2} \Theta(\varepsilon_{\mathbf{k}}- \hbar\omega_\lambda) \nonumber \\ & \times \int_{-\theta_0}^{\theta_0} \! d\theta_{{\mathbf{k}},{\mathbf{q}}} \; \frac{m^*}{k\hbar^2} \frac{q_+^e f(q_+^e) + q_-^e f(q_-^e)}{ \sqrt{\cos^2{\theta_{{\mathbf{k}},{\mathbf{q}}}} - \frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}}.\end{aligned}$$ Collision integral for optical phonon scattering ------------------------------------------------ In the following two sections, analytic expressions for collision integral in the case of zero- and first-order optical deformation potential coupling are given. For the Fr[ö]{}hlich interaction the integration over $\theta$ can not be reduced to a simple analytic form and is therefore evaluated numerically. ### Zero-order deformation potential For the zero-order deformation potential coupling in Eq. , the $\theta$-integration in the out-scattering part of the collision integral yields a factor of $2\pi$ resulting in the following form $$\begin{aligned} \label{eq:I_out_zero} & {I'}_\text{inel}^\text{out} [\phi_k] = - \frac{m^* D_\lambda^2}{2\hbar^2\rho\omega_\lambda} \frac{\phi_k}{{1 - f_{\mathbf{k}}^0}} \nonumber \\ & \times \bigg[ (1 - f_{{\mathbf{k}}+ {\mathbf{q}}}^0) + (1 - f_{{\mathbf{k}}- {\mathbf{q}}}^0) e^{\hbar\omega_\lambda / k_\text{B}T} \Theta(\varepsilon_{{\mathbf{k}}} - \hbar\omega_{\lambda} ) \bigg] N_{\lambda}^0 .\end{aligned}$$ For the in-scattering part, the $q$-independence of the deformation potential results in a vanishing ${\mathbf{q}}$-sum and the in-scattering contribution is zero, i.e. $$\label{eq:I_in_zero} {I'}_\text{inel}^\text{in}[\phi_k] = 0 .$$ ### First-order deformation potential For first-order deformation potentials, we find for the out-scattering part $$\begin{aligned} \label{eq:I_out_first} {I'}_\text{inel}^\text{out} & [\phi_k] = - \frac{m^{*2} \Xi_\lambda^2}{\hbar^4\rho\omega_\lambda} \frac{\phi_k}{{1 - f_{\mathbf{k}}^0}} \times \bigg[ (1 - f_{{\mathbf{k}}+ {\mathbf{q}}}^0)( 2\varepsilon_{\mathbf{k}}+ \hbar\omega_\lambda) \bigg. \nonumber \\ & + (1 - f_{{\mathbf{k}}- {\mathbf{q}}}^0) ( 2\varepsilon_{\mathbf{k}}- \hbar\omega_\lambda) e^{\hbar\omega_\lambda / k_\text{B}T} \Theta(\varepsilon_{{\mathbf{k}}} - \hbar\omega_{\lambda} ) \bigg] N_{\lambda}^0 .\end{aligned}$$ For the in-scattering part, the result for the first and second term inside the square brackets in Eq.  is $$\label{eq:I_in_first1} {I'}_\text{inel}^\text{in} [\phi_k] = - \frac{m^{*2} \Xi_\lambda^2}{\hbar^4\rho\omega_\lambda} \frac{1 - f_{{\mathbf{k}}+{\mathbf{q}}}^0}{1 - f_{\mathbf{k}}^0} \frac{\phi_{k+q}}{\sqrt{1 + \frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}} N_{\lambda}^0 (\varepsilon_{\mathbf{k}}+ \hbar\omega_\lambda)$$ and $$\begin{aligned} \label{eq:I_in_first2} {I'}_\text{inel}^\text{in} & [\phi_k] = - \frac{m^{*2} \Xi_\lambda^2}{\hbar^4\rho\omega_\lambda} \frac{1 - f_{{\mathbf{k}}- {\mathbf{q}}}^0}{1 - f_{\mathbf{k}}^0} \frac{\phi_{k-q}}{\sqrt{1 - \frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}} \nonumber \\ & \times (1 + N_{\lambda}^0 ) \left( \varepsilon_{\mathbf{k}}- \hbar\omega_\lambda \right) \Theta(\varepsilon_{{\mathbf{k}}} - \hbar\omega_{\lambda} ) \nonumber \\ & \times \frac{1}{\pi} \left( \frac{\pi}{2} + \arccos{\sqrt{\frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}} + \arcsin{\sqrt{\frac{\hbar\omega_\lambda}{\varepsilon_{\mathbf{k}}}}} \right) ,\end{aligned}$$ which accounts for absorption and emission, respectively. [56]{} natexlab\#1[\#1]{}bibnamefont \#1[\#1]{}bibfnamefont \#1[\#1]{}citenamefont \#1[\#1]{}url \#1[`#1`]{}urlprefix\[2\][\#2]{} \[2\]\[\][[\#2](#2)]{} , ****, (). , , , , , ****, (). , , , , ****, (). , ****, (). , , , , , , , ****, (). , , , , ****, (). , , , , , , , ****, (). , , , , , ****, (). , , , , , , ****, (). , , , , , ****, (). , , , , , ****, (). , , , , , , , , ****, (). , , , ****, (). , ****, (). , ****, (). , ****, (). , , , ****, (). , ****, (). , , , , , ****, (). , , , ****, (). , , , , , ****, (). , , , , , , , , , , ****, (). , ****, (). , ****, (). , , , ****, (). , , , , , ****, (). , , , , , , , , , , , ****, (). , ****, (). , , , , , , , , , , , ****, (). , ****, (). , , , , ****, (). , , , , ****, (). , ****, (). , , , , , , , ****, (). , ****, (). , ****, (). , ** (, , ). , ** (, , ). , , , ****, (). , , , ****, (). , , , , , , , , ****, (). , ****, (). , , , ****, (). , ** (, ). , in **, edited by (, , ), vol. , pp. . , ** (, , ), ed. , ****, (). , , , in **, edited by , , (, , ), vol. , p. . , ****, (). , , , (). , , , ****, (). , ****, ().
{ "pile_set_name": "ArXiv" }
--- address: 'TUM Institute for Advanced Study, Lichtenbergstr. 2a, 85747 Garching, Germany' author: - DARIO BUTTAZZO title: NATURAL SCALARS IN THE NMSSM --- FLAVOUR(267104)-ERC-73 Introduction ============ Is the Higgs boson recently discovered at the LHC alone, or is it a member of an extended family of scalar particles? While this is a relevant question on its own, it has a fundamental importance in the context of supersymmetry, where at least a second doublet is required. Also in view of the negative results in the searches for many supersymmetric partners up to TeV-range masses, the extra Higgs bosons may well be the lightest particles in the spectrum – perhaps with the exception of the LSP – making the search for these scalar states an important task for present and future experiments. The Higgs system of the NMSSM contains two scalar $\SU(2)_L$ doublets $H_u$ and $H_d$, and a complex singlet $S$, all parts of the corresponding chiral supermultiplets, coupled through a cubic term $\lambda H_u H_d S$ in the superpotential. [@Fayet] Taking naturalness as a guideline, there are two main reasons for considering the NMSSM: [@Barbieri] 1. it adds a new tree-level contribution to the Higgs mass, which makes a 126 GeV Higgs boson compatible with lighter stops with respect to the MSSM; 2. it reduces the fine-tuning of the electroweak scale $v$ (fixed by the weak gauge couplings in the MSSM) for moderate $\tan\beta$ and $\lambda\approx 1$. The aim here is to present an analytical study of the Higgs system of the general NMSSM in a most natural scenario, without specifying any particular form of the scalar potential and avoiding the use of benchmark points. [^1] In order to simplify the analysis, however, we shall make the following motivated assumptions: 1. we assume a negligibly small CP violation in the Higgs system, and therefore we ignore the two CP-odd states, which do not influence the physics of the CP-even states; 2. we neglect any effect from loops of supersymmetric particles other than the correction $\Delta_t$ to the quartic coupling of $H_u$ due to the top-stop loop; this is motivated by the choice of a spectrum with all the s-particles as close as possible to their “naturalness limit”; 3. we assume $\mu A_t \lesssim m_{\tilde t}^2$, due to naturalness arguments, where $m_{\tilde t}$ is the average stop mass, $A_t$ is its trilinear coupling, and $\mu$ is the quadratic term in the superpotential; 4. we do not include any invisible decay of the lightest Higgs boson – e.g. into a pair of neutralinos; this can easily be corrected for rescaling all the branching ratios and signal strengths by a common factor $\Gamma/(\Gamma + \Gamma_{\rm inv})$. Parameter space of a generic NMSSM ================================== Assuming a negligibly small violation of CP in the Higgs sector, the three neutral CP-even fields $\mathcal{H} = (H_u^0,H_d^0,S)^\mathsf{T}$ are related to the physical mass eigenstates $\mathcal{H}_{\rm ph} = (h_1,h_2,h_3)^\mathsf{T}$ by $$\label{rotation} \mathcal{H} = R_\alpha^{12}R_\gamma^{23}R_\sigma^{13}\mathcal{H}_{\rm ph} \equiv R\,\mathcal{H}_{\rm ph},$$ where $R_x^{ij}$ are rotations by an angle $x$ in the $(i,j)$ sector. Their squared mass matrix reads, in the $\mathcal{H}$ basis, $$\label{scalar_mass_matrix} {\cal M}^2=\left( \begin{array}{ccc} m_A^2 c^2_\beta+m_Z^2 s^2_\beta +\Delta_t^2/s_\beta^2 & \left(2 v^2 \lambda ^2-m_A^2-m_Z^2\right) c_\beta s_\beta & v M_1 \\ \left(2 v^2 \lambda ^2-m_A^2-m_Z^2\right) c_\beta s_\beta & m_Z^2 c^2_\beta + m_A^2 s^2_\beta & v M_2 \\ v M_1 & v M_2 & M_3^2 \end{array} \right),$$ where $$\label{mHcharged} m_A^2 = m_{H^{\pm}}^2 - m_W^2 +\lambda^2 v^2,$$ $m_{H^{\pm}}$ is the physical mass of the single charged Higgs boson, $\Delta_t^2$ is the well-known effect of the top-stop loop corrections to the quartic coupling of $H_u$, and $v \simeq 174$ GeV. Here and in the following we write $s_x = \sin x, c_x = \cos x$. We neglect the corrections to $\mathcal{M}_{11}$ and $\mathcal{M}_{12}$, which are suppressed as the second and first power of $\mu A_t/m_{\tilde t}^2$, respectively, and the analogous correction to . We leave unspecified the other parameters $M_1, M_2, M_3$ in , which are not directly related to physical masses and depend on the particular NMSSM under consideration – i.e. the form of the singlet potential. The matrix $\mathcal{M}$ is related to the physical scalar masses by $$R^\mathsf{T} {\cal M}^2 R = \mathrm{diag}(m_{h_1}^2, m_{h_2}^2, m_{h_3}^2). \label{diag_matrix}$$ In the following, we identify $h_1$ with the state found at the LHC, so that $m_{h_1} = 125.7$ GeV. For simplicity we shall always consider $h_1$ as the lightest CP-even state, although other cases with a lighter scalar are still compatible with current data. [@Barbieri2] Although the full matrix $\mathcal{M}^2$ depends on the specific model in consideration, its $2\times 2$ submatrix in the (1,2) sector provides, by the use of , three relations between the mixing angles and the physical masses [@Barbieri1; @Barbieri2] which do not depend on the unknown quantities $M_1$, $M_2$, $M_3$. The Higgs system of the NMSSM is thus completely determined by the parameters $m_{h_{1,2,3}}$, $m_{H^\pm}$, $\lambda$, $t_\beta$, $\Delta_t$. Due to the large number of free parameters, in order to simplify the analysis we will consider the limiting cases where only two states out of three are light: - Singlet decoupled: $m_{h_3} \gg m_{h_1}, m_{h_2}$ or $M_3^2 \gg v M_1, v M_2$, and $\sigma, \gamma \rightarrow 0$; - Doublet decoupled: $m_{h_2} \gg m_{h_1}, m_{h_3}$ or $m_A^2 \gg v M_1, v M_2$, and $\sigma, \delta = \alpha - \beta + \pi/2 \rightarrow 0$. ![\[fit\] Fit of the measured signal strengths of $h_1 = h_{\text{LHC}}$. Left: 3-parameter fit of $t_\beta$, $s_{\delta}$ and $s_{\gamma}^{2}$. The allowed regions at 95% C.L. are given for $s_{\gamma}^{2} = 0$ (green) and $s_\gamma^2 = 0.15$ (grey). Note that the two regions overlap in part. The dashed line shows the fit reach expected at LHC14 for $s_\gamma^2 = 0$. Right: fit of $s_{\gamma}^{2}$ in the case of $\delta = 0$ (solid) and its projection at LHC14 (dashed).](fit1.pdf "fig:"){width=".49\textwidth"}![\[fit\] Fit of the measured signal strengths of $h_1 = h_{\text{LHC}}$. Left: 3-parameter fit of $t_\beta$, $s_{\delta}$ and $s_{\gamma}^{2}$. The allowed regions at 95% C.L. are given for $s_{\gamma}^{2} = 0$ (green) and $s_\gamma^2 = 0.15$ (grey). Note that the two regions overlap in part. The dashed line shows the fit reach expected at LHC14 for $s_\gamma^2 = 0$. Right: fit of $s_{\gamma}^{2}$ in the case of $\delta = 0$ (solid) and its projection at LHC14 (dashed).](fit2.pdf "fig:"){width=".49\textwidth"} Higgs couplings =============== ATLAS CMS --------------------- ------- ------ $h\to \gamma\gamma$ 0.16 0.15 $h\to ZZ$ 0.15 0.11 $h\to WW$ 0.30 0.14 $Vh\to V b\bar b$ – 0.17 $h\to \tau\tau$ 0.24 0.11 $h\to \mu\mu$ 0.52 – : Projected uncertainties of the measurements of the signal strengths of $h_{\rm LHC}$, normalized to the SM, at the 14 TeV LHC with 300 fb$^{-1}$, both for ATLAS and CMS.[]{data-label="table"} From , $h_1 = h_{\text{LHC}}$ is related to the gauge eigenstates by $$\label{rotation_h1} h_1 = c_{\gamma} (-s_{\alpha} H_d^0 + c_\alpha H_u^0) + s_{\gamma} S,$$ and similar relations, also involving the angle $\sigma$, hold for $h_2$ and $h_3$. The angles $\delta = \alpha - \beta + \pi/2$ and $\gamma$ alone thus determine the couplings of $h_1$ to the fermions and to vector boson pairs, normalised to the corresponding couplings of the SM Higgs boson, $$\begin{aligned} \frac{g_{h_1tt}}{g^{\text{SM}}_{htt}} &= c_\gamma\Big(c_\delta +\frac{s_\delta}{t_\beta}\Big), & \frac{g_{h_1bb}}{g^{\text{SM}}_{hbb}} &= c_\gamma(c_\delta -s_\delta t_\beta ), & \frac{g_{h_1VV}}{g^{\text{SM}}_{hVV}} &= c_\gamma c_\delta. \label{h1couplings}\end{aligned}$$ A fit of all ATLAS, CMS and TeVatron data collected so far [@rates]^,^ [^2] on the various signal strengths of $h_{\text{LHC}}$ can then be used to put bounds on $\delta$ and $\gamma$ (as a function of $t_\beta$). We perform this fit adapting the code of Giardino [*et al.*]{} [@Giardino] As stated above, we do not include in and in the fit any loop effect from supersymmetric particles. The 95% C.L. allowed regions for $\delta$, at different fixed values of $\gamma$, and for $\gamma$ at $\delta = 0$, are shown in Figure \[fit\]. To quantify the impact of the future experimental improvement in the measurements of the various signal strengths of $h_{\rm LHC}$, as foreseen in the next run of the LHC, we have repeated the previous fit assuming the expected errors on the signal strengths [@projection] with a luminosity of 300 fb$^{-1}$ at $\sqrt{s} = 14$ TeV as in Table \[table\], and central values as in the Standard Model. The results of this projection are also shown in Figure \[fit\] as a comparison. Notice that, while a big improvement is expected in the – already quite precise – determination of $\delta$, the fit of $\gamma$ will only marginally improve. Singlet decoupled ================= If the singlet is decoupled, the only nonzero mixing angle is s\_\^2 = , where \[mhh\] m\_[hh]{}\^2 = m\_Z\^2 c\_[2]{}\^2 + v\^2\^2s\_[2]{}\^2 + \_t\^2, and both $m_{h_2}$ and $m_{H^\pm}$ can be expressed in terms of the three parameters $\lambda, t_\beta, \Delta_t$. The $\lambda\to 0$ limit corresponds to the MSSM case. The dependence on $\Delta_t$ is very mild and can be neglected if $\Delta_t$ itself is not too large, corresponding to a moderate level of fine-tuning. We therefore choose to fix $\Delta_t = 75$ GeV, which is compatible with an average stop mass of about 700 GeV. ![\[Sdec\]Singlet decoupled in the $(t_\beta, m_{h_2})$ plane. The orange region is excluded at 95% C.L. by the Higgs fit, the blue region is unphysical. Left: the colored dashed lines show the fit reach expected at LHC14, black isolines of $\lambda$ (solid) and $m_{H^\pm}$ (dashed). Right: isolines of the gluon fusion production cross-section $\sigma(gg\to h_2)$ at 14 TeV (solid) and of the decay branching ratio into top quark pairs ${\rm BR}(h_2\to t\bar t)$ (dashed).](Sdec1.pdf "fig:"){width="49.00000%"}![\[Sdec\]Singlet decoupled in the $(t_\beta, m_{h_2})$ plane. The orange region is excluded at 95% C.L. by the Higgs fit, the blue region is unphysical. Left: the colored dashed lines show the fit reach expected at LHC14, black isolines of $\lambda$ (solid) and $m_{H^\pm}$ (dashed). Right: isolines of the gluon fusion production cross-section $\sigma(gg\to h_2)$ at 14 TeV (solid) and of the decay branching ratio into top quark pairs ${\rm BR}(h_2\to t\bar t)$ (dashed).](Sdec2.pdf "fig:"){width="49.00000%"} The left panel of Figure \[Sdec\] shows the regions excluded by the fit, present and forseen, in the $(t_\beta,m_{h_2})$ plane, together with the isolines of $\lambda$ and $m_{H^\pm}$. A second light state is allowed for small $t_\beta$ and moderate values of $\lambda\lesssim 1$, unlike in the MSSM, [@Barbieri1] although masses below 300 GeV are disfavoured by the presence of a too light charged Higgs. Notice that with the improved measurements of the signal rates of $h_{\rm LHC}$ in the next stage of the LHC it will be possible to probe a large fraction of the parameter space solely by the Higgs fit. The couplings of $h_2$, on the other hand, are given by $$\begin{aligned} \frac{g_{h_2 tt}}{g_{htt}^{\rm SM}} &= s_\delta - \frac{c_\delta}{t_\beta}, & \frac{g_{h_2 bb}}{g_{hbb}^{\rm SM}} &= s_\delta + c_\delta t_\beta, & \frac{g_{h_2 VV}}{g_{hVV}^{\rm SM}} &= s_\delta,\end{aligned}$$ and allow to calculate its production cross-sections and most relevant branching ratios. The small values of $\lambda$ in the region allowed by the fit make the phenomenology of $h_2$ quite similar to the one of the heavier Higgs state of the MSSM. Its dominant decay mode is into fermions, either top or bottom quarks, depending on the mass $m_{h_2}$ and on $t_\beta$. The right panel of Figure \[Sdec\] shows the predictions for the NNLL gluon fusion production cross-section of $h_2$ at $\sqrt{s} = 14$ TeV, and for its branching ratio into top quark pairs ${\rm BR}(h_2\to t\bar t)$, in the same $(t_\beta, m_{h_2})$ plane. Doublet decoupled ================= ![\[Hdec\]Doublet decoupled in the $(t_\beta, m_{h_3})$ plane, for fixed $\lambda = 1$. The orange region is excluded at 95% C.L. by the Higgs fit. Left: the coloured dashed line shows the fit reach expected at LHC14, black isolines of $s_\gamma^2$. Right: isolines of the gluon fusion production cross-section $\sigma(gg\to h_3)$ at 14 TeV (solid) and of the decay branching ratio of $h_3$ into two Higgs bosons ${\rm BR}(h_3\to h_1 h_1)$, with $v_S = 2v$ (dashed).](Hdec1.pdf "fig:"){width="49.00000%"}![\[Hdec\]Doublet decoupled in the $(t_\beta, m_{h_3})$ plane, for fixed $\lambda = 1$. The orange region is excluded at 95% C.L. by the Higgs fit. Left: the coloured dashed line shows the fit reach expected at LHC14, black isolines of $s_\gamma^2$. Right: isolines of the gluon fusion production cross-section $\sigma(gg\to h_3)$ at 14 TeV (solid) and of the decay branching ratio of $h_3$ into two Higgs bosons ${\rm BR}(h_3\to h_1 h_1)$, with $v_S = 2v$ (dashed).](Hdec2.pdf "fig:"){width="49.00000%"} If the standard-like Higgs mixes with a singlet, and the second doublet is decoupled, the only nonzero mixing angle is s\_\^2 = . Since the off-diagonal entries $M_1$ and $M_2$ in are unknown, there is one free parameter more respect to the previous case. We therefore fix $\lambda = 1$, along with choosing $\Delta_t = 75$ GeV as before, in order to produce the plots of Figure \[Hdec\]. The left panel shows, as before, the 95% C.L. excluded region from the Higgs fit, together with its projection at 14 TeV, and the isolines of the mixing angle $s_\gamma^2$. Recall from that $s_\gamma^2$ increases with $\lambda$, thus enlarging the region allowed by the fit for smaller values of $\lambda$. [@Barbieri1] The mild improvement in the fit foreseen for the next run of the LHC in this particular case makes the direct searches for the heavy state $h_3$ crucial. The couplings of $h_3$, due to its singlet-like nature, are proportional to the ones of a standard Higgs boson with mass $m_{h_3}$, = = = -s\_. The branching ratios of $h_3$ are therefore the same as the standard ones, while the production cross-sections and decay widths are simply rescaled by $s_\gamma^2$ (still neglecting radiative corrections from supersymmetric particles). The dominant decay channel, if kinematically allowed, is the one of $h_3$ into two $h_1$ bosons, followed in importance by the one into two vector bosons. Unfortunately, all the triple scalar couplings depend on the particular form of the singlet potential, and are therefore model dependent. In the limit of large $\lambda$, however, this dependence can be parametrized simply by the vacuum expectation value of the singlet $v_S$. [@Barbieri1] In the right panel of Figure \[Hdec\] we show the branching ratio ${\rm BR}(h_3\to h_1 h_1)$ for $v_S = 2 v$, together with the gluon fusion production cross-section at $\sqrt{s} = 14$ TeV. Although not easily, searches in the $b\bar b\gamma\gamma$ and $b\bar b b\bar b$ channels could reach comparable sensitivities in the near future, probing regions of the parameter space difficultly accessible by other means. Finally, it is worth to mention that very large deviations of the triple $h_1$ coupling from the standard value can arise in some part of the parameter space, [@Barbieri1] perhaps making this measure also accessible at the LHC in the future. Conclusions =========== We have analysed the Higgs system of a most natural NMSSM, focussing on relations between physical parameters. The modified couplings of $h_1$, which influence the signal strengths measured at the LHC, provide a powerful tool to exclude regions of the parameter space. We have considered two limiting cases in which one of the states is much heavier than the others. In both cases a second light neutral CP-even scalar is consistent with all the constraints. If the singlet is decoupled, it will be possible to thoroughly explore the parameter space combining the refined measurements of the signal strengths of $h_{\rm LHC}$ with searches for the second Higgs decaying into fermions in the remaining allowed regions. The case where the doublet is decoupled is more difficult to test. The $h_1$ signal strengths are not very sensitive to the mixing with the second state, due to its singlet-like nature. Furthermore, the small production cross-section of $h_3$, together with its large branching ratio into a pair of $h_1$, whenever kinematically allowed, makes the search at the LHC challenging, although not impossible. It will in any case be interesting to follow the progression of the experimental searches for additional Higgs-like states, either direct or indirect, which are an independent way to probe weak-scale supersymmetry, complementary to the search for superpartners. Acknowledgments {#acknowledgments .unnumbered} =============== I would like to thank Riccardo Barbieri, Kristjan Kannike, Filippo Sala and Andrea Tesi for the fruitful collaborations, and the organizers of the Rencontres de Moriond for the pleasant and stimulating environment. Thanks also to Maxime Gouzevitch and Caterina Vernieri for interesting discussions. This work was supported by the ERC Advanced Grant project [Flavour]{} (agreement n. 267104). References {#references .unnumbered} ========== [99]{} P. Fayet, [*Nucl. Phys. B*]{} [**90**]{} (1975) 104-124. R. Barbieri, L. J. Hall, Y. Nomura and V. S. Rychkov, [*Phys. Rev. D*]{} [**75**]{} (2007) 035007, [arXiv:hep-ph/0607332](http://arxiv.org/abs/hep-ph/0607332); L. J. Hall, D. Pinner and J. T. Rudermann, [*JHEP*]{} [**1204**]{} (2012) 131, [arXiv:1112.2703 \[hep-ph\]](http://arxiv.org/abs/1112.2703). R. Barbieri, D. Buttazzo, K. Kannike, F. Sala and A. Tesi, [*Phys. Rev. D*]{} [**87**]{} (2013) 115018, [arXiv:1304.3670 \[hep-ph\]](http://arxiv.org/abs/1304.3670). R. Barbieri, D. Buttazzo, K. Kannike, F. Sala and A. Tesi, [*Phys. Rev. D*]{} [**88**]{} (2013) 055011, [arXiv:1307.4937 \[hep-ph\]](http://arxiv.org/abs/1307.4937). P. P. Giardino, K. Kannike, I. Masina, M. Raidal and A. Strumia, [arXiv:1303.3570\[hep-ph\]](http://arxiv.org/abs/1303.3570). ATLAS Collaboration, [twiki.cern.ch/twiki/bin/view/AtlasPublic/HiggsPublicResults](https://twiki.cern.ch/twiki/bin/view/AtlasPublic/HiggsPublicResults);\ CMS Collaboration, [twiki.cern.ch/twiki/bin/view/CMSPublic/PhysicsResultsHIG](https://twiki.cern.ch/twiki/bin/view/CMSPublic/PhysicsResultsHIG). ATLAS Collaboration, [ATL-PHYS-PUB-2012-004](http://cds.cern.ch/record/1484890); CMS Collaboration, [CMS-NOTE-2012-006](http://cds.cern.ch/record/1494600). [^1]: See Barbieri [*et al.*]{} [@Barbieri1; @Barbieri2] for more details. [^2]: See Barbieri [*et al.*]{} [@Barbieri1; @Barbieri2] for a detailed list of references.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We present a comprehensive high spatial-resolution imaging study of globular clusters (GCs) in NGC1399, the central giant elliptical cD galaxy in the Fornax galaxy cluster, conducted with the [*Advanced Camera for Surveys*]{} (ACS) aboard the [*Hubble Space Telescope*]{} (HST). Using a novel technique to construct drizzled PSF libraries for HST/ACS data, we accurately determine the fidelity of GC structural parameter measurements from detailed artificial star cluster experiments and show the superior robustness of the GC half-light radius, $r_h$, compared with other GC structural parameters, such as King core and tidal radius. The measurement of $r_h$ for the major fraction of the NGC1399 GC system reveals a trend of increasing $r_h$ versus galactocentric distance, $R_{\rm gal}$, out to about 10kpc and a flat relation beyond. This trend is very similar for blue and red GCs which are found to have a mean size ratio of $r_{\rm h, red}/r_{\rm h, blue}\!=\!0.82\pm0.11$ at all galactocentric radii from the core regions of the galaxy out to $\sim\!40$kpc. This suggests that the size difference between blue and red GCs is due to [*internal*]{} mechanisms related to the evolution of their constituent stellar populations. Modeling the mass density profile of NGC1399 shows that additional [*external*]{} dynamical mechanisms are required to limit the GC size in the galaxy halo regions to $r_h\!\approx\!2$ pc. We suggest that this may be realized by an exotic GC orbit distribution function, an extended dark matter halo, and/or tidal stress induced by the increased stochasticity in the dwarf halo substructure at larger galactocentric distances. We compare our results with the GC $r_h$ distribution functions in various galaxies and find that the fraction of extended GCs with $r_h\geq5$pc is systematically larger in late-type galaxies compared with GC systems in early-type galaxies. This is likely due to the dynamically more violent evolution of early-type galaxies. We match our GC $r_h$ measurements with radial velocity data from the literature and split the resulting sample at the median $r_h$ value into compact and extended GCs. We find that compact GCs show a significantly smaller line-of-sight velocity dispersion, $\langle\sigma_{\rm cmp}\rangle\!=\!225\!\pm\!25$ km s$^{-1}$, than their extended counterparts, $\langle\sigma_{\rm ext}\rangle\!=\!317\!\pm\!21$ km s$^{-1}$. Considering the weaker statistical correlation in the GC $r_h$-color and the GC $r_h$-$R_{\rm gal}$ relations, the more significant GC size-dynamics relation appears to be astrophysically more relevant and hints at the dominant influence of the GC orbit distribution function on the evolution of GC structural parameters.' author: - | Thomas H. Puzia$^{1,2}$, Maurizio Paolillo$^{3,4,5}$, Paul Goudfrooij$^{6}$, Thomas J. Maccarone$^{7}$,\ Giuseppina Fabbiano$^{8}$, Lorella Angelini$^{9}$ title: 'Wide-Field Hubble Space Telescope Observations of the Globular Cluster System in NGC1399' --- Introduction ============ Structural Parameters of Extragalactic GCs ------------------------------------------ Wide-field studies of massive galaxies provide important benchmarks for comparisons with globular cluster (GC) formation and evolution models as well as GC system assembly in the context of galaxy formation scenarios, not only because they define homogeneous and uniform datasets but also due to their simultaneous sampling of galaxy core and halo regions where various different physical processes affect the GC formation and survivability. In general, GC formation is influenced by small-scale physics that governs star-formation and feedback processes [e.g. @murray92; @harris94; @ee97; @hartwick09; @murray09] while stellar feedback as well as internal and external dynamical mechanisms determine their early evolution [@gieles08; @bastianetal08; @fall09; @chandar09; @elmegreen10; @mapelli13] and the latter, ultimately, their fate [e.g. @gnedin97; @vesperini97; @vesperini03; @chandar10; @bekki10]. The vast dynamical parameter ranges that need to be probed to study the complex interplay of these processes with numerical simulations are still very challenging for today’s computers [e.g. @kravtsov05; @li05; @bournaud08; @griffen10; @schulman12; @greif12]. One simple approach to understand the influence of some of these processes on GC formation and evolution is the empirical study of GC structural parameters and their variation as a function of galactocentric distance. Detailed GC structural parameters, such as core, half-light, and tidal radius, as well as central surface brightness, concentration, ellipticity, etc. were, until the past decade, only accessible within the Local Group (LG) due to the limited spatial resolution of ground-based instrumentation [e.g. @king68; @ill76; @kontizas82; @elson85; @elson88; @elson91; @elson92; @crampton85; @demers90; @trager95]. The launch of the [*Hubble Space Telescope*]{} (HST) catapulted this field to a whole new stratum, making vast numbers of GC systems accessible to high spatial resolution studies [@harris13]. In fact, HST still provides our only access to high spatial resolution observations at optical wavelengths. Several pioneering HST works quickly reached out with their GC half-light radius measurements beyond the LG as far as the Fornax galaxy cluster at $\sim\!20$ Mpc distance [e.g. @elson94; @fusipecci94; @kundu98; @kundu99; @puzia99; @puzia00; @zepf99]. Numerous subsequent studies have used the superior spatial resolution of HST and the relatively large field of view ($\sim\!202\arcsec \!\times\! 202\arcsec$) of the [*Advanced Camera for Surveys*]{} (ACS) to collect large imaging datasets of extragalactic GC systems, the most homogeneous of which was obtained by the ACS Virgo and Fornax Cluster Surveys (ACSVCS and ACSFCS, see @cote04 and @jordan07, respectively). These observations set the baseline for systematic studies of GC structural parameters in the central regions of early-type cluster galaxies. There was quickly mounting consensus among the early HST investigations that the observed central GCs had a rather broad half-light radius distribution with a peak somewhere in the range $\sim\!2\!-\!3$ pc, which led to the suggestion that this peak value may be used as a geometric distance indicator [e.g. @kundu01; @jordan05]. Another important finding was that the blue GCs are on average larger than the red GCs. In particular, within the central regions of galaxies typically observed with HST, blue GCs show $\sim\!20\%$ larger mean half-light radii compared to the red GC sub-population [e.g. @kundu98; @kundu01; @kundu99; @puzia99; @puzia00; @zepf99; @larsen01; @jordan05; @spitler06; @harris06; @harris09; @harris10; @blom12; @goudfrooij12]. However, one major limitation of most previous HST studies, targeting extragalactic GC systems such as the ACS Virgo and Fornax cluster surveys, was their limited field of view, using only one HST/ACS pointing per galaxy. Because of this, these HST studies focused on the core regions of elliptical galaxies covering the inner few kpc (i.e. $\lesssim 1 R_{\rm eff}$). The outer parts of rich GC systems in central cluster galaxies were so far missed and mainly observed with ground-based instrumentation at much lower spatial resolutions [e.g. @rhode01; @rhode04; @rhode07]. The only other ground-based study featuring a wide field of view [*and*]{} high spatial-resolution was performed by [@gomez07] using Magellan/IMACS under exceptional $\sim\!0.5\arcsec$ average seeing conditions to measure half-light radii of 364 radial-velocity confirmed GCs in NGC 5128 (S0/E) out to $\sim\!8 R_{\rm eff}$ of the spheroid light and found no significant correlation between GC half-light radius and projected galactocentric distance, i.e. $r_h\!\propto\!R^0$, outside $\sim\!1 R_{\rm eff}$. However, @gomez07 reported that at $\lesssim\!1 R_{\rm eff}$ the red GCs show a steeper $r_h\!-\!R$ relation and on average 30% smaller sizes than blue GCs. Other studies using more than single-pointing HST observations conducted GC half-light radius measurements in NGC4594 (Sa) out to about $6\,R_{\rm eff}$ of the bulge component [@spitler06; @harris10 658 GC candidates], in NGC4365 (E) out to $\sim\!2.4 R_{\rm eff}$ of the spheroid light [@blom12 659 GC candidates], and in 6 giant elliptical galaxies out to $\sim\!4\!-\!5 R_{\rm eff}$ of their spheroids [@harris09 altogether 3330 GC candidates]. In the case of NGC4594, the inner red GCs are $\sim\!17\%$ smaller than the blue ones, but because of a steeper size-radius relation of the red GC sub-population, this difference becomes insignificant at galactocentric radii $\gtrsim\!2.7 R_{\rm eff}$ of the bulge light. NGC4365 hosts on average $\sim\!32\%$ larger blue GCs compared to their red counterparts and shows a steep size-radius relation, $r_h\!\propto\!R^{(0.49\pm0.04)}$, for the entire GC sample, similar to Milky Way’s GC system [@vdB91]. However, @blom12 do not investigate whether this relation differs between GC sub-populations as a function of projected galactocentric radius. The composite GC system of the six giant ellipticals studied by @harris09 exhibits a mild relation of the form $r_h\!\propto\!R^{0.11}$ and a $\sim\!17\%$ size difference between red and blue GCs that is independent of projected galactocentric radius. Astrophysical Implications -------------------------- In general, the finding of a size difference between blue and red GCs has important astrophysical implications for the understanding of the formation and evolution of GCs and for the usefulness of the peak value of the GC size distribution as geometric distance indicator. Several studies, such as [@larsen03], [@jordan04], and [@harris09] put forward models to explain the size difference between blue and red GCs. Inspired by the Milky Way GC system where a shallow relation exists between GC half-light radius and the 3-dimensional galactocentric distance, $r_h\!\propto\!R_{\rm 3D}^{0.5}$ [@vdB91], @larsen03 suggested that the GC size difference between red and blue GCs in massive ellipticals could be due to the difference in their spatial distribution functions. Typically, the red GC sub-population would be more centrally concentrated than their blue counterpart and therefore on average smaller, being tidally more truncated by the stronger host galaxy potential. However, [@webb12b] have shown in detailed numerical simulations that the observed GC size difference is unlikely due to projection effects alone. In contrast to this [*external effect*]{}, two alternative [*internal effects*]{} were put forward. Firstly, [@jordan04] suggested that the combined effect of mass segregation and shorter stellar lifetimes of more metal-rich stars at a given mass may explain the GC size difference. This was strictly valid under the assumption that the GC [*half-mass*]{} radius distribution would be independent of metallicity and that metal-poor and metal-rich GCs were of the same age, which may be at odds with observations [e.g. @puzia02; @marinfranch09; @goudfrooij12]. This scenario was further developed by [@sippel12] and [@schulman12] in direct-integration N-body simulations of young, low-mass clusters with and without initial mass segregation, the absence of which was found to be enhancing the GC size difference. [@downing12] performed Monte-Carlo N-body simulations of massive star clusters and found that significant numbers of massive stellar remnants, i.e. single and binary black holes would boost this GC size difference. Secondly, [@harris09] suggested that more metal-rich proto-GC clouds could cool more efficiently and therefore collapse into a more concentrated quasi-equilibrium state before forming stars than clusters formed from low-metallicity gas. Any of these three scenarios comes with limiting assumptions and is likely not the single cause for the measured GC size difference as the variety of results described above indicates. To provide a larger and statistically robust dataset to constrain GC sizes as a function of galactocentric radius, we embarked on a wide-field observing campaign covering a large area with an HST/ACS mosaic out to several effective radii ($>\!5 R_{\rm eff}$) of the diffuse spheroid light around NGC1399, the central galaxy in the Fornax galaxy cluster that hosts one of the richest ($\ga6000$ GCs; Specific frequency[^1] $S_N\approx5$) and most extended GC systems in the nearby Universe [@dirsch03; @faifer04; @bassino06]. A significant part of the outer-halo GC system of NGC1399 is located hundreds of kpc away from its host and is probing the transition regime between galaxy and cluster potential [@ferguson89]. At the same time, the formation efficiencies of these outer-halo, blue GCs appear to be higher than those of the inner red GCs, $S_N(\mbox{red})\!\approx\!3$ while $S_N(\mbox{blue})\!\approx\!14$ [@forte05]. Spectroscopic radial-velocity studies of hundreds of GCs established a very complex multi-component system with the blue GCs being kinematically distinct from the red GC sub-population the latter of which shows dynamics similar to that of the host galaxy diffuse stellar component. The blue GCs, on the other hand, seem to have been partly accreted from satellite galaxies [@schuberth10]. It is this large auxiliary kinematic dataset that makes the GC system of NGC1399 an ideal target for a wide-field, high spatial-resolution study with HST/ACS [in comparison to M87, e.g. @epeng09; @madrid09] as several hundreds of member stellar systems are robustly separated from the fore- and background in radial velocity space. In our previous works, we used the dataset from this paper to study the Low Mass X-ray Binary (LMXB) population and the correlation of their properties with GC structural parameters [@paolillo11; @dago13], as well as the GC selection techniques based on neural algorithms [@brescia12]. Here we focus on the properties of the GC system itself. Our present paper is organized as follows: in §2 we present the HST/ACS observations and discuss the details of sub-pixel dithering, §3 includes a description of the preliminary photometry that enters our structural parameter fitting code, which is introduced and thoroughly tested in §4. We present our results in §5, where we show the large-scale variations of GC structural parameters within NGC1399. We discuss the implications in §6 and conclude this work in §7. Observations ============ ![Illustration of the 3x3 mosaic of our ACS observations overplotted on a DSS-2 image. Individual tiles and the main galaxies in the field are labeled. The orientation of the image, which measures $20\arcmin\times 20\arcmin$, is indicated in the upper right corner.[]{data-label="fov"}](fig1_n1399ACSfov.pdf){width="8.7cm"} Field Coverage and Orientation ------------------------------ All observations were taken as part of the program GO-10129 (PI:Puzia) with the [*Advanced Camera for Surveys*]{} [ACS; @ford03] onboard the [*Hubble Space Telescope*]{} (HST) in November 2004 and April 2005. The pointings were arranged in a $3\!\times\!3$ ACS mosaic with a few arcseconds overlap between the individual tiles as illustrated in Figure \[fov\]. To maximize common-field coverage with other imaging and spectroscopy observations (i.e. [*Chandra*]{} X-ray imaging, see @paolillo11, and VLT ground-based spectroscopy) the entire mosaic is rotated with a position angle of about $-30\,^{\rm o}$ with respect to the meridian and centered on the coordinates: RA (J2000) $\!=\!03^{\rm h} 38^{\rm m} 28.62^{\rm s}$ and Dec (J2000) $\!=\!-35^{\rm o} 28\arcmin\ \! 18.9\arcsec$. Due to scheduling constraints the north, north-east, and north-west tiles were observed with a position angle $-30.467\,^{\rm o}$, while the other six tiles were taken at a position angle $149.552\,^{\rm o}$. The full mosaic covers roughly $10\arcmin\!\times\!10\arcmin$ arcminutes and extends out to a maximum projected galactocentric distance of 8.76 or $51.3\pm1.0$ kpc with respect to NGC1399 [adopting the distance $D\!=\!20.13\!\pm\! 0.4$ Mpc, see @dunn06 also @blakeslee09]. This corresponds to a projected coverage of $\sim\!5.2$ effective radii of the NGC1399 diffuse galaxy light [@RC3] and $\sim\!4.9$ core radii of the globular cluster system density profile[^2] [@schuberth10]. Our filter choice considerations included the optimization of throughput, detector sensitivity, high spatial resolution, and a well-defined transformation to a standard photometric system. The filter that optimally balances these effects is F606W and was used for all our exposures. The ACS [*Wide-Field Channel*]{} (WFC) spatial sampling of the point-spread function (PSF) is sub-critical at the wavelength of our observations (F606W $\approx4600\!-\!7200$Å). If not accounted for, this would introduce aliasing artifacts and significantly degrade the spatial information in the final images, thus hampering the measurement of globular cluster structural parameters at the distance of Fornax. Each tile was, therefore, observed in a single orbit in four dithered sub-exposures of 527 seconds to allow sub-pixel resampling (see below), yielding a total integration time of 2108 seconds. Data Reduction and Image Combination {#ln:drizz} ------------------------------------ The basic data reduction of each ACS/WFC dither set was performed by the ACS data pipeline CALACS [@hack03]. The reduction steps included subtraction of masterbias and masterdark images, correction for flat-field and gain variations, as well as elimination of bad pixels. [rl]{}\[!t\] Pattern type & ACS-WFC-DITHER-BOX\ Primary pattern shape & PARALLELOGRAM\ Pattern purpose & DITHER\ Number of points & 4\ Point spacing & 0.285\ Line spacing & 0.285\ Coordinate frame & POS-TARG\ Pattern orient & 30.155 deg\ Angle between sides & 145.82 deg\ Center pattern & NO For the dithered observations we adopted a slightly modified Hubble Ultra-Deep Field dither pattern, for which the dither parameters are provided for reference in Table \[tbl-dither\]. Note that this dither pattern is not designed to cross the ACS inter-chip gap, but to maximize the sub-pixel shift integrity over the full ACS/WFC field of view. In its shape it follows the UDF dither pattern with a 67% larger step size. Each set of four dithered frames was combined into a single image using the [MultiDrizzle]{} routine v.2.7.0 [@koekemoer02]. The software takes care of correcting the geometric field distortions which affect individual ACS exposures and projects all dithered images onto a common grid in which the rectified frames are averaged. The averaged image is then “blotted” back into each distorted frame to identify and clean cosmic rays and bad pixels/columns by means of comparison of input vs. averaged image [see @fruchter02]. No background subtraction was performed at this stage of data processing. The main background contribution in our fields is due to the NGC 1399 diffuse light and is correctly accounted for in the following structural profile analysis (see Sect. \[ln:profilefit\]). Similar to the GOODS and UDF datasets, we use the Gaussian drizzle kernel and set the pixel scale to 0.03/pix on the final combined images. This provides a super-Nyquist sampling of the PSF with a FWHM of $\sim\!0.08\arcsec$ at 6000 Å [see also @beckwith06]. [@rhodes07] find that this combination of Gaussian drizzle kernel and 0.03/pix output pixel scale gives minimal aliasing in the final images. [@jee07] argue that a Lanczos drizzle kernel with a 0.05/pix output scale reduces the PSF width by $\sim\!3\%$ compared to the Gaussian kernel, at the expense that the Lanczos kernel introduces “cosmetic artifacts in the regions where flux gradients change abruptly" [@jee07]. Since most of our target GCs are likely to have structural parameters at the resolution limit of HST we are expecting strong varying profile gradients for the most compact objects. We find that noise correlation between neighbouring pixels produces moiré patterns in the vicinity of bright objects and strong gradients [see also @rhodes07], but this affects only a few blended sources in our dataset. After these considerations and careful visual inspection of the drizzled images we therefore decide to use the Gaussian drizzle kernel with [pixfrac]{}=0.8 in the subsequent analysis. The combined field is illustrated in Figure \[pdf:full\] and has an effective field of view of 99.053 arcmin$^2$. ![image](fig2_N1399_total_lowres.pdf){width="17.5cm"} Using the [MultiDrizzle]{} software, we also produce weight and error maps representing the final error budget for each pixel, which account for all uncertainties in the reduction process, including bias, flatfield, drizzling, and aliasing effects. These weight and error maps enter the photometry and structural parameter analysis. We note that the high spatial resolution of our drizzled images safeguards them from crowding effects, even in the central regions of NGC1399, and reveals in every pointing a wealth of detail in object morphology as illustrated in Figure \[pdf:example\]. The Photometric Input Catalog {#ln:photin} ============================= Aperture Photometry and Astrometry ---------------------------------- To obtain a rough estimate of the total magnitudes of all detected sources we perform aperture photometry with the [SExtractor]{} package [@bertin96] and measure instrumental magnitudes in apertures of successively growing diameter, i.e. photometric growth curve analysis. We use the asymptotic limit of these curves to compute mean photometric corrections from finite aperture sizes to “infinity”. Our tests show that an aperture with $0.24\arcsec$ radius maximizes the signal-to-noise ratio (S/N) of the final photometry. Leaving out saturated objects and spurious detections we obtain a mean aperture correction for this optimal aperture size to an “infinite” aperture radius of $\langle\Delta_{\rm F606W}\rangle=-0.14$ mag (with a standard deviation $\sigma\!=\!0.22$ mag). We also measure the mean photometric correction from the standard 0.5 aperture radius to “infinity” $\langle\Delta_{\rm F606W}\rangle=-0.07$ mag (with a standard deviation $\sigma\!=\!0.14$ mag), which compares well with the suggested value from [@sirianni05] of $\langle\Delta_{\rm F606W}\rangle=-0.088$ mag. We follow the prescriptions of [@sirianni05] to calibrate our F606W “infinite”-aperture magnitudes $m_i$ to the broadband $V$-filter in the VEGAMAG filter system. We include second-order color terms from the synthetic model of @sirianni05 that are applicable for the color range $V\!-\!I\!>\!0.4$ mag and obtain the final photometric calibration equation $$\label{photcalib} V_{\rm F606W} = m_i + 26.331 + 0.340\, (V\!-\!I) - 0.038\, (V\!-\!I)^2,$$ where we assume a mean $V\!-\!I\!=\!0.95\!\pm\!0.1$ mag for our globular cluster candidates [see e.g. @peng06]. All frames have a minimum background flux level of $\sim\!40\, e^-$ per sub-integration which would correspond to a CTE correction of the order $\la 0.02$ mag across each WFC chip [@riess04; @vera07]. Since the average background level in all ACS mosaic tiles is higher than the minimum background flux we do not correct for this negligible photometric offset. The Galactic foreground extinction in the direction of NGC1399 is $E(B\!-\!V)\!=\!0.013$ mag [@schlegel98; @schlafly11], which translates into $A_{\rm F606W}\!=\!0.038$ mag using the [@schlegel98] reddening curve. The total uncertainty of the photometric calibration in Equation \[photcalib\] formally amounts to $\sim\!0.089$ mag. However, when we consider the small CTE corrections, a color mismatch of $\sim0.1$ mag in Equation \[photcalib\] for GCs with extreme $V\!-\!I$ colors, and potential differential reddening of $\sim\!0.05$ mag across the ACS mosaic field, we estimate that our final photometry is accurate to $\Delta V_{\rm F606W}\approx 0.1$ mag. It is important to note that at this point we are not concerned with achieving photometry of the highest possible quality but providing first-guess input catalogs for our profile fitting routine. To be able to match source detections taken with other telescopes we compute an absolute astrometric solution for each of the nine ACS tiles. We select 40 bright unsaturated stars distributed homogeneously over the entire mosaic and match their positions with those of stars from the USNO-B1 catalog[^3] [@monet03] to obtain the world coordinate solution (WCS) for each tile. The final WCS accuracy across the entire mosaic is $\sim0.2\arcsec$. Object Classification {#ln:gcc} --------------------- In the following we describe the object detection and classification schemes that were used to define a photometrically-selected globular cluster candidate (GCC) sample for which we later measure structural parameters (see Sect. \[ln:sp\]). On the drizzled stack images we measure object coordinates, the background level, Kron radius[^4], isophotal area, FWHM, ellipticity, position angle, and the [SExtractor]{} quality flag parameter of each detection that had at least 20 pixels approximately $1.6\,\sigma$ above the background noise, corresponding to S/N $\approx 6$. The error images, produced during the drizzle procedure, were used as weight maps in the detection process to account for the varying NGC1399 surface brightness. Rather than trying to find the optimal source parameters to select high-probability GCCs, we adjust our classification parameters to reject clearly extended and/or amorphous background objects and image artifacts. Visual inspection of the individual frames shows that a very reliable rejection of clearly extended background sources and image artifacts is provided by the following parameter cuts: $\Delta V_{\rm F606W}\!<\!0.1$ mag, Kron radius $r_k<0.21$, FWHM $<0.75$, ellipticity ($1-b/a)<0.8$ (see Figure \[gcsel\]). The ellipticity criteria are based on Local Group GCs [see also @jordan09], while the FWHM cut is set at about $\sim\!10\times$ the one of the stellar PSF, and in [@brescia12] we showed that using more restrictive criteria may result in losing extended GCs, such as $\omega$Cen. The photometric uncertainty cut is to ensure reliable fitting [approximately equivalent to @paolillo11], since at less conservative cuts the galaxy background begins to dominate (see Section \[ln:bckg\] and Figure \[gclf\]). Additional criteria are the [SExtractor]{} flag parameter set to $<\!4$, which excludes objects with incomplete and/or corrupted photometry apertures that are very close to the frame edges, and the total isophotal area limit of $\la\!6000$ pixel[^5], which eliminates particularly extended galaxies and saturated foreground stars. The final input catalog contains 6634 sources. We show the $V_{\rm F606W}$ luminosity function of all detected and selected objects in Figure \[gclf\]. We point out that the above selection criteria serve only as preparation of our sample for the next step of the analysis, i.e. the profile fitting routine and are intended to minimize human interaction during the fitting process. In particular, they do not affect our final results. ![image](fig4a_magkron.jpg){width="6.5cm"} ![image](fig4b_magerr.jpg){width="6.5cm"} ![image](fig4c_magfwhm.jpg){width="6.5cm"} ![image](fig4d_magell.jpg){width="6.5cm"} ![Luminosity distribution of all F606W detections around NGC1399 in our $3\!\times\!3$ ACS mosaic (solid open histogram) and the background galaxy contribution estimated from the HUDF based on the same object detection criteria (dotted open histogram). Our pre-selected sample that enters the structural parameter measuring routine is shown as solid red histogram. To illustrate the influence of the background galaxy population on the faint end of our input catalog sample we statistically subtract the background galaxy distribution (dotted open histogram) from our initial photometric sample (solid open histogram) and show the result as the hatched blue histogram. The corresponding result after filtering with the photometric pre-selection (see Sect. \[ln:gcc\] for details) is shown as solid green histogram. This is in remarkable agreement with the expected classic GC luminosity function with $M_V({\rm TO})\!=\!-7.5$ mag and $\sigma_{\rm GCLF}\!=\!1.4$ mag, which is indicated as dotted curve [e.g. @richtler03]. The top axis shows absolute magnitudes assuming a distance of $D\!=\!20.13$ Mpc.[]{data-label="gclf"}](fig5_gclf.pdf){width="8.9cm"} Estimating the Background Galaxy Contribution {#ln:hudf} --------------------------------------------- We estimate the contribution of the background galaxy population to the luminosity distribution of our input catalog (see Fig. \[gclf\]) by applying the exact same photometry procedure to the F606W observations that were obtained as part of the Hubble Ultra-Deep Field (HUDF) program [@beckwith06]. The HUDF observations were conducted in 112 sub-exposures spread over 56 orbits with a total integration time of 135320 seconds and give us excellent access to high-quality background galaxy photometry. We obtained the drizzled HUDF image and the corresponding weight map from the Hubble Data Archive as reduced higher-level science products[^6], which were produced with virtually identical multidrizzle parameters compared with our procedure (see Sect. \[ln:drizz\]). To avoid unnecessary profile fitting of the many extended and amorphous sources in the HUDF we use SExtractor to measure their MAG\_BEST magnitudes, corrected for Galactic foreground extinction $E_{(B-V)}\!=\!0.008$ mag, and plot the corresponding luminosity function in Figure \[gclf\]. Since the final HUDF frame covers 11 arcmin$^2$ with a spatial sampling of 0.03/pix, we, therefore, scale the galaxy background number counts by a factor of 9.005 to match the survey area of our ACS mosaic. The plot shows the remarkable similarity of the faint-end of the luminosity distribution of our sample with the background galaxy population, modulo a small difference at faint magnitudes $V_{\rm F606W}\!\ga\!26$mag, which is likely the manifestation of cosmic variance; e.g. there is a variation in the number of background galaxy clusters in our ACS mosaic field. This is consistent with the results of [@hilker99] and [@drinkwater00] who find a background galaxy cluster at $z\!=\!0.11$ behind the core of the Fornax galaxy cluster. Analysis {#ln:sp} ======== At the distance of Fornax ($20.13\pm0.4$ Mpc) one arcsecond spans 97.6 pc. On our drizzled ACS frames one pixel with the angular size of 0.03 subtends therefore 2.93 pc at the distance of NGC1399. This is similar to the typical half-light radius for Milky Way globular clusters [@harris96]. HST’s confusion limit, $\delta$, at the pivot wavelength of the F606W filter, $\lambda_p\! =\! 606$ nm, can be estimated via $\delta\! =\! 1.22\, \lambda_p/D$, where $D\!=\!2.4$ meter of the HST primary mirror. We obtain $\delta\!=\!6.2$ pc at the distance of Fornax. However, because we are fitting analytical, multiple-component 2-D surface brightness profiles, our nominal spatial resolution is much better than the computed confusion limit. The exact numerical value of the spatial resolution limit is determined through detailed artificial cluster experiments, which are discussed in Section \[ln:art\] in detail. Observations of the integrated-light profile $\Sigma({\mbox{\boldmath$r$}})$ of resolved astronomical objects measure their surface brightness variations $\mu({\mbox{\boldmath$r$}})$ over the 2-D spatial extent ($\vec{r}\!=\!{\mbox{\boldmath$r$}}$ for spherically symmetric sources) convolved with the instrumental point-spread function $P({\mbox{\boldmath$r$}})$ and the detector diffusion kernel $D({\mbox{\boldmath$r$}})$, plus, in the simplest case, an additive noise term $N({\mbox{\boldmath$r$}})$: $$\label{eq:obs} \Sigma({\mbox{\boldmath$r$}})=2\pi\int\limits_{r_1}^{r_2} \left\{\mu({\mbox{\boldmath$r$}}) \otimes P({\mbox{\boldmath$r$}}) \otimes D({\mbox{\boldmath$r$}}) +\! N({\mbox{\boldmath$r$}})\right\}{\mbox{\boldmath$r$}}d{\mbox{\boldmath$r$}}$$ where $\mu({\mbox{\boldmath$r$}})$ is the sum of the source and background surface brightness $\mu_{\rm s}({\mbox{\boldmath$r$}})\!+\! \mu_{\rm b}({\mbox{\boldmath$r$}})$. The access to surface brightness profiles of distant objects (e.g. globular clusters in NGC1399) is therefore limited by the spatial resolution of the data (i.e. the width of functions $P({\mbox{\boldmath$r$}})$ and $D({\mbox{\boldmath$r$}})$), the brightness of the sky (i.e. where $\mu({\mbox{\boldmath$r$}})\!\approx\!\mu_{\rm b}$), and the noise properties of the data (i.e. $N({\mbox{\boldmath$r$}})$). Among today’s imaging instruments that operate at optical wavelengths, the ideal case of $P({\mbox{\boldmath$r$}})\otimes D({\mbox{\boldmath$r$}})\!\rightarrow \!\delta({\mbox{\boldmath$r$}})$ and $N({\mbox{\boldmath$r$}})\!\rightarrow\!0$ is best approximated by HST. In particular, the ACS/WFC camera provides a large field of view ($\sim\!202\arcsec\!\times\!202\arcsec$) over which the geometric variations of $P({\mbox{\boldmath$r$}})\otimes D({\mbox{\boldmath$r$}})$ are relatively stable and well understood [@anderson05; @jee07]. An additional major advantage of HST observations is the very low sky background with a typical surface brightness $\mu_{{\rm b}, V}\!\ga\!22.5$ mag arcsec$^{-2}$ (see also ACS Instrument Handbook). King Surface Brightness Profile ------------------------------- The reason for the great success of the King profile in parametrizing the surface brightness profiles of most Galactic globular clusters is their structural homology, and is a simple consequence of the fact that virtually all of these systems have ages far in excess of their relaxation times [e.g. @king68; @ill76; @dacosta79; @kukarkin79; @chun80; @trager95]. We note here [*en passant*]{} that this might not be the case for more extended sources [e.g. @misgeld11]. The King profile [@king62], which is defined as $$\label{eq:king} \mu_{\rm K}({\mbox{\boldmath$r$}})=k\left[\left(1+\frac{{\mbox{\boldmath$r$}}^2}{r_c^2}\right)^{-\frac{1}{2}}\! - \left(1+\frac{r^2_t}{r_c^2}\right)^{-\frac{1}{2}}\right]^2,$$ describes the surface number density in the range $0\leq |{\mbox{\boldmath$r$}}| \!<\! r_t$ and is zero for ${\mbox{\boldmath$r$}}\! \geq\! r_t$. Its shape is governed by the core radius $r_c$, at which the projected surface density is half the central stellar surface density, which itself is set by the cluster gravitational binding energy ($r_c\!\approx\!3\sigma /\sqrt{4\pi G\rho_o}$ for $r_t/r_c\!\gg\!1$, see e.g. @binney87). If the GC is tidally filling, $r_t$ can be considered the tidal radius, otherwise $r_t$ marks the limiting radius beyond which the stellar density drops to zero; this is sometimes referred to as the King radius ($r_k$). The profile is normalized to the central surface brightness by $k\!=\!\mu(0)(1\!-\!1/\sqrt{1\!+\!r^2_t/r_c^2})^{-2}$. The family of King profiles is parametrized by the concentration $c\!=\!r_t/r_c$, which is directly proportional to the central potential $W_o$ via $c\simeq 9.12\! +\! (W_o\!-\!4.215)^{3.064}$ for $W_o\leq12$ (@king66, see also @binney87). The basic assumption of this parametrization is a truncated (so-called “lowered”) Maxwellian phase-space distribution of GC member stars in addition to the premise of orbital isotropy. It is assumed that the King profile is a valid description of the surface-brightness profiles of extragalactic GCs [e.g. @harris02; @harris10; @sharina05; @jordan05; @huxor05; @gomez06; @barmby07; @mclaughlin08; @masters10]. In other words, we assume a universal homology among globular clusters and adopt the King structural parameters as a sufficient set to describe their light profiles. However, we have to keep in mind that such objects may not be well represented by isotropic, single-mass, isothermal spheres but may be better described by other profiles. [@mclaughlin05] show that other profiles such as the [@wilson75] profile or power-law profiles à la [@elson87] fit the outermost parts of Milky Way and Magellanic Clouds GCs as well or better than classic King profiles. Furthermore, [@webb12a] compare King62 models fits to King66 [@king66], Wilson75 [@wilson75], and Sérsic models [@sersic68] for GCs in M87, and find that King66 models significantly underestimate cluster sizes, while Wilson75 fits are in close agreement with King62 measurements. However, we keep in mind that GCs outside the Local Group may have experienced different dynamical evolution histories given that their host galaxies may have undergone more violent merging and accretion histories [e.g. @baumgardt03] that may give rise to a larger variety of unusual GC surface-brightness profiles. Our analysis will necessarily be less sensitive to the outer low-surface brightness outskirts of the NGC1399 GCs than to their half-light or core properties. Since all the aforementioned profiles are virtually identical in their inner parts [i.e. within their half-light radius, see @mclaughlin05] we adopt the King62 profile for the rest of the analysis. The main reason is that for marginally resolved GCs the profile choices become rather unconstrained and more complex models often diverge or give degenerate results [@barmby07; @mclaughlin08; @harris10], whereas the King62 profile provides the most robust measures of GC structural parameters for both marginally resolved and well resolved targets. The Fitting Routine {#ln:profilefit} ------------------- To derive the structural parameters of NGC1399 GCs we fit their surface brightness profiles using a modified version of the GALFIT package that includes the King profile as a fitting option [v3.0, @cpeng10 and references therein]. Previous software packages such as [ishape]{} [@larsen99], [gridfit]{} [@mclaughlin05] and [kingphot]{} [@jordan05] offer valid alternatives for measuring GC structural parameters. However, [ishape]{} generally uses fixed King concentration parameters and deals with elliptical sources in a semi-analytical way. These three routines do not allow for flexible fitting of multiple blended sources with various profile types plus a variable background component. Additional advantages of our code is the execution handling and speed, which allows us to efficiently conduct large amounts of artificial cluster experiments (see below). ![image](fig6d_gcfit_data.pdf){width="5.2cm"} ![image](fig6e_gcfit_model.pdf){width="5.2cm"} ![image](fig6f_gcfit_resid.pdf){width="5.2cm"} We account for blended sources within the fit region of each GC and simultaneously fit profiles to sources which are less than five magnitudes fainter than the target within a radius of two FWHM, and to sources less than three magnitudes fainter outside this region. At the same time, we match the contributions of the sky+galaxy surface brightness by fitting a surface within the same area. The code uses a $\chi^2$ minimization scheme to simultaneously optimize the fit to each source and the local background surface brightness. Because some objects are blended with nearby very extended sources, we additionally use various profiles types for those blended objects, such as clearly extended nearby dwarf galaxies for which we choose the Sérsic profile [@sersic68]. Extended objects that have isophotal areas $\ga\!10\times$ larger than the fitting area are well approximated by a simple sloped background contribution. A representative example of the fit quality for a typical GC in NGC1399 is shown in Figure \[pdf:fit\]. Constructing the PSF Library ---------------------------- Equation \[eq:obs\] shows that detailed knowledge of the local PSF over the entire image is mandatory to obtain meaningful measurements of profile parameters, and the most realistic representation of the convolution product $P({\mbox{\boldmath$r$}})\otimes D({\mbox{\boldmath$r$}})$ is provided in form of a library of empirically measured PSFs [see discussion in @georgiev09a]. Such a library of [*effective*]{} PSF (ePSF) profiles based on repeated ACS observations of dense stellar fields was presented for several HST/ACS filters by [@anderson05] and [@anderson06]. Because of the fully empirical approach to build such a library [@anderson00], this collection provides the best characterization of the ACS/WFC-PSF for our purposes, as it preserves the variations of high [*and*]{} low-contrast features of the PSF with high spatial on-chip sampling. This is superior to the PSF modeling techniques provided by the TinyTim simulator[^7] and other parametric PSF approximations [@jee07], as well as building the PSF library from the science images themselves where the relative foreground stellar density is not sufficiently high to obtain a clean PSF star sample. The ePSF library provides a set of $10\!\times\! 9$ PSF profiles uniformly covering the WFC field of view. Each ePSF is oversampled by a factor of four to account for shifts of the source centroid with respect to the pixel center and applies only to the individual distorted ACS exposures (“flt” files). In order to transform the ePSFs into the final drizzled images, we need to apply our data reduction process to the library itself. To this end we designed a custom software package [[MultiKing[^8]]{}, see @paolillo11] to overlay the Anderson PSF grid onto a set of empty WFC frames, reproducing the actual data frame properties (orientation, dither pattern, astrometry, etc.). The grid positioning was modified on each frame to preserve the sky coordinates of each PSF, properly accounting for geometric distortions that affect the WFC “flt” frames, as would be expected for a real source within a set of observations taken with our dithering pattern. Since each dither pattern is executed with slightly varying sub-integration pointings, this procedure was applied to each individual pointing of the ACS mosaic. Finally, the dithered ePSF frames were combined together in the same way as the science frames, producing a [*drizzled effective PSF*]{} (drPSF) library for each individual ACS tile. The specific stellar PSF at a random location within our final images is chosen to be the nearest drPSF within the template grid. We use these drPSF libraries for the subsequent analysis. Our code was already implemented in the study of [@goudfrooij12] who successfully used the drPSF approach to measure star cluster sizes in NGC1316. Artificial Cluster Experiments {#ln:art} ------------------------------ ![image](fig7a_rc_sim.jpg){width="7.4cm"} ![image](fig7b_rc_corr.jpg){width="7.4cm"} ![image](fig7c_rt_sim.jpg){width="7.4cm"} ![image](fig7d_rt_corr.jpg){width="7.4cm"} ![image](fig7e_rh_sim.jpg){width="7.4cm"} ![image](fig7f_rh_corr.jpg){width="7.4cm"} Every attempt to determine the structural parameters of extragalactic GC is affected by measurement uncertainties, parameter covariance, and other inherent systematic characteristics of the dataset and measuring technique. To test the robustness of our measurements (under the assumption that the King62 profile describes the NGC1399 GC profiles sufficiently well) and probe parameter correlations and systematics we used our [MultiKing]{} code to create and add artificial star clusters to our ACS science frames and attempt to recover their structural parameters with our profile fitting routines using the exact same approach as for the analysis of NGC1399 GCs. This process includes convolving the appropriate drPSFs of the corresponding ACS tile with King profiles of varying structural parameters and inserting the noise-corrected clusters at random locations in the eastern, southern, and central tile of the ACS mosaic. In this way we include 1500 artificial clusters per tile in 15 runs each to avoid effects of artificial crowding. The input structural parameters cover a broad dynamic range that aims to sample crucial values around the resolution and confusion limits more densely. In particular, it covers the typical sizes of Galactic and LMC globular clusters. The recovery quality of the core radius, $r_c$, half-light radius, $r_h$, and tidal radius, $r_t$, is illustrated in Figure \[gc\_sim\]. For each parameter we show the input vs. output correlation, together with a ”sliding-median” probability density estimate and the corresponding 1-$\sigma$ contours as well as the error-of-the-mean margin. The renormalized profile fit quality serves as a metric to divide our artificial cluster sample in low- and high-quality fits, the division of which is done at the renormalized reduced chi-square ${\hat\chi}^2\!=\!1$. This division generally corresponds to faint and bright sources. The corresponding histograms in the left panels of Figure \[gc\_sim\] compare the input with the recovered parameter distribution and indicate biases in our measuring process. Our cluster experiments are consistent with the results presented in [@carlson01]. In particular, all our bona-fide sample GCs have an integrated S/N $\ga\!100$ in agreement with the minimum prescription of @carlson01 to measure sizes of marginally resolved GCs[^9]. In the following, we discuss and quantify these systematics to provide numerical estimates of the reliability of the subsequent structural parameter analysis. ### The Recovery Fidelity of the King Core Radius {#ln:fidel_c} The top panels in Figure \[gc\_sim\] show how our code recovers the King core radius, $r_c$. From the left panel, i.e. input vs. output $r_c$ diagram, it is evident that the spatial resolution of our dataset becomes increasingly poorer at $r_c\la3$ pc, and we see nicely a ”leveling off” of the relation towards smaller spatial scales. The reader should be aware that the logarithmic scaling in this plot is chosen to show exactly this physical limit and exaggerates this effect optically. In the right panel we plot the $r_c$ difference relation, in the sense $\Delta r_c = r_{c,{\rm out}} - r_{c,{\rm in}}$, vs. the recovered core radius $r_{c,{\rm out}}$. The graph illustrates that the measurements can be robustly corrected with a small systematic offset of the form $\langle\Delta r_c\rangle = 0.81\pm0.04$ pc with a mean standard deviation of $\bar{\sigma}_{\rm 1-30 pc}=1.54$ pc, which is equivalent to the overall $r_c$ measurement uncertainty in this range. While the $\Delta r_c$ trend around the spatial resolution limit allows an almost linear correction it is clear that the scatter in $\Delta r_c$ increases towards larger $r_c$, which is due to the confusion limit of the data, e.g. blended sources, sky background fluctuations, etc. At this end we see a higher-order systematic trend that cannot be approximated with a simple offset. We, therefore, use the correction function $$\label{eqn:rc} \phi_{r_c} = 0.731 - 5.563\cdot10^{-2} r_{c,{\rm out}} + 3.742\cdot10^{-3} r_{c,{\rm out}}^2$$ to fit the overall trend in $\Delta r_c$ as a function of $r_{c,{\rm out}}$ and correct our $r_c$ measurements for $r_{c,{\rm out}}\!\in\![1,30]$ pc. The function is shown as dash-dotted curve in the upper right panel of Figure \[gc\_sim\] and approximates the probability density curve very well in the range $r_c\!\approx\!2\!-\!20$ pc, which we consider as our high-confidence range for the King core radius measurements. ### The Recovery Fidelity of the King Tidal Radius {#ln:rhcorr} The tidal radius, $r_t$, probes the outskirts of the GC light distribution. Our tests recover $r_t$ with good accuracy in the range between $\sim\!10$ and 100 pc (see middle panels in Fig. \[gc\_sim\]). The mean residual is $\langle\Delta r_t\rangle\!=\!2.86\pm0.25$ pc with an average standard deviation of $\bar{\sigma}_{\rm 10-100 pc}=9.06$ pc. The lower limit is set by the starting value of our fitting routine which is ten times the initial core radius value, so that some clusters with a large core radius and a slightly larger tidal radius end up with an overestimated tidal radius, because the numerical convergence of the code for fits with very similar core and tidal radii is internally defined by the core radius. Note, that the tidal radius cannot be smaller than the core radius. For objects more extended than $\sim\!100$ pc we run into background confusion and fit degeneracy problems, introduced by nearby diffuse galaxy components and satellite objects (which are fit as described in Sect. \[ln:profilefit\]). Hence, the fits become poorly defined beyond such large tidal radii, simply because there is not enough signal-to-noise in the low surface-brightness wings of the profiles. We approximate the corresponding residual trend with the following two component correction function $$\label{eqn:rt} \phi_{r_t}= \begin{cases} 0.646 + 22.458\, (r_t\!-\!5.581)^{-0.86} & \hspace{-0.15cm} \text{if $r_t\!\in\! [10,25)$,} \\ 2.078 + 6.675\cdot10^{-3}\, (r_t\!-\!1.11) + \\ + 2.591\cdot10^{-4}\,(r_t\!-\!1.11)^2 & \hspace{-0.15cm} \text{if $r_t\!\in\![25,100]$,} \end{cases}$$ which is valid in $r_t\!\in\![10,100]$ and robustly follows the probability density estimate out to the extreme edges of the parameter range. ### The Recovery Fidelity of the Half-Light Radius {#ln:rhfidelity} The GC half-light (or effective) radius, $r_h$, is a structural parameter that emerges from the correlation of the King core and tidal radius, as described by Equation \[eq:king\] and encircles 50% of the total GC light. The half-light radius is relatively stable throughout the GC dynamical evolution and is predicted to evolve much slower with time ($r_h\!\propto\! t^{2/3}$) than the tidal and core radius [@henon73; @henon75; @elson87rev; @murphy90; @murray92]. One major advantage of $r_h$, is its relatively effortless accessibility in more distant stellar systems and because of its slow evolution it provides the most reliable measure of the true size distribution function of extragalactic GC system. Formally, the GC half-light radius, $r_h$, is defined as $$\label{rhdef} 2\pi\int\limits_{0}^{r_h}\mu({\mbox{\boldmath$r$}}){\mbox{\boldmath$r$}}d{\mbox{\boldmath$r$}}= \pi\int\limits_{0}^{\infty}\mu({\mbox{\boldmath$r$}}){\mbox{\boldmath$r$}}d{\mbox{\boldmath$r$}}\; ,$$ and can be evaluated with the integral form of the King profile which can be written as $$\begin{aligned} \label{kingint} \lefteqn{2\pi\int\limits_{r_1}^{r_2}\mu_{\rm K}({\mbox{\boldmath$r$}}){\mbox{\boldmath$r$}}d{\mbox{\boldmath$r$}}= \left[ \frac{kr_c}{\varrho}\left\{\varrho\,\mbox{arctan}\left(\frac{{\mbox{\boldmath$r$}}}{r_c}\right) + \right. \right.} \\ &\left. \left. r_c\left({\mbox{\boldmath$r$}}- 2\sqrt{\varrho}\, \ln\left(2{\mbox{\boldmath$r$}}\sqrt{\varrho} + 2r_c^2\sqrt{1 + \frac{r_t^2}{r_c^2}} \sqrt{1 + \frac{{\mbox{\boldmath$r$}}^2}{r_c^2}}\right)\right)\right\}\right]_{r_1}^{r_2} \nonumber\end{aligned}$$ where $\varrho=r_c^2 + r_t^2$. Since the half-light radius $r_h$ cannot be written in a closed analytic form from Equations \[rhdef\] and \[kingint\], it has to be evaluated numerically. Hence, we determine $r_h$ from the direct numeric integration of the King profile for each individual cluster and, thus, probe immediately the influence of parameter correlations between $r_c$ and $r_t$ on the integrated luminosity. The bottom panels of Figure \[gc\_sim\] show that the average bias of the half-light radius is $\langle\Delta r_h\rangle\!=\!0.87\pm0.02$ pc with an average standard deviation of $\bar{\sigma}_{\rm 1.5-15 pc}\!=\!0.54$ pc. This is in excellent agreement with the results of [@harris09] who found $\sigma_{r_h}\!=\!1.1$ pc as mean uncertainty for size measurements of GCs at the distance of $\sim\!40$ Mpc, based on similar data of distant BCGs which are roughly twice as far away as NGC1399. The reduced mean uncertainty of $r_h$ is likely due to parameter correlations between $r_c$ and $r_t$, the uncertainties of which compensate each other to leave $r_h$ a very reliable parameter of GC size. The results of our artificial cluster experiments indicate that we can measure [*and*]{} correct $r_h$ reliably for $r_h\in[1.5,19]$ pc. We compute the corresponding correction function of the form $$\label{eqn:rh} \phi_{r_h}= \begin{cases} 0.33 + (r_h\!-\!0.354)^{-4.26} & \text{if $r_h\!\in\![1.5,7)$,} \\ 0.449 - 5.766\cdot10^{-2}\, (r_h\!-\!0.1) + \\ + 5.869\cdot10^{-3}\, (r_h\!-\!0.1)^2 & \text{if $r_h\!\in\![7,19]$.} \end{cases}$$ In summary, we now understand the fidelity and limitations of our structural parameter measurements and move on to test the influence of the variable background surface brightness in our ACS mosaic. ### The Influence of the Variable Galaxy Background {#ln:bckg} ![Illustration of the residuals around the correction functions (see Equations \[eqn:rc\], \[eqn:rt\] and \[eqn:rh\]) as a function of background surface brightness. High- and low-quality fits are depicted as grey and black dots, respectively, and are defined as fits with a reduced $\chi^2$ below and above unity. Mean residuals and dispersion are given in each panel. [@caon94] measures the $B$-band surface brightness profile of NGC1399 out to $\sim\!15\arcmin$ galactocentric radius. We use the numbers from [@sandage75] who reports $B\!-\!V\!\simeq\!0.95$ to $0.98$ mag in the range $54.3\arcsec\!- 107.8\arcsec$ from the center of NGC1399 to obtain a rough estimate of the $V$-band surface brightness.[]{data-label="pdf:residbkg"}](fig8_cth_residuals.jpg){width="8.9cm"} After correcting for biases in our measuring procedure we explore in the following the influence of the variable galaxy surface brightness in the studied field on our structural parameter measurements. To do so we compute the residuals with respect to the correcting functions $\phi_{r_i}$ in Figure \[gc\_sim\] in the form $\delta r_i\!=\! (r_{i,{\rm out}}\! -\! r_{i,{\rm in}}) - \phi_{r_i}$ where the index $i$ stands for the core, tidal, and half-light radius, respectively. The residuals are then plotted in Figure \[pdf:residbkg\] as a function of the background counts. All our measurements are shown in this figure, however, only the values within the confidence limits of Equations \[eqn:rc\], \[eqn:rt\] and \[eqn:rh\] are considered in the computation of the mean residual statistics. This exercise shows that our GC profile fitting routine accounts very robustly and without any significant residual systematic for the variable background. Our tests probe surface brightness levels fainter than $\mu_{V_{\rm F606W}}\ga21.4$ mag, which corresponds to galactocentric radii $r\ga 30\arcsec$, and we expect that the final measurements are reliable without further corrections to within the quoted uncertainties of our artificial cluster experiments within the above $\mu_{V_{\rm F606W}}$ range. Comparison with ACSFCS Measurements {#ln:litcomp} ----------------------------------- ![Comparison of GC half-light radius measurements from this work and the ACS Fornax Cluster Survey [ACSFCS, see @jordan07; @masters10] for the same target GCs in NGC1399. The grey line shows the one-to-one relation. The greyscale parametrizes the ACSFCS measurement uncertainties, $\Delta r_h$, which are computed as the square root of the square sum of the individual uncertainties in the F475W and F850LP filters. The red line is a third-order polynomial approximation to the data.[]{data-label="pdf:litcomp"}](fig9_compare_fcs_all_conc.pdf){width="8.9cm"} In the following we compare our measurements to the most recent GC half-light radius measurements in NGC1399 based on the ACS Fornax Cluster Survey [ACSFCS, see @jordan07] which observed the galaxy with one central pointing. The ACSFCS GC half-light radius measurements were conducted with the [kingphot]{} software [@jordan05] and are restricted to the brightest GCs with $z\!\leq\!23.35$mag and colors $0.6\!\leq\!(g\!-\!z)\!\leq\!1.7$ mag. The ACSFCS data are comprised of $2\times565+90$sec exposures in F850LP and $2\!\times\!380$sec exposures in F475W. Compared to our $4\!\times\!527$sec, optimally dithered F606W observations, the ACSFCS data have, therefore, a somewhat lower S/N at an equivalent GC luminosity (due to lower system throughput in F475W and F850LP) and have a more sparsely sampled PSF due to their 2-step dither pattern [@jordan05; @jordan07]. We take the ACSFCS GC half-light radii published as part of the study presented in [@masters10] and use the arithmetic mean of their GC half-light radius measurements in the F475W and F850LP filters and select only GC candidates that were assigned a GC probability of $p_{\rm GC}\!\geq\!0.5$ [see also @jordan09]. Figure \[pdf:litcomp\] shows the direct comparison between the two samples where we find no significant offset beyond $r_h\!\approx\!2$ pc. However, at smaller half-light radii, the influence of the correction function from Equation \[eqn:rh\] (see also Fig. \[gc\_sim\]) becomes increasingly apparent as the ACSFCS data tend to be biased towards larger values relative to our measurements. This is primarily due to the fact that the ACSFCS measurements are not corrected for measurement systematics by means of artificial cluster experiments as in our procedure (see Section \[ln:rhfidelity\]). We parametrize the grey shading of data points in Figure \[pdf:litcomp\] with the measurement uncertainties, $\sigma_{r_h}$, of the ACSFCS data. The main trend of the comparison is approximated by a third-order polynomial and depicts the shape of the correction function. The rms around this relation is 0.77 pc, and with our measurement uncertainty of 0.54 pc from the artificial cluster experiments described in section \[ln:rhcorr\], we obtain $\Delta_{\rm total}\!=\!\sqrt{\sigma_{r_h}({\rm ACS})^2 - \sigma_{r_h}({\rm this\ work})^2}\approx 0.55$pc, which we regard as the total statistical uncertainty when comparing individual GC half-light radius measurements from various studies using different techniques. Results ======= We use our surface brightness profile (SBP) fitting routine described in Section \[ln:profilefit\] to measure the structural parameters of all sources from our photometric input catalog (see Section \[ln:photin\]) and calibrate them with the correction functions $\phi_{r_i}$ derived in Section \[ln:art\] (see Equations \[eqn:rc\], \[eqn:rt\] and \[eqn:rh\]). Because of the higher measurement fidelity of the half-light radius, we use $r_h$ in the subsequent analysis and refer to it as GC size, unless stated otherwise. We point out that for addressing other specific scientific topics, such as measuring the binary star formation efficiency via LMXB population analysis, other parameters such as the core radius and central surface brightness proved to be more diagnostic than $r_h$ [@paolillo11]. Total Object Magnitudes ----------------------- ![Comparison of total object luminosities determined via aperture photometry and direct integration of their surface brightness profile. Both magnitudes are in the Vega system and are corrected for Galactic foreground reddening with $A_{\rm F606W}\!=\!0.038$ mag. The top ordinate indicated the absolute $M_V$ at the distance of Fornax. A blue dashed line shows the equality relation.[]{data-label="pdf:apint"}](fig10_ap_vs_int_mag.jpg){width="8.9cm"} To investigate correlations of structural parameters with GC brightness it is important to compute accurate total luminosities for our object sample. In Section \[ln:gcc\] we corrected our aperture photometry with a generic aperture correction term to compensate for the light outside the $r\!=\!0.24\arcsec$ photometry radius which delivered the highest photometric S/N and served as a first guess for the structural parameter fitting routines. With our structural parameter measurements in hand we can now directly integrate the surface brightness profile of all targets and determine their total luminosities and compare them with the traditionally determined aperture magnitudes. Figure \[pdf:apint\] shows the direct comparison of total Vega magnitudes measured via corrected aperture photometry and integrated SBP luminosities. Small and large dots are defined as in Figure \[gc\_sim\] for low and high-quality fits of the surface brightness profile. Figure \[pdf:apint\] shows that the vast majority of our sample aligns very well with the one-to-one relation, which is due to the fact that most of our sample objects are marginally resolved GCs. Clearly resolved objects scatter to the right of the one-to-one relation and have brighter integrated magnitudes and have too faint aperture photometry counterparts. Their total aperture magnitudes at the GCLF turnover $M_V({\rm GCLF})\!\simeq\!-7.5$ mag are up to $\sim\!0.5\!-\!2$ mag fainter than the corresponding integrated SBP luminosities. In general, this is due to an average correction that is applied to all GCs when measuring GC luminosities via aperture photometry. This is direct evidence that for partially resolved and clearly resolved objects an average aperture correction term is not sufficient to determine their total luminosities. Note also that there are virtually no outliers left of the one-to-one relation which is visual assurance of our SBP fitting quality. The study of [@kundu08] has previously claimed that certain GC parameter correlations, such as the color-luminosity relation for bright blue globular clusters may be the result of inappropriately applying average aperture corrections to multi-passband photometry object samples with widely varying structural parameters. Although our structural analysis is based on the F606W filter only, to avoid such problems in what follows we use the directly integrated SBP magnitudes for the subsequent analysis, and point to the works of [@epeng09] and [@harris09] for a more detailed discussion of this filter-dependent aperture correction issue. Radial Velocity Information {#ln:vrad} --------------------------- ![Radial velocity distribution of objects towards the core regions of the Fornax cluster for which [@schuberth10] provide $v_{\rm helio}$ measurements. The open histogram are all GCs with such measurements and the shaded histogram shows the $v_{\rm helio}$ distribution of matched GCs for which we measured structural parameters. The hatched histogram shows the $v_{\rm helio}$ distribution of foreground stars from [@schuberth10]. The solid and dotted red curves show the probability density estimates to the entire GC sample together with their 90% confidence limits.[]{data-label="pdf:vrad"}](fig11_vrad.pdf){width="8.9cm"} In the following we use radial velocity measurements from [@schuberth10] to define a clean GC sub-sample which is consistent with the systemic velocity and GCS velocity dispersion in the center of Fornax. Figure \[pdf:vrad\] shows the distribution of heliocentric radial velocities, $v_{\rm helio}$, for foreground stars and [*bona-fide*]{} GCs, as well as the sub-sample of GCs for which we measured structural parameters. We match 306 out of the 790 GCs for which [@schuberth10] provide $v_{\rm helio}$ values that have structural parameter measurements from our analysis. Most of the remaining objects are at larger galactocentric radii and a small fraction has bad profile fits due to detector edge effects and confusion with very bright nearby sources. The distribution of the matched GCs, illustrated in Figure \[pdf:vrad\], shows that they representatively sample the total radial velocity distribution of the @schuberth10 sample. We also match nine stars out of the 236 confirmed by @schuberth10 (see the hatched histogram around $v_{\rm helio}\approx0$ km/s in Figure \[pdf:vrad\]) and study the distribution of structural parameters of false positives introduced by the foreground stellar population. GC Half-Light Radius as a Function of Luminosity {#ln:rh_lum} ------------------------------------------------ Before analyzing GC size variations as a function of GC color and galactocentric radius we need to make sure that potential observational biases are not influencing our result. One such bias is a correlation between GC size and luminosity; such a correlation can introduce systematics in the size distribution function for photometrically selected samples due to changing $M/L$ ratios for stellar populations with different ages and/or metallicities. The population synthesis models of [@bc03] and [@maraston05] give roughly a factor two difference between the stellar $(M/L)_V$ ratios for 13 Gyr old stellar populations with metallicities \[Z/H\] $\!=\!-1.5$ and $-0.5$ dex, which roughly correspond to the mean metallicities of the GC sub-populations in the Milky Way and other massive spiral and elliptical galaxies [e.g. @peng06]. For a magnitude-limited sample such a $M/L$ difference would correspond to a $\sim\!0.75$ mag offset in completeness for a uniformly old GC population. We plot the GC size, i.e. half-light radius $r_h$, versus luminosity in Figure \[pdf:rh\_intmag\]. Running median curves with their 90% percentile limits show that there is no indication for any significant GC size-luminosity relation for the entire GC sample. At a constant $M/L$ ratio this corresponds to $L\!\propto\!r_h^3\rho$ and implies, therefore, that the stellar density is directly proportional to the GC size, i.e. $\rho\!\propto\!r_h^{-3}$. Linear and quadratic least-square fits (dashed blue lines in Fig. \[pdf:rh\_intmag\]) do not show any significant slopes for the entire sample and neither linear or higher-order fits are statistically preferred over one another. ![GC half-light radius as a function integrated $V_{\rm F606W,0}$ luminosity for outer clusters ([*top panel*]{}), inner clusters ([*middle panel*]{}) and the entire GC sample ([*bottom panel*]{}). Black dots are resolved GC while grey dots mark unresolved objects. Solid green curves show the sliding-median trends of resolved data together with their 90% percentile limits. Dashed curves are the corresponding trends for all objects. Linear and quadratic least-square fits to the resolved cluster data are shown as long-dash and short-dash lines, respectively. The shaded region at faint luminosities ($V_{\rm F606W,0}\!>\!25.5$ mag) indicates the region where the photometric pre-selection becomes significantly incomplete. Cyan data mark [*bona-fide*]{} GCs confirmed by their radial velocity.[]{data-label="pdf:rh_intmag"}](fig12_rh_intmagc.jpg){width="8.9cm"} Splitting the entire GC sample at a projected galactocentric radius of $R_{\rm gal}\!=\! 20$ kpc into an ’inner’ and ’outer’ sub-population, we spot a few interesting trends. Firstly, at intermediate luminosities ($22\!\la\!V_{\rm F606W,0}\!\la\!24.0$ mag) the ’inner’ sample contains fewer extended GCs with half-light radii $r_h\!\ga\!4$ pc than the ’outer’ sample. This is ruled out to be due to the varying galaxy background and/or completeness (see Section \[ln:bckg\]) as well as due to lower statistics in the galaxy center (as there are actually more GCs), and might be due to the preferred disruption of extended GCs in the inner regions of NGC1399. The fact that we see virtually no extended GCs more massive than the GCLF turn-over at $V_{\rm F606W,0}\approx24.0$ mag indicates that disruption or tidal limitation (see Section \[txt:innerouter\]) may occur more frequently for low-mass GCs and that high-mass GCs are more prone to dynamical friction and orbital decay [e.g. @lotz01]. Secondly, we observe a weak indication for a size-luminosity relation for GC brighter than $V_{\rm F606W,0}\approx22.0$ mag, predominantly for the ’inner’ sub-sample. We fit this sub-sample separately with a linear relation that yields a significant slope of $r_h\!\propto\!(-0.6\pm0.2)\, V_{\rm F606W,0}$ which is reminiscent of the transition from the GC regime without any size-luminosity relation below $M_\star\approx10^6 M_\odot$ to the size-stellar mass relation ($r_h\propto M_\star^{0.8}$) of more massive compact stellar systems such as UCDs [e.g. @taylor10; @misgeld10; @misgeld11]. This relation is indicated in the middle panel as a thin red curve and is a good representation to the general trend of the data. We point out that the sub-sample of confirmed GCs with $v_{\rm helio}$ measurements (cyan dots in Fig. \[pdf:rh\_intmag\]) is consistent with this trend. Despite the fact that we detect a weak size-luminosity relation for massive GCs we stress that the majority of our sample, in particular the intermediate-luminosity to faint-end part, does not show any such relation. We are therefore safe to apply a simple magnitude cut to our data without introducing systematics in the GC size-color relation which we discuss in the following. GC Half-Light Radius as a Function of Color {#ln:sizecol} ------------------------------------------- ![Globular cluster half-light radius, $r_h$, as a function of their photometric color. [*Top panel*]{}: The bottom sub-panel shows the distribution of half-light radii for all GCs with ground-based $(C\!-\!R)_0$ color information from [@schuberth10]. The two upper sub-panels show the size distribution divided in projected galactocentric distance at $R_{\rm gal}\!=\!20$ kpc for the inner and outer samples. [*Bottom panel*]{}: GC half-light radii as a function of their $(g\!-\!z)_0$ color based on HST photometry taken from [@kundu08]. Solid thick and thin curves illustrate the running median and the $1\sigma$ limits, respectively, of the $r_h$ distribution in each sub-panel. We show the unresolved clusters as grey dots, which are not considered in computing the solid curves. Including those unresolved sources results in corresponding relations shown as dash-dotted curves. Cyan data illustrate [*bona-fide*]{} GCs confirmed by their radial-velocity.[]{data-label="pdf:rh_color"}](fig13_rh_color.jpg){width="8.9cm"} We add photometric color information to our GC size measurements and search the $C\!-\!R$ color database presented in [@schuberth10] to find 1811 sources that match our final catalog within 1 matching radius. The @schuberth10 photometric catalog is a combination of 1) the [@dirsch03] Washington photometry, obtained for one central pointing with a field of view of $36\arcmin\times36\arcmin$ using the MOSAIC camera on the CTIO-Blanco 4m telescope and 2) the photometry from [@bassino06] which covers additional fields in the outskirts around NGC1399, also imaged with the MOSAIC camera. In addition, we combine our GC size measurements with the HST photometry of [@kundu08] and find 1258 matches within 1 search radius, all of which are within the field of view of only one central HST pointing. These two datasets have very different completeness limits and spatial resolution characteristics so that we use them only to search for differential trends in each dataset separately. We note that based on the ACSVCS data, [@jordan05] have demonstrated that the mean trend of increasing half-light radius towards bluer GC colors does not strongly depend on the host galaxy $(g\!-\!z)_{\rm gal}$ color, except for the very bluest galaxies with $(g\!-\!z)_{\rm gal}<1.52$ mag where the GC size difference appears to vanish (see more detailed discussion in Sect. \[ln:discussion\]). In that sense, the GC system of NGC1399 should be representative for most massive galaxies. In Figure \[pdf:rh\_color\] we show the trends of GC size, i.e. half-light radius $r_h$, versus $C\!-\!R$ color from the MOSAIC study and the $g\!-\!z$ color from the HST central pointing and find significant trends in both colors of increasing GC sizes towards bluer GC colors. This is a different depiction of the well-known size difference between blue and red GCs discussed in previous studies [e.g. @kundu98]. For both photometry samples of resolved clusters we find $r_h\!\propto\!(-0.44\pm0.15)\times(g\!-\!z)$ for the HST data and $r_h \!\propto\! (-0.78\pm0.15)\!\times\!(C\!-\!R)$ for the wide-field MOSAIC sample. For the entire MOSAIC sample the average $r_h$ gradient corresponds to a mean size difference of $\sim\!15\%$ between the peak colors $C\!-\!R=1.3$ and 1.8 mag. Since the MOSAIC data cover a wide field of view we determine the GC size variation as a function of color for two sub-samples split at 20 kpc in projected galactocentric distance into an ’inner’ and ’outer’ sample. We find that the $r_h$ gradient is stronger for the outer sample \[i.e. $r_h\!\propto\!(-0.89\pm0.22)\times(C\!-\!R)$\] compared to the inner variation \[i.e. $r_h \propto (-0.65\pm0.22)\times(C\!-\!R)]$, which corresponds to a physical size variation of $\sim\!12\%$ and $\sim\!17\%$, respectively. If we use only radial-velocity confirmed GCs we obtain a much more significant $r_h$ change, namely $r_h\!\propto\!(-0.52\!\pm\!0.50)\!\times\!(C\!-\!R)$ for the inner and $r_h\!\propto\!(-1.36\!\pm\!0.45)\!\times\!(C\!-\!R)$ for the outer MOSAIC sample. This corresponds to a physical difference of $\sim\!10\%$ and $\sim\!23\%$, respectively. The overlap between the radial-velocity information and the HST photometry sample from @kundu08 is too small to derive any robust $r_h$ gradient values. However, for illustration purposes we mark all [*bona-fide*]{} GCs confirmed by their $v_{\rm helio}$ as cyan dots in Figure \[pdf:rh\_color\] and find no significant differences in their GC size-color distributions down to the limiting magnitude of $V\!\approx\!23.5$ mag, which marks the typical limit of spectroscopic studies. GC Half-Light Radius as a Function of Projected Galactocentric Radius {#ln:sizediff} --------------------------------------------------------------------- ![image](fig14_rh_galrad.jpg){width="14.5cm"} Thanks to the wide field coverage of our ACS mosaic we are now in the position of determining the change of the [*classic*]{} size difference between blue and red GCs as a function of projected galactocentric radius, $R_{\rm gal}$, in much greater detail. To begin with, we use the photometric parameters summarized in Table \[tab:col\_sel\] to define the blue and red GC sub-sample. The top panel of Figure \[pdf:rh\_galrad\] shows the corresponding GC size versus $R_{\rm gal}$ distribution for all GC candidates. Taking the entire GC sample for which structural parameters were measured and calibrated (see Section \[ln:sp\]), we observe several interesting regimes with constant and gradually changing GC sizes. Firstly, GCs in the inner $\sim\!10$ kpc become on average larger as a function of $R_{\rm gal}$, while GCs at larger galactocentric distances ($\ga\!10$ kpc) show no significant GC size-$R_{\rm gal}$ relation. This is illustrated by the black curves which depict the sliding median together with error-of-the-mean margins. Secondly, plotting the median size trends for the blue and red GC sub-population separately (middle panel of Fig. \[pdf:rh\_galrad\]) reveals the well known GC size difference of $\sim\!20\%$ in the central parts of NGC1399, i.e. $R_{\rm gal}\la10$ kpc [e.g. @kundu98; @jordan05]. Except for the range $R_{\rm gal}\!\approx\!14\!-\!20$ kpc, this size difference prevails at large galactocentric distances out to $\sim\!30\!-\!40$ kpc. The bottom panel shows the ratio of the median GC sizes for blue and red clusters in the sense ${\rm med}(r_{\rm h, red})/{\rm med}(r_{\rm h, blue})$. This mean ratio for the whole $R_{\rm gal}$ range is $0.82\pm0.11$. [*The existence of a GC size difference at large $R_{\rm gal}$ is direct evidence that this difference cannot be solely due to a projection effect as suggested by [@larsen03]. Instead, it has to have its origin in at least one other internal or external parameter that determines the GC size and/or its evolution. The simulations of [@sippel12] suggest that this size difference is mainly due to GC internal evolution related to the impact of metallicity effects on stellar evolution combined with the GC dynamical evolution under the influence of mass segregation.*]{} In the middle panel of Figure \[pdf:rh\_galrad\] we show the comparison with the GC size-$R_{\rm gal}$ relations for blue and red GCs in M87 as derived by [@madrid09]. Similar conclusions have been reached by [@paolillo11], [@blom12], and [@webb12b]. Within the $R_{\rm gal}$ coverage of the single central ACS pointing that these authors have used for their analysis, the agreement between their M87 and our NGC1399 GC size trends is remarkably good. ![image](fig15_rh_galrad_Jacobi2.pdf){width="17.5cm"} ![image](fig16_rh_distrib_colsel2.pdf){width="15cm"} Discussion {#ln:discussion} ========== The Inner vs. Outer GC System of NGC1399 {#txt:innerouter} ---------------------------------------- The significant GC size-luminosity relation of the inner $10$kpc which disappears in the outer regions may indicate a transition in the predominance of various mechanisms at different galactocentric radii that shape the GC sizes and thus their evolution as a system. Since the transition does not depend on GC color, i.e. blue and red GCs show that same $r_h-R_{\rm gal}$ relation, external dynamical effects are the most probable explanation (e.g. dynamical friction of massive GCs that quickly sink into the core regions of the inner galaxy, tidal harassment of low-mass GCs by dwarf haloes in the outer halo regions, etc.). While detailed numerical modelling of these effects goes beyond this work, we point out that our dataset is ideal to conduct detailed analyses such as those presented in [@vesperini03] and [@webb12a; @webb13]. We note that the mean $r_h$ for all resolved sources within $20\arcsec\!<\!R_{\rm gal}\!<\!120$ is $1.95\pm0.06$ pc, i.e. significantly smaller than the mean value for the entire GC system of $\langle r_h\rangle\!=\!3.21\pm0.07$ pc. [lccc]{}\[!t\] Ground-based & $T1<23$& $T1<23$ & (1)\ data & $1.0\leq C\!-\!R<1.65$ & $1.65\leq C\!-\!R<2.2$ &\ \ HST data & $z<22.5$ & $z<22.5$ & (2)\ & $1.3\leq g\!-\!z<1.9$ & $1.9\leq g\!-\!z<2.5$ &\ To test whether the stellar mass distribution in NGC1399 is sufficient to produce the GC $r_h\!-\!R_{\rm gal}$ trend (see Figure \[pdf:rh\_galrad\]) we use the surface brightness profile data obtained as part of the Carnegie-Irvine Galaxy Survey [CGS, see @ho11; @li11] to compute the local [*instantaneous*]{} Jacobi radius of GC ($r_J$) as a function of galactocentric radius out to $\sim\!280\arcsec$ (i.e. $\sim\!28$kpc) which corresponds to the maximum sampling radius of CGS. The Jacobi radius marks the point at which the gravitation forces exerted on GC member stars due to the GC potential and that of its host galaxy are equal but opposite in direction. The Jacobi radius can be expressed as $$r_J=R_{\rm gal}\left(\frac{m_{\rm GC}}{2{\cal M}_{\rm gal}}\right)^{1/3} \label{eq:jacobi}$$ and is a robust representation of the [*instantaneous*]{} GC tidal radius that is induced by the surrounding tidal field [@innanen83; @bertin08; @renaud11; @webb13]. We proceed with computing the NGC1399 mass distribution profile using the CGS data[^10] and the recipes outlined in [@bell03] to convert photometric colors into stellar mass-to-light ratios as a function of galactocentric radius [see also @zibetti09; @into13]. We compute the ${\cal M}_\star/L_V$ profile through linear interpolation of predictions for a 13 Gyr old stellar population with variable metallicity, which is set by the measured photometric color profile of NGC1399, using the 2007 update of the [@bc03] SSP models. With the radial trend for ${\cal M}_\star/L_V$ we derive then the corresponding relation for the stellar mass $$\label{eq:massprof} \log({\cal M}_\star)\!=\!\log({\cal M}_\star/L_V)_r\!-\!0.4[m_V(r)\!-\!D_L\!-\!M_{V,\odot}]$$ enclosed in $R_{\rm gal}\!=\!r$, where $m_V(r)$ is the integrated magnitude derived from the galaxy surface brightness profile, $D_L$ is the luminosity distance, and $M_{V,\odot}\!=\!4.83$ mag is the absolute $V$-band magnitude of the Sun. With Equation \[eq:jacobi\] and the derived stellar mass profile of NGC1399 from Equation \[eq:massprof\], we determine the [*instantaneous*]{} Jacobi radii for GCs with a total mass of $m_{\rm GC}\!=\!10^3, 10^4$, and $10^5 M_\odot$ using the results from [@baumgardt10] and [@ernst13] who determine the typical ratios between half-light and Jacobi radius with minimum, mean, and maximum values of $\log(r_h/r_J)\!=\!-1.5,-1.0$, and $-0.5$ for Milky Way GCs. The extremes of the $r_h/r_J$ distribution are representative of GCs that are under- and overfilling their Roche lobes, respectively. We also compute the GC stellar masses using the differential ${\cal M}_\star/L$ predictions from the GALEV SSP models [@kotulla09], assuming uniformly old GC ages ($t_{\rm GC}=13$Gyr) and using the $g\!-\!z$ and $C\!-\!T1$ GC colors to account for ${\cal M}_\star/L$ variations as a function metallicity. For GCs which lack color information we adopt the median ${\cal M}_\star/L$ of the GC sample for which photometric colors are available. We overplot the corresponding expectation trends for $r_h$ as a function of galactocentric radius in Figure \[pdf:rh\_galrad\_jacobi\] and use the color shading to parametrize GC mass. Consistent with Figure \[pdf:rh\_intmag\] we see no preferred GC mass scale at a given galactocentric radius. We observe that none of the curves reproduces the break at 10kpc of the $r_h\!-\!R_{\rm gal}$ profile and its flatness at large galactocentric radii. The stellar mass density distribution is clearly not sufficient and requires an additional mechanism to limit GC sizes at large $R_{\rm gal}$. This could in principle be achieved by an exotic eccentricity distribution function of GC orbits, which would bring the outer clusters into the inner galaxy on preferentially radial orbits [see @webb13]. An alternative explanation for the observed situation could be an additional tidal limitation of GCs in the outskirts of the galaxy, which could be realized in two different ways: 1\) by an additional mass component in the form of a dark matter density profile of the NFW type $\rho(r)\!=\!\rho_0/[(r/R_s)(1+r/R_s)^2]$ [@navarro96] where the total mass inside radius $R_{\rm gal}$ is given by $$\begin{aligned} {\cal M}_{\rm DM}(<\!R_{\rm gal})&=& 4\pi \int_0^{R_{\rm gal}}r^2 \rho (r) dr\\ &=&4\pi \rho_0 R_s^3 \left[\ln\left(\frac{R_s+R_{\rm gal}}{R_s}\right)-\frac{R_{\rm gal}}{R_s+R_{\rm gal}}\right]\!. \nonumber\end{aligned}$$ The resulting relations for $\rho_0\!\approx\!4\cdot10^7 M_\odot/{\rm pc}^3$, $R_s\!\approx\!130$kpc, and $\log(r_h/r_J)\!=\!-1.0$ are illustrated in Figure \[pdf:rh\_galrad\_jacobi\] as magenta curves and show that even in the presence of a typical dark matter halo, i.e. considering ${\cal M}_\star\!+\!{\cal M}_{\rm DM}\!=\!{\cal M}_{\rm gal}$ in Equation \[eq:jacobi\], the $r_h\!-\!R_{\rm gal}$ GC relations are still monotonically increasing, albeit not as rapidly as in the case of considering ${\cal M}_\star$ only. Hence, an additional component is required to flatten out the $r_h\!-\!R_{\rm gal}$ profiles at large galactocentric radii. 2\) We, therefore, suggest that an increased stochastic distribution of low-mass dark matter haloes that are part of the galaxy cluster potential induce additional tidal stress on outer-halo GCs. Such a changing mass fraction in subhaloes as a function of galactocentric radius is observed in high-resolution $\Lambda$CDM simulations [e.g. @springel08] and would increase the “tidal variance" in outer-halo regions, thereby truncating the GC stellar density profiles. This may limit the GC sizes to a roughly constant value, something that shall be explored with dedicated high-resolution numerical simulations. Structural Parameter Distributions {#ln:distribs} ---------------------------------- [lccccccccccccc]{}\[!t\] NGC 1399 & 0.122 & 0.62 & (1) & $20.13\pm0.4$ & (7) & 51.3 & 0.061 & 0.21 & 32.9& 3.21& 0.0480 & 0.12\ NGC 4486 (M87) & 0.066 & 0.34 & (2) & $16.70\pm0.2$ & (8) & 12.3 & 0.064 & 0.22 & 41.5& 3.36& 0.0633 & 0.16\ NGC 4472 (M49) & 0.073 & 0.37 & (2) & $16.40\pm0.2$ & (8) & 11.6 & 0.072 & 0.25 & 56.1& 4.46& 0.0714 & 0.18\ NGC 4594 (M104) & 0.026 & 0.13 & (3) & $9.08\pm0.2$ & (9) & 15 & 0.024 & 0.08 & 55.3& 2.43& 0.0160 & 0.04\ NGC 5128 (CenA) & 0.170 & 0.86 & (4) & $3.84\pm0.35$ &(10)& 20 & 0.179 & 0.61 & 82.6& 1.54& 0.2127 & 0.54\ NGC 224 (M31) & 0.241 & 1.22 & (5) & $0.779\pm0.05$&(11)& 160 & 0.132 & 0.45 & 443.2& 1.67& 0.1316 & 0.33\ Milky Way & 0.197 &$\equiv1$&(6)& $\dots$&$\dots$&120&0.292&$\equiv1$&$\dots$&2.50&0.3974&$\equiv1$\ We show the $r_h$ distribution of NGC1399 GCs in Figure \[pdf:h\_distr\] together with corresponding measurements for Milky Way and M31 GCs, taken from the McMaster catalog [2010 update of @harris96] as well as [@peacock09] and [@huxor14], respectively. In addition, we compare our half-light radius measurements to the $r_h$ distributions of GCs in NGC 5128 [@woodley10], the Sombrero galaxy [M104, @harris10], and the two brightest Virgo ellipticals M49 and M87 which were studied by the ACSVCS [for details see @jordan09]. The bottom panel of Figure \[pdf:h\_distr\] shows the entire sample of NGC1399 GCs together with the distribution of radial-velocity confirmed GCs (dark histogram) and stars (cyan histogram). It is important to note that all spectroscopically confirmed foreground stars concentrate around $r_h\!\approx\!0$ pc, where unresolved objects are generally expected. We provide mean and median values of each $r_h$ distribution in each panel of Figure \[pdf:h\_distr\] and point out that there is a trend of decreasing $r_h$ with increasing host galaxy luminosity [@masters10] in which NGC1399 and its central GC system fit right in. Such a trend generally supports the notion that the host environment has an impact on the GC $r_h\!-\!R_{\rm gal}$ relation (see discussion above), and will depend on the sampled $R_{\rm gal}$ range. We also compare our sample to the measurements of [@masters10] who derived GC half-light radii for the central regions in NGC1399 from the ACS Fornax Cluster Survey data [@jordan07] which, similar to the ACSVCS in Virgo, sampled massive early-type galaxies in Fornax with one central HST/ACS pointing. Both our and the ACSFCS distributions show very similar shapes and drop-offs from $\sim\!1.5$ pc up to about 5 pc, beyond which our sample starts to include many more extended GCs. This is mainly due to the nine times larger field of view of our data and we point out that many of these extended sources are radial-velocity confirmed [*bona-fide*]{} GCs at large galactocentric radii. The comparison with other GC systems in the upper panels of Figure \[pdf:h\_distr\] shows that all half-light radius distributions have very similar shapes featuring a relatively steep increase in GC number density at low $r_h$ values, with a peak somewhere in the range of $2\!-\!3$ pc, and a shallower decline towards more extended objects. This distribution is dependent on the sampling of the GC luminosity function, galactocentric radius, as well as the amount of contamination, the measurement errors, and the GC selection criteria [e.g. @brescia12]. It is hard to compare the unresolved parts at $r_h\!\la\!1$ pc for galaxies further away than Sombrero (NGC4594 at $D\!\approx\!9$ Mpc) due to the resolution limit of HST (see shaded regions in Figure \[pdf:h\_distr\]). Despite this limitation there is ample information and some intriguing aspects of the GC size distributions for sources with $r_h\!\ga\!1.5$ pc. In NGC1399, these extended clusters predominantly reside at projected galactocentric radii, $R_{\rm gal}$, larger than 10 kpc (see Figure \[pdf:rh\_galrad\]). Since the observed GC populations in M49, M87, and M104 are all inside this radius (see Table \[tab:radfrac\]), we find a very small population of similarly extended GCs in the corresponding samples. This is, of course, an observational bias considering our and the earlier results by [@vdB91] and [@larsen03] who found correlations of the type $r_h\!\propto\!R_{\rm gal}^{n}$ with $n\!<\!1$, and [@jordan05] who suggested an analytic expression that approximates the $r_h$ distribution for the inner GC systems in Virgo ellipticals. Having sampled a significant population of GCs to large galactocentric radii in NGC1399 in combination with similar results for less rich GC system (see Figure \[pdf:h\_distr\]), we, therefore, suggest that all GC systems are comprised of two components of clusters: one standard GC population with a size distribution resembling the typical GC half-light radius of $2\!-\!3$ pc and a second, less rich component of more extended GCs that are predominantly found at larger galactocentric radii. Alternatively, there might be a combination of mechanisms (explored further below) that act on just one GC population, but their effects manifest themselves at different radii, so that the extended GCs are only observed at large $R_{\rm gal}$. To quantify the fraction of extended GCs in a GC system, we define the number ratio of GCs with sizes larger than 5 pc relative to the total GC population, $$\label{eq:S5} E_5=N_{\rm GC}(r_h\!\geq\!5 {\rm pc})/N_{\rm GC}({\rm all}),$$ and normalize this value to the Galactic GC system, i.e. $$\label{eq:S5hat} \hat{E}_5=\frac{ N_{\rm GC}(r_h\geq5 {\rm pc})}{N_{\rm GC}({\rm all})} \left(\frac{N_{\rm GC}(r_h\geq5 {\rm pc})_{\rm MW}}{N_{\rm GC} ({\rm all})_{\rm MW}}\right)^{-1}.$$ The results for all GC systems are summarized in Table \[tab:radfrac\] for the galactocentric sampling ranges of the corresponding dataset, which vary by about an order of magnitude. In order to representatively compare the GC samples we, therefore, restrict each dataset to within $R_{\rm gal}\!\leq\!10$kpc (about the maximum homogeneous sampling radius of the samples) as well as 2.5 effective radii of the host galaxy’s diffuse light (set by the maximum radial sampling of each dataset), measured in the near-infrared $K_s$ filter. We summarize the corresponding values as $E_{5/10}$ and $\hat{E}_{5/10}$ for $R_{\rm gal}\!\leq\!10$kpc as well as ${\cal E}_{5}$ and $\hat{{\cal E}}_{5}$ for $R_{\rm gal}\!\leq\!2.5\,r_e$ in Table \[tab:radfrac\]. We find a clear dichotomy in the ${\cal E}_{5}$ (and $E_{5/10}$) between late-type and early-type galaxies. While the three giant ellipticals NGC1399, M87 and M49 as well as M104 show ${\cal E}_{5}$ values clearly below 10%, the two late-type spirals, i.e. M31 and the Milky Way, as well as NGC5128 stand out with significantly higher ${\cal E}_{5}$ values, clearly above $\sim\!10\%$. We attribute this result to differences in the tidal environment properties throughout the dynamical evolution and merging history of these galaxies. Giant ellipticals experience in general a more violent evolution than spirals. It is unclear yet, how these numbers compare to other GC systems, but the fact that ${\cal E}_{5}$ values of NGC1399 and the two Virgo giant ellipticals M87 and M49 are remarkably similar, hints at physical processes acting that are acting in a similar way on the size evolution of their GC systems. This includes the somewhat surprising result for the Sombrero galaxy’s GC system with a similar ${\cal E}_{5}$ value as the giant ellipticals. Higher ${\cal E}_{5}$ values for the Milky Way and M31 might be the result of the dynamically more benign tidal field around such distant GCs and/or the younger, i.e. less evolved, nature of NGC5128, a recent merger remnant, and its GC system. How these numbers will play out for the GC systems in other Virgo cluster galaxies will be shown by the [*Next Generation Virgo Cluster Survey*]{} (NGVS) which achieves a spatial resolution of $\sim\!5$pc for the entire Virgo galaxy cluster out to its virial radius [@ferrarese12; @munoz14]. At least then it will be clear whether late-type galaxies have a systematically larger population of extended GCs than early-type galaxies, which host GC systems with a relatively smaller population of extended GCs. Of course, we expect a complex interplay between the formation paths of the compact and extended GCs. In fact, we expect multiple components in the GC size distribution depending on the star cluster formation history and the evolution of the host galaxy. However, in a simplistic picture we speculate that while the primary component GCs (i.e. compact GCs) are likely massive and old, and formed [*in-situ*]{}, the nature of secondary component GCs (i.e. extended GCs) is likely the result of a combination of populations of 1) dissolving star clusters triggered by recent formation of younger, low-mass GCs combined with increased tidal stress, e.g. in central regions of galaxy clusters or merger remnants [@gieles11; @goudfrooij12], 2) the accretion of more extended GCs from satellite galaxies which formed and survived in a more benign tidal environment [see also @georgiev09; @dacosta09; @smith13], and/or 3) disrupting cores of stripped dwarf galaxy nuclei [e.g. @oh00; @bekki03; @pfeffer13]. The corresponding detailed analysis of this scenario is the focus of a forthcoming paper. Kinematic Properties of Compact and Extended GCs ------------------------------------------------ ![([*Top panel*]{}): The plot shows all measurements for individual GCs as in Figure \[pdf:rh\_galrad\] with the green dash-dotted curve approximating the running median relation as defined in Equation \[eqn:split\]. We use this empirical separation to define formally compact and extended GCs. ([*Middle panel*]{}): The radial velocity of each GC matched with the [@schuberth10] sample as a function of galactocentric radius. The mean radial velocity of the sample is shown as horizontal dashed line. The symbols are parametrized by the GC color (see Table \[tab:col\_sel\]) and split into blue and red GCs (shown in corresponding colors) and those without color information (shown as black dots). Green circles indicate GCs that have sizes formally more compact than relation shown as dash-dotted relation in the top panel. ([*Bottom panel*]{}): Sliding-median relations of the line-of-sight radial velocity dispersion as a function of galactocentric radius for compact (GC$_{\rm cmp}$, [*green*]{}) and extended GCs (GC$_{\rm ext}$, [*black*]{}) with their corresponding 90% confidence limits shown as dotted curves.[]{data-label="pdf:rh_galrad_losvd"}](fig17_rh_galrad_LOSVD.jpg){width="8.9cm"} The availability of matched GC size and radial velocity measurements in NGC1399 allows us to investigate correlations between these two parameters. The radial velocities of all GCs with size measurements do not correlate in any statistically significant way as a function of projected galactocentric radius (Figure \[pdf:rh\_galrad\_losvd\]). We measure a total systemic heliocentric radial velocity of the entire sample as $\langle v_{\rm helio}\rangle\!=\!1456\!\pm\!17$ km s$^{-1}$ with a line-of-sight velocity dispersion of $\sigma\!=\!295$ km s$^{-1}$ in good agreement with [@schuberth10]. This is also in good agreement and consistent with previous measurements of the diffuse light, i.e. $\langle v_{\rm helio}\rangle\!=\!1425\!\pm\!4$ km s$^{-1}$ and $\sigma_0\!=\!353\pm19$ km s$^{-1}$ [@graham98], respectively. Next, we divide our sample in a population of compact (GC$_{\rm cmp}$) and extended GCs (GC$_{\rm ext}$) using as division the relation illustrated as green dash-dotted line in the top panel of Figure \[pdf:rh\_galrad\_losvd\], which is approximating the running-median $r_h$ curve of the entire sample (black curves). This linear separation can be numerically expressed as $$r_h [{\rm pc}] = \left\{ \begin{array}{l l} 0.012\,R_{\rm gal} + 0.8 & \quad \text{if $R_{\rm gal}<100$\arcsec}\\ 2 & \quad \text{if $R_{\rm gal}\geq100$\arcsec.} \end{array} \right. \label{eqn:split}$$ We scrutinize the GC size-$v_{\rm helio}$ relation for any correlations and find no significant slope for compact and extended GCs as a function of galactocentric radius $R_{\rm gal}$. However, looking at the line-of-sight velocity dispersion, $\sigma$, of each of those sub-samples, we find a surprisingly clear dichotomy between compact and extended GCs in terms of their mean velocity dispersion. While the compact GC sample exhibits $\langle\sigma_{\rm cmp}\rangle\!=\!225\!\pm\!25$ km s$^{-1}$, we compute a much higher value for the extended sample with $\langle\sigma_{\rm ext}\rangle\!=\!317\!\pm\!21$ km s$^{-1}$. This is consistent with the $\sigma$ differences found by [@schuberth10] between the blue and red GC sub-population in NGC1399 at a similar range in galactocentric radius. Plotting the sliding median of the $\sigma\!-\!R_{\rm gal}$ relation (bottom panel of Figure \[pdf:rh\_galrad\_losvd\]) reveals that this difference is most pronounced in the range $50\arcsec\!\la\! R_{\rm gal}\!\la\!320\arcsec$, which roughly corresponds to the physical range of $15\!\la\! R_{\rm gal}\!\la\!32$kpc. Outside this range, the difference seem to disappear, but we lack sample statistics to make definitive conclusions, and defer a more detailed analysis of this surprising result to a future study, when more comprehensive radial velocity samples become available. Here we just note that given the scatter of the rather weak correlation between GC size and color (see Section \[ln:sizecol\]), the significantly lower velocity dispersion of more compact (i.e. red) GCs compared to their more extended counterparts (i.e. blue GCs) appears to be the astrophysically stronger relation, which likely has its origin in the stronger influence of [*external*]{} tidal truncation effects compared to [*internal*]{} mechanisms that govern the GC size. This is also consistent with our result of the flatter GC size-color relation for the inner vs. outer GC sample (see Figure \[pdf:rh\_color\]). These findings indicate the preferential influence of [*external*]{} dynamical effects damping the size difference between red and blue GCs which is predominantly driven by [*internal*]{} evolution of their constituent stellar populations and is likely a corollary of the GC size-dynamics correlation. Future GC radial velocity samples of the inner GC system in NGC1399 will shed light on how the GC orbit distribution function influences these relations. [*Acknowledgments*]{} – Support for HST program GO-10129 was provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Incorporated, under NASA contract NAS5-26555. This research was supported by FONDECYT Regular Project Grant No. 1121005 and BASAL Center for Astrophysics and Associated Technologies (PFB-06). THP is thankful for the hospitality and support during his visits at the University of Napoli Federico II where parts of this work were completed; he also gratefully acknowledges support in form of a Plaskett Research Fellowship from the National Research Council of Canada. MP acknowledges financial support from the FARO2011 project of the University of Napoli Federico II. We are grateful to Anton Koekemoer and Andy Fruchter for their technical support and useful discussions on the MultiDrizzle code and to Chien Y. Peng for his help with the implementation and testing of the modified GALFIT routine. We thank Tom Richtler and Ylva Schuberth for providing their radial velocity measurements ahead of publication, as well as Luis Ho and Zhao-Yu Li for kindly making available to us their latest NGC1399 surface brightness profile measurements from the Carnegie-Irvine Galaxy Survey, again prior to publication. Avon Huxor has very kindly supplied M31 GC data prior to publication. We are grateful to the referee, Bill Harris, for providing a thoughtful and constructive report that helped improve the presentation of the results. We thank Jeremy Webb, Mark Gieles, Andres Jord[á]{}n, Eric Peng, Chunyan Jiang, Stephen Zepf, and Arunav Kundu for valuable discussions and providing data in electronic format. Some of the data presented in this paper were obtained from the Multimission Archive at the Space Telescope Science Institute (MAST). STScI is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. Support for MAST for non-HST data is provided by the NASA Office of Space Science via grant NNX09AF08G and by other grants and contracts. This research has made use of the NASA/ IPAC Infrared Science Archive, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. Figure \[pdf:full\] and \[pdf:example\] were created with the help of the ESA/ESO/NASA Photoshop FITS Liberator. This research has made use of NASA’s Astrophysics Data System. Facilities: . Anderson, J., & King, I. R. 2000, , 112, 1360 Anderson, J., & King, I. 2006, Instrument Science Report ACS 2006-001 (Baltimore: STScI), http://www.stsci.edu/hst/acs/documents/isrs/isr0601.pdf Anderson, J. 2005, The 2005 HST Calibration Workshop, eds. Koekemoer, A. M., Goudfrooij, P., & Dressel, L. L. Barmby, P., McLaughlin, D. E., Harris, W. E., Harris, G. L. H., & Forbes, D. A. 2007, , 133, 2764 Bassino, L. P., Faifer, F. R., Forte, J. C., Dirsch, B., Richtler, T., Geisler, D., & Schuberth, Y. 2006, , 451, 789 Bastian, N., Gieles, M., Goodwin, S. P., Trancho, G., Smith, L. J., Konstantopoulos, I., & Efremov, Y. 2008, , 389, 223 Baumgardt, H., & Makino, J. 2003, , 340, 227 Baumgardt, H., Parmentier, G., Gieles, M., & Vesperini, E. 2010, , 401, 1832 Beckwith, S. V. W., et al. 2006, , 132, 1729 Bekki, K., & Freeman, K. C. 2003, , 346, L11 Bekki, K. 2010, , 401, 2753 Bell, E. F., McIntosh, D. H., Katz, N., & Weinberg, M. D. 2003, , 149, 289 Bertin, E., & Arnouts, S. 1996, , 117, 393 Bertin, G., & Varri, A. L. 2008, , 689, 1005 Binney, J., & Tremaine, S. 1987, Princeton, NJ, Princeton University Press, p. 235 Blakeslee, J. P., et al. 2009, , 694, 556 Blom, C., Spitler, L. R., & Forbes, D. A. 2012, , 420, 37 Bournaud, F., Duc, P.-A., & Emsellem, E. 2008, , 389, L8 Brescia, M., Cavuoti, S., Paolillo, M., Longo, G., & Puzia, T. 2012, , 421, 1155 Bruzual, G., & Charlot, S. 2003, , 344, 1000 Caon, N., Capaccioli, M., & D’Onofrio, M. 1994, , 106, 199 Carlson, M. N., & Holtzman, J. A. 2001, , 113, 1522 Chandar, R. 2009, , 324, 315 Chandar, R., Whitmore, B. C., & Fall, S. M. 2010, , 713, 1343 Chun, M. S., Suh, Y. R., & Lee, Y. B. 1980, Journal of Korean Astronomical Society, 13, 27 Conn, A. R., Ibata, R. A., Lewis, G. F., et al. 2012, , 758, 11 C[ô]{}t[é]{}, P., et al. 2004, , 153, 223 Crampton, D., Cowley, A. P., Schade, D., & Chayer, P. 1985, , 288, 494 Da Costa, G. S. 1979, , 84, 505 Da Costa, G. S., Grebel, E. K., Jerjen, H., Rejkuba, M., & Sharina, M. E. 2009, , 137, 4361 Demers, S., Grondin, L., & Kunkel, W. E. 1990, , 102, 632 D’Ago et al. 2013, ApJ [*submitted*]{} Dirsch, B., Richtler, T., Geisler, D., Forte, J. C., Bassino, L. P., & Gieren, W. P. 2003, , 125, 1908 Downing, J. M. B. 2012, arXiv:1204.5363 Drinkwater, M. J., et al. 2000, , 355, 900 Dunn, L. P., & Jerjen, H. 2006, , 132, 1384 Elmegreen, B. G., & Efremov, Y. N. 1997, , 480, 235 Elmegreen, B. G., & Hunter, D. A. 2010, , 712, 604 Elson, R., Hut, P., & Inagaki, S. 1987, , 25, 565 Elson, R. A. W. 1991, , 76, 185 Elson, R. A. W. 1992, , 256, 515 Elson, R. A. W., & Freeman, K. C. 1985, , 288, 521 Elson, R. A., & Walterbos, R. A. M. 1988, , 333, 594 Elson, R. A. W., & Schade, D. J. 1994, , 437, 625 Elson, R. A. W., Fall, S. M., & Freeman, K. C. 1987, , 323, 54 Ernst, A., & Just, A. 2013, , 429, 2953 Faifer, F. R., Bassino, L. P., Forte, J. C., Dirsch, B., Richtler, T., & Geisler, D. 2004, Boletin de la Asociacion Argentina de Astronomia La Plata Argentina, 47, 132 Fall, S. M., Chandar, R., & Whitmore, B. C. 2009, , 704, 453 Ferguson, H. C., & Sandage, A. 1989, , 346, L53 Ferrarese, L., C[ô]{}t[é]{}, P., Cuillandre, J.-C., et al. 2012, , 200, 4 Ford, H. C., et al. 2003, , 4854, 81 Forte, J. C., Faifer, F., & Geisler, D. 2005, , 357, 56 Fruchter, A. S., & Hook, R. N. 2002, , 114, 144 Fusi Pecci, F., et al. 1994, , 284, 349 Georgiev, I. Y., Puzia, T. H., Hilker, M., & Goudfrooij, P. 2009a, , 392, 879 Georgiev, I. Y., Hilker, M., Puzia, T. H., Goudfrooij, P., & Baumgardt, H. 2009b, , 396, 1075 Georgiev, I. Y., Puzia, T. H., Goudfrooij, P., & Hilker, M. 2010, arXiv:1004.2039 Gieles, M., & Bastian, N. 2008, , 482, 165 Gieles, M., Heggie, D. C., & Zhao, H. 2011, , 413, 2509 Gnedin, O. Y., & Ostriker, J. P. 1997, , 474, 223 G[ó]{}mez, M., Geisler, D., Harris, W. E., Richtler, T., Harris, G. L. H., & Woodley, K. A. 2006, , 447, 877 G[ó]{}mez, M., & Woodley, K. A. 2007, , 670, L105 Goudfrooij, P. 2012, , 750, 140 Graham, A. W., Colless, M. M., Busarello, G., Zaggia, S., & Longo, G. 1998, , 133, 325 Greif, T. H., Bromm, V., Clark, P. C., et al. 2012, , 424, 399 Griffen, B. F., Drinkwater, M. J., Thomas, P. A., Helly, J. C., Hack, W., Busko, I., & Jedrzejewski, R. 2003, Astronomical Data Analysis Software and Systems XII, 295, 453 Harris, W. E. 1996, , 112, 1487 Harris, W. E., Harris, G. L. H., Barmby, P., McLaughlin, D. E., & Forbes, D. A. 2006, , 132, 2187 Harris, W. E. 2009, , 699, 254 Harris, W. E., Spitler, L. R., Forbes, D. A., & Bailin, J. 2010, , 401, 1965 Harris, G. L. H., Rejkuba, M., & Harris, W. E. 2010, Publications of the Astronomical Society of Australia, 27, 457 Harris, W. E., Harris, G. L. H., & Alessi, M. 2013, , 772, 82 Harris, W. E., & Pudritz, R. E. 1994, , 429, 177 Harris, W. E., Harris, G. L. H., Holland, S. T., & McLaughlin, D. E. 2002, , 124, 1435 Hartwick, F. D. A. 2009, , 691, 1248 H[é]{}non, M. 1973, , 24, 229 H[é]{}non, M. 1975, IAU Symp.  69: Dynamics of the Solar Systems, 69, 133 Hilker, M., Infante, L., Vieira, G., Kissler-Patig, M., & Richtler, T. 1999, , 134, 75 Ho, L. C., Li, Z.-Y., Barth, A. J., Seigar, M. S., & Peng, C. Y. 2011, , 197, 21 Huxor, A. P., Tanvir, N. R., Irwin, M. J., Ibata, R., Collett, J. L., Ferguson, A. M. N., Bridges, T., & Lewis, G. F. 2005, , 360, 1007 Huxor, A. P., et al. 2014, MNRAS [*in preparation*]{} Illingworth, G., & Illingworth, W. 1976, , 30, 227 Innanen, K. A., Harris, W. E., & Webbink, R. F. 1983, , 88, 338 Into, T., & Portinari, L. 2013, , 430, 2715 Jee, M. J., Blakeslee, J. P., Sirianni, M., et al. 2007, , 119, 1403 Jensen, J. B., Tonry, J. L., Barris, B. J., et al. 2003, , 583, 712 Jord[á]{}n, A. 2004, , 613, L117 Jord[á]{}n, A., et al. 2005, , 634, 1002 Jord[á]{}n, A., et al. 2007, , 169, 213 Jord[á]{}n, A., et al. 2009, , 180, 54 King, I. 1962, , 67, 471 King, I. R. 1966, , 71, 64 King, I. R., Hedemann, E. J., Hodge, S. M., & White, R. E. 1968, , 73, 456 Koekemoer, A. M., Fruchter, A. S., Hook, R. N., & Hack, W. 2002, The 2002 HST Calibration Workshop.  eds. S. Arribas, A. Koekemoer, and B. Whitmore, p.337 Kontizas, M., Danezis, E., & Kontizas, E. 1982, , 49, 1 Kotulla, R., Fritze, U., Weilbacher, P., & Anders, P. 2009, , 396, 462 Kozhurina-Platais, V., Goudfrooij, P., & Puzia, T. H. 2007, Instrument Science Report ACS 2007-04 (Baltimore: STScI), www.stsci.edu/hst/acs/documents/isrs/isr0704.pdf Kravtsov, A. V., & Gnedin, O. Y. 2005, , 623, 650 Kron, R. G. 1980, , 43, 305 Kukarkin, B. V., & Kireeva, N. N. 1979, Soviet Astronomy, 23, 261 Kundu, A. 2008, , 136, 1013 Kundu, A., & Whitmore, B. C. 1998, , 116, 2841 Kundu, A., Whitmore, B. C., Sparks, W. B., Macchetto, F. D., Zepf, S. E., & Ashman, K. M. 1999, , 513, 733 Kundu, A., & Whitmore, B. C. 2001, , 121, 2950 Larsen, S. S. 1999, , 139, 393 Larsen, S. S., Brodie, J. P., Huchra, J. P., Forbes, D. A., & Grillmair, C. J. 2001, , 121, 2974 Larsen, S. S., & Brodie, J. P. 2003, , 593, 340 Li, Y., Mac Low, M.-M., & Klessen, R. S. 2005, , 626, 823 Li, Z.-Y., Ho, L. C., Barth, A. J., & Peng, C. Y. 2011, , 197, 22 Lotz, J. M., Telford, R., Ferguson, H. C., et al. 2001, , 552, 572 Madrid, J. P., Harris, W. E., Blakeslee, J. P., & G[ó]{}mez, M. 2009, , 705, 237 Mapelli, M., & Bressan, A. 2013, , 430, 3120 Maraston, C. 2005, , 362, 799 Mar[í]{}n-Franch, A., et al. 2009, , 694, 1498 Masters, K. L., Jord[á]{}n, A., C[ô]{}t[é]{}, P., et al. 2010, , 715, 1419 McLaughlin, D. E., & van der Marel, R. P. 2005, , 161, 304 McLaughlin, D. E., Barmby, P., Harris, W. E., Forbes, D. A., & Harris, G. L. H. 2008, , 384, 563 Mei, S., Blakeslee, J. P., C[ô]{}t[é]{}, P., et al. 2007, , 655, 144 Misgeld, I., Hilker, M., & Mieske, S. 2010, arXiv:1010.3138 Misgeld, I., & Hilker, M. 2011, , 414, 3699 Monet, D. G., et al. 2003, , 125, 984 Mu[ñ]{}oz, R. P., Puzia, T. H., Lan[ç]{}on, A., et al. 2014, , 210, 4 Murphy, B. W., Cohn, H. N., & Hut, P. 1990, , 245, 335 Murray, N. 2009, , 691, 946 Murray, S. D., & Lin, D. N. C. 1992, , 400, 265 Navarro, J. F., Frenk, C. S., & White, S. D. M. 1996, , 462, 563 Oh, K. S., & Lin, D. N. C. 2000, , 543, 620 Paolillo, M., Puzia, T. H., Goudfrooij, P., et al. 2011, , 736, 90 Peacock, M. B., Maccarone, T. J., Waters, C. Z., Kundu, A., Zepf, S. E., Knigge, C., & Zurek, D. R. 2009, , 392, L55 Peng, C. Y., Ho, L. C., Impey, C. D., & Rix, H.-W. 2010, , 139, 2097 Peng, E. W., et al. 2009, , 703, 42 Peng, E. W., Jord[á]{}n, A., C[ô]{}t[é]{}, P., et al. 2006, , 639, 95 Pfeffer, J., & Baumgardt, H. 2013, , 433, 1997 Puzia, T. H., Kissler-Patig, M., Brodie, J. P., & Huchra, J. P. 1999, , 118, 2734 Puzia, T. H., Kissler-Patig, M., Brodie, J. P., & Schroder, L. L. 2000, , 120, 777 Puzia, T. H., Zepf, S. E., Kissler-Patig, M., et al. 2002, , 391, 453 Renaud, F., Gieles, M., & Boily, C. M. 2011, , 418, 759 Rhode, K. L., & Zepf, S. E. 2001, , 121, 210 Rhode, K. L., & Zepf, S. E. 2004, , 127, 302 Rhode, K. L., Zepf, S. E., Kundu, A., & Larner, A. N. 2007, , 134, 1403 Rhodes, J. D., et al. 2007, , 172, 203 Richtler, T. 2003, Stellar Candles for the Extragalactic Distance Scale, 635, 281 Riess, A., & Mack, J. 2004, Instrument Science Report ACS 2004-006 (Baltimore: STScI), also available at http://www.stsci.edu/hst/acs/documents/isrs/isr0406.pdf Robin, A. C., Reyl[é]{}, C., Derri[è]{}re, S., & Picaud, S. 2003, , 409, 523 Sandage, A. 1975, , 202, 563 Schlafly, E. F., & Finkbeiner, D. P. 2011, , 737, 103 Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, , 500, 525 Schuberth, Y., Richtler, T., Hilker, M., Dirsch, B., Bassino, L. P., Romanowsky, A. J., & Infante, L. 2010, , 513, A52 Schulman, R. D., Glebbeek, V., & Sills, A. 2012, , 420, 651 Sersic, J. L. 1968, Atlas de Galaxias Australes, Observatorio Astronomico, Cordoba Sippel, A. C., Hurley, J. R., Madrid, J. P., & Harris, W. E. 2012, , 427, 167 Sharina, M. E., Puzia, T. H., & Makarov, D. I. 2005, , 442, 85 Sirianni, M., et al. 2005, , 117, 1049 Smith, R., S[á]{}nchez-Janssen, R., Fellhauer, M., et al. 2013, , 429, 1066 Spitler, L. R., Larsen, S. S., Strader, J., Brodie, J. P., Forbes, D. A., & Beasley, M. A. 2006, , 132, 1593 Springel, V., Wang, J., Vogelsberger, M., et al. 2008, , 391, 1685 Taylor, M. A., Puzia, T. H., Harris, G. L., Harris, W. E., Kissler-Patig, M., & Hilker, M. 2010, , 712, 1191 Trager, S. C., King, I. R., & Djorgovski, S. 1995, , 109, 218 van den Bergh, S., Morbey, C., & Pazder, J. 1991, , 375, 594 de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H. G., Jr., Buta, R. J., Paturel, G., & Fouque, P. 1991, Volume 1-3, XII,  Springer-Verlag Vesperini, E., & Heggie, D. C. 1997, , 289, 898 Vesperini, E., & Zepf, S. E. 2003, , 587, L97 Webb, J. J., Sills, A., & Harris, W. E. 2012a, , 746, 93 Webb, J. J., Harris, W. E., & Sills, A. 2012b, , 759, L39 Webb, J. J., Harris, W. E., Sills, A., & Hurley, J. R. 2013, , 764, 124 Wilson, C. P. 1975, , 80, 175 Woodley, K. A., & G[ó]{}mez, M. 2010, PASA, 27, 379 Zepf, S. E., Ashman, K. M., English, J., Freeman, K. C., & Sharples, R. M. 1999, , 118, 752 Zibetti, S., Charlot, S., & Rix, H.-W. 2009, , 400, 1181 [^1]: The specific frequency of a GC system is defined as twice the number of GCs brighter than the turn-over luminosity of the GC luminosity function, given by $N_{\rm GC}$, relative to the absolute $V$-band luminosity of the host galaxy, $M_V$, which is normalized to $-15$ mag. This quantity is defined as the specific frequency of a GC system $S_N\!=\!N_{\rm GC}10^{0.4(M_V+15)}$; see also [@georgiev10] and [@harris13] for other GC system scaling relations. [^2]: [@schuberth10] approximate the radial GC system number density distribution with a cored power-law profile of the form $N(R)\!\propto\!((R/R_0)^2\!+\!1)^{-\alpha}$, where the core radius is $R_0\!=\!1.74\arcmin\pm0.27\arcmin$ and the power-law exponent $\alpha=0.84\pm0.02$. [^3]: <http://tdc-www.harvard.edu/software/catalogs/ub1.html> [^4]: The Kron radius is defined as $r_k=\sum rI(r)/\sum I(r)$. A circular aperture of radius $2r_k$ encloses $\geq90$% of an object’s flux independent of its magnitude [@kron80]. [^5]: We note that in [@paolillo11] and [@brescia12] the selection criteria were somewhat different, although broadly consistent, as those works had a different objective. [^6]: http://archive.stsci.edu/prepds/udf/udf\_hlsp.html [^7]: http://www.stsci.edu/software/tinytim/tinytim.html [^8]: The IDL source code to produce the drPSF library grid images is available at http://people.na.infn.it/$\sim\!$paolillo/Software.html. [^9]: We also note that [@carlson01] claim that S/N $>500$ is required in order to fully recover all King model parameters for every type of GC out to a distance of $\sim\!40$Mpc, i.e. twice as far as NGC1399. On the other hand, they state that S/N $\!\approx\!100$ is appropriate for, e.g. Virgo galaxies, or less concentrated systems, and that GC half-light radii are recovered with even better accuracy. Furthermore our spatial sampling (pixel size) is $\sim\!3$ times better than what was used in their study. [^10]: http://cgs.obs.carnegiescience.edu/CGS/Home.html
{ "pile_set_name": "ArXiv" }
--- abstract: 'Strong experimental evidence has indicated that tumor growth belongs to the molecular beam epitaxy universality class. This type of growth is characterized by the constraint of cell proliferation to the tumor border, and surface diffusion of cells at the growing edge. Tumor growth is thus conceived as a competition for space between the tumor and the host, and cell diffusion at the tumor border is an optimal strategy adopted for minimizing the pressure and helping tumor development. Two stochastic partial differential equations are introduced in this work in order to correctly model the physical properties of tumoral growth in (1+1) and (2+1) dimensions. The advantages of these models is that they reproduce the correct geometry of the tumor and are defined in terms of polar variables. Analysis of these models allow us to quantitatively estimate the response of the tumor to an unfavorable perturbation during the growth.' author: - Carlos Escudero title: Stochastic models for tumoral growth --- In Nature, one can find a huge number of systems that develop a rough interface in the process of growing. Many of them have been adequately understood by means of the use of some tools from fractal geometry, as scaling analysis, as well as via modelling with stochastic partial differential equations (SPDEs) and discrete models [@barabasi]. While these concepts do not constitute an exclusive theoretical framework for surfaces in the physical world, they can be applied to win a deeper understanding of many processes in biology [@eden]. Due to the many possible important applications in medicine, tumor growth constitutes one of the most interesting subjects of study to which scaling analysis can be applied. Actually, a very important research on tumor growth has been recently carried out. It has been found a strong empirical evidence that a broad class of tumors belong to the same universality class: the molecular beam epitaxy (MBE) universality class [@bru1; @bru2]. MBE dynamics are characterized by a number of features which include a linear growth rate, the constraint of growth activity to the outer border of the tumor, and surface diffusion at the growing edge; all of them have been observed experimentally. Surface diffusion has been identified as an optimal mechanism for favoring tumor growth. The host tissue exerts pressure on tumors which opposes their growth, but surface diffusion drives the cells to the concavities of the interface keeping this pressure to a minimum. These findings suggested studying the effect of the immune response on the tumor, and it has been established that an enhancement of the immune response increases the pressure on the tumor surface, and therefore limits its development [@bru3]. A very important consequence of this fact is its possible application to improve cancer therapy, something that has been already exploited with positive results [@bru4]. All these achivements underline the fundamental importance of understanding the physics of tumor growth. Before introducing any model, it is of fundamental importance to point out that the application of a physical model to a complex biological process implies a high simplification of many of its features. The movement of the cells is actually much more complex than that simply described by diffusion, it is affected by chemotaxis and haptotaxis. The dynamics of tumor-host interactions is determined by many complex cellular and extracellular processes, which include normal epithelial and mesenchymal cells as well as the extracellular matrix in addition to the immune response. This interaction is very complex and highly variable. Normal cells adjacent to the tumor are often induced to promote tumor growth by releasing proteolytic enzymes to break down the extracellular matrix or release growth factors that enhance tumor proliferation. In addition, it appears that tumor cells may under some circumstances transform into mesenchymal cells producing new populations of relatively normal appearing cells that support tumor growth. Tumors are extensively infiltrated by immune cells which may constitute as much as one third of its volume. Both the tumor phenotype and the tumor environment are very heterogeneous. The former is the result of accumulating random mutations, variable enviromental selection forces and perhaps restriction of proliferate capacity in non-stem cell components of the tumor. In addition, the tumor environment is extremely heterogeneous primarily due to disordered angiogenesis and blood flow. These facts underline the huge complexity of the problem, and remind us that the equations appearing below do not constitute a fundamental description of tumor growth, but a statistical approach to some of its properties. The continuum equation which describes the MBE universality class (also known as the Mullins-Herring equation [@mullins]) is the following SPDE $$\partial_t h=-K\nabla^4 h + F + \eta({\bf x},t),$$ where $h$ is the interface height, $K$ is the surface diffusion coefficient, and $\eta({\bf x},t)$ is a gaussian noise with zero mean and correlations given by $<\eta({\bf x},t)\eta({\bf x'},t')>=D\delta({\bf x}-{\bf x'})\delta(t-t')$. The term $F$ has the dimensions of a velocity, and in this case should be interpreted as the product of the mean cell radius and the cell division rate. The critical exponents can be extracted from this equation simply by power counting. If we neglect for a moment the velocity $F$, and we perform the transformations $x \to b x$, $t \to b^z t$, and $h \to b^\alpha h$ in the one-dimensional case, we see that the only values of $z$ and $\alpha$ that yield scale invariance are $z=4$ and $\alpha=3/2$. Since in this case $\alpha>1$, the system is super-rough and it is characterized by the set of critical exponents: $\alpha=3/2$, $\alpha_{loc}=1$, $z=4$, $\beta=3/8$, and $\beta^*=1/8$ [@lopez], which were found to be compatible with those measured in experiments on tumor growth [@bru1; @bru2]. Another additional nice feature of the Mullins-Herring equation is its simplicity: since it is linear, it can be exactly solved by means of a Fourier transformation. Another properties of this equation, however, make it not so suitable for describing tumor growth. It describes the growth of a surface from a planar substrate of fixed size, while actual tumors show radial growth with their size continuously increasing in time. It is thus important to derive a SPDE able to describe MBE physics with the correct geometry and spatiotemporal properties of tumor growth. In order to get the correct theoretical description of tumoral growth we will need to borrow some elements from differential geometry. On the other hand, these geometrical concepts are common in the formulation of stochastic growth equations in reparametrization invariance form [@marsili]. The equation of growth of a general riemannian surface reads $$\partial_t \vec{r}({\bf s},t)=\hat{n}({\bf s},t)\Gamma[\vec{r}({\bf s},t)]+\vec{\Phi}({\bf s},t),$$ where the $d+1$ dimensional surface vector $\vec{r}({\bf s},t)=\{r_\alpha({\bf s},t)\}_{\alpha=1}^{d+1}$ runs over the surface as ${\bf s}=\{s^i\}_{i=1}^d$ varies in a parameter space (in the following, latin indices vary from 1 to $d$ and greek indices from 1 to $d+1$). In this equation $\hat{n}$ stands for the unitary vector normal at the surface at $\vec{r}$, $\Gamma$ contains a deterministic growth mechanism that causes growth along the normal $\hat{n}$ to the surface, and $\vec{\Phi}$ is a random force acting on the surface. In our case the deterministic part should include a term modelling cell diffusion in the tumor border. When surface diffusion occurs to minimize the surface area the corresponding term in the equation is [@marsili]: $$\label{surdiff} \Gamma_s=-K\Delta_{BL}H,$$ where $\Delta_{BL}$ is the Beltrami-Laplace operator $$\Delta_{BL}=\frac{1}{\sqrt{g}}\partial_i(\sqrt{g}g^{ij}\partial_j),$$ $g_{ij}$ is the metric tensor and g is its determinant, $\partial_i=\partial/\partial s^i$ is a covariant derivative, and $H=\hat{n} \cdot \Delta_{BL} \vec{r}$ is the mean curvature. Summation over repeated indices is always assumed along this work. Finally, the unitary normal vector is given by $\hat{n}=g^{-1/2}\partial_1\vec{r}\times\cdots\times\partial_d\vec{r}$. In the case of the (1+1)-dimensional Monge form (or what is the same, the parametrization corresponding to a planar substrate) we have $\vec{r}=(x,h(x))$, the unitary normal vector takes this times the form $$\hat{n}=\frac{1}{\sqrt{1+(\partial_x h)^2}}(-\partial_x h,1),$$ the metric tensor is given by (note that for this particular case the metric tensor is a scalar) $\tilde{g}=1+(\partial_x h)^2$. Thus, the resulting mean curvature is $$H=\frac{\partial_x^2 h}{[1+(\partial_x h)^2]^{3/2}}.$$ The corresponding contribution to the drift reads $$\Gamma_s=-K\frac{3[-1+5(\partial_x h)^2](\partial_x^2 h)^3-10[\partial_x h + (\partial_x h)^3]\partial_x^2 h \partial_x^3 h + [1+(\partial_x h)^2]^2\partial_x^4 h }{[1+(\partial_x h)^2]^{9/2}},$$ that constitutes an expression far more complex than that of the Mullins-Herring equation. However, we can linearize this expression about the derivatives of $h$ to get $\Gamma_s=-K\partial^4_x h$, recovering the familiar drift of the MBE equation. This is the so called small gradient expansion, and is valid if sharp changes in the interface are absent [@marsili]. The other contribution to the dynamics comes from particle input $\Gamma_F=\hat{n}\cdot <\vec{F}>$, where $\vec{F}$ is the flux of cells generated at the interface. If we suppose that cell generation is an isotropic process we find $<\vec{F}>=F\hat{n}$, which implies $\Gamma_F=F$. However, the input of new cells $\vec{F}$ is a random process, giving rise to a stochastic contribution to the dynamics under the form of a noise $\eta=\hat{n}\cdot\vec{\Phi}$, where $\vec{\Phi}=\vec{F}-<\vec{F}>$. In summary, this implies that the stochastic force fulfills $\vec{\Phi}=\hat{n}\eta$ and $<\eta>=0$. Rearranging all the terms we recover the one-dimensional Mullins-Herring equation for MBE growth: $$\label{mbe} \frac{\partial h}{\partial t}=-K\frac{\partial^4 h}{\partial x^4}+F+\eta(x,t),$$ where the noise, $\eta(x,t)$, has been assumed to be gaussian with correlation given by $<\eta(x,t)\eta(x',t')>=D\delta(x-x')\delta(t-t')$. The drift of this equation comes originally from Eq.(\[surdiff\]), which expresses the “diffusion of the mean curvature” of the surface. This corresponds, also, to a homogeneization of the pressure; let us show this fact as follows. Normal cells adjacent to the tumor exert force against new born tumoral cells. Concavities are sourrended by a higher number of normal cells than convexities, and thus feel more pressure. Tumoral cells move along the tumor edge driven by the surface forces, what causes the effect of a diffusion: tumoral cells are redistributed from convexities to concavities (this fact will be shown explicitly below by means of linear stability analysis). The “equilibrium” distribution corresponds to the spherically symmetric form, which presumably implies the homogeneization of the pressure all along the tumor edge; this same form implies the minimization of the mean curvature of the surface. The specific form of the cuartic derivative in this equation has been deduced phenomenologically [@bru1; @bru2]. Now that we have identified the physical mechanisms that have led us to Eq.(\[mbe\]), we are in position to derive SPDEs describing the same physics but with geometrical properties compatible with those of a tumor. For the case of (1+1)-dimensional circular model in polar coordinates we have $\vec{r}=(r(\theta)\cos(\theta),r(\theta)\sin(\theta))$, the unitary normal vector reads $$\hat{n}=\frac{1}{\sqrt{r^2+(\partial_\theta r)^2}}(-r\sin(\theta),r\cos(\theta)),$$ the corresponding metric tensor is given this time by $\tilde{g}=r^2+(\partial_\theta r)^2$, and the mean curvature is $$H=\frac{r^2+2(\partial_\theta r)^2-r\partial_\theta^2 r}{[r^2 + (\partial_\theta r)^2]^{3/2}}.$$ We are now ready to derive the diffusive drift $$\Gamma_s=-\frac{K}{r^4} \left( \frac{\partial^2 r}{\partial \theta^2} + \frac{\partial^4 r}{\partial \theta^4} \right),$$ correspondingly linearized with respect to the different derivatives of $r(\theta)$. The term containing the second derivative of $r$ is irrelevant in the renormalization group sense, so we can neglect it to obtain $$\Gamma_s=-\frac{K}{r^4} \frac{\partial^4 r}{\partial \theta^4}.$$ The stochastic term comes from the force in the same way as in the last case $\vec{\Phi}=\hat{n}\eta$. The noise $\eta$ is a gaussian variable with zero mean and correlation given by $$<\eta({\bf s},t)\eta({\bf s'},t')>=n_\alpha({\bf s},t)n_\beta({\bf s},t)D^{\alpha \beta}\frac{\delta({\bf s}-{\bf s'})} {\sqrt{g}}\delta(t-t').$$ We now can write the SPDE for tumor growth in (1+1)-dimensions $$\label{tumor2d} \frac{\partial r}{\partial t}= -\frac{K}{r^4} \frac{\partial^4 r}{\partial \theta^4} + F + \frac{1}{\sqrt{r}}\eta(\theta,t),$$ where the noise $\eta(\theta,t)$ is gaussian, with zero mean, and correlation given by $<\eta(\theta,t)\eta(\theta',t')>=D\delta(\theta-\theta')\delta(t-t')$. As indicated above, we have assumed that cell generation is isotropic, which implies $D^{\alpha \beta}=D\delta^{\alpha \beta}$. It is important to note that this time the noise is multiplicative, and that it *must* be interpreted according to Itô, since all the deterministic contributions to the drift have been already extracted [@marsili]. Another desirable characteristic of this equation is that the variable $\theta$ only varies in $[0,2\pi]$ at any time, which represents an advantage with respect to using a different coordinate, as for instance the arc length. The arc length (a magnitude more similar to $x$ in Eq.(\[mbe\])) vary in an interval which depends on time (because the tumor grows), what makes more difficult to study the scaling properties of the model. We can determine the critical exponents by power counting. The arc lenght of a circumference is $l=r\theta$, and taking into account that it scales as $l \to bl$, we deduce that the angle scales as $\theta \to b^{1-\alpha}\theta$. The other two variables scale as $r \to b^\alpha r$ and $t \to b^z t$; direct substitution reveals that Eq.(\[tumor2d\]) is in the MBE universality class. Of course, the range of validity of this equation assumes that the interface shows neliglible overhangs in the radial direction compared with the size of the tumor, a fact that has been observed experimentally in many cases [@bru1]. Now we deal with a nonlinear equation, in contrast to Eq.(\[mbe\]), that cannot be solved simply by means of a Fourier transformation. We can instead employ different techniques in order to get some insight into it. For a big enough tumor we can approximate the first moment of the radius by the solution the mean-field version of Eq.(\[tumor2d\]), i.e., neglecting the noise term. On the other hand, the condition of possibility for formulating a continuous equation as an adequate description of a tumor is that it is composed by a sufficiently high number of cells. So describing the tumor with a continuous equation is in the same order of approximation of considering the mean-field level for computing the first moment of the radius. The first step in the analysis is to note that the deterministic version of Eq.(\[tumor2d\]) admits radially symmetric solutions of the form $r(\theta,t)=R(t)=Ft+R_0$, where $R_0$ is the radially symmetric initial condition. It is easy to show the linear stability of this solution by substituting $r(\theta,t)=R(t)+\rho(\theta,t)$ in Eq.(\[tumor2d\]) with $D=0$, where $\rho$ is a small perturbation. The resulting equation for $\rho$ is $$\label{fourier2d} \frac{\partial \rho}{\partial t}=\frac{-K}{(Ft+R_0)^4}\frac{\partial^4 \rho}{\partial \theta^4}.$$ Since the function $\rho$ is $2\pi$-periodic in the $\theta$ variable we can express it exactly in terms of a Fourier series $$\rho(\theta,t)=\sum_{n=-\infty}^{\infty}\rho_n(t)e^{in\theta},$$ and by direct substitution in Eq.(\[fourier2d\]) we see that the Fourier modes obey $$\frac{d\rho_n}{dt}=\frac{-Kn^4}{(Ft+R_0)^4}\rho_n.$$ We can integrate exactly this equation to obtain $$\label{perturbation2d} \rho_n(t)=\rho_n(t_0)\exp \left( \frac{-Kn^4}{3F} \left[\frac{1}{(R_0+Ft_0)^3}-\frac{1}{(R_0+Ft)^3} \right] \right),$$ where $$\rho_n(t_0)=\frac{1}{2\pi}\int_0^{2\pi}\rho(\theta,t_0)e^{-in\theta}d\theta,$$ and thus we see that the perturbation decreases in time provided $t>t_0$ [@note1]. It is also important to note that Eq.(\[perturbation2d\]) might be interpreted as the response of the tumor to an external perturbation. Stochastic generation of new cells drives the tumor away from the radially symmetric form, while surface diffusion tries to restore it; redistribution of cell density after radial symmetry breaking follows the law Eq.(\[perturbation2d\]). Our next step will be to derive the corresponding equation for the growth of a (2+1)-dimensional interface. We can parametrize the two-dimensional surface by means of the vector $\vec{r}=(r(\theta,\phi)\sin(\theta)\cos(\phi),r(\theta,\phi)\sin(\theta)\sin(\phi),r(\theta,\phi)\cos(\theta))$. Implying that the metric tensor is $$\tilde{g}=\left( \begin{array}{cc} r^2+(\partial_\theta r)^2 & \partial_\theta r \partial_\phi r \\ \partial_\theta r \partial_\phi r & r^2 \sin^2(\theta)+(\partial_\phi r)^2 \end{array} \right).$$ The mean curvature can be derived from the metric tensor, however, the expression is so cumbersome that it cannot be handled with simplicity. We can instead linearize this expression about the different derivatives of $r$ to get $$H=\frac{1}{r^2}\left(-2r+\frac{\partial_\theta r}{\tan(\theta)}+\partial_\theta^2 r+\frac{\partial_\phi^2 r}{\sin^2(\theta)}\right).$$ Collecting the results we see that for the (2+1)-dimensional model for spherical growth the SPDE reads $$\label{tumor3d} \frac{\partial r}{\partial t}= -\frac{K}{r^4}\left(\frac{\partial^4 r}{\partial \theta^4}+ \frac{2}{\sin^2(\theta)}\frac{\partial^4 r}{\partial \theta^2 \partial \phi^2}+ \frac{1}{\sin^4(\theta)}\frac{\partial^4 r}{\partial \phi^4}\right)+ F + \frac{1}{r\sqrt{|\sin(\theta)|}}\eta(\theta,\phi,t),$$ where the noise $\eta(\theta,\phi,t)$ is gaussian, with zero mean, and correlation given by $<\eta(\theta,\phi,t)\eta(\theta',\phi',t')>=D\delta(\theta-\theta')\delta(\phi-\phi')\delta(t-t')$, and it *must* be again interpreted according to Itô. As happened with the (1+1)-model, we get the desirable characteristic that the variables $\theta,\phi \in [0,2\pi]$. In this case we see again that there exists the radially symmetric solution for the deterministic version of Eq.(\[tumor3d\]): $r(\theta,\phi,t)=R(t)=Ft+R_0$, where $R_0$ is the radially symmetric initial condition. We can analize its linear stability by substituting the solution $r(\theta,\phi,t)=R(t)+\rho(\theta,\phi,t)$ in Eq.(\[tumor3d\]) with $D=0$, where $\rho$ is a small perturbation $2\pi$-periodic in both variables $\theta$ and $\phi$, and thus can be represented in the form of a Fourier series $$\rho(\theta,\phi,t)=\sum_{n,m=-\infty}^{\infty}\rho_{n,m}(t)e^{in\theta+im\phi}.$$ The Fourier modes obey the ordinary differential equation $$\frac{d\rho_{n,m}}{dt}=\frac{-K}{(Ft+R_0)^4}\left(n^4+\frac{8}{3}m^2n^2+\frac{8}{3}m^4\right)\rho_{n,m},$$ that can be integrated to yield $$\rho_{n,m}(t)=\rho_{n,m}(t_0)\exp \left(\frac{-K}{3F}\left[\frac{1}{(R_0+Ft_0)^3}-\frac{1}{(R_0+Ft)^3}\right] \left[n^4+\frac{8}{3}m^2n^2+\frac{8}{3}m^4\right]\right),$$ where $$\rho_{n,m}(t_0)=\frac{1}{4\pi^2}\int_0^{2\pi}\int_0^{2\pi}\rho(\theta,\phi,t)e^{-in\theta-im\phi}d\theta d\phi,$$ implying the stability of the solution provided $t>t_0$ [@note2]. This is, as in the former case, the law that imposes density redistribution after radial symmetry breaking due to stochastic generation of new cells. In conclusion, we have derived the equations of growth of the (1+1)- and (2+1)-dimensional tumor interfaces containing the physics of MBE in the appropiate geometry. These equations provide us a description of the tumor in a coordinate system that is polar, and since all the coordinates are angles we have the additional advantage that they vary in intervals that are independent of time. They correctly predict the constant velocity growth regime found experimentally during the initial phase of growth, and a linear stability analysis of radial solutions allowed us to quantitatively estimate the law of density distribution of new generated cells. However, latter stages of growth are characterized by certain decelaration of the growth rate; this fact is not captured by the present model, and it will be studied in the future. We have assumed all along this work that the tumor is composed of an enough large number of cells so that the hydrodynamic description by means of continuous equations makes sense. If we want to describe small tumors a kinetic approach to the problem becomes necessary, as for instance a master equation formulation. Master equation descriptions of growth models are already present in the literature [@sos], and may be adapted for the present case of a tumor. Furthermore, we can project the master equation for the cell population into a SPDE via field theoretic arguments [@escudero] in order to recover a more similar theoretical approach to the one presented here. These and other questions will be the object of future research. This work has been partially supported by the Ministerio de Ciencia y Tecnología (Spain) through Project No. BFM2001-0291 and by UNED. [99]{} A.-L. Barabási and H. E. Stanley, [*Fractal Concepts in Surface Growth*]{} (Cambridge University Press, Cambridge, England, 1995). M. Eden, in [*Symposium on Information Theory in Biology*]{}, edited by H. P. Yockey, (Pergamon Press, New York, 1958). A. Brú, J. M. Pastor, I. Fernaud, I. Brú, S. Melle, and C. Berenguer, Phys. Rev. Lett. [**81**]{}, 4008 (1998). A. Brú, S. Albertos, J. L. Subiza, J. L. García-Asenjo, and I. Brú, Biophys. J. [**85**]{}, 2948 (2003). A. Brú, S. Albertos, J. A. L. García-Asenjo, and I. Brú, Phys. Rev. Lett. [**92**]{}, 238101 (2004). A. Brú, S. Albertos, F. García-Hoz, and I. Brú, J. Clin. Res. [**8**]{}, 9 (2005). W. W. Mullins, J. Appl. Phys. [**28**]{}, 333 (1957); C. Herring, J. Appl. Phys. [**21**]{}, 301 (1950). J. M. López, M. A. Rodríguez, and R. Cuerno, Physica A [**246**]{}, 329 (1997). M. Marsili, A. Maritan, F. Toigo, and J. R. Banavar, Rev. Mod. Phys. [**68**]{}, 936 (1996). Except for the mode $n=0$, which is trivial since it implies perturbations homogeneous in $\theta$. Except for the mode $m=n=0$, which is trivial since it implies perturbations homogeneous in $\theta$ and $\phi$. D. D. Vvedensky, A. Zangwill, C. N. Luse, and M. R. Wilby, Phys. Rev. E [**48**]{}, 852 (1993). C. Escudero, J. Buceta, F. J. de la Rubia, and K. Lindenberg, Phys. Rev. E [**69**]{}, 021908 (2004).
{ "pile_set_name": "ArXiv" }
--- abstract: 'The classical matrix-tree theorem discovered by G.Kirchhoff in 1847 relates the principal minor of the $n \times n$ Laplace matrix to a particular sum of monomials of matrix elements indexed by directed trees with $n$ vertices and a single sink. In this paper we consider a generalization of this statement: for any $k \ge n$ we define a degree $k$ polynomial $\det_{n,k}$ of matrix elements and prove that this polynomial applied to the Laplace matrix gives a sum of monomials indexed by acyclic graphs with $n$ vertices and $k$ edges.' address: 'National Research University Higher School of Economics, 119048, 6 Usacheva str., Moscow, Russia, and Independent University of Moscow, 119002, 11 B.Vlassievsky per., Moscow, Russia' author: - Yurii Burman title: 'Abstract matrix-tree theorem' --- Introduction and the main results ================================= Principal definitions {#SSec:Def} --------------------- Denote by $\Gamma_{n,k}$ the set of all directed graphs with $n$ vertices numbered $1 {, \dots ,} n$ and $k$ edges numbered $1 {, \dots ,} k$. We will write $e = [ab]$ if $e$ is an edge from vertex $a$ to vertex $b$; in particular, $[aa]$ means a loop attached to the vertex $a$. We will treat elements of $\Gamma_{n,k}$ as sequences of edges: $G = (e_1 {, \dots ,} e_k) \in \Gamma_{n,k}$ means a graph where the edge $e_\ell$ has number $\ell$, for all $\ell = 1 {, \dots ,} k$. By a slight abuse of notation $e \in G$ will mean that $e$ is an edge of $G$ (regardless of number). Let $G \in \Gamma_{n,k}$ and $e \in G$. By $G \setminus e$, $G/e$ and $G_e^{\vee}$ we will denote the graph $G$ with $e$ deleted, $e$ contracted and $e$ reversed, respectively. Note for correctness that since $G \setminus e \in \Gamma_{n,k-1}$, one has to change the edge numbering in $G$ after deleting $e$: namely, if $e$ bears number $s$ in $G$ then the numbers of the edges are preserved if they are less than $s$ and lowered by $1$ otherwise. For $G/e \in \Gamma_{n-1,k-1}$ the same renumbering is applied both to the edges and to the vertices. The contracted edge $e$ should not be a loop. A graph $H \in \Gamma_{n,m}$ is called a subgraph of $G \in \Gamma_{n,k}$ (notation $H \subseteq G$) if $H$ is obtained from $G$ by deletion of several (possibly zero) edges. Denote by $\Graph_{n,k}$ a vector space over $\Complex$ spanned by $\Gamma_{n,k}$. The direct sum $\Graph_n {\stackrel{\mbox{\scriptsize def}}{=}}\bigoplus_{k=0}^\infty \Graph_{n,k}$ bears the structure of an associative algebra: one defines a product of the graphs $G_1 = (e_1 {, \dots ,} e_{k_1}) \in \Gamma_{n,k_1}$ and $G_2 = (h_1 {, \dots ,} h_{k_2}) \in \Gamma_{n,k_2}$ as $G_1*G_2 {\stackrel{\mbox{\scriptsize def}}{=}}(e_1 {, \dots ,} e_{k_1}, h_1 {, \dots ,} h_{k_2}) \in \Gamma_{n,k_1+k_2}$; then $*$ is extended to the whole $\Graph_n$ as a bilinear operation. Note that $G_1*G_2 \ne G_2*G_1$ (the edges are the same but the edge numbering is different), so the algebra $\Graph_n$ is not commutative. We call a graph $G \in \Gamma_{n,k}$ [*strongly connected*]{} if every two its vertices can be joined by a directed path. A graph is [*strongly semiconnected*]{} if every its connected component (in the topological sense) is strongly connected; equivalently, if every its edge is a part of a directed cycle. A strongly semiconnected graph may contain isolated vertices (i.e. vertices not incident to any edge); by $\SSC_{n,k}^{\{i_1 {, \dots ,} i_s\}}$ we denote the set of strongly semiconnected graphs $G \in \Gamma_{n,k}$ such that the vertices $i_1 {, \dots ,} i_s$, and only they, are isolated. By $\SSC_{n,k} {\stackrel{\mbox{\scriptsize def}}{=}}\bigcup_{I \subset \{1 {, \dots ,} n\}} \SSC_{n,k}^I$ we will denote the set of all strongly semiconnected graphs. We call a graph $G \in \Gamma_{n,k}$ [*acyclic*]{} if it contains no directed cycles. Recall that a vertex $a$ of the graph $G$ is called a [*sink*]{} if $G$ has no edges starting from $a$. Note that an isolated vertex is a sink but a vertex with a loop attached to it is not. We denote by $\AC_{n,k}^{\{i_1 {, \dots ,} i_s\}}$ the set of acyclic graphs $G \in \Gamma_{n,k}$ such that the vertices $i_1 {, \dots ,} i_s$, and only they, are sinks. By $\AC_{n,k} {\stackrel{\mbox{\scriptsize def}}{=}}\bigcup_{I \subset \{1 {, \dots ,} n\}} \AC_{n,k}^I$ we will denote the set of all acyclic graphs. \[Ex:SSC\] If a vertex of a strongly semiconnected graph $G \in \SSC_{n,k}^I$ is not isolated then there is at least one edge starting from it; so if $I = \{i_1 {, \dots ,} i_s\}$ and $\SSC_{n,k}^I \ne {\varnothing}$ then $k \ge n-s$. Let $k=n-s$. If $G \in \Gamma_{n,k}^I$ then for any vertex $i \notin I$ there is exactly one edge $[i, \sigma(i)]$ starting at it and exactly one edge $[j, \sigma(j)] = [j,i]$ finishing at it (that is, $\sigma(j) = i$). Hence $\sigma$ is a bijection $\{1 {, \dots ,} n\} \setminus I \to \{1 {, \dots ,} n\} \setminus I$ (a permutation of $k = n-s$ points). Geometrically $G$ is a union of disjoint directed cycles passing through all vertices except $i_1 {, \dots ,} i_s$. \[Ex:Forest\] Let $n > k$; then any graph $G \in \Gamma_{n,k}$ contains at least $n-k$ connected components. If $G$ is acyclic then every its connected component contains a sink. So for $I = \{i_1 {, \dots ,} i_s\}$ if $\AC_{n,k}^I \ne {\varnothing}$ then $k \ge n-s$. Let $k = n-s$. Then the elements of $\AC_{n,k}^I$ are forests of $s$ components, each component containing exactly one vertex $i_\ell \in I$ (for some $\ell = 1 {, \dots ,} s$), which is its only sink. This component is a tree and every its edge is directed towards the sink $i_\ell$. Determinants and minors ----------------------- Let $W = (w_{ij})$ be a $n \times n$-matrix; denote by $\langle W \vert: \Graph_{n,k} \to \Complex$ a linear functional acting on the basic element $G \in \Gamma_{n,k}$ as $$\langle W \mid G\rangle {\stackrel{\mbox{\scriptsize def}}{=}}\prod_{[ij] \in G} w_{ij}.$$ Note that $\langle W \mid G\rangle$ is independent of the edge numbering in $G$; in particular, $\langle W \mid G_1*G_2 - G_2*G_1\rangle = 0$ for all $G_1, G_2$. For a function $f: \bigcup_s \Gamma_{n,s} \to \Complex$ and a graph $G \in \Gamma_{n,k}$ introduce the notation $$\label{Eq:SumSubgr} \SumSub(f;G) {\stackrel{\mbox{\scriptsize def}}{=}}\sum_{H \subseteq G} f(H).$$ For a set of graphs $\mathfrak B \subset \Gamma_{n,k}$ denote $$\begin{aligned} &U(\mathfrak B) {\stackrel{\mbox{\scriptsize def}}{=}}\sum_{G \in \mathfrak B} G \in \Graph_{n,k}, \\ &X(\mathfrak B) {\stackrel{\mbox{\scriptsize def}}{=}}\sum_{G \in \mathfrak B} (-1)^{\beta_0(G)} G \in \Graph_{n,k}; \end{aligned}$$ $\beta_0(G)$ here means the $0$-th Betti number of $G$, i.e. the number of its connected components (in the topological sense). \[Df:Minors\] The element $$\det_{n,k}^I {\stackrel{\mbox{\scriptsize def}}{=}}\frac{(-1)^k}{k!} X(\SSC_{n,k}^I) \in \Graph_{n,k}$$ is called a [*universal diagonal $I$-minor*]{} of degree $k$; in particular, $\det_{n,k}^{\varnothing}$ is called a [*universal determinant*]{} of degree $k$. The element $$\det_{n,k}^{i/j} {\stackrel{\mbox{\scriptsize def}}{=}}\frac{(-1)^k}{k!} X(\{G \in \Graph_{n,k} \mid ([ij])*G \in \SSC_{n,k+1}^{\varnothing}\})$$ is called a [*universal (codimension $1$) $(i,j)$-minor*]{} of degree $k$. \[Ex:Det\] Example \[Ex:SSC\] implies that if $I = \{i_1 {, \dots ,} i_s\}$ and $k < n-s$ then $\det_{n,k}^I = 0$. Let $k = n$ and $I = {\varnothing}$. By Example \[Ex:SSC\] the graphs $G \in \SSC_{n,n}^{\varnothing}$ are in one-to-one correspondence with permutations $\sigma$ of $\{1 {, \dots ,} n\}$. It is easy to see that $(-1)^{\beta_0(G)}$ is equal to $(-1)^n$ if $\sigma$ is even and to $-(-1)^n$ if it is odd. Geometrically $G$ is a union of disjoint directed cycles. If the order of vertices in all the cycles is fixed, then there are $n!$ ways to assign numbers $\{1 {, \dots ,} n\}$ to the edges; this implies the equality $$\langle W \mid \det_{n,n}^{\varnothing}\rangle = \sum_\sigma (-1)^{\text{parity of $\sigma$}} w_{1\sigma(1)} \dots w_{n \sigma(n)} = \det W$$ for any matrix $W = (w_{ij})$. Similarly, for any set $I = \{i_1 {, \dots ,} i_s\}$ the value $\nolinebreak[0]\langle W \mid \det_{n,n-s}^I\rangle$ is equal to the diagonal minor of the matrix $W$ obtained by deletion of the rows and the columns $i_1 {, \dots ,} i_s$. Also $\langle W \mid \det_{n,n-1}^{i/j}\rangle$ is equal to the codimension $1$ minor of $W$ obtained by deletion of the row $i$ and the column $j$. This explains the terminology of Definition \[Df:Minors\]. The elements $\det_{n,k}^I$ exhibit some properties one would expect from determinants and minors: \[Pp:DetProp\] 1. \[It:RowCol\] (generalized row and column expansion) $$\label{Eq:DetViaMinors} \det_{n,k}^{\varnothing}= \frac{1}{k} \sum_{i,j=1}^n ([ij]) * \det_{n,k-1}^{i/j}.$$ 2. \[It:PDer\] (partial derivative with respect to a diagonal matrix element) Let matrix elements $w_{ij}$, $i,j = 1 {, \dots ,} n$, of the matrix $W$ be independent (commuting) variables. Then for any $i = 1 {, \dots ,} n$ and any $m = 1 {, \dots ,} k$ one has $$\label{Eq:Deriv} \frac{\partial^m}{\partial w_{ii}^m} \langle W \mid \det_{n,k}^{\varnothing}\rangle = \langle W \mid \det_{n,k-m}^{\varnothing}+ \det^{\{i\}}_{n,k-m}\rangle.$$ See [@Epstein Lemma 86] for a formula similar to (with $m=1$ and a finite difference instead of a derivative). Main results {#SSec:Main} ------------ Let $G \in \Gamma_{n,k}$, $p \in \{1 {, \dots ,} k\}$ and $i,j \in \{1 {, \dots ,} n\}$. Denote by $R_{ab;p} G \in \Gamma_{n,k}$ the graph obtained from $G$ by replacement of its $p$-th edge by the edge $[ab]$ bearing the same number $p$. Consider now a linear operator $B_p: \Graph_{n,k} \to \Graph_{n,k}$ acting on every basic element $G \in \Gamma_{n,k}$ as follows: $$B_p(G) = \begin{cases} G, &\text{if the $p$-th edge of $G$ is not a loop},\\ -\sum_{b \ne a} R_{ab;p} G, &\text{if the $p$-th edge of $G$ is the loop $[aa]$}. \end{cases}$$ In particular, $B_p = 0$ if $n=1$ (and $k > 0$). \[Df:Laplace\] The product $\Lapl {\stackrel{\mbox{\scriptsize def}}{=}}B_1 \dots B_k: \Graph_{n,k} \to \Graph_{n,k}$ is called [*Laplace operator*]{}. If $n = 1$ and $k > 0$ then $\Delta = 0$; also take $\Delta = \name{id}$ by definition if $k = 0$. The operators $B_p$, $p = 1 {, \dots ,} k$, are commuting idempotents: $B_p^2 = B_p$ and $B_p B_q = B_q B_p$ for all $p, q = 1 {, \dots ,} k$. Therefore, $\Lapl$ is an idempotent, too: $\Lapl^2 = \Lapl$. Let $W = (w_{ij})_{i,j=1}^n$ be a $n \times n$-matrix, like in Example \[Ex:Det\] and Proposition \[Pp:DetProp\]. Denote by $\widehat{W}$ the corresponding Laplace matrix, i.e. a matrix with nondiagonal elements $w_{ij}$ ($1 \le i \ne j \le n$) and diagonal elements $-\sum_{j \ne i} w_{ij}$ ($1 \le i \le n$). It follows from Definition \[Df:Laplace\] that $$\langle \widehat{W} \mid X \rangle = \langle W \mid \Lapl(X) \rangle$$ for any $X \in \Graph_{n,k}$. This equation explains the name “Laplace operator” for $\Lapl$. Note that since $\Lapl(X)$ is a sum of graphs containing no loops, one is free to change diagonal entries of $W$ in the right-hand side; in particular, one can use $\widehat{W}$ instead. The main results of this paper are the following two theorems: \[Th:Diag\] $$\label{Eq:LaplDiag} \Lapl(\det_{n,k}^I) = \frac{(-1)^n}{k!} U(\AC_{n,k}^I).$$ and \[Th:Codim1\] $$\label{Eq:LaplCodim1} \Lapl(\det_{n,k}^{i/j}) = \frac{(-1)^n}{k!} U(\AC_{n,k}^{\{i\}}).$$ Applying the functional $\langle \widehat{W}\vert\relax$ to equation with $k = n-s$ and to equation with $k = n-1$ and using Examples \[Ex:Det\] and \[Ex:Forest\] one obtains \[Cr:Diag\] The diagonal minor of the Laplace matrix obtained by deletion of the rows and columns numbered $i_1 {, \dots ,} i_s$ is equal to $\frac{(-1)^{n-s}}{(n-s)!}\langle W \mid U(\AC_{n,n-s}^I)\rangle$, that is, to $(-1)^{n-s}$ times the sum of monomials $w_{a_1 b_1} \dots w_{a_{n-s} b_{n-s}}$ such that the graph $([a_1 b_1] {, \dots ,} [a_{n-s} b_{n-s}])$ is a $s$-component forest where every component contains exactly one vertex $i_\ell$ for some $\ell = 1 {, \dots ,} s$, and all the edges of the component are directed towards $i_\ell$. and \[Cr:Codim1\] The minor of the Laplace matrix obtained by deletion of its $i$-th row and its $j$-th column is equal to $(-1)^{n-1}$ times the sum of monomials $w_{a_1 b_1} \dots w_{a_{n-1} b_{n-1}}$ such that the graph $([a_1 b_1] {, \dots ,} [a_{n-1} b_{n-1}])$ is a tree with all the edges directed towards the vertex $i$. Corollaries \[Cr:Diag\] and \[Cr:Codim1\] are particular cases of the celebrated matrix-tree theorem first discovered by G.Kirchhoff [@Kirch] in 1847 (for symmetric matrices and diagonal minors of codimension $1$) and proved in its present form by W.Tutte [@TutteMTT]. Consider now the following functions on $\Gamma_{n,k}$: $$\begin{aligned} \sigma(G) &= \begin{cases} (-1)^{\beta_1(G)}, &G \in \Gamma_{n,k} \text{ is strongly semiconnected},\\ 0 &\text{otherwise}, \end{cases}\\ \text{and\hspace{3.5cm}}\\ \alpha(G) &= \begin{cases} (-1)^k, &G \in \Gamma_{n,k} \text{ is acyclic},\\ 0 &\text{otherwise}. \end{cases} \end{aligned}$$ Theorem \[Th:Diag\] follows from the two equivalent statements (see Section \[Sec:Proofs\] for details): \[Th:Direct\] $\SumSub(\alpha;G) = (-1)^k \sigma(G)$ for $G \in \Gamma_{n,k}$. $\SumSub(\sigma;G) = (-1)^k\alpha(G)$ for $G \in \Gamma_{n,k}$. Here $\SumSub$ is summation over subgraphs, as defined by . These theorems are essentially [@Bernardi Proposition 6.16]. We will nevertheless give their proofs in Section \[Sec:Proofs\] thus answering a request for a direct proof expressed in [@Bernardi] (the original proof in [@Bernardi] is a specialization of a much more general identity). A digression: undirected graphs and the universal Potts partition function -------------------------------------------------------------------------- Denote by $\Upsilon_{n,k}$ the set of all [*undirected*]{} graphs with $n$ vertices numbered $1 {, \dots ,} n$ and $k$ edges numbered $1 {, \dots ,} k$. Denote by $\lmod \,\cdot\,\rmod: \Gamma_{n,k} \to \Upsilon_{n,k}$ the “forgetful” map replacing every edge by its undirected version; the edge numbering is preserved. By $\Undir_{n,k}$ denote a vector space spanned by $\Upsilon_{n,k}$; then $\lmod \,\cdot\,\rmod$ is extended to the linear map $\Graph_{n,k} \to \Undir_{n,k}$. The notion of a subgraph and the notation $\SumSub$ (see ) for undirected graphs are similar to those for $\Graph_{n,k}$. One can also define the operators $B_p: \Undir_{n,k} \to \Undir_{n,k}$, $p = 1 {, \dots ,} k$, and the Laplace operator $\Delta: \Undir_{n,k} \to \Undir_{n,k}$ for undirected graphs exactly as in Definition \[Df:Laplace\]. For any $G \in \Undir_{n,k}$ consider the two-variable polynomial: $$\label{Eq:DefPotts} Z_G(q,v) = \SumSub(q^{\beta_0(H)} v^{\#\text{of edges of $H$}};G).$$ called [*Potts partition function*]{}. It is related [@Sokal Eq.(2.26)] to the Tutte polynomial $T_G$ of the graph $G$ as $$T_G(x,y) = (x-1)^{-\beta_0(G)} (y-1)^{-n} Z((x-1)(y-1), y-1;G).$$ Values of $Z_G$ in some points have a special combinatorial interpretation, in particular \[Pp:SpecVal\] $$\begin{aligned} Z_G(-1,1) &= (-1)^{\beta_0(G)} 2^{\#\text{\upshape of loops of $G$}} \# \{\Phi \in \SSC_{n,k} \subset \Gamma_{n,k} \mid \lmod \Phi\rmod = G\}.\\ Z_G(-1,-1) &= (-1)^n \# \{\Phi \in \AC_{n,k} \subset \Gamma_{n,k} \mid \lmod \Phi\rmod = G\}. \end{aligned}$$ (Recall that by $\SSC_{n,k}$ and $\AC_{n,k}$ we denote the sets of all strongly semiconnected and acyclic graphs in $\Gamma_{n,k}$, respectively.) \[Cr:NumSSC\] For any graph $G \in \Upsilon_{n,k}$ one has $$\# \{\Phi \in \SSC_{n,k} \subset \Gamma_{n,k} \mid \lmod \Phi\rmod = G\} = (-1)^{\beta_0(G)} Z_{\widehat G}(-1,1)$$ where $\widehat G$ is the graph $G$ with all the loops deleted. The definition of the Potts partition function implies immediately that $Z_G(q,v) = (v+1)^{\#\text{of loops of $G$}} Z_{\widehat G}(q,v)$. Consider now the [*universal Potts partition function*]{} $${\mathcal Z}_{n,k}(q,v) {\stackrel{\mbox{\scriptsize def}}{=}}\sum_{G \in \Upsilon_{n,k}} Z_G(q,v)G \in \Undir_{n,k}$$ and its “shaved” version $$\widehat{{\mathcal Z}}_{n,k}(q,v) {\stackrel{\mbox{\scriptsize def}}{=}}\sum_{G \in \Upsilon_{n,k}} Z_{\widehat G}(q,v)G \in \Undir_{n,k}.$$ \[Pp:LaplTutte\] $$\Delta\widehat{{\mathcal Z}}_{n,k}(-1,1) = (-1)^k {\mathcal Z}_{n,k}(-1,-1).$$ Note that by Proposition \[Pp:SpecVal\] the right-hand side of the equality contains only graphs without loops, as does the left-hand side. Corollary \[Cr:NumSSC\] implies that $$\widehat{{\mathcal Z}}_{n,k}(-1,1) = (-1)^k k!\sum_{I \subset \{1 {, \dots ,} n\}} \lmod \det_{n,k}^I\rmod.$$ Apply now the Laplace operator $\Delta$ to both sides of the equality. Apparently, $\Delta$ commutes with the forgetful map: $\lmod \Delta(x)\rmod = \Delta(\lmod x\rmod)$ for any $x \in \Graph_{n,k}$. Therefore by Theorem \[Th:Diag\] and Proposition \[Pp:SpecVal\] $$\begin{aligned} \Delta\widehat{{\mathcal Z}}_{n,k}(-1,1) &= (-1)^k k!\sum_{I \subset \{1 {, \dots ,} n\}} \lmod \Delta \det_{n,k}^I\rmod = (-1)^{n+k} \sum_{I \subset \{1 {, \dots ,} n\}} \lmod U(\AC_{n,k}^I)\rmod \\ &= (-1)^k {\mathcal Z}_{n,k}(-1,-1). \end{aligned}$$ Proposition \[Pp:LaplTutte\] admits several generalizations. The author is planning to write a separate paper considering action of the Laplace operator on the universal Potts functions and their oriented-graph versions. An application: invariants of $3$-manifolds ------------------------------------------- Universal determinants have an application in $3$-dimensional topology, due to M.Polyak. We describe it briefly here; see [@Toronto] and the MSc. thesis [@Epstein] for detailed definitions, formulations and proofs. A [*chainmail graph*]{} is defined as a planar graph, possibly with loops but without parallel edges; the edges (including loops) are supplied with integer weights. We denote by $w_{ij} = w_{ji}$ the weight of the edge joining vertices $i$ and $j$; $w_{ii}$ is the weight of the loop attached to the vertex $i$. If the edge $[ij]$ is missing then $w_{ij} = 0$ by definition. There is a way (see [@Toronto]) to define for every chainmail graph $G$ a closed oriented $3$-manifold $\Manif(G)$; any closed oriented $3$-manifold is $\Manif(G)$ for some $G$ (which is not unique). To the chainmail graph $G$ with $n$ vertices one associates two $n \times n$-matrices: the adjacency matrix $W(G) = (w_{ij})$ and the Laplace (better to say, Schroedinger) matrix $L(G) = (l_{ij})$ where $l_{ij} {\stackrel{\mbox{\scriptsize def}}{=}}w_{ij}$ for $i \ne j$ and $l_{ii} {\stackrel{\mbox{\scriptsize def}}{=}}w_{ii} - \sum_{j \ne i} w_{ij}$. If all $w_{ii} = 0$ (such $G$ is called a balanced graph) then $L(G)$ is the usual (symmetric, degenerate) Laplace matrix $\widehat{W}$ from Section \[SSec:Main\]. 1. The rank of the homology group $H_1(\Manif(G),\Integer)$ is equal to $\dim \name{Ker} L(G)$. 2. If $L(G)$ is nondegenerate (so that $\Manif(G)$ is a rational homology sphere and $H_1(\Manif(G),\Integer)$ is finite) then $$\label{Eq:3ManifDet} \lmod H_1(\Manif(G),\Integer)\rmod = \lmod \det L(G)\rmod = \lmod\langle L(G) \mid \det_{n,n}^{\varnothing}\rangle\rmod.$$ 3. If $L(G)$ is nondegenerate then $$\label{Eq:3ManifTheta} \langle W(G) \mid \Theta_n\rangle = 12 \det L(G) \bigl(\lambda_{CW}(\Manif(G)) - \frac{1}{4} \name{sign}(L(G))\bigr)$$ where $\lambda_{CW}$ is the Casson–Walker invariant [@Walker] of the raional homology sphere $\Manif(G)$, $\name{sign}$ is the signature of the symmetric matrix $L(G)$, and $\Theta_n$ is an element of $\Graph_{n,n+1} \oplus \Graph_{n,n-1}$ defined as $$\Theta_n \stackrel{\mathrm{def}}{=}\det_{n,n+1}^{\varnothing}- \sum_{1 \le i \ne j \le n} ([ij])*\det_{n,n-2}^{\{i,j\}} - \sum_{i=1}^n \det_{n,n-1}^{\{i\}}.$$ Conjecturally, and begin a series of formulas for invariants of $3$-manifolds. See [@Toronto] for details. Applying $\Delta$ to the element $\Theta_n$ and using Theorem \[Th:Diag\] and Corollary \[Cr:Diag\] one obtains $\Delta \Theta_n = -2 U(\AC_{n,n-1})$. Therefore if $G$ is balanced then $\langle L(G) \mid \Theta_n\rangle$ is equal to $-2$ times the codimension $1$ diagonal minor of $L(G)$. The last assertion is [@Epstein Theorem 84]. Acknowledgements {#acknowledgements .unnumbered} ---------------- The research was inspired by numerous discussions with prof. Michael Polyak (Haifa Technion, Israel) whom the author wishes to express his most sincere gratitude. The research was funded by the Russian Academic Excellence Project ‘5-100’ and by the grant No. 15-01-0031 “Hurwitz numbers and graph isomorphism” of the Scientific Fund of the Higher School of Economics. Proofs {#Sec:Proofs} ====== We start with proving Proposition \[Pp:DetProp\] (Section \[SSec:PrDetProp\]), to continue with Theorems \[Th:Direct\] and \[Th:Direct\]’ (Sections \[SSec:Equiv\] and \[SSec:PrDirect\]). Theorem \[Th:Diag\] will then follow from Theorem \[Th:Direct\]’ (Section \[SSec:DiagFromDirect\]), and Theorem \[Th:Codim1\], from Theorem \[Th:Diag\] and assertion \[It:RowCol\] of Proposition \[Pp:DetProp\] (Section \[SSec:PrCodim1\]). For two vertices $a, b \in G \in \Gamma_{n,k}$ we will write $a \succeq b$ if $G$ contains a directed path starting at $a$ and finishing at $b$; also $a \succeq a$ for any $a$ by definition. Proof of Proposition \[Pp:DetProp\] {#SSec:PrDetProp} ----------------------------------- Let $G$ be a strongly semiconnected graph, and $[ij] \in G$ be its edge carrying number $1$. Since $G$ is strongly semiconnected, $j \succeq i$ in $G \setminus ([ij])$, and therefore $\beta_0(G \setminus ([ij])) = \beta_0(G)$. Then $G$ enters the left-hand side and the $(i,j)$-th term of the sum at the right-hand side of with the same coefficient. Denote by $\SSC_{n,k}^{[i:q]}$ the set of all graphs $G \in \SSC_{n,k}^{\varnothing}$ having $q$ loops ($0 \le q \le k$) attached to vertex $i$. The graph $\hat G$ obtained from $G$ by deletion of all these loops (with the relevant renumbering of the remaining edges) belongs either to $\SSC_{n,k-q}^{[i:0]} \subset \SSC_{n,k-q}^{\varnothing}$ or, if $q > 0$, to $\SSC_{n,k-q}^{\{i\}}$. Vice versa, if $q > 0$ and $\hat G \in \SSC_{n,k-q}^{[i:0]} \cup \SSC_{n,k-q}^{\{i\}}$ then $G \in \SSC_{n,k}^{[i:q]}$. Deletion of a loop does not break a graph, so $\beta_0(G) = \beta_0(\hat G)$. If $G \in \SSC_{n,k}^{[i:q]}$ then there are $\binom{k}{q}$ ways to assign numbers to the loops of $G$ attached to $i$. Since $\langle W \mid G\rangle$ does not depend on the edge numbering, one has for $q > 0$ $$\langle W \mid X(\SSC_{n,k}^{[i:q]})\rangle = \binom{k}{q} w_{ii}^q \langle W \mid X(\SSC_{n,k-q}^{[i:0]}) + X(\SSC_{n,k-q}^{\{i\}})\rangle,$$ so that $$\begin{aligned} \langle W \mid \det_{n,k}^{\varnothing}\rangle &= \frac{1}{k!}\sum_{q=0}^k \langle W \mid X(\SSC_{n,k}^{[i:q]})\rangle\nonumber \\ &= \frac{1}{k!} \langle W \mid X(\SSC_{n,k}^{[i:0]})\rangle + \sum_{q=1}^k \frac{w_{ii}^q}{q!(k-q)!} \langle W \mid X(\SSC_{n,k-q}^{[i:0]}) + X(\SSC_{n,k-q}^{\{i\}})\rangle\nonumber \\ &= \sum_{q=0}^k \frac{w_{ii}^q}{q!(k-q)!} \langle W \mid X(\SSC_{n,k-q}^{[i:0]}) + X(\SSC_{n,k-q}^{\{i\}})\rangle - \langle W \mid \det_{n,k}^{\{i\}}\rangle.\label{Eq:Develop} \end{aligned}$$ The expressions $\langle W \mid X(\SSC_{n,k-q}^{[i:0]}) + X(\SSC_{n,k-q}^{\{i\}})\rangle$ and $\langle W \mid \det_{n,k}^{\{i\}}\rangle$ do not contain $w_{ii}$. So, applying the operator $\frac{\partial^m}{\partial w_{ii}^m}$ to equation and using the equation again with $k-m$ in place of $k$ one gets . Theorems \[Th:Direct\] and \[Th:Direct\]’ are equivalent. {#SSec:Equiv} --------------------------------------------------------- Denote by $E(G)$ the set of edges of the graph $G \in \Gamma_{n,k}$. The functions $\alpha$ and $\sigma$ do not depend on the edge numbering; so the summation in the left-hand side of both theorems is performed over the set $2^{E(G)}$ of subsets of $E(G)$. The equivalence of the theorems is now a particular case of the Moebius inversion formula [@Rota]. Namely, for any finite set $X$ the Moebius function of the set $2^X$ partially ordered by inclusion is $\mu(S,T) = (-1)^{\#(S \setminus T)}$, where $S, T \subseteq X$. Therefore one has $$\begin{aligned} \SumSub(&\sigma;G) = (-1)^k \alpha(G) \\ &\Longleftrightarrow \SumSub(\mu(G,H) (-1)^{\#\text{edges of H}} \alpha(H);G) = \sigma(G) \\ &\Longleftrightarrow \SumSub((-1)^{k-\#\text{edges of H}} (-1)^{\#\text{edges of H}}\alpha(H);G) = \sigma(G) \\ &\Longleftrightarrow \SumSub(\alpha(H);G) = (-1)^k \sigma(G). \end{aligned}$$ Proof of Theorem \[Th:Direct\] {#SSec:PrDirect} ------------------------------ To prove the theorem we use simultaneous induction by the number of vertices and the number of edges of the graph $G$. If $\mathcal R$ is some set of subgraphs of $G$ (different in different cases) and $\chi_{\mathcal R}$ is the characteristic function of this set then for convenience we will write $\SumSub(f,\mathcal R) {\stackrel{\mbox{\scriptsize def}}{=}}\SumSub(f \chi_{\mathcal R},G) = \sum_{H \in \mathcal R} f(H)$ for any function $f$ on the set of subgraphs. Consider now the following cases: ### $G$ is disconnected. Let $G = G_1 {\sqcup \dots \sqcup} G_m$ where $G_i$ are connected components. A subgraph $H \subset G$ is acyclic if and only if the intersection $H_i {\stackrel{\mbox{\scriptsize def}}{=}}H \cap G_i$ is acyclic for all $i$. Hence $\alpha(H) = \alpha(H_1) \dots \alpha(H_m)$, and therefore $\SumSub(\alpha,G) = \SumSub(\alpha, G_1) \dots \SumSub(\alpha, G_m)$. By the induction hypothesis $\SumSub(\alpha, G_i) = (-1)^{k_i}\sigma(G_i)$ where $k_i$ is the number of edges of $G_i$. So $$\SumSub(\alpha,G) = \SumSub(\alpha, G_1) \dots \SumSub(\alpha, G_m) = (-1)^{k_1 {+ \dots +} k_m} \sigma(G_1) \dots \sigma(G_m) = (-1)^k \sigma(G).$$ Now it will suffice to prove Theorem \[Th:Direct\] for connected graphs $G$. ### $G$ is connected and not strongly connected. {#SSec:NotSSC} In this case $G$ contains an edge $e$ which is not contained in any directed cycle. For such $e$ if $H \subset G$ is acyclic and $e \notin H$ then $H \cup \{e\}$ is acyclic, too. The converse is true for any $e$: if an acyclic $H \subset G$ contains $e$ then $H \setminus \{e\}$ is acyclic. Therefore $$\SumSub(\alpha,G) = \sum_{\substack{H \subset G \setminus \{e\},\\ H \text{ is acyclic}}} (-1)^{\#\text{edges of $H$}} + (-1)^{\#\text{edges of $H \cup \{e\}$}} = 0 = \sigma(G).$$ So it will suffice to prove Theorem \[Th:Direct\] for strongly connected graphs $G$. ### $G$ is strongly connected and contains a crucial edge. {#SSec:SSCCrucial} We call an edge $e$ of a strongly connected graph $G$ crucial if $G \setminus \{e\}$ is not strongly connected. Suppose $e = [ab] \in G$ is a crucial edge. Denote by $\mathcal R_e^-$ (resp., $\mathcal R_e^+$) the set of all subgraphs $H \subset G$ such that $e \notin H$ (resp., $e \in H$). Let $H \in \mathcal R_e^-$ be acyclic. Since $G \setminus \{e\}$ is not strongly connected and contains one edge less than $G$, one has by Clause \[SSec:NotSSC\] above $$\label{Eq:SumNoE} \SumSub(\alpha,\mathcal R_e^-) = \SumSub(\alpha,G \setminus \{e\}) = 0.$$ Let now $H \in \mathcal R_e^+$ be acyclic; such $H$ contains no directed paths joining $b$ with $a$. Since $G \setminus \{e\}$ is not strongly connected, $G \setminus \{e\}$ does not contain a directed path joining $a$ with $b$ either. It means that such path in $H$ will necessarily contain $e$, and therefore the graph $H/e \subset G/e$ (obtained by contraction of the edge $e$) is acyclic. The converse is true for any $e$: if $e \in H$ and $H/e \subset G/e$ is acyclic then $H \subset G$ is acyclic, too. The graph $G/e$ is strongly connected, contains one edge less (and one vertex less) than $G$, and $\beta_1(G/e) = \beta_1(G)$, so $\sigma(G/e) = \sigma(G)$. The graph $H/e$ contains one edge less than $H$, so $\alpha(H/e) = -\alpha(H)$. Now by the induction hypothesis $$\SumSub(\alpha,R_e^+) = -\SumSub(\alpha,G/e) = -(-1)^{k-1} \sigma(G/e) = (-1)^k \sigma(G),$$ and then implies $$\SumSub(\alpha,G) = \SumSub(\alpha,R_e^-) + \SumSub(\alpha,R_e^+) = 0 + (-1)^k \sigma(G) = (-1)^k \sigma(G).$$ ### $G$ is strongly connected and contains no crucial edges. Let $e = [ab]\in G$ be an edge and not a loop: $b \ne a$. Recall that $G_e^{\vee}$ will denote a graph obtained from $G$ by reversal of the edge $e$. Since $e$ is not a crucial edge, $G \setminus \{e\} = G_e^{\vee} \setminus \{e\}$ is strongly connected. So $G_e^{\vee}$ is strongly connected, too, implying $\sigma(G_e^{\vee}) = \sigma(G)$. \[Lm:Reversal\] If the graph $G$ is strongly connected and $e = [ab] \in G$ is not a crucial edge then $\SumSub(\alpha,G) = \sigma(G)$ if and only if $\SumSub(\alpha,G_e^{\vee}) = \sigma(G_e^{\vee}) = \sigma(G)$. Acyclic subgraphs $H \subset G$ are split into five classes: 1. \[It:NoEAtoB\] $e \notin H$, but $a \succeq b$ in $H$ (that is, $H$ contains a directed path joining $a$ with $b$). 2. \[It:NoEBtoA\] $e \notin H$, but $b \succeq a$ in $H$. 3. \[It:NoENoPath\] $e \notin H$, and both $a \not\succeq b$ and $b \not\succeq a$ in $H$. 4. \[It:EAtoB\] $e \in H$, and $a \succeq b$ in $H \setminus \{e\}$. 5. \[It:ENoPath\] $e \in H$, and $a \not\succeq b$ in $H \setminus \{e\}$. Obviously, $H \in {\mathrm{\ref{It:NoEAtoB}}}$ if and only if $H \cup \{e\} \in {\mathrm{\ref{It:EAtoB}}}$. The number of edges of $H \cup \{e\}$ is the number of edges of $H$ plus $1$, so $$\label{Eq:1Plus4} \SumSub(\alpha, {\mathrm{\ref{It:NoEAtoB}}} \cup {\mathrm{\ref{It:EAtoB}}}) = \sum_{H \in \mathrm{\ref{It:NoEAtoB}}} (-1)^{\#\text{ of edges of $H$}} \,(1-1) = 0.$$ Also, $H \in \mathrm{\ref{It:NoENoPath}}$ if and only if $H \cup \{e\} \in \mathrm{\ref{It:ENoPath}}$, and similar to one has $\SumSub(\alpha,{\mathrm{\ref{It:NoENoPath}}} \cup {\mathrm{\ref{It:ENoPath}}}) = 0$, and therefore $$\label{Eq:AllEq2} \SumSub(\alpha,G) = \SumSub(\alpha,{\mathrm{\ref{It:NoEAtoB}}} \cup {\mathrm{\ref{It:NoEBtoA}}} \cup {\mathrm{\ref{It:NoENoPath}}} \cup {\mathrm{\ref{It:EAtoB}}} \cup {\mathrm{\ref{It:ENoPath}}}) = \SumSub(\alpha,{\mathrm{\ref{It:NoEBtoA}}}).$$ Like in Clause \[SSec:SSCCrucial\] if $H \in {\mathrm{\ref{It:ENoPath}}}$ then $H/e \subset G/e$ is acyclic, and vice versa, if $e \in H$ and $H/e \subset G/e$ is acyclic then $H \in {\mathrm{\ref{It:ENoPath}}}$. The graph $G/e$ is strongly connected, so by the induction hypothesis $\SumSub(\alpha,{\mathrm{\ref{It:ENoPath}}}) = -\SumSub(\alpha,G/e) = -(-1)^{k-1} \sigma(G/e) = (-1)^k \sigma(G)$, hence $\SumSub(\alpha,{\mathrm{\ref{It:NoENoPath}}}) = -(-1)^k\sigma(G)$. If $e \notin H$ and $H$ is acyclic, then $H$ is an acyclic subgraph of the strongly connected graph $G \setminus \{e\}$. The graph $G$ is strongly connected, too, so $e$ enters a cycle, and $\beta_1(G \setminus \{e\}) = \beta_1(G)-1$, which implies $\sigma(G \setminus \{e\}) = -\sigma(G)$. The graph $G \setminus \{e\}$ contains $k-1 < k$ edges, so by the induction hypothesis $$\SumSub(\alpha, {\mathrm{\ref{It:NoEAtoB}}} \cup {\mathrm{\ref{It:NoEBtoA}}} \cup {\mathrm{\ref{It:NoENoPath}}}) = \SumSub(\alpha,G \setminus \{e\}) = (-1)^{k-1} \sigma(G \setminus \{e\}) = (-1)^k \sigma(G),$$ and therefore $$\label{Eq:Sum12} \SumSub(\alpha, {\mathrm{\ref{It:NoEAtoB}}} \cup {\mathrm{\ref{It:NoEBtoA}}}) = 2(-1)^k \sigma(G).$$ A subgraph $H \subset G$ of class \[It:NoEAtoB\] is at the same time a subgraph $H \subset G_e^{\vee}$ of class \[It:NoEBtoA\]. So, applied to $G_e^{\vee}$ gives $\SumSub(\alpha, {\mathrm{\ref{It:NoEAtoB}}}) = \SumSub(\alpha, G_e^{\vee})$. If follows now from and that $$\SumSub(\alpha,G) + \SumSub(\alpha,G_e^{\vee}) = 2(-1)^k \sigma(G) = (-1)^k (\sigma(G) + \sigma(G_e^{\vee})),$$ which proves the lemma. To complete the proof of Theorem \[Th:Direct\] let $a$ be a vertex of $G$, and let $e_1 {, \dots ,} e_m$ be the complete list of edges finishing at $a$. Consider the sequence of graphs $G_0 = G$, $G_1 = G_{e_1}^{\vee}$, $G_2 = (G_1)_{e_2}^{\vee}$, …, $G_m = (G_{m-1})_{e_m}^{\vee}$. The graphs $G_0$ and $G_1$ are strongly connected; the graph $G_m$ is not, because $a \not\succeq b$ for any $b \ne a$ in it. Take the maximal $\ell$ such that $G_\ell$ is strongly connected. Since $\ell < m$, the graph $G_{\ell+1}$ exists and is not strongly connected, and therefore $G_\ell \setminus \{e_{\ell+1}\} = G_{\ell+1} \setminus \{e_{\ell+1}\}$ is not strongly connected either. So, the edge $e_{\ell+1}$ is crucial for the graph $G_\ell$, and by Clause \[SSec:SSCCrucial\] one has $\SumSub(\alpha,G_\ell) = (-1)^k \SumSub(G_\ell) = (-1)^k \sigma(G)$. The graphs $G_0=G {, \dots ,} G_\ell$ are strongly connected, so for any $i = 0 {, \dots ,} \ell-1$ the edge $e_{i+1}$ is not crucial for the graph $G_i$. Lemma \[Lm:Reversal\] implies now $$\begin{aligned} \SumSub(\alpha,G_{\ell-1}) = (-1)^k \sigma(G_{\ell-1}) &\Longrightarrow \SumSub(\alpha,G_{\ell-2}) = (-1)^k \sigma(G_{\ell-2})\\ &{\Longrightarrow \dots \Longrightarrow} \SumSub(\alpha,G) = (-1)^k \sigma(G). \end{aligned}$$ Theorem \[Th:Direct\] is proved. Theorem \[Th:Diag\] follows from Theorem \[Th:Direct\]’ {#SSec:DiagFromDirect} ------------------------------------------------------- Note first that the operation $B_i$, and hence $\Lapl$, preserves the sinks of the graph: if $\Lapl H = \sum_G x_G G$ and $x_G \ne 0$ then $G$ has the same sinks as $H$. Therefore if $I = \{i_1 {, \dots ,} i_s\}$ then $\Lapl(\det_{n,k}^I) = \sum_G x_G G$ where all the graphs $G$ in the right-hand side have the sinks $i_1 {, \dots ,} i_s$ and have no loops. Let $G$ be a graph with sinks $i_1 {, \dots ,} i_s$ and without loops, and let $\Phi \in \SSC_{n,k}^I$ (a strongly semiconnected graph with the isolated vertices $i_1 {, \dots ,} i_s$). Denote by $\widehat{\Phi}$ the graph $\Phi$ with the loops deleted. A contribution of $\Phi$ into $x_G$ is equal to $\frac{1}{k!} (-1)^{\beta_0(\Phi) - \#\text{ of loops in $\Phi$} + n}$ if $\widehat{\Phi} \subset G$ and is $0$ otherwise. The number of edges of $\widehat{\Phi}$ is $k - \#\text{ of loops of $\Phi$}$. The graph $\widehat{\Phi}$ is strongly semiconnected if and only if $\Phi$ is. The Euler characteristics of $\widehat{\Phi}$ is $$\beta_0(\widehat{\Phi}) - \beta_1(\widehat{\Phi}) = n - \#\text{ of edges of $\widehat{\Phi}$} = n-k + \#\text{ of loops of $\Phi$}$$ and $\beta_0(\widehat{\Phi}) = \beta_0(\Phi)$. Therefore, the contribution of $\Phi$ into $x_G$ is $$(-1)^{n+k+\beta_0(\widehat{\Phi})+\#\text{ of edges of $\widehat{\Phi}$}}\frac{1}{k!} = (-1)^{k+\beta_1(\widehat{\Phi})}\frac{1}{k!}$$ if $\widehat{\Phi} \subset G$ is strongly semiconnected and $0$ otherwise. Summing up, $$x_G = \frac{(-1)^k}{k!}\SumSub(\sigma;G) = \frac{1}{k!} \alpha(G)$$ by Theorem \[Th:Direct\]’. This proves Theorem \[Th:Diag\]. Proof of Theorem \[Th:Codim1\] {#SSec:PrCodim1} ------------------------------ Note that $\det_{n,k}^{i/i} = \det_{n,k}^{\varnothing}+ \det_{n,k}^{\{i\}}$. Applying the operator $\Lapl$ to equation and using Theorem \[Th:Diag\] with $I = {\varnothing}$ and $I = \{i\}$ one obtains $$\begin{aligned} 0 &= \sum_{i,j=1}^n \Lapl(([ij]) * \det_{n,k}^{i/j}) = \sum_{i=1}^n \Lapl ([ii]) * \Lapl(\det_{n,k}^{\varnothing}+ \det_{n,k}^{\{i\}}) + \sum_{\substack{i,j=1\\i \ne j}}^n ([ij]) * \Lapl(\det_{n,k}^{i/j}) \\ &= \sum_{\substack{i,j=1\\i \ne j}}^n ([ij]) * (\Lapl(\det_{n,k}^{i/j}) - \Lapl(\det_{n,k}^{\{i\}})) = \sum_{\substack{i,j=1\\i \ne j}}^n ([ij]) * (\Lapl(\det_{n,k}^{i/j}) - \frac{(-1)^k}{k!} U(\AC_{n,k}^{\{i\}})). \end{aligned}$$ The $(i,j)$-th term of the identity above consists of graphs where the edge $[ij]$ carries the number $1$. Therefore different terms of the identity cannot cancel, so every single term is equal to $0$. [9]{} J.Awan and O.Bernardi, Tutte polynomials for directed graphs, ArXiv:1610.01839v2. B.Epstein, A combinatorial invariant of $3$-manifolds via cycle-rooted trees, MSc. thesis (under supervision of prof.M.Polyak), Technion, Haifa, Israel, 2015. G.Kirchhoff, Über die Auflösung der Gleichungen, auf welche man bei der Untersuchung det linearen Verteilung galvanischer Ströme gefurht wird, [*Ann. Phys. Chem.*]{}, [**72**]{} (1847), S. 497–508. M.Polyak, From $3$-manifolds to planar graphs and cycle-rooted trees, talk at [*Arnold’s legacy*]{} conference, Fields Institute, Toronto, 2014. G.-C.Rota, On the foundations of combinatorial theory I: Theory of Mobius functions, [*Z. Wahrsch. Verw. Gebiete*]{}, [**2**]{} (1964) pp. 340–368. A.Sokal, The multivariate Tutte polynomial (alias Potts model) for graphs and matroids, in: [*Surveys in Combinatorics 2005*]{}, Cambridge University Press, Jul 21, 2005 — Mathematics — 258 pages. W.T.Tutte, The dissection of equilateral triangles into equilateral triangles, [*Math. Proc. Cambridge Phil. Soc.*]{}, [**44**]{} (1948) no. 4, pp. 463–482. K.Walker, [*An extension of Casson’s invariant*]{}, Annals of Mathematics Studies, [**126**]{}, Princeton University Press, 1992. D.J.A.Welsh and C.Merino, The Potts model and the Tutte polynomial, [*J. of Math. Physics*]{}, [**41**]{} (2000), no. 3, pp. 1127–1152.
{ "pile_set_name": "ArXiv" }
--- abstract: 'An ensemble of particles in thermal equilibrium at temperature $T$, modeled by Nosè-Hoover dynamics, moves on a triangular lattice of oriented semi-disk elastic scatterers. Despite the scatterer asymmetry a directed transport is clearly ruled out by the second law of thermodynamics. Introduction of a polarized zero mean monochromatic field creates a directed stationary flow with nontrivial dependence on temperature and field parameters. We give a theoretical estimate of directed current induced by a microwave field in an antidot superlattice in semiconductor heterostructures.' author: - 'G. Cristadoro$^{(a),(b)}$ and D.L. Shepelyansky$^{(b)}$' date: 'October 20, 2004' title: Nonequilibrium stationary states with ratchet effect --- According to the second law of thermodynamics there is no stationary directed transport in spatially periodic asymmetric systems in thermal equilibrium [@smol; @feynman]. However, a time periodic parameter variation may drive such a system out of equilibrium leading to the emergence of stationary transport whose direction depends non trivially on parameters. Such directed transport appears in systems with noise, fluctuations and dissipation and is now called Brownian motor or ratchet (see e.g. reviews [@belinicher; @hanggi; @reimann]). The ratchet effect has a generic nature and it has been observed in various physical systems including semiconductor heterostructures [@linke], cold atoms in a laser field [@grynberg], vortices in superconductors [@mooij; @nori; @ustinov] and macroporous silicon membranes under pressure oscillations [@muller]. It has also important applications in biological systems as discussed in [@hanggi; @prost]. In spite of a great recent interest to ratchets the theoretical research is mainly concentrated on one-dimensional models (see e.g. [@reimann]). Also, since the ratchet behavior is usually rather complex, an overdamped limit is used very often to obtain analytical parameter dependence even if in this regime a directed transport is absent for [*ac*]{} zero mean force [@reimann]. To understand in a better way the global properties of ratchets and their dependence on such important physical parameters as temperature $T$ and driving strength ${\bf f}$, we analyze here a generic case when [*ac*]{}-driving affects a Maxwell thermostat ensemble of noninteracting particles moving in an asymmetric two-dimensional (2D) periodic structure. This structure is composed of triangular 2D-lattice of rigid semi-disks of radius $r_d$ as shown in Fig. 1 (insert). The distance $R$ between disk centers is fixed to be $R=2 r_d$ and we assume that collisions with semi-disks are elastic. Free particle motion between semi-disks is affected by a polarized monochromatic force ${\bf f} = f (\cos \theta, \sin \theta) \cos \omega t$ with frequency $\omega$, strength $f$ and polarization angle $\theta$ to $x$-axis. It is also assumed that particles are in thermal equilibrium and at $f=0$ their velocities are given by the Maxwell distribution at temperature $T$. Fig. 1 shows that in this system [*ac*]{}-force generates stationary directed transport. In numerical computations we put $r_d$, particle mass $m$, unit of time and Boltzmann’s constant $k$ to be equal to unity. -0.3cm =8.5cm -0.3cm In order to put particles in thermal equilibrium we choose the elegant method of the Nosè-Hoover thermostat (see e.g. [@hover; @hover1; @klages] and Refs. therein). In this method the motion of a particle is affected by an effective friction $\gamma$ which keeps the average kinetic energy $\langle \mathbf{p}^2/2 \rangle$ equal to a given thermostat temperature $T$. In this way the dynamics of particle is described by the equations: $$\label{enh} \mathbf{\dot q} = \mathbf{p}/m \; , \;\; \mathbf{\dot p} = \mathbf{F} - \gamma \mathbf{p} \; , \;\; \dot{\gamma} = [\mathbf{p}^2/(2mT) - 1]/\tau^2$$ where $\mathbf{q}, \mathbf{p}$ are particle coordinate and momentum, $\mathbf{F}$ is a sum of [*ac*]{}-force and force of elastic collisions with semi-disks, and $\tau$ is the time scale of relaxation to equilibrium. =8.5cm -0.2cm It is known that the Nosè-Hoover thermostat works well only if the dynamics is sufficiently chaotic [@hover; @hover1; @klages]. In some cases, e.g. for the Galton board, the Nosè-Hoover thermostat gives noticeable deviations from the Maxwell distribution [@klages]. To check that in our case this method really gives a thermal equilibrium we analyze the steady state distribution in the momentum space obtained by numerical Runge-Kutta integration of Eqs.(\[enh\]). Our results show that a small [*ac*]{}-force is needed to make chaotic dynamics between semi-disks more homogeneous and to produce a stable Maxwell thermal equilibrium which is not sensitive to variation of relaxation rate $1/\tau$ (Fig.3, insert). At large force the numerical data show that 2D steady state in the momentum space is still close to the Maxwell distribution (Fig. 2) even if the [*ac*]{}-driving produces a clear ratchet effect shown in Fig. 1. The dependence of steady state on temperature closely follows the Maxwell distribution in momentum space $p=| \mathbf{p} |$ as shown in Fig. 3. Thus we may conclude that the dynamics given by the Nosè-Hoover equations allows efficiently investigate the effects of [*ac*]{}-driving on particles in thermal equilibrium. The numerical data show that this driving generates a strong ratchet effect (Fig. 1) with directed transport wich depends on temperature and parameters of driving force. =8.5cm -0.2cm To understand the properties of this directed transport we first analyze the dependence of averaged friction $\langle \gamma \rangle$ on driving strength $f$ and temperature $T$. The value of $\langle \gamma \rangle$ is obtained by averaging over long time interval during numerical integration of Eqs.(\[enh\]) for one trajectory. We also checked that averaging over a few trajectories gives the same result. The data are shown in Fig. 4. They are well described by a global scaling given by $$\label{gamma} \langle \gamma \rangle = C r_d m^{-1/2} f^2/T^{3/2} \; .$$ Small deviations seen at low $T $ appear because of strong driving force which starts to modify significantly the particle velocity distribution in this regime. The numerical constant $C$ is only weakly dependent on $\theta$ and $\omega$ changing by 50% to 30% when $\theta$ changes from $0$ to $\pi/2$ and $\omega$ changes by a factor 10, respectively. The dependence (\[gamma\]) clearly tells that in presence of driving force the thermostat creates an effective friction force $\mathbf{f}_f = - \langle \gamma \rangle \mathbf{p}$ acting on particle propagation with an effective friction constant $\langle \gamma \rangle$. Surprisingly, this friction coefficient varies with $f$ and $T$ according to Eq.(\[gamma\]) but in a large range remains independent of the relaxation time $\tau$ (Fig. 4 insert). We note that the particle dynamics in absence of thermostat but in presence of friction force $\mathbf{f}_f = - \gamma \mathbf{p}$ with constant friction coefficient $\gamma$ has been analyzed in [@chep] where it was shown that [*ac*]{}-force generates a directed transport on semi-disk lattice. To understand the origin of the dependence (\[gamma\]) we put forward the following heuristic arguments. The driving force gives a diffusive energy growth during a dissipative time scale $1/\gamma$ so that $$\label{diff} (\Delta E)^2 \sim D_E /\gamma \; , \;\; D_E \sim f^2 v l \; ,$$ where the diffusion rate in energy is $D_E \sim {\dot{E}}^2 \tau_c \sim f^2 v l$ and the mean-free path $l \sim R \sim r_d \sim 1$ determines the collision time $\tau_c = l/v$. In the Maxwell equilibrium the particle velocity is $v \sim (T/m)^{1/2}$ and the fact that the driving force does not modify the velocity distribution implies that $\Delta E \sim T$ so that the diffusive growth is stopped by effective friction $\gamma \sim D_E/T^2 \sim r_d m^{-1/2}f^2/T^{3/2}$ in agreement with (\[gamma\]). In fact there is a close relation to results [@chep] where the thermostat is absent but a friction force $\mathbf{f}_f = - \gamma \mathbf{p}$ with constant $\gamma$ affects particle dynamics. In that case [*ac*]{}-driving force heats a particle up to energy $E \sim (r_d f^2/m^{1/2}\gamma)^{2/3}$ while in presence of thermostat the energy is fixed by temperature $T \sim E$ that imposes a convergence to the stationary state with effective friction given by (\[gamma\]) [@note]. =3.2in -0.3cm The dependence of average velocity $v_f$ of the ratchet flow on $\langle \gamma \rangle$ for various values of driving strength $f$ is shown in Fig. 5. Globally, the flow velocity $v_f$ grows with increase of $\gamma$. Two regimes are clearly seen: $v_f \approx r_d\langle \gamma \rangle$ for $\langle \gamma \rangle < \gamma_c$ and $v_f \approx (r_d^2 f \langle \gamma \rangle /m)^{1/3}/10$ for $\langle \gamma \rangle > \gamma_c \approx f^{1/2}/[30(r_d m)^{1/2}]$. In fact this dependence is very close to the one found in a model with fixed $\gamma$ [@chep]. As a result, from (\[gamma\]) we obtain the dependence of flow velocity on temperature: $$\begin{aligned} \label{flow} v_f /v & \approx & r_d f /50T \;\;\;,\; (T<T_c); \\ \label{flow1} v_f /v & \approx & (r_d f/8T)^2 \;,\; (T>T_c);\end{aligned}$$ where $v=(2T/m)^{1/2}$ is the thermal velocity and $T_c \approx r_d f$ is linked to $\gamma_c$ obtained from Fig. 5. The transition between two regimes takes place when the energy given by [*ac*]{}-force to particle between two collisions becomes larger than thermal energy ($T<T_c$). In that case the effect of driving is strong and $v_f \sim f \tau_c/m \sim r_d f/(mT)^{1/2}$ leading to (\[flow\]). For $T>T_c$ thermal fluctuations are strong and the ratchet effect appears only in the second order of force $f$ giving (\[flow1\]). The numerical factors in (\[flow\]),(\[flow1\]) are taken for the case $\theta=0$ from Figs. 4,5. We note that Eqs. (\[gamma\])-(\[flow1\]) are derived in the regime of relatively weak friction $\langle \gamma \rangle \ll \omega$ and relaxation rate $1/\tau \ll \omega$. Another important point is that the dependence (\[flow\]),(\[flow1\]) is robust with respect to variation of scatter geometry, e.g. introduction of additional disk scatterer in the center of unit cell eliminates all collisionless paths but gives no significant modifications (see Fig. 5). =3.2in =3.2in -0.3cm The dependence of flow directionality, determined through angle $ \varphi$ ($\mathbf{v_f}=v_f (\cos \varphi, \sin \varphi)$), on the polarization of [*ac*]{}-force is shown in Fig. 5 (insert). In average, it is satisfactory described by the relation $\varphi = \pi - 2\theta$ (similar dependence was seen in [@chep]). On a qualitative ground, we may say that at $\theta=0$ due to friction a particle becomes trapped between semi-disks of a unit cell that gives a directed transport to the left while for $\theta=\pi/2$ vertical oscillations push particle to the right in presence of friction. The linear dependence $\varphi = \pi - 2\theta$ interpolates between these two limits. However, a more quantitative derivation is needed. It is interesting to apply the approach developed above to other types of thermostat. It is possible to realize the semi-disk Galton board with antidot superlattices for 2D electron gas in semiconductor heterostructures. With such structures the Galton board of disks has already been implemented (see e.g. [@weiss]) and effects of microwave radiation has been studied [@kvon1]. For disk antidots like those in [@weiss; @kvon1] the ratchet effect is absent due to symmetry of antidot. However, for semi-disk antidot lattice strong ratchet effect should appear. To find its properties we should take into account that in this case we have the Fermi-Dirac thermostat with the Fermi energy $E_F \gg T$. Due to that in (\[diff\]) the particle velocity $v$ is equal to the Fermi velocity $v_F=(2E_F/m)^{1/2}$ independent of $T$. This modification gives the average friction $\gamma_F$ for the Fermi gas $$\label{fgamma} \gamma_F = C f^2 v_F r_d/T^2 \approx v_f/r_d\; ,$$ where we kept the same numerical constant $C \sim 1/50$. In fact (\[fgamma\]) follows from $D_E \sim f^2 v_F r_d$ (see (\[diff\])) and $\gamma_F \sim D_E/T^2$. The second equality in (\[fgamma\]) appears due to the fact that $E_F \gg T_c$ implying the regime (\[flow1\]) with $v_f \sim \gamma_F r_d$. Of course, only a small fraction $T/E_F$ of electrons near $E_F$ contributes to this ratchet flow. Hence, the current $I$ per one semi-disk row is $$\label{current} I \sim e r_d n_e v_f T/E_F \sim C e r_d^3 \sqrt{n_e} f^2 / ( T \hbar) \; .$$ where we used that for the 2D electron Fermi gas $E_F = \pi n_e \hbar^2/m$. We note that in semiconductor antidot lattices like in [@weiss; @kvon1] the effective mass $m$ is about 15 times smaller compared to the electron mass. For a typical parameters of semi-disk antidot lattice with electron density $n_e \sim 10^{12} cm^{-2}$, $r_d \sim 1\mu m$, field strength per electron charge $f/e \sim 1 V/cm $ and $T \sim 10 K$ we obtain $v_f/v_F \sim 10^{-4}$. At these parameters $E_F \sim 150 K$, $v_F \sim 3 \cdot 10^7 cm/sec$ and the current $I \sim 10^{-9} A$ is sufficiently large to be observed experimentally. The result (\[current\]) is based on the semiclassical estimate for the diffusion rate $D_E$ which assumes that the energy of microwave photon is larger than the level spacing $\Delta$ inside one unit cell: $\hbar \omega > \Delta \approx 2\pi \hbar^2/(m r_d^2)$. In the opposite limit $\hbar \omega \ll \Delta$, [*ac*]{}-driving is in the quantum adiabatic regime when the excitation in energy is very weak. Thus for $r_d \sim 1 \mu m$ we have $\Delta \approx 5 \cdot 10^{-6} eV \approx 0.05 K$ and the directed transport appears only for $\omega/2\pi > 1 GHz$. In experiments [@linke] the frequency was deeply in the adiabatic regime with $\omega/2\pi \sim 100Hz$ and the directed transport was absent at zero mean force. We also note that in the quantum case the ratchet transport should disappear as soon as the amplitude of oscillations $f/m \omega^2$ induced by [*ac*]{}-force becomes smaller than the wavelenght $\hbar/m v_F$ at the Fermi level. Thus the ratchet survives only for $\omega<\sqrt{f v_F/\hbar}$. For field strenght of $1 V/cm$ this gives an approximate bound at $30 GHz$. In summary, we showed that zero mean [*ac*]{}-force applied to particles being in thermal equilibrium in asymmetric periodic potential creates a directed transport flow. Its direction is efficiently changed by polarization of the force. We also established the dependence of the flow velocity $v_f$ on temperature and driving field strength for the Maxwell \[Eqs. (\[flow\]),(\[flow1\]),(\[fgamma\])\] and the Fermi-Dirac \[Eqs. (\[fgamma\]),(\[current\])\] thermostats. We thank Alexei Chepelianskii, Kvon Ze Don and Sergey Vitkalov for useful discussions. [99]{} M. v. Smoluchowski, Physik. Zeitschr. [**13**]{}, 1069 (1912). R.P.Feynman, R.B.Leighton, and M.Sands, [*The Feynman Lectures on Physics*]{}, [**1**]{}, chapter 46, Addison Wesley, Reading MA (1963). V.I.Belinicher, and B.I.Sturman, Sov. Phys. Usp. [**23**]{}, 199 (1980). R.D. Astumian and P. Hänggi, Physics Today [**55**]{} (11), 33 (2002). P. Reimann, Phys. Rep. [**361**]{}, 57 (2002). H. Linke, T.E. Humphrey, A. Löfgren, A.O. Sushkov, R. Newbury, R.P. Taylor, and P. Omling, Science [**286**]{}, 2314 (1999). C. Mennerat-Robilliard, D. Lucas, S. Guibal, J. Tabosa, C. Jurczak, J.-Y. Courtois, and G. Grynberg, Phys. Rev. Lett. [**82**]{}, 851 (1999). J.B. Majer, J. Peguiron, M. Grifoni, M. Tusveld, and J.E. Mooij, Phys. Rev. Lett. [**90**]{}, 056802 (2003). J.E. Villegas, S. Savel’ev, F. Nori, E.M. Gonzalez, J.V. Anguita, R. Garcia, and J.L. Vicent, Science [**302**]{}, 1188 (2003). A.V. Ustinov, C. Coqui, A. Kemp, Y. Zolotaryuk and M. Salerno, Phys. Rev. Lett. [**93**]{}, 087001 (2004). S. Matthias and F. Müller, Nature [**424**]{}, 53 (2003). F. Jülicher, A. Ajdari, and J. Prost, Rev. Mod. Phys. [**69**]{}, 1269 (1997). W. G. Hoover, Phys. Rev. A [**31**]{}, 1695 (1985). W. G. Hoover, [*Time reversibility, computer simulation, and chaos*]{}, World Scientific, Singapore (1999). K. Rateitschak, R. Klages, W. G. Hoover, J. Stat. Phys. [**101**]{}, 61 (2000). A. D. Chepelianskii and D. L. Shepelyansky, cond-mat/0402675. We note that a numerical constant $C \approx 0.02$ in (\[gamma\]) (see Fig. 4) corresponds to average dependence $v^2=2T=(f^2/\gamma)^{2/3}/7$ from Fig. 5 in [@chep]. D. Weiss, M. L. Roukes, A. Menschig, P. Grambow, K. von Klitzing, and G. Weimann, Phys. Rev. Lett. [**66**]{}, 2790 (1991). A.A. Bykov, G.M. Gusev, Z.D.Kvon, V.M.Kudryashev, and V.G.Plyukhin, Pis’ma Zh. Eksp. Teor. Fiz. [**53**]{}, 407 (1991) \[JETP Lett. [**53**]{}, 427 (1991)\].
{ "pile_set_name": "ArXiv" }
--- abstract: 'The ExoMol database (www.exomol.com) provides extensive line lists of molecular transitions which are valid over extended temperatures ranges. The status of the current release of the database is reviewed and a new data structure is specified. This structure augments the provision of energy levels (and hence transition frequencies) and Einstein $A$ coefficients with other key properties, including lifetimes of individual states, temperature-dependent cooling functions, Landé $g$-factors, partition functions, cross sections, $k$-coefficients and transition dipoles with phase relations. Particular attention is paid to the treatment of pressure broadening parameters. The new data structure includes a definition file which provides the necessary information for utilities accessing ExoMol through its application programming interface (API). Prospects for the inclusion of new species into the database are discussed.' address: 'Department of Physics and Astronomy, University College London, London, WC1E 6BT, UK' author: - 'Jonathan Tennyson[^1]' - 'Sergei N. Yurchenko' - 'Ahmed F. Al-Refaie' - 'Emma J. Barton' - 'Katy L. Chubb' - 'Phillip A. Coles' - 'S. Diamantopoulou' - 'Maire N. Gorman' - Christian Hill - 'Aden Z. Lam' - Lorenzo Lodi - 'Laura K. McKemmish' - Yueqi Na - Alec Owens - 'Oleg L. Polyansky' - 'Clara Sousa-Silva' - 'Daniel S. Underwood' - Andrey Yachmenev - Emil Zak title: 'The ExoMol database: molecular line lists for exoplanet and other hot atmospheres' --- infrared ,visible ,Einstein $A$ coefficients ,transition frequencies ,partition functions ,cooling functions ,lifetimes ,cross sections ,$k$ coefficients ,Landé $g$-factors Introduction ============ Hot molecules exist in many environments in space including cool stars [@09Bexxxx.exo], failed stars generally known as brown dwarfs [@aha97; @09Bexxxx.exo] and exoplanets [@13TiEnCo.exo]. The atmospheric properties of these objects are known to be strongly influenced by the spectra of the molecules they contain. On Earth, spectra of hot molecules are observed in flames [@14BoWeHy; @15CoLixx], discharge plasmas [@02ChWaLi.SO; @15BiCrGl], explosions [@11CaLiPi.H2O] and in the hot gases emitted, for example, from smoke stacks [@12EvFaCl.CO2]. In addition, high-lying states can be important in non local thermodynamic equilibrium (LTE) environments both in space, for example, emissions observed from comets [@jt330; @jt349], and on Earth. The spectra of key atmospheric molecules at room temperature have been the subject of systematically maintained databases such as HITRAN [@jt350; @jt453; @jt546] and GEISA [@jt504; @jtGEISA]. As will be amply illustrated below, the spectra of hot molecules contain many, many more transitions and so far attempts to compile systematic databases have been limited. Databases of hot molecular spectra do exist for other specialized applications, such as the EM2C database for combustion applications [@12RiSoxx] or one due to Parigger [[*et al*]{}]{} for studies of laser-induced plasmas [@15PaWoSu]. Planets and cool stars share some common fundamental characteristics: they are faint, their radiation peaks in the infrared and their atmosphere is dominated by strong molecular absorbers. Modelling planetary and stellar atmospheres is therefore difficult as their spectra are extremely rich in structure with hundreds of thousands to many billions of spectral lines which may be broadened by high-pressure and temperature effects. The ExoMol project [@jt528] aims to provide the molecular line lists that astronomers need in order to understand the physics and chemistry of astronomical bodies cool enough to form molecules in their atmospheres. In particular for extrasolar planets, brown dwarfs and cool stars [@jt143; @aha97; @13TiEnCo.exo], as well as circumstellar structures such as planetary envelopes and molspheres [@08Tsuji]. In practice these data are also useful for a wide range of other scientific disciplines; examples include studies of the Earth’s atmosphere [@14ScGaHe; @15LaPoTs], hypersonic non-equilibrium [@15LoSuBo], analysis of laboratory spectra [@14FoGoHe; @jt616; @16NiReTaKa.CH4], measurements of hot reactive gases [@13GrFaNi] and the proposed remote analysis of molecular composition using laser oblation [@15TaCa]. The original ExoMol data structure was very focused in its goal, with the scope of the data limited to generating lists of transitions [@jt548]. However, it has become obvious that the potential applications of the ExoMol spectroscopic data are much more diverse. For example, the ExoMol data can be used to compute partition functions [@jt571], cross sections [@jt542], lifetimes [@jt624], Landé $g$-factors [@jtgfac] and other properties. Our aim is to systematically provide this additional data to maximise its usefulness. To do this requires significant extension of the ExoMol data structure, which is the major purpose of this paper. At the same time this implementation should facilitate the adoption of an application programming interface (API) between the database and programs using it. Similar enhancements are actively being pursued by other related databases such as HITRAN [@jt559; @HAPI; @Honline]. A major new feature is the inclusion, albeit at a fairly crude level, of pressure-broadening parameters. These have been shown to be important for models of exoplanets [@jt521; @16HeMaxx] and are known to be vital for many other applications. Section 2 summarizes current data coverage of the ExoMol database. Section 3 reviews the individual line lists included in the database with particular emphasis on any changes made to them since they were originally published. Section 4 describes the expansion of data that is now provided by the ExoMol database. Section 5 gives a formal description of the new data structures implemented to cover this provision and to provide the API functionality. Section 6 briefly describes the utility programs available as part of the ExoMol database and Section 7 the ExoMol website. Finally we discuss future prospects for the database. Database coverage ================= The ExoMol project aims at complete coverage of the spectroscopic properties of molecules which are deemed to be important in hot astrophysical environments. Coverage concerns (a) the molecular species considered, including isotopologues; (b) the frequency range considered and (c) the upper temperature range for which the data is reasonably complete. Both the required temperature and frequency range completeness are to some extent a judgement on what is required for astronomical and other studies. For example, a molecule like nitric acid (HNO$_3$) will dissociate at a relatively low temperature so coverage above about 500 K is unlikely to be important. On the other hand, several diatomics such SiO and CO are known to feature in stellar spectra, and coverage to temperatures over 6000 K is necessary. The general ExoMol approach is molecule-by-molecule. That is, a comprehensive line list is created for a particular molecule which is then made available in the database. For more challenging larger systems such as NH$_3$, PH$_3$ and HNO$_3$, it has been our policy to produce initial, room-temperature line lists, see [@jt466; @jt556; @jt603] respectively. This allows us to improve the model, validate the available experimental data and, in some cases, make spectral assignments. In all instances, the subsequent hot line lists [@jt500; @jt592; @jt614] are both more complete and more accurate and should therefore be used even for studies at low temperatures. Thus far ExoMol has not considered ultraviolet (UV) absorption; however, there is increasing interest in the consequences of UV radiation on exoplanets [@14FoBiLa; @14LoPaFr; @13BeBaxx] so this may need to be reviewed in future. As discussed in the comments below, molecules are still being added to the database so that coverage by species is steadily increasing. Tables \[tab:exomoldata\] and \[tab:otherdata\] summarize the molecules for which the ExoMol database currently provides data. The division is between species that have explicitly been studied as part of the ExoMol project (Table \[tab:exomoldata\]) and those for which the data has been taken from other studies (Table \[tab:otherdata\]). The ExoMol project has a specific methodology based on the use of spectroscopically-determined potential energy surfaces and [*ab initio*]{} dipole surfaces, which are combined with explicit variational treatments of the nuclear motion problem. For open shell systems these treatments involve explicit inclusion of spin-orbit and related curve-coupling effects; the project has developed a nuclear motion program, [Duo]{}, for treating coupled diatomic curves [@jt609]. For closed shell molecules a range of codes [@lr07; @jt338; @jt339; @07YuThJe.method; @15YaYuxx.method; @jt588] are used. These are all essentially based on finding near-exact solutions of the ro-vibrational Schrödinger equation for a given potential energy surface but the level of approximation increases with the size of the molecule considered. Data from other sources arise from a variety of methodologies which range from completely [*ab initio*]{}, appropriate for systems with very few electrons, to largely empirical. Table \[tab:otherdata\] gives a pointer to the method used in each case; for full details the reader should consult the cited reference. Molecule $N_{\rm iso}$ $T_{\rm max}$ $N_{elec}$ $N_{\rm lines}$ $^a$ DSName Reference ---------- --------------- --------------- ------------ ---------------------- ---------- -------------------------------------- BeH 1 2000 1 16 400 Yadin Yadin [[*et al*]{}]{} [@jt529] MgH 3 2000 1 10 354 Yadin Yadin [[*et al*]{}]{} [@jt529] CaH 1 2000 1 15 278 Yadin Yadin [[*et al*]{}]{} [@jt529] SiO 5 9000 1 254 675 EJBT Barton [[*et al*]{}]{} [@jt563] HCN/HNC 2$^a$ 4000 1 399 000 000 Harris Barber [[*et al*]{}]{} [@jt570] CH$_4$ 1 1500 1 9 819 605 160 YT10to10 Yurchenko & Tennyson [@jt564] NaCl 2 3000 1 702 271 Barton Barton [[*et al*]{}]{} [@jt583] KCl 4 3000 1 1 326 765 Barton Barton [[*et al*]{}]{} [@jt583] PN 2 5000 1 142 512 YYLT Yorke [[*et al*]{}]{} [@jt590] PH$_3$ 1 1500 1 16 803 703 395 SAlTY Sousa-Silva [[*et al*]{}]{} [@jt593] H$_2$CO 1 1500 1 10 000 000 000 AYTY Al-Refaie [[*et al*]{}]{} [@jt593] AlO 4 8000 3 4 945 580 ATP Patrascu [[*et al*]{}]{} [@jt598] NaH 2 7000 2 79 898 Rivlin Rivlin [[*et al*]{}]{} [@jt605] HNO$_3$ 1 500 1 6 722 136 109 AlJS Pavlyuchko [[*et al*]{}]{} [@jt614] CS 8 3000 1 548 312 JnK Paulose [[*et al*]{}]{} [@jt615] CaO 1 5000 5 21 279 299 VBATHY Yurchenko [[*et al*]{}]{} [@jt618] SO$_2$ 1 2000 1 1 300 000 000 ExoAmes Underwood [[*et al*]{}]{} [@jt635] : Datasets created by the ExoMol project and included in the ExoMol database.[]{data-label="tab:exomoldata"} $N_{\rm iso}$ Number of isotopologues considered;\ $T_{\rm max}$ Maximum temperature for which the line list is complete;\ $N_{elec}$ Number of electronic states considered;\ $N_{\rm lines}$ Number of lines: value is for the main isotope.\ $^a$ A line list for H$^{13}$CN/HN$^{13}$C due to Harris [[*et al*]{}]{} [@jt447] is also available. ----------------------------------------------------------------------------------------------------------------------------------------------- Molecule $N_{\rm iso}$ $T_{\rm max}$ $N_{elec}$ $N_{\rm DSName Reference Methodology lines}$ ---------- --------------- --------------- ------------ --------------- ---------- ---------------------------------------------- ------------- H$_3^+$ 2$^a$ 4000 1 3 070 571 NMT Neal [[*et al*]{}]{} [@jt181] ExoMol H$_2$O 2$^b$ 3000 1 505 806 202 BT2 Barber [[*et al*]{}]{} [@jt378] ExoMol NH$_3$ 2$^c$ 1500 1 1 138 323 351 BYTe Yurchenko [[*et al*]{}]{} [@jt500] ExoMol HeH$^+$ 4 10000 1 1 431 Engel Engel [[*et al*]{}]{} [@jt347] Ab initio HD$^+$ 1 12000 1 10 119 CLT Coppola [[*et al*]{}]{} [@jt506] Ab initio LiH 1 12000 1 18 982 CLT Coppola [[*et al*]{}]{} [@jt506] Ab initio LiH$^+$ 1 12000 1 332 CLT Coppola [[*et al*]{}]{} [@jt506] Ab initio ScH 1 5000 6 1 152 827 LYT Lodi [[*et al*]{}]{} [@jt599] Ab initio MgH 1 3 30 896 13GhShBe GharibNezhad [[*et al*]{}]{} [@13GhShBe.MgH] Empirical CaH 1 2 6000 11LiHaRa Li [*et al*]{} [@11LiHaRa.cah] Empirical NH 1 1 10 414 14BrRaWe Brooke [[*et al*]{}]{} [@14BrRaWe.NH] Empirical CH 2 4 54 086 14MaPlVa Masseron [[*et al*]{}]{} [@14MaPlVa.CH] Empirical CN 1 1 195 120 14BrRaWe Brooke [[*et al*]{}]{} [@14BrRaWe.CN] Empirical CP 1 1 28 735 14RaBrWe Ram [[*et al*]{}]{} [@14RaBrWe.CP] Empirical HCl 1 1 2588 11LiGoBe Ram [[*et al*]{}]{} [@11LiGoBe.HCl] Empirical CrH 1 2 13 824 02BuRaBe Burrows [[*et al*]{}]{} [@02BuRaBe.CrH] Empirical FeH 1 2 93 040 10WEReSe Wende [[*et al*]{}]{} [@10WEReSe.FeH] Empirical TiH 1 3 181 080 05BuDuBa Burrows [[*et al*]{}]{} [@05BuDuBa.TiH] Empirical ----------------------------------------------------------------------------------------------------------------------------------------------- : Datasets not created as part of the ExoMol project but included in the ExoMol database.[]{data-label="tab:otherdata"} $N_{\rm iso}$ Number of isotopologues considered;\ $T_{\rm max}$ Maximum temperature for which the line list is complete;\ $N_{elec}$ Number of electronic states considered;\ $N_{\rm lines}$ Number of lines: value is for the main isotope.\ $^a$ There is a H$_2$D$^+$ line list available from Sochi and Tennyson [@jt478].\ $^b$ The VTT line list for HDO due to Voronin [[*et al*]{}]{} [@jt469] is also available.\ $^c$ There is a room temperature $^{15}$NH$_3$ line list due to Yurchenko [@15Yurche.NH3].\ The molecules listed in Tables \[tab:exomoldata\] and \[tab:otherdata\] do not at present provide a comprehensive set of species. A number of key species are available from other sources. HITEMP [@jt480] in principle provides a source of data for CO, OH, NO and CO$_2$. In practice there are new hot line lists for CO [@15LiGoRo.CO], OH [@16BrBeWe.OH], and CO$_2$ [@11TaPe.CO2; @14HuGaFr.CO2] which are more recent than the data given in the current release of HITEMP. High quality line lists for ozone are also available from elsewhere [@13BaMiSt.O3]. Line lists for some other missing species may also be found in the UGAMOP (https://www.physast.uga.edu/ugamop/) and Kurucz [@11Kurucz.db] databases. The coverage of these databases for a (possible) exoplanet characterization mission has recently been reviewed by two of us [@jt578]. Finally we note that while the majority of exoplanet spectroscopy is performed at rather low resolution, very precise spectroscopic data can be required [@13DeBrSn.exo; @13BiDeBr.exo; @14BrDeBi.exo], which is proving an issue for particular key species [@15HoDeSn.TiO]. Therefore, ExoMol aims to be both complete, and as accurate as possible. Individual line lists ===================== Below we consider some of the line lists presented in the ExoMol database and listed in Tables \[tab:exomoldata\] and \[tab:otherdata\]. We restrict our discussion to issues not covered in the original publication. One general issue is that Medvedev and co-workers [@highv; @15LiGoRo.CO] identified a numerical problem with the intensities of high overtone transitions computed with the standard compilation of the diatomic vibration-rotation program [Level]{} [@lr07]. Our line lists computed with [Level]{} have been adjusted to remove transitions which appeared to have been affected by this issue; such cases are noted below. The transitions removed are all very weak and it is anticipated that these changes will have very little effect on practical applications. We checked similar calculations performed with our in-house rovibronic program [Duo]{} [@jt609] and found similar behaviour: when the value of the transition dipole moment becomes comparable with the double precision error ($\approx 10^{-16}$ a.u.), the corresponding electric dipole intensities essentially represent numerical noise and have to be removed. MgH --- The ExoMol MgH line list only considers transitions within the X $^2\Sigma^+$ ground electronic state [@jt529]. A line list containing A $^2\Pi$ – X $^2\Sigma^+$ and B$^\prime$ $^2\Sigma^+$ – X $^2\Sigma^+$ transitions has been given by GharibNezhad [[*et al*]{}]{} [@13GhShBe.MgH]. Both line lists are included in the ExoMol database. Furthermore Szidarovszky and Császár [@15SzCsxx.MgH] showed that, due to the relatively low dissociation energy of MgH, consideration of quasi-bound states in the partition sum significantly alters the partition function at higher temperatures. However, for reasons of self-consistency we recommend using the partition function given by Yadin [[*et al*]{}]{}[@jt529]. CaH --- The ExoMol CaH line list only considers transitions within the X $^2\Sigma^+$ ground electronic state [@jt529]. A line list containing E $^2\Pi$ – X $^2\Sigma^+$ transitions has been given by Li [*et al*]{} [@11LiHaRa.cah]. Both line lists are included in the ExoMol database. SiO --- The original SiO ExoMol line lists of Barton [[*et al*]{}]{} [@jt563] computed using [Level]{} have been truncated by removing transitions with $\Delta v \geq 6$. HCN/HNC ------- The combined HCN/HNC ExoMol line list of Barber [[*et al*]{}]{} [@jt570] used a calculated energy separation (isomerization energy) of 5705 [cm$^{-1}$]{}between the ground states of HCN and HNC. Recent work by Nguyen [[*et al*]{}]{}[@15NgBaRu.HCN] suggests that this value is too high. We have done some additional [*ab initio*]{} calculation using [MOLPRO]{} [@12WeKnKn.methods] at multi-reference configuration interaction (MRCI) level of theory. The complete basis set extrapolation (CBS) value CBS\[56\]z was found to equal 5398.62 [cm$^{-1}$]{}. Subtracting from this value the zero point energy difference between HCN and HNC wells of 88 [cm$^{-1}$]{}, gives the value for the energy separation of 5310 [cm$^{-1}$]{}, closer to the value of 5236 $\pm$ 50 [cm$^{-1}$]{} recommended by Nguyen [[*et al*]{}]{} [@15NgBaRu.HCN]. This value has been adopted by revising the states file and calculated partition function. CH$_4$ ------ Although the YT10to10 line list [@jt564] represents a major step forward in the modelling of hot methane spectra [@jt572], it does not represent a complete solution to the problem. Rey [[*et al*]{}]{} [@14ReNiTy.CH4] have also produced a line list, using similar procedures to those adopted by ExoMol, valid to higher temperatures (2000 K) but covering a more limited spectral range. The temperature coverage can be checked using available, high-temperature partition functions [@08WeChBo.CH4; @15NuKrRe.CH4]. Yurchenko [[*et al*]{}]{} [@jtCH4] have extended the YT10to10 line list to higher temperatures (2000 K) and improved the predicted frequencies by replacing computed energy levels with empirical ones provided by Boudon [@14Boudon.CH4]. The resulting line list has 35 billion transitions and Yurchenko [[*et al*]{}]{} propose that most of the lines in the line list can be represented by background cross sections. These temperature-dependent, but pressure-independent cross sections, generated using the vast majority of the lines, can then be used to supplement a reduced line list of 203 million lines selected to be the strongest in each spectral region. If this methodology proves successful, it will be adopted by the ExoMol database for other very extensive line lists in future releases. At present these data are not included in the database. NaCl ---- The original NaCl ExoMol line lists of Barton [[*et al*]{}]{} [@jt583] computed using [Level]{} have been truncated by removing transitions with $\Delta v \geq 8$. KCl --- The original KCl ExoMol line lists of Barton [[*et al*]{}]{} [@jt583] computed using [Level]{} have been truncated by removing transitions with $\Delta v \geq 12$. PN -- The original PN ExoMol line lists of Yorke [[*et al*]{}]{} [@jt590] computed using [Level]{} have been truncated by removing transitions with $\Delta v \geq 6$. AlO --- A number of ExoMol users pointed out that the vibrational labels used in the original AlO line list were not the best ones. Our analysis of AlO state lifetimes [@jt623] reinforced this impression. The `.states` file has therefore been updated with revised vibrational state labels. We note that these labels are only approximate quantum numbers and this change should not affect any results obtained with the line list. H$_3^+$ ------- The H$_3^+$ line list of Neal [[*et al*]{}]{} [@jt181] (NMT) is the oldest in the ExoMol database. While it has continued to demonstrate its accuracy and predictive power [@jt512; @jt587], perhaps surprisingly so, there are other issues with it. The line list was constructed using Jacobi coordinates which do not allow for a full treatment of the symmetry. This led to the use of approximate nuclear spin statistical weights and, more problematically, to the removal of very weak transitions. Most of these transitions are symmetry-forbidden and should have zero dipole intensity. However “forbidden” rotational transitions [@86PaOk.H3+; @jt72] have proved important for both astrophysical [@02GoMcGe.H3+; @05OkGeGo.H3+] and laboratory [@jt306; @jt340] studies. Methods are available which allow for the proper treatment of symmetry [@jt358; @03ScAlHia.H3+]. A new line list [@jtH3+] is nearly complete which extends the range of the NMT line list. The upper energy limit is increased to 25 000 [cm$^{-1}$]{} and the highest rotational quantum number state considered in the calculations is $J=40$. This line list will remove the symmetry problem as well as improve both the accuracy and coverage for H$_3^+$, thus making it useful for both higher and lower temperatures. H$_2$O ------ The BT2 line list [@jt378] has been outstandingly successful: it was used for the original detection of water in an exoplanet atmosphere [@jt400], forms the basis of the well-used BT-Settl brown-dwarf model [@12AlHoFr] and provides the hot water line list for HITEMP [@jt480]. However, although it is more complete, for many levels it is less accurate than the Ames line list of Partridge and Schwenke [@97PaScxx.H2O]. Since the construction of BT2 there has been significant progress in deriving experimental energy levels for H$_2$$^{16}$O [@jt539; @jt562], improvements in the [*ab initio*]{} dipole moment surface [@jt509] which have been used to construct high accuracy room temperature line lists [@jt522] and further improvements in representing the potential energy surface of the molecule [@jt519]. A new H$_2$$^{16}$O line list, named POKAZATEL, [@jtpoz] which builds on these advances will be released soon. The completeness of this line list is illustrated by the following figures: The highest rotational quantum number used in the calculations of the wavefunctions was $J=72$ — close to the highest $J$ for wich the bound states still do exist. The highest upper energy levels limit is equal to 40 000 [cm$^{-1}$]{}, which is also close to the dissociation limit of water [@jt549]. While the BT2 and Ames line lists are seriously incomplete for temperatures above 3000 K, POKAZATEL considers every ro-vibrational transition in the molecule and therefore is appropriate for studies at higher temperatures. Line lists for hot H$_2$$^{17}$O and H$_2$$^{18}$O will also be released very shortly [@jtiso]. Again these will take advantage of the available set of experimental energy levels [@jt454; @jt482] obtained using the MARVEL procedure [@jt412; @12FuCsi.method]. These line lists will be as complete as BT2 and consider levels with $J \leq 50$ and energies up to 30 000 [cm$^{-1}$]{}. NH$_3$ ------ Ammonia is a ten electron system like water, so in terms of electronic calculations one might expect similar improvements here too. $^{14}$NH$_3$ has also been the subject of a systematic study of experimental energy levels [@jt608] and improvements in its [*ab initio*]{} treatment [@11HuScLe.NH3; @jt634]. New experimental data characterizing higher-lying energy levels is also available [@jt616; @jt633]. Again the Ames group have performed high accuracy studies on this system [@11HuScLe.NH3; @11HuScLe2.NH3; @12SuBrHu.NH3], although there is no corresponding line list. Work is therefore in progress to compute a new ammonia line list which will replace the existing BYTe line list [@jt500]; in particular this work the option of using the improved nuclear motion capability of the program [TROVE]{} working in curvilinear coordinates [@15YaYuxx.method]. Recent, excellent [*ab initio*]{} calculations [@jt634] which reproduce very highly excited energy levels of ammonia up to 18 000 [cm$^{-1}$]{}, will be used as the starting point for a semi-empirical fit of the PES. This should give a line list which is both more accurate and more extensive than BYTe. CrH --- Chromium hydride is an important astronomical molecule for a variety of reasons, including classification of brown dwarfs [@99KiReLi.CrH; @jt430] and measurements of magnetic fields [@13KuBexx.CrH]. The ExoMol database currently provides a substantially empirical line list due to Burrows [[*et al*]{}]{}[@02BuRaBe.CrH]; we are in the process of computing a more complete CrH line list covering 8 states and the 4 main isotopologues [@jtCrH]. TiH --- The current titanium hydride line list due to Burrows [[*et al*]{}]{}[@05BuDuBa.TiH] is in the process of being updated with a more extensive ExoMol one. Other species ------------- Besides the species listed above, line lists are in an advanced stage of construction for VO [@jt621], H$_2$O$_2$ [@jtHOOH] (see Al-Refaie [[*et al*]{}]{}[@jt620] for a preliminary, room temperature version), SO$_3$ [@jtSO3] (see also Underwood [[*et al*]{}]{}[@jt554]), C$_2$H$_4$, CH$_3$Cl (see also Owens [[*et al*]{}]{} [@jt612]), C$_2$, SiH, NS, NO, NaO, AlH and SH [@jtAlH], MnH, and PO, PS and PH [@jtPS]. Finally we note that work is in progress looking at TiO [@jtTiO]. TiO is a major absorber in cool stars [@00AlHaS1.TiO; @13DaKuPl.TiO]. There are TiO line lists available from Schwenke [@98Scxxxx.TiO] and Plez [@98Plxxxx.TiO]. However it would appear that these line lists are incomplete; in particular there are missing bands that have been observed in the laboratory [@99RaBeDu.TiO]. Furthermore, recent exoplanet studies suggest that the line frequencies are too inaccurate for high resolution studies [@15HoDeSn.TiO]. Data Provided ============= The original aim of the ExoMol project was the provision of extensive line lists of energy levels, and hence transition frequencies, and Einstein $A$ coefficients from which transition intensities and related information can be computed. The original ExoMol data structure [@jt548] provided precisely for these data in a concise manner. The demands on and utility of the data provided by the project has led us to expand the scope of the database. In this section we outline the new data provision and in the following sections we formally define the new data structures employed for this purpose. Extended states file -------------------- There are other properties of a molecular state that can be computed by ExoMol which we now propose to store. The first of these is the radiative lifetime of each individual state, which can be computed in a straightforward fashion from the Einstein $A$ coefficients [@jt624]. Such lifetimes have been shown to be important in both laboratory [@jt306; @jt340] and astronomical [@02GoMcGe.H3+; @05OkGeGo.H3+] environments. Indeed the radiative lifetime of a state is a key component in determining the critical density of species in that state and hence whether it exists in local thermodynamic equilibrium in a given environment. Our recent study of lifetimes based on use of ExoMol data suggested that methane should show particularly interesting lifetime effects [@jt624]. These computed lifetimes can be compared with experimental measurements, where available, providing a check on our calculated dipole moments. This is particularly useful for rovibronic spectra, where measurements of absolute transition intensities are unusual. The second state-dependent property are Landé $g$-factors. These provide the behaviour of the states in the presence of a weak magnetic field as given by the Zeeman effect. Molecular spectra can provide important information on magnetic fields in stars, brown dwarfs and exoplanets [@13KuBexx.CrH; @15AfBexx]. For open shell diatomics, Landé $g$-factors are given by a fairly straightforward formula in terms of standard Hund’s case (a) quantum numbers [@02BeSoxx.diatom]. The vibronic wavefunctions computed by [Duo]{} [@jt609] contain the necessary information to compute these $g$-factors and this has currently recently been implemented within [Duo]{} [@jtgfac]. The magnetic fields in some astronomical environments are strong enough that the Paschen-Back effect becomes important [@13KuBexx.CrH; @07ShFlBe.CN]. In this case the shifts in the energy levels do not have a simple dependence on the magnetic field [@05BeBrFl.diatom] and they will have to be explicitly computed on a case-by-case basis. Landé $g$-factors are only given for systems whose spectroscopic model includes open shell electronic states (ie ones with unpaired electrons) since these are the systems whose states which show significant splitting in a magnetic field. Besides (approximate) quantum numbers, which may appear in several forms, it is also often desirable to store more than one estimate of the energy level of the state. A general methodology, known as MARVEL, is available for obtaining self-consistent sets of empirical energy levels from networks of transitions [@jt412; @12FuCsi.method], which means that sets of empirical energy levels are available for key molecules [@jt454; @jt482; @jt539; @jt576; @13FuSzMa.H3+; @jtC2]. In this case it is preferable to replace computed energy levels with empirical ones, as has already been done for the HCN/HNC [@jt570] and the CS [@jt615] line lists. It is also recommended to keep the original computed levels as well as the uncertainties which are generally available for the empirical levels. The inclusion of approximate quantum numbers and extra energy levels included has been decided on a molecule-by-molecule basis; information on this is contained in the isotopologue definition file (see Section 5.3). Cross sections, $k$-coefficients and pressure effects ----------------------------------------------------- Many of the line lists stored in the database are huge. This makes finding other forms of storing the data desirable. ExoMol data has been used to generate temperature-dependent cross sections as a function of wavenumber [@jt542]. Such cross sections now form part of the database which will also be expanded to contain pressure-dependent cross sections and $k$-coefficient tables generated from them. Cross sections depend on pressure, as well as temperature, due to collisional broadening by the species in the atmosphere. The provision of temperature-, pressure- and broadening species-dependent cross sections would be both computationally very demanding and again result in very large datasets. Instead we have chosen to primarily provide parameters which characterise pressure profiles for key species plus software allowing cross sections to be generated for selected temperature-pressure parameters on the user’s own computer. There is a widespread recognition that for detailed studies, such as those performed in the Earth’s atmosphere at high resolution, Voigt profiles only provide an approximate solution to the problem [@jt584]. However Voigt profiles are in widespread use and are easily computed [@79Hum]; they are therefore used to represent pressure broadening effects in ExoMol. The new release of ExoMol includes, where possible, pressure broadening parameters. So far, these have not been considered for vibronic spectra. HITRAN provides a source of air-broadening and self-broadening parameters [@07GoRoGa.broad] so in the absence of other sources these are taken from HITRAN 2012 [@jt557]. In gas giants H$_2$ and He are the major broadeners. There is no systematic source of parameters for these broadeners so we have to treat each system on a case-by-case basis. While some work has been performed for broadening of hot water by H$_2$ and He [@jt544; @jt636], more data would clearly be helpful in this area. Hedges and Madhusudhan [@16HeMaxx] recently put this in context by quantifying the effects of various factors involved in modelling line broadening in exoplanetary atmospheres, including completeness of broadening parameters, on molecular absorption cross sections. ExoMol will provide Lorentzian half-widths and their temperature-exponents for key molecule-broadener systems. The availability of these parameters varies with species and broadener. In any case there are not individual values for every molecular line and often parameters have only been measured or calculated for a small fraction of transitions. Therefore, rather than pre-determine values for every molecular line and increase the size of the line lists, which are very large for some molecules, a separate pressure broadening parameters file is provided. This file contains three types of parameters: experimental, theoretical and semi-empirical. Experimental, theoretical and semi-empirical parameters from the literature are presented with their respective full or partial quantum number assignments. Additional semi-empirical parameters are determined by compiling all experimental and theoretical parameters as a function of $J{^{\prime\prime}}$, the total rotational quantum number of the lower level of the transition, and computing an average value for each $J{^{\prime\prime}}$. To avoid introducing additional error, no extrapolation is attempted beyond the $J{^{\prime\prime}}_{\rm max}$ for which data are available, the parameters are simply assumed to be constant from this point. This is a very basic model similar to that used by other line-width studies [@97BuMaHu.broad; @jt483; @14AmBaTr.broad]. Constructing the pressure broadening parameters file in this way ensures that parameters are provided for every spectral line. Table \[tab:broad\_ref\] lists the main sources used to provide pressure broadening parameters. Clearly further and improved parameters would be welcome. [llll]{} \ \ \ & Broadener & Reference & Methodology\ [  – *Continued from previous page*]{}\ Molecule & Broadener & Reference & Methodology\ \ H$_{2}$O & H$_{2}$ & Lavrentieva [[*et al*]{}]{} [@14LaVoNa.h2opb] & Semi-empirical\ & & Lavrentieva [[*et al*]{}]{} [@14LaDuMa.h2opb] & Semi-empirical\ & & Steyert [[*et al*]{}]{} [@Steyert2004183] & Experimental\ & & Brown & Plymate [@Brown1996263] & Experimental\ & & Brown [[*et al*]{}]{} [@05BrBeDe.h2opb] & Experimental\ & & Gamache [[*et al*]{}]{} [@Gamache1996471] & Theoretical\ & & Dick [[*et al*]{}]{} [@Dick2009619] & Experimental\ & & Faure [[*et al*]{}]{} [@jt544] & Theoretical\ & & Langlois [[*et al*]{}]{} [@Langlois1994272] & Experimental\ & & Dutta [[*et al*]{}]{} [@93DuJoGo.h2opb] & Experimental\ & & Golubiatnikov [@05Goxxxx.h2opb] & Semi-empirical\ & & Zeninari [[*et al*]{}]{} [@04ZePaCo.h2opb] & Semi-empirical\ & & Drouin and Wiesenfeld [@12DrWixx.h2opb] & Theoretical\ H$_{2}$O & He & Lavrentieva [[*et al*]{}]{} [@14LaVoNa.h2opb] & Semi-empirical\ & & Lavrentieva [[*et al*]{}]{} [@14LaDuMa.h2opb] & Semi-empirical\ & & Steyert [[*et al*]{}]{} [@Steyert2004183] & Experimental\ & & Gamache [[*et al*]{}]{} [@Gamache1996471] & Theoretical\ & & Dick [[*et al*]{}]{} [@Dick2009619] & Experimental\ & & Lazarev [[*et al*]{}]{} [@95LaPnSu.h2opb] & Experimental\ & & Dutta [[*et al*]{}]{} [@93DuJoGo.h2opb] & Experimental\ & & Golubiatnikov [@05Goxxxx.h2opb] & Semi-empirical\ & & Zeninari [[*et al*]{}]{} [@04ZePaCo.h2opb] & Semi-empirical\ H$_{2}$O & Air & Mérienne [[*et al*]{}]{} [@03MeJeHe.h2opb]$^*$ & Experimental\ & & Gamache & Hartmann [@04GaHaxx.h2opb]$^*$ & Semi-empirical\ & & Gasster [[*et al*]{}]{} [@88GaToGo.h2opb]$^*$ & Experimental\ & & Payne [[*et al*]{}]{} [@08PaDeCa.h2opb]$^*$ & Experimental\ & & Gamache [@05Gaxxxx.h2opb]$^*$ & Semi-empirical\ & & Gamache & Laraia [@09GaLaxx.h2opb] & Semi-empirical\ H$_{2}$O & H$_{2}$O & Mérienne [[*et al*]{}]{}  [@03MeJeHe.h2opb]$^*$ & Experimental\ & & Gamache & Hartmann [@04GaHaxx.h2opb]$^*$ & Semi-empirical\ & & Markov [@94Maxxxx.h2opb]$^*$ & Experimental\ & & Golubiatnikov [[*et al*]{}]{} [@08GoKoKr.h2opb]$^*$ & Experimental\ & & Cazzoli [[*et al*]{}]{} [@08CaPuBu.h2opb]$^*$ & Semi-empirical\ CH$_{4}$ & H$_{2}$ & Pine [@92Pixxxx.ch4pb] & Experimental\ & & Margolis [@93Maxxxx.ch4pb] & Experimental\ & & Fox [[*et al*]{}]{} [@88FoJeSt.ch4pb] & Experimental\ & & Strong [[*et al*]{}]{} [@93StTaCa.ch4pb] & Experimental\ CH$_{4}$ & He & Pine [@92Pixxxx.ch4pb] & Experimental\ & & Varanasi & Chudamani[@90VaChxx.ch4pb] & Experimental\ & & Gabard [[*et al*]{}]{} [@04GaGrGr.ch4pb] & Experimental\ & & Grigoriev [[*et al*]{}]{} [@01GrFiTo.ch4pb] & Semi-empirical\ & & Fox [[*et al*]{}]{} [@88FoJeSt.ch4pb] & Experimental\ CH$_{4}$ & Air & Predoi-Cross [[*et al*]{}]{} [@06PrBrDe.ch4pb]$^*$ & Experimental\ & & Smith [[*et al*]{}]{} [@09aSmBePr.ch4pb]$^*$ & Experimental\ & & Brown [[*et al*]{}]{} [@03BrBeCh.ch4pb]$^*$ & Semi-empirical\ & & Anthony [[*et al*]{}]{} [@08AnNiWr.ch4pb] & Theoretical\ CH$_{4}$ & CH$_{4}$ & Predoi-Cross [[*et al*]{}]{} [@05PrBrDe.ch4pb]$^*$ & Experimental\ & & Smith [[*et al*]{}]{} [@09bSmBePr.ch4pb]$^*$ & Experimental\ & & Brown [[*et al*]{}]{} [@03BrBeCh.ch4pb]$^*$ & Semi-empirical\ NH$_{3}$ & H$_{2}$ & Pine [[*et al*]{}]{} [@93PiMaBu.nh3pb]$^*$ & Experimental\ & & Hadded [[*et al*]{}]{} [@01HaArOr.nh3pb]$^*$ & Experimental\ & & Sharp & Burrows [@07ShBuxx.nh3pb]$^*$ & Semi-empirical\ NH$_{3}$ & He & Pine [[*et al*]{}]{} [@93PiMaBu.nh3pb]$^*$ & Experimental\ & & Hadded [[*et al*]{}]{} [@01HaArOr.nh3pb]$^*$ & Experimental\ & & Sharp & Burrows [@07ShBuxx.nh3pb]$^*$ & Semi-empirical\ NH$_{3}$ & Air & Brown & Peterson [@94BrPexx.nh3pb]$^*$ & Experimental\ NH$_{3}$ & NH$_{3}$ & Brown & Peterson [@94BrPexx.nh3pb]$^*$ & Experimental\ PH$_{3}$ & H$_{2}$ & Bouanich [[*et al*]{}]{} [@Bouanich2004195] & Semi-empirical\ & & Levy [[*et al*]{}]{} [@Levy1993172] & Experimental\ & & Sergent-Rozey [[*et al*]{}]{} [@SergentRozey198866] & Experimental\ & & Salem [[*et al*]{}]{} [@Salem200423] & Experimental\ & & Pickett [[*et al*]{}]{} [@Pickett1981197] & Experimental\ & & Levy [[*et al*]{}]{} [@Levy199420] & Experimental\ PH$_{3}$ & He & Salem [[*et al*]{}]{} [@Salem2005247] & Experimental\ & & Levy [[*et al*]{}]{} [@Levy1993172] & Experimental\ & & Sergent-Rozey [[*et al*]{}]{} [@SergentRozey198866] & Experimental\ & & Pickett [[*et al*]{}]{} [@Pickett1981197] & Experimental\ & & Levy [[*et al*]{}]{}  [@Levy199420] & Experimental\ PH$_{3}$ & Air & Butler [[*et al*]{}]{} [@06BuSaKl.ph3pb]$^*$ & Experimental\ & & Kleiner [[*et al*]{}]{} [@Kleiner2003293]$^*$ & Semi-empirical\ PH$_{3}$ & PH$_{3}$ & Butler [[*et al*]{}]{} [@06BuSaKl.ph3pb]$^*$ & Experimental\ & & Kleiner [[*et al*]{}]{} [@Kleiner2003293]$^*$ & Semi-empirical\ H$_{2}$CO & H$_{2}$ & Nerf [@75Nexxxx.hrcopb] & Experimental\ H$_{2}$CO & He & Nerf [@75Nexxxx.hrcopb] & Experimental\ H$_{2}$CO & N$_{2}$ & Jacquemart [[*et al*]{}]{} [@10JaLaKw.hrcopb] & Semi-empirical\ H$_{2}$CO & H$_{2}$O& Jacquemart [[*et al*]{}]{} [@10JaLaKw.hrcopb]$^*$ & Semi-empirical\ HCN & Air & Yang [[*et al*]{}]{} [@08YaBuGo.hcnpb]$^*$ & Experimental\ & & Devi [[*et al*]{}]{} [@04DeBeSm.hcnpb]$^*$ & Semi-empirical\ & & Risland [[*et al*]{}]{} [@03RiDeSm.hcnpb]$^*$ & Semi-empirical\ HCN & HCN & Devi [[*et al*]{}]{} [@04DeBeSm.hcnpb]$^*$ & Semi-empirical\ & & Devi [[*et al*]{}]{} [@03DeBeSm.hcnpb]$^*$ & Semi-empirical\ CS & Air & Blanquet [[*et al*]{}]{} [@99BaWaBo.cspb]$^*$ & Semi-empirical\ CS & CS & Misago [[*et al*]{}]{} [@09MiLeBo.cspb]$^*$ & Semi-empirical\ Another compact form of data input to exoplanet modelling codes involves the use of the $k$-coefficient approximation [@92FuLixx.method; @95Kratzx.method; @96IrCaTa.method], for example, by the NEMESIS code [@08IrTeKo.model] and cross sections, as used by Tau-REx [@13HoTeTi.exo; @jt593; @jt611]. Temperature- and pressure-dependent $k$-coefficients are also been provided for certain key species. Partition and cooling functions ------------------------------- In practice the ExoMol project always required partition functions, the study of which has often been performed independently [@jt169; @jt263; @jt304; @jt571], and on occasion adopted from other studies [@08WeChBo.CH4]. Temperature-dependent partition functions are now formally included as part of the data structure. Of course in various environments such as the early Universe and regions of star or planetary formation, the transformation of energy into radiation through molecular emissions provide an important source of cooling. Temperature-dependent cooling functions are important [@jt489; @jt506; @jt551] and can also be computed from ExoMol line lists [@jt624]. These are also now included in the data structure. Dipoles ------- While Einstein $A$ coefficients are sufficient for the vast majority of radiative transport applications, the construction of these loses the phase information contained in the individual, complex-valued transition dipole moments. Incorporating the phase information in the ExoMol database will make it useful for theoretical modelling of the effects of an electric field on molecular structure. The areas of application includes cooling and trapping of molecular beams (see Ref. [@Lemeshko13] and references therein), manipulating the long-range molecular interactions and collisional dynamics [@Campbell09; @Lemeshko09; @Lemeshko12], molecular orientation and alignment [@Boca00; @Friedrich00; @Lemeshko13], as well as rotational spectroscopy [@Aldegunde08; @Aldegunde09; @Stuhl12; @Garttner13]. As shown in Section \[dipole\_file\], the absolute values of transition dipole moments can be computed from the Einstein $A$ coefficients and it is sufficient to complement this with only the sign of transition dipole moment for each molecular line. Spectroscopic Models -------------------- In addition to the spectroscopic data provided by the ExoMol database, each ExoMol line list has been constructed from a detailed spectroscopic model of the given molecule. In general this model includes a (spectroscopically-determined) potential energy surface and [*ab initio*]{} dipole moment surfaces, and input to the appropriate nuclear motion program. For systems with multiple electronic states, the spectroscopic model also includes coupling between electronic states, e.g. spin-orbit, electronic angular momentum etc. These spectroscopic models are a valuable product of the ExoMol procedure and there are several reasons for preserving them. Firstly, the spectroscopic model is ultimately the source of error in the final line list and is thus important when investigating discrepancies between experiment or astronomical observations and theoretical predictions. Secondly, they can be used as a starting point for future models or line lists. For example, new experimental data can be incorporated through refinement of potential energy curves; improved [*ab initio*]{} methodologies can be incorporated by improvements to the dipole moment surface, or further electronic states can be added to the model. Finally, they can be useful for other applications besides calculation of the original line list. Current work within the ExoMol group is investigating the sensitivity of spectral lines to a possible variation in the proton-to-electron mass ratio [@15OwYuTh.NH3; @15OwYuPo.H3Op]. An important part of each spectroscopic model is the input file to the appropriate nuclear motion program (e.g. [ Level]{} [@lr07], [Duo]{} [@jt609], [DVR3D]{} [@jt338], [TROVE]{} [@07YuThJe.method]). For the diatomic codes [ Level]{} and [Duo]{}, the program input generally includes full details of the appropriate (potential, dipole and coupling) curves. For the polyatomic nuclear motion programs [DVR3D]{} and [TROVE]{}, the PES and DMS are provided as standalone functions, usually written in Fortran, with the program inputs as separate files. These spectroscopic models are not part of the formal ExoMol data structure described in the next section but can be accessed from the webpage of the appropriate isotopologue in files labelled `.model`. Data Structures =============== The data structure outlined below represents a significant extension of the original ExoMol data structure described by Tennyson [[*et al*]{}]{}[@jt548]. However the core structure of the two files, the states and transitions files, used in the original specification remains unchanged meaning that utilities designed to work with the previous structure will still work. A summary of the contents of the database is stored in a master file, exomol.all. This file points towards the files which store the actual data. Table \[tab:files\] gives an overview of these files, which are available for each isotopologue under the updated ExoMol structure. Each of these files is specified in turn below. Figure \[f:tree\] gives a schematic representation of the file structure of the exomol data. ![Structure of the ExoMol data files. There is a single master file describing the database, $N_{\rm mol}$ molecules and $N_{\rm iso}$ files per molecule.[]{data-label="f:tree"}](Tree.jpg){width="79.50000%"} File extension $N_{\rm files}$ File DSname Contents ---------------- ----------------- -------------------- ------------------------------------------------------------------- `.all` 1 Master Single file defining contents of the ExoMol database.. `.def` $N_{\rm tot}$ Definition Defines contents of other files for each isotopologue. `.states` $N_{\rm tot}$ States Energy levels, quantum numbers, lifetimes, (Landé $g$-factors). `.trans` $^a$ Transitions Einstein $A$ coefficients, (wavenumber). `.broad` $N_{\rm mol}$ Broadening Parameters for pressure-dependent line profiles. `.cross` $^b$ cross sections Temperature or temperature and pressure-dependent cross sections. `.kcoef` $^c$ $k$-coefficients Temperature and pressure-dependent $k$-coefficients. `.pf` $N_{\rm tot}$ Partition function Temperature-dependent partition function, (cooling function). `.dipoles` $N_{\rm tot}$ Dipoles Transition dipoles including phases. `.overview` $N_{\rm mol}$ Overview Overview of datasets available. `.readme` $N_{\rm iso}$ Readme Specifies data formats. `.model` $N_{\rm iso}$ Model Model specification. : Specification of the ExoMol file types. (Contents in brackets are optional.)[]{data-label="tab:files"} $N_{\rm files}$ total number of possible files.\ $N_{\rm mol}$ Number of molecules in the database\ $N_{\rm tot}$ is the sum of $N_{\rm iso}$ for the $N_{\rm mol}$ molecules in the database\ $N_{\rm iso}$ Number of isotopologues considered for the given molecule\ $^a$ There are $N_{\rm tot}$ sets of `.trans` files but for molecules with large numbers of transitions the `.trans` files are subdivided into wavenumber regions.\ $^b$ There are $N_{\rm cross}$ sets of `.cross` files for isotopoplogue.\ $^c$ There are $N_{\rm kcoef}$ sets of `.kcoef` files for each isotopoplogue.\ File naming convention ---------------------- The naming convention for the files in the ExoMol database has the following specification. All file names, with the exception of the Pressure Broadening file, have a common part built using the template `Iso-slug__LineList__`. The Iso-slug“ is a machine-readable, plain ASCII-text XML-safe” unique identifier for the molecular isotopologue, for example, `14N-1H3` is the iso-slug for $^{14}$NH$_3$. Ions are denoted by an appended ’`_mZ`’ or ’`_pZ`’ for charges $-Z$ and $+Z$ respectively (with the value of Z omitted if it is unity); so `1H3_p` is the iso-slug for $^1$H$_3^{+}$. Linelist" is the line list publication name, see the column “DSName” in Tables \[tab:exomoldata\] and \[tab:otherdata\]; for example, the published names for the $^{14}$NH$_3$ and H$_3^{+}$ line lists are BYTe and NMT respectively. For external datasets that have been cast in ExoMol format, this name is an 8 character year-authors string which follows the convention of the IUPAC Task Group on water spectroscopy [@jt454]. The extensive bibliography files, in BibTeX format, provided on the ExoMol website for each molecule in the database, uses the same naming convention. This common part of the file name can be followed by additional identifiers specific to different file types. For example, the transition file `.trans` for polyatomic molecules is usually split into smaller wavenumber ranges in order to simplify the manipulation of data. In this case the `Iso-slug__LineList__` is followed by `NUMIN-NUMAX`, which specifies the wavenumber range in [cm$^{-1}$]{}. The absorption cross section `.cross` file names also contain a wavenumber range, representing the full coverage of the line list, followed by the temperature in Kelvin, pressure in bar and the binning interval, or resolution (in [cm$^{-1}$]{}), for which the cross sections have been computed. The file names have the form `NUMIN-NUMAX__Temperature[K]__Pressure[bar]__resolution` after the common part. In case of the Pressure Broadening file `.broad`, the file name does not contain the string `LineList__` as the data in the broadening files is not line list specific and has been compiled from a number of sources. Here the string `Iso-slug__` is simply followed by `Broadener` identifying the broadener. For example, the current BT2 line list package consists of the definition file `1H2-16O__BT2.def`, the States file `1H2-16O__BT2.states`, sixty Transitions files: `1H2-16O__BT2__00000-00500.trans`, `1H2-16O__BT2__00500-01000.trans`, …`1H2-16O__BT2__29500-30000.trans`, Partition function file `1H2-16O__BT2.pf`, Pressure Broadening files: `1H2-16O__H2.broad`, `1H2-16O__He.broad`, `1H2-16O__air.broad`, `1H2-16O__self.broad` and eighteen cross section files: `1H2-16O__BT2__00000-30000__296K__0bar__0.01.cross`, `1H2-16O__BT2__00000-30000__400K__0bar__0.01.cross`, …`1H2-16O__BT2__00000-30000__900K__0bar__0.01.cross`, `1H2-16O__BT2__00000-30000__1000K__0bar__0.01.cross`, `1H2-16O__BT2__00000-30000__1200K__0bar__0.01.cross`, …`1H2-16O__BT2__00000-30000__3000K__0bar__0.01.cross`, Here `1H2-16O` is the iso-slug for H$_2{}^{16}$O, `BT2` is the name of the water line list [@jt378], `00000-00500` is the wavenumber range 0–500 cm$^{-1}$, H$_2$, He, air and self are the broadeners, 296 K is the temperature, 0 bar is the pressure and 0.01 is the binning interval for the cross sections in wavenumbers. For most cases a single `.broad` file is provided for a given molecule and designated for the parent (most abundant) isotopologue. Should broadening parameters be required for other isotopologues, this file should be employed. In a few cases where isotopic substitution lowers the symmetry, e.g. HDO as compared to H$_2$O, then extra isotopologue-specific broadening files may be provided, although thus far there are no such files. |exomol.all| [lrll]{} Field & Fortran Format & C Format & Description\ \ `ID` & A13 & %13s & Always the ASCII string “EXOMOL.master”\ $V_{\rm all}$ & I8 & %8d & Version number, format YYYYMMDD, recording\ & & & last update of the whole database\ $N_{\rm mol}$ & I4 & %4d & Number of molecules in the database\ \ \ \ $N_{\rm name}$ & I3 & %3d& Number of molecule names listed.\ \ `MolName` & A27 & %27s & (Common) name of the molecule\ \ `MolFormula` & A27 & %27s & Molecule chemical formula\ $N_{\rm iso}$ & I4 & %4d& Number of isotopologues considered\ \ `MolKey` & A27 & %27s & Inchi key of the isotopologue\ `IsoFormula` & A27 & %27s & Isotopologue chemical formula\ `Iso-slug` & A160 & %160s & Isotopologue slug identifier, see text for details\ `DSName` & A10 & %10s & Isotopologue dataset name\ $V$ & I8 & %8d & Version number with format YYYYMMDD\ \ Metadata& & & Free format\ \[tab:all\] The master file --------------- The file exomol.all is designed to be machine searchable making it easy to check the current contents of the database and when any of it was last updated. The file is structured as a list of molecules with the isotopologues associated with that molecule. The actual ExoMol data files are stored by isotopologue. Table \[tab:all\] specifies the format of the file while Table \[tab:alleg\] gives a portion of the current contents for water with two isotopologues, H$_2$$^{16}$O and HD$^{16}$O. The final entry gives web-searchable meta-data. |exomol.all| -------------------------------------- ---------------------------------------------------- ` EXOMOL.master ` ` # ID ` ` 20160315 ` ` # Version number with format YYYYMMDD ` ` 30 ` ` # Number of molecules in the database ` ` 1 ` ` # Number of molecule names listed ` ` water ` ` # Name of the molecule ` ` H2O ` ` # Molecule chemical formula ` ` 2 ` ` # Number of isotopologues considered ` ` XLYOFNOQVPJJNP-DYCDLGHISA-N ` ` # Inchi key of molecule ` ` 1H-2H-16O ` ` # Iso-slug ` ` (1H)(2H)(16O) ` ` # IsoFormula ` ` VTT ` ` # Isotopologue dataset name ` ` 20160314 ` ` # Version number with format YYYYMMDD ` ` XLYOFNOQVPJJNP-UHFFFAOYSA-N ` ` # Inchi key of molecule ` ` 1H2-16O ` ` # Iso-slug ` ` (1H)2(16O) ` ` # IsoFormula ` ` BT2 ` ` # Isotopologue dataset name ` ` 20160220 ` ` # Version number with format YYYYMMDD ` -------------------------------------- ---------------------------------------------------- : Extract from the master file showing the three line header and the portion for the water molecule.[]{data-label="tab:alleg"} The definition file ------------------- A new addition to the ExoMol format is the inclusion of the definition file with extension `.def`. The definition file gives information on what ExoMol provides for a particular isotopologue and describes how the data can be used. The definition file serves multiple purposes: - **Standardized ExoMol file usage**: The ExoMol format [@jt548] `.states` file contains 4 standard fields: The ID of the state, the energy, the total degeneracy and the $J$ quantum number. However additional fields such as the symmetry and vibrational quantum numbers are not standardized and differ between molecules. For usage that requires these additional quantum numbers, a subroutine or function must be created for each molecule to be read. The definition file therefore provides a means to overcome this by providing a single subroutine with the information to read any state file provided by the ExoMol project. This allows codes to easily integrate the ExoMol format only once and provide support for all molecules in the ExoMol project without additional changes. - **Improved database function**: The definition file also provides detailed information on all fields in the ExoMol format. This means that database functionality such as sorting, filtering, splitting, selecting etc on specific or multiple fields can be easily performed on single or multiple molecules. - **Facilitate updates**: Finally, updates of particular molecules can be easily propagated to any user of the ExoMol format. Each definition file is assigned a version number; this allows the user to check whether there are any updates to a particular molecule, with the possibility of an automatic download of any modified files and immediate usage without change to codes. The format is defined in Table \[tab:def\_format\]. The data format can be subdivided into different sections (highlighted in bold) describing a different aspect of the molecule in the ExoMol format. Each section may contain multiple subsections that define further aspects of that section. The file format therefore takes a tree structure. An example definition file for the CS line list [@jt615] is given in Table. \[tab:CS\_example\] and for the BT2 water line list [@jt378] in Table. \[tab:BT2\_example\]. ### Header The first section describes metadata related to internal molecular data. $V$ is the version number of the definition file, it utilises a numerical dating format $YYYYMMDD$ in order to allow for easy chronological ordering of different versions by integer comparison. The `MolKey` is the 27 character InchiKey [@inchikey] used to identify the molecule and $N_{\rm atom}$ describes the number of atoms in the molecule. This allows code to modify its behaviour based on whether it is dealing with diatomic molecules $N_{\rm atom}=2$ or polyatomics $N_{\rm atom}>2$. [lrll]{} \ \[tab:def\_format\]\ & Fortran Format & C Format & Description\ [  – *Continued from previous page*]{}\ Field & Fortran Format & C Format & Description\ \ \ `ID` & A10 & %10s & Always the ASCII string “EXOMOL.def”\ `IsoFormula` & A27 & %27s & Isotopologue chemical formula\ `Iso-slug` & A160 & %160s & Isotopologue identifier, see text for details\ `DSName` & A10 & %10s & Isotopologue dataset name\ $V$ & I8 & %8d & Version number with format YYYYMMDD\ `MolKey` & A27 & %27s & Standard inchi key of the molecule\ $N_{\rm atom}$ & I4 & %4d & Number of atoms\ \ $I_{\rm atom}$ & I3 & %3d & Isotope number\ Atom & A3 & %3s & Element symbol\ \ $m_{Da} \quad m_{kg}$ & [F12.6,1X,ES14.8]{} & [%12.6f %14.8e]{} & Isotopologue mass in Da and kg\ $I_{\rm sym}$ & A6 & %6s & [Molecular symmetry Group (if $N_{\rm atom}=2$ then C or D)]{}\ $N_{\rm irrep}$ & I4 & %4d & Number of irreducible representations\ \ $I_{\rm irrep}$ & I3 & %3d & Irreducible representation ID\ Symmetry & A6 & %6s & Irreducible representation label\ $g_{\rm ns}$ & I3 & %3d & Nuclear spin degeneracy\ \ $T_{\rm max}$ & F8.2 & %8.2f & Maximum temperature of the line list\ $N_{\rm broad}$ & I3 & %3d & No. of pressure broadeners available\ $D_{\rm avail}$ & I2 & %2d & Dipole availability (1=Yes 0=No)\ $N_{\rm cross}$ & I3 & %3d & No. of cross section files available\ $N_{\rm kcoef}$ & I3 & %3d & No. of $k$-coefficient files available\ \ Life$_{\rm avail}$ & I2 & %2d & Flag denoting lifetime availability (1=Yes 0=No)\ $g_{\rm avail}$ & I2 & %2d & [Flag denoting Landé $g$-factor availability (1=Yes 0=No)]{}\ $N_{\rm states}$ & I10 & %10d & No. of states in `.states` file\ $N_{\rm cases}$ & I3 & %3d & No. of quanta cases\ \ Case Label & A8 & %8s & Label of the quanta case\ $N_{\rm quanta}$ & I3 & %3d & No. of quanta defined\ \ Quanta Label & A3 & %3s &\ $F_{\rm Fortran}\quad F_{\rm C}$ & A8,1x,A8 & %8s %8s& Fortran and C format for quanta\ Description & A40 & %40s & Short description of quanta\ \ $N_{\rm trans}$ & I15 & %15d & Total number of transitions\ $N_{\rm files}$ & I4 & %4d & Number of transition files\ $\tilde{v}_{\rm max}$ & F8.2 & %8.2f & Maximum wavenumber (in [cm$^{-1}$]{})\ \ $T^Q_{\rm max}$ & F8.2 & %8.2f & Maximum temperature of partition functions\ $T_{\rm Step}$ & F5.2 & %5.2f & Temperature step size\ $C_{\rm avail}$ & I2 & %2d & Cooling function availability (1=Yes 0=No)\ \ $\gamma_{D}$ & F6.4 & %6.4f & [Default value of Lorentzian half-width for all lines (in [cm$^{-1}$]{}/bar)]{}\ $n_{D}$ & F5.3 & %5.3f & [Default value of the temperature exponent for all lines]{}\ \ Broadener Label & A8 & %8s & Label for particular broadener\ & %20s & File name of particular broadener\ $J_{\rm max}$ & I7/F7.1 & %7d/%7.1f & [Maximum $J{^{\prime\prime}}$ for which pressure broadened parameters are provided]{}\ $\gamma_{L}$ & F6.4 & %6.4f & [Value of Lorentzian half-width for lines with $J{^{\prime\prime}}> J_{\rm max}$ (in [cm$^{-1}$]{}/bar)]{}\ $n_{L}$ & F5.3 & %5.3f & [Value of the temperature exponent for lines with $J{^{\prime\prime}}> J_{\rm max}$]{}\ $N_Q$ & I4 & %4d & Number of defined quantum number sets\ \ Set Code & A2 & %2s & A code that defines this set of quantum numbers\ $N_{\rm lines}$ & I6 & %6d & No. of lines in the broad that contain this code\ $N_{\rm Quanta}$ & I4 & %4d & No. of quantum numbers defined\ \ Quantum Label & A4 & %4s &\ ### Atom Definition This section describes each of the atoms in the molecule. The $N_{\rm atom}$ field describes how many of them are defined. An integer mass number and a string of the molecule allows a definite description of the isotopologue. ### Isotopologue Information This section gives mass and symmetry information about the molecules. $m$ is the molecular mass given in both Da and kg; the mass is used for modeling Doppler broadening. Next is the symmetry information of the molecule. The molecular symmetry group [@04BuJexx.method] and number of irreducible representations (irreps), $N_{\rm irrep}$, are described here, after which $N_{\rm irrep}$ lines defining each irrep in that group are given. The $I_{\rm sym}$ field gives the ID of the symmetry used in the `.states` file. However, for linear molecules including diatomics the symmetry group label will only take values of *C* or *D* corresponding to $C_{\infty v}$ or $D_{\infty h}$ respectively and $N_{\rm sym}$ is set to 0. $g_{\rm ns}$ corresponding to nuclear spin degenracy will be provided for *C*. For *D*, $g_{\rm even}$ and $g_{\rm odd}$ are provided; these correspond to the nuclear-spin degeneracy factor needed for when the rovibronic wavefunction (i.e. the wavefunction without nuclear spin) is respectively even or odd with respect to interchange of identical atoms. Otherwise we follow symmetry labels and conventions as set out by Bunker and Jensen [@04BuJexx.method]. In this case $g_{\rm ns}$ is be provided for each irreducible representation. ### ExoMol Information This section relates to the availability of certain aspects of the line list. First and foremost it describes the maximum temperature for which the line list is applicable ($T_{\rm max}$). Secondly, for molecules with pressure broadening parameters, it describes the number of broadeners ($N_{\rm broad}$) available; this is used later on to determine what additional structures should be parsed in the file. Finally, there is a boolean field which describes whether we provide dipoles for the particular molecule. ### State file information This section is the most important as it describes how the `.states` file format is structured. The first two fields describe whether lifetime information and Landé $g$-factors are available. These will be placed after the initial ‘standard’ fields in the states file. Then there is the $N_{\rm states}$ field which gives the number of states in the file. Finally the number of quanta cases, $N_{\rm cases}$. A quanta case describes what the following quantum numbers represent. For example, commonly for polyatomics the ExoMol `.states` file provides two representations of quantum numbers; the standard normal mode quantum numbers and the local mode ‘TROVE’ quantum numbers. For this example $N_{\rm cases}=2$; see Down [[*et al*]{}]{}[@jt546] for instance. Each case is given a label and must describe the number of quanta defined $N_{\rm quanta}$, after which $N_{\rm quanta}$ lines will follow, each describing a particular quantum number. Quantum numbers are described by a string label comprising a string format identifier and a short, human-readable description. The string format identifier utilises the FORTRAN string formatting to allow for the construction of formats more easily for FORTRAN which has stricter string handling than C/C++. An overview of the new `.states` format is given in Section \[ss:states\]. ### Transition and partition information This section simply describes the availability of `.trans` and `.pf` files. In particular, for the partition function file (`.pf`) it describes the maximum temperature provided ($T^Q_{\rm max}$), the step size of the temperature ($T_{\rm Step}$) and whether a cooling function is also present. ### Pressure broadening information Finally, the pressure broadening information is provided here. If $N_{\rm broad}=0$ this simply contains default values of Lorentzian half-width $\gamma_{L}$ and temperature exponent $n_{L}$. If $N_{\rm broad}>0$ this section has the following structure: the broadener is defined first, immediately followed by the associated file name, $J_{\rm max}$, and the asymptotic values of the Lorentzian half-width $\gamma_{L}$ and temperature exponent $n_{L}$ for the $a0$ quantum number set (see Section 5.6). $N_Q$ quantum number sets are explicitly defined by a label (e.g. $a0$) that corresponds to codes in the broadener file, the number of lines in the broadener file with that code and the number of quantum numbers in that set. After this each additional quantum number is also described; the total angular momentum quantum number of the lower state $J"$ is compulsory. The quantum numbers relate to the quanta defined previously in the states section followed by either a prime or double prime. Further details relating to pressure broadening are given in Section \[ss:broad\]. The final field of the definition file is used for keywords to facilitate the search of the database by web search engines. |12C-32S\_\_JnK.def| -------------------------------------------- -------------------------------------------------------------------------------- ` EXOMOL.def ` `# ID ` ` (12C)(32S) ` `# IsoFormula ` ` 12C-32S ` `# Iso-slug ` ` JnK ` `# Isotopologue dataset name ` ` 20160217 ` `# Version number with format YYYYMMDD ` ` DXHPZXWIPWDXHJ-UHFFFAOYSA-N ` `# Inchi key of molecule ` ` 2 ` `# Number of atoms ` ` 12 ` `# Isotope number 1 ` ` C ` `# Element symbol 1 ` ` 32 ` `# Isotope number 2 ` ` S ` `# Element symbol 2 ` ` 43.972071 7.30173406e-26 ` `# Isotopologue mass (Da) and (kg) ` ` C ` `# Symmetry group ` ` 1 ` `# Number irreducible representations ` ` 1 ` `# Irreducible representation ID ` ` 0 ` `# Irreducible representation label ` ` 1 ` `# Nuclear spin degeneracy ` ` 3000 ` `# Maximum temperature of linelist ` ` 2 ` `# No. of pressure broadeners available ` ` 0 ` `# Dipole availability (1=yes, 0=no) ` ` 0 ` `# No. of cross section files available ` ` 0 ` `# No. of k-coefficient files available ` ` 0 ` `# Lifetime availability (1=yes, 0=no) ` ` 0 ` `# Lande g-factor availability (1=yes, 0=no) ` ` 11497 ` `# No. of states in .states file ` ` 1 ` `# No. of quanta cases ` ` dcs ` `# Quantum case label ` ` 5 ` `# No. of quanta defined ` ` i ` `# Quantum label 1 ` ` I12 %12d ` `# Format quantum label 1 ` ` State ID ` `# Description quantum label 1 ` ` E ` `# Quantum label 2 ` ` F12.6 %12.6f ` `# Format quantum label 2 ` ` State energy in cm-1 ` `# Description quantum label 2 ` ` g ` `# Quantum label 3 ` ` I6 %6d ` `# Format quantum label 3 ` ` State degeneracy ` `# Description quantum label 3 ` ` J ` `# Quantum label 4 ` ` I7 %7d ` `# Format quantum label 4 ` ` Total rotational angular momentum ` `# Description quantum label 4 ` ` v ` `# Quantum label 5 ` ` I4 %4d ` `# Format quantum label 5 ` ` State vibrational quantum number ` `# Description quantum label 5 ` ` 199045 ` `# Total number of transitions ` ` 1 ` `# No. of transition files ` ` 10996.09 ` `# Maximum wavenumber (in cm-1) ` ` 3000.0 ` `# Maximum temperature of partition function ` ` 1.00 ` `# Temperature step size ` ` 0 ` `# Cooling function availability (1=yes, 0=no) ` ` 0.0700 ` `# Default value of Lorentzian half-width for all lines (in cm-1/bar) ` ` 0.500 ` `# Default value of temperature exponent for all lines ` ` air ` `# Label for a particular broadener ` ` 12C-32S__air.broad ` `# Filename of particular broadener ` ` 46 ` `# Maximum J for which pressure broadening parameters provided ` ` 0.0698 ` `# Value of Lorentzian half-width for J" > Jmax ` ` 0.750 ` `# Value of temperature exponent for lines with J" > Jmax ` ` 1 ` `# Number of defined quantum number sets ` ` a0 ` `# A code that defines this set of quantum numbers ` ` 47 ` `# No. of lines in the broad that contain this code ` ` 0 ` `# No. of quantum numbers defined ` ` self ` `# Label for a particular broadener ` ` 12C-32S__self.broad ` `# Filename of particular broadener ` ` 46 ` `# Maximum J for which pressure broadening parameters provided ` ` 0.0620 ` `# Value of Lorentzian half-width for J" > Jmax ` ` 0.500 ` `# Value of temperature exponent for lines with J" > Jmax ` ` 1 ` `# Number of defined quantum number sets ` ` a0 ` `# A code that defines this set of quantum numbers ` ` 47 ` `# No. of lines in the broad that contain this code ` ` 0 ` `# No. of quantum numbers defined ` -------------------------------------------- -------------------------------------------------------------------------------- : File : the definition file for $^{12}$C$^{32}$S.[]{data-label="tab:CS_example"} |BT2.def| ---------------------------------------------- ------------------------------------------------------------------------------------------ ` EXOMOL.def ` ` # ID ` ` (1H)2(16O) ` ` # IsoFormula ` ` 1H2-16O ` ` # Iso-slug ` ` BT2 ` ` # Isotopologue dataset name ` ` 20160220 ` ` # Version number with format YYYYMMDD ` ` XLYOFNOQVPJJNP-UHFFFAOYSA-N ` ` # Inchi key of molecule ` ` 3 ` ` # Number of atoms ` ` 1 ` ` # Isotope number 1 ` ` H ` ` # Element symbol 1 ` ` 16 ` ` # Isotope number 2 ` ` O ` ` # Element symbol 2 ` ` 18.010565 2.99072463e-26 ` ` # Isotopologue mass (Da) and (kg) ` ` C2v ` ` # Symmetry group ` ` 4 ` ` # Number of irreducible representations ` ` 1 ` ` # Irreducible representation ID ` ` A1 ` ` # Irreducible representation label ` ` 1 ` ` # Nuclear spin degeneracy ` ` 2 ` ` # Irreducible representation ID ` ` A2 ` ` # Irreducible representation label ` ` 1 ` ` # Nuclear spin degeneracy ` ` 3 ` ` # Irreducible representation ID ` ` B1 ` ` # Irreducible representation label ` ` 3 ` ` # Nuclear spin degeneracy ` ` 4 ` ` # Irreducible representation ID ` ` B2 ` ` # Irreducible representation label ` ` 3 ` ` # Nuclear spin degeneracy ` ` 3000.0 ` ` # Maximum temperature of linelist ` ` 0 ` ` # No. of pressure broadeners available ` ` 0 ` ` # Dipole availability (1=yes, 0=no) ` ` 0 ` ` # No. of cross section files available ` ` 0 ` ` # No. of k-coefficient files available ` ` 0 ` ` # Lifetime availability (1=yes, 0=no) ` ` 0 ` ` # Lande g-factor availability (1=yes, 0=no) ` ` 221097 ` ` # No. of states in .states file ` ` 1 ` ` # No. of quanta cases ` ` nltcs ` ` # Quantum case label ` ` 14 ` ` # No. of quanta defined ` ` i ` ` # Quantum label 1 ` ` I12 %12d ` ` # Format quantum label 1 ` ` State ID ` ` # Description quantum label 1 ` ` E ` ` # Quantum label 2 ` ` F12.6 %12.6f ` ` # Format quantum label 2 ` ` State energy in cm-1 ` ` # Description quantum label 2 ` ` g_tot ` ` # Quantum label 3 ` ` I6 %6d ` ` # Format quantum label 3 ` ` Total degeneracy of the level ` ` # Description quantum label 3 ` ` J ` ` # Quantum label 4 ` ` I7 %7d ` ` # Format quantum label 4 ` ` Total rotational angular momentum ` ` # Description quantum label 4 ` ` +/- ` ` # Quantum label 5 ` ` A1 %1s ` ` # Format quantum label 5 ` ` Total parity ` ` # Description quantum label 5 ` ` Gamma ` ` # Quantum label 6 ` ` I2 %2d ` ` # Format quantum label 6 ` ` Symmetry block number (1) ` ` # Description quantum label 6 ` ` N ` ` # Quantum label 7 ` ` I10 %10d ` ` # Format quantum label 7 ` ` Reference within the symmetry block ` ` # Description quantum label 7 ` ` nucspin ` ` # Quantum label 8 ` ` A1 %1s ` ` # Format quantum label 8 ` ` Nuclear spin isomer label(2) ` ` # Description quantum label 8 ` ` Gamma_rve ` ` # Quantum label 9 ` ` A2 %2s ` ` # Format quantum label 9 ` ` Rovibrational symmetry label ` ` # Description quantum label 9 ` ` v1 ` ` # Quantum label 10 ` ` I2 %2d ` ` # Format quantum label 10 ` ` v1 symmetric stretch quantum number(3 ` ` # Description quantum label 10 ` ` v2 ` ` # Quantum label 11 ` ` I2 %2d ` ` # Format quantum label 11 ` ` v2 bend quantum number ` ` # Description quantum label 11 ` ` v3 ` ` # Quantum label 12 ` ` I2 %2d ` ` # Format quantum label 12 ` ` v2 asymmetric stretch quantum number ` ` # Description quantum label 12 ` ` Ka ` ` # Quantum label 13 ` ` I2 %2d ` ` # Format quantum label 13 ` ` Ka rotational quantum number ` ` # Description quantum label 13 ` ` Kc ` ` # Quantum label 14 ` ` I2 %2d ` ` # Format quantum label 14 ` ` Kc rotational quantum number ` ` # Description quantum label 14 ` ` 505806255 ` ` # Total number of transitions ` ` 16 ` ` # No. of transition files ` ` 29971.78 ` ` # Maximum wavenumber (in cm-1) ` ` 3000.0 ` ` # Maximum temperature of partition function ` ` 1.0 ` ` # Step size of temperature ` ` 1 ` ` # Cooling function availability (1=yes, 0=no) ` ` 0.0700 ` ` # Default value of Lorentzian half-width for all lines (in cm-1/bar) ` ` 0.500 ` ` # Default value of temperature exponent for all lines ` ---------------------------------------------- ------------------------------------------------------------------------------------------ : File : the definition file for H$_2$$^{16}$O.[]{data-label="tab:BT2_example"} States file {#ss:states} ----------- The basic ExoMol data structure is a comprehensive list of the molecular states involved. This file contains a numbered list of energy terms, in cm$^{-1}$ with the zero defined by the lowest level, immediately followed by the total statistical weight $g_\mathrm{tot} = (2J+1) \times g_{\rm ns}$ in the HITRAN convention [@HITRAN-A] which retains the full nuclear-spin degeneracy of all atoms. These three columns are sufficient, in conjunction with the Transitions file, to model absorption and/or emission intensities. The degeneracy is followed by the total angular momentum quantum number $J$ (integer or half-integer) The next columns are used for the state lifetimes (s) and/or Landé $g$-factors with availability determined in the definition file. This is followed by good (rigorous) quantum number(s) such as the state symmetry, parity or a total symmetry (irreducible representation). Then follows any approximate quantum numbers that are specified. There is some flexibility over this portion of the file which is defined on a case-by-case basis, see Ref. [@VAMDCcasebycase] for example. For polyatomic molecules the approximate quantum numbers include, for example, the vibrational normal mode $v_i$, local mode $n_i$ quantum numbers, rotational quantum numbers $K$, $K_a$, $K_c$, polyad numbers, vibrational and rotational symmetries etc. For the diatomic molecules these are the spin components $F_i$, $\Sigma$ or $\Omega$, electronic term designations, e.g $^{3}{}\Pi_{g}$, vibrational quantum numbers $v$, rotational angular momentum $N$ (Hund’s case (b)), projection of the electronic angular momentum, $\Lambda$, etc. The spin components quantum numbers can be integral, half-integral or even strings F1, F2, $\ldots$ If the theoretical energy is replaced by an experimental counterpart, the former is usually kept as an extra field. Regardless of molecule, the structure of any ExoMol state file can be determined from its complimentary definition file. The structure in order is: 1. Mandatory fields 2. Lifetime if Life$_{\rm avail}=1$ 3. Landé $g$-factor if $g_{\rm avail}=1$ 4. Quanta Cases in order of definition The detailed specification of the mandatory section of the states file is given in Table \[Tab:states\]. Using the state file information from Table \[tab:CS\_example\], the mandatory definition in the FORTRAN string format can easily be programmatically generated: "(I12,1x,F12.6,1x,I6,1x,I7,1x,ES12.4,1x,A1,1x,A1,1x,A9,1x,I2,1x,I2,1x,I2,1x, I2)" This format corresponds to the sample state file given in Table \[tab:levels\]. Note that for lifetimes it is necessary to deal with states with infinite radiative lifetimes (e.g. the ground state) and for which lifetimes are not available: for high-lying states not all downward transitions are considered so it is not possible to compute the lifetime. Here, and elsewhere in the database, infinity is specified by the string INF’ and unknown numbers by NaN’. Field Fortran Format C Format Description ------------------ ---------------- ------------- ------------------------------------------- $i$ `I12` `%12d` State ID $E$ `F12.6` `%12.6f` State energy in $\mathrm{cm^{-1}}$ $g_\mathrm{tot}$ `I6` `%6d` State degeneracy $J$ `I7/F7.1` `%7d/%7.1f` $J$-quantum number (integer/half-integer) ($\tau$) `ES12.4` `%12.4e` Lifetime in s (optional) ($g$) `F10.6` `%10.6f` Landé $g$-factor (optional) : Specification of the mandatory part of the states file and extra data options.[]{data-label="Tab:states"} [ID: state identifier: a non-negative integer index, starting at 1\ $J$: total angular momentum quantum, excluding nuclear spin\ Fortran format, $J$ integer: `(I12,1x,F12.6,1x,I6,I7,1x,ES12.4,1x,F10.6)`\ or $J$ half-integer: `(I12,1x,F12.6,1x,I6,F7.1,1x,ES12.4,1x,F10.6)`\ ]{} |40Ca-16O\_\_VBATHY.state| $i$ $g$ $J$ $\tau$ $g$ State $v$ $\Lambda$ $\Sigma$ $\Omega$ ------ -------------- ----- ----- ----------- ---------- --- --- ------------------- ----- ----------- ---------- ---------- -- -- -- -- 4051 8242.235601 15 7 6.509E-01 0.106733 + f ` a3Pi ` 0 1 1 2 4052 8294.404892 15 7 2.106E-03 0.018230 + f ` a3Pi ` 0 1 0 1 4053 8369.059680 15 7 4.594E-01 0.000045 + f ` a3Pi ` 0 1 -1 0 4054 8627.261523 15 7 6.479E-05 0.017893 + f ` Ap1Pi ` 0 1 0 1 4055 8781.692232 15 7 1.881E-01 0.106728 + f ` a3Pi ` 1 1 1 2 4056 8833.629438 15 7 1.466E-03 0.018239 + f ` a3Pi ` 1 1 0 1 4057 8907.535275 15 7 1.595E-01 0.000048 + f ` a3Pi ` 1 1 -1 0 4058 9167.417952 15 7 4.406E-05 0.017996 + f ` Ap1Pi ` 1 1 0 1 4059 9314.686162 15 7 1.019E-01 0.106678 + f ` a3Pi ` 2 1 1 2 4060 9365.515048 15 7 1.139E-03 0.018390 + f ` a3Pi ` 2 1 0 1 4061 9434.128287 15 7 7.821E-02 0.000113 + f ` a3Pi ` 2 1 -1 0 4062 9533.168336 15 7 2.160E-03 0.029567 + f ` b3Sigma+ ` 0 0 1 1 4063 9545.896549 15 7 1.114E-02 0.005414 + f ` b3Sigma+ ` 0 0 0 0 4064 9709.239641 15 7 3.342E-05 0.018256 + f ` Ap1Pi ` 2 1 0 1 4065 9841.636252 15 7 8.905E-02 0.106743 + f ` a3Pi ` 3 1 1 2 4066 9893.716709 15 7 1.039E-03 0.018284 + f ` a3Pi ` 3 1 0 1 4067 9966.036778 15 7 6.823E-02 0.000061 + f ` a3Pi ` 3 1 -1 0 4068 10103.239512 15 7 1.106E-03 0.026224 + f ` b3Sigma+ ` 1 0 1 1 4069 10115.858820 15 7 2.887E-03 0.008900 + f ` b3Sigma+ ` 1 0 0 0 4070 10240.534926 15 7 2.729E-05 0.018317 + f ` Ap1Pi ` 3 1 0 1 : Extract from the state file for $^{40}$Ca$^{16}$O, .[]{data-label="tab:levels"} [$i$: State counting number.\ $\tilde{E}$: State energy in [cm$^{-1}$]{}.\ $g$: State degeneracy.\ $J$: Total angular momentum.\ $\tau$: Lifetime in s.\ $g$: Landé $g$-factor\ $+/-$: Total parity.\ $e/f$: rotationless-parity.\ $v$: State vibrational quantum number.\ $\Lambda$: Projection of the electronic angular momentum.\ $\Sigma$: Projection of the electronic spin.\ $\Omega$: $\Omega=\Lambda+\Sigma$, projection of the total angular momentum.\ ]{} Transitions file ---------------- The transitions file has a simple structure, see Table \[tab:trans\]. Two pointers, $i$ and $f$ point to rows in the `.states` file to identify the upper and lower states involved plus the full information characterising these states. The $A$ gives the Einstein $A$ coefficient for this transition. This file can be very large but, optionally, there is a fourth column $\tilde{\nu}$ which gives the transition wavenumber. Otherwise this must be computed from the states file as the difference of the upper and lower energy levels. A sample transitions file is given in Table \[tab:transeg\]. Many of the molecules in the database are characterised by a very large number of transitions. To make the use of `.trans` tractable these files are often split into wavenumber regions, as discussed above. The `.trans` files are also compressed using `.bz2` format; a utility is provided for reading this format without needing to uncompress the file. Field Fortran Format C Format Description -------------------- ---------------- ---------- ------------------------------------------------ $i$ `I12` `%12d` Upper state ID $f$ `I12` `%12d` Lower state ID $A$ `ES10.4` `%10.4e` Einstein $A$ coefficient in $\mathrm{s^{-1}}$ $\tilde{\nu}_{fi}$ `E15.6` `%15.6e` Transition wavenumber in cm$^{-1}$ (optional). : Specification of the transitions file.[]{data-label="tab:trans"} Fortran format: `(I12,1x,I12,1x,ES10.4,1x,ES15.6)`\ |40Ca-16O\_\_VBATHY.trans| $i$ $f$ $A_{if}$ $\tilde{\nu}_{if}$ ------- ------- ------------ -------------------- 10571 10884 9.5518E-06 120.241863 21053 21375 1.9515E-05 120.242886 8726 9672 1.8658E-04 120.243522 11655 11950 5.0065E-06 120.243733 93209 93967 5.7055E-03 120.244192 2228 3175 7.3226E-07 120.244564 46727 46432 1.0599E-04 120.244658 44436 44774 1.4626E-04 120.245583 29037 28723 1.8052E-04 120.245669 4458 4805 1.0431E-08 120.246396 69313 68434 5.0531E-06 120.248178 22640 22985 1.1281E-07 120.248891 57027 56721 7.1064E-06 120.250180 : Extract from the transitions file for $^{40}$Ca$^{16}$O, .[]{data-label="tab:transeg"} $i$: Upper state counting number;\ $f$: Lower state counting number; $A_{if}$: Einstein $A$\ coefficient in s$^{-1}$; $\tilde{\nu}_{if}$: transiton wavenumber in [cm$^{-1}$]{}. Pressure broadening file {#ss:broad} ------------------------ Field Fortran Format C format Description ------------------------- ---------------- ----------- ----------------------------------------------------------------------------------- code A2 %2s Code identifying quantum number set following $J{^{\prime\prime}}$ $\gamma_{\textrm{ref}}$ F6.4 %6.4f Lorentzian half-width at reference tem perature and pressure in [cm$^{-1}$]{}/atm $n$ F6.3 %6.3f Temperature exponent $J{^{\prime\prime}}$ I7/F7.1 %7d/%7.1f Lower $J$-quantum number integer/half-integer : Specification of the mandatory part of the pressure broadening parameters file.[]{data-label="tab:broad_format"} Fortran format, $J$ integer: `(A2,1x,F6.4,1x,F6.3,1x,I7)`\ or $J$ half-integer: `(A2,1x,F6.4,1x,F6.3,1x,F7.1)`\ Like the `.states` file, the first four fields of the `.broad` file are mandatory for all records, see Table \[tab:broad\_format\]. This includes one quantum number, $J{^{\prime\prime}}$, which is always known and hence guarantees that at least a semi-empirical Lorentzian half-width ($\gamma_{\textrm{ref}}$) and temperature dependence, represented by exponent $n$ (see Eq. \[eq:TP\_dep\]), can be generated for every molecular line. Additional upper and lower state labels, which give molecule and quantum number dependent behavior, follow the compulsory fields. These are given in the definitions file along with the maximum $J{^{\prime\prime}}$ for which approximate parameters have been generated, $J{^{\prime\prime}}_{\rm max}$, and the respective values of Lorentzian half-width and temperature exponent, $\gamma_{\rm max}$ and $n_{\rm max}$. When $J{^{\prime\prime}}> J{^{\prime\prime}}_{\rm max}$, $\gamma_{\rm max}$ and $n_{\rm max}$ should be used for the Lorentzian half-width and temperature exponent respectively. The `.broad` file has a hierarchical structure; values of $\gamma_{\textrm{ref}}$ and $n$ with full quantum assignments are presented first, followed by values with partial quantum assignments, then finally values with $J{^{\prime\prime}}$ dependence only. This represents the preferential order in which the values of $\gamma_{\textrm{ref}}$ and $n$ should be used. This hierarchical structure is demonstrated in Table \[tab:broad\_H2O\_example\] where a portion of the H$_2$O-H$_2$ `.broad` file is presented. Once a suitable $\gamma_{\rm ref}$ and $n$ have been identified, the Lorentzian half-width of a spectral line at temperature $T$ and pressure $P$ can be calculated as: $$\label{eq:TP_dep} \gamma(T) = \gamma_{\textrm{ref}} \times \left(\frac{T_{\textrm{ref}}}{T}\right)^{n} \times \left(\frac{P}{P_{\textrm{ref}}}\right).$$ Here, $T_{\textrm{ref}}$ = 296 K and $P_{\textrm{ref}}$ = 1 bar. A separate `.broad` file is supplied for each molecule-broadener system and the broadener is specified in the file name. Example .broad files for CS-Air and H$_2$O-H$_2$ are given in Table \[tab:broad\_CS\_example\] and Table \[tab:broad\_H2O\_example\]. For particular molecule-broadener systems where no pressure broadening information is available, the default values of $\gamma_{\textrm{ref}}$ and $n$ given in the isotopologue’s definition file may be used. |12C-32S\_\_air.broad| ----- -------- ------- ---- a0 0.0860 0.096 0 a0 0.0850 0.093 1 a0 0.0840 0.091 2 a0 0.0840 0.089 3 a0 0.0830 0.087 4 ... a0 0.0720 0.067 35 a0 0.0720 0.066 36 ... ----- -------- ------- ---- : File : Air `.broad` file for $^{12}$C$^{32}$S: portion of the file (upper part); field specification (lower part).[]{data-label="tab:broad_CS_example"} Field Fortran Format C format Description ------------------------- ---------------- ---------- ---------------------------------------------------------------------------------- code A2 $\%$2s Code identifying quantum number set following $J{^{\prime\prime}}$\* $\gamma_{\textrm{ref}}$ F6.4 %6.4f Lorentzian half-width at reference temperature and pressure in [cm$^{-1}$]{}/bar $n$ F5.3 %5.3f Temperature exponent $J{^{\prime\prime}}$ I7/F7.1 $\%$7d Lower $J$-quantum number : File : Air `.broad` file for $^{12}$C$^{32}$S: portion of the file (upper part); field specification (lower part).[]{data-label="tab:broad_CS_example"} Code definition: a0 = none |1H2-16O\_\_H2.broad| ----- -------- ------- ---- ---- --- --- --- --- --- --- --- --- --- --- ... b2 0.0356 0.300 9 8 1 9 0 8 0 0 0 0 1 0 b2 0.0522 0.300 9 8 1 9 0 8 0 0 0 1 0 0 b2 0.0521 0.300 7 8 1 6 1 7 0 0 0 0 0 1 ... a5 0.0600 0.546 3 4 0 3 1 4 a5 0.0618 0.551 3 4 2 1 1 4 a5 0.0569 0.525 2 3 2 1 3 0 ... a1 0.0301 0.268 14 15 a1 0.0291 0.230 15 16 a1 0.0282 0.218 16 17 ... a0 0.0242 0.165 24 a0 0.0239 0.160 25 a0 0.0236 0.150 26 ... ----- -------- ------- ---- ---- --- --- --- --- --- --- --- --- --- --- : : H$_{2}$O - H$_{2}$ broad file: portion of the file (upper part); field specification (lower part).[]{data-label="tab:broad_H2O_example"} Field Fortran Format C format Description ------------------------- ---------------- ---------- ---------------------------------------------------------------------------------- code A2 $\%$2s Code identifying quantum number set following $J{^{\prime\prime}}$\* $\gamma_{\textrm{ref}}$ F2.4 %2.4f Lorentzian half-width at reference temperature and pressure in [cm$^{-1}$]{}/bar $n$ F2.3 %2.3f Temperature exponent $J{^{\prime\prime}}$ I7 $\%$7d Lower $J$-quantum number $J{^\prime}$ I7 $\%$7d Upper $J$-quantum number $K_a{^{\prime\prime}}$ I2 $\%$2d Lower rotational quantum number $K_c{^{\prime\prime}}$ I2 $\%$2d Lower rotational quantum number $K_a{^\prime}$ I2 $\%$2d Upper rotational quantum number $K_c{^\prime}$ I2 $\%$2d Upper rotational quantum number $v_1{^{\prime\prime}}$ I2 $\%$2d Lower vibrational quantum number $v_2{^{\prime\prime}}$ I2 $\%$2d Lower vibrational quantum number $v_3{^{\prime\prime}}$ I2 $\%$2d Lower vibrational quantum number $v_1{^\prime}$ I2 $\%$2d Upper vibrational quantum number $v_2{^\prime}$ I2 $\%$2d Upper vibrational quantum number $v_3{^\prime}$ I2 $\%$2d Upper vibrational quantum number : : H$_{2}$O - H$_{2}$ broad file: portion of the file (upper part); field specification (lower part).[]{data-label="tab:broad_H2O_example"} Code definitions: b2 = $J{^\prime}$, $K_a{^{\prime\prime}}$, $K_c{^{\prime\prime}}$, $K_a{^\prime}$, $K_c{^\prime}$, $v_1{^{\prime\prime}}$, $v_2{^{\prime\prime}}$, $v_3{^{\prime\prime}}$, $v_1{^\prime}$, $v_2{^\prime}$, $v_3{^\prime}$; a5 = $J{^\prime}$, $K_a{^{\prime\prime}}$, $K_c{^{\prime\prime}}$, $K_a{^\prime}$, $K_c{^\prime}$; a1 = $J{^\prime}$; a0 = no further quantum numbers. Dipoles file {#dipole_file} ------------ The dipole file mirrors the simple structure of the transitions file, see Table \[tab:dips\]. Two pointers, $i$ and $f$ point to rows in the `.states` file to identify the upper and lower states involved, as specified by the state ID variable, plus the full information characterising these states. $D$ gives the signed transition dipole (in Debye). In this form the combination of the `.states` and `.dipole` files can be used to simulate the effect of weak linear polarized electric fields (Stark) on the molecule, both static and time-dependent, where the polarization of the field is along $Z$. The $|D|$ value is related to the Einstein $A$ coefficient from the transitions file via a simple transformation: $$A_{if} = \frac{64\times 10^{-36} \pi^4}{3 h} |D|^2 (2J^{\prime\prime} +1) \tilde\nu_{if}^3$$ where $h$ is Planck’s constant, and $J^{\prime\prime}$ is the total angular momentum of the lower state. However the phase of the (potentially complex) $D$, which is important for calculating the Stark effect, is lost in this transformation. Therefore, in the dipole file we provide the actual value of $D$. The ExoMol format of the `.states`/`.dipole` files combination can then also be used for modelling higher-order field effects, including optical activity, polarization phenomena, nonlinear optical properties, and effects of strong fields including the effects of strong magnetic fields [@04Barron.method]. In this case the matrix elements of the corresponding properties will be stored as additional columns after the $D$ column. Table \[tab:dipeg\] gives an example of a dipole file for CaO. Field Fortran Format C Format Description ------- ---------------- ---------- ------------------------ $i$ `I12` `%12d` Upper state ID $f$ `I12` `%12d` Lower state ID $D$ `F12.8` `%12.8d` Dipole moment in Debye : Specification of the dipole file.[]{data-label="tab:dips"} Fortran format: `(I12,1x,I12,1x,F12.8)`\ |40Ca-16O\_\_VBATHY.dipole| $i$ $f$ $\mu_{if}$ ----- ----- ----------------- 33 1 0.52209235E+01 34 1 0.14271796E+00 35 1 0.50123653E-01 36 1 0.16344445E-01 37 1 -0.48627678E-03 38 1 -0.86965402E-03 39 1 0.31660861E-03 40 1 0.21892847E-03 41 1 -0.15794191E-03 : Extract from the dipole file for $^{40}$Ca$^{16}$O, .[]{data-label="tab:dipeg"} $i$: Upper state counting number; $f$: Lower state counting number; $\mu_{if}$: transition dipole in Debye. Cross section file ------------------ Absorption cross sections are provided by ExoMol in files with the extension `.cross`. The two columns of this file are described in Table \[tab:cross\] and are wavenumber ($\tilde{\nu}$, in $\mathrm{cm^{-1}}$) and cross section value ($\sigma$, in $\mathrm{cm^2 molec^{-1}}$). The two fields are separated by a single space. The stored cross sections may be temperature or temperature and pressure dependent as denoted by the header which includes the temperature in K and the pressure in bar. The absorption cross section is given as a sequence of average values, $\sigma_i$ within wavenumber bins, $\tilde{\nu}_i$ of specified width, $\Delta\tilde{\nu}$. That is, in general, cross section values are given by $$\sigma_i = \sum_j \frac{S_j}{\Delta\tilde{\nu}} \int_{\tilde{\nu}_i - \Delta\tilde{\nu}/2}^{\tilde{\nu}_i + \Delta\tilde{\nu}/2} f(\tilde{\nu}; \tilde{\nu}_{0,j}\cdots)\,\mathrm{d}\tilde{\nu},$$ where the sum is taken over individual absorption lines of strength $S_j$ and central wavenumber $\tilde{\nu}_{0,j}$, and $f(\tilde{\nu}; \tilde{\nu}_{0,j}\cdots)$ is a normalized lineshape function which is calculated using transition-dependent parameters, $(\tilde{\nu}_{0,j}\cdots)$. In the case that the bin width is much smaller than the line-width, $\Delta\tilde{\nu} \ll \mathrm{HWHM}$, this formula reduces to $$\sigma_i = \sum_j S_j f(\tilde{\nu}; \tilde{\nu}_{0,j}\cdots).$$ See Hill [[*et al*]{}]{} [@jt542] for further details. Table \[tab:cross\_example\] gives an example of a cross section file for water. Field Fortran Format C Format Description ----------------- ---------------- ---------- ----------------------------------------------------- $\tilde{\nu}_i$ `F12.6` `%12.6f` Central bin wavenumber, $\mathrm{cm^{-1}}$ $\sigma_i$ `ES14.8` `%14.8e` Absorption cross section, $\mathrm{cm^2molec^{-1}}$ : Specification of the `.cross` cross section file format[]{data-label="tab:cross"} Fortran format: `(F12.6,1x,ES14.8)`\ |1H2-16O\_\_BT2\_\_0-30000\_\_296K\_\_0bar\_\_0.01.cross| $\tilde{\nu}$/cm$^{-1}$ $\sigma$/$\mathrm{cm^2 molec^{-1}}$ ------------------------- ------------------------------------- 1525.530000 8.58106599E-36 1525.540000 4.67629889E-37 1525.550000 4.53342436E-36 1525.560000 6.55402974E-34 1525.570000 2.08219623E-31 1525.580000 1.90580250E-33 1525.590000 4.18283497E-32 1525.600000 1.30145956e-26 1525.610000 5.31710192E-27 1525.620000 7.96854826E-28 1525.630000 1.04806968E-18 1525.640000 7.80260131E-18 1525.650000 3.18257478E-22 1525.660000 1.27578728E-34 1525.670000 6.09403363E-30 1525.680000 1.50281502E-29 1525.690000 2.79935737E-29 : Extract from : Cross section function file for H$_2{}^{16}$O.[]{data-label="tab:cross_example"} $k$-coefficients file --------------------- $k$-coefficients are provided by ExoMol in files with the extension `.kcoef`. The format of this file is described in Table \[tab:kcoef\]. The $k$-coefficients are provided in $N_{\rm bins}$ wavelength bins spanning a specified wavelength range $\lambda_{\rm min}$ to $\lambda_{\rm max}$. The spacing between the bin centers is $\Delta\lambda$ while each bin is $\delta\lambda$ wide. Clearly, for the case that the bins do not overlap $\Delta\lambda = \delta\lambda$. Each bin contains $N_g$ values of $k$ corresponding to the $g$-ordinate values (which will be an increasing sequence of $k$-coefficients). Figure \[tab:kcoef\_example\] gives an example of a $k$-coefficient file for water. Field Fortran Format C Format Description --------------------- --------------------- ---------------- ------------------------------------------------------------------------- $P$ `F12.6` `%12.6f` Pressure in bar $T$ `F8.2` `%8.2f` Temperature in K $\lambda_{\rm min}$ `F12.6` `%12.6f` Minimum central wavelength of bins, in $\mu$m, $\lambda_{\rm max}$ `F12.6` `%12.6f` Maximum central wavelength of bins, in $\mu$m, $\Delta\lambda$ `F12.6` `%12.6f` Spacing of wavelength bin centres, in $\mu$m, $\delta\lambda$ `F12.6` `%12.6f` Bin widths, in $\mu$m, $N_{\rm bins}$ `I6` `%6d` No. of wavelength bins provided $N_g$ `I6` `%6d` No. of $g$-ordinate values at which $k$ is provided within each bin $g_{\rm ord}$ $N_g$`*(F10.8, 1X)` $N_g$`*%10.8f` $N_g$ $g$-ordinate values, $0\leq g\leq 1$ to be read on a single line. : Specification of the `.kcoef` cross section file format; the $k$-coefficients are in cm$^2$ molecule$^{-1}$.[]{data-label="tab:kcoef"} |1H2-16O\_\_BT2\_\_0-30000\_\_296K\_\_0.01.kcoef| ---------------------------------------------------------------------------------------- -- ` EXOMOL.def begin{figure} ` ` 0.005000 ` ` 240.00 ` ` 0.300000 ` ` 30.000000 ` ` 0.025000 ` ` 0.025000 ` ` 1189 ` ` 20 ` ` 0.00343570 0.01801404 0.04388279 ... 0.98198600 0.99656430 ` ` 0.3 6.41554300e-23 6.42944800e-23 6.60117000e-23 ... 6.32264600e-21 ` ` 0.3025 6.67216472e-23 6.68662592e-23 6.86521680e-23 ... 7.09400881e-21 ` ` ... ` ` 30.000000 1.47557489e-24 1.47877304e-24 1.51826910e-24 ... 1.45420858e-22 ` ---------------------------------------------------------------------------------------- -- : Extract from : $k$-coefficiant file for H$_2{}^{16}$O.[]{data-label="tab:kcoef_example"} Partition function file ----------------------- Temperature-dependent properties such as the partition function and the cooling function are provided by the partition function file with extension `.pf`. This file gives partition function, $Q$, and, if available ($C_{\rm avail}=1)$, cooling function, $W$, values as a function of temperature, $T$. Usually $T$ is given in steps of $T_{\rm Step}$ (usually 1 K) up to $T^Q_{\rm max}$. Table \[tab:pf\] gives the specification for the `.pf` file while Table \[tab:pfeg\] gives an example for the CaH molecule. ExoMol, like HITRAN [@HITRAN-A], follows the convention that the nuclear spin factors are derived from the full atomic nuclear spin degeneracies. Thus, for example, for H$_2{}^{17}$O, H has spin $\frac{1}{2}$ so has degeneracy $(2I_H+1) = 2$ and $^{17}$O has spin $\frac{5}{2}$ so $(2I_O+1)=6$. This means that para water has degenercy 1 and ortho water has degeneracy 18. This is in contrast to standard compilations of astronomical partition functions such as those of Irwin [@81Irwin.partfunc], Sauval and Tatum [@84SaTaxx.partfunc], and Barklem and Collet [@16BaCoxx.partfunc] who normalise to unit atomic nuclear spin degeneracy factors. Final results are independent of which convention is chosen provided they are used self-consistently. Field Fortran Format C Format Description -------- ---------------- ---------- ------------------------------------------------------------------- $T$ `F8.1` `%8.1d` Temperature in K $Q(T)$ `F15.4` `%15.4d` Partition function (dimensionless). $W(T)$ `ES12.4` `%12.4e` Cooling function in ergs s$^{-1}$ molecule$^{-1}$ (if available). : Specification of the `.pf` partition function file.[]{data-label="tab:pf"} Fortran format: `(F8.1,1x,F15.4,1x,ES12.4)` or `(F8.1,1x,F15.4)`\ |40Ca-1H\_\_Yadin.pf| $T$/K $Q(T)$ $W(T)$ ------- --------- ------------ 1.0 4.0001 1.5662E-24 2.0 4.0274 6.9307E-22 3.0 4.2080 5.1097E-21 4.0 4.5752 1.3731E-20 5.0 5.0664 2.6214E-20 6.0 5.6252 4.3743E-20 7.0 6.2198 6.7576E-20 8.0 6.8342 9.8328E-20 9.0 7.4602 1.3618E-19 10.0 8.0937 1.8114E-19 11.0 8.7323 1.0279E-18 12.0 9.3745 2.6263E-18 13.0 10.0193 5.0095E-18 14.0 10.6663 8.1986E-18 15.0 11.3148 1.2208E-17 16.0 11.9646 1.7050E-17 : Extract from the partition function file for $^{40}$Ca$^1$H, , including cooling function, $W(T)$.[]{data-label="tab:pfeg"} Utility Programs ================ As part of the database, ExoMol provides a number of sample utilities for manipulating data presented in the ExoMol format. This covers the most common applications of the line lists, such as computing absorption/emission cross sections [@jt542] or coefficients (‘stick’ intensities), partition functions, cooling functions and lifetimes. These utilities are available as Fortran 95 and Python programs combined into an <span style="font-variant:small-caps;">ExoCross</span> package. ExoCross allows users to utilise different line profiles to generate cross-sections, including the Voigt profile, using the line parameters compiled in the ExoMol database (file `.broad`) when available. Other standard line profiles include the temperature-dependent Doppler profile, rectangular-shape profiles (i.e. averaged cross-sections over the wavenumber bin), as well as general Gaussian and Lorentzian profiles with line-widths as free input parameters. The utilities can be obtained directly from the ExoMol website together with examples of input files. As mentioned above a Python utility, `extract__trans.py` is provided which reads the `.trans` in `.bz2` format without requiring it to be uncompressed. The molecular data produced by ExoMol is increasingly being used in a wide variety of applications. In many of these applications, pre-existing software packages utilise different data formats for their input files. The detailed specification of our format above should be sufficient to allow conversion to any required format using a suitable parser program. We have written a number of utility programs (in Python) that convert the ExoMol format to other common formats. These can be used to produce the full line list in different formats such as HITRAN, for which the Python utility `exomol2hitran.py` is available. Oscillator strengths can be generated by the utility exomol2gf.py. However, we should stress that the ExoMol format is much more compact than most other alternatives (notably HITRAN format), primarily due to the separation of information into the states and trans file. The size of the ExoMol line lists means that it can be prohibitive to store the full data in another format; for example, the methane 10to10 line list would require 14 TB to be stored in HITRAN format. It is therefore recommended that the utility programs provided be integrated within the larger program, such that the new format data is produced for a small number of lines at a time which are immediately utilised by the program, thus avoiding the storage of data. Website ======= The ExoMol website (www.exomol.com) is the main source of the ExoMol data. Data can be accessed two ways: (a) by first selecting a molecule and then a specific isotopologue-dataset name combination or (b) by selecting a particular data type, such as partition function. We note that cross sections are also being made available via the Virtual Atomic and Molecular Data Centre (VAMDC) [@jt481; @jt630]. This API takes the form of a “Table Access Protocol” service at `www.exomol.com/tap/sync`. The exomol.all file described in Section 5.2 is available at www.exomol.com/exomol.all and the data files comprising each dataset can be accessed at a URL of the form `www.exomol.com/db/<MolFormula>/<IsoSlug>/<DSName>/<Filename>`, for example `www.exomol.com/db/SiO/28Si-16O/EBJT/28Si-16O__EBJT.pf`. A separate API for machine-access of the ExoMol data in the format described in this paper will be available at `www.exomol.com/api` and documented on the website. This will allow the automatic retrieval of data files by the `HTTP GET` method. Conclusions =========== The ExoMol database presented here is a molecule-by-molecule set of comprehensive line lists for modelling spectra and other properties of hot gases. The choice of molecules is dictated by the need to model the atmospheres of exoplanets and other hot astronomical objects, but the spectroscopic data have much wider applications than this. We are still in the process of adding molecules to the database and are receptive to suggestions of other key species to include. In particular the discovery [@11LeOdKu.exo; @11RoDeDe.exo] and observation of spectra [@jt629] of very hot, rocky planets will undoubtedly lead to the consideration of novel species [@12ScLoFe.exo; @15ItIkMa.exo]. In this paper we have significantly expanded the flexibility and scope of the data model used by ExoMol. The original ExoMol format of a `.states` and `.trans` file has been already adopted as a possible input format by some programs, such as PGOPHER [@PGOPHER]. The extended format proposed here is both backward-compatible and designed to allow direct extraction of data from the database using an application programming interface (API). We believe that this will be helpful to scientists and relieve them of the need to manually deal with datasets containing many billions of spectral lines. Our new data structure allows the inclusion of important new molecular parameters including those to determine pressure broadening, state-resolved radiative lifetimes and Landé $g$-factors. Acknowledgments {#acknowledgments .unnumbered} =============== This work was supported by the ERC under the Advanced Investigator Project 267219. [271]{} natexlab\#1[\#1]{}\[2\][\#2]{} , () . , , , , () . , , , () . , , , () . , , () . , , , , , () . , , , () . , , , , () . , , , () . , , , , , , , () . , , , , , , , , () . , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , () . , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , () . , , , , , , () . , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , () . , et al., (). , , () . , , , , , , () . , , () . , , , , () . , () . , , , , , , , () . , , , , , () . , , , , , , () . , , , , , , () . , , , , , () . , , , , , , , () . , , , , () . , , , , , , , , , in: (Ed.), , volume of **, p. . . , , , in: , volume of **, , , pp. . , , , , , () . , , , () . , , , , () . , , , (). , , , , () . , , , , , , () . , , , , , () . , , , , () . , , (). , , , , , , () . , , , () . , , , () . , , , () . , , , , () . , , , () . , , , , , () . , , () . , , () . , , , , () . , , , , . , , , , , , , () . , , , , () . , , , () . , , () . , , , () . , , , , , , () . , , , () . , , , , , , , () . , , () . , , , , , , , () . , , , , () . , , , , , () . , , , () . , , , , , () . , , , , () . , , , , , , () . , , , , , , , , (). , , , , , , () . , , , () . , , , , () . , , , , () . , , , () . , , , () . , , , () . , , , , , () . , , , , , , () . , , , , , , , , , , () . , , , , , , () . , , , , () . , , , , () . , , , , , () . , , , , () . , , , , , , , () . , , () . , , , , , () . , () . , , , , , , , , , () . , , , , , , , , () . , , , , , , , () . , , () . , , , , , () . , , , , , , , , , , , , , () . , () . , , () . , , , , , , () . , , , , , , , () . , , , , , () . , , , , , , () . , , , , () . , , () . , , , , () . , , , , , () . , , , , , () . , , , () . , , , () . , , , , , () . , , , (). , , . , , , (). , , , , , , , , , , , , () . , , , , , , , , , () . , , () . , , () . , , , , , , , () . , , , , , () . , , , , , , , , , , , , () . , , , , , () . , , , () . , , , () . , , , , , (). , , , , , , , , , , , , , () . , , , () . , , () . , , , , , , , , , , , , , , , , , , , , , () . , , , , , , , , , , , , , () . , , , () . , , () . , , , , , , , , () . , , , , , , (). , , , , , , , , , () . , , , , , (). , , , , , , , , , , , , , , , , , , , , , () . , , , , , , , , , , , , , , , , , , , , () . , , , () . , , () . , , , , , () . , , , () . , , , , , , , , (). , , , , , () . , , , () . , , , , , , , () . , , , , , , , , , , () . , , , , , , , , () . , , () . , , , (). , , , () . , , , , (). , , , , , () . , , , , , () . , , , () . , , , , , () . , , , , , () . , , , (). , , , , (). , , , () . , , , , , , , , , () . , () . , () . , , , , () . , , () . , , () . , , , , () . , , , , () . , , , , , , , , , , , , , , , , , , , , () . , , , , , () . , , , , , , (). , , , , , , , , , , , , , , , , , , , , , , () . , () . , , , , , , , , , () . , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , () . , , , , () . , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , (). , , , , , , , , , () . , , , , , , , () . , , , , , , , () . , , , , , () . , , , , , () . , , , () . , , , , , () . , , () . , , , , , () . , , , () . , , , () . , , , () . , , , , () . , () . , , , , , , () . , , () . , , , , , , , () . , , , , , , , , , () . , , () . , , , , () . , , , , , , , () . , () . , , () . , () . , , , () . , , , , () . , () . , () . , , , , () . , , , , , () . , , () . , , , , () . , , , , , () . , , , , , () . , , , , () . , , , , , , , () . , , , , , , () . , , , , , () . , , , , () . , , , , () . , , , , , () . , , () . , , () . , , , , , () . , , , () . , , , , , () . , , , , , () . , , , () . , , , () . , , , , , () . , , , , () . , , , , , , , , , , , , , , , , () . , () . , , , , , , () . , , , , , , , , () . , , , , , , () . , , , , , , () . , , , , , , () . , , , () . , , , () . , , () . , () . , , , , () . , , , , , , , , , , () . , , , () . , , , , , , () . , , () . , , () . , , , () . , , , , () . , , , , () . , , , , () . , , , , , , () . , , () . , , , () . , , () . , , () . , , , , () . , , () . , , , , , () . , , , , () . , , , , () . , , , , , , () . , , , , , () . , , , , , . , , , , , () . , , , , , , . , , , . , () . , , () . , , (). , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , () . , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , (). , , , , , , , , , , , , , () . , , , , , , , () . , , , , , , , , , (). , , , () . , , , , , , () . , , . . [^1]: Corresponding author
{ "pile_set_name": "ArXiv" }
--- author: - 'Yoshihide [Tanaka]{}, Akira [Oguri]{} and Hiroumi [Ishii]{}' title: ' Perturbation Study of the Conductance through an Interacting Region Connected to Multi-Mode Leads ' --- \#1\#2 Introduction ============ Low-dimensional electron systems have been one of the current interests in the fields of the condensed matter physics and materials science. For instance, in some of the organic conductors, the electron correlation has been considered to be important to understand the physical properties. [@ishiguro] The Kondo effect in quantum dots [@NL; @GR; @Kawabata] has also been studied intensively from theoretical [@MWL1-2; @HDW2] and experimental [@RalphBuhrman; @Goldharber-Gordon; @Kouwenhoven] sides. When the average number of the electrons in a quantum dot is odd, the perfect transmission due to the Kondo resonance situated at the Fermi energy occurs at low temperatures. Recently, artificial molecules which are realized by arranging two or more quantum dots have also been studied. [@oosterkamp; @tokura] Theoretically, the crossover from the high-temperature Coulomb-blockade to low-temperature Fermi-liquid behaviors of the quantum dots has been studied using advanced numerical methods such as the numerical renormalization group [@Izumida1-3; @Izumida4-5] and quantum Monte Carlo methods. [@Sakai; @ao; @ao6] In a previous work, one of the authors has studied the conductance of small interacting systems connected to two single-mode leads, [@ao7; @ao9] and has calculated the conductance of a Hubbard chain of finite size $N$ using the order $U^2$ self-energy. The results obtained in the electron-hole symmetric case depend strongly on whether $N$ is even or odd. In the even cases, the conductance decreases with increasing $N$ showing a tendency toward a Mott-Hubbard insulator. On the other hand, in the odd cases the perfect transmission due to the Kondo resonance occurs.[@ao9] The purpose of the present work is to generalize the formulation to the multi-mode systems where the interacting system is connceted to noninteracting leads with a number of channels. As in the single-mode case, at $T=0$ the contributions of the vertex corrections on the dc conductance $g$ vanish. Then the conductance is determined by the value of the single-particle Green’s function at the Fermi energy $\omega=0$, and thus $g$ can be written in terms of the transmission coefficient of the free quasiparticles described by an effective Hamiltonian also in the multi-mode case. We apply the formulation to a two-dimensional Hubbard model of finite size consisting of $N_{\rm C}=N \times M$ sites, where $N$ and $M$ are the size in the $x$- and $y$-direction, respectively. Two noninteracting leads of $M$ channels are attached to this cluster along the $x$-direction. This system may be considered as a model for two-dimensional materials such as an array of quantum dots and a carbon nanotube. For instance, for $M=2$ it may also be considered as a model for a ladder system of nanometer size. Since there is no translational symmetry in the systems we are considering, the self-energy has $N_{\rm C} \times N_{\rm C}$ matrix elements. We calculate all these elements within the second order perturbation expansion in $U$ in the electron-hole symmetric case, and determine the parameters of the effective Hamiltonian up to terms of order $U^2$. The results of the conductance show maximums at finite $U$ for some values of $t_y/t_x$, where $t_x$ ($t_y$) is the nearest-neighbor transfer integral in the $x$-direction ($y$-direction). This is caused by resonances occurring in some of the subbands. These behaviors can also be understood through the $U$ dependence of the eigenvalues of the effective Hamiltonian. We note that the second-order perturbation theory has been applied by a number of groups to study transport through the single Anderson impurity[@HDW2; @YMF; @MiiMakoshi; @TakagiSaso] and systems consisting of a number of sites.[@PFA; @KKNO] Also, the band calculations has been applied to obtain the conductance of nanometer systems. [@Lang] Our approach can, in principle, be extended to realistic systems by evaluating the order $U^2$ self-energy using the orbitals which are obtained with the band calculations. In §2, we describe the outline of the general formulation and introduce the effective Hamiltonian. In §3, we apply the method to a two-dimensional Hubbard model, and present the results of the conductance. Summary is given in §4. In Appendix, we provide the derivation of the expressions of the conductance and the total charge displacement. Formulation {#sec:Model} =========== In this section, we give the outline of a general formulation of the conductance through interacting systems connected to Fermi-liquid reservoirs with a number of channels. Our formulation is applicable for various systems which have the time reversal symmetry, i.e., eqs. (\[eq:g\_multi\]) and (\[eq:Friedel\_multi\]) hold for a wide range of the systems described by the Hamiltonian eq. (\[eq:H\]). We provide the details of the derivations in Appendix, and demonstrate the application to a Hubbard model in the next section. We start with a system which consists of three regions as illustrated in Fig. \[fig:multi\]; a finite interacting region (${\rm C}$) situated in the middle, and two noninteracting reservoirs on the left (${\rm L}$) and right (${\rm R}$). The central region contains $N_{\rm C}$ resonant levels, and the inter-electron interaction $U_{j_4 j_3; j_2 j_1}$ is switched on in the central region. Each of the two reservoirs is infinitely large and has a continuous energy spectrum. The central region and two reservoirs are connected by $M_{\rm L}$ and $M_{\rm R}$ channels which are described by the mixing matrix elements $v_{{\rm L},m}^{\phantom{\dagger}}$ and $v_{{\rm R},m}^{\phantom{\dagger}}$, respectively. The complete Hamiltonian is given by $$\begin{aligned} {\cal H} \ &=& \ {\cal H}_{\rm L} + {\cal H}_{\rm R} + {\cal H}_{\rm C}^0 + {\cal H}_{\rm C}^{\rm int} + {\cal H}_{\rm mix} \label{eq:H} \;, \\ {\cal H}_{\rm L} &=& \sum_{ij\in {\rm L}} \sum_{\sigma} \left(\,-t_{ij}^{\rm L} - \mu\, \delta_{ij} \,\right) c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{j \sigma} \;, \label{eq:H_L} \\ {\cal H}_{\rm R} &=& \sum_{ij\in {\rm R}} \sum_{\sigma} \left(\,-t_{ij}^{\rm R} - \mu\, \delta_{ij} \,\right) c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{j \sigma} \;, \label{eq:H_R} \\ {\cal H}_{\rm C}^{0} &=& \sum_{ij\in {\rm C}} \sum_{\sigma} \left(\,-t_{ij}^{\rm C} - \mu\, \delta_{ij} \,\right) c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{j \sigma} \;, \\ {\cal H}_{\rm C}^{\rm int} &=& {1 \over 2} \sum_{\{j\} \in {\rm C}}\sum_{\sigma \sigma'} U_{j_4 j_3; j_2 j_1}\, c^{\dagger}_{j_4 \sigma} c^{\dagger}_{j_3 \sigma'} c^{\phantom{\dagger}}_{j_2 \sigma'} c^{\phantom{\dagger}}_{j_1 \sigma} \;, \label{eq:H_int} \\ {\cal H}_{\rm mix} &=& - \sum_{m=1}^{M_{\rm L}} \sum_{\sigma} v_{{\rm L},m}^{\phantom{\dagger}} \left(\, c^{\dagger}_{{\cal I}_m \sigma}\, c^{\phantom{\dagger}}_{{\cal L}_m \sigma} + c^{\dagger}_{{\cal L}_m \sigma}\, c^{\phantom{\dagger}}_{{\cal I}_m \sigma} \, \right) \nonumber \\ & & - \sum_{m=1}^{M_{\rm R}} \sum_{\sigma} v_{{\rm R},m}^{\phantom{\dagger}} \left(\, c^{\dagger}_{{\cal R}_m \sigma}\, c^{\phantom{\dagger}}_{{\cal N}_m \sigma} + c^{\dagger}_{{\cal N}_m\, \sigma} c^{\phantom{\dagger}}_{{\cal R}_m \sigma} \, \right) . \label{eq:H_mix_multi}\end{aligned}$$ Here $c^{\dagger}_{j \sigma}$ ($c^{\phantom{\dagger}}_{j \sigma}$) creates (destroys) an electron with spin $\sigma$ at site $j$, and $\mu$ is the chemical potential. $t_{ij}^{\rm L}$, $t_{ij}^{\rm R}$, and $t_{ij}^{\rm C}$ are the intra-region hopping matrix elements in each of the regions ${\rm L}$, ${\rm R}$, and ${\rm C}$, respectively. The labels $1$, $2$, $\ldots$, $N_{\rm C}$ are assigned to the sites in the central region. In eq. (\[eq:H\_mix\_multi\]), ${\cal I}_m$ (${\cal L}_m$) is the label assigned to the $m$-th site at the sample side (lead side) of the interface on the left. ${\cal N}_m$ (${\cal R}_m$) is the label assigned to the $m$-th site at the sample side (lead side) of the interface on the right. Note that the number of the channels is less than the number of the interacting sites, i.e, $M_{\rm L} \leq N_{\rm C}$ and $M_{\rm R}\leq N_{\rm C}$. For instance, in the case of $M_{\rm L}=M_{\rm R}=N_{\rm C}$, all the sites in the central region are connected to both of the reservoirs, i.e., ${\cal I}_m = {\cal N}_m$ for $m=1,\,2,\,\ldots,\, N_{\rm C}$. We assume that all the hopping matrix elements are real, and the interaction has the time reversal symmetry: $U_{4 3; 2 1}$ is real and $U_{4 3; 2 1}=U_{3 4; 1 2}=U_{1 2; 3 4 }=U_{4 2; 3 1}=U_{1 3; 2 4}$. We will be using units $\hbar=1$ unless otherwise noted. The single-particle Green’s function is defined by $$G_{jj'}({\rm i}\varepsilon_n) = - \int_0^{\beta} \! {\rm d}\tau \left \langle T_{\tau} \, c^{\phantom{\dagger}}_{j \sigma} (\tau) \, c^{\dagger}_{j' \sigma} (0) \right \rangle \, {\rm e}^{{\rm i}\, \varepsilon_n \tau} . \label{eq:G_Matsubara}$$ Here $\beta= 1/T$, $\varepsilon_n = (2n+1)\pi/\beta$, $c_{j \sigma}(\tau) = {\rm e}^{\tau {\cal H}} c_{j \sigma} {\rm e}^{- \tau {\cal H}}$, and $\langle \cdots \rangle$ denotes the thermal average $\mbox{Tr} \left[ \, {\rm e}^{-\beta {\cal H} }\, {\cdots} \,\right]/\mbox{Tr} \, {\rm e}^{-\beta {\cal H} }$. The spin index has been omitted from the left-hand side of eq. (\[eq:G\_Matsubara\]) assuming the expectation value to be independent of whether spin is up or down. Since the interaction is switched on only for the electrons in the central region, the Dyson equation can be written as $$G_{ij}(z) = G^0_{ij}(z) + \sum_{i'j' \in {\rm C}}\,G^0_{ii'}(z)\, \Sigma_{i'j'}(z) \, G_{j'j}(z) \;. \label{eq:Dyson}$$ Here $G^0_{ij}(z)$ is the unperturbed Green’s function corresponding to the noninteracting Hamiltonian ${\cal H}^0 \equiv {\cal H}_{\rm L} + {\cal H}_{\rm R} + {\cal H}_{\rm C}^0 + {\cal H}_{\rm mix}$. The summations with respect to $i'$ and $j'$ run over the sites in the central region, and $\Sigma_{i'j'}(z)$ is the self-energy due to the interaction ${\cal H}_{\rm C}^{\rm int}$. Because of the time reversal symmetry of ${\cal H}$, these functions are symmetric against the interchange of the site indices: $G_{ij}(z) = G_{ji}(z)$ and $\Sigma_{ij}(z) = \Sigma_{ji}(z)$. Note that at $T=0$ the single-particle excitation at the Fermi energy $z= {\rm i}0^+$ does not decay, i.e., $\mbox{Im}\,\Sigma_{ij}^+ (0)=0$.[@LangerAmbegaokar] In what follows, we will treat $z$ as a complex variable, and use the symbol $+$ ($-$) in the superscript as a label for the retarded (advanced) function: $\Sigma_{ij}^{\pm}(\varepsilon) \equiv \Sigma_{ij}(\varepsilon \pm {\rm i}0^+)$. As in the single-mode case, the dc conductance at $T=0$ can be expressed in terms of the single-particle Green’s function at $\omega=0$ \[see Appendix \[sec:CONDUCTANCE\]\]; $$g = {2 e^2 \over h} \, \mbox{Tr} \left[\,4\, \mbox{\boldmath $\Gamma$}_{\rm R} (0) \, \mbox{\boldmath $G$}_{{\cal N}{\cal I}}^{+}(0)\, \mbox{\boldmath $\Gamma$}_{\rm L}(0) \, \mbox{\boldmath $G$}_{{\cal I}{\cal N}}^{-}(0) \,\right] . \label{eq:g_multi}$$ Here $\mbox{\boldmath $\Gamma$}_{\rm L}(\omega)$ and $\mbox{\boldmath $\Gamma$}_{\rm R}(\omega)$ are $M_{\rm L} \times M_{\rm L}$ and $M_{\rm R} \times M_{\rm R}$ matrices, respectively. These two matrices are caused by the coupling with the left (${\rm L}$) and right (${\rm R}$) leads, and the elements are given by $\Gamma_{\alpha;mm'} (\omega) = - \mbox{Im} \left[ v_{\alpha,m}\, F_{\alpha,mm'}^{+}(\omega) \,v_{\alpha,m'} \right]$ with $F_{\alpha,mm'}(z)$ being the Green’s functions at the interface of the isolated lead ($\alpha ={\rm L},{\rm R}$). In eq. (\[eq:g\_multi\]), $\mbox{\boldmath $G$}_{{\cal N}{\cal I}}^+$ and $\mbox{\boldmath $G$}_{{\cal I}{\cal N}}^-$ are $M_{\rm R} \times M_{\rm L}$ and $M_{\rm L} \times M_{\rm R}$ matrices the elements of which are given by $G_{{\cal N}_{l}{\cal I}_{m}}^+$ and $G_{{\cal I}_{l}{\cal N}_{m}}^-$, respectively. Another quantity which can be related to the scattering coefficients is the displacement of the total charge defined by [@LangerAmbegaokar; @Langreth] $$\begin{aligned} \Delta N_{\rm tot} \ &=& \ \sum_{i\in {\rm C}} \sum_{\sigma} \langle c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{i \sigma} \rangle \nonumber\\ & & + \sum_{i\in {\rm L}} \sum_{\sigma} \left[\, \langle c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{i \sigma} \rangle - \langle c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{i \sigma} \rangle_{\rm L}^{\phantom{0}} \,\right] \nonumber\\ & & + \sum_{i\in {\rm R}} \sum_{\sigma} \left[\, \langle c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{i \sigma} \rangle - \langle c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{i \sigma} \rangle_{\rm R}^{\phantom{0}} \,\right]. \label{eq:dn_def}\end{aligned}$$ Here $\,\langle \cdots \rangle_{\rm L}^{\phantom{0}}$ and $\,\langle \cdots \rangle_{\rm R}^{\phantom{0}}$ denote the ground-state average of isolated leads described by ${\cal H}_{\rm L}$ and ${\cal H}_{\rm R}$, respectively. At $T=0$, following the derivation of the Friedel sum rule by Langer and Ambegaokar,[@LangerAmbegaokar] $\Delta N_{\rm tot}$ can be written in terms of a $(M_{\rm L} +M_{\rm R}) \times (M_{\rm L} +M_{\rm R})$ matrix $\mbox{\boldmath $S$}$, $$\begin{aligned} & &\Delta N_{\rm tot} = {1 \over \pi {\rm i}}\, \log [\, \det \mbox{\boldmath $S$} \,] \;, \label{eq:Friedel_multi} \\ \nonumber \\ & &\mbox{\boldmath $S$} \, = \, \mbox{\boldmath $1$} - \, 2 {\rm i} \left [ \matrix { \mbox{\boldmath $\Gamma$}_{\rm L}(0) & \mbox{\boldmath $0$} \cr \mbox{\boldmath $0$} & \mbox{\boldmath $\Gamma$}_{\rm R}(0) \cr } \right ] \left [ \matrix { \mbox{\boldmath $G$}_{{\cal I}{\cal I}}^{+}(0) & \mbox{\boldmath $G$}_{{\cal I}{\cal N}}^{+}(0) \cr \mbox{\boldmath $G$}_{{\cal N}{\cal I}}^{+}(0) & \mbox{\boldmath $G$}_{{\cal N}{\cal N}}^{+}(0) \cr } \right ] . \nonumber \\ \label{eq:S_multi}\end{aligned}$$ Here $\mbox{\boldmath $G$}_{{\cal I}{\cal I}}^+$ and $\mbox{\boldmath $G$}_{{\cal N}{\cal N}}^+$ are $M_{\rm L} \times M_{\rm L}$ and $M_{\rm R} \times M_{\rm R}$ matrices, and the elements are given by $G_{{\cal I}_{l}{\cal I}_{m}}^+$ and $G_{{\cal N}_{l}{\cal N}_{m}}^+$, respectively. The outline of the Friedel sum rule in the single-mode case is given in Appendix \[sec:Friedel\]. The extension to multi-mode case eq. (\[eq:Friedel\_multi\]) is straightforward because in the case of the Friedel sum rule we do not have to take into account the contributions vertex corrections. For convenience, in eq. (\[eq:S\_multi\]) we have assumed that there is no common element in the sets ${\cal I}$ and ${\cal N}$. This is not an important assumption, and the expression without the assumption can be obtained in the similar way. Note that eq. (\[eq:Friedel\_multi\]) is also written in terms of the Green’s function at $T=0$ and $\omega=0$. Thus, due to the property $\mbox{Im}\,\Sigma_{ij}^+ (0)=0$, the values of $g$ and $\Delta N_{\rm tot}$ at $T=0$ can be expressed in terms of the transmission and reflection coefficients defined with respect to a one-particle Hamiltonian[@ao9] $$\widetilde{\cal H}_{\rm qp} = {\cal H}_{\rm L} + {\cal H}_{\rm R} + {\cal H}_{\rm C}^{\rm eff} + {\cal H}_{\rm mix}\;. \label{eq:H_eff}$$ Here ${\cal H}_{\rm C}^{\rm eff} = \sum_{ij\in {\rm C},\sigma} \left(\,-\widetilde{t}_{ij}^{\rm C} - \mu\, \delta_{ij} \,\right) c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{j \sigma}$ with $-\widetilde{t}_{ij}^{\rm C} = -t_{ij}^{\rm C} + \mbox{Re}\, \Sigma_{ij}^+ (0)$. This effective Hamiltonian describes free quasiparticles of the local Fermi-liquid. As shown in the next section, the eigenvalues of the partial Hamiltonian ${\cal H}_{\rm C}^{\rm eff}$ have important information about the ground-state properties. Application to a Two-dimensional Hubbard Model {#sec:Hubbard2D} ============================================== In this section, we apply the formulation to a two-dimensional Hubbard model connected to reservoirs and calculate the conductance using the order $U^2$ self-energy in the electron-hole symmetric case. Model ----- The schematic picture of the model is illustrated in Fig. \[fig:2dHubbard\]. The system size in the $y$-direction, $M$, is finite. In the $x$-direction, the central region consists of $N$ columns, and two noninteracting leads are connected at $x=1$ and $x=N$. Thus, the total number of the interacting sites is $N_{\rm C} = N\times M$. The parameters of the Hamiltonian eq. (\[eq:H\]) are specified as follows. The mixing matrix element is assumed to be uniform in the $y$-direction: $v_{{\rm L},m}^{\phantom{0}} =v_{\rm L}^{\phantom{0}}$ and $v_{{\rm R},m}^{\phantom{0}} = v_{\rm R}^{\phantom{0}}$ for $m=1,2,\ldots,M$. We assume that the off diagonal part of $t_{ij}^{\rm C}$ describes the nearest-neighbor hopping; $t_x$ and $t_y$ in the $x$- and $y$-directions, respectively. Along the $y$-direction, we use the periodic boundary condition. Furthermore, we assume $U_{j_4 j_3; j_2 j_1}$ to be an onsite repulsion $U$, and concentrate on the electron-hole symmetric case taking the parameters to be $\mu =0$ and $\epsilon_d + U/2 =0$, where $\epsilon_d$ is the onsite energy of the interacting sites $-t_{ii}^{\rm C}=\epsilon_d$. Then the Dyson equation can be written in a $N_{\rm C}\times N_{\rm C}$ matrix form: $ \left\{ \widehat{\mbox{\boldmath ${\cal G}$}}(z) \right\}^{-1} = \left\{ \widehat{\mbox{\boldmath ${\cal G}$}}^0(z) \right\}^{-1} - \widehat{\mbox{\boldmath $\Sigma$}}(z)$ with $$\begin{aligned} \left\{ \widehat{\mbox{\boldmath ${\cal G}$}}^0(z) \right\}^{-1} &=& z \, \widehat{\mbox{\boldmath $1$}} - \widehat{\mbox{\boldmath ${\cal H}$}}_{\rm C}^0 - \widehat{\mbox{\boldmath ${\cal V}$}}_{\rm mix}(z) \;. \label{eq:81}\end{aligned}$$ Here $\widehat{\mbox{\boldmath $1$}}$ is the $N_{\rm C} \times N_{\rm C}$ unit matrix, and the matrices can be written in the partitioned forms; $$\begin{aligned} \widehat{\mbox{\boldmath ${\cal H}$}}_{\rm C}^0 &=& \left[ \matrix{ \mbox{\boldmath $h$}_y^0& -t_x \mbox{\boldmath $1$} & & \mbox{\boldmath $0$} \cr -t_x \mbox{\boldmath $1$} & \ddots & \ddots & \cr & \ddots & \ddots & -t_x \mbox{\boldmath $1$} \cr \mbox{\boldmath $0$} & & -t_x \mbox{\boldmath $1$} & \mbox{\boldmath $h$}_y^0 \cr } \right], \label{eq:85} \\ \widehat{\mbox{\boldmath ${\cal V}$}}_{\rm mix}(z) &=& \left[ \matrix{ v_{\rm L}^2 \mbox{\boldmath $F$}_{\rm L}(z) & & & & \cr & & & & \cr & & \mbox{\boldmath $0$} & & \cr & & & & \cr & & & & v_{\rm R}^2 \mbox{\boldmath $F$}_{\rm R}(z) \cr } \right], \label{eq:87} \\ \widehat{\mbox{\boldmath $\Sigma$}}(z) &=& \left[ \matrix{ \mbox{\boldmath $\Sigma$}_{11}(z) & \mbox{\boldmath $\Sigma$}_{12}(z) & \ldots & \mbox{\boldmath $\Sigma$}_{1N}(z) \cr \mbox{\boldmath $\Sigma$}_{21}(z) & \ddots & \ddots & \vdots \cr \vdots & \ddots & \ddots & \vdots \cr \mbox{\boldmath $\Sigma$}_{N1}(z) & \ldots & \ldots & \mbox{\boldmath $\Sigma$}_{NN}(z) \cr } \right], \label{eq:86}\end{aligned}$$ where $\mbox{\boldmath $1$}$ is the $M\times M$ unit matrix, and $\mbox{\boldmath $h$}_y^0$ is the $M\times M$ hopping matrix in the $y$-direction; $$\mbox{\boldmath $h$}_y^0 \,= \left [ \matrix{ 0 & -t_y & & & & -t_y \cr -t_y & 0 & -t_y & & \mbox{\boldmath $0$} & \cr & -t_y & 0 & \ddots & & \cr & & \ddots & \ddots & \ddots & \cr & \mbox{\boldmath $0$} & & \ddots & \ddots & -t_y \cr -t_y & & & & -t_y & 0 \cr } \right]. \label{eq:59}$$ We note that the Hartree-Fock term of the self-energy is already included in the unperturbed Green’s function $\widehat{\mbox{\boldmath ${\cal G}$}}^0(z)$ defined by eq. (\[eq:81\]). Therefore, $\widehat{\mbox{\boldmath $\Sigma$}}(z)$ is the self-energy correction beyond the mean-field theory, which is described by the many-body perturbation theory with respect to $ {\cal H}_{\rm C}^{\rm int} = U \sum_{i=1}^{N_{\rm C}} \left[\, n_{i \uparrow}\, n_{i \downarrow} - ( n_{i \uparrow} + n_{i \downarrow} )/2 \,\right] $ where $n_{i \sigma} = c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{i \sigma}$. We note that the mixing with the noninteracting leads is included in the unperturbed Hamiltonian, and thus $\widehat{\mbox{\boldmath $\Sigma$}}(z)$ depends on $v_{\rm L}$ and $v_{\rm R}$ through $\widehat{\mbox{\boldmath ${\cal G}$}}^0(z)$. In eq. (\[eq:86\]) the partitioned element $\mbox{\boldmath $\Sigma$}_{ll'}(z)$ is a $M \times M$ matrix the $(m,m')$ element of which corresponds to the self-energy correction between the sites located at $\mbox{\boldmath $r$}=(l, m)$ and $\mbox{\boldmath $r$}'=(l', m')$, where $l$ and $m$ correspond to the $x$ and $y$ coordinates, respectively. The local Green’s functions at the interfaces of the isolated leads, $\mbox{\boldmath $F$}_{\rm L}(z)$ and $\mbox{\boldmath $F$}_{\rm R}(z)$ in eq. (\[eq:87\]), depend on the excitation spectrum of the leads, i.e., ${\cal H}_{\rm L}$ and ${\cal H}_{\rm R}$. We concentrate on the case $\mbox{\boldmath $F$}_{\rm L}=\mbox{\boldmath $F$}_{\rm R}$ ($\equiv \mbox{\boldmath $F$}$) assuming the same excitation spectrum for the left and right leads. Then the matrices of the level width are given by $\mbox{\boldmath $\Gamma$}_{\rm L}(\omega) = -v_{\rm L}^2\, \mbox{Im}\, \mbox{\boldmath $F$}(\omega+{\rm i} 0^+)$ and $\mbox{\boldmath $\Gamma$}_{\rm R}(\omega) = -v_{\rm R}^2 \,\mbox{Im}\,\mbox{\boldmath $F$}(\omega+{\rm i} 0^+)$. For the function $\mbox{\boldmath $F$}(z)$, we consider two types models: I) semi-infinite tight-binding leads, and II) leads of a constant density of states. A schematic picture of the type I lead is illustrated in Fig. \[fig:2dHubbard\_2\]: the hopping matrix element is given by the nearest neighbor one, $t_x$ and $t_y$, as that in the central region. Therefore the Green’s function for the type I lead $\mbox{\boldmath $F$}^{\rm I}$ satisfies a $M\times M$ matrix equation $\mbox{\boldmath $F$}^{\rm I}(z) = \left[\, z \mbox{\boldmath $1$} - \mbox{\boldmath $h$}_y^0 - t_x^2\, \mbox{\boldmath $F$}^{\rm I}(z) \,\right]^{-1}$. In the case of the type II leads, we assume that the local density of states at the interfaces $\rho$ is a constant and the bandwidth is infinity. Then the corresponding retarded Green’s function becomes pure imaginary independent of the frequency $\omega$. Specifically, we consider a simple diagonal matrix of the form $\mbox{\boldmath $F$}^{\rm II}(\omega+{\rm i} 0^+) = -{\rm i}\,\pi \rho \,\mbox{\boldmath $1$}$. Thus for the type II leads, the effects of the mixing are parametrized by the constant $\Gamma_{\alpha} = \pi \rho \, v^2_{\alpha}$ for $\alpha ={\rm L}$ and ${\rm R}$. The subband structure of the system is determined by the eigenstates of eq. (\[eq:59\]): $\mbox{\boldmath $h$}_y^0 \, \mbox{\boldmath $\chi$}_n = \epsilon_n \, \mbox{\boldmath $\chi$}_n$ for $n=1,2,\ldots, M$. Due to the translational symmetry in the $y$-direction, the self-energy $\mbox{\boldmath $\Sigma$}_{ll'}(z)$ and the $M \times M$ matrix Green’s function $\mbox{\boldmath $G$}_{ll'}(z)$, which are the $(l,l')$ partitioned element of $\widehat{\mbox{\boldmath ${\cal G}$}}(z)$, can be diagonalized using the eigenstates; $$\begin{aligned} \mbox{\boldmath $\Sigma$}_{ll'}(z) &=& \sum_{n=1}^M \mbox{\boldmath $\chi$}_n \, \Sigma_{ll'}^{(n)}(z) \, \mbox{\boldmath $\chi$}_n^{\dagger} \;, \label{eq:self_mode} \\ \mbox{\boldmath $G$}_{ll'}(z) &=& \sum_{n=1}^M \mbox{\boldmath $\chi$}_n \, G_{ll'}^{(n)}(z) \, \mbox{\boldmath $\chi$}_n^{\dagger} \;. \label{eq:Green_mode}\end{aligned}$$ Therefore, the conductance eq. (\[eq:g\_multi\]) can be decomposed into the sum of the contributions of the subbands: $$g \, = \, {2 e^2 \over h} \, \sum_{n=1}^M 4\, \Gamma_{\rm R}^{(n)} (0) \, G_{N1}^{(n)+}(0)\, \Gamma_{\rm L}^{(n)}(0) \, G_{1N}^{(n)-}(0) \;. \label{eq:g_multi_mode}$$ Here $\Gamma_{\alpha}^{(n)}(0)$ is defined by $\mbox{\boldmath $\Gamma$}_{\alpha}(z) = \sum_{n=1}^M \mbox{\boldmath $\chi$}_n \, \Gamma_{\alpha}^{(n)}(0) \, \mbox{\boldmath $\chi$}_n^{\dagger} $ for $\alpha={\rm L},{\rm R}$. Similarly, the Friedel sum rule eq. (\[eq:Friedel\_multi\]) can be rewritten as $$\begin{aligned} & &\Delta N_{\rm tot} = {2 \over \pi} \sum_{n=1}^{M} {1 \over 2 {\rm i}} \,\log \left[ \det \mbox{\boldmath $S$}^{(n)} \right] \;, \label{eq:Friedel_mode} \\ \nonumber \\ & &\mbox{\boldmath $S$}^{(n)} = \left [ \, \matrix { 1 & 0 \cr 0 & 1 \cr } \, \right ] \nonumber \\ & & \qquad - \, 2 {\rm i} \left [ \matrix { \Gamma_{\rm L}^{(n)}(0) & 0 \cr 0 & \Gamma_{\rm R}^{(n)}(0) \cr } \right ] \left [ \matrix { G_{11}^{(n)+}(0) & G_{1N}^{(n)+}(0) \cr G_{N1}^{(n)+}(0) & G_{NN}^{(n)+}(0) \cr } \right ] . \nonumber \\ \label{eq:S_mode}\end{aligned}$$ Here $\delta^{(n)} \equiv 1/(2{\rm i})\, \log \left[ \det\mbox{\boldmath $S$}^{(n)} \right]$ corresponds to the phase shift of the $n$-th subband, and the charge displacement in each subband is given by $\Delta N_{\rm tot}^{(n)} = 2\, \delta^{(n)}/\pi$. As mentioned, at $T=0$ the conductance is determined by the value of the Green’s function at $\omega=0$, and $\mbox{Im}\,\widehat{\mbox{\boldmath $\Sigma$}}^+(0) = \widehat{\mbox{\boldmath $0$}}$ due to the Fermi-liquid property. Thus the effective Hamiltonian $\widehat{\mbox{\boldmath ${\cal H}$}}_{\rm C}^{\rm eff} =\widehat{\mbox{\boldmath ${\cal H}$}}_{\rm C}^0 + \mbox{Re}\, \widehat{\mbox{\boldmath $\Sigma$}}^+(0)$, which was introduced in the previous section, can be used to calculate the conductance and the displacement of the total charge.[@ao9] We now consider the $M\times M$ matrix $\mbox{\boldmath ${\cal H}$}_{{\rm C};ll'}^{\rm eff}$, which is the partitioned element of $\widehat{\mbox{\boldmath ${\cal H}$}}_{\rm C}^{\rm eff}$. The matrix $\mbox{\boldmath ${\cal H}$}_{{\rm C};ll'}^{\rm eff}$ can also be diagonalized as $$\begin{aligned} \mbox{\boldmath ${\cal H}$}_{{\rm C};ll'}^{\rm eff} \, &=& \, - \sum_{n=1}^M \mbox{\boldmath $\chi$}_n \, \widetilde{t}^{(n)}_{ll'} \, \mbox{\boldmath $\chi$}_n^{\dagger} \;, \label{eq:H_eff_part} \\ -\widetilde{t}^{(n)}_{ll'} \, &=& \, \epsilon_n \, \delta_{ll'} \, -t_x \left[\,\delta_{l,l'+1} \, + \, \delta_{l+1,l'} \,\right] + \mbox{Re}\,\Sigma_{ll'}^{(n)+}(0) \;, \nonumber \\ \label{eq:25}\end{aligned}$$ where $1\leq l, l' \leq N$. Thus each of the modes can be mapped onto a tight-biding model in one-dimension with the renormalized hopping matrix element $\widetilde{t}^{(n)}_{ll'}$. As it will be seen later, the behavior of eigenvalues of the $N \times N$ effective Hamiltonian defined by $\widetilde{\mbox{\boldmath ${\cal H}$}}_{\rm C}^{(n)} = \{ -\widetilde{t}^{(n)}_{ll'} \}$ is related to the transport properties. We calculate the conductance using the perturbation expansion with respect to $U$. In the electron-hole symmetric case, the order $U^2$ self-energy can be described by the diagram shown in Fig. \[fig:diagramSelf\]; $$\begin{aligned} & &\Sigma_{jj'}^+(0) \, = \nonumber \\ & & \quad - U^2 \int_{-\infty}^{\infty} \! \int_{-\infty}^{\infty} \frac{{\rm d}\varepsilon\, {\rm d}\varepsilon'}{(2\pi)^2} \, G^{0}_{jj'}({\rm i} \varepsilon) \, G^{0}_{jj'}({\rm i} \varepsilon') \, G^{0}_{j'j}({\rm i} \varepsilon + {\rm i} \varepsilon') \;, \nonumber \\ \label{eq:Self_2v} \end{aligned}$$ where $1\leq j, j' \leq N_{\rm C}$. The explicit form of the unperturbed Green’s function $G^{0}_{jj'}({\rm i} \varepsilon)$ can be obtained by taking the inverse of eq. (\[eq:81\]). Since ${\cal H}_{\rm mix}$ is included in the unperturbed part, $G^{0}_{jj'}$ and $\Sigma_{jj'}^+$ depend on the mixing matrix elements $v_{\rm L}^{\phantom{0}}$ and $v_{\rm R}^{\phantom{0}}$. Note that the retarded function at $\omega=0$ can be obtained from the Matsubara function, i.e., $\Sigma_{jj'}^+(0) = \Sigma_{jj'}({\rm i}\varepsilon)|_{\varepsilon \to 0^+}$. The imaginary part of eq. (\[eq:Self\_2v\]) vanishes, i.e., $\mbox{Im} \Sigma_{jj'}^+(0)=0$ owing to the Fermi-liquid property. Furthermore, due to the electron-hole symmetry, $\mbox{Re}\, \Sigma_{jj'}^+(0) = 0$ if $j$ and $j'$ belong to the same sublattice. We calculate all the nonzero elements of $\mbox{Re}\, \Sigma_{jj'}^+(0)$ carrying out the integrations in eq. (\[eq:Self\_2v\]) numerically. Then the full Green’s function $\mbox{\boldmath $G$}_{N1}^{+}(0)$ is evaluated substituting the order $U^2$ self-energy into the Dyson equation. In what follows we will assume the inversion symmetry $v_{\rm L}^{\phantom{0}} =v_{\rm R}^{\phantom{0}}$ ($\equiv v$), and take the parameter to be $v/t_x= 0.9$ in the case of the type I leads. We have actually done some calculations also for $v/t_x= 0.7$, but the results are similar to those for $v/t_x= 0.9$ qualitatively. Correspondingly, in the case of the type II leads we set $\Gamma_{\rm L} = \Gamma_{\rm R}$ ( $\equiv\ \Gamma$ ) and take the value to be $\Gamma/t_x= 0.75$. We will mainly discuss the results obtained for the leads of the width $M=4$. Noninteracting systems ---------------------- We now discuss some properties in the noninteracting case in order to show the situations we are considering clear. Fig. \[fig:40\] shows the conductance $g_{\rm 1d}^0(E_F)$ for noninteracting electrons in the one-dimensional systems connected to the leads of the type I (solid line) and II (dashed line). In this figure, $E_F$ is the Fermi-energy, and the size of the central region is taken to be $N=4$. The coupling with the leads are taken to be $v/t_x=0.9$ for the type I leads, and $\Gamma/t_x= 0.75$ for the type II leads. There are $N$ ($=4$) peaks of the resonance, and the peaks become sharp if the mixing matrix element $v/t_x$ or $\Gamma/t_x$ decreases. In two dimensional lattice with the periodic boundary condition in the $y$-direction, the conductance for noninteracting electrons is given by $g_{\rm 2d}^0 = \sum_{n=1}^M \, g_{\rm 1d}^0(\epsilon_n)$ in the electron-hole symmetric case. For instance, in the case of $M=4$, the eigenvalues $\epsilon_n$ are given by $\epsilon_1= -\epsilon_4 = 2t_y$, and $\epsilon_2=\epsilon_3=0$. Therefore, when $t_y$ decreases, the conductance shows peaks corresponding to the resonance occurring in the one-dimensional system $g_{\rm 1d}^0(\epsilon_n)$. As an example of the the subband structure in the case of the type I leads, we show the dispersion relation of an ideal system $E_{nk_x} = -2t_x \cos k_x + \epsilon_n$ in Fig. \[fig:09\]. Here the size along the $y$-direction is taken to be $M=4$, and the hopping matrix elements are taken to be $t_y/t_x=1.0$ (solid line) and $t_y/t_x=0.5$ (dashed line). In the case of $t_y/t_x =1.0$ the highest and lowest subbands, i.e., modes $1$ and $4$, do not contribute to the conductance because the Fermi energy is situated at the band edge of the subbands. These two subband become conducting for $0.0<t_y/t_x <1.0$. We note that the modes $2$ and $3$ are degenerate and several curves are overlaps on the line corresponding to $-2t_x \cos k_x$ in Fig. \[fig:09\]. Hubbard model connected to the type I leads ------------------------------------------- In this subsection, we discuss the results for the Hubbard model connected to the type I leads. The conductance $g$ for the isotropic hopping $t_y/t_x=1.0$ is shown as a function of $U$ in Fig. \[fig:30\], where the size of the interacting region in the $y$-direction is $M=4$ and that in the $x$-direction is $N=4,6,8$, and $10$. The conductance decreases with increasing $U$. The $N$ dependence is similar to that in the one-dimensional case, i.e., $g$ decreases monotonically with increasing $N$ showing a tendency toward the Mott-Hubbard insulator.[@ao9] As mentioned in the above, the modes $1$ and $4$ are not conducting in the case of $t_y/t_x=1.0$, and the conductance is determined by the contributions of the modes $2$ and $3$. This situation is changed in the anisotropic cases $t_y/t_x <1.0$. In Fig. \[fig:31\], the conductance is plotted for several values of $t_y/t_x$, where the size of the interacting region is taken to be $N=4$ and $M=4$. There is a broad peak at finite $U$ for $0.65 \lesssim t_y/t_x \lesssim 0.75$. This shows that there are parameter regions where the total conductance increases with $U$ even in the half-filled case. To clarify this behavior, in Fig. \[fig:35\] the contributions of each of the conducting modes are plotted separately taking $t_y/t_x$ to be $0.7$. The resonant tunneling occurs in the modes $1$ and $4$ at $U/(2 \pi t_x) \simeq 3.3$, while the contributions of the modes $2$ and $3$ decreases monotonically with increasing $U$. Note that the pair of the subbands whose wavenumber in the $y$-direction are $k_y$ and $-k_y$ give the same contributions to the conductance. Furthermore in the electric-hole symmetric case, the contributions of the subbands whose eigenvalues are $\epsilon_n$ and $-\epsilon_n$ are the same. The occurrence of the resonant tunneling is linked with the behavior of the eigenvalues of effective Hamiltonian which is defined in terms of the renormalized hopping matrix element $-\widetilde{t}^{(n)}_{ll'}$ given by eq. (\[eq:25\]). In Fig. \[fig:50\], the eigenvalue of the $N\times N$ matrix $\widetilde{\mbox{\boldmath ${\cal H}$}}_{\rm C}^{(n)}$ is plotted as a function of $U$ for the mode $1$ (solid line), where $t_y/t_x=0.7$, $N=4$ and $M=4$. The second lowest eigenvalue crosses the Fermi energy at $U/(2 \pi t_x) \simeq 4.6$. This corresponds to the peak seen at $U/(2 \pi t_x) \simeq 3.3$ in Fig. \[fig:35\], although the values of $U$ do not coincide. This difference is mainly due to the contribution of the real part of the mixing term eq. (\[eq:87\]): $\mbox{Re} \widehat{\mbox{\boldmath ${\cal V}$}}_{\rm mix}^+(0)$ causes the energy shift, and the position of the resonance corresponding to the second eigenvalue moves toward the low-energy side. While the transmission probability is defined with respect to the effective Hamiltonian of the whole system $\widetilde{\cal H}_{\rm qp}$ in eq. (\[eq:H\_eff\]), the eigenvalues of the partial Hamiltonian $\widetilde{\mbox{\boldmath ${\cal H}$}}_{\rm C}^{(n)}$ are useful to investigate the behaviors of the conductance. Note that $\widetilde{t}^{(n)}_{ll'}$ depends on the mixing matrix elements $v_{\rm L}^{\phantom{0}}$ and $v_{\rm R}^{\phantom{0}}$ because the unperturbed Green’s functions used to calculate the self-energy eq. (\[eq:Self\_2v\]) are defined with respect to the whole system. In Fig. \[fig:50\], the eigenvalues for the modes $2$ and $3$ are also plotted (dashed lines). In the present case, the eigenvalues of the modes $2$ and $3$ are the same, and the eigenvalues of the modes $1$ and $4$ are symmetric with respect to the Fermi energy $\omega=0$. The energy gap between the second and third eigenvalues of the mode $2$ (or $3$) becomes large with increasing $U$, and it seems to show the tendency toward the Mott-Hubbard insulator. We have also examined the conductance in the case of $M=6$, where the subbands are classified into two groups; $\epsilon_1=-\epsilon_6=2t_y$ and $\epsilon_2=\epsilon_3=-\epsilon_4=-\epsilon_5=t_y$. In Fig. \[fig:57\], the total conductance (solid line) and the contributions of the two groups of subbands (dashed line) are shown for $t_y/t_x=0.65$, where the size of the interacting region is taken to be $N=6$ and $M=6$. The total conductance show two bump like behaviors at $U/(2\pi t_x)\simeq 1.5$ and $5.1$, which correspond to the resonance occurring in each group of the subbands. The resonance occurs at different values of $U$ for the different groups of the subbands. This seems to show a general tendency that the resonance does not occur simultaneously in the different groups of the subbands. Thus, when the resonance occurs at one group, the remaining $M-2$ or $M-4$ subbands behave monotonically. Therefore the presence of the visible maximum in the $U$ dependence of the total conductance is expected only for small $M$ where the contributions of the resonating subbands are comparable to the those of the remaining subbands. The number of the eigenvalues which cross the Fermi energy is expected to increase with the size in the $x$-direction $N$. To clarify this feature, we calculate the conductance for $N=10$ keeping the width $M=4$ unchanged. The results are shown in Fig. \[fig:24\] for several values of $t_y/t_x$. In each of the cases, at least one peak is visible at finite $U$. Specifically in the case of $t_y/t_x=0.5$, a shoulder like structure is seen at $U/2\pi t_x \simeq 0.5$ in addition to the peak at $U/2\pi t_x \simeq 3.8$. To see this behavior precisely, the contributions of each of the subbands are plotted in Fig. \[fig:51\]. Two resonant peaks are seen at $U/(2 \pi t_x) \simeq 0.5$ and $3.8$ in the modes $1$ and $4$. The $U$ dependence of the eigenvalues of $\widetilde{\mbox{\boldmath ${\cal H}$}}_{\rm C}^{(n)}$ for $n=1$ (mode $1$) is shown in Fig. \[fig:52\]. The 4th and 5th lowest eigenvalues cross the Fermi energy at $U/(2 \pi t_x) \simeq 0.7$ and $4.1$, respectively. These correspond to the two resonant peaks seen in Fig. \[fig:51\] as in the case of $N=4$. The energy shift caused by the mixing with the leads seems to decrease with increasing $N$. Hubbard model connected to the type II leads -------------------------------------------- We next discuss the conductance of the Hubbard model connected to the type II leads. In this lead, the local density of states at the interface of the lead is a constant and the bandwidth is infinity. Then $\widehat{\mbox{\boldmath ${\cal V}$}}_{\rm mix}^+(\omega)$ becomes pure imaginary and independent of $\omega$, i.e., $\mbox{Re}\, \widehat{\mbox{\boldmath ${\cal V}$}}_{\rm mix}^+(\omega) \equiv 0$ and $\partial \widehat{\mbox{\boldmath ${\cal V}$}}_{\rm mix}^+(\omega) /\,\partial \omega \equiv 0$. Therefore the energy shift caused by the mixing with the leads is expected to be small compared to that in the case of the leads I. In Fig. \[fig:54\] (a), the conductance in the case of $t_y/t_x=0.6$ and $0.78$ is shown as a function of $U$, where the size of the interacting region is chosen to be $N=4$ and $M=4$. There is a resonance peak at finite $U$ in each line; at $U/(2 \pi t_x) \simeq 4.3$ for $t_y/t_x=0.6$ and $U/(2 \pi t_x) \simeq 0.6$ for $t_y/t_x=0.78$. The contributions of each of the subbands are plotted separately in Fig. \[fig:54\] (b). The peak seen in the total conductance is due to the resonance tunneling in the modes $1$ and $4$. Qualitatively, these results are similar to those of the type I leads. In Fig. \[fig:55\], the eigenvalues of the effective Hamiltonian $\widetilde{\mbox{\boldmath ${\cal H}$}}_{\rm C}^{(n)}$ for $n=1$ (mode $1$) are plotted for (a) $t_y/t_x=0.6$ and (b) $t_y/t_x=0.78$. The second lowest eigenvalue for $t_y/t_x=0.6$ becomes zero at $U/(2 \pi t_x) \simeq 4.1$. This is close to the position of the resonance peak seen in Fig. \[fig:54\] (b). The lowest eigenvalue for $t_y/t_x=0.78$ does not cross the Fermi energy for any values of $U$, although it is nearly zero for small $U$. Nevertheless, this eigenstate corresponds to the resonance peak seen at $U/(2 \pi t_x) \simeq 0.6$ in Fig. \[fig:54\] (b). This is because the lowest eigenvalue is shifted slightly up due to the mixing, and it becomes to cross the Fermi energy. We next discuss the phase shift $\delta^{(n)}$. As mentioned, the charge displacement of the $n$-th subband can be deduced from the phase shift using the Friedel sum rule $\Delta N_{\rm tot}^{(n)} = 2\, \delta^{(n)}/\pi$, where the factor $2$ corresponds to the spin degeneracy. Specifically, in the case of the type II leads, the compensation theorem holds.[@Anderson] Then the second and the third term in the right-hand of eq. (\[eq:dn\_def\]) vanish because $\widehat{\mbox{\boldmath ${\cal V}$}}_{\rm mix}^+(\omega)$ is independent of $\omega$ (see also Appendix \[sec:Friedel\]). Thus the number of electrons in the type II leads is unchanged when the leads are connected to the interacting region, and the left-hand side of eq. (\[eq:Friedel\_mode\]), $\Delta N_{\rm tot}$, becomes equal to the number of electrons in the interacting region $\sum_{i\in {\rm C},\sigma} \langle c^{\dagger}_{i \sigma} c^{\phantom{\dagger}}_{i \sigma} \rangle$. In Fig. \[fig:56\], the phase shift $\delta^{(n)}/\pi$ is plotted as a function of $U$ for the modes $1$ and $4$, where the parameters are taken to be $N=4$, $M=4$, and $t_y/t_x=0.78$. Here the principal value of the phase shift is determined in the limit of $U \to 0$ by comparing with the charge displacement obtained independently. Reflecting the resonance seen in Fig. \[fig:54\] (b), $\delta^{(n)}$ for the mode $1$ starts to increase rapidly at $U/(2 \pi t_x) \simeq 0.6$. Correspondingly in the mode $4$, which is the lowest subband in this case, the phase shift $\delta^{(4)}$ decreases. These behaviors show how the filling of the subband changes when the resonance peak crosses the Fermi energy. In Fig. \[fig:56\], the sum of the two phase shifts is a constant, i.e., $2\,\delta^{(1)}/\pi + 2\,\delta^{(4)}/\pi = 2\,N$ with $N=4$. The phase shifts for the modes $2$ and $3$ are the same, and it is a constant independent of $U$, i.e., $2\,\delta^{(n)}/\pi= N$ for $n=2$ and $3$ with $N=4$. This means that these two subbands are half filled. Note that the total charge displacement is given by $\Delta N_{\rm tot}=M\times N$ in the electron-hole symmetric case. Remarks ------- The results presented in this section are obtained using the order $U^2$ self-energy given by eq. (\[eq:Self\_2v\]). Therefore, the results plotted for rather large values of $U$ may not be sufficient in quantitative sense. Nevertheless, some typical features of the results are seen even for small $U$. For instance, the resonance at finite $U$ occurs for small value of $U/(2 \pi t_x) \lesssim 1$ in the case of $t_y/t_x=0.78$ as shown in Fig. \[fig:54\] (a). Therefore we believe that the qualitative features of the results, such as the relation between the resonance and the eigenvalues of the effective Hamiltonian, hold also for large $U$. The perturbation approach gives us a correct, in principle, description of the conductance and the total charge displacement with respect to the local Fermi-liquid ground state. Summary ======= We have studied the conductance through small interacting systems connected to multi-mode leads based on a local Fermi-liquid approach. At $T=0$, the conductance and the total charge displacement are determined by the value of the Green’s function at the Fermi energy. Since the excitations at the Fermi energy do not decay at $T=0$, there exists the one-particle Hamiltonian which reproduces these two quantities exactly. We have applied this formulation to a two-dimensional Hubbard model of finite size in the electron-hole symmetric case. We have calculated all the matrix elements of the order $U^2$ self-energy and determine the parameters of the effective Hamiltonian up to terms of order $U^2$. Specifically, we have examined two types of the noninteracting leads: I) semi-infinite tight-binding leads, and II) leads of a constant density of states. The results are similar, qualitatively, in both types of the leads. The conductance shows maximums in the $U$ dependence for some ranges of $t_y/t_x$, where $t_x$ and $t_y$ are the hopping matrix element in the $x$- and $y$-directions, respectively. This means that there exists parameter regions, where the total conductance increases with $U$, even in the half-filled case. By decomposing the total conductance into the sum of the contributions of the subbands, it is clarified that the peaks of the conductance are caused by the resonance occurring in some group of the subbands which have the similar symmetric properties. The phase shift of the subbands obtained from the Friedel sum rule shows a typical changes when the resonance occurs. The resonance generally does not occur simultaneously in different groups of the subbands, and the subbands of off-resonance behave monotonically. Therefore the maximum in the $U$ dependence of the total conductance is expected only for the mesoscopic systems in which the number of the conducting modes is small enough and the contributions of the subbands of on-resonance are comparable to the those of the remaining subbands. The perturbation approach examined in this work can be applied to interacting electrons in disordered systems. Especially, the analysis of the eigen values of the effective Hamiltonian, which is used in the present study to investigate the behaviors of the resonance, may be combined with the Thouless-number[@LicciThouless] and finite-size scaling[@MackinnonKramer] methods. The application along this line seems to be interesting in relation to the metal-insulator transition observed in two-dimensional systems.[@Kravchenko] Furthermore, extensions to the finite temperatures[@ao11] and nonequilibrium situations[@ao10] are left for future studies. Acknowledgements {#acknowledgements .unnumbered} ================ We would like to thank K. Murata, K. Tanigaki, and S. Nonoyama for valuable discussions. Numerical computation was partly performed at computation center of Nagoya University and at Yukawa Institute Computer Facility. This work was supported by the Grant-in-Aid for Scientific Research from the Ministry of Education, Science and Culture, Japan. Conductance {#sec:CONDUCTANCE} =========== In this appendix, we provide the derivation of the dc conductance in the multi-mode case eq. (\[eq:g\_multi\]) by generalizing the derivation given in the single-mode case.[@ao6] The current operator corresponding to the mixing term eq. (\[eq:H\_mix\_multi\]) is given by $$\begin{aligned} J_{\rm L} &=& {\rm i}\, e \sum_{m=1}^{M_{\rm L}} \sum_{\sigma} v_{{\rm L},m}^{\phantom{\dagger}} \left(\, c^{\dagger}_{{\cal I}_m \sigma}\, c^{\phantom{\dagger}}_{{\cal L}_m \sigma} - c^{\dagger}_{{\cal L}_m \sigma}\, c^{\phantom{\dagger}}_{{\cal I}_m\sigma} \, \right) \label{eq:J_L} \;,\\ J_{\rm R} &=& {\rm i}\, e \sum_{m=1}^{M_{\rm R}} \sum_{\sigma} v_{{\rm R},m}^{\phantom{\dagger}} \left(\, c^{\dagger}_{{\cal R}_m \sigma}\, c^{\phantom{\dagger}}_{{\cal N}_m \sigma} - c^{\dagger}_{{\cal N}_m \sigma}\, c^{\phantom{\dagger}}_{{\cal R}_m \sigma} \, \right) \label{eq:J_R} \;.\end{aligned}$$ Here $J_{\rm L}$ is the total current flowing into the sample from the left lead, and $J_{\rm R}$ is the current flowing out to the right lead from the sample. These currents and the total charge in the sample $\rho_{\rm C}^{\phantom{0}} = e \sum_{j\in {\rm C}, \sigma} c^{\dagger}_{j \sigma} c^{\phantom{\dagger}}_{j \sigma}$ satisfy the equation of continuity $\partial \rho_{\rm C}^{\phantom{0}} / \partial t \,+ J_{\rm R} - J_{\rm L} = 0$. In the linear response theory, the dc conductance is given by $$\begin{aligned} g \ &=& \ \lim_{\omega \to 0} { K_{\alpha\alpha'}^+(\omega) - K_{\alpha\alpha'}^+(0) \over {\rm i} \omega } \label{eq:Kubo_M} \;, \\ K_{\alpha\alpha'}({\rm i} \nu_l) &=& \int_0^{\beta} d\tau \, \langle\, T_{\tau}\, J_{\alpha}(\tau)\, J_{\alpha'}(0) \, \rangle \, {\rm e}^{{\rm i}\, \nu_l \tau} \;, \end{aligned}$$ where $\alpha, \alpha' = {\rm L}, {\rm R}$, and $K_{\alpha\alpha'}^+(\omega)$ is the retarded function which is obtained from the Matsubara function $K_{\alpha\alpha'}({\rm i} \nu_l)$ by the analytic continuation ${\rm i}\nu_l \to \omega + {\rm i}0^+$. The dc conductance eq. (\[eq:Kubo\_M\]) corresponds to the $\omega$-linear imaginary part of $K_{\alpha\alpha'}^+(\omega)$, and it is independent of $\alpha$ and $\alpha'$ owing to the current conservation.[@Fisher; @Lee] Moreover, $K_{\alpha'\alpha}(z) = K_{\alpha\alpha'}(z)$ because of the time reversal symmetry, and thus the imaginary part of $K_{\alpha\alpha'}^+(\omega)$ corresponds to the discontinuity of $K_{\alpha\alpha'}(z)$ along the real axis in the complex $z$-plane. Therefore, the dc conductance is equal to the coefficient of the $\nu\, \mbox{sgn}\, \nu$ term of $K_{\alpha\alpha'}({\rm i}\nu)$,[@Shiba] where $\mbox{sgn}\, \nu$ is the sign function. In what follows, we extract this singular term from $K_{\alpha\alpha'}({\rm i}\nu)$ taking the current operators to be $\alpha={\rm R}$ and $\alpha'={\rm L}$. At $T=0$, $K_{\rm RL}({\rm i}\nu)$ is written, treating the Matsubara frequencies to be continuous variables, as \[see also Fig. \[fig:multi\]\] $$\begin{aligned} K_{\rm RL}({\rm i} \nu) &=& K_{\rm RL}^{(a)}({\rm i} \nu) \ + \ K_{\rm RL}^{(b)}({\rm i} \nu) \; , \label{eq:K_nu_M} \\ K_{\rm RL}^{(a)}({\rm i} \nu) &=& e^2 \sum_{\sigma} \int_{-\infty}^{+\infty} \! {{\rm d}\varepsilon \over 2 \pi} \nonumber \\ & & \times \mbox{Tr} \Bigl[\ \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal R}{\cal L}}({\rm i}\varepsilon + i\nu) \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal I}{\cal N}}({\rm i}\varepsilon) + \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal N}{\cal I}}({\rm i}\varepsilon + {\rm i}\nu) \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal L}{\cal R}}({\rm i}\varepsilon) \nonumber \\ & & \quad - \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal R}{\cal I}}({\rm i}\varepsilon + {\rm i}\nu) \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal L}{\cal N}}({\rm i}\varepsilon) - \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal N}{\cal L}}({\rm i}\varepsilon + {\rm i}\nu) \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal I}{\cal R}}({\rm i}\varepsilon) \ \Bigr] \;, \label{eq:K_nu_a_M} \\ K_{\rm RL}^{(b)}({\rm i} \nu) &=& e^2 \sum_{\sigma \sigma '} \sum_{ \{j\} \in {\rm C}} \int_{-\infty}^{+\infty} \! {{\rm d}\varepsilon {\rm d}\varepsilon' \over (2 \pi)^2} \; \Gamma_{\sigma \sigma ' ;\, \sigma ' \sigma}^{j_1 j_2;\, j_3 j_4} ({\rm i} \varepsilon + {\rm i} \nu, {\rm i} \varepsilon'+ {\rm i} \nu ; \, {\rm i} \varepsilon' , {\rm i} \varepsilon ) \nonumber \\ & & \times \Bigl\{ \mbox{\boldmath $G$}_{{\rm C}{\cal L}}({\rm i}\varepsilon + i\nu) \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal I}{\rm C}}({\rm i}\varepsilon) \, - \mbox{\boldmath $G$}_{{\rm C}{\cal I}}({\rm i}\varepsilon + {\rm i}\nu) \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal L}{\rm C}}({\rm i}\varepsilon) \Bigr\}_{j_1j_4} \nonumber \\ & & \times \Bigl\{ \mbox{\boldmath $G$}_{{\rm C}{\cal N}}({\rm i}\varepsilon') \, \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal R}{\rm C}}({\rm i}\varepsilon' + {\rm i}\nu) - \mbox{\boldmath $G$}_{{\rm C}{\cal R}}({\rm i}\varepsilon') \, \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal N}{\rm C}}({\rm i}\varepsilon' + {\rm i}\nu) \Bigr\}_{j_3j_2} \;. \label{eq:K_nu_b_M}\end{aligned}$$ Here $\mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}}$ and $\mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}}$ are diagonal matrices corresponding to $v_{{\rm L},m}^{\phantom{\dagger}}$ and $v_{{\rm R},m}^{\phantom{\dagger}}$, respectively. The subscript ${\rm C}$ denotes a set consisting of $N_{\rm C}$ sites in the central region. $\mbox{\boldmath $G$}_{{\rm C}{\cal L}}(z)$ is a $N_{\rm C} \times M_{\rm L}$ matrix, and the $(j,m)$ element is given by $G_{j,{\cal L}_m}(z)$ with $j \in {\rm C}$. Also, $\mbox{\boldmath $G$}_{{\cal R}{\rm C}}(z)$ is a $M_{\rm R} \times N_{\rm C}$ matrix, and the $(m,j)$ element is given by $G_{{\cal R}_m,j}(z)$ with $j \in {\rm C}$. The matrix Green’s functions in eqs. (\[eq:K\_nu\_a\_M\]) and (\[eq:K\_nu\_b\_M\]) are defined in this way. $\Gamma_{\sigma \sigma ' ;\, \sigma ' \sigma}^{j_1 j_2;\, j_3 j_4} ({\rm i} \varepsilon_1, {\rm i} \varepsilon_2 ;\, {\rm i} \varepsilon_{3}, {\rm i} \varepsilon_{4} )$ is the vertex corrections due to the inter-electron interaction, and illustrated in Fig. \[fig:vertex\]. Since we are considering the interaction which is switched on only in the central region, the Green’s function satisfies following relations at the interfaces; $$\begin{aligned} \left\{ \!\!\! \begin{array}{ll} \mbox{\boldmath $G$}_{{\cal R} \gamma}(z) = - \, \mbox{\boldmath $F$}_{\rm R}(z) \, \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal N}\gamma }(z) &\mbox{for} \ \ \gamma = {\cal L},\,{\cal I},\,{\rm C},\,{\cal N} \\ \mbox{\boldmath $G$}_{\gamma{\cal L}}(z) = - \, \mbox{\boldmath $G$}_{\gamma{\cal I}}(z) \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \mbox{\boldmath $F$}_{\rm L}(z) &\mbox{for} \ \ \gamma = {\cal I},\,{\rm C},\,{\cal N},\, {\cal R} \end{array} \right. \nonumber \\ . \label{eq:rec_lead_M}\end{aligned}$$ Using these relations, eqs. (\[eq:K\_nu\_a\_M\]) and (\[eq:K\_nu\_b\_M\]) are rewritten as $$\begin{aligned} K_{\rm RL}^{(a)}({\rm i} \nu) &=& e^2 \sum_{\sigma} \int_{-\infty}^{+\infty} \! {{\rm d}\varepsilon \over 2 \pi} \, \mbox{Tr} \biggl[\, \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \left[\, \mbox{\boldmath $F$}_{\rm R}({\rm i}\varepsilon + {\rm i}\nu) - \mbox{\boldmath $F$}_{\rm R}({\rm i}\varepsilon) \,\right] \, \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal N}{\cal I}}({\rm i}\varepsilon + {\rm i}\nu) \nonumber \\ & & \qquad \qquad \qquad \qquad \times \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \left[\, \mbox{\boldmath $F$}_{\rm L}({\rm i}\varepsilon + {\rm i}\nu) - \mbox{\boldmath $F$}_{\rm L}({\rm i}\varepsilon) \,\right] \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal I}{\cal N}}({\rm i}\varepsilon) \, \biggr] \;, \label{eq:bubble_M} \\ K_{\rm RL}^{(b)}({\rm i} \nu) &=& e^2 \sum_{\sigma \sigma '} \sum_{\{j\} \in {\rm C}} \int_{-\infty}^{+\infty} \! {{\rm d}\varepsilon {\rm d}\varepsilon' \over (2 \pi)^2} \; \Gamma_{\sigma \sigma ' ;\, \sigma ' \sigma}^{j_1 j_2;\, j_3 j_4} ({\rm i} \varepsilon + {\rm i} \nu, {\rm i} \varepsilon'+ {\rm i} \nu ;\, {\rm i} \varepsilon' , {\rm i} \varepsilon ) \nonumber \\ & & \times \Bigl\{ \mbox{\boldmath $G$}_{{\rm C}{\cal I}}({\rm i}\varepsilon + {\rm i}\nu) \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \left[\, \mbox{\boldmath $F$}_{\rm L}({\rm i}\varepsilon + {\rm i}\nu) - \mbox{\boldmath $F$}_{\rm L}({\rm i}\varepsilon) \,\right] \, \mbox{\boldmath $v$}_{\rm L}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal I}{\rm C}}({\rm i}\varepsilon) \Bigr\}_{j_1j_4} \nonumber \\ & & \times \Bigl\{ \mbox{\boldmath $G$}_{{\rm C}{\cal N}}({\rm i}\varepsilon') \, \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \left[\, \mbox{\boldmath $F$}_{\rm R}({\rm i}\varepsilon' + {\rm i}\nu) - \mbox{\boldmath $F$}_{\rm R}({\rm i}\varepsilon') \,\right] \, \mbox{\boldmath $v$}_{\rm R}^{\phantom{\dagger}} \, \mbox{\boldmath $G$}_{{\cal N}{\rm C}}({\rm i}\varepsilon' + {\rm i}\nu) \Bigr\}_{j_3j_2} \;. \label{eq:vertex_M}\end{aligned}$$ We can now extract the singular $\nu\, \mbox{sgn}\, \nu$ term of $K_{\rm RL}({\rm i}\nu)$ using eqs. (\[eq:bubble\_M\]) and (\[eq:vertex\_M\]), as in the case of the single Anderson impurity.[@Shiba; @ao3] It can be shown that there is no singular $\nu\, \mbox{sgn}\, \nu$ term in $K_{\rm RL}^{(b)}({\rm i}\nu)$ due the two factors $[ \mbox{\boldmath $F$}_{\rm L}({\rm i}\varepsilon + {\rm i}\nu) - \mbox{\boldmath $F$}_{\rm L}({\rm i}\varepsilon) ]$ and $[ \mbox{\boldmath $F$}_{\rm R}({\rm i}\varepsilon' + {\rm i}\nu) - \mbox{\boldmath $F$}_{\rm R}({\rm i}\varepsilon') ]$ which have different frequencies. Thus, the $\nu\, \mbox{sgn}\, \nu$ term comes only from $K_{\rm RL}^{(a)}({\rm i}\nu)$. It can be obtained from eq. (\[eq:bubble\_M\]) and yields eq. (\[eq:g\_multi\]). Friedel sum rule {#sec:Friedel} ================ We provide here the outline of the derivation of the Friedel sum rule in the single-mode case eq. (\[eq:Friedel\]) following Langer and Ambegaokar.[@LangerAmbegaokar] The Green’s function in the left lead $ij \in {\rm L}$ is written as $$\begin{aligned} G_{ij}^{+}(\omega) &=& F^{\rm L}_{ij}(\omega) \ + \ F^{\rm L}_{i0}(\omega)\, v_{\rm L}^{\phantom{\dagger}} \, G_{11}^{+}(\omega)\, v_{\rm L}^{\phantom{\dagger}} \, F^{\rm L}_{0j}(\omega) \;, \\ F_{ij}^{\rm L}(\omega) &=& \sum_n { \phi_{n,{\rm L}}^{\phantom{\dagger}}(i) \,\phi_{n,{\rm L}}^*(j) \over \omega - \epsilon_{n,{\rm L}} + {\rm i}0^+ } \;.\end{aligned}$$ Here $F_{ij}^{\rm L}(\omega)$ is the retarded Green’s function of the isolated lead, and $\epsilon_{n,{\rm L}}$ and $\phi_{n,{\rm L}}^{\phantom{\dagger}}(i)$ are the eigenvalue and the corresponding eigenfunction of the one-particle Hamiltonian ${\cal H}_{\rm L}$. Using the orthogonality relation $\,\sum_{i \in {\rm L}} \phi_{n,{\rm L}}^{\phantom{\dagger}}(i)\, \phi_{n',{\rm L}}^*(i) = \delta_{nn'}$, we obtain $$\begin{aligned} \sum_{i \in {\rm L}} \left[\, G_{ii}^{+} - F^{\rm L}_{ii}\,\right] \ = \ - \, {\partial F^{\rm L}_{00} \over \partial \omega} \, v_{\rm L}^{\phantom{\dagger}} \, G_{11}^{+} \, v_{\rm L}^{\phantom{\dagger}} \;. \label{eq:dn_L}\end{aligned}$$ Here $F^{\rm L}_{00} \equiv F_{\rm L}^{+}$, which is the Green’s function at the interface. Using eq. (\[eq:dn\_L\]) and the corresponding relation for the right lead, the displacement of the total charge defined by eq. (\[eq:dn\_def\]) is written, at $T=0$, as $$\begin{aligned} \Delta N_{\rm tot} &=& - \, {2 \over \pi} \ \mbox{Im} \, \int_{-\infty}^0 {\rm d}\omega \Biggl(\, \, \sum_{i \in {\rm C}} G_{ii}^+ \ + \ \sum_{i \in {\rm L}} \left[\, G_{ii}^+ - F^{\rm L}_{ii}\,\right] \ + \ \sum_{i \in {\rm R}} \left[\, G_{ii}^+ - F^{\rm R}_{ii}\,\right] \, \Biggr) \nonumber\\ &=& - \, {2 \over \pi} \ \mbox{Im} \, \int_{-\infty}^0 {\rm d}\omega \ \mbox{Tr} \left[\, \mbox{\boldmath ${\cal G}$}^+ \, - \, \mbox{\boldmath ${\cal G}$}^+ \, {\partial \mbox{\boldmath ${\cal V}$}_{\rm mix}^+ \over \partial \omega} \,\right] \nonumber\\ &=& - \, {2 \over \pi} \ \mbox{Im} \, \int_{-\infty}^0 {\rm d}\omega \left(\, {\partial \over \partial \omega} \ \mbox{Tr} \left[\, \log \left\{ \mbox{\boldmath ${\cal G}$}^+ \right\}^{-1} \,\right] \ + \ \mbox{Tr} \left[\, \mbox{\boldmath ${\cal G}$}^+ \, {\partial \mbox{\boldmath $\Sigma$}^+ \over \partial \omega} \, \right] \, \right) . \label{eq:dn}\end{aligned}$$ Here we have used a matrix notation as that was used in the text eqs. (\[eq:81\])-(\[eq:86\]) for the multi-mode case: the Green’s function in the central region is denoted by $\mbox{\boldmath ${\cal G}$}(z) = \{G_{ij}(z)\}$ with $ij \in {\rm C}$ and the Dyson equation is given by $ \left\{\mbox{\boldmath ${\cal G}$}(z)\right\}^{-1} = z \, \mbox{\boldmath $1$} - \mbox{\boldmath ${\cal H}$}_{\rm C}^0 - \mbox{\boldmath ${\cal V}$}_{\rm mix}(z) - \mbox{\boldmath $\Sigma$}(z)$, where $\mbox{\boldmath ${\cal H}$}_{\rm C}^0 = \{ -t_{ij}^{\rm C}-\mu \,\delta_{ij} \}$, and $\mbox{\boldmath ${\cal V}$}_{\rm mix}(z)$ and $\mbox{\boldmath $\Sigma$}(z)$ are the single-mode version of eqs. (\[eq:87\]) and (\[eq:86\]), respectively. It has been shown that the contribution of the second term in the third line of eq. (\[eq:dn\]) vanishes.[@LuttingerWard; @Luttinger] Thus, eq. (\[eq:dn\]) can be rewritten as $$\begin{aligned} \Delta N_{\rm tot} &=& {1\over \pi {\rm i}}\, \log \left[ \det\left\{ \mbox{\boldmath ${\cal G}$}^-(0) \right\}^{-1} / \det\left\{ \mbox{\boldmath ${\cal G}$}^+(0) \right\}^{-1} \right] \label{eq:detS} \\ &=& {1\over \pi {\rm i}}\, \log \, \det\left[ \mbox{\boldmath $1$} + 2 {\rm i} \,\mbox{Im} \left\{ \mbox{\boldmath ${\cal V}$}_{\rm mix}^+(0) \right\} \mbox{\boldmath ${\cal G}$}^+(0) \right] . \label{eq:dn_det}\end{aligned}$$ For obtaining eq. (\[eq:dn\_det\]), we have used the Fermi-liquid property $\mbox{Im} \mbox{\boldmath $\Sigma$}^+(0)$. The matrix $\mbox{\boldmath ${\cal V}$}_{\rm mix}$ has only two non-zero elements as eq. (\[eq:87\]). Therefore, the determinant in eq. (\[eq:dn\_det\]) can be written in terms of a $2\times 2$ matrix as $$\begin{aligned} & &\Delta N_{\rm tot} = {1 \over \pi {\rm i}}\, \log [\, \det \mbox{\boldmath $S$} \,] \;, \label{eq:Friedel} \\ \nonumber \\ & & \mbox{\boldmath $S$} = \left [ \matrix { 1 & 0 \cr 0 & 1 \cr } \right ] - \, 2 {\rm i} \left [ \matrix { \Gamma_{\rm L}(0) & 0 \cr 0 & \Gamma_{\rm R}(0) \cr } \right ] \left [ \matrix { G_{1 1}^{+}(0) & G_{1 N}^{+}(0) \cr G_{N 1}^{+}(0) & G_{N N}^{+}(0) \cr } \right ] . \nonumber\\ \label{eq:S}\end{aligned}$$ [99]{} See, for instance, T. Ishiguro, K. Yamaji and G. Saito: [*Organic Superconductors Second Edition*]{}, (Springer-Verlag, Berlin, 1998). T. K. Ng and P. A. Lee: Phys. Rev. Lett. [**61**]{} (1988) 1768. L. I. Glazman and M. E. Raikh: Pis’ma Zh.  Eksp. Teor. Fiz. [**47**]{} (1988) 378, \[JETP Lett. [**47**]{} 452 (1988)\]. A. Kawabata: J. Phys. Soc. Jpn. [**60**]{} (1991) 3222. Y. Meir, N. S. Wingreen and P. A. Lee: Phys. Rev. Lett. [**66**]{} (1991) 3048; [*ibid.*]{} [**70**]{} (1993) 2601. S. Hershfield, J. H. Davies and J. W. Wilkins: Phys. Rev. B [**46**]{} (1992) 7046. D. C. Ralph and R. A. Buhrman: Phys. Rev. Lett. [**72**]{} (1994) 3401. D. Goldharber-Gordon, H. Shtrikman, D. Mahalu, D. Abusch-Magder, U. Meirav and M. A. Kastner: Nature [**391**]{} (1998) 156. S. M. Cronenwett, T. H. Oosterkamp and L. P. Kouwenhoven: Science [**281**]{} (1998) 540. T. H. Oosterkamp, T. Fujisawa, W.G. van der Wiel, K. Ishibashi, R. V. Hijman, S. Tarucha and L. P. Kouwenhoven: Nature [**395**]{} (1998) 873. Y. Tokura, D. G. Austing and S. Tarucha: J. Phys.: Condens. Matter [**56**]{} (1999) 6023. W. Izumida, O. Sakai and Y. Shimizu: J. Phys. Soc. Jpn. [**66**]{} (1997) 717; [*ibid.*]{} [**67**]{} (1998) 2444. W. Izumida and O. Sakai: Phys. Rev. B [**62**]{} (2000) 10260; W. Izumida, O. Sakai and S. Suzuki: J. Phys. Soc. Jpn. [**70**]{} (2001) 1045. O. Sakai, S. Suzuki, W. Izumida and A. Oguri: J. Phys. Soc. Jpn. [**68**]{} (1999) 1640. A. Oguri, H. Ishii and T. Saso: Phys. Rev. B [**51**]{} (1995) 4715. A. Oguri: Phys. Rev. B [**56**]{} (1997) 13422 \[Errata: [**58**]{} (1998) 1690\]. A. Oguri: Phys. Rev. B [**59**]{} (1999) 12240. A. Oguri: Phys. Rev. B [**63**]{} (2001) 115305 \[Errata: [**63**]{} (2001) 249901\]. A. Yeyati, A. Martín-Rodero and F. Flores: Phys. Rev. Lett. [**71**]{} (1993) 2991. T. Mii and K. Makoshi: Jpn. J. Appl. Phys. [**35**]{} (1996) 3706. O. Takagi and T. Saso: J. Phys. Soc. Jpn. [**68**]{} (1999) 1997. P. L. Pernas, F. Flores and E. V. Anda: J. Phys.: Condens. Matter [**4**]{} (1992) 5309. Y. Kawahito, H. Kasai, H. Nakanishi and A. Okiji: J. Appl. Phys. [**85**]{} (1999) 947. N. D. Lang and Ph. Avouris: Phys. Rev. Lett. [**84**]{} (1999) 358. P. W. Anderson: Phys. Rev. [**124**]{} (1961) 41. D. C. Licciardello and D. J. Thouless: J. Phys. C [**8**]{} (1975) 4157. A. MacKinnon and B. Kramer: Z. Phys. B [**53**]{} (1983) 1. S. V. Kravchenko, G. V. Kravchenko, J. E. Furneaux, V. M. Pudalov and M. D’Iorio: Phys. Rev. B [**50**]{} (1994) 8039. A. Oguri: J. Phys. Soc. Jpn. [**70**]{}, (2001) 2666. A. Oguri: Phys. Rev. B [**64**]{}, (2001) 153305. D. S. Fisher and P. A. Lee: Phys. Rev. B [**23**]{} (1981) 6851. P. A. Lee and D. S. Fisher: Phys. Rev.  Lett. [**47**]{} (1981) 882. J. S. Langer and V. Ambegaokar: Phys. Rev. [**121**]{} (1961) 1090. D. C. Langreth: Phys. Rev. [**150**]{} (1966) 516. H. Shiba: Prog. Theor. Phys. [**54**]{} (1975) 967. A. Oguri: Phys. Rev. B [**52**]{} (1995) 16727. J. M. Luttinger and J. C. Ward: Phys. Rev. [**118**]{} (1960) 1417. J. M. Luttinger: Phys. Rev. [**119**]{} (1960) 1153.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Local linear gyrokinetic simulations show that electron temperature gradient (ETG) instabilities are the fastest growing modes for $k_y \rho_i \gtrsim 0.1$ in the steep gradient region for a JET pedestal discharge (92174) where the electron temperature gradient is steeper than the ion temperature gradient. Here, $k_y$ is the wavenumber in the direction perpendicular to both the magnetic field and the radial direction, and $\rho_i$ is the ion gyroradius. At $k_y \rho_i \gtrsim 1$, the fastest growing mode is often a novel type of toroidal ETG instability. This toroidal ETG mode is driven at scales as large as $k_y \rho_i \sim (\rho_i/\rho_e) L_{Te} / R_0 \sim 1$ and at a sufficiently large radial wavenumber that electron finite Larmor radius effects become important; that is, $K_x \rho_e \sim 1$, where $K_x$ is the effective radial wavenumber. Here, $\rho_e$ is the electron gyroradius, $R_0$ is the major radius of the last closed flux surface, and $1/L_{Te}$ is an inverse length proportional to the logarithmic gradient of the equilibrium electron temperature. The fastest growing toroidal ETG modes are often driven far away from the outboard midplane. In this equilibrium, ion temperature gradient instability is subdominant at all scales and kinetic ballooning modes are shown to be suppressed by $\mathbf{ E} \times \mathbf{ B} $ shear. ETG modes are very resilient to $\mathbf{ E} \times \mathbf{ B}$ shear. Heuristic quasilinear arguments suggest that the novel toroidal ETG instability is important for transport.' address: - '$^1$Rudolf Peierls Centre for Theoretical Physics, University of Oxford, Oxford OX1 3PU, UK' - '$^2$Culham Centre for Fusion Energy, Culham Science Centre, Abingdon OX14 3DB, UK' - '$^3$Merton College, Merton Street, Oxford, OX1 4JD, UK' - '$^4$Department of Physics, University of Maryland, College Park, Maryland 20742, USA' - '$^5$Institute for Fusion Studies, University of Texas at Austin, Austin, Texas, 78712, USA' - '$^6$National Institutes for Quantum and Radiological Science and Technology, Rokkasho, Aomori 039-3212, Japan' - '$^7$École Polytechnique Fédérale de Lausanne (EPFL), Swiss Plasma Center (SPC), CH-1015 Lausanne, Switzerland' - '\*See the author list of Joffrin E. et al. 2019 Nucl. Fusion 59, 112021' author: - 'Jason F. Parisi$^{1,2,3}$, Felix I. Parra$^{1}$, Colin M. Roach$^{2}$, Carine Giroud$^{2}$, William Dorland$^{4,1}$, David R. Hatch$^{5}$, Michael Barnes$^{1}$, Jon C. Hillesheim$^{2}$, Nobuyuki Aiba$^{6}$, Justin Ball$^{7}$, Plamen G. Ivanov$^{1,2}$, and JET Contributors\*' bibliography: - 'EverythingPlasmaBib.bib' title: 'Toroidal and slab ETG instability dominance in the linear spectrum of JET-ILW pedestals' --- Introduction {#sec:introduction} ============ H-mode is currently the most favored high confinement operating regime in tokamaks. In H-mode, plasma confinement significantly improves once plasma heating exceeds a certain threshold [@Wagner1982; @Ryter1996]. H-mode was first discovered in ASDEX [@Wagner1982], and subsequently in most other tokamaks [@Burrell1989; @Bush1990; @Greenwald1997; @Keilhacker1999]. The precursor to H-mode, L-mode [@Burrell1992], has fairly constant equilibrium gradients across its radius, whereas H-mode is characterized by the presence of a pedestal with decreased turbulent particle and heat diffusivities, and therefore significantly increased equilibrium gradients. These increased gradients drive MHD instabilities, which set hard limits on the maximum achievable pressure gradient [@Keilhacker1984; @Connor1998; @Huysmans2005; @Snyder2009]. Turbulent transport caused by microinstabilities driven unstable by equilibrium gradients that steepen during the inter-ELM (inter-edge-localized mode) period [@Hill1997] can constrain other pedestal dynamics such as MHD stability [@Rogers1998; @Hatch2017], scrape off layer and divertor physics [@Neuhauser2002], and neoclassical transport [@Pusztai2016], and hence studying H-mode inter-ELM pedestal microstability is of great interest. To study the pedestal microinstabilities, we use gyrokinetics [@Taylor1968; @Catto1978; @Antonsen1980; @Frieman1982; @Parra2008; @Abel2013] — an asymptotic approach to solving the Fokker-Planck kinetic equation. Gyrokinetics is well-suited for studying anisotropic turbulence in highly magnetized plasmas. One of its main results, the gyrokinetic equation, is a nonlinear partial differential equation for the time evolution of the perturbed gyroaveraged distribution function. We will use the linearized gyrokinetic equation in conjunction with Maxwell’s equations to study microinstabilities in JET pedestals. The local $\delta f$ gyrokinetic code GS2 [@Dorland2000] is used to simulate the pedestal plasmas presented in this article. We study the stability of a JET ITER-like wall (JET-ILW) inter-ELM magnetic equilibrium with different ion and electron temperature profiles. The ion and electron temperatures are obtained using impurity charge exchange emission and Thomson scattering, respectively. Since $\mathbf{ E} \times \mathbf{ B} $ shear is hypothesized to play a key role in pedestal formation [@Burrell1992; @Hahm1995; @Biglari1990], we focus on the radial region near the maximum value of the equilibrium $\mathbf{ E} \times \mathbf{ B}$ shear. The region of maximum $\mathbf{ E} \times \mathbf{ B}$ shear is estimated by balancing the radial electric field with the pressure gradient. Gyrokinetic studies of pedestals have been performed before. Local gyrokinetic analysis of MAST found the main instabilities at $k_{\perp} \rho_i \sim 1$ to be kinetic ballooning modes (KBMs) in the steep pressure gradient region and microtearing modes (MTMs) in the less steep pressure gradient region inside the pedestal top, throughout the inter-ELM recovery of the pedestal [@Dickinson2012]. A follow up study using DBS and cross-polarization scattering found that $k_{\perp} \rho_i \approx 3 - 4$ turbulence at the pedestal top in MAST was most consistent with the electron temperature gradient (ETG) instability [@Hillesheim2016]. Using the Gyrokinetic Toroidal Code [@Lin1998], PIC simulations in the steep gradient region of DIII-D discharges found electrostatic electron-driven modes peaking at poloidal angle $\theta = \pm \pi/2$ [@Fulton2014]. More recently, in JET-ILW discharges where the ion temperature was not measured and was assumed to be equal to the measured electron temperature, nonlinear global gyrokinetic calculations were performed using the GENE code [@Jenko2000; @Gorler2011]. These global simulations predict pedestal heat transport fluxes that are comparable with experiment, and suggest that pedestal fluxes will be increasingly dominated by ion temperature gradient (ITG) turbulence as the heating power increases [@Hatch2017]. Hatch et al. also proposed that impurity seeding reduces ion-scale and ETG instability transport via ion-dilution and increased collisionality [@Hatch2017]. In [@Hatch2016], it was again demonstrated that the sum of neoclassical, MTM, and ETG turbulent transport was in good agreement with another JET-ILW pedestal measurement. Another recent work that used experimental ion temperature profiles found that ITG was suppressed in JET Carbon discharges, but not in JET-ILW cases, where ITG turbulence carried a substantial fraction of the total heat flux [@Hatch2019]. The difference between JET Carbon and JET-ILW was attributable to a decreased density gradient in JET-ILW discharges, which increased the growth rates of slab ITG and ETG instabilities. In this work, we identify a novel type of toroidal ETG instability that appears in regions of steep equilibrium temperature gradients. These sub-ion Larmor scale modes have a radial wavenumber larger than its poloidal wavenumber, and have been observed (but not explained) in previous pedestal simulations [@Told2008; @Jenko2009; @Told2012; @Fulton2014; @Baumgaertel2011; @Kotschenreuther2019]. The particularly large radial wavenumber means that the radial magnetic drift plays an important role in these toroidal ETG modes. We find that this toroidal ETG has a large critical gradient threshold, which occurs due to the pedestal’s magnetic geometry and the radial magnetic drift. Moreover, because of the large equilibrium temperature gradients, we show theoretically and numerically that both toroidal and slab ETG modes are extended from perpendicular scales of $k_y \rho_e \sim 1$ in the core, to $k_y \rho_i \sim 1$ in the pedestal, where $k_y$ is the binormal wavenumber, defined in , and $\rho_s$ is the gyroradius for a species $s$. We primarily examine microinstability at a single radial location in the steep gradient region of JET-ILW shot 92714 [@Giroud2018], a highly-fueled deuterium discharge with deuterated ethylene ($\mathrm{C}_2 \mathrm{D}_4$) injection. For this discharge, at all scales where instability occurs — $0.005 \lesssim k_y \rho_i \lesssim 400$ — we find that electron temperature gradient-driven modes are the fastest growing modes. For $k_y \rho_i \gtrsim 1$, the novel toroidal ETG mode is usually the fastest growing mode. We also show that the gradients of the measured ion temperature profiles are insufficiently steep to drive ITG instability. With the measured ion temperature profiles, the ion temperature gradient is close to the critical gradient needed for linear instability and hence subdominant. We also show that if ion temperature gradients are made sufficiently steep, toroidal and slab ITG modes become unstable at $k_y \rho_i \ll 1$, but are suppressed by $\mathbf{ E} \times \mathbf{ B}$ shear. Our findings suggest that the toroidal and slab ITG mode are stable in many radial pedestal locations, even in the steep gradient region that we examine. The layout of this paper is as follows: we first introduce gyrokinetics and the notation used throughout this paper in . We then present JET-ILW density, temperature, and rotation profiles from an inter-ELM pedestal in . Here, we also give a broad overview of the growth rates and unusual mode structures for the fastest growing modes in this pedestal, including a discussion of electromagnetic effects and $\mathbf{ E} \times \mathbf{ B}$ shear. At a wide range of scales, we find an ETG mode with unusual character. This mode typically has a radial wavenumber that is significantly larger than the poloidal wavenumber, and is insensitive to finite $\beta$ effects and $\mathbf{ E} \times \mathbf{ B}$ shear. Motivated by the results of and using the notation of , we then make analytic predictions about microinstability in steep gradient regions in . This theoretical analysis explains the existence of the novel toroidal ETG modes that we see in . We then examine ETG and ITG (or lack thereof) instability in linear gyrokinetic simulations in , respectively. The effect of $\mathbf{ E} \times \mathbf{ B} $ shear is discussed further in . Finally, we conclude in . Experimentally-minded readers might wish to jump to , while those more theoretically inclined and with a background in gyrokinetics might wish to begin at . Gyrokinetics {#sec:gyrokineticreviews} ============ In this section, we introduce the system of gyrokinetic equations and notation used throughout this paper. This section can be skipped for readers well-acquainted with gyrokinetics, or who mainly wish to see gyrokinetic simulations results in . Gyrokinetics [@Taylor1968; @Catto1978; @Antonsen1980; @Frieman1982; @Parra2008; @Abel2013] is used to investigate turbulence and transport using an asymptotic expansion in the ratio of $ \rho_{*s} \equiv \rho_s / L_{Ts} \ll 1$. We express the gradients by the equilibrium length scales $L_{Q} \equiv -( \partial \ln Q / \partial r )^{-1}$, where $Q$ can be the equilibrium density, temperature, or pressure, and the distance $r$ is half of the diameter of the flux surface at the midplane. Assuming $k_{\perp} \rho_i \sim 1$ and $\omega \ll \Omega_s$, gyrokinetics describes plasma behavior on spatial scales comparable to the ion gyroradius, and on timescales much longer than the gyro period. The quantity $k_{\perp}$ is the perpendicular turbulence wavenumber, $\omega $ is the frequency for turbulence quantities, $\Omega_s = Z_s e B / m_s c $ is the gyrofrequency, $Z_s$ is the charge number, $e$ is the proton charge, $B$ is the magnetic field strength, $m_s$ is the species mass, and $c$ is the speed of light. The gyrokinetic ordering is $\rho_{*s} \sim \omega / \Omega_s \sim \nu_s / \Omega_s \sim k_{\parallel} / k_{\perp} \ll 1$, where $\nu_s$ is a typical collision frequency for species $s$, and $k_{\parallel}$ is the turbulence parallel wavenumber. To obtain a rough estimate for the radial electric field (see ), we will impose that the radial electric field is comparable to the pressure gradient, which implies a low flow ordering [@Hinton1976; @Helander2002; @Parra2015] for the electric field, $| \mathbf{E} | \sim T_{0e} / e L_{Te}$, that is, the equilibrium $\mathbf{ E} \times \mathbf{ B} $ drift is small compared with the thermal velocity $v_{ts} = \sqrt{2T_{0s}/m_s} $ by a factor of $\rho_{*s}$, where $T_{0s}$ is the leading order temperature. We expand the magnetic field in $\rho_{*s} $, $\mathbf{B} + \mathbf{B}_1 + \mathbf{B}_2 + \ldots $, where $\mathbf{B}_n = \rho_{*s}^n \mathbf{B} $. The lowest order magnetic field is written as $\mathbf{B} = I(r) \nabla \zeta + \nabla \zeta \times \nabla \psi $, where $\zeta$ is the toroidal angle, $\psi$ is the poloidal flux divided by $2\pi$, and $I(r)$ is a flux function. For $n \geq 1$, we further split $\mathbf{B}_n $ into long-wavelength and turbulence components, $\mathbf{B}_n = \mathbf{B}_n^{lw} + \mathbf{B}_n^{tb} $. Long wavelength quantities, $g ^{lw}$, spatially change on equilibrium length scales, $\nabla g ^{lw} \sim g ^{lw} / L_{Ts}$, and temporally change on slow time scales, $\partial g ^{lw}/ \partial t \sim g ^{lw} / \tau_E$, where $\tau_E$ is the energy confinement time and $t$ is the time variable. Turbulence quantities, $g ^{tb} $, spatially change on equilibrium length scales along the mean magnetic field, $ \hat{\mathbf{ b} } \cdot \nabla g ^{tb} \sim g ^{tb}/ L_{Ts}$, but on gyroradius scales across the mean field, $\nabla_{\perp} g ^{tb} \sim g ^{tb} / \rho_s$, and temporally change on fast time scales, $\partial g ^{tb} / \partial t \sim \omega g ^{tb} $. Here, $\hat{\mathbf{ b} } = \mathbf{B} / B $, and $\nabla_{\perp}$ is a spatial derivative perpendicular to $\mathbf{B} $. We ignore the correction, $\mathbf{B}_1 ^{lw} $, which is mainly due to the effect of the neoclassical pressure anisotropy on the magnetic field. One can show that the turbulent component of $\mathbf{B}_1 $ can be written as $\mathbf{B}_1 ^{tb} = \nabla \overline{ A}_{\parallel 1 } ^{tb} \times \hat{\mathbf{ b} } + \overline{ B}_{\parallel 1 } ^{tb} \hat{\mathbf{ b} } $, where $\overline{B}_{\parallel 1} ^{tb} $ and $\overline{ A}_{\parallel 1} ^{tb} $ are the leading order parallel components of the turbulent magnetic field and magnetic vector potential, respectively. We reserve the overline notation for these turbulent quantities because later we will write their Fourier components without an overline, which will keep the notation tidier. For the electric field $\mathbf{E}$, we also expand in $\rho_{*s} $, $\mathbf{E} = \mathbf{E}_0 + \mathbf{E}_1 + \ldots $, where $\mathbf{E}_n \sim \rho_{*s}^n T_{0s} / e L_{Ts} $. We split $\mathbf{E}_n $ into long wavelength and turbulent parts, $\mathbf{E}_n = \mathbf{E}_n ^{lw} + \mathbf{E}_n ^{tb} $. To lowest order, $\mathbf{E}_0 $ is electrostatic; $\mathbf{E}_0 ^{lw} = - \nabla \phi_0 $, and $\mathbf{E}_0 ^{tb} = - \nabla_{\perp} \overline{\phi} ^{tb}_1 $. Here, $\phi_0$ is the leading order electric potential and $\phi ^{tb} _1$ is the leading order turbulent electric potential, where $\phi ^{tb} _1 \sim \rho_{*s} \phi_0$. Since $\phi_0$ is a flux function, $\mathbf{E}_0 \cdot \hat{\mathbf{ b} } = 0 $. To leading order, the parallel components of the electric field are $E_{\parallel } ^{lw} = - \hat{\mathbf{ b} } \cdot \nabla \phi_1 ^{lw} $ and $E_{\parallel } ^{tb} = - \hat{\mathbf{ b} } \cdot \nabla \overline{ \phi}_1 ^{tb} - (1/c) (\partial \overline{A}_{\parallel 1 } ^{tb} / \partial t )$. The electrostatic potential $\phi ^{lw} _1$ is mainly due to neoclassical physics. We expand the distribution function in $\rho_{*s}$, $f_{s} = F_{Ms} + f_{1s} + \ldots$, where the lowest order distribution function, $F_{Ms}$, is a stationary Maxwellian, $$F_{Ms} (r, v) = n_{0s} (r) \Big{(} \frac{m_s}{2 \pi T_{0s} (r)} \Big{)}^{3/2} \exp \Big{(} - \frac{m_s v^2 }{2 T_{0s} (r)} \Big{)},$$ with particle speed $v$, and flux functions $n_{0s}$ and $T_{0s}$, where $n_{0s}$ is the leading order density. The Maxwellian is stationary because the mean flow is subsonic. Higher order corrections to the distribution function can be split into long-wavelength and turbulent quantities, $f_{ns} = f ^{lw}_{ns} + f ^{tb}_{ns}$, where neoclassical corrections would be included in $f ^{lw}_{ns} $. To describe phase space, we will employ gyrokinetic variables. These are the guiding center, $\mathbf{R}_s$, the kinetic energy, $\mathcal{E} = v^2 /2$ , the magnetic moment, $\mu = v_{\perp}^2 / 2B$ where $\mathbf{v}_{\perp} = \mathbf{v} - \mathbf{v} \cdot \hat{\mathbf{ b} } \hat{\mathbf{ b} } $, and the gyrophase, $\varphi$, which is a particle’s angular location during its gyromotion. The guiding center is given by $\mathbf{R}_s = \mathbf{r} - \bm{\rho}_s $, the gyroradius position is given by $\bm{\rho}_s = \hat{\mathbf{ b} } \times \mathbf{v} / \Omega_s $, and the quantity $\mathbf{r} $ is the particle position. The first order turbulent component of the distribution function can be written as $$f^{tb}_{1s} (\mathbf{R}_s , \mathcal{E}, \mu, \varphi, t) = \overline{{h}}_s \big{(} \mathbf{R}_s , \mathcal{E}, \mu, t \big{)} - \frac{Z_s e \overline{ \phi} ^{tb}_1 }{T_{0s}} F_{Ms} (\mathbf{r} , \mathcal{E}, t). \label{eq:hdefinition}$$ Note that the function $\overline{ h}_s$ is independent of the gyrophase — our task is to find an evolution equation for $\overline{ h}_s$. To find $\overline{ h}_s$, we substitute into the first order Fokker-Planck equation. Because only the variable $\varphi $ varies over a single gyroperiod, it is convenient to average the Fokker-Planck equation over the gyromotion using a gyroaverage, defined as $\langle \ldots \rangle = (1/2\pi) \int_0^{2\pi} \ldots d \varphi |_{\mathbf{R}_s , \mathcal{E}, \mu}$, evaluated at fixed $\mathbf{R}_s, \mathcal{E}$, and $\mu$. Gyroaveraging the first order Fokker-Planck equation and taking its turbulent component, we obtain the low flow electromagnetic gyrokinetic equation, $$\fl \eqalign{ \bigg{(} \frac{\partial }{\partial t} + \Omega_E \frac{\partial}{\partial \zeta} \bigg{)} \overline{h}_{s} + & (v_{\parallel } \hat{\mathbf{b} } + \mathbf{ v}_{Ms} + \langle \mathbf{ v}_{\chi} ^{tb} \rangle ) \cdot \nabla_{\mathbf{R_s} } \overline{ h}_{s} + \sum_s \bigg{\langle} C^{(l)}_{ss'} \bigg{\rangle} \\ & = \frac{Z_s e F_{Ms}}{T_{0s}} \bigg{(} \frac{\partial }{\partial t} + \Omega_E \frac{\partial}{\partial \zeta} \bigg{)} \langle \overline{ \chi} ^{tb}_1 \rangle - \langle \mathbf{ v}_{\chi} ^{tb} \rangle \cdot \nabla_{\mathbf{R}_s } F_{Ms}, } \label{eq:lowfloweq}$$ where $\Omega_E (r) = - c \partial \phi_0 / \partial \psi$ is the $\mathbf{ E} \times \mathbf{ B} $ toroidal angular velocity, $C^{(l)}_{ss'}$ is a linearized Fokker-Planck collision operator, $\nabla_{\mathbf{R}_s } \equiv \partial / \partial \mathbf{R}_s $, the magnetic drift is $$\mathbf{v}_{Ms} = \frac{ \hat{\mathbf{ b} }}{\Omega_s} \times \left[ \left( v_{\parallel }^2 + \frac{v_{\perp}^2}{2} \right) \nabla \ln B + v_{\parallel}^2 \frac{4 \pi }{B^2} \frac{\partial p_0}{\partial r} \nabla r \right]. \label{eq:magneticdrift}$$ Here, $p_0 = \sum_s p_{0s}$ is the total pressure and $p_{0s} = n_{0s} T_{0s}$ is the lowest order pressure. The parallel velocity is $v_{\parallel } = \mathbf{v} \cdot \hat{\mathbf{ b} }$, and the gyrokinetic drift is $ \mathbf{ v}_{\chi}^{tb} = (c/B) \hat{\mathbf{b} } \times \nabla \overline{ \chi} ^{tb}_1$. Here, $\overline{ \chi} ^{tb}_1 $ is the leading order turbulent gyrokinetic potential defined as $$\overline{\chi} ^{tb}_1 = \overline{ \phi }^{tb}_1 - \frac{ v_{\parallel } \overline{A} ^{tb} _{\parallel 1}}{c} + \frac{m_s}{Z_s e} \int_0^{\mu} \overline{ B}_{\parallel 1} ^{tb} (\mathbf{R}_s + \bm{\rho}_s (\mu') ) d\mu'.$$ In Equation (3), $\Omega_E(r)$ can be approximated around the radial location $r_c$ of interest by $\Omega_E(r_c) + (r - r_c) (\partial \Omega_E/ \partial r)$ because the characteristic size of the eddies is small compared with $L_{Te}$. In the low flow ordering that we use, the term $(r - r_c) (\partial \Omega_E/ \partial r)$, which represents the $\mathbf{E} \times \mathbf{B}$ shear, should be neglected because it is of the same size as other terms that we have not kept. Even so, we perform some simulations with $\mathbf{E} \times \mathbf{B}$ shear. We will justify using this small term in . To close the system of equations, we need to find $ \overline{ \phi} ^{tb} _1 $, $\overline{A}_{\parallel 1} ^{tb} $, and $\overline{ B}_{\parallel 1} ^{tb} $ using $ \overline{ h}_s$. To find $\overline{ \phi} ^{tb} _1$, we use the first order turbulent quasineutrality condition, $$\sum_s \frac{Z_s^2 e^2 \overline{ \phi} ^{tb} _1}{T_{0s}} n_{0s} = \sum_s Z_s e \int \overline{ h}_s (\mathbf{r} - \bm{\rho}_s, \mathcal{E}, \mu) d^3 v. \label{eq:turbQN}$$ The parallel vector potential, $\overline{ A} ^{tb}_{\parallel 1}$, is found using the parallel component of Ampère’s law, $$- \nabla^2_{\perp} \overline{A}_{\parallel 1} ^{tb} = \frac{4 \pi e}{c} \sum_s Z_s \int v_{\parallel } \overline{ h}_s (\mathbf{r} - \bm{\rho}_s, \mathcal{E}, \mu) d^3 v.$$ Finally, $\overline{ B}_{\parallel 1} ^{tb}$ is determined by perpendicular pressure balance, $$\frac{B \overline{ B}_{\parallel 1} ^{tb} }{4 \pi} + \sum_s \int m_s B \int_0^{\mu} \overline{h}_s (\mathbf{r} - \bm{\rho}_s (\mu'), \mathcal{E}, \mu) d \mu' d^3 v = 0.$$ Note that the integral over $\mu'$ only affects the $\mu$ dependence of $\bm{\rho}_s$. Throughout this paper, we will examine the stability properties of the gyrokinetic equation in the linear local limit. To understand how these linear instabilities then cause turbulent transport, one needs to keep the nonlinear term of , which we will neglect in this work. The local limit, $k_{\perp} L_{Ts} \gg 1$, is useful for analytic treatment and numerically efficient simulations. If $k_{\perp} L_{Ts} \gg 1$, modes can be Fourier analyzed in the perpendicular domain. In JET shot 92174 at the radial location we examine, $L_{Te} \simeq 0.02 \mathrm{m}$, and thus the local approximation is good provided that $k_{\perp} \rho_i \gg \rho_i / L_{Te} \simeq 0.12$. To describe the properties of the turbulent pieces, $\overline{ \phi }^{tb}_1, \overline{A}_{\parallel 1} ^{tb}, \overline{B}_{\parallel 1} ^{tb}$, and $\overline{ h}_s $, we use the flux coordinates $(x, y, \theta)$. The coordinate $x$ is a local flux surface label defined around the flux surface $r_c$ (note that it is different from the flux label $r$), $y$ is a field line label, and $\theta$ is a poloidal coordinate that labels the position along the magnetic field line. The coordinates $x$ and $y$ are given by $$x = \frac{q(r_c)}{r_c B_a}(\psi(r) - \psi(r_c)), \;\;\; y = \frac{1}{B_a} \frac{\partial \psi}{\partial r} \bigg{|}_{r_c} \alpha,$$ where $B_a$ is the toroidal magnetic field strength evaluated at $r_c$ and $R_c$, $R_c$ is the distance from the axis of symmetry of the tokamak to the center of the flux surface $r_c$ at the midplane, $\alpha = \zeta - q \theta + \nu (r, \theta)$, and $\nu (r, \theta)$ is a function $2 \pi$-periodic in $\theta$, $$\fl \nu (r, \theta) = - I(r) \bigg{(} \int_0^{\theta} d \theta' \bigg{[} \frac{1}{R^2(\theta') \mathbf{B} (\theta') \cdot \nabla \theta' } - \frac{1}{2\pi} \oint \frac{d \theta''}{R^2(\theta'') \mathbf{B}(\theta'') \cdot \nabla \theta''} \bigg{]} \bigg{)}.$$ The safety factor, $q(r)$, is given by $2\pi q(r) = \oint I(r) d \theta / R^2 \mathbf{B} \cdot \nabla \theta $. We choose to define the poloidal angle $\theta$ as $$\theta = 2\pi l /L_{\theta}, \label{eq:thetaequation}$$ where $l$ is the arclength along the magnetic field, and $L_{\theta}$ is the distance along a field line for one complete poloidal turn. Spatial anisotropy, $k_{\perp}/k_{\parallel} \gg 1$, implies that $\partial / \partial x \sim \partial / \partial y \gg (2\pi / L_{\theta}) \partial / \partial \theta$. In the linear local limit, we Fourier analyze $\overline{ \phi}^{tb}_1$ locally in the perpendicular plane and in time, $$\overline{\phi} ^{tb}_1 (x, y, \theta, t) = \sum_{k_{x}, k_{y}, \omega } \phi ^{tb} _1 (k_{x}, k_{y}, \theta, \omega) \exp(ik_{x} x + i k_{y} y - i \omega t). \label{eq:phifourier}$$ The electromagnetic fluctuations $\overline{A}_{\parallel } ^{tb} $ and $\overline{B}_{\parallel } ^{tb} $ are Fourier analyzed in a similar way. It will also be useful to Fourier analyze $\overline{h}_s$, $$\fl \overline{h}_s (X_s, Y_s, \theta, \mathcal{E} ,\mu, t) = \sum_{k_{x}, k_{y}, \omega} h_s (k_{x}, k_{y}, \theta, \mathcal{E}, \mu, \omega) \exp(ik_{x} X_s + i k_{y} Y_s - i \omega t), \label{eq:hfourier}$$ where $X_s = x - \bm{\rho}_s \cdot \nabla x $ and $Y_s = y - \bm{\rho}_s \cdot \nabla y$ are guiding center variables. In the next section we present the profiles for the JET shot that we are examining, as well as an overview of the gyrokinetic results. These gyrokinetic results will motivate the work for the rest of the paper. Pedestal Gyrokinetic Simulations of JET Shot 92174 {#sec:jetprofiles} ================================================== In this section, we present the significant linear microstability features of a single JET-ILW inter-ELM pedestal discharge at a single radial location. This equilibrium exhibits properties such as temperature, magnetic geometry, injected neutral beam power, and fueling that are typical for JET-ILW inter-ELM H-mode pedestals: key experimental parameters for this discharge are $I_p = 1.4$ MA, $B_{T0} = 1.9$ T, $H_{98(y,2)} = 1.0$, $n_G = 0.7$, $P_{\mathrm{NBI}} = 17.4$ MW, $\beta_N = 2.5$, and $R_D = 0.9 \times 10^{22}$ electrons/s. Here, $I_p$ is the poloidal current, $B_{T0}$ is the toroidal magnetic field at $R = 2.96$m, $H_{98(y,2)}$ is the H factor relative to the IPB98${}_{(y,2)}$ scaling [@Doyle2007], $n_G$ is the Greenwald density fraction [@Greenwald1988] defined as the line averaged density divided by the Greenwald density limit, $P_{\mathrm{NBI}}$ is the neutral beam injection power, $\beta_N$ is the normalized $\beta$ factor [@Troyon1984], and $R_D$ is the deuterium electron flow rate. In , we show the pedestal equilibrium temperature, density, and flow profiles, which will have significant implications for microstability. In , we present an overview of linear results from gyrokinetic simulations, run both with and without finite $\beta$ effects. From these results, we justify an electrostatic study. Here, we find a range of modes, including an unusual toroidal ETG instability that is driven at a very wide range of perpendicular scales, and has a radial wavenumber that is typically much larger than its poloidal wavenumber. A significant portion of the paper will be devoted to understanding this mode. We show that this mode is largely unaffected by finite $\beta$ effects and $\mathbf{ E} \times \mathbf{ B}$ shear, and in subsequent sections, that it could play an important role in transport. Finally, in , we present the prominent features of the electrostatic growth rate spectrum. JET-ILW Profiles {#subsec:JETprofiles} ---------------- In this paper, we focus on simulation results from JET shot 92174. We run linear gyrokinetic simulations with a single deuterium ion species and no impurities, assuming that $n_{0e} = n_{0i}$ (note that experimentally $Z_{\mathrm{eff}} = 1.8$, where $Z_{\mathrm{eff}} = \sum_i n_i Z_i^2 / n_e$). The three other pedestals that we have analyzed (82550, 92167, 92168) give qualitatively similar results, which is notable, given that the nature of these discharges varies quite significantly. The experimental and simulation parameters and linear gyrokinetic growth rates for these additional three discharges are shown in \[app:dischargeparams\]. ![Pedestal profiles and their gradients for JET shot 92174. Crosses indicate simulation location of $r/a = 0.9743$. (a): Ion and electron temperatures profiles. (b): Density profiles. (c): Flow profiles. (d): Temperature and density gradients profiles. (e): $\eta_s$ profiles, where the parameter $\eta_s$ is defined as $\eta_s \equiv L_n/L_{Ts}$. (f): $\mathbf{ E} \times \mathbf{ B}$ shear profiles.[]{data-label="fig:1profiles"}](JETelectronandiontempgradientandtempsANDflows92174ITORwithetasMar30threebytwo.eps){width="\textwidth"} The temperature and density profiles for shot 92174 and associated gradients, are shown in (a) as functions of $r/a$. The distance $a$ is the value of $r$ at the last closed flux surface (LCFS). In (d), we also show the toroidal velocity of ${}^{12}_6 C^+$, $u_{\zeta C}$, at the outboard midplane, normalized by the ion thermal speed $v_{ti} = \sqrt{2 T_{0i}/m_i}$. We assume that this velocity is a good proxy for the toroidal ion velocity, $u_{\zeta i} $. We normalize the gradient length scales using the major radius of the last closed flux surface, $R_0$, which is the radial distance to the center of the last closed flux surface at the midplane. The profiles in are consistent with an emerging JET-ILW pedestal paradigm [@Leyland2015; @Hatch2017; @Kotschenreuther2017; @Hatch2019], whereby enhanced gas puffing reduces the edge density gradient [@Wolfrum2017] and shifts the density pedestal outwards [@Hillesheim2016; @Dunne2017], making microinstabilities more virulent [@Maggi2017]. Weaker density gradients also reduce the $\mathbf{ E} \times \mathbf{ B}$ shear, which has often been observed to be important for microinstability suppression in the pedestal [@Hatch2017; @Kotschenreuther2017; @Hatch2019]. It is hypothesized that heat transport from more strongly-driven microinstabilities with less shear suppression is responsible for a reduced temperature at the pedestal top [@Hatch2019]. In this work, the electron temperature and density are determined from the High Resolution Thomson Scattering profiles [@Pasqualotto2004; @Frassinetti2012]. To improve the data statistics, a composite profile is constructed from profiles collected in a time window of 80-99% in the ELM interval period. The profiles of the ion temperature and rotation are measured with the edge Charge Exchange Recombination Spectroscopy diagnostic [@Delabie2016] for fully stripped carbon-12 (${}_6^{12}C^+$), with a time integration of 7.2ms. These ion profiles are collected on a longer 60-99 % ELM interval period time window. The ${}_6^{12}C^+$ and ion temperature and rotation profiles in the pedestal can differ substantially, as found in some recent DIII-D experiments [@Haskey2018; @Haskey2018b; @Camenen2010]. Since the ITG instability is sensitive to $T_{0i}$ and $R_0/L_{Ti}$, the ITG instability results in should be viewed in the context of potentially large uncertainties in ion temperature measurements, which might significantly underestimate the ion temperature gradient. For this reason, while we have mainly used $T_{0i} > T_{0e}$ and $R_0/L_{Te} > R_0/L_{Ti}$ in our simulations and theory, we have also explored the impact on gyrokinetic microinstabilities of assuming $T_{0i} = T_{0e}$ and $R_0/L_{Ti} = R_0/L_{Te}$, which can be found in . However, unless explicitly mentioned otherwise, we use the measured ion temperature profiles. To obtain an estimate for the radial electric field, we use the most general ion flow [@Hinton1976; @Helander2002], $$\mathbf{ u}_i = - c \frac{\partial \phi_0 }{\partial \psi} R^2 \nabla \zeta - \frac{c}{Z_i e n_{0 i}} \frac{\partial p_{0i}}{\partial \psi} R^2 \nabla \zeta + \frac{ \mathbf{B}}{n_{0i} } K_i (\psi) \frac{\partial T_{0i}}{\partial \psi}. \label{eq:lowflowequation}$$ Here, $R$ is the major radius, and the flux function, $K_i(\psi)$, is determined by neoclassical theory [@Hinton1976; @Helander2002]. Based on the experimental data in , we find that $u_{\zeta C} \lesssim (\rho_{P i}/L_{Ti}) v_{ti}$. The quantity $\rho_{P s} = (B/B_P) \rho_s$ is the poloidal gyroradius for a species $s$, where $B_P$ is the poloidal magnetic field strength. Thus, the flow velocity of the ${}^{12}_6C^+$ impurity species is comparable to the size of the ion diamagnetic flow, $u_{\zeta i p}$, $$\frac{u_{\zeta i p}}{v_{ti}} = - \frac{R c}{Z_i e n_{0i} v_{ti}} \frac{\partial p_{0i}}{\partial \psi} \sim \frac{\rho_{Pi}}{L_{pi}} \sim \frac{1}{3}. \label{eq:uzeta}$$ Note that this implies that there are only several poloidal gyroradii in a pressure length scale, $L_{pi}$. To obtain a rough estimate of the radial electric field, we use the fact that the measurement of $u_{\zeta_i}$ suggests that the overall flow, the $\mathbf{ E} \times \mathbf{ B} $ flow, the diamagnetic flow in , and the term proportional to $K_i (\psi)$ are all of the same order. Thus, $$- \frac{\partial \phi_0}{\partial \psi} \approx \frac{1}{Z_i e n_{0i}} \frac{\partial p_{0i}}{\partial \psi}. \label{eq:pressgradientradEfield}$$ Then, the radial shear in the $\mathbf{ E} \times \mathbf{ B} $ rotation, $\gamma_E (\psi)$, is approximately $$\gamma_E \equiv - \frac{c r}{q} \frac{\partial}{\partial r} \bigg{(} \frac{\partial \phi_0}{\partial \psi} \bigg{)} \approx \frac{r}{q} \frac{\partial }{\partial r} \bigg{(} \frac{c}{Z_i e n_{0i}} \frac{\partial p_{0i}}{\partial \psi} \bigg{)}. \label{eq:gammaEapproximation}$$ The location of the simulations was chosen to have equilibrium length scales characteristic of the steep gradient region in the pedestal, and an $\mathbf{ E} \times \mathbf{ B}$ shear value close to the maximum possible for a given equilibrium, using the estimate in . The radial location for JET shot 92174, shown in , is $r/a = 0.9743$. To simulate this discharge, we use the following simulation parameters: $\rho_i = 0.27 \; \mathrm{cm}, \; \nu_{ee} a/ v_{ti} = 0.14, \; a / L_{Te} = 42, \; a/ L_{Ti} = 11, \; a / L_{n} = 10, \; \rho_i / L_{Te} = 0.12, \; T_{0e}/T_{0i} = 0.56, \; \hat{s} = 3.36, \; q = 5.1, \; R_0 = 2.86$ m, $a = 0.91$ m, $R_c = 2.91$ m, and $r_c = 0.89$ m, where $\nu_{ss'} = \sqrt{2} \pi n_{0s'} Z_s^2 Z_{s'}^2 e^4 \ln (\Lambda_{ss'}) / \sqrt{m_s} T_{0s}^{3/2} $, $\ln (\Lambda_{ss'})$ is the Coulomb logarithm, and $\hat{s} = (r / q) \partial q / \partial r$ is the magnetic shear. In the instances where we included $\mathbf{ E} \times \mathbf{ B}$ shear and electromagnetic effects, we used $\gamma_E a / v_{ti} = 0.56$ and $\beta = 0.0031$. Here, the quantity $\beta = 8 \pi (p_{0i} + p_{0e})/B_a^2$, where $B_a = 1.99$ T for this equilibrium. Gyrokinetic Simulation Results {#subsec:gyroresults} ------------------------------ In this section, we present results obtained from linear gyrokinetic simulations (both electromagnetic and electrostatic) for this radial location and pedestal. For the chosen pedestal and radial location, we will establish that linear *electrostatic* simulations without $\mathbf{ E} \times \mathbf{ B}$ shear give similar growth rate spectra to linear *electromagnetic* simulations with $\mathbf{ E} \times \mathbf{ B}$ shear. The electrostatic limit of is taken by requiring that the turbulent electric field is primarily electrostatic, $|\nabla \overline{\phi}_1^{tb}| \gg (1/c) | \partial \overline{A}_{\parallel 1} ^{tb}/ \partial t | $, and that the turbulent magnetic field is small, $| \mu \overline{B} ^{tb}_1 | \ll | Z_s \overline{\phi} ^{tb}_1| e / m_s$ [^1]. It is no coincidence that the electrostatic regime without $\mathbf{ E} \times \mathbf{ B}$ shear and the electromagnetic case with $\mathbf{ E} \times \mathbf{ B}$ shear give similar results; electromagnetic modes are suppressed by $\mathbf{ E} \times \mathbf{ B}$ shear, leaving electrostatic modes that are unaffected by $\mathbf{ E} \times \mathbf{ B}$ shear as the dominant instabilities. Therefore, it is reasonable to study this pedestal with linear electrostatic simulations without $\mathbf{ E} \times \mathbf{ B}$ shear. We will choose to study the electrostatic limit without $\mathbf{ E} \times \mathbf{ B}$ shear rather than an electromagnetic case with $\mathbf{ E} \times \mathbf{ B}$ shear because the former is analytically and numerically simpler. We now proceed to give an overview of gyrokinetic results for the electrostatic pedestal. We performed these local simulations in ballooning space, which can be represented in a flux-tube [@Beer1995]. Because the novel toroidal ETG instability we have found is often driven at large distances along the field line from $\theta = 0$, we require a large range of $\theta$ values, and hence we typically choose a flux-tube with 64 gridpoints in each $2\pi$ period in $\theta$, with nine periods. This is equivalent to a ballooning space calculation extending to nine poloidal turns in the extended ballooning angle. The standard velocity space grid had 20 passing pitch angles, 17 trapped pitch angles, and 12 energy gridpoints [@Barnes2010]. Resolution scans were performed in all of these parameters. ![The Miller equilibrium and numerical equilibrium for JET shot 92174 used for gyrokinetic simulations. (a): Equilibrium and Miller flux surfaces in $R_M, Z_M$ space, (b): Equilibrium and Miller poloidal magnetic field versus $\theta_M$, (c): Equilibrium toroidal and magnetic fields.[]{data-label="fig:Millerfits"}](millerfitsBtorandBpol.eps){width="90.00000%"} While GS2 is capable of reading in numerical equilibria, we fit the magnetic equilibrium with Miller geometry. A Miller equilibrium is a prescription to generate flux surfaces that satisfy the Grad Shafranov equation locally by fitting to nine parameters [@Miller1998]. In we show the difference between the exact flux surface at $r/a = 0.9743$ and the Miller fits that we use. The Miller angle $\theta_M$, which is in general not equal to the poloidal angle $\theta$ defined in , is given by $\theta_M=\arcsin(Z/\kappa r)$, where $\kappa$ is the flux surface elongation, $R_M$ is the Miller major radius, and $Z_M$ is the Miller distance above the midplane [@Miller1998]. The Miller parameters for this radial location are $\Delta = dR_M /d r = -0.345$, $\kappa = 1.55$, $a (d\kappa/dr) = 0.949$, $ \delta = 0.263$, $a (d \delta/ dr) = 0.737$, $\beta' = \beta a (d \ln p_0 / d r) = -0.161$, where $\delta$ is the triangularity. Electromagnetic effects have been shown to be important for microinstability in the pedestal [@Dickinson2012; @Hatch2017; @Hatch2016; @Hatch2019; @Kotschenreuther2019]. While we have neglected electromagnetic effects in most of this study, we have scoped out the potential effects of nonzero $\beta$. As an initial study, this is well-justified since we will show that a linear electromagnetic gyrokinetic simulation with $\mathbf{ E} \times \mathbf{ B}$ shear gives similar results to a linear *electrostatic* gyrokinetic simulation without $\mathbf{ E} \times \mathbf{ B}$ shear. To demonstrate this equivalence, we first show the results of gyrokinetic simulations with and without finite $\beta$ effects in . To include finite $\beta$ effects, we included values of $\beta$ and $\beta'$ consistent with the Miller equilibrium. ![(a): GS2 growth rate ($\gamma$) and (b): real frequency ($\omega_R$) for JET shot 92174 with $\theta_0 = 0$ with and without finite $\beta$. (c): eigenmodes for $k_y \rho_i = 0.2$. (d): growth rates for an electromagnetic simulation with different $\theta_0$ values at $k_y \rho_i \sim 0.1$.[]{data-label="fig:electromagneticspectraearly"}](zoomedgs2realfreqandgrowthratesandeigenmodesshotbetascan92174annotationsJan15twobytwo.eps){width="100.00000%"} In , we show the effect of finite $\beta$ on the growth rates (a), real frequencies (b), and eigenmodes (c) for $\theta_0 = 0$, where $\theta_0$ is the ballooning angle, defined as $\theta_0 = k_x / \hat{s} k_y$. Throughout this paper, the eigenmodes are separately normalized such that $|\phi ^{tb} _1|$ has a maximum of 1. When finite $\beta$ effects are included, a KBM appears, as shown by the small bump at $k_y \rho_i \sim 0.1$ in (a) of the growth rates. This KBM has a standard ballooning eigenmode structure, centered at $\theta = \theta_0 = 0$. However, when $\beta = 0$, there is no KBM, and instead at $k_y \rho_i \sim 0.1$ there are modes with a much lower growth rate and a complicated mode structure in $\theta$ (see (c)). These eigenmodes have maxima in bad curvature regions and have tearing parity in both $\mathrm{Re}(\phi ^{tb} _1)$ and $\mathrm{Im}(\phi ^{tb} _1)$. More details regarding these long wavelength electron modes can be found in \[sec:smallkrimodes\]. Much of the rest of the growth rate spectrum is quite unaffected by finite $\beta$ effects. At $k_y \rho_i \approx 1-5$ for $\theta_0 = 0$, there is a peculiar bump in (a), whose corresponding instability will be the focus of much of this paper. We identify this mode as toroidal ETG. We have undertaken extensive tests described later in to confirm that it is a novel type of toroidal ETG; for now, we will refer to it as a toroidal ETG mode without justification. Finally, for $k_y \rho_i \gtrsim 5$ and $\theta_0 = 0$, the fastest growing mode becomes a slab ETG mode, which again, we will justify later in . Clearly the toroidal ETG mode is unaffected by finite $\beta$, and the slab ETG growth rates decrease by roughly 20%, but the mode structure is qualitatively the same. Thus, apart from the KBM, the electromagnetic and electrostatic growth rates and modes are very similar. Once $\mathbf{ E} \times \mathbf{ B}$ shear is included in the simulations, the electromagnetic and electrostatic growth rate spectra become qualitatively the same. This is because $\mathbf{ E} \times \mathbf{ B}$ shear is found to easily suppress the KBM. Recall that the KBM is the main difference between the electromagnetic and electrostatic simulations without $\mathbf{ E} \times \mathbf{ B}$ shear. Further evidence for the effectiveness of the $\mathbf{ E} \times \mathbf{ B}$ shear for suppressing the KBM is that the KBM is stable for all $|\theta_0| > \theta_{0c} \approx 0.5$, as shown in (d), where we show the growth rates for a range of $\theta_0$ values at scales $ 0.01 <k_y \rho_i < 0.3$ in a simulation with finite $\beta$. The dependence on $\theta_0$ is important because $\mathbf{ E} \times \mathbf{ B}$ shear causes a mode’s radial wavenumber to vary with time as $\Delta k_x = k_y \gamma_E t$, giving a change of $\theta_0$ of $\Delta \theta_0 = \gamma_E t / \hat{s}$. If a mode is shown to be unstable only for a very narrow range of $\theta_0$ values, $|\theta_0| < |\theta_{0c}|$, it is highly susceptible to $\mathbf{ E} \times \mathbf{ B} $ shear because in a time of order $1/\gamma_E$ its $\theta_0$ changes significantly. After a time $t_{C} \sim \hat{s} \theta_{0c}/ \gamma_E$, $\mathbf{ E} \times \mathbf{ B}$ shear will have suppressed the KBM; in our simulations, $t_C \approx 3$. Thus, to suppress instability we require $\gamma t_C \lesssim 1$, leading to $\gamma_E / \hat{s} \gamma \gtrsim \theta_{0c}\approx 0.5$. We will discuss the $\mathbf{ E} \times \mathbf{ B}$ shear and its effects on all the other instabilities we find in more detail in . Finally, the perpendicular wavenumber of the KBM is close to the limit where local simulations are valid, which is when $k_{\perp} \rho_i \gg 0.12$, and hence, results from our KBM simulations should be viewed in the context of uncertainties that are present due to being close to the local approximation. Linear Features of the Electrostatic Pedestal {#subsec:linearfeatures} --------------------------------------------- In this section, we describe the most prominent features of the electrostatic growth rate spectrum. ![(a): electrostatic growth rates for 2 values of $\theta_0$. (b): eigenmodes for 2 values of $k_y \rho_i$ at $\theta_0 = 0$. (c): eigenmodes for 2 values of $k_y \rho_i$ at $\theta_0 = 0.5$.[]{data-label="fig:electrostaticspectra"}](92174growthratetheta0Presentationreveal3paperplotDec21.eps){width="100.00000%"} A notable feature of the growth rate spectrum shown earlier in is the bump at $k_y \rho_i \approx 1 - 5$ in (a), which we claimed was a novel toroidal ETG instability. In (a), we show the growth rates for two values of $\theta_0$. Focusing first on $\theta_0 = 0$, we again identify the bump at $k_y \rho_i \approx 1 - 5$, which has a peak growth rate at $k_y \rho_i \simeq 3$. Once $k_y \rho_i \gtrsim 5$, the mode switches to a slab ETG instability. In (b), we show the eigenmodes for two $k_y \rho_i$ values in the $\theta_0 = 0$ growth rate spectrum, one at $k_y \rho_i = 2.4$ (near the top of the toroidal ETG bump) and one at $k_y \rho_i = 51.4$. The eigenmode associated with $k_y \rho_i = 2.4$ is fairly localized at large $\theta$, whereas the eigenmode associated with $k_y \rho_i = 51.4$ is centered at $\theta = 0$ and has a large parallel wavenumber. The $k_y \rho_i = 2.4$ mode is the novel toroidal ETG mode, and the $k_y \rho_i = 51.4$ mode is a slab ETG mode. In our up-down symmetric equilibrium fit, there is a subtlety for the novel toroidal ETG eigenmodes when $\theta_0 = 0$: there are two independent modes that grow at the same rate, and are localized at opposite signs of $\theta$. Indeed, for toroidal ETG, there must be two independent modes with $\theta_0 = 0$, since the gyrokinetic equation is invariant under the transformation $\theta \to - \theta$, $\theta_0 \to - \theta_0$ [@Peeters2005]. Thus, henceforth, when plotting the eigenmodes for $\theta_0 \simeq 0$, we choose a small value of $\theta_0$, $\theta_0 = 0.05$, which causes the mode at one location to grow slightly faster than the mode at the other, but barely changes the growth rate compared with $\theta_0 = 0$. This results in a well-defined single eigenmode, like the one in (a), rather than two separate modes, like the ones shown in (b). The relative size and phase of the modes at opposite values of $\theta$ depend on the initial condition. To distinguish between the toroidal and slab ETG modes in (a) and (a), we used a set of criteria discussed extensively in . Briefly, the toroidal ETG eigenmodes are localized far along the field line for smaller $k_y \rho_i$ values, and are at a $\theta$ location with the opposite sign of $\theta_0$ for larger $k_y \rho_i $ values. Sensitivity scans to equilibrium parameters, shown in , reveal that the slab and toroidal ETG branches have different dependences on parameters such as $R_0/L_{Ti}$ and $R_0/L_{n}$. For a given $k_y \rho_i$, slab ETG modes also tend to have a much larger $k_{\parallel}$ than toroidal ETG modes. While the novel toroidal ETG mode is the fastest growing instability for $1 \lesssim k_y \rho_i \lesssim 5$ when $\theta_0 = 0$, we find that when $\theta_0$ differs slightly from 0, the toroidal ETG mode is the fastest growing for $1 \lesssim k_y \rho_i \lesssim 400$. We show a simple example of the growth rate spectrum for $\theta_0 = 0.5$ in (a), where the toroidal ETG mode is the fastest growing mode for that particular value of $\theta_0$ for all $k_y \rho_i \gtrsim 1$. In (c), we show the eigenmodes for $\theta_0 = 0.5$ for $k_y \rho_i = 2.4$ and $k_y \rho_i = 51.4$. For $k_y \rho_i = 2.4$, the eigenmodes for $\theta_0 = 0$ and $\theta_0 = 0.5$ have a similar structure, both being localized at $|\theta| \simeq 8$. However, the eigenmode at $k_y \rho_i = 51.4$ is dramatically different to the $\theta_0 = 0$ mode at $k_y \rho_i = 51.4$; the eigenmode for $\theta_0 = 0.5$ is localized at $\theta \simeq -1$, and has, in fact, the same novel toroidal ETG character we identified earlier. In we will explain these toroidal ETG modes in much more detail, including the reasons why they move in $\theta$ for different values of $k_y \rho_i$, as evidenced by the eigenmodes for $\theta_0 = 0.5$ at $k_y \rho_i = 2.4$ and $k_y \rho_i = 51.4$. For completeness, we briefly describe the modes we find at larger scales. For this JET discharge and the surface $r/a = 0.9743$, we find that the instabilities are electron-driven between $0.005 \lesssim k_y \rho_i \lesssim 400$. For $0.005 \lesssim k_y \rho_i \lesssim 0.1$ the modes have electron tails similar to those described in [@Hallatschek2005], and for $0.1 \lesssim k_y \rho_i \lesssim 1.0$, there are complicated modes that appear to be a form of ETG we do not yet fully understand. Both the electron tails and complicated ETG modes will be excluded from in-depth analysis in the main text, but are described in \[sec:smallkrimodes\]. In the next section, we introduce the theory needed to understand these novel toroidal ETG modes as well as the slab ETG modes at $k_y \rho_i \gtrsim 1$. We will see that the existence of these modes follows naturally from the steep temperature gradients in pedestals. Linear Gyrokinetics With Large Gradients {#sec:linGKwithlargegrads} ======================================== In this section, we analyze the consequences of large equilibrium gradients for linear collisionless electrostatic gyrokinetic stability, which will considerably change the character of the toroidal ETG instability. We have already motivated the local and linear limits in , and the electrostatic limit in . We now motivative the *collisionless* limit of the electron gyrokinetic equation, which will be used for the theoretical analysis. The collisionless limit for electrons is justified by the small electron collision frequency, $\nu_{ee} \ll v_{te} / q R $. For JET shot 92174 at $r/a = 0.9743$, $\nu_{ee} \simeq 4 \times 10^{4}$ Hz, and $v_{te} / qR \simeq 3 \times 10^{5}$ Hz. In gyrokinetic simulations, we found ETG instabilities to be insensitive to whether collisions are kept. However, for ITG scale instabilities at lower frequencies, electron collisions can decrease the ITG growth rates and cause electrons to be non-adiabatic, as we will see in . Using the equations laid out in , we take the linear electrostatic collisionless local limit of the gyrokinetic equation in . Analytically and computationally, this limit is more straightforward, and includes key elements of the pedestal microinstability linear physics that we wish to explain. Motivated by the steep pedestal gradients, we explore the implications of steep equilibrium temperature gradients on ETG instability in . Simple arguments based on balancing terms with the same order of magnitude reveal how these steep gradients affect the perpendicular scales of the instability and how magnetic shear determines the parallel toroidal ETG mode structure, allowing the toroidal ETG mode to compete with the slab ETG mode. In , we convert the gyrokinetic equation derived in to an algebraic equation in order to analyze slab and toroidal ETG instabilities in the presence of large equilibrium gradients. This is then used to derive an analytical ETG dispersion relation that supports our simplified arguments. Electrostatic Collisionless Local Limit {#sec:electrostatcollisionlesslocal} --------------------------------------- In this section, we take the electrostatic, linear, collisionless form of the gyrokinetic equation. In this limit, is $$\fl \eqalign{ \frac{\partial \overline{ h}_s}{\partial t} & + v_{\parallel} \hat{\mathbf{b} } \cdot \nabla_{\mathbf{R}_s} \overline{ h}_s + \mathbf{v}_{Ms} \cdot \nabla_{\mathbf{R}_s} \overline{ h}_s = \frac{Z_s e F_{Ms}}{T_{0s}} \frac{\partial \langle \overline{\phi} ^{tb}_1 \rangle}{\partial t} \\ & + \frac{c}{B} (\nabla_{\mathbf{R}_s } \langle \overline{ \phi}^{tb}_1 \rangle \times \hat{\mathbf{b}} ) \cdot \nabla r \bigg{[} \frac{\partial \ln n_s}{\partial r} + \frac{\partial \ln T_s}{\partial r} \bigg{(} \frac{m_s \mathcal{E}}{T_{0s}} - \frac{3}{2} \bigg{)} \bigg{]} F_{Ms}. } \label{eq:GKgeneral}$$ We have absorbed the toroidal mean flow in the convective derivative as a constant Doppler shift, and neglected the equilibrium $\mathbf{ E} \times \mathbf{ B}$ shear, which is consistent with the low flow ordering in , and is justified in with simulation results. Substituting the expressions for $\phi^{tb}_1$ and $h_s$ in into gives a Fourier-analyzed gyrokinetic equation, $$\fl \eqalign{ - i \omega h_s & + \frac{2\pi v_{\parallel }}{L_{\theta}} \frac{\partial h_s}{\partial \theta} + i \mathbf{v}_{Ms} \cdot \mathbf{k}_{\perp} h_s = - i \omega \frac{Z_s e F_{Ms}}{T_{0s}} \phi ^{tb}_1 J_0 \bigg{(} \frac{k_{\perp} v_{\perp}}{\Omega_s} \bigg{)} \\ & + i \omega_{*s} \bigg{[} 1 + \eta_s \bigg{(} \frac{m_s \mathcal{E}}{T_{0s}} - \frac{3}{2} \bigg{)} \bigg{]} \frac{Z_s e F_{Ms}}{T_{0s}} \phi ^{tb}_1 J_0 \bigg{(} \frac{k_{\perp} v_{\perp}}{\Omega_s} \bigg{)}, } \label{eq:GKgeneralFourierAnalyzednotkpll}$$ where $J_0$ is a Bessel function of the first kind that comes from gyroaveraging $\overline{ \phi }^{tb} _1$. The perpendicular wavenumber $\mathbf{k}_{\perp}$ is $$\eqalign{ \mathbf{k}_{\perp} = k_x \nabla x + k_y \nabla y = & \bigg{[}k_x - k_y\bigg{(}\hat{s} \theta - \frac{r}{q} \frac{\partial \nu}{\partial r} \bigg{)} \bigg{]} \nabla x \\ & + \frac{\partial \psi}{\partial r} \frac{1}{B_a} k_y \bigg{[} \nabla \zeta + \bigg{(}\frac{\partial \nu }{\partial \theta} - q\bigg{)} \nabla \theta \bigg{]}, } \label{eq:kperpxywithshear}$$ where every function is evaluated at $r_c$. We have also introduced the drift frequency, $\omega_{*s}$, and the stability parameter, $\eta_s$, $$\omega_{*s} \equiv - \frac{c}{B} \frac{T_{0s}}{Z_s e L_{ns}} (\mathbf{ k}_{\perp} \times \hat{\mathbf{b} } )\cdot \nabla r = \frac{c}{B_a} \frac{T_{0s}}{Z_s e L_{ns}} k_y, \;\;\;\; \eta_s \equiv \frac{ L_{ns}}{ L_{Ts}}.$$ Note that the factor $ (\mathbf{ k}_{\perp} \times \hat{\mathbf{b} } )\cdot \nabla r$ in $\omega_{*s} $ is only proportional to $k_y$. The system of equations is closed by the first order turbulent quasineutrality condition in , $$\eqalign{ \frac{e \phi^{tb}_1 n_{0e}}{T_{0e}} \bigg{(} \frac{Z_i T_{0e}}{T_{0i}} +1 \bigg{)} & + 2\pi \int \frac{B}{|v_{\parallel }|} h_e J_0 \bigg{(} \frac{k_{\perp} v_{\perp}}{\Omega_e} \bigg{)} d \mathcal{E} d \mu \\ & - 2\pi \int \frac{B}{|v_{\parallel }|} h_i J_0 \bigg{(} \frac{k_{\perp} v_{\perp}}{\Omega_i} \bigg{)} d \mathcal{E} d \mu = 0, } \label{eq:turnQNexplicit}$$ where we used that the Jacobian of the gyrokinetic transformation is $\mathcal{J} = \partial (\mathbf{r}, \mathbf{v} )/ \partial(\mathbf{R}, \mathcal{E}, \mu, \varphi ) \simeq B/ |v_{\parallel }|$ [@Parra2008]. We proceed to demonstrate how the presence of large equilibrium gradients changes the perpendicular scales at which ETG can be strongly driven, and how in the presence of these steep gradients, magnetic shear can act to determine the poloidal location where the ETG mode has its maximum amplitude. Slab Versus Toroidal ETG In Large Gradient Regions {#sec:slabversustoroidalETG} -------------------------------------------------- In this section, we describe a novel type of toroidal ETG with anisotropic perpendicular wavenumbers. contains two branches of electron temperature gradient driven instability, slab [@Rudakov1961; @Coppi1967] and toroidal [@Coppi1977; @Horton1981]. These modes have been covered extensively [@Jenko2000; @Rudakov1961; @Coppi1967; @Coppi1977; @Horton1981; @Cowley1991]. Here, we give a very brief overview. In the slab branch, the density perturbation is caused by a competition between the parallel streaming and the radial $\mathbf{ E} \times \mathbf{ B} $ drift. For sufficiently large $\eta_s$, a large parallel compression causes $\phi ^{tb}_1$ to grow in time. For smaller values of $\eta_s$, the radial $\mathbf{ E} \times \mathbf{ B} $ drift term dominates and we obtain stable electron drift waves. The toroidal instability is caused by magnetic drifts, rather than parallel streaming, creating a compression that again, gives rise to a destabilizing electric field for sufficiently large $\eta_s$. In both cases, at the onset of instability, increasing the temperature gradients causes the linear instability to be more virulent. Motivated by the large temperature gradients in (b), we proceed to demonstrate that $$\frac{R_0}{L_{Te}}, \frac{R_0}{L_{Ti}} \gg 1,$$ has major implications for ETG stability. First, we present an intuitive, albeit non-rigorous argument that will turn out to be incorrect. We then develop a more careful argument, which reveals the distinctive new character of ETG modes in steep gradients, which is very different to the more familiar lower gradient regime typical of the core. Throughout this section, we shall assume that $\theta_0 = 0$. We will investigate the physics of $\theta_0 \neq 0$ in . First, we present the intuitive, albeit incorrect argument. For the electrons, since $R_0 / L_{Te} \gg 1$, we naively expect that the ratio determining the relative strength of the drive frequency to the magnetic drift frequency to be large. Therefore, in the pedestal, one might naively think that the drive for toroidal ETG is weak and independent of $k_{\perp}$, $$\frac{\omega_{*e} \eta_e }{\mathbf{ v}_{Me} \cdot \mathbf{ k}_{\perp}} \sim \frac{R_0}{L_{Te}} \gg 1. \label{eq:drivetodriftstrength}$$ Here, we use $\mathbf{ v}_{Me} \cdot \mathbf{ k}_{\perp} \sim k_{\perp} v_{te}^2/ \Omega_e R_0$ and $k_y \sim k_{\perp}$. Comparing the size of the drive frequency to the parallel streaming frequency, we obtain $$\frac{\omega_{*e} \eta_e}{k_{\parallel} v_{te}} \sim \frac{k_y}{k_{\parallel} } \frac{\rho_e}{L_{Te}}. \label{eq:drifttoparallelstreamelectron}$$ As we will show in , the ratios in must be of order unity for a large toroidal and slab ETG growth rate, respectively (see ). Thus, suggests that the magnetic drifts are small for every $k_{\perp}$, whereas in , $k_{\parallel} $ can become large to drive slab instability. One would therefore expect slab ETG to be the dominant electron microinstability at all scales. The above argument, however, suffers from a deficiency. It is naive to make the assumption $\omega_{*e} \eta_e / \mathbf{ v}_{Me} \cdot \mathbf{ k}_{\perp} \sim R_0/L_{Te}$ (see ) in the presence of magnetic shear, because $\mathbf{k}_{\perp}$ varies along a field line (see ). At large values of $|\theta|$, the *radial* component of the magnetic drift frequency becomes increasingly large and can compete with the linear drive $\omega_{*e} \eta_e$, to allow the toroidal branch to become unstable. Toroidal modes, with $\mathbf{ v}_{Me} \cdot \mathbf{ k}_{\perp} \sim \omega_{*e} \eta_e$, are therefore possible because the competition between the slab and toroidal modes has a $\mathbf{k}_{\perp}$ dependence, which arises from the fact that $\mathbf{ v}_{Me} \cdot \mathbf{ k}_{\perp}$ depends on both $k_x$ and $k_y$, whereas $\omega_{*e} $ only depends on $k_y$. For convenience, we define the radial component of $\mathbf{k}_{\perp} $ in as $$K_{x} = k_x - k_y \bigg{(} \hat{s} \theta - \frac{r}{q} \frac{\partial \nu}{\partial r} \label{eq:effectivewavenumber} \bigg{)}.$$ We now show that toroidal ETG modes with $k_{\perp} \sim K_x \gg k_y$ can indeed compete with the slab ETG at sufficiently small $k_y \rho_i $. Motivated by the eigenmodes in that are localized far along a field line, we will make $K_x$ large by taking $\hat{s} \theta \gg k_x/k_y = \hat{s} \theta_0$ and $\hat{s} \theta \gg (r/q) \partial \nu / \partial r$. Thus, for $\hat{s} \theta$ large, we find $$k_{\perp} \sim K_{x} \sim k_y \hat{s} \theta.$$ Hence, for $\hat{s} \theta \gg 1$, the magnetic drift term that drives toroidal ETG can become comparable to the drive term, $$\frac{\omega_{*e} \eta_e }{\mathbf{ v}_{Me} \cdot \mathbf{ k}_{\perp}} \sim \frac{k_y}{k_{\perp}} \frac{R_0}{L_{Te}} \sim \frac{1}{\hat{s} \theta} \frac{R_0}{L_{Te}} \sim 1, \label{eq:sizeofomsteetaetoomegakappae} \label{eq:shatrlte}$$ Thus, for sufficiently small $k_x$, the toroidal mode must be driven far along the field line, $$\hat{s} \theta \sim \frac{R_0}{L_{Te}} \gg 1. \label{eq:shatthetaRLTerln}$$ Through detailed analysis in later sections, we will indeed see that this explains the toroidal ETG modes, which are often unstable at large distances along the field line (see ). Recall that here $\theta$ is the ballooning angle, which has a range $-\infty<\theta<\infty$. When is satisfied, we will demonstrate with a local gyrokinetic dispersion relation in that when $\mathbf{v}_{Me} \cdot \mathbf{k}_{\perp} \sim \omega_{*e} \eta_e $, the toroidal ETG growth rate becomes comparable to the slab ETG growth rate. This would seem to suggest that toroidal ETG exists for all $k_y$. However, for large $k_y$ and small $k_x$, $k_{\perp} \rho_e \sim \hat{s} \theta k_y \rho_e $ becomes so large that finite Larmor radius (FLR) effects from the electron gyromotion become important. Thus, if $R_0/L_{Te} \gg 1$ and $\hat{s} \theta \gg 1$, for strongly driven toroidal ETG, $K_x$ has a maximum of the order of $$K_x \rho_e \sim \hat{s} \theta k_y \rho_e \sim 1. \label{eq:effectiveKXsize}$$ If $K_x \rho_e $ is much larger than in , then the growth rate will be strongly electron FLR damped. Motivated by , for a toroidal mode we expect ion FLR damping to be very strong at $k_y \rho_e \ll 1$ with $k_{\perp} \rho_e \sim 1$. Thus, our analytic treatment of toroidal ETG will assume $h_i = 0$ because $|J_0 (k_{\perp} \rho_i)| \ll 1$. Using , we obtain a scale for $k_y$, $$k_y \rho_e \sim \frac{L_{Te}}{R_0}. \label{eq:kyrhoeLTeRscale}$$ Given that the pedestal profiles have $R_0/L_{Te} \gtrsim \rho_i / \rho_e$ in the steep pedestal regions, toroidal ETG can be unstable even at scales as large as $k_y \rho_i \lesssim 1$. Therefore, $R_0/L_{Te} \gg 1$ extends the minimum $k_y$ scale at which toroidal ETG modes can be strongly driven to ion gyroradius scales or larger. To obtain the parallel width of a toroidal ETG mode $\Delta \theta$, we balance the parallel streaming term with the change in the magnetic drift over the mode width, $$\frac{v_{te}}{qR_0} \frac{\partial h_e}{\partial \theta} \sim \Delta \theta \frac{\partial}{\partial \theta} (\mathbf{k}_{\perp} \cdot \mathbf{v}_{Me}) h_e. \label{eq:parallellengthbalance}$$ This is based on the conjecture that the magnetic drift profiles limit the parallel width of the mode. The quantity $\Delta \theta$ captures the width of the mode envelope, rather than the oscillations within it, which would be captured by $k_{\parallel}$. For the Taylor expansion of the magnetic drift frequency in to be valid, $\Delta \theta$ must be small, and as a result, any scalings that we obtain from will only be valid as long as $\Delta \theta \ll 1$. Assuming that $$\frac{\partial h_e}{\partial \theta} \sim \frac{h_e}{\Delta \theta}, \;\;\;\; \frac{\partial}{\partial \theta} (\mathbf{k}_{\perp} \cdot \mathbf{v}_{Me}) \sim \mathbf{k}_{\perp} \cdot \mathbf{v}_{Me},$$ and that magnetic drifts balance the drive frequency, as in , $$\mathbf{v}_{Me} \cdot \mathbf{k}_{\perp} \sim \omega_{*e} \eta_e,$$ we obtain a scaling for the mode width, $$\Delta \theta \sim \sqrt{\frac{v_{te}}{q R_0 \omega_{*e} \eta_e} } \sim \sqrt{\frac{1}{q k_y \rho_e } \frac{L_{Te}}{R_0} }, \label{eq:deltathetascaling}$$ where we use $\omega_{*e} \eta_e \sim k_y \rho_e v_{te} / L_{Te}$. Hence, higher values of $R_0/L_{Te}$, $k_y \rho_e $, and $q$ make the mode narrower. Using $\hat{s} \theta \sim R_0 / L_{Te}$, we obtain $$\frac{\Delta \theta}{\theta} \sim \hat{s} \sqrt{\frac{1}{q k_y \rho_e } } \left( \frac{L_{Te}}{R_0} \right)^{3/2}.$$ In the pedestal, the quantity $\Delta \theta / \theta $ is small, whereas in the core, $\Delta \theta / \theta$ is of order unity. Results from gyrokinetic scans in $q$, $R_0 / L_{Te}$ and $k_y \rho_e$ are in fair agreement with the scalings in . We report these scans in . To summarize thus far, pedestal toroidal ETG — where $R_0/L_{Te} \gg 1$ — has a very different character to core toroidal ETG — where $R_0/L_{Te} \sim 1$. In the pedestal, toroidal ETG can be driven strongly at wavenumbers as small as $k_y \rho_e \sim L_{Te} / R_0 \ll 1$, but with a large effective radial wavenumber $K_x \rho_e \sim 1$, due to the mode being driven far along the field line, $\hat{s} \theta \sim R_0/L_{Te} \gg 1$. For pedestal toroidal ETG, the radial component of the magnetic drift is essential for instability. In contrast, core toroidal ETG only becomes unstable at much larger poloidal wavenumbers $k_y \rho_e \sim 1$, and has a much smaller radial wavenumber $K_x \rho_e \ll 1$ due to $\theta \approx 0$. For core toroidal ETG, the in-surface poloidal magnetic drift is essential to the instability drive. Slab ETG is also shifted to larger perpendicular scales by $R_0/L_{Te} \gg 1$. Re-examining , and requiring a strong slab drive, $$\frac{\omega_{*e} \eta_e}{k_{\parallel} v_{te}} \sim \frac{k_y \rho_e }{k_{\parallel} R_0} \frac{R_0}{L_{Te}} \sim 1.$$ Thus, the scale for which slab ETG can be strongly driven is $$k_y \rho_e \sim k_{\parallel} R_0 \frac{L_{Te}}{R_0}. \label{eq:krescaleslab}$$ We place bounds on $k_y \rho_e $ for the ‘pure’ slab ETG branch by considering two linear effects that can constrain the parallel mode extent. The first constraint on the slab ETG mode is that the mode is not too strongly FLR damped, which according to , requires $$\theta \lesssim \frac{1}{\hat{s}} \frac{1}{k_y \rho_e}. \label{eq:FLRconstrainttheta}$$ A mode that oscillates only a few times before reaching the maximum value of $\theta$ in has a parallel wavenumber $k_{\parallel} \sim k_y \rho_e \hat{s} / q R_0$. Using , we find that such a mode would have $R_0/L_{Te} \sim \hat{s}/q$. Electron temperature gradients smaller than this value would be FLR damped. Since the gradients in the pedestal satisfy $R_0/L_{Te} \gg \hat{s}/q$, we conclude that the FLR damping constraint on the electron temperature gradient for the slab ETG mode is irrelevant in pedestals. The second constraint on the slab ETG mode determines how far the mode can extend in the parallel direction while still retaining a parallel streaming frequency that is faster than the magnetic drift frequency. From , the largest $\theta$ value a mode can have before $\mathbf{v}_{Me} \cdot \mathbf{k}_{\perp}$ and $\omega_{*e} \eta_e$ become comparable is $$\theta \lesssim \frac{1}{\hat{s}} \frac{R_0}{L_{Te}}.$$ A mode that oscillates only a few times before reaching this value of $\theta$ has a parallel wavenumber of order $$k_{\parallel} \sim \frac{\hat{s}}{qR_0} \frac{L_{Te}}{R_0}. \label{eq:kplldeltatheta}$$ A slab ETG mode with such a $k_{\parallel}$ is the mode with the smallest $k_y \rho_e $ value because, for smaller values of $k_y \rho_e$, the mode would have to extend into the region of $\theta$ where the magnetic drift is large. Thus, due to the magnetic drift condition, slab ETG modes must satisfy $$k_y \rho_e \gtrsim \frac{\hat{s}}{q} \left( \frac{L_{Te}}{R_0} \right)^2. \label{eq:magneticdriftconstraintkre}$$ Thus, for a fast-growing ‘pure’ slab ETG mode, we require $$\frac{\hat{s}}{q} \left( \frac{L_{Te}}{R_0} \right)^2 \lesssim k_y \rho_e \lesssim 1.$$ Even though our simple estimates suggest that slab ETG modes can grow for wavenumbers as small as $k_y \rho_e \sim (\hat{s}/q) (L_{Te}/R_0)^2 \sim 1/30000$, we should point out that kinetic ion physics is important at such large scales, and hence the slab ETG will be modified at these very long wavelengths. In principle, the above arguments are also valid for toroidal and slab ITG in the collisionless limit with identical gradients. However, in the JET pedestal equilibrium we have studied, $R_0/L_{Te} > R_0/L_{Ti}$, which causes the ITG growth rates to decrease substantially. Furthermore, in the pedestal the electrons are sufficiently collisional to be non-adiabatic on ITG timescales; as we will show in , these electron collisions also decrease the ITG growth rate. Indeed, we will see that the less steep measured ion temperature gradients and collisions result in ITG being the subdominant mode at all scales. For $k_y \rho_i \lesssim 1$, ITG is likely stable, and hence we do not expect ITG to cause significant transport in the equilibrium and radial location studied in this paper. For other JET pedestal equilibria that we studied in less detail, it was also true that $R_0/L_{Te} > R_0 / L_{Ti}$ in the steep gradient region; these equilibria had qualitatively similar growth rate spectra to the equilibrium studied in this paper (see ). We now proceed to obtain an ETG dispersion relation using the approximations in the previous sections. Its solutions will provide useful insights on toroidal ETG stability, which will be used heavily in subsequent sections. ETG Dispersion Relation {#subsec:dispersionrelation} ----------------------- ![The functions $\Gamma_0$ and $\Gamma_1$ that appear in .[]{data-label="fig:gammafunsreminder"}](PlotofGamma0.eps){width="55.00000%"} Formally solving for $h_s$ gives $$h_s = \frac{- \arc{ \omega}_s + \arc{ \omega}_{*s} \bigg{[} 1 + \eta_s \left( \hat{v}_{\parallel }^2 + \hat{v}_{\perp}^2 - 3/2 \right) \bigg{]} }{- \arc{ \omega}_s + \arc{ k}_{\parallel s} \hat{v}_{\parallel } + \sigma \hat{v}_{\parallel}^2 + \arc{\omega}_{\nabla B s} \hat{v}_{\perp}^2/2} \frac{Z_s e}{T_{0s}} \phi ^{tb} _1 F_{Ms} J_0 \left( \sqrt{2 b_s} \hat{v}_{\perp} \right), \label{eq:hsomegakappanormalized}$$ where the parallel wavenumber is the operator $$i k_{\parallel} h_s \equiv \hat{\mathbf{ b} } \cdot \nabla h_s,$$ and we define $b_s$ and $\hat{v}$ as $$b_s = \frac{k_{\perp}^2 T_{0s}}{m_s \Omega_s^2}, \;\;\;\;\; \hat{v} = \frac{v}{v_{ts}}.$$ We have non-dimensionalized quantities using the modulus of the curvature magnetic drift frequency $\omega_{\kappa s}$, $$\eqalign{ \sigma \equiv \frac{\omega_{\kappa s}}{|\omega_{\kappa s} |}, \;\; \arc{\omega} \equiv \frac{\omega }{|\omega_{\kappa s} |}, \;\; \arc{\omega}_{\nabla B s} \equiv \frac{\omega_{\nabla B s} }{|\omega_{\kappa s} |}, \;\; \arc{\omega}_{*s} \equiv \frac{\omega_{*s}}{|\omega_{\kappa s} |}, \;\; \arc{k }_{\parallel} \equiv \frac{k_{\parallel} v_{ts}}{|\omega_{\kappa s} |}, }$$ where $$\fl \omega_{\kappa s} \equiv \frac{v_{ts}^2 \mathbf{ k}_{\perp}}{\Omega_s} \cdot \left(\hat{\mathbf{b} } \times \left( \nabla \ln B + \frac{4\pi}{B^2} \frac{\partial p_0}{\partial r} \nabla r \right) \right), \;\;\; \omega_{\nabla B s} \equiv \frac{v_{ts}^2 \mathbf{ k}_{\perp}}{\Omega_s} \cdot (\hat{\mathbf{b} } \times \nabla \ln B ).$$ We write the total magnetic drift frequency as $$\mathbf{v}_{Ms} \cdot \mathbf{k}_{\perp} = \omega_{\kappa s} \hat{v}_{\parallel}^2 + \omega_{\nabla B s} \frac{\hat{v}_{\perp}^2}{2}.$$ It is important to note that is valid for any value of $\theta_0$, since in this work we are paying particular attention to the radial component of $k_{\perp}$ (see ) due to its importance for the toroidal ETG instability in steep temperature gradient regions. Thus, $b_s$, $\omega_{\kappa s}$, and $\omega_{\nabla B s}$ depend on $\theta_0$; this differs from many previous works where only the $\nabla y$ component of the magnetic drift frequency was retained. ![Solutions to with $\eta_e = 4.28$. (a): growth rates for different $\omega_{*e} \eta_e$ and $b_e$ with $k_{\parallel } =0$. (b): growth rates versus $k_{\parallel }$ for different values of $\omega_{*e} / \omega_{\kappa e}$ with $b_e = 0$ and $\omega_{\kappa e} > 0$. (c): growth rates $k_{\parallel }$ for different values of $\omega_{*e} / \omega_{\kappa e}$ and $b_e$. Here, we set $\omega_{\kappa e} = \omega_{\nabla B e}$.[]{data-label="fig:toroidaltoslab"}](theoryscan.eps){width="\textwidth"} As a simplified model, we will take $k_{\parallel}$ to be a number. We obtain the ETG dispersion relation by substituting into quasineutrality, as demonstrated in \[app:disprlnfull\]. For a single ion species, this gives $$\frac{T_{0e}}{T_{0i}} Z_i + 1 - \sum_s D_s = 0, \label{eq:localdispreln}$$ where $D_s$ is given by $$\fl \eqalign{ D_s = & i Z_s^2 \frac{T_{0e} n_{0s}}{T_{0s} n_{0e}} \int_0^{\infty} d \lambda \frac{ \Gamma_0(\hat{b}_s^{\sigma}) }{(1 + i \sigma \lambda)^{1/2}} \frac{ 1}{(1 + i \arc{\omega}_{\nabla B s} \lambda / 2)} \exp \bigg{(} i \lambda \arc{\omega} - \frac{(\lambda \arc{k }_{\parallel})^2}{4(1 + i \sigma \lambda)} \bigg{)} \\ & \times \Bigg{[} -\arc{\omega} + \arc {\omega}_{*s} \Bigg{(} 1 + \eta_s \bigg{\{} \frac{1}{1 + i \arc{\omega}_{\nabla B s} \lambda / 2} - \frac{3}{2} \\ & \;\;\;\;\;\; + \frac{2(1 + i \sigma \lambda) - (\arc{k}_{\parallel}\lambda)^2}{4 (1+i \sigma \lambda)^2} - \hat{b_s^{\sigma}} \frac{1-\Gamma_1(\hat{b}_s^{\sigma})/\Gamma_0(\hat{b}_s^{\sigma})}{1 + i \arc{\omega}_{\nabla B s} \lambda / 2} \bigg{\}} \Bigg{)} \Bigg{]}. } \label{eq:newDR}$$ The quantities $\Gamma_{\nu}$ and $\hat{b}_s^{\sigma}$ are defined as $$\Gamma_{\nu} (x) = I_{\nu} (x) \exp(-x), \;\;\;\; \hat{b}^{\sigma}_s \equiv \frac{b_s}{1 + i \arc{\omega}_{\nabla B s} \lambda/2},$$ where $I_{\nu}$ is a modified Bessel function of the first kind. We plot $\Gamma_0$ and $\Gamma_1$ in ; the function $\Gamma_0$ will be used extensively in this work. We have numerically solved in the adiabatic ion limit, $h_i = 0$, $$\frac{T_{0e}}{T_{0i}} Z_i + 1 - D_e = 0, \label{eq:toroidalETGsolved}$$ which is justifed by $k_{\perp} \rho_i \gg 1$. For information on the numerical techniques used to solve , refer to \[app:disprlnfull\]. In , we solve , performing a scan in $\omega_{*e} \eta_e / \omega_{\kappa e}$ and $k_{\parallel} v_{te} / \omega_{*e} \eta_e$. Note that while for we have set $\omega_{\kappa e} = \omega_{\nabla B e}$, when we solve with the geometry for the discharge 92174 in forthcoming sections, we use the correct values of $\omega_{\kappa e}$ and $\omega_{\nabla B e}$ (for example, see ). For the toroidal ETG mode, we observe two stability limits in $\omega_{*e} \eta_e / \omega_{\kappa e}$. (a) shows that for $b_e = 0$, toroidal ETG instability only occurs when $1.4 \lesssim \omega_{*e} \eta_e / \omega_{\kappa e} \lesssim 42$, and we found no instability when $\omega_{*e} \eta_e / \omega_{\kappa e} < 0$. We observe in (b) and (c) that increasing $k_{\parallel} $ causes the ETG instability to transition from the toroidal ETG branch to the slab ETG branch for the values of $\omega_{*e} \eta_e / \omega_{\kappa e}$ where the toroidal mode is unstable. Generally, increasing $b_e$ strongly decreases the growth rate for both the toroidal and slab branches, although small increasing values of $b_e$ can can increase the growth rate, shown by comparing the $\omega_{*e} \eta_e / \omega_{\kappa e} = 21$ values in (b) and (c). The $h_i = 0$ limit is generally an accurate description of toroidal and slab ETG instability in the JET pedestal discharges we analyzed, as will be described in . This is not surprising given that for the toroidal ETG instability we require $K_x \rho_e \sim 1$, which means that $h_i \approx 0$ because of the large argument of $J_0$ (see ). For the fastest growing slab ETG instability we usually find that $k_y \rho_i \gg 1$, again resulting in $h_i \approx 0$. However, the $h_i = 0$ approximation might not always be justified for $k_y \rho_i \sim 1$ slab ETG instability, where FLR damping has not substantially decreased the size of the ion kinetic response. In the next section, we proceed to use gyrokinetic simulations to study ETG stability in the pedestal. Of particular interest, consistent with the predictions of this section, we will find both toroidal and slab ETG modes at scales $k_y \rho_i \sim (\rho_i/\rho_e) L_{Te}/R_0 \lesssim 1$, and long poloidal wavelength toroidal ETG being unstable at $\hat{s} \theta \sim R_0 / L_{Te}$ (for $\theta_0 = 0$). ![Electrostatic GS2 growth rates for JET shot 92174 for $0.15 \leq k_y \rho_i \leq 7.0$ and sensitivity scans, all with $\theta_0 = 0$. (a): $R_0/L_{Te}$ scans. (b): $R_0/L_{Ti}$ and $R_0/L_n$ scans. ‘Standard’ denotes simulations performed with the following parameters: $R_0/L_{Te} = 130,\; R_0/L_{Ti} = 34,\; R_0/L_n = 31$. All of the fastest growing ‘Standard’ modes at scales $k_y \gtrsim 0.1$ are ETG-like instabilities.[]{data-label="fig:krionetofivesensitivityscan"}](92174RLTeTinscanNov15.eps){width="\textwidth"} ETG Stability in JET Shot 92174 {#sec:GKsims} =============================== In this section, we describe ETG instability in electrostatic gyrokinetic simulations of JET shot 92174 at $r/a = 0.9743$. The layout of this section is as follows. We first discuss the character of the toroidal and slab ETG instability in the pedestal in . In , we describe the parallel dynamics of the toroidal ETG mode, detailing how its parallel location and mode width are determined. In , the effects of a nonzero $\theta_0$ for the toroidal ETG mode are analyzed, including an estimate for the quasilinear diffusion coefficient. Then in , we study the critical temperature gradient for the toroidal ETG mode described in . Toroidal ETG Versus Slab ETG Instability {#subsec:toroidalETG} ---------------------------------------- ![(a): Ballooning eigenmodes for toroidal and slab ETG in GS2 simulations. (b): Toroidal ETG eigenmodes in $\theta_M$ space with $k_y \rho_i = 1.1$, using the transformation in at two locations: (1) $x/\rho_i = -0.1, y/\rho_i = 0.0$, and (2) $x/\rho_i = 0 , y/\rho_i = 0$. Location (1) is where the mode amplitude is maximum.[]{data-label="fig:krionetofiveplot"}](BallooningModesPlotApril22020BallooningModesONLYPhaseConstant.eps){width="\textwidth"} Gyrokinetic simulations show toroidal and slab ETG instability as the fastest growing modes for $k_y \rho_i \gtrsim 0.1$ for JET shot 92174. Unlike ETG instability in the core, where the linear growth rate typically peaks at $k_y \rho_e \sim 1$, we find instances of maximum toroidal ETG growth rates at spatial scales as large as $k_y \rho_i \sim (\rho_i/\rho_e) L_{Te}/R_0 \lesssim 1$, strongly supporting the arguments in . We emphasize that very similar modes have been seen in previous works [@Told2008; @Jenko2009; @Told2012; @Fulton2014; @Baumgaertel2011; @Kotschenreuther2019], but have not been explained until now. For $\theta_0 \neq 0$, we find toroidal ETG as the fastest growing mode at all spatial scales between $k_y \rho_i \sim 1$ and $k_y \rho_e > 1$, which we will discuss in . In , we show the growth rates of modes with $\theta_0 = 0$, where we find two dominant ETG modes: for this specific pedestal location, the toroidal ETG branch is the fastest growing mode for $1 \lesssim k_y \rho_i \lesssim 5$. Once $k_y \rho_i$ is sufficiently large ($k_y \rho_i \approx 5$) the toroidal ETG is FLR damped, and the slab ETG branch grows faster. The slab ETG branch is not FLR damped as quickly as the toroidal branch because the slab branch generally satisfies $K_x \sim k_y$. ![Real space images at the outboard midplane ($\theta_M = \vartheta = 0$), and at $\vartheta = 1.6$, $\theta_M = 2.1$, of a single toroidal ETG ballooning mode with $k_y \rho_i = 1.1$ and $\theta_0 = 0.0$ from GS2 simulations, demonstrating a relatively large radial wavenumber at both $\theta_M$ locations, and that the mode has a larger amplitude at $\theta_M = 2.1$ than at the outboard midplane. These were obtained using the transformation in . We define the coordinates $\delta R = R - R_M(r_c, \theta_r)$ and $\delta Z = Z - Z_M(r_c, \theta_r)$, where $\theta_r = -0, 2.1$ is the Miller poloidal angle of the image. The gyroradius $\rho_i$ is evaluated on the usual $r/a = 0.9743$ flux surface. Both plots are normalized to the same colorbar. Each box is evaluated on the same $x-y$ grid, and therefore each box is the same size in these variables. The small red boxes on the flux surface are a realistic size for what the plot domains would be in the experiment. The maximum absolute mode amplitude at $\theta_M = 0$ is about 25% of the mode amplitude at $\theta_M = 2.1$. The specific $\theta_M = 2.1$ location was chosen as this was the location of the maximum value of $\overline{ \phi } ^{tb} _1$, which can be seen in (b).[]{data-label="fig:krionetofiveplotrealspace"}](ModeCrossSec4Apr3){width="\textwidth"} We use several criteria to distinguish between the toroidal and slab ETG modes in the pedestal. First, as predicted in , toroidal ETG modes have $\Delta \theta / \theta \ll 1$, and have a $\theta$ location that satisfies $\hat{s} \theta \sim L_{Te} / R_0$ for $|\theta_0|$ sufficiently small. Parameter scans can also be used to determine whether the location along a field line of a suspected toroidal ETG mode changes as predicted by . In contrast, slab ETG modes tend to have a much larger $k_{\parallel}$ (at a fixed $k_y \rho_i$), and to have eigenmodes that are centered around $\theta = 0$. In , we show both toroidal and slab ETG eigenmodes in (a). To go from ballooning angle $\theta$ to the physical poloidal angle $\vartheta$, where $-\pi \leq \vartheta \leq \pi$, we use the ballooning transform, $$\fl \eqalign{ \overline{\phi} ^{tb}_1 (\vartheta, x&, y) = \sum_{p = -\infty}^{\infty} \phi ^{tb}_1 (\vartheta - 2\pi p) \exp \left( i k_y x \hat{s} \left(\vartheta - 2\pi p - \frac{r}{\hat{s} q} \frac{\partial \nu}{\partial r} \right) - i k_y y \right) \\ & + \sum_{p = -\infty}^{\infty} {\phi^{tb}_1}^{*} (\vartheta - 2\pi p) \exp \left(- i k_y x \hat{s} \left(\vartheta - 2\pi p - \frac{r}{\hat{s} q} \frac{\partial \nu}{\partial r}\right) + i k_y y \right), } \label{eq:undoballooningtransform}$$ where $*$ denotes a complex conjugate. In (b), the toroidal ETG eigenmode is plotted against the Miller angle $\theta_M$ for $x/\rho_i = 0, \; y/\rho_i = 0$ and for $x/\rho_i = -0.1, \; y/\rho_i = 0.0$. We have normalized the mode such that the maximum of $\overline{\phi } ^{tb} _1 $ is 1, and we have chosen the mode’s phase such that the maximum is located at $y = 0$. The maximum value of $\overline{ \phi} ^{tb} _1$ occurs at $x/\rho_i = -0.1$. In , we show the real space picture of the mode at the outboard midplane ($\theta_M = 0$) and where the amplitude is maximum, at $\theta_M = 2.1$. As expected, the toroidal ETG modes have $K_x \gg k_y$ at both the outboard midplane and at $\theta_M = 2.1$, and the maximum amplitude is far away from the outboard midplane. To make the plots in , we first evaluated for $k_y \rho_i = 1.1$ on a uniform $x,y$ grid. We then performed a change of variables from $x,y$ to $R,Z$ using the Miller formulas for $R_M$ and $Z_M$. Finally, we changed from $\vartheta$ to $\theta_M$ variables. ![Growth rates. (a): GS2 scan in $R_0/L_{Te}$, (b): GS2 scan in $T_{0i}/T_{0e}$, and (c): theory scan in $T_{0i}/T_{0e}$. These scans show the value of $k_y \rho_i$ for the peak growth rate of the toroidal ETG mode shifting. For $T_{0i}/T_{0e}$ scans, $T_{0i}$ was fixed and $T_{0e}$ was allowed to vary. (b): growth rates from GS2 simulations with consistent collisionality. (c): the collisionless dispersion relation in was solved, along with a Fourier-transformed value of $k_{\parallel}$ for each $k_y \rho_i$ mode, described in . The numbers in parentheses in the legend for (a) are the multiples of the correct $R_0/L_{Te}$ value.[]{data-label="fig:kymaxmovement"}](TiTeandRLTescanCollisionsandCollisionlessWithTheoryMar242020.eps){width="\textwidth"} To investigate the character of the toroidal and slab ETG modes, we have performed a scan in equilibrium gradients, as shown in the linear gyrokinetic spectrum in . Our simulations indicate that the fastest growing toroidal ETG modes are driven strongly by $R_0/L_{Te}$ because they depend strongly on this parameter, as shown in (a). Conversely, these modes are relatively insensitive to $R_0/L_n$, and do not depend on $R_0/L_{Ti}$. Modifying $R_0/L_n$ mainly affects the slab ETG growth rate, determining at which $k_y \rho_i $ it will exceed the toroidal ETG growth rate. Kinetic ion physics is usually unimportant for toroidal ETG instability because $k_{\perp} \rho_i \gg 1$. This is demonstrated by the linear spectrum for the toroidal ETG being unchanged when the non-adiabatic part of the ion distribution function is artificially set to zero, $h_i = 0$, shown (b). The simulation results in also show higher $R_0/L_{Te}$ and smaller $T_{0i}/T_{0e}$ shifting the maximum growth rate of the toroidal ETG instability to a smaller $k_y \rho_i$, as predicted by . Unlike the wavenumber of the fastest growing modes, the size of the maximum growth rate in the range of wavelengths shown depends of $T_{0e}/T_{0i}$ in a non-trivial way. We show in that this dependence is consistent with a theory that we describe in . ![(a): two eigenmodes obtained from two separate GS2 simulations, and the function $\Gamma_0(b_e)$ for $k_y \rho_i = 3.4$. When $\omega_{M e} \to - \omega_{M e}$, the mode moves to a location where the sign of $\omega_{M e} $ allows instability, where $\omega_{M e}$ refers to either $\omega_{\kappa e}$ or $\omega_{\nabla B e}$. (b) The quantities $\omega_{*e} \eta_e / \omega_{\kappa e}$ and $\omega_{*e} \eta_e / \omega_{\nabla B e}$. The eigenmodes in (a) have their maxima in bad curvature regions, corresponding to $\omega_{*e} \eta_e / \omega_{M e} > 0$. (c): finding the growth rates for the ETG dispersion relation in for two signs of $\omega_{*e} \eta_e / \omega_{M e} $ in JET shot 92174. Note how the maximum growth rates in (c) roughly align with the eigenmode maximum in (a). Horizontal red and blue lines denote the eigenmode location for the two signs of $\omega_{Me}$ in (a). Here, $\omega_{*e} < 0$, $\eta_e = 4.28$, $k_y \rho_i = 3.4$, $k_{\parallel} = 0$, $\theta_0 = 0$.[]{data-label="fig:omegaMeflip"}](nversusthetafixedflippingsignofommeandGamma0WITHMATLABGROWTHRATESMar242020ConsistentOmegakappaKP3p4.eps){width="\textwidth"} To understand the $\theta$ location of the toroidal ETG eigenmodes, we solve the dispersion relation in locally for JET shot 92174 at each value of $\theta$ by choosing $k_y$ and $k_{\parallel}$, and by using $\omega_{\kappa e}, \omega_{\nabla B e}$ and $b_e$ from the Miller equilibrium. This is an approximation that assumes the mode’s growth rate is local in $\theta$. Note that $\mathbf{k}_{\perp}$ in is a function of $\theta$. By solving the dispersion relation, we obtain a set of frequencies as a function of $\theta$. (c) shows the growth rates along $\theta$ with $k_{\parallel} = 0$ (for the present discussion, consider only the curve labeled ’Standard’; the curve labeled ‘$\omega_{Me} \rightarrow - \omega_{Me}$’ will be discussed in ). For $\theta_0 = 0$, we find that the maximum growth rates are at $|\theta| \simeq 7.7$ with the standard sign of $\omega_{\kappa e}$ and $\omega_{\nabla B e}$. This $\theta$ location is very close to the $\theta$ where GS2 toroidal ETG eigenmodes have their maximum amplitude, as shown by comparison of (a) and (c). Therefore, the parallel location of the toroidal ETG is fairly well described by our model. ![Electrostatic slab eigenmodes from GS2 for $ k_y \rho_i > 7.0$ instabilities at $\theta_0 = 0.05$. The corresponding linear growth rates are shown in the inset.[]{data-label="fig:kribiggerthanfive"}](nversustheta92174slabETGscalesEIGENMODEShighresDec7smoother.eps "fig:"){width="90.00000%"} \[fig:torEmodes\] One prediction of was that the toroidal ETG mode is driven most strongly at $\hat{s} \theta \gg 1$ when $R_0/L_{Te} \gg 1$. This causes the $k_y \hat{s} \theta \nabla x$ term in $\mathbf{k}_{\perp} $ in to become particularly large. In , we show that the toroidal ETG eigenmodes are indeed driven at $\hat{s} \theta \gg 1$. As an experiment, we set the $k_y \hat{s} \theta \nabla x$ component of $\mathbf{v}_{Me}$ to zero. As expected, the toroidal ETG mode was not driven, and slab ETG was the fastest growing mode. In JET shot 92174, slab ETG instability is the fastest growing mode for $k_y \rho_i \gtrsim 5$ when $\theta_0 = 0$ — however, the ‘slab’ ETG we observe is not always the conventional slab ETG with $\omega_{\kappa e} = \omega_{\nabla B e} = 0$. By artificially turning the magnetic drift off in gyrokinetic simulations, we observed that the slab ETG growth rate was reduced by factors of order unity. As shown in , the slab ETG eigenmodes have quite a wide $\theta$ extent, especially for smaller $k_y \rho_i$ where FLR effects are less strong, and hence the magnetic drift, which increases for increasing $\theta$, can have a strong impact on the character of the slab ETG in the pedestal. As $k_y \rho_i $ increases, FLR effects become stronger and the slab ETG eigenmode becomes more localized near $\theta = 0$. Hence, when we refer to the ‘slab’ ETG in the pedestal simulations described in this paper, we refer to the modes with a $k_{\parallel}$ much larger than the toroidal ETG, but also sometimes with a significant magnetic drift contribution. ![A stability plot for the toroidal ETG mode, combining theory and GS2 simulations. For (a), the small red $b_e = 0$ stable region corresponding to $0 < \omega_{*e} \eta_e / \omega_{\kappa e} \lesssim 1.8$, is obtained from . The blue $b_e = 0$ stable region is also obtained from , and corresponds to $\omega_{*e} \eta_e / \omega_{\kappa e} \gtrsim 42$. This is valid for $\theta_0 = 0$ and $k_y \rho_i = 1.1$. (b): quantity $\Gamma_0(b_e)$ versus $\theta$ for $k_y \rho_i = 1.1$. (c): the associated eigenmodes from GS2 with different temperature gradients, demonstrating that these modes are centered close to local maxima in $\Gamma_0(b_e)$, and that increasing $R_0/L_{Te}$ moves the mode to larger $\hat{s} \theta$, predicted in . Only for (c), we artificially lowered $\hat{s} \to 1.68$ to make the mode more mobile in $\theta$. Dashed vertical lines show the local maxima of $\Gamma_0(b_e)$ in bad curvature regions.[]{data-label="fig:generalstabilityplot"}](omsteetaeoveromegaMewithtrickintervalsJan14withGamma0.pdf){width="90.00000%"} The toroidal ETG modes are not affected by kinetic ion physics due to their large radial wavenumber $K_x \rho_i \gg 1$, but the ions can modify the slab ETG modes slightly when $k_y \rho_i \sim 1$, as we demonstrate in , where we show results with the full ion kinetic response and with $h_i = 0$. This is consistent with the fact that slab modes with $k_y \rho_i \sim 1$ have $K_x \rho_i \sim 1$. We have checked that $h_i$ becomes unimportant at larger values of $k_y \rho_i$. Note that the slab ETG modes in are asymmetric. This asymmetry is not a result of our choice of $\theta_0$ because we observe it in modes with $\theta_0 = 0$. Due to the symmetry of the gyrokinetic equation described in [@Peeters2005], for $\theta_0 = 0$, if one obtains an asymmetric mode, there must be two modes with opposing asymmetry that grow at the same rate. We have run our simulations with a small value of $\theta_0$ to avoid getting a linear combination of these two modes — the final result would depend on the initial conditions in this case. Thus far, using the method described above to solve the dispersion relation in , we found we could predict the parallel location of the toroidal ETG modes. We next describe the physics that determines the parallel location and width of the toroidal ETG mode in more detail. Location And Width Of The Toroidal ETG Mode {#sec:locwidth} ------------------------------------------- We now discuss the parallel location and width of the toroidal ETG mode. The parallel location of the toroidal ETG mode is subject to four main constraints: 1. **The mode can only be driven in bad curvature regions,** $ \omega_{*e} \eta_e / \omega_{\kappa e} > 0$, which eliminates roughly half of the parallel domain. 2. **The mode is only unstable when** $\mathbf{ A > \omega_{*e} \eta_e /\omega_{\kappa e} > C}$. According to the results in (a), for toroidal ETG instability the value of $\omega_{*e} \eta_e /\omega_{\kappa e}$ must be above some critical value $C$ for instability, but not larger than another critical value $A$. Consistent with (a), we observe that no toroidal ETG modes with $\theta_0 = 0$ can exist at $|\theta| \lesssim 6$; this is because $\omega_{*e} \eta_e / \omega_{\kappa e} $ is too large and the bad curvature region is too narrow, as shown in (a) (note that for smaller values of $R_0/L_{Te}$, the $\theta_0 = 0$ toroidal ETG mode *can* have its maximum amplitude at $|\theta| \lesssim 6$ because $\omega_{*e} \eta_e/ \omega_{\kappa e} $ is smaller — see ). Note that we discuss ‘good’ and ‘bad’ curvature using the quantity $\omega_{*e} \eta_e / \omega_{\kappa e} $ rather than $\omega_{*e} \eta_e / \omega_{\nabla B e} $ because in the regions where the toroidal ETG mode is typically most unstable (at large $|\theta|$), $\omega_{\kappa e}/ \omega_{\nabla B e} \simeq 1$ (see (b), for example). There are important exceptions, which occur for $\theta_0 \neq 0$ with larger values of $k_y \rho_i$, which we discuss briefly in . 3. **The parallel extent of bad curvature regions must be sufficiently wide.** We require that the ‘bad curvature’ regions not be too narrow in the parallel direction; if this is the case, then the mode acquires a large value of $k_{\parallel}$ and becomes damped. 4. **The mode maximum is close to a local maximum in $\Gamma_0 (b_e)$.** The maximum amplitude for the fastest growing toroidal ETG mode (at a given $k_y \rho_i$) is usually centered close to a local maximum in $\Gamma_0 (b_e)$ (or equivalently a local minimum in $b_e$) to limit FLR damping. We choose to plot the quantity $\Gamma_0 (b_e)$ rather than $b_e$ to demonstrate the importance of FLR damping at different $\theta$ locations. This is because $\Gamma_0 (b_e) \in [0,1]$, and therefore it is easier to convey the size of FLR damping, whereas $b_e$ is unbounded and can become extremely large. Furthermore, the term $\Gamma_0 (b_e)$ appears directly in the dispersion relation in , and thus is a good measure of the size of FLR effects. ![(a): linear growth rates from GS2 for different $\hat{s}$ values with $R_0/L_{Te} = 520$. (b): corresponding eigenmodes for $k_y \rho_i = 2.6$.[]{data-label="fig:shiftingthetawithshat"}](gammaandshatmovingNov15.eps){width="\textwidth"} As an experiment, we artificially reversed the signs of the magnetic drifts in GS2. As expected, the toroidal ETG modes only grew in regions that were previously ‘good curvature’ regions, which due to the sign reversal of $\omega_{\kappa e}$, are turned into ‘bad curvature’ regions. This is shown in , being substantiated both by GS2 simulations ((a)) and the results of our model ETG dispersion relation ((c)). Since $\omega_{*e} \eta_e$ is fixed for a given $k_y \rho_i $, the $\theta$ location will be such that $\omega_{\kappa e}$ and $b_e$ have the right value for maximum growth subject to FLR and curvature constraints. These constraints are shown in (a) and (b). According to (a) and the above arguments, the smallest $|\theta|$ that a mode with $\theta_0 = 0$ can occupy is $|\theta| \simeq 6.5$. We denote this minimum $\theta$ location as $\theta_{\mathrm{min} }$. The toroidal ETG mode cannot occupy a smaller $|\theta|$ value because either $\omega_{*e} \eta_e / \omega_{\kappa e} < 0$, $\omega_{*e} \eta_e /\omega_{\kappa e}$ is too large, or the bad curvature region is too narrow. From these considerations, there are several obvious parameters that can change where the mode is located. As already predicted in , a larger $R_0/L_{Te}$ causes a mode to be unstable at larger $\theta$ values; in (c) we show that increasing $R_0/L_{Te}$ increases the $\theta$ location of the mode. In (c), we use a smaller value of $\hat{s}$ (1.68 instead of 3.36), since we found that, for larger values of $\hat{s}$, increasing $R_0/L_{Te}$ was not particularly effective at shifting the mode to larger values of $|\theta|$ — this is because $b_e$ increases nonlinearly with $\hat{s}$, and once $\hat{s}$ is sufficiently large, a toroidal ETG mode becomes significantly more FLR damped as it moves along $\theta$. The parallel location of the modes with different values of $R_0/L_{Te}$ agrees well with the curvature and FLR constraints discussed above. Smaller $\hat{s}$ and $k_y \rho_i$ also force the mode to larger $\theta$ — as predicted in , the shifting of modes due to $\hat{s}$ and $k_y \rho_i$ is shown in , respectively. ![(a): growth rates versus $k_y \rho_i$. (b): corresponding eigenmodes and the functions $\Gamma_0(b_e)$ and $\omega_{*e} \eta_e / \omega_{\kappa e} $. The toroidal ETG mode shifts due to changing $k_y \rho_i $, predicted by . Here we have set $\hat{s} = 0.45$ and $R_0/L_{Te} = 520$, allowing the mode to be very mobile in $\theta$. The values of $\Gamma_0(b_e)$ are evaluated for $k_y \rho_i = 5.9$.[]{data-label="fig:krishiftingthemode"}](modejumpingJan14.eps){width="\textwidth"} \(a) illustrates that the toroidal ETG growth rate is relatively insensitive to $\hat{s}$, until $\hat{s}$ exceeds a threshold value. Recall that $\omega_{*e} \eta_e / \omega_{\kappa e} \sim R_0 / L_{Te} \hat{s} \theta$. This implies that if $\hat{s}$ changes, then a toroidal mode would move in $\theta$ to have a $R_0 / L_{Te} \hat{s} \theta$ that maximizes its growth rate. As $\hat{s}$ increases, the $\theta$ location will decrease. However, the mode cannot be driven linearly unstable below $\theta_{\mathrm{min}}$, so at a critical value of $\hat{s}$ the mode will become increasingly stabilized by FLR effects while the mode maximum remains at fixed $\theta = \theta_{\mathrm{min} }$. In (a), we show that increasing $\hat{s}$ beyond some critical $\hat{s}$ indeed decreases the growth rate of the toroidal ETG mode. This increase in $\hat{s}$ once the mode was at $\theta_{\mathrm{min} }$ increased $k_{\perp}$, and hence caused its growth rate to be lower than the slab ETG mode — this occurred for a value of $\hat{s}$ somewhere between $\hat{s} = 3.4$ and $\hat{s} = 10$ in (b). The $\theta$ location of the mode also depends strongly on $k_y \rho_e$, as shown in (b) where we ran GS2 simulations with a smaller value of $\hat{s} = 0.45$ and an increased value of $R_0/L_{Te}$, which makes the location of the mode more sensitive to changes in $k_y$. Clearly, the eigenmodes are centered very close to a local minimum in $b_e$. The toroidal ETG modes are close to this minimum because of a competition between the size of the magnetic drift and FLR effects; as shown in , the growth rates are very sensitive to $b_e$. Careful inspection of the growth rates in (a) reveals that there is a change in mode type as the mode jumps to a new $\theta$ location — this can be seen by discontinuities in $\partial \gamma / \partial k_y$. ![(a): toroidal ETG eigenmodes for different values of $k_y \rho_i $, and numerical definition of $\Delta \theta$ used in subsequent subplots. (b): Numerical (solid) and predicted (dashed) $\Delta \theta$ versus $q$ scaling, (c): $\Delta \theta$ versus $R_0/L_{Te}$ scaling, (d): $\Delta \theta$ versus $k_y \rho_i $ scaling.[]{data-label="fig:narrowingthemode"}](modewidthscanwithqscanJan142020LOGLOG.eps){width="\textwidth"} We now examine the scalings for the mode width from by comparing them with toroidal ETG eigenmodes from GS2 simulations. We calculate the width $\Delta \theta$ as the length in $\theta$ for the half height of the mode; this is shown in (a). predicts that the mode width $\Delta \theta$ scales with $R_0/L_{Te}$, $k_y \rho_i$, and $q$ as $\Delta \theta \sim \sqrt{L_{Te} / R_0 k_y \rho_e q}$. Scans in these quantities, shown in , demonstrate increasing $R_0/L_{Te}$, $k_y \rho_i$, and $q$ narrows the toroidal ETG mode structure. However, the scaling exponents do not appear to be quantitatively correct. The theoretical scaling $\Delta \theta \sim \sqrt{L_{Te} / R_0 k_y \rho_e q}$ in is not perfect because the mode changes location. Indeed, since the parallel location of the mode is sensitive to $q, k_y \rho_i$, and $R_0/L_{Te}$, changing the location of the mode by changing these parameters changes the local derivative of $\mathbf{v}_{Me} \cdot \mathbf{k}_{\perp}$, and hence changes $\Delta \theta$. Additionally, because we have used a Taylor expansion assuming that the variation in $\mathbf{v}_{Me} \cdot \mathbf{k}_{\perp}$ is proportional to $\Delta \theta$, this expansion breaks down when $\Delta \theta$ becomes too large. As the toroidal ETG instability is FLR damped at increasing $k_y$, the mode switches to the slab branch, with an accompanying increase in $k_{\parallel} $. The switch from toroidal to slab at fixed $k_y$ is shown in the simple dispersion relation used to plot (c). At this transition, $k_{\parallel} $ for the slab mode is much larger than the toroidal mode and the eigenmodes move from being quite localized around a large value of $\theta$, to oscillating rapidly about smaller $\theta$, as shown in (a). ![(a): the Fourier transformed coefficient $|\hat{\phi} ^{tb} _1 (m)|^2$ spectrum for 2 modes from GS2 with different values of $k_y \rho_i$. (b): eigenmodes. (c): the $k_{\parallel} $ associated with the largest coefficient $|\hat{\phi} ^{tb} _1 (m)|^2$ in (a). (d): growth rates. All of these plots have $\theta_0 = 0.02$.[]{data-label="fig:transitionfromtoroidaltoslab"}](kpllplotwithmanythings92174Jun8kpllnormalized.eps){width="90.00000%"} To demonstrate this transition, we need to define $k_{\parallel}$. Our choice of $\theta$ in is such that $\theta$ is proportional to the length along the magnetic field line. Thus, Fourier analyzing in $\theta$ is equivalent to obtaining the spectrum in $k_{\parallel}$. To carry out the Fourier transform, we first interpolate $\phi ^{tb}_1 (\theta)$ onto a regular $\theta$ grid, since GS2’s $\theta$ grid is not usually regularly spaced. Next, we apply a Fast Fourier Transform [@Cooley1965] to obtain the Fourier transform of $\phi ^{tb}$, $$\hat{\phi} ^{tb}_1 (m) = \int_{-\infty}^{\infty} \phi ^{tb}_1 (\theta) \exp(-i m \theta) d \theta. \label{eq:phifourierkpll}$$ The relation between $m$ and $k_{\parallel} $ is $$k_{\parallel} = \frac{2\pi}{L_{\theta}}m.$$ (a) shows that the power spectrum $|\hat{\phi} ^{tb}_1|^2 $ changes significantly at the transition between toroidal and slab ETG. The toroidal ETG spectrum is Gaussian whereas the slab spectrum is more complicated, with at least two peaks. It is noteworthy that the toroidal ETG has a non-zero $k_{\parallel}$ for its fastest growing mode since theory predicts toroidal ETG with the highest growth rate at $k_{\parallel} = 0$, shown in . Previous studies of toroidal ETG have also found $k_{\parallel} = 0$ as the fastest growing mode [@Dorland2000]. We now use to calculate the toroidal ETG growth rates for a range of $k_y \rho_i $. Our analytic model requires $k_{\parallel}$ as an input, which we obtain from GS2 by choosing the value of $k_{\parallel}$ that corresponds to the largest amplitude in the poloidal Fourier transform $\hat{\phi} ^{tb} _1$. Once we have obtained $k_{\parallel} $ from the GS2 data for each value of $k_y \rho_i$, we solve the model dispersion relation in for each value of $\theta$, inputting the correct value of $k_{\perp}$, $\omega_{\kappa e}$, and $\omega_{\nabla B e}$ at each $\theta$ location. For each $k_y \rho_i$ value, we take the growth rate from the $\theta$ location with the highest growth rate to be the growth rate of the toroidal ETG mode for that $k_y \rho_i $. There is excellent agreement between the $\theta$ location with the highest growth rate by solving and the eigenmode maximum from GS2. This method for calculating $k_{\parallel}$ gave a toroidal ETG growth rate reasonably close to the values obtained from GS2 shown in , as well as the $k_y \rho_i$ location of the peak. Surprisingly, this method also gives a very good approximation to the slab ETG growth rate even though slab ETG modes are very extended (see ). The theory presented in this paper cannot self-consistently calculate $k_{\parallel} $ and thus we have used solutions with a $k_{\parallel} $ associated with the numerical simulations. Until now, our analysis has been performed with $\theta_0 = 0$. In the next section, we extend our analysis to toroidal ETG with a nonzero value of $\theta_0$. ![The growth rates obtained in theory and in GS2. For the toroidal ETG growth rate, we found the $\theta$ with the highest growth rate for , which occurred at $\theta = 7.7$, and for the slab ETG growth rate, we evaluated the dispersion relation at $\theta = 0.0$ (note that $\omega_{\kappa e}$ is nonzero at $\theta=0$). The $k_{\parallel} $ input for the toroidal ETG was obtained by Fourier transforming the GS2 eigenmodes for each $k_y$, and for the ‘$1.25 k_{\parallel}$’ series, we multiplied all $k_{\parallel} $ values by 1.25.[]{data-label="fig:growthversussimulation"}](growthrateplottoroidalETGMar2020.eps){width="49.00000%"} Effects of $\theta_0$ {#sec:effectsofkx} --------------------- We now consider ETG instability for $\theta_0 \neq 0$. The growth rate of microinstabilities and MHD ballooning instabilities has a complicated dependence on $\theta_0$. Previous works have found that nonzero $\theta_0$ can substantially change the growth rates for toroidal ITG [@Migliano2013; @Kotschenreuther2017], ETG [@Jenko2009; @Told2012], and MTMs [@Hatch2016]. For MHD ballooning modes, it was found that for smaller pressure gradients, increasing $|\theta_0|$ is stabilizing, but once the gradients become sufficiently large, increasing $|\theta_0|$ is destabilizing [@Miller1995]. ![The effect of $\theta_0$ on growth rates and eigenmodes. (a): growth rates with three values of $\theta_0$. Vertical dashed lines indicate the $k_y \rho_i$ values for the eigenmodes that are shown in (b), (c), (d), and (e). (b), (c), (d), and (e): eigenmodes for $k_y \rho_i = 2.11, 6.34, 21.15, 49.35$ and different $\theta_0$. (f): $\omega_{*e} \eta_e / \omega_{\kappa e}$ for different $\theta_0$; for $|\theta_0|$ sufficiently large, new good curvature regions near $\theta = 0$ appear. (g) and (h): $\Gamma_0(b_e)$ for different $\theta_0$ at two values of $k_y \rho_i$. Vertical solid lines on rows 2 - 5 indicate the maximum amplitude of a selected toroidal ETG eigenmode for a given $\theta_0$; if the eigenmode is not shown for a given $k_y \rho_i$, then the fastest growing mode for that $k_y \rho_i$ is not a toroidal ETG mode. Rows 2-5 share the same $\theta$ axis. Consistent coloring and linestyle series is used throughout the plot, determined by the legend in (a).[]{data-label="fig:theta0growthrate"}](92174growthratesdifferenttheta0higherkyTHREETHETA0Nov13.eps){width="\textwidth"} As briefly discussed in , we find that increasing $|\theta_0|$ can substantially increase the toroidal ETG growth rate, shown in (a). For many values of $\theta_0$, the toroidal ETG mode can be the fastest growing mode not only at ion scales, $k_y \rho_i \sim 1$, but at scales smaller than the electron gyroradius: $k_y \rho_e > 1$. To be precise, we find that at low values of $k_y \rho_i $ ($k_y \rho_i \lesssim 2$), the toroidal ETG has a similar growth rate for all values of $\theta_0$, whereas for larger values of $k_y \rho_i $, the toroidal ETG growth rate becomes very strongly dependent on $\theta_0$. We proceed to explain why. For $k_y \rho_i \lesssim 2$, the location and growth rate of the toroidal ETG mode are fairly independent of $\theta_0$, as shown in (a) and (b). For such small values of $k_y \rho_i$, FLR damping is weak at many $\theta$ locations, that is, $k_{\perp} \rho_e \ll 1$ (and hence $\Gamma_0 (b_e) \approx 1$) in many distinct bad curvature regions. Since $\Gamma_0 (b_e) \approx 1$ in multiple regions, the fastest growing mode will be located at $\theta$ where $\omega_{*e} \eta_e / \omega_{\kappa e}$ is optimal. The value of $\omega_{*e} \eta_e / \omega_{\kappa e}$ is modified by $\theta_0$, shown in (f). The modification is particularly noticeable for $|\theta| \lesssim 6$, where there are regions of much smaller values of $\omega_{*e} \eta_e / \omega_{\kappa e}$ when $\theta_0$ is nonzero. For example, for $\theta_0 = -1.05$, (f) shows that $\omega_{*e} \eta_e / \omega_{\kappa e}$ has values as small as $\omega_{*e} \eta_e / \omega_{\kappa e} \simeq 15-30$ for $1 \lesssim \theta \lesssim 2$. While this value of $\omega_{*e} \eta_e / \omega_{\kappa e}$ is appropriate to have an unstable toroidal ETG mode, at larger values of $|\theta|$ there exists an even smaller value of $\omega_{*e} \eta_e / \omega_{\kappa e}$ (recall that smaller $\omega_{*e} \eta_e / \omega_{\kappa e}$ typically gives higher growth rates as long as $\omega_{*e} \eta_e / \omega_{\kappa e} \gtrsim 2-3$, see ). Again considering the $\theta_0 = -1.05$ mode, we see that $\omega_{*e} \eta_e / \omega_{\kappa e} \simeq 3-10$ for $-8 \lesssim \theta \lesssim -7$. Because we are currently considering relatively small values of $k_y \rho_i $, the FLR damping at $\theta = -7.7$ is not much stronger than at $\theta = 1.5$ (see (g)). Therefore, a mode at $\theta \simeq -7.7$ grows faster than a mode at $\theta \simeq 1.5$. The $k_y \rho_i = 2.11$ modes in (b) (all with $\theta_0 \leq 0$) have their maximum amplitude at $\theta = -7.7$ rather than $\theta = 7.7$ because FLR damping is slightly weaker at $\theta = -7.7$. Because *both* the $\omega_{*e} \eta_e / \omega_{\kappa e}$ profiles and the $\Gamma_0 (b_e)$ profiles are not strongly dependent on $\theta_0$ for $|\theta| \gtrsim 6$ (see (f)), the location of the toroidal ETG modes and their associated growth rates are almost independent of $|\theta_0|$ for $k_y \rho_i \lesssim 2$, although the sign of the $\theta$ location does depend on $\mathrm{sign}(\theta_0)$. ![Growth rate-associated quantities from GS2 simulations. (a): contour plot of growth rates versus $\theta_0$ and $k_y \rho_i$. (b): contour plot of $\gamma / k_{\perp}^2$ versus $\theta_0$ and $k_y \rho_i$. (c): location of the maximum of $|\phi _1 ^{tb} |$, $\theta_{\mathrm{Max} }$. (d): the maximum value of $\gamma / k_{\perp}^2$ (over all $\theta_0$ values) for each value of $k_y \rho_i$.[]{data-label="fig:theta0growthrateandmodeclassification"}](growthratesalltheta0contourandmodetypeFeb9.eps){width="\textwidth"} We now consider what happens for larger values of $k_y \rho_i$. Here, the $\Gamma_0$ profiles are much more strongly dependent on $\theta_0$, as shown in (h). For $\theta_0 = 0$, as $k_y \rho_i $ increases the toroidal ETG mode cannot grow at a smaller value of $|\theta|$ because either $\omega_{*e} \eta_e / \omega_{\kappa e}$ is too large, or the bad curvature region is too narrow, causing the mode to have a stabilizing value of $k_{\parallel}$. Hence, the $\theta_0 = 0$ toroidal ETG mode becomes increasingly FLR damped as $k_y \rho_i $ increases and at $k_y \rho_i \simeq 5$, the slab ETG mode overtakes the FLR damped toroidal ETG mode to become the fastest growing mode (see (a)). However, for nonzero $\theta_0$, the toroidal ETG mode *can* grow at a smaller value of $|\theta|$ where FLR damping is much weaker, and have a high growth rate because $\omega_{*e} \eta_e / \omega_{\kappa e}$ is sufficiently small. A consequence of the toroidal ETG mode moving to a bad curvature region with reduced FLR damping is that modes can be unstable in a wide range of poloidal locations, even close to the inboard midplane of the tokamak, a region that has traditionally been considered to have ‘good curvature’ for all values of $\theta_0$ (see (b), where even the toroidal ETG mode with $\theta_0 = 0$ is unstable close to the inboard midplane). However, the maximum eigenmode amplitude for the fastest growing mode is typically close to $\theta \; \mathrm{mod} \; 2 \pi \simeq \pm \pi/2$, which is mainly due to local magnetic shear making a local maximum in $\Gamma_0$ at $\theta \; \mathrm{mod} \; 2 \pi \simeq \pm \pi/2$. As shown in (c), (d), and (e), for nonzero $\theta_0$ and larger values of $k_y \rho_i$, the mode moves to a $\theta$ location that satisfies $\theta \theta_0 < 0$. This can be explained by including $\theta_0$ in the scaling for $\omega_{*e} \eta_e / \omega_{\kappa e}$, $$\frac{\omega_{*e} \eta_e}{\omega_{\kappa e}} \sim \frac{k_y}{k_{\perp}} \frac{R_0}{L_{Te}} \sim \frac{1}{\hat{s} (\theta_0 - \theta)} \frac{R_0}{L_{Te}} \sim 1.$$ Hence, at larger values of $k_y \rho_i$ when a mode needs to move to a location with a smaller $|\theta|$ value, it will choose the location where $\theta \theta_0 < 0$ in order to make $\omega_{*e} \eta_e / \omega_{\kappa e}$ small. To summarize, for smaller values of $k_y \rho_i$ (here $k_y \rho_i \lesssim 2$), FLR effects are relatively weak in multiple bad curvature regions, allowing the toroidal ETG mode to choose between multiple $\theta$ locations in order to find the optimal value of $\omega_{*e} \eta_e / \omega_{\kappa e}$. For the equilibrium considered in this paper, this occurs for $|\theta| \gtrsim 6$. However, when $k_y \rho_i$ is much larger and $\theta_0 = 0$, FLR damping prevents instability at higher values of $|\theta|$, even though bad curvature regions still exist there. For larger $k_y \rho_i $ and $\theta_0 \neq 0$, instability becomes possible at lower $|\theta|$ values due to modest FLR damping in select regions near $\theta = 0$. To gauge the relative importance of toroidal and slab ETG modes for transport, we calculate the quantity $\gamma / k_{\perp}^2$ for all modes at $1 \lesssim k_y \rho_i \lesssim 230$ and $|\theta_0| < \pi$. The quantity $\gamma / k_{\perp}^2$ is a rough quasilinear estimate for the transport diffusion coefficient of the mode. To estimate $k_{\perp}$ for each mode, we find the $\theta$ location with the largest eigenmode amplitude, and calculate $k_{\perp}$ at that location. In (a), we show the growth rates versus $\theta_0$ and $k_y \rho_i$. There is a notable maximum in the growth rate at $k_y \rho_i \approx 80$ and $\theta_0 = 0$ (which corresponds to a slab ETG mode). In (b) we show the quantity $\gamma / k_{\perp}^2$ — normalized and presented as the dimensionless parameter $\gamma a / v_{ti} k_{\perp}^2 \rho_i^2$ — versus $\theta_0$ and $k_y \rho_i$. We observe that $\gamma / k_{\perp}^2$ has its largest values across a wide range of $k_y \rho_i $ and $\theta_0$ scales, $5 \lesssim k_y \rho_i \lesssim 100$ and $|\theta_0| \lesssim 1.5$. Most of these modes are toroidal ETG, although when $\theta_0 = 0$ and $k_y \rho_i \gtrsim 5$, the fastest growing mode is a slab ETG mode. We stress that the quantity $\gamma / k_{\perp}^2$ is only an approximate measure, and that nonlinear simulations will be needed to ascertain which modes are most important for transport. In (c), we plot the $|\theta|$ location of the maximum of $|\phi ^{tb} _1|$, denoted as $|\theta_{\mathrm{Max}}|$; we see that modes with large values of $\gamma / k_{\perp}^2$ tend to have $0 \lesssim |\theta_{\mathrm{Max}}| \lesssim \pi/2$. In (d), for each $k_y \rho_i$ we plot the normalized value of $\gamma/ k_{\perp}^2$ that is maximum over $\theta_0$. This plot demonstrates that there is a comparable quasilinear diffusion coefficient estimate for all fastest growing modes between $1 \lesssim k_y \rho_i \lesssim 100$, and hence suggests that a wide range of $k_y \rho_i$ values might be important for transport. While significant heat might be transported by toroidal ETG modes, they are unlikely to transport particles because the ions are very close to adiabatic (see (b)). However, since the ions are not fully adiabatic for the slab ETG at lower $k_y \rho_i$ (see (b)), the long wavelength slab ETG instability could cause particle transport. Finally, the ‘extended ETG’ modes, which are the fastest growing modes for $0.1 \lesssim k_y \rho_i \lesssim 1$ (see \[sec:smallkrimodes\]), can also have a large non-adiabatic ion response, and thus they too, may cause particle transport. Next, we show how the values of $\theta_0$, $\theta_{\mathrm{min} }$, and $\hat{s}$ determine the critical temperature gradient of the toroidal ETG mode. Critical $R_0 / L_{Te}$ {#sec:critR0LTe} ----------------------- We now discuss the critical temperature gradient for the toroidal ETG instability that we are studying. We find critical $R_0/L_{Te}$ values as large as $R_0/L_{Te} \approx 40$ for toroidal ETG modes in the pedestal (see (a) and (b), and (a)), significantly larger than in the core. Unless mentioned otherwise, the quantity $\eta_e$ will be kept fixed, to prevent the ETG from becoming stable due to $\eta_e$ being less than its critical value. We want to understand the dependence of the critical $R_0/L_{Te}$ on different parameters. Recall from (a) that there exists a stability boundary $\omega_{*e} \eta_e / \omega_{\kappa e}$ for the toroidal ETG mode; that is, for instability we require $$\frac{\omega_{*e} \eta_e}{\omega_{\kappa e}} > C. \label{eq:Cequation}$$ For $b_e = 0$, $C \simeq 2$. Given that $\omega_{*e} \eta_e / \omega_{\kappa e} \sim R_0 / \hat{s} \theta L_{Te}$, and that $\hat{s}$ and $R_0/L_{Te}$ are fixed parameters, the only free parameter in our scaling theory for the ratio $\omega_{*e} \eta_e / \omega_{\kappa e}$ for a given equilibrium is $\theta$ (note that $C$ in is weakly dependent on $\theta$, because $C$ depends on $b_e$, which in turn depends on $\theta$). For the toroidal ETG mode to be unstable we then require $$\frac{R_0}{\hat{s} L_{Te}} \frac{1}{C} \gtrsim \theta \gtrsim \theta_{\mathrm{min} }. \label{eq:thetaboundsinstability}$$ The quantity $\theta_{\mathrm{min}}$ is determined by the profiles of $\omega_{*e} \eta_e/ \omega_{\kappa e}$ and $\Gamma_0$ (see discussion at start of ). If a simulation only resolves up to $\theta < \theta_{\mathrm{min}}$ in ballooning space (or equivalently insufficiently large values of $|K_x|$), a toroidal ETG mode might incorrectly appear to be stable. Numerical results have shown that $ \theta_{\mathrm{min} } $ is only very weakly dependent on $R_0 / L_{Te}$, but can be strongly dependent on $\theta_0$, and on $\hat{s}$ for large values of $\hat{s}$. For now we set $\theta_0 = 0$, but will soon consider the $\theta_0 \neq 0$ case. Thus, from we obtain a critical gradient, $R_0 / L_{Te}^{\mathrm{crit} }$, $$\frac{R_0}{L_{Te}^{\mathrm{crit}} } \approx \hat{s} \theta_{\mathrm{min} } C. \label{eq:RLTecrit}$$ ![Stability plots of the toroidal ETG mode with $k_y \rho_i = 2.8$. (a): growth rate scan in $R_0/L_{Te}$ with $\eta_e$ and $\eta_i$ fixed for three values of $\hat{s}$. (b): growth rate scan in $R_0/L_{Te}$ with $R_0/L_n$ and $R_0/L_{Ti}$ fixed for three values of $\hat{s}$. (c): eigenmodes corresponding to values of $R_0/L_{Te}$ denoted by the markers in (a). (d): eigenmodes corresponding to values of $R_0/L_{Te}$ denoted by the markers in (b). (e): the quantity $\omega_{*e} \eta_e / \omega_{\kappa e}$ for three values of $\hat{s}$, where $R_0/L_{Te} = 26, \; \eta_e = 4.3$. (f): the quantity $\Gamma_0(b_e)$ for three values of $\hat{s}$.[]{data-label="fig:RLTeandShatThetaminScan"}](RLTescangrowthratearrayJan162020.eps){width="\textwidth"} We first demonstrate the $\hat{s}$ and $\theta_{\mathrm{min} }$ scaling of the critical temperature gradient by performing a scan in $R_0 / L_{Te}$ for three different values of $\hat{s}$, shown in (a). Here, $\eta_e$ and $\eta_i$ are held fixed to avoid the $\eta_s$ stability boundary. This scan is performed in GS2 for $k_y \rho_i = 2.8$ with the standard pedestal equilibrium we have used before, except for changing the value of $\hat{s}$. In (a), we see that $\theta_{\mathrm{min} } \simeq 2$ for $\hat{s} = 3.4$, as shown by the eigenmode in (c). For this value of $\hat{s}$, the eigenmode can have a relatively small value of $\theta_{\mathrm{min} }$ because of the bad curvature region ($\omega_{*e} \eta_e / \omega_{\kappa e} > 0$) that appears at $\theta \simeq 2$ in (e). Once $\hat{s}$ is decreased, the smallest possible value for the mode appears to be $\theta_{\mathrm{min} } \simeq 8.5$, as shown in (c) and (e). Due to the scaling of $R_0/L_{Te}^{\mathrm{crit}}$ in , a much larger value of $\theta_{\mathrm{min} }$ causes $R_0/L_{Te}^{\mathrm{crit}}$ to increase, shown in (a). Both the cases $\hat{s} = 0.8$ and $\hat{s} = 1.7$ have the same value of $\theta_{\mathrm{min} } \simeq 8.5$, but the $\hat{s} = 1.7$ case has a much higher $R_0/L_{Te}^{\mathrm{crit}}$ due to its value of $\hat{s}$ being larger. Thus, we have demonstrated that increasing both $\hat{s}$ and $\theta_{\mathrm{min} }$ increases $R_0/L_{Te}^{\mathrm{crit}}$ for the toroidal ETG mode. As mentioned above, there is another critical value of $R_0/L_{Te}$ that occurs due to $\eta_e$ being too small[@Jenko2001]. (b) shows a scan in $R_0/L_{Te}$ and $\hat{s}$ with $R_0/L_{n}$ and $R_0 / L_{Ti}$ fixed, allowing $\eta_e$ to vary; here, we see that the critical value of $\eta_e$ for the toroidal ETG mode is $\eta_e \approx 1.3$. Interestingly, for smaller values of $R_0/L_{Te}$ we find a very weakly driven slab ITG mode. ![Stability plots of the toroidal ETG mode with $k_y \rho_i = 2.8$. (a): growth rate scan in $R_0/L_{Te}$ with $\eta_e$ and $\eta_i$ fixed for four values of $\theta_0$. (b): corresponding eigenmodes at locations indicated by markers in (a). (c): the quantity $\Gamma_0(b_e)$ for different values of $\theta_0$. (d): the ratio $\omega_{*e} \eta_e / \omega_{\kappa e}$ for different values of $\theta_0$, using $R_0/L_{Te} = 26$.[]{data-label="fig:RLTeandShatScanetaifixed"}](RLTescangrowthratearraytheta0andetaivariedshat3p4Jan17.eps){width="\textwidth"} The above arguments assumed that $|\theta_0| \ll |\theta|$. The critical temperature gradient is also modified by $\theta_0$. As discussed previously, larger values of $|\theta_0|$ can allow a new region of bad curvature to appear at small values of $|\theta|$, as shown in (d). Allowing $\theta_0 \neq 0$, for instability, we require $$\frac{R_0}{L_{Te}} \gtrsim \hat{s} |\theta - \theta_0| C. \label{eq:theta0critscaling}$$ We expect that for nonzero $\theta_0$, $\theta$ and $\theta_0$ have opposite signs because the mode will grow faster where $\omega_{*e} \eta_e/ \omega_{\kappa e} \sim R_0/L_{Te} \hat{s} |\theta- \theta_0|$ is smallest, giving the critical temperature gradient $$\frac{R_0}{L_{Te}^{\mathrm{crit} }} \approx \hat{s}( |\theta_{\mathrm{min} }| + |\theta_0| ) C. \label{eq:theta0critscalingtheta0large}$$ Consistent with this idea, we see that for $|\theta| \lesssim 6$ the only accessible bad curvature regions appear when $\theta \theta_0 < 0$ and when $|\theta_0|$ is sufficiently large. To demonstrate the scaling in , we performed a scan in $\theta_0$ and $R_0/L_{Te}$ at fixed $\hat{s}$, $\eta_e$, and $\eta_i$, shown in (a); we observe that $R_0/L_{Te}^{\mathrm{crit} }$ indeed increases with $\theta_0$ as expected. Furthermore, the assumption that $\theta_{\mathrm{min} } \theta_0 < 0$ is also shown to be correct, as seen by the eigenmodes in (b). Finally, we briefly discuss the effect of the difference between $\omega_{\kappa e}$ and $\omega_{\nabla B e}$ on toroidal ETG stability. Throughout this paper, we have exclusively used $\omega_{*e} \eta_e / \omega_{\kappa e}$ for our analysis, which is justifiable if $\omega_{\kappa e} \simeq \omega_{\nabla B e}$ in the parallel vicinity of where the toroidal mode is most unstable. While this is true for $|\theta| \gtrsim \pi$ (see (b)), for $|\theta| \lesssim \pi$, the value of $\omega_{\kappa e} / \omega_{\nabla B e}$ in bad curvature regions can be as large as 1.5 in a sufficiently-wide parallel region for some values of $\theta_0$. Thus, we might expect as much as a 25% increase in the linear stability boundary compared to the case where one artificially sets $\omega_{\kappa e} = \omega_{\nabla B e}$. Therefore, for certain values of $k_y \rho_i$ and $\theta_0$, the stability boundary for the toroidal ETG mode is increased when $\omega_{\kappa e} > \omega_{\nabla B e}$, which is consistent with previous work [@Bourdelle2003]. To summarize, we have demonstrated that the value of $R_0/L_{Te}^{\mathrm{crit}} $ for toroidal ETG depends on $\hat{s}, \theta_{\mathrm{min} }$, and $\theta_0$. Most relevant to the Miller equilibrium of JET discharge 92174, scans in $\theta_0$ at fixed $\hat{s} = 3.4$ showed $R_0/L_{Te}^{\mathrm{crit}} \approx 13 - 32$, depending on the value of $\theta_0$. This is a much higher value of $R_0/L_{Te}^{\mathrm{crit}} $ than is typically observed in the core (for example, $R_0/L_{Te}^{\mathrm{crit}} \approx 3$ for Cyclone Base Case toroidal ITG). This new type of stability boundary for toroidal ETG directly results from the importance of the radial component of the magnetic drift, in contrast to the core, where the $\nabla y$ component of the drift is usually considered more important. ITG Instability in JET Shot 92174 {#sec:ITG} ================================= In this section, we discuss the ITG instability in JET shot 92174. Previous works have emphasized the importance of ITG instability in the pedestal [@Hatch2017; @Hatch2019; @Wang2012; @Saarelma2013; @Chen2018]. In this work, we find that with the measured $T_{0i}$ profiles, the ITG growth rate is extremely small compared with the ETG instability growth rate. This is due to $R_0 / L_{Ti}$ and $\eta_i$ being relatively small, and electron collisions that decrease the ITG growth rates. If we increase the ion temperature profiles to make them equal to the electron temperature profiles and we ignore the $\mathbf{ E} \times \mathbf{ B}$ shear, the ITG instability is the fastest growing mode at very large scales, $k_y \rho_i \sim L_{Ti}/R_0$. This finding is entirely consistent with ’s results, as the same arguments can equally be applied to ITG (since $R_0/L_{Ti} \gg 1$). While this section will discuss ITG for $\theta_0 = 0$, we also performed a scan in $\theta_0$, to see if any other $\theta_0 \neq 0$ values could be unstable at $k_y \rho_i \lesssim 1$ using the measured ion temperature profile. We found no significant increase in growth rates due to $\theta_0$ with the measured ion profiles. ![Linear ITG and ETG GS2 growth rates at (a): $k_y \rho_i \sim L_{Ti}/R_0$ (ITG scales) and (b): $k_y \rho_e \sim L_{Te}/R_0$ (ETG scales). Dashed series indicates an ITG mode, solid is a mode driven by electron temperature gradients. For the ITG scales, the growth rates and $k_y \rho_i $ have been multiplied by $\rho_i / \rho_e$. The series ‘$T_{0i} = T_{0e}$’ indicates that $T_{0e} = T_{0i}$, $L_{Ti} = L_{Te}$; ‘Measured $T_{0i}$’ indicates that values of $T_{0i}$ and $L_{Ti}$ are taken from the measured ion profiles. Here, $\rho_e / \rho_i \approx 82$ for the measured $T_{0i}$ and $T_{0e}$ profiles.[]{data-label="fig:ITGscan"}](gs2growthspectraisomorphicmultipliedNov24.eps){width="90.00000%"} Due to the symmetry of the collisionless ITG and ETG dispersion relations when $h_e = 0$ for ITG and $h_i = 0$ for ETG, the growth rates of ITG and ETG are isomorphic: $\gamma_{\mathrm{ITG}} = \gamma_{\mathrm{ETG} } \rho_e/\rho_i$ at wavenumbers $k_{y \mathrm{ITG} } = k_{y \mathrm{ETG} } \rho_e/\rho_i$. Here we investigate how the non-adiabatic electron response and a difference in equilibrium profiles in the pedestal break this isomorphism. According to the isomorphism, ITG instability is driven at $k_y \rho_i \sim L_{Ti}/R_0 \ll 1$, and the ETG instability is driven at $k_y \rho_i \sim (\rho_i/\rho_e) L_{Te}/R_0$, as demonstrated in . In , we show the growth rates of ITG at ‘ITG’ scales, $k_y \rho_i \sim L_{Ti}/R_0$, and the growth rates of ETG at ‘ETG’ scales, $k_y \rho_i \sim (\rho_i/\rho_e) L_{Te}/R_0$, for JET shot 92174. The isomorphism between ITG and ETG is confirmed, with the ‘$T_{0i} = T_{0e}, h_e = 0$’ and ‘$T_{0i} = T_{0e}, h_i = 0$’ cases having the same isomorphic growth rates. Here, ‘$T_{0i} = T_{0e}$’ means that both the ion and electron temperatures and their gradients are set equal to each other — specifically, $T_{0i}$ is increased to match $T_{0e}$. This affects the ion collision frequencies, which are decreased self-consistently. Electron collisions have a significant effect on the toroidal and slab ITG growth rates. As shown in (a), there is a substantial difference between the collisional and collisionless simulations, indicated by ‘$T_{0i} = T_{0e}$’ and ‘$T_{0i} = T_{0e}$, Collisionless’ cases. In the simulations we have performed, electron collisions reduce the toroidal and slab ITG growth rates. It is not obvious that electron collisions should always decrease the ITG growth rates. We now describe gyrokinetic simulations with the measured ion profiles. Compared with the equal profile case, ‘$T_{0i} = T_{0e}$’, once measured equilibrium profiles are included, the ITG growth rates decrease substantially. In (a), ‘Measured $T_{0i}$’ is a simulation with the measured ion temperature profiles; the fastest growing modes at ITG scales are electron-driven modes with large electron tails [@Hallatschek2005] (see \[sec:smallkrimodes\]), switching to a toroidal ETG mode once $k_y \rho_i \gtrsim 0.1$. In order to find the subdominant ITG instability, we must set $h_e = 0$ (otherwise electron-driven modes dominate), as shown in the ‘Measured $T_{0i}$, $h_e=0$’ line. The ITG instability barely grows in the runs with adiabatic electrons, although there were well-resolved toroidal ITG eigenmodes. Using GS2’s eigensolver function [@Dickinson2014], we could not find any toroidal ITG instability for $k_y \rho_i \sim L_{Ti}/R_0$ when using the measured profiles and kinetic electrons, indicating that ITG is stable at $k_y \rho_i \ll 1$. However, at ETG scales ($k_y \rho_e \sim L_{Te}/R_0$), we did find weakly growing slab ITG modes by using adiabatic electrons, shown in (b) (‘Measured $T_{0i}$, $h_e = 0$’), a result that was corroborated by very weakly growing slab ITG modes found using GS2’s eigensolver. Therefore, for the measured profiles, ITG is extremely subdominant in JET shot 92174. Moreover, we will see in that the slab ITG is easily quenched by $\mathbf{ E} \times \mathbf{ B}$ shear. Heuristically, we can understand the stability of the toroidal ITG mode using a similar stability analysis performed for the toroidal ETG mode in . In (b), we show the toroidal ETG mode being stabilized at $\eta_e \simeq 1.2$. Due to the isomorphism between toroidal ITG and toroidal ETG in the collisionless case where the other species is adiabatic, we can reasonably predict that toroidal ITG also has a similar critical $\eta_i \approx 1$. Examining the $\eta_i$ profile in (c), we find that $\eta_i \simeq 0.8-1.2$ in the steep gradient region of the pedestal ($r/a \approx 0.97 - 0.99$). Hence, $\eta_i$ is very close to (and likely slightly below) its critical value in all regions of the pedestal for $\theta_0 = 0$, and it is unsurprising that the toroidal ITG mode is very weakly-driven. A broader question that merits examination is the physics that keeps $\eta_i$ close to its critical value, while $\eta_e$ is far above its critical value (although this is subject to uncertainties in the ion temperature profile, which could change $\eta_i$). Finally, the suppression of ITG instability in pedestals is not inconsistent with what has been observed in previous analyses; for example, [@Viezzer2017] found that the ion heat diffusion was close to neoclassical in ASDEX-U inter-ELM pedestal discharges. To summarize, we find that with the measured ion temperature profiles, the ITG mode is stable for $k_y \rho_i \ll 1$, and there is very weakly-driven ITG at $k_y \rho_i \sim 1$. When the ion temperature profile is set equal to the electron profile and ITG modes become linearly unstable at very long wavelengths, the isomorphism between ITG with $h_e = 0$ and ETG with $h_i = 0$ holds. Electron collisions appear to decrease the ITG growth rate significantly. The detailed mechanism for this stabilizing impact of electron collisions requires further investigation. $\mathbf{ E} \times \mathbf{ B} $ Shear {#sec:ExBeffects} ======================================= In this paper we chose to perform most simulations without $\mathbf{ E} \times \mathbf{ B} $ shear, since in simulations with $\mathbf{ E} \times \mathbf{ B} $ shear, the electrostatic modes were barely modified compared to the simulations without $\mathbf{ E} \times \mathbf{ B} $ shear. In this section, we present the results of gyrokinetic simulations with $\mathbf{ E} \times \mathbf{ B}$ shear. First, we discuss the validity of keeping $\mathbf{ E} \times \mathbf{ B}$ shear even though it is small in the low flow ordering. In addition to the results we presented in where KBMs were shown to be suppressed by $\mathbf{ E} \times \mathbf{ B}$ shear, we also show the effect of $\mathbf{ E} \times \mathbf{ B}$ shear on ETG and ITG modes. We will see that while KBMs usually easily suppressed by $\mathbf{ E} \times \mathbf{ B}$ shear, ETG modes are barely affected. ITG instability is easily stabilized when using the measured ion temperature profile, but is not fully-suppressed when the ion temperature profile is made equal to the electron temperature profile. In our local linear simulations with $\mathbf{ E} \times \mathbf{ B}$ shear, we use a new $\mathbf{ E} \times \mathbf{ B}$ shear algorithm [@Christen2019], and also tested that the results were qualitatively similar with the previous GS2 algorithm [@Hammett2006]. With the newer algorithm, a typical simulation with $\mathbf{ E} \times \mathbf{ B}$ shear contained a single poloidal mode, 150 radial wavenumbers with a spacing of $\Delta k_x \approx k_y$, and a $\mathbf{ E} \times \mathbf{ B}$ shear value of $\gamma_E a / v_{ti} = 0.56$. With the previous algorithm, the range of $k_x$ values was held fixed, but the $\Delta k_x$ spacing was reduced by a factor of 10. In the low flow ordering, if one retains the $\mathbf{ E} \times \mathbf{ B} $ shear, one should also keep neoclassical corrections to the Maxwellian [@Parra2015; @Lee2014two], but for simplicity, we have neglected neoclassical corrections throughout this paper. When analyzing high $k_{\perp}$ modes for this equilibrium, it is inconsequential whether or not the $\mathbf{ E} \times \mathbf{ B}$ shear is kept, and we expect the neoclassical corrections to be similarly unimportant. However, for small $k_{\perp}$, we find the small $\mathbf{ E} \times \mathbf{ B}$ shear can suppress instabilities and hence one might expect that neoclassical corrections are also important. The parallel flow is one of the main physical features of neoclassical corrections. Therefore to estimate the effect of these corrections, we will use previous studies on the parallel velocity gradient (PVG) instability [@Catto1973; @Newton2010; @Barnes2011b; @Schekochihin2012; @Lee2015; @Ball2019]. The PVG growth rate is $$\gamma_{\mathrm{PVG} } \sim \frac{d u_{\zeta i p}}{d r} k_y \rho_i.$$ In regions where we see ITG stabilization by $\mathbf{ E} \times \mathbf{ B}$ shear, $k_y \rho_i \sim 0.1$, and the PVG growth rate is much smaller than the $\mathbf{ E} \times \mathbf{ B}$ shear rate. From the measured ${}_6^{12}C^+$ rotation profiles at $r/a = 0.974$, we find that $|du_{\zeta i p} / dr| a / v_{ti} \approx 1.4$, and thus $\gamma_{\mathrm{PVG} } a / v_{ti} \approx 0.14$. Therefore, given that $\gamma_E a / v_{ti} = 0.56 > \gamma_{\mathrm{PVG} } a / v_{ti}$, this PVG mode is likely stabilized by the $\mathbf{ E} \times \mathbf{ B}$ shear. Hence, we do not expect that the neoclassical flows will significantly modify a mode’s growth rate, although the effect of neoclassical terms at these small scales merits further investigation. The $\mathbf{ E} \times \mathbf{ B}$ shear is usually more effective for low than for high $k_{\perp}$ modes, as shown in . This is because the growth rate of the electrostatic instabilities that we are investigating typically scales with $\omega_{*s} \eta_s \sim k_y \rho_s v_{ts} / L_{Ts} $, and because of the differences in a mode’s radial extent for different instabilities. If the typical timescale for an instability, $1/ \gamma$, is comparable to the $\mathbf{ E} \times \mathbf{ B} $ shearing time, $1/\gamma_E$, the $\mathbf{ E} \times \mathbf{ B}$ shear can be effective. However, when $1/\gamma_E \gg 1/\gamma \sim L_{Ts}/ k_y \rho_s v_{ts}$, the $\mathbf{ E} \times \mathbf{ B}$ shear is unable to shear the mode sufficiently quickly. Hence, $\mathbf{ E} \times \mathbf{ B} $ shear suppresses modes at smaller $k_y$, and barely modifies short wavelength modes. Additionally, modes that are radially localized ($K_x \gg k_y$) are harder to shear than those with a wider radial width. ![Density time traces of KBM and ITG instabilities with and without $\mathbf{ E} \times \mathbf{ B}$ shear. (a): the KBM is suppressed by the $\mathbf{ E} \times \mathbf{ B}$ shear consistent with the measured ion temperature profile. (b) the ITG is not fully suppressed by the $\mathbf{ E} \times \mathbf{ B}$ shear when the ion temperature and gradient is equal to the electron temperature and gradient. The two separate values of $\gamma_E a / v_{ti}$ correspond to its consistent value for the measured ion temperature profile ($\gamma_E a / v_{ti} = 0.56$) and when the ion temperature profile is equal to the electron temperature profile ($\gamma_E a / v_{ti} = 2.24$). (c): the effective growth rates of the ITG instability for the three separate values of $\gamma_E a / v_{ti}$ in (b).[]{data-label="fig:flowshearresults"}](n2versustallflowshearpaperplotallshotsALLTHREEflowshearPAPERleftnoflowshearrightwithflowshearNO4tprimApril2TwoSubplots.eps){width="\textwidth"} We now apply these two criteria (growth rate versus shearing rate, and radial extent of the mode) to explain our observations for which modes are suppressed by $\mathbf{ E} \times \mathbf{ B}$ shear. The KBM we discussed in is easily suppressed by $\mathbf{ E} \times \mathbf{ B}$ shear because it is radially extended and is stable for a wide range of $\theta_0$ values (see (d)). The KBM was shear suppressed even though $\gamma_{\mathrm{KBM} } > \gamma_E$. This suppression is demonstrated in (a), where the mode’s density is shown to decay in time. Determining the effect of the $\mathbf{ E} \times \mathbf{ B}$ shear on toroidal and slab modes separately is challenging. To understand why this is the case, it will be useful to define an ‘effective’ $\theta_0$ that now depends on time, $$\Theta_0(\gamma_E,t) = \theta_0 - k_y \frac{\gamma_E}{\hat{s}} t, \label{eq:Theta0time}$$ such that the time-dependent radial wavenumber is $$K_{x} = k_y \bigg{(} \hat{s}(\theta_0 - \theta) - \frac{r}{q} \frac{\partial \nu}{\partial r} \bigg{)} - k_y \gamma_E t = k_y \bigg{(} \hat{s} (\Theta_0-\theta) - \frac{r}{q} \frac{\partial \nu}{\partial r}. \label{eq:effectivewavenumberwithflowshear} \bigg{)}.$$ The fact that the mode has different $\Theta_0$ values at different times considerably complicates understanding the effect of $\mathbf{ E} \times \mathbf{ B}$ shear on toroidal and slab ETG in the pedestal separately: for $k_y \rho_i \gtrsim 5$ in the absence of $\mathbf{ E} \times \mathbf{ B}$ shear, while for $\theta_0 = 0$ the fastest growing modes are slab ETG modes, for $\theta_0 \neq 0$ the fastest growing modes are almost always toroidal ETG modes. Since $\mathbf{ E} \times \mathbf{ B} $ shear changes $\Theta_0$ with time as described in , if at $t = 0$ a mode is a slab ETG mode (i.e. it has $\theta_0 = 0$), after a period of time it will become a toroidal ETG mode. Therefore, we can only determine if the $\mathbf{ E} \times \mathbf{ B}$ shear suppresses both slab and toroidal modes. We now consider the effect of $\mathbf{ E} \times \mathbf{ B}$ shear on the ITG instability. Our simulations indicate that the effectiveness of $\mathbf{ E} \times \mathbf{ B}$ at suppressing ITG is sensitive to several parameters. We first test the effectiveness of $\mathbf{ E} \times \mathbf{ B} $ shear with the measured ion temperature profiles, which requires using adiabatic electrons, since electron temperature gradient-driven modes are the fastest growing at all scales (see ). We test the $\mathbf{ E} \times \mathbf{ B} $ shear on an ITG mode with $k_y \rho_i = 0.7$, which has a modest growth rate of $\gamma a /v_{ti} \simeq 0.1$. In simulations with $\mathbf{ E} \times \mathbf{ B}$ shear, the mode is easily suppressed. This is expected, since $1/\gamma_E \ll 1/\gamma$ for this ITG mode, and hence, both toroidal and slab ITG are suppressed by $\mathbf{ E} \times \mathbf{ B}$ shear at $k_y \rho_i = 0.7$ with the measured ion temperature profiles. We also test the effectiveness of the $\mathbf{ E} \times \mathbf{ B} $ shear at suppressing the ITG instability when the ion temperature profiles are made equal to the electron temperature profiles (that is, $T_{0i} = T_{0e}$ and $L_{Ti} = L_{Te}$). To investigate this, we perform a GS2 simulation with $\mathbf{ E} \times \mathbf{ B}$ shear for a single toroidal ITG mode with $k_y \rho_i = 0.04$. Recall that we estimate the radial electric field by balancing it with the pressure gradient as in , which requires that $\gamma_E$ is roughly proportional to the second derivative of the pressure gradient, as in . Therefore, when we quadruple $1/L_{Ti}$ for the case where the ion and electron temperature profiles are made equal, to be consistent with the temperature profile we must also roughly quadruple the value of $\gamma_E$. In (b), we show the time trace of the density for three simulations of the ITG mode with $T_{0i} = T_{0e}, \; L_{Ti} = L_{Te}$, where the value of $\gamma_E$ varies in each simulation. We show the ITG mode in the absence of $\mathbf{ E} \times \mathbf{ B}$ shear, the mode with $\gamma_E a/v_{ti} = 0.56$, which is consistent with the measured ion temperature gradients, and the mode with $\gamma_E a/v_{ti} = 2.24$, which is consistent with the steepened ion temperature gradients. To calculate the effective growth rate, we used a similar technique to that in [@Roach2009], which involves fitting the mode amplification in time. As shown in (c), while the consistent value of $\mathbf{ E} \times \mathbf{ B}$ shear, $\gamma_E a/v_{ti} = 2.24$, reduces the growth rate by 70 %, it does not fully suppress the ITG instability. We also found a range of additional parameters that determined how successfully the $\mathbf{ E} \times \mathbf{ B}$ shear suppressed the high gradient ITG mode such as $T_{0i}/T_{0e}$; more work is required to understand the resilience of strongly-driven pedestal ITG to $\mathbf{ E} \times \mathbf{ B}$ shear. We now discuss the ETG instability. We found that $\mathbf{ E} \times \mathbf{ B} $ shear was insufficient to quench the ETG modes. Even tripling the value of $\gamma_E$ at $k_y \rho_i = 2.8$ barely changed the growth rates of the toroidal and slab ETG modes. The ineffectiveness of the $\mathbf{ E} \times \mathbf{B} $ shear for ETG modes is due to $\gamma \gg \gamma_E$ for these modes. There is likely no experimentally-realizable value of $\gamma_E$ that would suppress these ETG modes in the pedestal. Thus, to summarize, we establish the following hierarchy for the efficiency of $\mathbf{ E} \times \mathbf{ B} $ shear at reducing the growth rates of linear modes. KBMs are completely suppressed by $\mathbf{ E} \times \mathbf{ B} $ shear, and ITG is also fully suppressed when using the measured ion temperature profiles. Using profiles with ion gradients as steep as the electron gradients, while the toroidal ITG growth rate is significantly reduced by $\mathbf{ E} \times \mathbf{ B}$ shear, it is not necessarily stabilized. ETG is very resistant to $\mathbf{ E} \times \mathbf{ B}$ shear. Discussion {#sec:discussion} ========== In the steep gradient region of the fully developed pedestal of a JET H-mode discharge (92174) where measurements indicate that $T_{0i} > T_{0e}$ and $R_0/L_{Te} > R_0/L_{Ti}$, local gyrokinetic simulations demonstrate that electron-driven modes are the fastest growing modes at all length scales perpendicular to $\mathbf{B} $. Linearly, KBMs are quenched by $\mathbf{ E} \times \mathbf{ B}$ shear, as is ITG when the measured ion temperature profiles are used. This leaves ETG at $0.1 \lesssim k_y \rho_i \lesssim 400$. Using $R_0/L_{Te} \gg 1$, we predicted that a novel type of toroidal ETG would be driven at $k_y \rho_i \sim 1$ and $K_x \rho_e \sim 1$, which we have confirmed in gyrokinetic simulations. This toroidal ETG at $k_y \rho_i \sim 1$ in the linear growth rate spectrum seems to be a robust feature of steep temperature gradient regions, having been seen in all three other pedestals we examined (see , and \[app:dischargeparams\] for experimental information), as well as in other works: DIII-D [@Fulton2014; @Kotschenreuther2019], NCSX [@Baumgaertel2011], and ASDEX-U [@Told2008; @Jenko2009; @Told2012; @Kotschenreuther2019]. It is also likely that a toroidal ITG mode of a similar nature has been observed at $k_y \rho_i \sim L_{Ti}/ R_0$ in [@Bowman2018]. A notable success of this work is that a simple theoretical model predicted the linear growth rates of the toroidal and slab ETG and the poloidal location of the toroidal ETG mode fairly well. If the ion temperature profile is set equal to the electron temperature profile, ITG modes grow fastest for $k_y \rho_i \lesssim 0.5$, and ETG modes grow fastest for $0.5 \lesssim k_y \rho_i \lesssim 400$. With equal ion and electron temperature profiles, one might be concerned about significant transport caused by the toroidal ITG at scales as small as $k_y \rho_i \sim L_{Ti}/ R \ll 1$, since nonlinearly these instabilities might produce large eddies that cause substantial heat transport. However, our simple estimate of $\gamma_E$ by balancing the radial electric field with the pressure gradient found that $\mathbf{ E} \times \mathbf{ B}$ shear could fully suppresses the ITG instability for certain temperature ratios $T_{0i}/T_{0e}$ when the ion temperature gradients are as steep as the electron temperature gradients. While the $\mathbf{ E} \times \mathbf{ B}$ shear frequency is too small to damp the ETG, impurities are known to damp ETG [@Jenko2001; @Reshko2008]. Therefore, further investigation might explore the effect of impurities on toroidal ETG instability in pedestals. Work has already shown that impurities can produce non-negligible ion-scale pedestal transport [@Kotschenreuther2017; @Hatch2019]. With the measured ion temperature profiles, it is likely that the nonlinear state of JET shot 92174’s pedestal is dominated by electron-driven transport. Indeed, the novel toroidal ETG modes we have described in this work could be important for transport, as evidenced by the heuristic estimate of $\gamma/k_{\perp}^2$ in . Careful work will be required to resolve these modes in nonlinear simulations. We have not included results from nonlinear simulations in this paper because the linear results of this work demonstrate how challenging these simulations are to correctly resolve. For example, to resolve the fastest growing linear modes — toroidal ETG modes — from $1 \lesssim k_y \rho_i \lesssim 100$ in a nonlinear simulation requires significant $k_x$ resolution, as well as a sufficiently large number of independent $\theta_0$ modes. In addition, the slab ETG modes require increasingly fine $\theta$ grids to resolve at higher values of $k_y \rho_i$, which significantly increases computational cost. Caution is required in attempting to infer transport properties from these linear results: the modes we observe span a wide range of perpendicular scales, and complex multiscale interactions could be important [@Maeyama2015; @Howard2016b; @Maeyama2017; @Hardman2019; @Hardman2019b]. While in this paper we have focused on a single radial location for a single discharge, we have also investigated the growth rates at various radial locations using gyrokinetic simulations. These simulations have demonstrated a significant sensitivity of the growth rates to the radial location because of the sensitivity of the instabilities to local gradients. Nevertheless, certain features such as (i) the dominance of ETG at all scales, and (ii) the toroidal ETG at $k_y \rho_e \sim L_{Te}/R_0$ were robust features. Due to the sensitivity of microstability to the radial location, we caution against using the local growth rates at any given flux surface to infer global properties about the pedestal, such as its width or height. We have observed that some pedestals have consistently lower growth rates than others, but more work, particularly nonlinear simulations, is required to connect gyrokinetic analysis with predictions of pedestal structure. Acknowledgements {#acknowledgements .unnumbered} ================ We gratefully acknowledge useful conversations with I. G. Abel, T. Adkins, O. Beeke, N. Christen, D. Dickinson, M. R. Hardman, W. Guttenfelder, Y. Kawazura, A. Mauriya, J. Ruiz Ruiz, S. Saarelma, A. A. Schekochihin, and D. A. St-Onge. JFP is supported by EPSRC Scholarship No 3000207032. FIP and MB are supported in part by the Engineering and Physical Sciences Research Council (EPSRC) \[Grant Number EP/R034737/1\]. CMR is supported in part by the RCUK Energy-Programme (grant number EP/EPI501045). PGI is supported by an EU H2020 grant, agreement No 3000207035. Computational time provided by Plasma HEC Consortium EPSRC (grant number EP/L000237/1). This work has been carried out within the framework of the EUROfusion Consortium and has received funding from the Euratom research and training programme 2014-2018 and 2019-2020 under grant agreement No 633053 and from the RCUK Energy Programme (grant number EP/T012250/1 and EP/P012450/1). Withal, this work has been carried out within the framework of the Contract for the Operation of the JET Facilities and has received funding from the European Union’s Horizon 2020 research and innovation programme. The views and opinions expressed herein do not necessarily reflect those of the European Commission. The authors acknowledge EUROfusion, the EUROfusion High Performance Computer (Marconi-Fusion), the use of ARCHER through the Plasma HEC Consortium EPSRC grant numbers EP/L000237/1 and EP/R029148/1 under the projects e281-gs2, and software support from Joseph Parker through the Plasma-CCP Network under EPSRC grant number EP/M022463/1. This work also received computational support from Joseph Parker of CoSeC (the Computational Science Centre for Research Communities), funded through CCP-Plasma/Plasma-HEC EPSRC grants EP/R029148/1 and EP/M022463/1. JFP acknowledges travel support from Merton College, Oxford. We are grateful for the hospitality of the Wolfgang Pauli Institute, University of Vienna. Other Discharges {#app:dischargeparams} ================ Discharge 82550 92167 92168 92174 ---------------------------------------- -------- -------- -------- -------- Experimental Parameters $I_p$ \[MA\] 2.5 1.4 1.4 1.4 $B_{T0}$ \[T\] 2.7 1.9 1.9 1.9 $H_{98(y,2)}$ 0.7 0.9 1.0 1.0 $n_G$ 0.8 0.6 0.7 0.7 $R_D$ \[electrons/s $\times 10^{22}$\] 2.3 0.8 0.4 0.9 $q_{95}$ 3.3 4.3 4.4 4.2 $\mathrm{Z}_{\mathrm{eff} } $ 1.2 1.8 1.8 1.8 $P_{\mathrm{NBI}}$ \[MW\] 14.4 17.4 17.6 17.4 $\beta_N$ 1.1 2.2 2.6 2.5 Simulation Parameters $r/a$ 0.9660 0.9784 0.9713 0.9743 $q$ 3.65 5.14 5.07 5.08 $\hat{s}$ 4.92 3.93 4.62 3.36 $a/L_{Te}$ 57 41 29 42 $a/L_{Ti}$ 12 19 16 11 $a/L_{n}$ 23 8 10 10 $\kappa$ 1.61 1.54 1.54 1.55 $\delta$ 0.30 0.26 0.26 0.26 $a \beta'$ -0.09 -0.06 -0.07 -0.08 $dR_M/dr$ -0.17 -0.34 -0.36 -0.35 $a(d \kappa / dr)$ 1.11 1.15 0.81 0.95 $a(d \delta / dr)$ 0.97 0.85 0.67 0.74 : Experimental and simulation parameters for the discharges in this work.[]{data-label="tab:expdetails"} Here we present the results of gyrokinetic analysis for three other JET-ILW H-mode pedestal discharges. The basic experimental and simulation parameters for these JET-ILW discharges in addition to the discharge discussed in the main text (shot 92174) are shown in . Discharge 82550 is a very highly-fueled deuterium discharge with high triangularity and low ion temperature, 92167 is a highly-fueled deuterium discharge, 92168 is is a weakly-fueled deuterium discharge, and 92174 is a highly-fueled deuterium discharge with deuterated ethylene ($\mathrm{C}_2 \mathrm{D}_4$) injection. In , the quantity $q_{95}$ is the safety factor measured at the location where the normalized poloidal flux is equal to $0.95$. For more information on these data types, refer to the JET data handbook. ![GS2 gyrokinetic pedestal electrostatic growth rates for 4 JET equilibria with $\theta_0 = 0$ for different ranges of $k_y \rho_i $. (a) $1 \lesssim k_y \rho_i \lesssim 135$. (b) $0.1 \lesssim k_y \rho_i \lesssim 1.0$. (c) $1 \lesssim k_y \rho_i \lesssim 5$. (d) $5 \lesssim k_y \rho_i \lesssim 50$.[]{data-label="fig:allspectra"}](gs2growthspectrashotslaballscalesfullrangeno92168.eps){width="\textwidth"} shows results from local gyrokinetic microinstability analysis at the radial location with the maximum pressure gradient (and therefore close to the maximum $\gamma_E$) in the four JET-ILW H-mode pedestals described in . These are all electrostatic, linear GS2 simulations performed without $\mathbf{ E} \times \mathbf{ B}$ shear and with $\theta_0 = 0$. While JET shot 92168 does not appear to have the characteristic toroidal ETG bump at $k_y \rho_i \sim 1$, an analysis of the eigenmodes demonstrates that toroidal ETG modes are indeed the fastest growing modes for $1 \lesssim k_y \rho_i \lesssim 7$ with $\theta_0 = 0$. Electrostatic modes at $k_y \rho_i \lesssim 1.0$ {#sec:smallkrimodes} ================================================ ![Eigenmodes for $k_y \rho_i \lesssim 1$ and $\theta_0 = 0.05$ for JET shot 92174. In (a), (b), and (c), the quantity $\omega_{*e} \eta_e/ \omega_{\kappa e}$ is plotted only when it is positive. In (a)-(d), the crimson lines are $\mathrm{Re}(\phi ^{tb} _1)$, the blue lines are $\mathrm{Im}(\phi ^{tb} _1)$, the gold dashed lines are $|\phi ^{tb} _1|$, and the black dashed lines are $\omega_{*e} \eta_e/ \omega_{\kappa e}$. (a): $k_y \rho_i = 0.97$, toroidal ETG with large amplitude far down the field line. (b): $k_y \rho_i = 0.62$, extended ETG, (c): $k_y \rho_i = 0.34$, extended ETG, and (d): $k_y \rho_i = 0.09$: modes with electron tails. Growth rates for $k_y \rho_i \lesssim 1.0$ modes with scans in temperature gradients, collisions, and kinetic/adiabatic ions: (e): $k_y \rho_i < 0.14$ modes, and (f): $0.14 < k_y \rho_i < 1.4$ modes.[]{data-label="fig:lowereigenmodes"}](eigenmodesANDgrowthratesintermediatescalespaperplotwithtailsMar30smalltheta0withparity.eps){width="\textwidth"} For completeness, we briefly detail the electrostatic modes at $k_y \rho_i \lesssim 1.0$. We describe their eigenmode structure as well as growth rate sensitivity scans in temperature gradients and collisionality. All of these simulations are performed with $\theta_0 = 0.05$. For $0.1 \lesssim k_y \rho_i \lesssim 1.0$, we observe modes that become increasingly extended in $\theta$ with decreasing values of $k_y \rho_i$. For $k_y \rho_i \approx 1$, the fastest growing mode is still the toroidal ETG mode described throughout this paper, shown in (a). Once $k_y \rho_i$ decreases, the eigenmodes become more complicated and more extended in $\theta$, as shown by (b) and (c); we refer to these modes as ‘extended ETG’. We also plot the quantity $\omega_{*e} \eta_e / \omega_{\kappa e}$ when it is positive in (a), (b), and (c) — we observe that the extended ETG tends to have maxima of $|\phi ^{tb} _1|$ in bad curvature regions. This leads us to speculate that the extended ETG modes are a more complicated version of the toroidal ETG modes described throughout this paper. The extended ETG modes in (b) and (c) have tearing parity for both $\mathrm{Re}(\phi ^{tb} _1)$ and $\mathrm{Im}(\phi ^{tb} _1)$. We normalize the eigenmodes in (a), (b), (c), and (d) such that the maximum of $|\phi ^{tb} _{1}|$ is 1, and such that the value of $\phi ^{tb} _1 $ is purely real at that location. In (f), we perform a growth rate sensitivity scan for these modes; the growth rate of these extended modes is very sensitive to $R_0/L_{Te}$ and only slightly sensitive to $R_0/L_{Ti}$ and collisions for smaller values of $k_y \rho_i$. The extended ETG modes are stable when run with adiabatic ions for $k_y \rho_i \lesssim 0.2$. For $k_y \rho_i \lesssim 0.1$, we observe extremely extended eigenmodes, shown in (d) — the mode extends as far as $\theta \approx 100$ before the typical $|\phi_1 ^{tb}|$ value is less than 10 % of the eigenmode maximum value. The modes are reminiscent of modes with extended electron tails [@Hallatschek2005]. There is no apparent relationship between the maxima of $|\phi ^{tb} _1|$ and bad curvature regions, unlike for the extended toroidal ETG modes. The mode shown in (d) has tearing parity for both $\mathrm{Re}(\phi ^{tb} _1)$ and $\mathrm{Im}(\phi ^{tb} _1)$. Sensitivity scans in (e) show that these modes are very sensitive to $R_0/L_{Te}$, but insensitive to $R_0/L_{Ti}$. The modes with electron tails were stable for collisionless simulations. Full Dispersion Relation {#app:disprlnfull} ======================== Using in the quasineutrality , we find with $$\fl \eqalign{ D_s & \equiv \bigg{(} \frac{e \phi ^{tb} _1 n_{0e}}{Z_s T_{0e}} \bigg{)}^{-1} \int h_s d^3 v = \frac{2 i Z_s^2}{\pi^{1/2} v_{ts}^3} \frac{T_{0e}}{T_{0s}} \frac{n_{0s}}{n_{0e}} \int_0^{\infty} d \lambda \int_0^{\infty} dv_{\perp} v_{\perp} \int_{-\infty}^{\infty} d v_{\parallel } \\ & \times \exp\bigg{(}i \lambda \left( \arc{\omega} - \sigma \hat{v}_{\parallel }^2 - \arc{\omega}_{\nabla B s} \frac{\hat{v}_{\perp}^2}{2} - \arc { k}_{\parallel} \hat{v}_{\parallel } \right) - \hat{v}_{\parallel }^2 - \hat{v}_{\perp}^2\bigg{)} \\ & \times \bigg{[} - \arc {\omega } + \arc {\omega}_{*s} \bigg{(} 1 + \eta_s \left(\hat{v}_{\parallel }^2 + \hat{v}_{\perp }^2 - \frac{3}{2}\right) \bigg{)} \bigg{]} J_0^2 \left( \sqrt{2 b_s} \hat{v}_{\perp} \right) }$$ where we have used [@Biglari1989] $$\fl i \int_0^{\infty} d\lambda \exp\bigg{(}i \lambda \left(\arc{\omega } - \sigma \hat{v}_{\parallel }^2 - \arc{\omega}_{\nabla B s} \hat{v}_{\perp}^2/2 - \arc {k}_{\parallel} \hat{v}_{\parallel } \right)\bigg{)} = \frac{1}{-\arc{\omega }+ \arc {k}_{\parallel} \hat{v}_{\parallel } +\sigma \hat{v}_{\parallel }^2 + \arc{\omega}_{\nabla B s} \hat{v}_{\perp}^2/2 }.$$ To find growing solutions and obtain a converged integral, we require that $ \mathrm{Im}(\arc {\omega }) > 0$. Evaluating the integral in $\hat{v}_{\parallel }$ gives $$\fl \eqalign{ D_s = & 2 i \frac{Z_s^2}{ v_{ts}^2} \frac{T_{0e} n_{0s}}{ T_{0s} n_{0e}} \int_0^{\infty} d \lambda \int_0^{\infty} dv_{\perp} v_{\perp} \frac{1 }{(1 + i \sigma \lambda)^{1/2}} \exp\bigg{(} i \lambda \arc {\omega } - \hat{v}_{\perp }^2 (1 + i \arc{\omega}_{\nabla B s} \lambda / 2) \bigg{)} \\ & \times \exp\bigg{(}- \frac{(\lambda \arc { k}_{\parallel})^2}{ 4(1 + i \sigma \lambda)}\bigg{)} \\ & \times \bigg{[} - \arc {\omega } + \arc {\omega}_{*s} \bigg{(} 1 + \eta_s \left( \frac{2(1 + i \sigma \lambda) - (\arc {k}_{\parallel}\lambda)^2}{4 (1+i \sigma \lambda)^2} + \hat{v}_{\perp }^2 - \frac{3}{2} \right) \bigg{)} \bigg{]} J_0^2 \left( \sqrt{2 b_s} \hat{v}_{\perp} \right) . }$$ The integral in $\hat{v}_{\perp}$ gives , where we used the integrals $$\int_0^{\infty} x J_0^2 (a x) \exp(-b x^2) d x = \frac{1}{2b} I_0 \bigg{(} \frac{a^2}{2 b} \bigg{)} \exp(-a^2/2b) = \frac{1}{2b} \Gamma_0 \left( \frac{a^2}{2b} \right), \label{eq:BessFunIntegral}$$ and $$\eqalign{ \int_0^{\infty} x^3 J_0^2 (a x) \exp(-b x^2) d x & = \frac{-(a^2 - 2b) \Gamma_0 \bigg{(} a^2/2b \bigg{)} + a^2 \Gamma_1 \bigg{(} a^2/2b \bigg{)}}{4 b^3}, }$$ which is found by differentiating with respect to $b$. We proceed to explain the numerical technique used to calculate the $\lambda$ integral in . The $\lambda$ integral in along the real $\lambda$ axis is highly oscillatory when $\gamma \to 0$, and standard numerical integration methods can make substantial errors in the low growth rate limit. Similarly, a straightforward change of variables such as $\lambda \to i \overline{ \lambda}$ will fail for nonzero $k_{\parallel}$ and $b_s$ due to exponential singularities caused by $k_{\parallel}$ and $b_s$ (at $\lambda = \sigma i$ and $2i / \arc{\omega}_{\nabla B s} $, respectively). To avoid these problems, we introduce a numerically robust path of integration that avoids singularities and significantly reduces the number of oscillations. In the limit $\lambda \to \infty$, the exponential in reduces to, $$\exp \left[ i \left( \arc{\omega} + \frac{\arc{k}_{\parallel}^2}{4 \sigma} \right) \lambda \right].$$ Thus, if we wish to minimize oscillations, we should choose our path such that the imaginary component of the exponential is constant. This is achieved with the integral path $$\lambda = i \left( \arc{\omega}^* + \frac{\arc{k}_{\parallel}^2}{4 \sigma} \right) \overline{\lambda} + a, \label{eq:integrationpath}$$ where $a$ is a constant that we need to choose to improve integral convergence. ![Contour paths $C_0$ and $C_1$, constructed to avoid the poles along the imaginary $\lambda$ axis at $\sigma i$ and $2 \sigma i$, as well as minimizing exponential oscillations.[]{data-label="fig:newstabcontour"}](StabContourApr2Sing){width="\textwidth"} Therefore, we choose an integration path composed of two different paths, $C_0$ and $C_1$. The first path, $C_0$, goes a short distance $a$ along the real $\lambda$ axis. The second path, $C_1$, is the one given in . The total integration path is shown in . The integration path in gives the same result as the original path because the integrand in Equation (51) decays as $|\lambda| \rightarrow \infty$. The constant $a$ needs to be sufficiently large to avoid the singularities at $\lambda = \sigma i$ and $2 i/ \arc{\omega}_{\nabla B s}$. A value $a = 0.5$ is usually sufficiently large. [^1]: Even though the last term in is formally small in $\beta$ in the electrostatic limit, we keep it in all our electrostatic simulations because the large pressure gradients in the pedestal can make it important.
{ "pile_set_name": "ArXiv" }
**[Octonions and vacuum stability]{}** **M.V. Gorbatenko** Russian Federal Nuclear Center - VNIIEF, Sarov, Nizhni Novgorod region E-mail: . ### Abstract {#abstract .unnumbered} The paper addresses one of nontrivial octonion related facts. According to paper gr-qc/0409095, the most stable space-time state is the one described by real Dirac matrices in 11-dimensional space of signature $1\left( { -} \right)\& 10\left( { +} \right)$. The internal subspace is 7-dimensional, and its stability is due to a high “zero” energy packing density when using an oblique-angled basis from fundamental vectors of lattice $E_{8} $ for the spinor degrees of freedom. The nontrivial fact consists in the following: Dirac symbols with octonion matrix elements can be used to describe states of the space of internal degrees of freedom if and only if the space corresponds either to stable vacuum states or states close to the just mentioned ones. The coincidence of the internal space dimension and signature for absolutely different and independent approaches to the consideration of this issue seems to predetermine the internal space vacuum properties and the apparatus, which is able to constitute the basis of the unified interaction theory. 1. Introduction {#introduction .unnumbered} =============== 4-dimensional Riemannian space with metric tensor $g_{\alpha \beta} $ is considered. The Greek letters take on values $\alpha ,\beta ,... = 0,1,2,3$. The metric tensor is assumed nonsingular, so Dirac symbols (DS) $\gamma _{\alpha} $ according to $$\label{eq1} \gamma _{\alpha} \gamma _{\beta} + \gamma _{\beta} \gamma _{\alpha} = 2g_{\alpha \beta}$$ can be introduced at every point. This paper considers algebraic properties of DS at some spatial point. The coordinate system is assumed to be locally Cartesian and tensor $g_{\alpha \beta} $ equal to $$\label{eq2} g_{\alpha \beta} = diag\left( { - 1,1,1,1} \right).$$ In the DS theory there are a number of problems, the solution to which directly affects the physical interpretation of DS involving constructions, however, such that there is no complete understanding in regard to their solution method. Mention two of them. It is well known that symbols $\left\{ {\gamma _{\alpha} } \right\}$ can be realized as Dirac matrices (DM) above any number field (real, complex, quaternion numbers) as well as above the octonion body. The realization in the form of square matrices $4 \times 4$ is meant. On the other hand, DS can be realized in the form of real square matrices $N \times N$, $N \ge 4$. Of interest is the question: What is the relation between these two realization types? In particular, what are the characteristics of the subspaces of internal degrees of freedom that are introduced additionally in each complication of the number body used? For physics, the octonion realization is of a special interest, in particular, for the reason that using any number body except for the octonion one does not allow us even to pose the question of explanation of the irreversibility of actual processes on the basis of time-reversible fundamental laws. The irreversibility phenomenon may be explained only in transition to the formulation of the physical laws in terms of octonions. But in this most interesting case some of the theorems do not hold, on the basis of which the polarization density matrix is introduced and conclusions on the correspondence between tensors and bispinors are reached. The problem is to give an answer to the question: To what extent are those results for the correspondence between tensors and bispinors, which have been found for real matrix realizations of DM, valid for the octonion DM? This paper makes an attempt to give answers to the two above-formulated problems. 2. DM realization above a real field in Riemannian spaces of a dimension higher than four {#dm-realization-above-a-real-field-in-riemannian-spaces-of-a-dimension-higher-than-four .unnumbered} ========================================================================================= The Riemannian spaces of dimension $n \ge 4$ have been studied in connection with construction of matrix spaces (MS), that is the Riemannian spaces, in which the internal degrees of freedom properties are introduced through Dirac matrices $\gamma _{A} $. Subscripts $A,B,...$ take on values $A,B = 1,2,...,n$, while the $\gamma _{A} $’s themselves are realized as square matrices $N \times N$ and satisfy relations $$\label{eq3} \gamma _{A} \gamma _{B} + \gamma _{B} \gamma _{A} = 2g_{AB} \cdot E$$ Here$E$ is the unit matrix in the space of internal degrees of freedom. The internal degrees of freedom are related, first, with transformations $$\label{eq4} \gamma _{A} \to {\gamma} '_{A} = S\left( {x} \right)\gamma _{A} S^{ - 1}\left( {x} \right),$$ and, second, with the transition to Riemannian spaces of larger dimensions and different signatures. The MS theory in multidimensional Riemannian spaces with real realizations of DM is discussed in detail in refs. \[1\]-\[3\]. These papers also prove the following: The realization of DM above a real field frequently entails the notion of so-called maximum MS, in which the set of the quantities, the generatrices for which are DM, coincides with the set of all matrices of a given dimension. The maximum MS have odd dimension $n$. $$n = 2k + 1,$$ where $k$ is a positive integer. Their signature is therewith of form\ $\left( {k + 1} \right)\left( { +} \right)\& k\left( { -} \right)$ or differs from that by a number of “minuses”, which is a multiple of four. The DM, which can be introduced in the Riemannian space possessing the above properties are square matrices $N \times N$, with $N$ relating to $k$ as $$N = 2^{k}.$$ In any MS, either anti-Hermitizing matrix $D$ or Hermitizing matrix $C$ can be introduced. The $D$ or $C$ are determined as $$D\gamma _{a} D^{ - 1} = - \gamma _{a}^{ +} ;\quad C\gamma _{a} C^{ - 1} = \gamma _{a}^{ +} ;\quad a = 0,1,2,...,N - 1.$$ The matrix $D$ or matrix $C$ can be used to introduce a Hermitean matrix set, whose existence, in its turn, is needed to introduce the concept of polarization density matrix. 3. Complex numbers and quaternions {#complex-numbers-and-quaternions .unnumbered} ================================== The results relating to determination of properties of those multidimensional Riemannian spaces, in which the real DM algebra is mapped isomorphously to the DM algebra in 4-dimensional space of signature $\left( { - + + +} \right)$ in realization of the latter above the real, complex and quaternionic fields are summarized in Table 1. Table 1. Parameters characterizing the isomorphism between DM realized above different number fields and DM realized above the real number field = [|p[203pt]{}|p[114pt]{}|]{} a & a\ **With the help of real matrices**& **With the help of matrices** $4 \times 4$**, but using different number bodies**\ Matrices $4 \times 4$. 16 parameters.& Real number field\ Subset of matrices $\left( {4 \times 4} \right) \otimes \left( {2 \times 2} \right)$. 32 parameters. In this realization method, an additional internal subspace of dimension 1 is actually introduced. Gauge group $U\left( {1} \right)$& Complex field. The transition to the matrix notation is performed using isomorphism $${\rm I} \Leftrightarrow E_{4 \times 4} \otimes i\sigma _{2}$$\ Subset of matrices $\left( {4 \times 4} \right) \otimes \left( {4 \times 4} \right)$. 64 parameters. In this realization method, an additional internal subspace of dimension 3 is actually introduced. Gauge group $SU\left( {2} \right)$& Quaternion field. The transition to the matrix notation is performed using isomorphism $$\left. {\begin{array}{l} {{\rm I}_{1} \Leftrightarrow E_{4 \times 4} \otimes i\rho _{2} \sigma _{1}} \\ {{\rm I}_{2} \Leftrightarrow E_{4 \times 4} \otimes i\sigma _{2}} \\ {{\rm I}_{3} \Leftrightarrow E_{4 \times 4} \otimes i\rho _{2} \sigma _{3}} \\ \end{array}} \right\}$$\ As it follows from Table 1, the dimension of the internal space that appears in the transition from one number field to another is the same as the number of imaginary units in the number field. The Riemannian spaces are therewith subspaces of maximum MS. 4. Octonion DS {#octonion-ds .unnumbered} ============== Irrespective of the fact that octonions are discussed extensively in the literature (see, e.g., \[4\], \[5\]), nevertheless, here we present some information about these unusual numbers. Naturally, we will do this briefly and only to the extent, which is needed for the consistency of the discussion. The algebra of octonion imaginary units $\left\{ {e_{N}} \right\}$ is determined as $$\label{eq5} e_{M} e_{N} = - \delta _{MN} e_{0} + C_{MNK} e_{K}$$ Here: $M,N,K = 1,2,...,7$; $C_{MNK} $ are quantities completely antisymmetric in their indices; nonzero components are: $$\label{eq6} C_{123} = C_{145} = C_{246} = C_{347} = C_{176} = C_{257} = C_{365} = 1.$$ Quantity $\Delta \left[ {A,B,C} \right]$is called the associator of three octonions $A,B,C$: $$\label{eq7} \Delta \left[ {A,B,C} \right] = \frac{{1}}{{2}}\left\{ {\left( {AB} \right)C - A\left( {BC} \right)} \right\}.$$ The whole specificity of the octonion algebra against the matrix algebra is that the associators (\[eq7\]) are nonzero. Perform the linear real transformation of symbols $e_{M} $ of the following form: $$\label{eq8} \left. {\begin{array}{l} {e_{M} \to {e}'_{M} = G_{MN} \cdot e_{N}} \\ {e_{0} \to {e}'_{0} = e_{0}} \\ \end{array}} \right\}.$$ Consider properties of tensor $G_{MN} $ in the 7-dimensional Euclidean space, in which the base vectors are symbols $e_{M} $. The substitution of ${e}'_{M} $ into $$\label{eq9} {e}'_{M} {e}'_{N} = - \delta _{MN} e_{0} + C_{MNK} {e}'_{K} ,$$ which symbols ${e}'_{M} $ should satisfy, leads to the following two relations: $$\label{eq10} \left. {\begin{array}{l} {G_{MK} G_{NK} = \delta _{MN}} \\ {G_{MA} G_{NB} C_{ABC} = C_{MNS} G_{SC}} \\ \end{array}} \right\}.$$ Quantities $G_{MN} $ produce 14-parametric group $G_{2} $ of rank 2. According to the universally adopted classification, group $G_{2} $ is attributed to the exceptional Lie group category. Detailed information about the group $G_{2} $ can be found, e.g., in ref. \[6\]. It is known in advance that in the case of DS realization above the octonion body any isomorphous mapping of the appearing DS apparatus to the matrix apparatus cannot exist in principle. So the question is quite appropriate: Do the matrix realizations of DS have any bearing on the octonion DS whatsoever? To answer this question, make it our aim to construct the DS realization in the form of real DM in a multidimensional Riemannian space, which would satisfy the following requirement: $ \bullet $ When in algebraic operations with octonion DS $C_{ABC} $ play actually no role, the algebraic operations should map to the algebra of real DM of an appropriate dimension. This is true for the algebraic operations with DS $\left\{ {\gamma _{\alpha} } \right\}$ near real DM $\left\{ {\bar {\gamma} _{\alpha} } \right\}$. To meet this requirement, suppose that in the scheme under discussion there is the smallness parameter $0 < \lambda < < 1$, such that all matrix elements $\left( {\gamma _{\alpha} - \bar {\gamma} _{\alpha} } \right)$ modulo are of the order of $\lambda $. Write the matrices $\gamma _{\alpha} $ as $$\label{eq11} \gamma _{\alpha} = \bar {\gamma} _{\alpha} + f_{\alpha ;0} \cdot e_{0} + f_{\alpha ;N} \cdot e_{N} ;\quad \left( {N = 1,2,...,7} \right)$$ Matrices (\[eq11\]) will satisfy relation (\[eq1\]), if small matrices $\left\{ {f_{\alpha ;0} ,f_{\alpha ;N}} \right\}$ are of the form $$\label{eq12} f_{\alpha ;0} = \left[ {s_{0} ,\bar {\gamma} _{\alpha} } \right]_{ -} ;\quad f_{\alpha ;N} = \left[ {s_{N} ,\bar {\gamma} _{\alpha} } \right]_{ - } ,$$ where $\left\{ {s_{0} ,s_{N}} \right\}$are arbitrary small real matrices $4 \times 4$. Upon substitution of (\[eq12\]) into (\[eq11\]) it turns out that octonion DS are written in the form $$\label{eq13} \gamma _{\alpha} = \bar {\gamma} _{\alpha} + \left[ {s_{0} ,\bar {\gamma }_{\alpha} } \right]_{ -} \cdot e_{0} + \left[ {s_{N} ,\bar {\gamma }_{\alpha} } \right]_{ -} \cdot e_{N} .$$ The substitution of (\[eq13\]) into (\[eq1\]) shows that in the first smallness order the $C_{ABC} $ drop out and play no role. This means that the algebra with generatrices satisfying relation $$\label{eq14} \left[ {e_{M} ,\;e_{N}} \right]_{ +} = - 2\delta _{MN} e_{0} ,$$ $$\label{eq15} \left[ {e_{M} ,\;e_{N}} \right]_{ -} = 2C_{MNK} e_{K} ,$$ can be mapped in the first order of smallness to the algebra of real DM, in which instead of seven imaginary units $\left\{ {e_{N}} \right\}$, seven matrix imaginary units $\left\{ {{\rm I}_{N}} \right\}$ are used. The specific form of the real DM satisfying either above-formulated requirement can be as follows: $$\label{eq16} \left. {\begin{array}{l} {e_{1} \Leftrightarrow {\rm I}_{1} = E_{4 \times 4} \otimes i\rho _{2} \sigma _{1} \otimes \sigma _{1} \quad e_{2} \Leftrightarrow {\rm I}_{2} = E_{4 \times 4} \otimes i\sigma _{2} \otimes \sigma _{1} \quad}\\ e_{3}\Leftrightarrow {\rm I}_{3} = E_{4 \times 4} \otimes i\rho _{2} \sigma _{3} \otimes \sigma _{1} \quad \quad {e_{4} \Leftrightarrow {\rm I}_{4} = E_{4 \times 4} \otimes i\rho _{1} \sigma _{2} \otimes \sigma _{3} \quad}\\ e_{5} \Leftrightarrow {\rm I}_{5} = E_{4 \times 4} \otimes i\rho _{2} \otimes \sigma _{3} \quad e_{6} \Leftrightarrow {\rm I}_{6} = E_{4 \times 4} \otimes i\rho _{3} \sigma _{2} \otimes \sigma _{3} \\ {e_{7} \Leftrightarrow {\rm I}_{7} = E_{4 \times 4} \otimes E_{4 \times 4} \otimes i\sigma _{2} .} \\ \end{array}} \right\}$$ The resultant multidimensional Riemannian space has dimension 11 and signature $1\left( { -} \right)\& 10\left( { +} \right)$. The DM in the space is written as: $$\label{eq17} \left. {\begin{array}{l} {\bar {\gamma} _{0} = - i\rho _{2} \sigma _{1} \otimes E_{4 \times 4} \otimes E_{2 \times 2} ;\quad} {\bar {\gamma} _{1} = \rho _{1} \otimes E_{4 \times 4} \otimes E_{2 \times 2}}\\ {\bar {\gamma} _{2} = \rho _{2} \sigma _{2} \otimes E_{4 \times 4} \otimes E_{2 \times 2} ;\quad \bar {\gamma} _{3} = \rho _{3} \otimes E_{4 \times 4} \otimes E_{2 \times 2} ;} \\ {\bar {\gamma} _{4} = i\rho _{2} \sigma _{3} \otimes i\rho _{2} \sigma _{1} \otimes \sigma _{1} ;\quad \bar {\gamma} _{5} = i\rho _{2} \sigma _{3} \otimes i\sigma _{2} \otimes \sigma _{1} }\\ { \bar {\gamma} _{6} = i\rho _{2} \sigma _{3} \otimes i\rho _{2} \sigma _{3} \otimes \sigma _{1} \quad {\bar {\gamma} _{7} = i\rho _{2} \sigma _{3} \otimes i\rho _{1} \sigma _{2} \otimes \sigma_{3}}}\\ { \bar {\gamma} _{8} = i\rho _{2} \sigma _{3} \otimes i\rho _{2} \otimes \sigma _{3} ;\quad \bar {\gamma} _{9} = i\rho _{2} \sigma _{3} \otimes i\rho _{3} \sigma _{2} \otimes \sigma _{3} ;} \\ {\bar {\gamma} _{10} = i\rho _{2} \sigma _{3} \otimes E_{4 \times 4} \otimes i\sigma _{2} .} \\ \end{array}} \right\}$$ Expressions (\[eq13\]) have the meaning of the ones in the first order of smallness for the first four DM among eleven DM. The expressions for all the eleven DM in the first order of smallness are derived from $$\label{eq18} \gamma _{\alpha} = \bar {\gamma} _{\alpha} + \left[ {s,\bar {\gamma }_{\alpha} } \right]_{ -} ;\quad \gamma _{N + 3} = \bar {\gamma} _{N + 3} + \left[ {s,\bar {\gamma} _{N + 3}} \right]_{ -} ;\quad N = 1,2,...,7,$$ where $$\label{eq19} s = s_{0} \cdot E + s_{N} \cdot {\rm I}_{N} .$$ Thus, in the linear approximation the octonion DM $\gamma _{\alpha} $ can be treated as ordinary matrices, if for the basic matrices, in the vicinity of which the expansion proceeds, real DM are used in 11-dimensional Riemannian space of signature $1\left( { -} \right)\& 10\left( { +} \right)$. Pay attention to the fact that except for the reality no other properties of DM in 11-dimensional Riemannian space have been used in this consideration. This means that instead of system (\[eq17\]) that DM system can be used in the consideration, which has been derived in \[7\] from system (\[eq17\]) through transition to the oblique-angled basis produced by simple root vectors of Lie algebra $E_{8} $. In the general case the following rule remains valid: If it was possible to realize DS with the help of octonion DM $4 \times 4$, then after that one can transfer from one realization to another using transformations $G_{2} $. 5. Discussion {#discussion .unnumbered} ============= Although the octonion DS can be written in the form of matrices $4 \times 4$ in the general case, but the algebra of the matrices possesses no associativity and, hence, cannot be mapped to the algebra of ordinary real matrices in a multidimensional space. In the linear approximation, however, the algebra of octonion DM is mapped to that of real DM in 11-dimensional Riemannian space of signature $1\left( { -} \right)\& 10\left( { +} \right)$. One of possible DM systems in this space is of form (\[eq17\]). The result obtained is of interest for several reasons. is that the correspondence found by us between octonion DS and real DM in a multidimensional Riemannian space leads to the Riemannian space, in which the most stable vacuum state appears. Ref. \[7\] shows that the most stable vacuum state both among the internal subspaces of dimensions other than 7 and among DM of different spinor basis structure is the DM realization in the form of real matrices, in which the oblique-angled basis from the set of fundamental vectors of lattice $E_{8} $ is used. In this realization, the internal space dimension is 7; the specific form of the lattice DM is presented in ref. \[7\] and the matrix of transition from the orthonormal basis to the lattice one is given, e.g., in \[7\], \[8\]. is that using any number body, except for the octonion body, in physical theories does not allow us even to pose the question of explanation of the irreversibility of actual processes on the basis of time-reversible fundamental laws with writing the latter in terms of any number field. The irreversibility phenomenon may be explained only in transition to the formulation of the physical laws in terms of octonions. In this connection note that it has been long since the physicists have paid attention to the existence of an evident contradiction: on the one hand, the dynamic equations describing fundamental interactions possess time reversibility; on the other hand, actual processes that occur in the Nature are irreversible. R. Penrose in \[9\] writes: “...It is hard to understand how our immense Universe could “sink” into one or another of the states with being unable to even imagine in what time direction to start! ...the only explanation ... remains: not all accurate physics laws are symmetric in time!...”. If DM are realized above the octonion body, then the transition amplitudes automatically cease to be associative. While this just means that the reversibility in time does disappear at the level of fundamental processes in the microworld. In fact, if $A_{1} ,A_{2} ,A_{3} ,...$ are amplitudes of the transitions from initial state $t_{0} $ to states arising at times $t_{1} ,t_{2} ,t_{3,...} $, then the amplitude for one of the paths of the process proceeding in the time-forward direction should be found according to rule $$\label{eq20} \left( {\left( {A_{1} \cdot A_{2}} \right) \cdot A_{3}} \right)...,$$ while the conjugate amplitude for the process running in the time-backward direction should be found according to rule $$\label{eq21} \left( {A_{1} \cdot \left( {A_{2} \cdot A_{3} ...} \right)} \right).$$ At the level of real, complex and quaternionic numbers expressions (\[eq20\]), (\[eq21\]) lead to the same probabilities of transitions. But as soon as octonions come into use, the equality between expressions (\[eq20\]), (\[eq21\]), generally speaking, disappears. Moreover, the body of octonion numbers is the only one possessing this property. This means that we may necessarily resort to the octonion quantities for explanation of the irreversibility of processes. The above considerations and results justify the multiple attempts to consider the octonion wave functions for half-integer spin particles. We only point out to refs. \[5\], \[10\], \[11\] as typical papers from the standpoint of the method for consideration of octonion Dirac matrices. The method of these papers is valid only to the quadratic approximation, as in these papers there is either explicit or implicit transition to so-called split octonions (introduction of the outer imaginary unit commutating with all octonions) or the octonion composition rule is replaced by the open product. Similar (or equivalent) techniques restore the associativity of the modified number body and allow the standard matrix apparatus to be employed. However, in so doing a most interesting part of the octonion specificity is lost. A method for description of the half-integer spin particle dynamics is the method of mapping of tensors to bispinors developed in a number of papers (see, e.g., \[3\]). In the method, one of principal objects is bispinor matrix $Z$. For the octonion implementation of DS, the matrix $Z$ exists in the linear approximation and, as it follows from (\[eq13\]), coincides with $S^{ - 1}$. Through multiplication on the right by the projectors, states with different quantum numbers can be separated from the bispinor matrix. For example, one of the subgroups of group $G_{2} $ is $SU\left( {3} \right)$. In the general case there is no bispinor matrix, however, the results obtained using the methods for consideration of the transformations of DM $4 \times 4$, which are suggested in ref. \[12\], remain valid. Thus, the vacuum stability requirement can be made consistent with using the most general number body. In so doing any violation of the bounds of the 7-dimensional internal Euclidean space will result in vacuum instability (and appearance of tachyons as a consequence). The work was carried out under partial financial support by the International Science and Technology Center (ISTC Project \#1655). References {#references .unnumbered} ========== \[1\] M.V. Gorbatenko, A.V. Pushkin. *VANT; Ser.: Teor. i Prikl. Fiz*. **1(\[eq1\])**, 49 (1984). \[2\] M.V. Gorbatenko. *TMF*. **103**, **1**, 32 (1995). \[3\] M.V. Gorbatenko, A.V. Pushkin. *VANT; Ser.: Teor. i Prikl. Fiz*. **2-3**, 61 (2000). \[4\] B.A. Rozenfeld. *Non-Euclidean geometries.* Moscow. Gostekhizdat Publishers (1955). \[5\] D.F. Kurdgelaidze. *Introduction to nonassociative classical field theory*. Tbilisi. Metsnnegeva (1987). \[6\] R.E. Behrends, J. Dreitlein, C. Fronsdal, W. Lee. *Rev. Mod. Phys.,* **34**, No. 1, 1 (January 1962). \[7\] M.V. Gorbatenko, A.V. Pushkin. *Physical Vacuum Properties and Internal Space Dimension.* gr-qc/0409095. To be published in GRG. \[8\] J.H. Conway, N.J.A. Sloane. *Sphere Packings, Lattices and Groups.* Springer-Verlag. New York (1988). \[9\] R. Penrose. *Singularities and asymmetry in time.* In: “The General Relativity”. Moscow, Publishers (1983), p. 233. \[10\] Sirley Marques-Bonham. *The Dirac equation in a non-Riemannian manifold III: An analysis using the algebra of quaternions and octonions.* J. Math. Phys. **32** (\[eq5\]), 1383 (1991). \[11\] S. Margues, C.G. Oliveira. *An extension of quaternionic metrics to octonions.* J. Math. Phys. **26 (\[eq12\])**, 3131 (1985). \[12\] N.D. Sen Gupta. *On the Invariance Properties of the Dirac Equation.* Nuovo Cimento, Vol. XXXVI, N.4 (1965).
{ "pile_set_name": "ArXiv" }
--- abstract: 'This article is an introductory review of the physics of quantum spin liquid (QSL) states. Quantum magnetism is a rapidly evolving field, and recent developments reveal that the ground states and low-energy physics of frustrated spin systems may develop many exotic behaviors once we leave the regime of semi-classical approaches. The purpose of this article is to introduce these developments. The article begins by explaining how semi-classical approaches fail once quantum mechanics become important and then describes the alternative approaches for addressing the problem. We discuss mainly spin $1/2$ systems, and we spend most of our time in this article on one particular set of plausible spin liquid states in which spins are represented by fermions. These states are spin-singlet states and may be viewed as an extension of Fermi liquid states to Mott insulators, and they are usually classified in the category of so-called $SU(2)$, $U(1)$ or $Z_2$ spin liquid states. We review the basic theory regarding these states and the extensions of these states to include the effect of spin-orbit coupling and to higher spin ($S>1/2$) systems. Two other important approaches with strong influences on the understanding of spin liquid states are also introduced: (i) matrix product states and projected entangled pair states and (ii) the Kitaev honeycomb model. Experimental progress concerning spin liquid states in realistic materials, including anisotropic triangular lattice systems ($\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ and EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$), kagome lattice systems (ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$) and hyperkagome lattice systems (Na$_{4}$Ir$_{3}$O$_{8}$), is reviewed and compared against the corresponding theories.' author: - Yi Zhou - Kazushi Kanoda - 'Tai-Kai Ng' bibliography: - 'refall.bib' title: Quantum Spin Liquid States --- Introduction ============ Quantum spin liquid (QSL) states in dimensions of $d>1$ have been a long-sought dream in condensed matter physics. The general idea is that when acting on spin systems, quantum mechanics may lead to exotic ground states and low-energy behaviors that cannot be captured by traditional semi-classical approaches. The difficulty in implementing this idea is that we have no natural place to start once we have left the comfort zone of semi-classical approaches, at least in dimensions larger than one. Except for a few exactly solvable models, we must rely heavily on numerical or variational approaches to “guess" the correct ground state wavefunctions and on a combination of sophisticated numerical and analytical techniques to understand the corresponding low-energy excitations. Several excellent reviews are available on QSLs [@Lee08; @Balents10] and frustrated magnetism [@BookDiep; @BookLacroix]. This article complements those mentioned above by providing a pedagogical introduction to this subject and reviews the current status of the field. We explain, at an introductory level, why sophisticated approaches are needed to study QSL states, how these approaches are implemented in practice, and what new physics may be expected to appear. The experimental side of the story and the drawbacks or pitfalls of the theoretical approaches are also discussed. We concentrate mainly on spin $1/2$ systems and study in detail one particular set of plausible spin liquid states that are usually termed resonating valence bond (RVB) states. The spins are treated as fermions in these states, which may be viewed as an extension of Fermi liquid states to Mott insulators. They are usually classified in the category of $SU(2)$, $U(1)$ or $Z_2$ spin liquid states. Because of the intrinsic limitations of the fermionic RVB approach, many other approaches to spin liquid states have been developed by different authors. These approaches often lead to other exotic possibilities not covered by the simple fermionic approach. Two of these approaches are introduced in this article for completeness: (i) matrix product states and projected entangled pair states and (ii) the Kitaev honeycomb model. The article is organized as follows. In section II, we introduce the semi-classical approach to simple quantum antiferromagnets, and we explain the importance of the spin Berry phase and how one can include it in a semi-classical description to obtain the correct theory. In particular, we show how it leads to the celebrated Haldane conjecture. The existence of end excitations as a natural consequence of the low-energy effective theory of these systems is discussed. One-dimensional quantum spin systems are of great interest at present because they provide some of the simplest realizations of symmetry-protected topological (SPT) phases in strongly correlated systems. The limitations of the semi-classical approach when applied to systems with frustrated interactions are discussed in section III, where we introduce the alternative idea of constructing variational wavefunctions directly. We introduce Anderson’s famous idea of the RVB wavefunction for spin $1/2$ systems and discuss how this idea can be implemented in practice. The difficulty of incorporating the $SU(2)$ spin algebra in the usual many-body perturbation theory is noted, and the trick of representing spins by particles (fermions or bosons) with [*constraints*]{} to avoid this difficulty is introduced. The non-trivial $SU(2)$ gauge structure in the fermion representation of RVB states and the resulting rich structure of the low-energy effective field theories for these spin states ($SU(2)$, $U(1)$ and $Z_2$ spin liquids) are discussed. An interesting linkage of the $U(1)$ spin liquid state to the (metallic) Fermi liquid state through a Mott metal-insulator transition is introduced. The difficulty of finding controllable approaches for studying spin liquid states has led to an extension of the RVB approach and a search for alternative approaches. Some of these approaches are reviewed briefly in section IV, including (i) the extension of the RVB approach to include the effect of spin-orbit coupling and to higher spin ($S>1/2$) systems, (ii) the concepts of matrix product states and projected entangled pair states, and (iii) the Kitaev honeycomb model. The main message of this section is that a larger variety of exotic spin states become possible when we leave the paradigm of spin $1/2$ systems with rotational symmetry. The $U(1)$ and $Z_2$ spin liquid states belong to merely a very small subset of the plausible exotic states once we leave the paradigm of semi-classical approaches. Section V is devoted to a survey of experimental research on spin liquid states. Special attention is paid to the $U(1)$ spin liquid state, on which most experimental efforts have been focused. The best studied examples are a family of organic compounds denoted by $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ (ET) [@Kanoda03] and Pd(dmit)$_{2}$(EtMe$_{3}$Sb) (dmit salts) [@Itou08]. Both materials are Mott insulators near the metal-insulator transition and become superconducting (ET) or metallic (dmit) under modest pressure. Despite the large magnetic exchange $J\approx 250$ K observed in these systems, there is no experimental indication of long-range magnetic ordering down to a temperature of $\sim30$ mK. A linear temperature dependence of the specific heat and a Pauli-like spin susceptibility have been found in both materials at low temperature, suggesting that the low-energy excitations are spin-$1/2$ fermions with a Fermi surface [@SYamashita; @Matsuda12]. This Fermi-liquid-like behavior is further supported by their Wilson ratios, which are close to one. In addition to ET and dmit salts, the kagome compound ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$ [@Kagome07] and the three-dimensional hyperkagome material Na$_{4}$Ir$_{3}$O$_{8}$ [@Na4Ir3O8] are also considered to be candidates for QSLs with gapless excitations. Experimental surveys on these QSL candidate materials are presented in this article, including their thermodynamics, thermal transport and various spin spectra. We also briefly introduce the discoveries of a few new materials and discuss the existing discrepancies between experiments and theories. The paper is summarized in section VI. From semi-classical to non-linear-$\sigma$ model approaches for quantum antiferromagnets ======================================================================================== Here, we consider simple Heisenberg antiferromagnets on bipartite lattices (with sublattices $A$ and $B$) with the Hamiltonian $$\label{hm1} H=J\sum_{\langle i,j\rangle}\mathbf{S}_i\cdot\mathbf{S}_j,$$ where $J>0$ and $\langle i,j\rangle$ describes a pair of nearest neighbor sites in the bipartite lattice. In a bipartite lattice, any two nearest neighbor sites always belong to different sublattices. $\mathbf{S}$ is a quantum spin with magnitude $S=n/2$, where $n=$ positive integer. Examples of bipartite lattices include 1D spin chains, 2D square or honeycomb lattices, and 3D cubic lattices. Two-spin problem ---------------- The semi-classical approach begins with the assumption that the quantum spins are “close" to classical spins, and it is helpful to start by first analyzing the corresponding classical spin problem. For simplicity, we start by considering only two classical spins coupled by the Heisenberg interaction $$H=J\mathbf{S}_A\cdot\mathbf{S}_B. \\\\\ (J>0).$$ The classical spins obey Euler’s equation of motion: $$\label{eqm1} {\partial\mathbf{S}_{A(B)}\over\partial t}=J\mathbf{S}_{B(A)}\times\mathbf{S}_{A(B)}.$$ This equation can be solved most easily by introducing the magnetization and staggered magnetism vectors $\mathbf{M}(\mathbf{N})=\mathbf{S}_A+(-)\mathbf{S}_B$, where it is easy to show from Eq.  (\[eqm1\]) that $$\begin{aligned} \label{eqm3} {\partial \mathbf{M}\over\partial t} & = & 0, \\ \nonumber {\partial\mathbf{N}\over\partial t} & = & J\mathbf{M}\times\mathbf{N}, \end{aligned}$$ indicating that classically, the staggered magnetization vector $\mathbf{N}$ rotates around the (constant) total magnetization vector $\mathbf{M}$. Let $\mathbf{S}_{A(B)}=S_{A(B)}\hat{r}_{A(B)}$, where $S_{A(B)}$ are the magnitudes of the spins $\mathbf{S}_{A(B)}$ and $\hat{r}_{A(B)}$ are unit vectors indicating the directions of $\mathbf{S}_{A(B)}$; then, the classical ground state has $\hat{r}_A=-\hat{r}_B$ with $\mathbf{M}=0$, i.e., the two spins are antiferromagnetically aligned. Note that the equation of motion given in Eq.  implies that ${\partial(\mathbf{N}^2)\over\partial t}=0$, i.e., the magnitude of $\mathbf{N}$ remains unchanged during its motion. Therefore, if we write $\mathbf{N}=N\hat{n}$, where $N$ is the magnitude of $\mathbf{N}$ and $\hat{n}$ is the unit vector denoting its direction, we find that only $\hat{n}$ changes under the equation of motion given in Eq. . The effects of quantum mechanics can be seen most easily by observing that the equation of motion given in Eq.  describes the dynamics of a free rotor (a rigid rod with one end fixed such that the rod can rotate freely around the fixed end). A free rotor can be represented by a vector $\mathbf{r}=r_0\hat{r}$, where $r_0=$ constant is the length of the rod and $\hat{r}$ is the unit radial vector describing the orientation of the rod. The rod has an angular momentum of $$\label{rotor1} \mathbf{L}=\mathbf{r}\times\mathbf{p}=r_0\hat{r}\times\mathbf{p},$$ where $\mathbf{p}=mr_0\dot{\hat{r}}$ is the momentum and $m$ is the mass. Using Eq. , we obtain \[rotor2\] $$\begin{aligned} \label{rotor2a} \hat{r}\times\mathbf{L}=-r_0\mathbf{p}=-mr_0^2\dot{\hat{r}}. \end{aligned}$$ We also have $$\label{rotor2b} \dot{\mathbf{L}}=0$$ (conservation of angular momentum). Comparing Eqs. and , we find that the equation of motion for two spins is equivalent to the equation of motion for a free rotor if we identify $\mathbf{L}\rightarrow\mathbf{M}$, $\hat{r}\rightarrow\hat{n}$ and $J=I^{-1}$, where $I=mr_0^2$ is the moment of inertia of the rotor. The quantum Hamiltonian of the free rotor is $$H_{rotor}={1\over2 I}\mathbf{L}^2,$$ and its solution is well known. The eigenstates are the spherical harmonics $Y_{lm}(\theta,\phi)$ (where $\theta$ and $\phi$ specify the direction of the unit vector $\hat{r}$) with eigenvalues $$\mathbf{L}^2=l(l+1)\hbar^2,\\\\\\\\\\\ L_z=m\hbar,$$ and corresponding energies $E_l=l(l+1)\hbar^2/2I$, where $l$ and $m$ are integers such that $l\geq0$ and $l\geq|m|$. In particular, $\mathbf{L}(\mathbf{M})=0$ for the ground state of the quantum rotor, but the direction of the vector $\mathbf{r}(\mathbf{N})$ is completely uncertain ($Y_{00}(\theta,\phi)={1\over\sqrt{4\pi}}$) as a result of quantum fluctuations, indicating a breakdown of the classical solution, in which $\mathbf{n}$ is fixed in the ground state. (Alternatively, one can gain this understanding from the Heisenberg uncertainty principle, $\langle\delta\hat{r}\rangle\langle\delta\mathbf{L}\rangle>\hbar$. With $\mathbf{L}=0$ in the ground state, $\delta\mathbf{L}\equiv0$ and $\delta\hat{r}\rightarrow\infty$, the direction of the vector $\hat{r}$ becomes completely uncertain.) A moment of thought indicates that our mapping of the spin problem to the rotor problem cannot be totally correct. What happens if $\mathbf{S}_A$ is an integer spin and $\mathbf{S}_B$ is a half-odd-integer spin? Elementary quantum mechanics tells us that the ground state should carry half-odd-integer angular momentum. The possibility of such a scenario is missing in our rotor mapping, in which the spin magnitudes $S_{A(B)}$ do not appear. Berry’s phase ------------- The missing piece in our mapping of the two-spin problem to the rotor model is the Berry’s phase [@Berry84], which is carried by spins but is absent in rotors. The correct spin-quantization rule is recovered only after this piece of physics is properly added into the rotor problem. First, let us review the Berry’s phase carried by a single spin. We recall that for a spin tracing out a closed path $\mathbf{C}$ on the surface of the unit sphere, the spin wavefunction acquires a Berry’s phase $\gamma(\mathbf{C})=S\Omega(\mathbf{C})$, where $S$ is the spin magnitude and $\Omega(\mathbf{C})$ is the surface area under the closed path $\mathbf{C}$ on the unit sphere (see Fig. \[fig:Berry\]). $S\Omega(\mathbf{C})$ can be represented more conveniently by imagining the spin trajectory as the trajectory of a particle carrying a unit charge moving on the surface of the unit sphere. In this case, the Berry’s phase is simply the phase acquired by the charged particle if a magnetic monopole of strength $S$ (i.e., $\mathbf{B}(\mathbf{r})=(S/r^2)\hat{r}$) is placed at the center of the sphere. The Berry’s phase acquired is the magnetic flux enclosed by the closed path $\mathbf{C}$. ![ Berry’s phase with a magnetic monopole.[]{data-label="fig:Berry"}](Berry.eps){width="4.0cm"} Let $S\mathbf{A}_M(\mathbf{r})$ be the vector potential associated with the monopole, i.e., $\nabla\times\mathbf{A}_M=\hat{r}/r^2$; then, in the “charge + gauge field" representation, the effect of the Berry’s phase can be described by a vector-potential term in the action: $$\label{WZ1} S_{B}=\hbar S\Omega(C)=\hbar S\int dt\mathbf{A}_M(\hat{r})\cdot\dot{\hat{r}}.$$ This is an example of a Wess-Zumino term for quantum particles. A more rigorous derivation of the Wess-Zumino action is given in Appendix A, where the action for a single spin in a magnetic field is derived [*via*]{} a path integral approach. We now revisit the two-spin problem. With the Berry’s phases included, the Lagrangian of the corresponding rotor problem becomes $$\label{La2} L={1\over2J}(\hat{n}\times\dot{\hat{n}})^2+\hbar S_A\mathbf{A}_{M}(\hat{r}_A)\cdot\dot{\hat{r}}_A+\hbar S_B\mathbf{A}_{M}(\hat{r}_B)\cdot\dot{\hat{r}}_B,$$ where $\mathbf{N}=N\hat{n}=\mathbf{S}_A-\mathbf{S}_B$. To simplify the problem, we adopt the semi-classical approximation $\hat{r}_A=-\hat{r}_B$ in the Berry’s phase terms, which is a reasonable approximation for states close to the classical ground state. With this approximation, we obtain $$\label{La2a} L\rightarrow{1\over2J}(\hat{n}\times\dot{\hat{n}})^2+\hbar\Delta S\mathbf{A}_M(\hat{n})\cdot\dot{\hat{n}},$$ where $\hat{n}=\hat{r}_A$ and $\Delta S=S_A-S_B$. The Hamiltonian of the system is $$\label{h2} H_M={J\over2}\left(\mathbf{\Pi}-\hbar\Delta S\mathbf{A}_M(\hat{n})\right)^2,$$ where $\mathbf{\Pi}=\dot{\hat{n}}/J$ is the canonical momentum of the rotor. $H_M$ is the Hamiltonian of a charged particle moving on the surface of a unit sphere with a magnetic monopole of strength $|\Delta S|$ located at the center of the sphere. The eigenstates of the Hamiltonian are well known and are called the monopole spherical harmonics [@Wu76]. The most interesting feature of the monopole spherical harmonics is that they allow half-odd-integer angular momentum states (which occur when $|\Delta S|$ is a half-odd-integer). The ground state carries an angular momentum of $L=|\Delta S|$ and is $(2|\Delta S|+1)$-fold degenerate, corresponding to the degeneracy of a quantum spin of magnitude $|\Delta S|$, in agreement with the exact result for the two-spin problem. Non-linear-$\sigma$-model ------------------------- The two-spin problem tells us that there are two important elements that we must keep track of when a classical spin problem is replaced with the corresponding quantum spin problem: a) quantum fluctuations, originating from the (non)-commutation relation between the canonical coordinates ($\mathbf{N}$) and momenta ($\mathbf{M}$), and b) Berry’s phase, which dictates the quantization of the spins. In the following, we generalize the rotor approach to the many-spin systems described by the antiferromagnetic (AFM) Heisenberg model, keeping in mind the above two elements. Following Haldane [@Haldane83PLA; @Haldane83PRL], we here consider Heisenberg antiferromagnets on a bipartite lattice described by the Hamiltonian given in Eq. . As in the two-spin problem, we introduce the magnetization vectors $\mathbf{M}(\mathbf{x}_i)$ and the staggered magnetization vectors $\mathbf{N}(\mathbf{x}_i)$ such that $$\begin{aligned} \label{sab} \mathbf{S}^A_i & = & \mathbf{M}(\mathbf{x}_i)+\mathbf{N}(\mathbf{x}_i), \\ \nonumber \mathbf{S}^B_i & = & \mathbf{M}(\mathbf{x}_i)-\mathbf{N}(\mathbf{x}_i), \end{aligned}$$ where $\mathbf{S}^{A(B)}$ denote spins on the $A(B)$ sublattices of the bipartite lattice. We assume that the ground state of the quantum system is “classical-like" with nearly anti-parallel spins on two nearest neighboring sites such that $\mathbf{M}(\mathbf{x}_i)\ll \mathbf{N}(\mathbf{x}_i)$, where both $\mathbf{M}(\mathbf{x})$ and $\mathbf{N}(\mathbf{x})$ are very slowly varying functions in space. (We show that this assumption can be justified in the following section.) The classical equation of motion for the spin at lattice site $i$ is $$\label{eqma} {\partial\mathbf{S}^{A(B)}_i\over\partial t}=J\left(\sum_{j=\text{NN}(i)}\mathbf{S}^{B(A)}_j\right)\times\mathbf{S}^{A(B)}_i,$$ where $j=\text{NN}(i)$ means that $j$ represents the nearest neighbor sites of $i$. Using Eq. , after some straightforward algebra and taking the continuum limit, we obtain $$\begin{aligned} \label{eqmb} {\partial\mathbf{N}(\mathbf{x})\over\partial t} & \sim & Jz\mathbf{M}(\mathbf{x})\times\mathbf{N}(\mathbf{x}), \\ \nonumber {\partial\mathbf{M}(\mathbf{x})\over\partial t} & \sim & -{Ja^2\over2}(\nabla^2\mathbf{N}(\mathbf{x}))\times\mathbf{N}(\mathbf{x}), \end{aligned}$$ where $a$ is the lattice spacing and $z=2d$ is the coordination number. We have assumed a square (cubic)-type lattice and have adopted the slowly varying approximation $$\mathbf{M}(x_{i+1})+\mathbf{M}(x_{i-1})\sim2\mathbf{M}(x_i),$$ $$\mathbf{N}(x_{i+1}))+\mathbf{N}(x_{i-1})\sim2\mathbf{N}(x_i)+a^2\partial^2_x\mathbf{N}(x_i),$$ etc. in deriving the above result. We have also assumed $\mathbf{M}(\mathbf{x})$ to be small and have neglected all non-linear terms in $\mathbf{M}(\mathbf{x})$ in deriving Eq. . To proceed further, we consider the situation in which all spins have the same magnitude $S$. Then, it is easy to see from Eq.  that $\mathbf{N}(\mathbf{x})^2+\mathbf{M}(\mathbf{x})^2=S^2$ and $\mathbf{N}(\mathbf{x})\cdot\mathbf{M}(\mathbf{x})=0$. Assuming that $M=|\mathbf{M}(\mathbf{x})|\ll N=|\mathbf{N}(\mathbf{x})|\sim S$, we find from Eq.  that $M\sim\omega/(zJ)$ and $\omega\sim\sqrt{z}JaS|\mathbf{k}|$, where $\omega$ and $\mathbf{k}$ are the frequency and wavevector, respectively, of the fluctuations in $\mathbf{N}$. In particular, $M\ll N$ when $ak\ll \sqrt{z}$, i.e., when $\mathbf{N}(\mathbf{x})$ is slowly varying in space. In the following, we adopt the approximation $N\sim S$ and write $\mathbf{N}(\mathbf{x})=S\hat{n}(\mathbf{x})$, where $\hat{n}^2=1$. Eliminating $\mathbf{M}(\mathbf{x})$ from Eq. , we obtain $$\label{nlsm1} {\partial^2\hat{n}(\mathbf{x},t)\over\partial t^2}={z(SJa)^2\over2}\nabla^2\hat{n}(\mathbf{x},t),$$ corresponding to the following classical action for the vector field $\hat{n}$: $$\label{nlsm2} S_{\sigma}={1\over2}\int dt\int d^dx \left({1\over J}\left({\partial\hat{n}\over\partial t}\right)^2-{zJ(Sa)^2\over2}(\nabla\hat{n})^2\right),$$ with the constraint $\hat{n}^2=1$. $S_{\sigma}$ is the non-linear-$\sigma$ model (NL$\sigma$M) for the unit vector field $\hat{n}(\mathbf{x})$. Comparing Eqs. and , we see that the NL$\sigma$M can be viewed as a continuum model describing coupled rotors $\hat{n}(\mathbf{x})$. The first term in the action gives the kinetic energy for the rotors, which we have discussed in detail for the two-spin model. The second term represents the coupling between nearest neighboring rotors in the lattice spin model. We note that the term for the coupling between rotors has a magnitude of $\sim S^2$ and dominates over the kinetic energy in the limit of large $S$. A more systematic derivation of the NL$\sigma$M starting from Eq.  can be achieved by writing $$\mathbf{S}_i=\eta_i S \hat{n}(x_i)\sqrt{1-\left|\frac{\mathbf{M}(x_i)}{S}\right|^2}+\mathbf{M}(x_i),$$ where $\eta_i=e^{i\pi x}$ and we still have $\mathbf{N}(\mathbf{x})\cdot\mathbf{M}(\mathbf{x})=0$. Assuming that $\mathbf{M}(\mathbf{x})$ is small, we can integrate out $\mathbf{M}(\mathbf{x})$ in a power series expansion of $\mathbf{M}(\mathbf{x})$ in the path integral. The NL$\sigma$M for $\hat{n}(\mathbf{x})$ is thus obtained to the leading (Gaussian) order [@Auerbach94]. ### Topological term We next consider the Berry’s phase contribution to the action. Following Appendix A, the total Berry’s phase contribution is $$\label{WZ2} S_{T}=\sum_iS_B(\hat{r}_i)\sim \hbar S\sum_i(-1)^i\Omega({\hat{n}_i}),$$ where $S\Omega(\hat{r}_i)=S\int dt\mathbf{A}_M(\hat{r}_i)\cdot \dot{\hat{r}}_i$ is the Berry’s phase for a single spin and $(-1)^i=1(-1)$ for sites on even (odd) sublattices. In the last step, we have assumed that the spins are almost anti-parallel. In the continuum limit, we obtain $$\label{WZ3} S_{T}\sim {\hbar S\over2^d}\int d^dx\left({\partial\over\partial_x^1}\cdots{\partial\over\partial_x^d}\right) \Omega(\hat{n}(\mathbf{x})).$$ $S_T$ is sensitive to the boundary conditions (see the discussion below), and we assume closed (periodic) boundary conditions in the following. The case of open boundary conditions is discussed afterward. For periodic boundary conditions, it is easy to see that $S_T$ is zero unless the integrand has a non-trivial topological structure. To evaluate $\partial_x\Omega$, we recall that $\Omega(\hat{n})$ measures the area on the surface of the sphere bounded by the trajectory $\hat{n}(t)$. Thus, the variation $\delta \Omega(\hat{n})$ due to a small variation in the trajectory $\delta\hat{n}$ is simply $$\delta\Omega(\hat{n})=\int dt \delta\hat{n}\cdot(\hat{n}\times\partial_t\hat{n}),$$ and $$\label{WZ4} S_{T}={\hbar S\over2^d}\int d^dx\int dt\left[\left({\partial\over\partial_x^1}\cdots{\partial\over\partial_x^d}\right)\hat{n}\right]\cdot(\hat{n}\times\partial_t\hat{n}).$$ The total effective action describing the quantum antiferromagnet is $S=S_{\sigma}+S_T$. The topological term is nonzero in one dimension and is usually written in the form $$\label{topo1} {S_{T}\over\hbar}={\theta\over8\pi}\sum_{\mu,\nu=0,1}\int d^2x\varepsilon_{\mu\nu}\hat{n}\cdot(\partial_{\mu}\hat{n}\times\partial_{\nu}\hat{n}),$$ where $x_0=t$, $x_1=x$, $\theta=2\pi S$ and $\varepsilon_{\mu\nu}$ is the rank-2 Levi-Civita antisymmetric tensor [@Haldane85; @Affleck85]. The Pontryagin index $$\label{topo2} Q={1\over8\pi}\sum_{\mu,\nu=0,1}\int d^2x\varepsilon_{\mu\nu}\hat{n}\cdot(\partial_{\mu}\hat{n}\times\partial_{\nu}\hat{n})= integer$$ measures how many times the 2\[=1(space)+1(time)\]-dimensional spin configuration $\hat{n}$ has wrapped around the unit sphere. In two dimensions, $$S_T\rightarrow{\hbar\theta\over2}\int dy {\partial Q(y)\over\partial y}=0,$$ where $Q(y)$ is the Pontryagin index that arises from summing over all spin configurations in the $y^{th}$ column of the two-dimensional lattice. The sum is zero for smooth spin configurations because $Q$ is an integer and thus cannot “change smoothly" [@Haldane88-NLSM; @Wen88; @Fradkin88; @Dombre88]. For the same reason, $S_T$ vanishes for any number of dimensions greater than one. However, one should be cautioned that this conclusion is valid only when we restrict ourselves to smooth spin configurations $\hat{n}(\mathbf{x},t)$ when computing $S_T$. The Berry’s phase may have a nonzero contribution if we also allow singular spin configurations in the theory. This is the case in $2+1$D, where monopole-like spin configurations are allowed in 3D space [@Haldane88-NLSM; @ReadSachdev90]. Quantum spin chains and the Haldane conjecture ---------------------------------------------- We now study the predictions of the effective action for quantum spin chains. In one dimension, the quantum spin chains are described by the path integral $$\int D[\hat{n}(x,t)]e^{{i\over\hbar}(S_{\sigma}(\hat{n})+S_T(\hat{n}))}.$$ We first consider the topological term. We note that $S_T=2\hbar\pi SQ$ and $e^{{i\over\hbar}S_T}=(-1)^{2SQ}$ ($Q=$ integer). In particular, $e^{{i\over\hbar}S_T}\equiv 1$ for integer spin chains, and the Berry’s phase has no effect on the effective action. However, $e^{{i\over\hbar}S_T}=\pm1$ for half-odd-integer spin chains, depending on whether $Q$ is even or odd. There is no further distinction between spin chains with different spin values $S$ in $S_T$. This result leads to the first part of the Haldane conjecture, namely, that fundamental differences exist between integer and half-odd-integer spin chains [@Haldane88-NLSM]. To proceed further, we first consider integer spin chains, where $e^{{i\over\hbar}S_T}\equiv 1$ and the system is described by the “pure" NL$\sigma$M $S_{\sigma}$. ### Integer spin chains We start by asking the following question: what are the plausible ground states described by $S_{\sigma}$? For this purpose, it is more convenient to consider a lattice version of $S_{\sigma}$: $$\label{nlsm3} S_{\sigma}\rightarrow{1\over2}\int dt\sum_i \left({1\over J}\left({\partial\hat{n}_i\over\partial t}\right)^2+JS^2\hat{n}_i\cdot\hat{n}_{i+1}\right),$$ with the corresponding Hamiltonian $$\label{hrotor} H_{\sigma}={J\over2}\sum_i\left((\mathbf{L}_i)^2-S^2\hat{n}_i\cdot\hat{n}_{i+1}\right),$$ where $\mathbf{L}_i$ is the angular momentum operator for the $i^{th}$ rotor. The Hamiltonian contains two competing terms, and we expect that it may describe two plausible phases, a strong coupling phase, in which the kinetic energy (first) term dominates, and a weak coupling phase, in which the potential energy (second) term dominates. A natural control parameter for this analysis is the spin magnitude $S$, which dictates the magnitude of the potential energy. In the first case (small $S$), in which the potential energy term is small, we expect that the ground state can be viewed, to a first approximation, as a product of local spin singlets, i.e., $\mathbf{L}=0$ states, $$|G\rangle=|0\rangle_1|0\rangle_2\cdots|0\rangle_N,$$ where $|0\rangle_i$ represents the $\mathbf{L}=0$ state for the rotor on site $i$. The lowest-energy excitations are $\mathbf{L}=1$ states separated from the ground state by an excitation gap $\sim\hbar^2J$. This picture is believed to be correct as long as the magnitude of the potential energy term is much smaller than the excitation energy for the $\mathbf{L}=1$ state. In the second case, in which the potential energy term dominates (large $S$), we expect that the ground state is a magnetically ordered (Néel state) with $\hat{n}_i=\hat{n}_0$ at all sites $i$, where the excitations are Goldstone modes of the ordered state (spin waves). It turns out that this naive expectation is valid only in dimensions of $d>1$. In one dimension, the magnetically ordered state is not stable because of quantum fluctuations associated with the Goldstone mode (Mermin-Wigner-Hohenberg Theorem), and the ground state is always quantum disordered [@Mermin66; @Hohenberg67], i.e., a spin liquid state. This result can be shown more rigorously through a renormalization group (RG) analysis of the NL$\sigma$M. We do not go through this analysis in detail in this article; instead, we simply assume that this is the case and examine its consequences. Readers interested in the RG analysis can consult, for example, references [@Polyakov75; @Polyakov87; @Brezin76]. Physically, this result means that after some renormalization, the ground state of integer spin chains can always be viewed as a product state of local spin singlets, irrespective of the spin magnitude $S$. The lowest-energy excitations are gapped spin triplet ($\mathbf{L}=1$) excitations. This is the Haldane conjecture for integer spin chains. ### Half-odd-integer spin chains The RG analysis cannot be straightforwardly applied to half-odd-integer spin chains because of the appearance of the topological term $S_T$. To understand why, let us again take the RG to the strong coupling limit and examine what happens in this case. To zeroth order, the Hamiltonian of the system consists only of the kinetic energy term. However, the rotors are moving under the influence of effective monopole potentials originating from $S_T$. In particular, all half-odd-integer spin chains have the same $S_T$ with an effective magnetic monopole strength of $1/2$, corresponding to that of a spin-$1/2$ chain. In this case, the ground state of a single rotor has an angular momentum of $\mathbf{L}=1/2$ and is two-fold degenerate (see the discussion after Eq. ). The total degeneracy of the ground state is $2^N$, where $N$=number of lattice sites. This enormous degeneracy implies that the coupling between rotors cannot be neglected when we consider the rotor Hamiltonian given in Eq. , and the strong coupling expansion simply tells us that the system behaves like a coupled-spin-$1/2$ chain [@Read90]. Fortunately, the antiferromagnetic spin-$1/2$ chain can be solved using the exact Bethe ansatz technique [@Giamarchi03]. The exact Bethe ansatz solution tells us that the antiferromagnetic spin-$1/2$ Heisenberg chain is critical, namely, the ground state has no long-range magnetic order but has a gapless excitation spectrum. Unlike integer spin chains, where the lowest-energy excitations carry spin $S=1$, the elementary excitation of this system has spin $S=1/2$. Combining this with the continuum theory leads to the Haldane conjecture for half-odd integer spin chains, namely, that they are all critical with elementary $S=1/2$ excitations. ### Open spin chains and end states The Haldane conjecture has been checked numerically for quantum spin chains with different spin magnitudes and has been found to be correct in all cases that have been studied thus far. One may wonder whether the difference in spin magnitudes may manifest at all in some low-energy properties of quantum spin chains. The answer is yes, when we consider open spin chains. Recall that we have always assumed periodic boundary conditions in deriving $S_T$. In fact, a periodic boundary condition is needed to define the Pontryagin index for the topological term $S_T$. For an open chain of length $L$, $S_T$ is replaced by [@Haldane83PLA; @Affleck90; @Ng94] $$\begin{aligned} \label{WZo} S_{T}^{(o)} & = &{\hbar\over2}\int^L_0 d^dx{\partial S_B(\hat{n}(x))\over \partial x} \\ \nonumber & = & 2\pi\hbar SQ+{\hbar S\over2}\left(\Omega(\hat{n}(L))-\Omega(\hat{n}(0))\right), \end{aligned}$$ where $2\pi SQ=\theta Q$ is the usual topological $\theta$ term that we obtain when $\Omega(\hat{n}(0))=\Omega(\hat{n}(L))$, i.e., when we consider periodic boundary conditions. An open chain differs from a closed chain in the existence of an additional boundary Berry’s phase term with an effective spin magnitude of $S/2$. We now examine the effect of this additional Berry’s phase term. First, we consider integer spin chains. Following the previous discussion, we expect the spin chain to be described by the strong coupling limit of the effective Hamiltonian given in Eq. , except that the rotors at the two ends of the spin chain are subjected to monopole potentials of strength $S/2$, resulting in effective free spins of magnitude $S/2$ located at the ends of the spin chain. The two spins are coupled by a term $J_{eff}\sim JS^2e^{-L/\xi}$ when the coupling between rotors is considered, where $\xi\sim E^{-1}_g$ is the correlation length and $E_g$ is the spin gap. These end states can also be understood based on a wavefunction proposed by Affleck, Lieb, Kennedy and Tasaki (the AKLT state) for $S=1$ spin chains [@AKLT] (see section IV) and have been observed experimentally in $S=1$ spin chain materials [@s1end]. In modern terminology, the end states of integer spin chains are a manifestation of symmetry-protected topological (SPT) order [@Gu09; @Chen12; @Pollmann12], which manifests itself as a boundary action that is protected by rotational (SO(3)) symmetry.[^1] For half-odd-integer spin chains, the analysis is a bit more complicated. We start by rewriting Eq.  for $S_T^{(o)}$ as follows [@Ng94]: $$\begin{aligned} \label{WZo1} S_{T}^{(o)} & = & {\hbar\over2}\left(4\pi{1\over2}Q+S\left(\Omega(\hat{n}(L))-\Omega(\hat{n}(0))\right)\right) \\ \nonumber & = & {\hbar\over2}\left(4\pi{1\over2}Q+{1\over2}\left(\Omega(\hat{n}(L))-\Omega(\hat{n}(0))\right)\right. \\ \nonumber & & +\left.(S-{1\over2})\left(\Omega(\hat{n}(L))-\Omega(\hat{n}(0))\right)\right) \end{aligned}$$ where we have replaced $S$ with $1/2$ in the usual topological (Pontryagin index) term and have divided the boundary Berry’s phase term into two parts; the first part, when combined with the Pontryagin index term, is the total Berry’s phase contribution for an open $S=1/2$ spin chain, and the second part is the additional contribution when $S>1/2$. Performing the strong coupling expansion as before, we find that the system behaves as an open spin-$1/2$ chain coupled to two end spins with a magnitude of ${1\over2}+{1\over2}(S-{1\over2})$. The problem of impurity end spins coupled to a spin-$1/2$ chain has been analyzed using the bosonization technique, through which it was found that after the screening induced by the spin-$1/2$ chain (essentially a Kondo effect), a free spin with a magnitude of ${1\over2}(S-{1\over2})$ is left at each end of the spin chain [@Eggert92]. Note that the existence of end states in half-odd-integer spin chains is rather non-trivial because the bulk spin excitations are [*gapless*]{}. As a result, the end spins at the two ends of a half-odd-integer spin chain are coupled by a term $J_{eff}\sim JS^2/(L\ln L)$, where $L$ is the length of the spin chain. The excitation energy of the end state is logarithmically lower than the energy of the bulk spin excitations, which have an energy of $\sim J/L$ [@Ng94]. These predictions for open chains and end states based on the NL$\sigma$M plus topological $\theta$ term analysis have been verified numerically by means of density matrix renormalization group (DMRG) calculations [@Qin95]. Higher dimensions and frustrated quantum antiferromagnets --------------------------------------------------------- The NL$\sigma$M approach to quantum antiferromagnets has been extended to higher dimensions and to frustrated quantum antiferromagnets. For simple antiferromagnets, $S_T$ vanishes in dimensions of $d>1$, and we need only consider the NL$\sigma$M, i.e., $S_{\sigma}$. As discussed before, $S_{\sigma}$ describes two plausible phases, the weak coupling phase, in which the ground state is antiferromagnetically ordered, and the strong coupling phase, in which the ground state is gapped. The weak coupling phase is favored for large spin magnitudes $S$. Various numerical and analytical studies have consistently demonstrated that the ground state is always Néel ordered for simple quantum antiferromagnets on a $2d$ square lattice, even for the smallest possible spin value of $S=1/2$ [@Manousakis91]. For this reason, physicists have turned to frustrated spin models to look for exotic spin liquid states. The NL$\sigma$M approach has generated interesting results when applied to weakly frustrated spin models, where the main effect of frustration is to reduce the effective coupling strength between rotors (for example, $J_1-J_2$ models, in which a next-nearest neighbor antiferromagnetic coupling is added to the Heisenberg model on a square lattice). In this case, it has been shown that spin-Peierls order can be obtained when discontinuous monopole-like spin configurations are included in the calculation of $S_T$ [@ReadSachdev90]. However, the method becomes questionable when applied to strongly frustrated spin systems, in which effective rotors become difficult to define locally, for example, the antiferromagnetic Heisenberg model on a kagome lattice. Generally speaking, a continuum theory is reliable only if the short-distance physics is captured correctly by the underlying classical or mean-field theory. A continuum theory becomes unreliable if the short-distance physics it assumes is not correct. This seems to be the case for the NL$\sigma$M approach when applied to strongly frustrated spin systems. In the following sections, we consider alternative methods of treating quantum spin systems, keeping in mind the physics that we have previously discussed. Resonant valence bond (RVB) states ================================== The semi-classical approach, which is based on fluctuations around a presumed classical (Néel) order, is difficult to apply in frustrated lattice models. The difficulties arise from two main sources. First, different degenerate or quasi-degenerate classical ground states may exist in a frustrated spin system. It is difficult to include these quasi-degenerate classical ground states in the NL$\sigma $M description. Second, the effect of Berry’s phases becomes intractable because of the complicated (classical) spin trajectory. The term geometric frustration (or frustration for short) was introduced by Gerard Toulouse in the context of frustrated magnetic systems [@Toulouse77; @Toulouse77a]. Indeed, frustrated magnetic systems had long been studied prior to that time. Early work included a study conducted by G. H. Wannier [@Wannier50] on the classical Ising model on a triangular lattice with antiferromagnetically coupled nearest neighbor spins, which serves as the simplest example of geometric frustration [@BookDiep]. Because of the AFM coupling, two nearest neighboring spins $A$ and $B$ tend to be anti-parallel. Then, a third spin $C$ that is a neighbor of both $A$ and $B$ is *frustrated* because its two possible orientations, up and down, both have the same energy. The classical ground state has a high level of degeneracy. As a result, we cannot choose a classical spin order as the starting point for constructing the NL$\sigma $M for the quantum $S=1/2$ $XXZ$ model $$H=J_z\sum_{\langle i,j\rangle}S_i^{(z)}S_j^{(z)}+J_{\perp}\sum_{\langle i,j\rangle}\left(S_i^{(x)}S_j^{(x)}+S_i^{(y)}S_j^{(y)}\right)$$ with $J_z>>J_{\perp}$ because there exist infinite spin configurations with the same classical energy. We note that the spin-spin correlation has been found to decay following a power law at zero temperature in the exact solution for the classical Ising model [@Stephenson70]. ![ Geometric frustration. The spin $C$ is frustrated because either the up or down orientation will give rise to the same energy in the AFM Ising limit.[]{data-label="fig:frustration"}](frustration.eps){width="3.6cm"} In this case, an alternative approach is a variational wavefunction, in which we essentially must *guess* the ground state wavefunction based on experience or physical intuition. A very important idea related to this approach is the resonating valence bond (RVB) concept for spin-$1/2$ systems suggested by Anderson. The term RVB was first coined by Pauling [@Pauling] in the context of metallic materials. Anderson revived interest in this concept in 1973 when he constructed a non-degenerate quantum ground state for an $S=1/2$ AFM system on a triangular lattice [@Anderson73]. A valence bond is a spin singlet state constructed from two $S=1/2$ spins at sites $i$ and $j$, given by $$(i,j)=\frac{1}{\sqrt{2}}(\left\vert \uparrow _{i}\downarrow _{j}\right\rangle -\left\vert \downarrow _{i}\uparrow _{j}\right\rangle ), \label{valencebond}$$ and an RVB state is a tensor product of valence bond states, whose wavefunction is given by$$\left\vert \Psi _{RVB}\right\rangle =\sum_{i_{1}j_{1}\cdots i_{n}j_{n}}a_{(i_{1}j_{1}\cdots i_{n}j_{n})}\left\vert (i_{1},j_{1})\cdots (i_{n},j_{n})\right\rangle , \label{RVB1}$$ where $(i_{1},j_{1})\cdots (i_{n},j_{n})$ are dimer configurations covering the entire lattice. The wavefunction is summed over all possible ways in which the lattice can be divided into pairs of lattice sites (i.e., dimers). The quantities $a_{(i_{1}j_{1}\cdots i_{n}j_{n})}$ are variational parameters determined by minimizing the ground-state energy of a given Hamiltonian. For a quantum disordered antiferromagnet, it has been proposed that the valence bond pairs in the RVB construction are dominated by short-range pairs, resulting in liquid-like states with no long-range spin order. The corresponding spin correlation function $\langle\mathbf{S}_i.\mathbf{S}_j\rangle$ in the RVB state may be short in range, with a finite correlation length (usually called short-range RVB (sRVB)), or may decay with distance following a power law (algebraic spin liquid states). The state is called a valence-bond solid (VBS) state if a single dimer configuration dominates in the ground state. An algebraic spin liquid state is usually invariant under all symmetry operations allowed by the lattice, whereas a VBS state usually breaks the translational or rotational lattice symmetry. ![ A spin-singlet dimer configuration covering a lattice. An RVB state is a superposition of such configurations.[]{data-label="fig:RVB"}](RVB.eps){width="6.4cm"} The wavefunction given in Eq. , which is parameterized by $a_{(i_{1}j_{1}\cdots i_{n}j_{n})}$, has too many variational degrees of freedom even after the translational and rotational symmetries of the wavefunction are considered and must be simplified for practical purposes. A solution has been proposed by Baskaran, Zou and Anderson [@BZA], who noted that the Bardeen-Cooper-Schrieffer (BCS) states for superconductors are direct product states of spin-singlet Cooper pairs and suggested that good RVB wavefunctions can be constructed from BCS wavefunctions [*via*]{} Gutzwiller projection, denoted by $P_{G}$: $$\begin{aligned} \left\vert \Psi _{RVB}\right\rangle &=&P_{G}\left\vert \Psi _{BCS}\right\rangle , \label{RVBPG} \\ \left\vert \Psi _{BCS}\right\rangle &=&\prod\limits_{\mathbf{k}}(u_{\mathbf{k% }}+v_{\mathbf{k}}c_{\mathbf{k}\uparrow }^{\dag }c_{-\mathbf{k}\downarrow }^{\dag })\left\vert 0\right\rangle , \notag\end{aligned}$$ where $c_{\mathbf{k}\uparrow }^{\dag }$ and $c_{-\mathbf{k}\uparrow }^{\dag } $ are electron creation operators and the numerical coefficients $u_{\mathbf{k}}$ and $v_{\mathbf{k}}$ are determined from a trial BCS mean-field Hamiltonian $H_{BCS}$ through the Bogoliubov-de Gennes equations, i.e., the RVB wavefunction is fixed by the parameters determining $H_{BCS}$. The number of electrons at each lattice site may take a value of 0, 1 or 2 in the original BCS wavefunctions. The Gutzwiller projection $P_{G}$ removes all wavefunction components with doubly occupied sites from the BCS state and freezes the charge degrees of freedom. A half-filled Mott insulator state is obtained if the total number of electrons is equal to the number of lattice sites. We note that the technique of Gutzwiller projection is currently being widely applied to other mean-field wavefunctions $\left\vert \Psi _{MF}\right\rangle $ to study Mott insulating states in diverse physical systems. Interesting and energetically favorable wavefunctions are often obtained when $\left\vert \Psi _{MF}\right\rangle $ is chosen properly. In addition to representing spins by electrons or fermions, one may also use Schwinger bosons to represent spins to construct RVB wavefunctions (see also the discussion after Eq. ). It is easy to recognize that in general, almost any mean-field wavefunction $\left\vert \Psi _{MF}\right\rangle $ can be employed to construct a corresponding spin state as follows: $$\left\vert \Psi _{Spin}\right\rangle =P_{G}\left\vert \Psi _{MF}\right\rangle , \label{RVBPG2}$$ where $\left\vert \Psi _{MF}\right\rangle $ is the ground state of a trial mean-field Hamiltonian $H_{trial}(c,c^{\dag};a_1,...,a_N)$, where $c_{i\sigma }^{\dag }(c_{i\sigma })$ can represent either fermions or bosons and $a_1,...,a_N$ are variational parameters determined by minimizing the energy of the parent spin Hamiltonian.[^2] The invention of Gutzwiller projection techniques enables us to construct a large variety of variational spin wavefunctions, of which the best is the one with the lowest energy. The most important difference between the fermion and boson constructions is that they lead to very different sign structures in the spin wavefunction $\left\vert \Psi _{RVB}\right\rangle $. In a bosonic wavefunction, when two spins (note that only spin degrees of freedom remain after Gutzwiller projection) at different sites are interchanged, the wavefunction does not change, whereas the wavefunction does change sign when two spins are interchanged in a fermionic wavefunction. These different sign structures represent very different quantum entanglement structures in the corresponding RVB wavefunctions. A famous example is Marshall’s sign rule [@Marshall55] for the AFM Heisenberg model on a bipartite lattice, where the Heisenberg exchange exists only between bonds linking sites in different sublattices. Marshall’s theorem tells us that the ground state for such an AFM system is a spin-singlet state with positive-definite coefficients in the Ising basis $\left\{(-1)^{N_{A\downarrow }}\left\vert \sigma _{1}\cdots \sigma_{N} \right\rangle \right\} $, where $N_{A\downarrow }$ is the number of down spins in sublattice $A$ and $N$ is the number of lattice sites. Using this result, Liang, Doucot and Anderson [@Liang88] proposed the use of the following trial ground-state RVB wavefunction for spin-$1/2$ Heisenberg antiferromagnets on a square lattice: $$\begin{aligned} \left\vert \Psi _{LDA}\right\rangle &=&\sum_{i_{\alpha }\in A,j_{\beta }\in B}h(i_{1}-j_{1})\cdots h(i_{n}-j_{n}) \notag \\ &&\times (-1)^{N_{A\downarrow }}\left\vert (i_{1},j_{1})\cdots (i_{n},j_{n})\right\rangle , \label{LDA} \end{aligned}$$ where $h(r)$ represents a positive-definite function of the bond length $r$. This particular wavefunction can be conveniently represented as a Gutzwiller-projected wavefunction in the Schwinger boson representation, whereas the representation of the same wavefunction in terms of fermions is far from straightforward [@Read89a]. However, it has been shown that the projected BCS wavefunction given in Eq.  (\[RVBPG\]) will satisfy Marshall’s sign rule provided that the spatial Fourier transformation of $u_{\mathbf{k}}$ and $v_{\mathbf{k}}$ (= $u_{ij}$ and $v_{ij}$) connects only sites in different sublattices in a bipartite lattice [@LiTao07; @Yunoki06] It has been noted by Ma [@Ma88] that the sum of states $\left\vert(i_{1},j_{1})\cdots (i_{n},j_{n})\right\rangle $, with $i_{\alpha }\in A$ and $j_{\beta }\in B$, forms an overcomplete set for spin-singlet states in a bipartite lattice. Because $h$ is a positive function, it can be interpreted as a weight factor in a Monte Carlo simulation based on loop gas statistics. Such a calculation has been performed for large lattices by Liang *et al.* [@Liang88], and a very accurate ground-state wavefunction for the AFM Heisenberg model on a square lattice was obtained. The wavefunction can give rise to either long-range or short-range spin correlations depending on the choice of $h(r)$. ![ A spinon excitation on top of an RVB ground state.[]{data-label="RVB-spinon"}](RVB-spinon.eps){width="6.4cm"} Once a proper RVB ground-state wavefunction has been constructed, the next natural question is what are the low-energy dynamics, or the elementary excitations on top of the ground states? A natural candidate for excitation is to break a spin-singlet pair in the ground state to form a spin-triplet excited state with two unpaired spins. For a long-range magnetically ordered state, it has been found that the two unpaired spins will bind together closely in space and that the resulting elementary excitations will be localized spin-triplet excitations with well-defined energy and momentum. This is nothing but a spin wave or magnon excitation, as guaranteed by the Goldstone theorem. By contrast, for a QSL state with short-range spin correlation, it has been proposed that the two unpaired spins may interact only weakly with each other and can be regarded as independent spin-$1/2$ elementary excitations called spinons. The existence of $S=1/2$ spinon excitations is one of the most important predictions in QSLs and is crucial to the experimental verification of QSLs. The process through which a spin-$1$ magnon turns into two independent spin-$1/2$ spinons is an example of *fractionalization*. Whether fractionalization of spin excitations actually occurs in a particular spin system is a highly non-trivial question. A systematic way to examine whether fractionalization may occur in a spin model was first proposed by X.G. Wen [@Wen89-Confinment; @Wen91] based on the concept of confinement/deconfinement in lattice gauge theory.[^3] This approach is explained in the following subsection, where the gauge theory for QSLs is introduced. RVB theory and gauge Theory --------------------------- This subsection presents a brief survey of how RVB theory is implemented in practice, especially how low-energy effective field theories for QSL states are constructed, which is crucial for characterizing QSLs. We discuss a few common examples of QSLs and define the $SU(2)$, $U(1)$ and $Z_2$ spin liquid states. The nature of the $U(1)$ QSL state is then further illuminated by relating it to a Fermi liquid state through a Mott metal-insulator transition. We shall see that analytical approaches have strong limitations and should be complemented by numerical approaches in practice. One complication associated with the RVB construction is that there exist, in general, different mean-field states $\left\vert \Psi_{MF}\right\rangle $ that correspond to the same RVB spin wavefunction after Gutzwiller projection. This redundancy originates from the enlarged Hilbert space in the boson/fermion representation for spins and is called *gauge redundancy* or *gauge symmetry*. Gutzwiller projection removes this redundancy, resulting in a unique state in spin Hilbert space. To see how this occurs, we consider the fermion representation of $S=1/2$ spin operators [@Abrikosov65; @BZA; @Baskaran88]: \[slaveparticle\] $$\vec{S}_{i}=\frac{1}{2}\sum_{\alpha \beta }f_{i\alpha }^{\dag }\vec{\sigma}_{\alpha \beta }f_{i\beta }, \label{spin-fermion}$$ where $\alpha ,\beta =\uparrow ,\downarrow $ are spin indices, $f_{i\alpha }^{\dag }(f_{i\alpha })$ is the fermion creation (annihilation) operator, and $\vec{\sigma}=(\sigma ^{1},\sigma^{2},\sigma ^{3})$ represents the Pauli matrices. It is easy to confirm that the three components of $\vec{S}_{i}$ satisfy the $SU(2)$ Lie algebra relation, $[S_{i}^{\lambda },S_{j}^{\mu }]=i\epsilon _{\lambda \mu \nu }S_{i}^{\nu }\delta _{ij}$, where $\lambda ,\mu ,\nu =1,2,3$ and $\epsilon _{\lambda \mu \nu }$ is the antisymmetric tensor. Hence, Eq.  is a representation of $SU(2)$ spins. However, the local Hilbert space for two fermions contains four Fock states, $\left\{ \left\vert 0\right\rangle,f_{\uparrow }^{\dag }\left\vert 0\right\rangle =\left\vert\uparrow \right\rangle ,f_{\downarrow }^{\dag }\left\vert 0\right\rangle =\left\vert \downarrow \right\rangle ,f_{\uparrow}^{\dag }f_{\downarrow }^{\dag }\left\vert 0\right\rangle =\left\vert \uparrow \downarrow \right\rangle \right\} $; this is larger than the physical spin Hilbert space for spin-$1/2=\left\{ \left\vert \uparrow \right\rangle ,\left\vert \downarrow \right\rangle\right\} $, and we need to impose the single-occupancy constraint $$\sum_{\alpha }f_{i\alpha }^{\dag }f_{i\alpha }=1 \label{n-constraint}$$ to remove the unphysical states to obtain a proper spin representation. This is what the Gutzwiller projection does. The construction presented in Eq.  is equally applicable for bosons (the Schwinger boson representation) because the $SU(2)$ Lie algebra is independent of the statistics of the represented particles. In the following, we focus on the fermion representation approach because it has been found to be a more fruitful approach for constructing QSLs. Readers who are interested in the Schwinger boson approach may refer to reference [@Arovas88] for details. There are multiple choices of $\left\{ f_{i\alpha }\right\} $ available to represent spin operators even once the single-occupancy constraint is satisfied and the statistics of the particles have been chosen. For example, a new set of $\left\{f_{i\alpha }\right\} $ can be obtained through an $U(1)$ gauge transformation: $$f_{i\alpha }\rightarrow f_{i\alpha }^{\prime }=e^{i\theta(i)}f_{i\alpha }.$$ It is easy to verify that $\left\{ f_{i\alpha }^{\prime }\right\}$ forms another representation of spin operators by replacing $f_{i\alpha }$ with $% f_{i\alpha }^{\prime }$ in Eq. , independent of whether the $f$s are fermions or bosons. This multiplicity is called gauge redundancy or gauge symmetry in the literature. We call it gauge redundancy here because symmetry usually refers to situations in which there are multiple physically distinct states with the same properties, e.g., there is a degeneracy in energy. However, the gauge degree of freedom we discuss here is not a “real" symmetry among different physical states. Here, two gauge-equivalent states are the [*same*]{} state in the spin Hilbert space. They just “look" different when they are represented by particles that live in an enlarged Hilbert space. There is no way to distinguish them physically [@Wen02-PSG]. The gauge redundancy in the fermion representation of $S=1/2$ spins extends beyond $U(1)$. There exists an additional $SU(2)$ gauge structure because of the particle-hole symmetry in the fermion representation, which is absent in the Schwinger boson representation. An elegant way of showing this $SU(2)$ gauge structure was suggested by Affleck, Zou, Tsu and Anderson [@AZHA88], who introduced the following $2\times 2$ matrix operator: $$\Psi =\left( \begin{array}{cc} f_{\uparrow } & f_{\downarrow }^{\dag } \\ f_{\downarrow } & -f_{\uparrow }^{\dag }% \end{array}\right) . \label{matrixpsi}$$ It is straightforward to show that the spin operator can be re-expressed in terms of $\Psi $ as $$\vec{S}_{i}=\text{tr}\left( \Psi _{i}^{\dagger }\vec{\sigma}\Psi_{i}\right) \text{.} \label{SSU2}$$ The single-occupancy condition given in Eq.  also leads to the identities $$\label{v-constraint} f_{i\uparrow }f_{i\downarrow }=f_{i\uparrow }^{\dag }f_{i\downarrow }^{\dag }=0.$$ Together with Eq. , Eq.  can be rewritten in the following compact vector form: $$\text{tr}\left( \Psi _{i}\vec{\sigma}\Psi _{i}^{\dagger }\right) =0. \label{n-constraintSU2}$$ We now consider the following $SU(2)$ gauge transformation of $\Psi $: $$\Psi _{i}\rightarrow \Psi _{i}^{^{\prime }}=\Psi_{i}W_{i},W_{i}\in SU(2). \label{SU2transform}$$ The spin operator $\vec{S}_{i}$ in Eq.  remains invariant under this transformation because $W_{i}W_{i}^{\dag }=1$. The single-occupancy constraint given in Eq.  is also invariant because $W_{i}\vec{\sigma}W_{i}^{\dag }$ represents a rotation of vector $\vec{\sigma}$ but all components of tr$\left( \Psi _{i}\vec{\sigma}\Psi _{i}^{\dagger}\right) $ are zero, i.e., $\Psi _{i}\rightarrow \Psi_{i}^{^{\prime}}=\Psi_{i}W_{i}$ is also a valid representation for $S=1/2$ spins. We show now how RVB theory is implemented in an analytical fermionic approach. For concreteness, we consider an AFM Heisenberg model on a lattice: $$H=J\sum_{\left\langle ij\right\rangle }\vec{S}_{i}\cdot \vec{S}_{j}, \label{HAFM}$$ where $\left\langle ij\right\rangle $ denotes a nearest neighbor bond and $ J>0$. The spin exchange $\vec{S}_{i}\cdot \vec{S}_{j}$ can be written in terms of fermionic (spinon) operators: $$\vec{S}_{i}\cdot \vec{S}_{j}=\frac{1}{4}\sum_{\alpha \beta }\left( 2f_{i\alpha }^{\dag }f_{i\beta }f_{j\beta }^{\dag }f_{j\alpha }-f_{i\alpha }^{\dag }f_{i\alpha }f_{j\beta }^{\dag }f_{j\beta }\right) , \label{SiSj1}$$ where we have used the relation $\vec{\sigma}_{\alpha \beta }\cdot\vec{\sigma}_{\alpha ^{\prime }\beta ^{\prime }}=2\delta _{\alpha \beta ^{\prime }}\delta _{\alpha ^{\prime }\beta }-\delta _{\alpha\beta }\delta _{\alpha ^{\prime }\beta ^{\prime }}$. The constraint given in Eq.  or can be imposed by inserting delta functions into the imaginary-time path integral. The corresponding partition function is $$\begin{aligned} Z &=&\int D[f,\bar{f}]\exp [-S(f,\bar{f})]\prod_{i}\delta \left( \sum\nolimits_{\alpha }\bar{f}_{i\alpha }f_{i\alpha }-1\right) \notag \\ &&\times \delta \left( \sum\nolimits_{\alpha \beta }\epsilon _{\alpha \beta }f_{i\alpha }f_{i\beta }\right) \delta \left( \sum\nolimits_{\alpha \beta }\epsilon _{\alpha \beta }\bar{f}_{i\alpha }\bar{f}_{i\beta }\right) , \label{Z1}\end{aligned}$$ where the action $S(f,\bar{f})$ is given by $$S(f,\bar{f})=\int_{0}^{\beta }d\tau \left( \sum_{i\alpha }\bar{f}_{i\alpha }\partial _{\tau }f_{i\alpha }-H\right) . \label{S1}$$ The delta functions can be represented by the integration over real auxiliary fields $a_{0}^{l}(i)$ on all sites $i$, $l=1,2,3$. Using the relation $\delta \left( x\right) =\int \frac{dk}{2\pi }e^{ikx}$, we obtain$$Z=\int D[f,\bar{f};a]\exp [-S(f,\bar{f};a)], \label{Z2}$$with$$\begin{aligned} S(f,\bar{f};a) &=&S(f,\bar{f})-i\left\{ \sum_{i}a_{0}^{3}\left( \sum\nolimits_{\alpha }\bar{f}_{i\alpha }f_{i\alpha }-1\right) \right. \notag \\ &&\left. +\left[ (a_{0}^{1}+ia_{0}^{2})\sum\nolimits_{\alpha \beta }\epsilon _{\alpha \beta }f_{i\alpha }f_{i\beta }+h.c.\right] \right\} . \label{S2}\end{aligned}$$ It is generally believed (but has not been proven) that the partition function $Z$ will remain invariant under a Wick rotation of the fields $% a_{0}^{l}$ in the path integral, namely, we can replace $ia_{0}^{l}$ with $a_{0}^{l}$. Then, the action becomes $$\begin{aligned} S(f,\bar{f};a) &=&S(f,\bar{f})-\left\{ \sum_{i}a_{0}^{3}\left( \sum\nolimits_{\alpha }\bar{f}_{i\alpha }f_{i\alpha }-1\right) \right. \notag \\ &&\left. +\left[ (a_{0}^{1}+ia_{0}^{2})\sum\nolimits_{\alpha \beta }\epsilon _{\alpha \beta }f_{i\alpha }f_{i\beta }+h.c.\right] \right\} . \label{S3}\end{aligned}$$ The action given in Eq.  serves as the starting point for theoretical analysis. The path integral is difficult to solve, and approximate methods are generally needed. We start with a mean-field theory in which we assume that the path integral is dominated by saddle points characterized by equal-time expectation values of the operators $\sum\nolimits_{\alpha }f_{i\alpha }^{\dag }f_{i\alpha}$, $\sum\nolimits_{\alpha \beta }\epsilon _{\alpha \beta }f_{i\alpha}f_{i\beta}$ and $a_0^{l}(i)$: $$\begin{aligned} \chi _{ij} &=&\sum\nolimits_{\alpha }\left\langle f_{i\alpha }^{\dag }f_{j\alpha }\right\rangle , \notag \\ \Delta _{ij} &=&\sum\nolimits_{\alpha \beta }\epsilon _{\alpha \beta }\left\langle f_{i\alpha }f_{j\beta }\right\rangle , \notag \\ a_0^{l} & = & \langle a_0^{l}(i)\rangle, \label{MFparameter}\end{aligned}$$ where $\epsilon_{\alpha\beta}$ is the totally antisymmetric tensor ($\epsilon_{\uparrow\downarrow}$=1), $l=1,2,3$. It is easy to verify that $\chi _{ij}$ and $\Delta _{ij}$ satisfy the relations $\chi _{ij}=\chi _{ji}^{\ast }$ and $\Delta_{ij}=\Delta _{ji}$. Note that any time-dependent fluctuations in $\Delta_{ij}, \chi_{ij}$ and $a_{0}^{l}(i)$ are ignored in mean-field theory. With these approximations, we arrive at the following mean-field Hamiltonian:$$\begin{aligned} H_{MF} &=&\sum_{\left\langle ij\right\rangle }-\frac{3}{8}J\left[ (\chi _{ji}\sum\nolimits_{\alpha }f_{i\alpha }^{\dag }f_{j\alpha }\right. \notag \\ &&+\left. \Delta _{ij}\sum\nolimits_{\alpha \beta }\epsilon _{\alpha \beta }f_{i\alpha }^{\dag }f_{j\beta }^{\dag }+h.c)-|\chi _{ij}|^{2}-|\Delta _{ij}|^{2}\right] \notag \\ &&+\sum_{i}\left\{ a_{0}^{3}\left( \sum\nolimits_{\alpha }f_{i\alpha }^{\dag }f_{i\alpha }-1\right) \right. \notag \\ &&\left. +\left[ (a_{0}^{1}+ia_{0}^{2})\sum\nolimits_{\alpha \beta }\epsilon _{\alpha \beta }f_{i\alpha }f_{i\beta }+h.c.\right] \right\} , \label{HMF}\end{aligned}$$ where $\chi _{ij}$, $\Delta _{ij}$ and $a_{0}^{l}$ are determined by minimizing the ground-state energy with the exact constraint condition(\[n-constraint\]) replaced with the average constraint $$\sum_{\alpha }\langle f_{i\alpha }^{\dag }f_{i\alpha }\rangle=1. \label{a-constraint}$$ The spin exchange term $\vec{S}_{i}\cdot \vec{S}_{j}$ in Eq.  can be evaluated within the mean-field assumption using the Wick theorem. Maintaining spin rotation invariance in the calculation, we obtain $$\label{mfenergy} \left\langle \vec{S}_{i}\cdot \vec{S}_{j}\right\rangle =-\frac{3}{8}\left( \chi _{ij}^{\ast }\chi _{ij}+\Delta _{ij}^{\ast }\Delta _{ij}\right) .$$ in mean-field theory. Physically, the mean-field theory outlined above is equivalent to assuming that the ground state of the spin system is given by a mean-field wavefunction $\left\vert \Psi_{MF}\right\rangle$ without Gutzwiller projection. The spin exchange energy evaluated in this way is usually not a good estimate of the energy of the “real" spin wavefunction. In practice, this mean-field theory provides an effective way to obtain a BCS Hamiltonian to construct a Gutzwiller-projected wavefunction. Whether the spin wavefunction obtained through Gutzwiller projection is a good wavefunction for the spin Hamiltonian can only be tested by evaluating the energy of the wavefunction numerically (see section \[GutzwillerP\]). In the following section, we assume that the Gutzwiller-projected wavefunction $P_G\left\vert \Psi_{MF}\right\rangle$ is a sufficiently good starting point to locate the true ground state of the spin Hamiltonian. In this case, we expect that the ground and low-energy states constructed from $H_{MF}$ are adiabatically connected to the corresponding Gutzwiller-projected wavefunctions and that we may construct an effective low-energy Hamiltonian/Lagrangian of the spin system from fluctuations around $H_{MF}$ through the usual path integral technique. The fluctuations in $\Delta_{ij},\chi_{ij}$ and $a^l_0(i)$ describe spin-singlet excitations and are usually called *gauge fluctuations*. Before discussing gauge fluctuations, we first discuss the effect of gauge redundancy on the mean-field states. To illustrate, we consider two mean-field QSL states with different structures of the mean-field parameters $\left\{ \chi _{ij},\Delta _{ij},a_{0}^{l}(i)\right\} $. We place the states on a simple square lattice. The first state is the uniform RVB state with \[examples\] $$\begin{aligned} \label{urvb} \chi _{ij} &=&0, \\ \nonumber \Delta _{ij} &=&\left\{ \begin{array}{cc} \Delta , & \text{NN bonds,} \\ \nonumber 0, & \text{others,}% \end{array}% \right. \\ \nonumber a_{0}^{l} & = &0 \ (l=1,2,3).\end{aligned}$$The second example considered is the zero-flux state given by$$\begin{aligned} \label{zeroflux} \chi _{ij} &=&\left\{ \begin{array}{cc} \chi , & \text{NN bonds,} \\ \nonumber 0, & \text{others,}% \end{array}% \right. \\ \Delta _{ij} &=&0, \\ \nonumber a_{0}^{l} &=& 0 \ (l=1,2,3).\end{aligned}$$ $\Delta$ and $\chi $ are real numbers. We show that irrespective of their very different appearances, these two mean-field ansatze actually give rise to the same spin state after Gutzwiller projection. The two states are gauge equivalent because they can be transformed into each other through a proper gauge transformation. The Hamiltonian given in Eq.  retains a local $SU(2)$ structure, which originates from the gauge redundancy in the fermion representation of spin. This local $% SU(2)$ symmetry becomes explicit if we introduce a doublet field $\psi =\left( f_{\uparrow },f_{\downarrow }^{\dag }\right) ^{T}$ and a $2\times 2$ matrix$$u_{ij}=\left( \begin{array}{cc} \chi _{ij} & \Delta _{ji}^{\ast } \\ \Delta _{ij} & -\chi_{ji}% \end{array}\right) .$$The mean-field Hamiltonian can be written in a compact manner as$$\begin{aligned} H_{MF} &=&\sum_{\left\langle ij\right\rangle }\frac{3}{8}J\left[ \frac{1% }{2}\text{Tr}(u_{ij}^{\dag }u_{ij})-(\psi _{i}^{\dag }u_{ij}\psi _{j}+h.c.)% \right] \notag \\ &&+\sum_{i,l}a_{0}^{l}\psi _{i}^{\dag }\tau ^{l}\psi _{i}, \label{HMFSU2}\end{aligned}$$where the $\tau ^{l},\ l=1,2,3$, are the Pauli matrices. From Eq.  we can clearly see that the Hamiltonian $H_{MF}$ is invariant under a local $% SU(2)$ transformation $W_{i}$: $$\begin{aligned} \psi _{i} &\rightarrow &W_{i}\psi _{i}, \notag \\ u_{ij} &\rightarrow &W_{i}u_{ij}W_{j}^{\dag }. \label{SU2MF}\end{aligned}$$This $SU(2)$ gauge transformation is the same as that in (\[SU2transform\]), where $\Psi =\left( \psi ,i\sigma _{2}\psi ^{\dag }\right) ^{T}$. Because of this $SU(2)$ gauge structure, if we regard the ansatz $\left( u_{ij},a_{0}^{l}\tau ^{l}\right)$ as labeling a physical spin wavefunction $|\Psi _{spin}^{\left(u_{ij},a_{0}^{l}\tau ^{l}\right)}\rangle=P_{G}|\Psi_{MF}^{(u_{ij},a_{0}^{l} \tau ^{l})}\rangle$, then such a label is not a one-to-one label. Two ansatze related by an $SU(2)$ gauge transformation, $\left(u_{ij},a_{0}^{l}\tau ^{l}\right) $ and $\left(u'_{ij},a_{0}^{'l}\tau ^{l}\right)=\left(W(u_{ij}),W(a_{0}^{l}\tau ^{l})\right) $, label the same physical spin wavefunction:$$\begin{aligned} |\Psi _{spin}(\{\alpha _{i}\})\rangle &=&P_{G}|\Psi_{MF}^{(W(u_{ij}),W(a_{0}^{l}\tau ^{l}))}\rangle \notag \\ &=&P_{G}|\Psi _{MF}^{(u_{ij},a_{0}^{l}\tau ^{l})}\rangle \label{Psi-PG2}\end{aligned}$$ where $W(u_{ij})=W_{i}u_{ij}W_{j}^{\dagger }$ and $W(a_{0}^{l}(i)\tau^{l})=W_{i}a_{0}^{l}(i)\tau ^{l}W_{i}^{\dagger }$, $W_{i}\in SU(2)$. The uniform RVB state and the zero-flux state discussed above denote the same physical spin state because they are related by a gauge transformation, $$W_i=\exp\left(i{\pi\over4}\tau^2\right).$$ More generally, the existence of gauge redundancy implies that the low-energy fluctuations in spin systems have a similar redundancy. To measure gauge fluctuations, we introduce the loop variables $$P(C_i)=u_{ij}u_{jk}\cdots u_{li},$$ where $i,j,k,\cdots,l$ denote a loop of lattice sites that passes through site $i$. $P(C_i)$ measures gauge fluxes and has the general form $$P(C_i)=A(C_i) \tau^0 +\mathbf{B}(C_i)\cdot\vec{\tau},$$ where $\tau^0$ is the identity matrix and $\vec{\tau}=\{\tau^1,\tau^2,\tau^3\}$ represents the Pauli matrices, $A(C_i)$ and $\mathbf{B}(C_i)$ measure the $U(1)$ and $SU(2)$ components, respectively, of the gauge flux. For a translationally invariant mean-field state, we can find a gauge with $\mathbf{B}(C_i)=\hat{n}B(C_i)$, where $A(C_i)$ and $B(C_i)$ are proportional to the area of the loop. Under a gauge transformation, $$P(C_i)\rightarrow W_iP(C_i)W_i^{\dag},$$ and the “direction" of $\hat{n}$ changes. The presence of gauge redundancy means that we may perform gauge transformations to change the “local" directions of $\hat{n}$, but the physical spin state remains unchanged. For a given mean-field state, it is useful to distinguish between two kinds of gauge transformations: those that change the mean-field ansatz $\left\{ u_{ij},a_{0}^{l}(i)\right\} $ and those that do not. The latter constitute a subgroup of the original $SU(2)$ symmetry called an invariant gauge group (IGG) [@Wen02-PSG]: $$IGG\equiv \left\{ W_{i}|W_{i}u_{ij}W_{j}^{\dagger }=u_{ij},W_{i}\in SU(2)\right\} . \label{IGG}$$ It can be shown rather generally that for a stable QSL state, physical gapless gauge excitations exist only for those fluctuations belonging to the IGG of the corresponding mean-field ansatz. Therefore, it is important to understand the structure of the IGGs in spin liquid states. Within the fermionic $SU(2)$ formalism, there are only three plausible kinds of IGG: $SU(2)$, $U(1)$ and $Z_{2}$. We call the corresponding spin liquids $SU(2)$, $U(1)$ and $Z_{2}$ spin liquids. $SU(2)$ spin liquids have $\mathbf{B}(C_i)=0$ with $IGG=SU(2)$. They are rather unstable because of the existence of a large amount of gapless $SU(2)$ gauge field fluctuations. $U(1)$ spin liquids have $\mathbf{B}(C_i)$ pointing in only one direction for all loops $C_i$. The condensation of fluxes in one “direction" provides an Anderson-Higgs mechanism for $SU(2)$ fluxes in “directions" perpendicular to $\mathbf{B}(C)$ and turns the IGG into $U(1)$. The low-energy fluctuations are $U(1)$ gauge field fluctuations. $Z_2$ spin liquids have $\mathbf{B}(C_i)$ pointing in different directions for different loops that pass through the same site $i$. The gauge fluctuations are all gapped because the Anderson-Higgs mechanism now applies to fluxes in all directions. A few examples of mean-field ansatze for these three types of spin liquid states are presented in the following subsections. U(1) gauge fluctuations ----------------------- We briefly discuss the $U(1)$ gauge theory in regard to two examples of spin liquids that are believed to exist in nature (see section V). The first example is the zero-flux state given in Eq. , for which $\Delta_{ij}=a_{0}^{l}=0$ and $\chi_{ij}=\chi$ in the mean-field ansatz. It is easy to see that $\mathbf{B}(C_i)\equiv\mathbf{0}$ and that the IGG of such a QSL is $SU(2)$, i.e., the zero-flux state describes a $SU(2)$ spin liquid. The low-energy fluctuations are $SU(2)$ gauge fluctuations. Here, we do not consider the full $SU(2)$ gauge fluctuations; we consider only the phase fluctuations of $\chi _{ij}$, i.e., $U(1)$ gauge fluctuations. The consideration of only $U(1)$ gauge fluctuations for the zero-flux state can be justified in a slave-rotor theory for the Hubbard model [@Lee05] or in a phenomenological Landau Fermi-liquid-type approach for spin liquid states [*near the metal-insulator transition*]{} (see the next subsection). Upon writing $\chi_{ij}=\chi e^{ia_{ij}}$, where $a_{ij}$ denotes phase fluctuations, it is straightforward to see that $$P(C_i)\propto\exp\left(i\Phi(C_i)\mathbf{\tau}^3\right),$$ where $\Phi(C_i)=\left(a_{ij}+a_{jk}+\cdots+a_{li}\right)$ is the total $U(1)$ gauge flux enclosed by the loop, i.e., the phase fluctuations of $\chi_{ij}$ represent one component of the $SU(2)$ gauge fluctuations. The effective Lagrangian describing these low-energy phase fluctuations is $$\begin{aligned} L^{(0)} &=&\sum_{i\alpha }\bar{f}_{i\alpha }(\partial _{\tau }-a_{0})f_{i\alpha } \notag \\ &&+\frac{3}{8}\sum_{\left\langle ij\right\rangle }\left( J\chi e^{ia_{ji}}\sum\nolimits_{\alpha }\bar{f}_{i\alpha }f_{j\alpha}+h.c.\right), \label{Leff1}\end{aligned}$$ and the corresponding Lagrangian in the continuum limit is $$\begin{aligned} L^{(0)} &=&\int d\vec{r}\sum_{\alpha }\bar{f}_{\alpha }(\vec{r})(\partial _{\tau }-a_{0})f_{\alpha }(\vec{r}) \notag \\ &&+\frac{1}{2m^{\ast }}\bar{f}_{\alpha }(\vec{r})(-i\triangledown +\vec{a}% )^{2}f_{\alpha }(\vec{r}), \label{Leff2}\end{aligned}$$ where $m^{\ast }$ is the effective mass for the spinon energy dispersion determined by $J\chi$ and the vector field $\vec{a}(\vec{r})$ is given by the lattice gauge field $a_{ij}$ through$$a_{ij}=(\vec{r}_{i}-\vec{r}_{j})\cdot \vec{a}\left( \frac{\vec{r}_{i}+\vec{r}% _{j}}{2}\right) .$$ Thus, the low-energy effective field theory describes non-relativistic spin-$1/2$ fermions (spinons) coupled to the $U(1)$ gauge field $\left(a_0(\vec{r}), \vec{a}(\vec{r})\right)$ in the continuum limit. The other spin liquid state we introduce here is the $\pi$-flux state [@Affleck-Marston88; @Kotliar88] on a square lattice given by $\Delta_{ij}=a_0^{l}=0$ and $$\begin{aligned} \label{piflux} \chi _{i,i+\hat{\mu}} &=&\left\{ \begin{array}{cc} \chi , & \mu=x, \\ i\chi(-1)^{i_x}, & \mu=y. \end{array}\right. \end{aligned}$$ It is easy to see that $P(C_i)\propto\exp\left(i\pi\mathbf{\tau}^3\right)$ per square plaquette in the mean-field ansatz, i.e., the $\pi$-flux state has $IGG=U(1)$ and is a $U(1)$ spin liquid. The zero-flux and $\pi$-flux states are physically distinct states because of their different IGGs. Their mean-field spinon dispersions are also qualitatively different. The zero-flux state has a mean-field dispersion of $E_0(\vec{k})=-J\chi(\cos k_x+\cos k_y)$, whereas the $\pi$-flux state has $E_{\pi}(\vec{k})=\pm J\chi\sqrt{\cos^2k_x+\cos^2k_y}$ with a reduced Brillouin zone. The continuum theory describes non-relativistic fermions with a large Fermi surface in the zero-flux state and describes Dirac fermions with four Fermi points ($\mathbf{k}=(\pm\pi/2,\pm\pi/2)$) in the $\pi$-flux state [@Affleck-Marston88]. The effective continuum theory for the $\pi$-flux state has the form $$\label{l22} L^{(\pi)}=\sum_{\mu\sigma}\left(\bar{\psi}_{+\sigma}(\partial_{\mu}-ia_{\mu})\tau_{\mu}\psi_{+\sigma}+ \bar{\psi}_{-\sigma}(\partial_{\mu}-ia_{\mu})\tau_{\mu}\psi_{-\sigma}\right),$$ where $\mu=0,1,2$. The two-component Dirac spinor fields $\psi_{\pm\sigma}$ describe two inequivalent Dirac nodes in the spinon spectrum [@Affleck-Marston88]. The two effective low-energy Lagrangians $L^{(0)}$ and $L^{(\pi)}$ describe two different types of spin liquid states that are believed to exist in nature. We discuss these states again in section V. The continuum action $L$ serves as the starting point for studying the stability and low-energy properties of spin liquid states. Integrating out the fermion fields (at each momentum shell) gives rise to a Maxwellian potential energy term in the gauge field: $$\frac{1}{2g^{2}(\Lambda)}(\triangledown \times \vec{a})^{2},$$ where $g(\Lambda)$ is a running gauge coupling constant in the sense of renormalization group theory, which depends on the energy or momentum scale $\Lambda$. If $g(\Lambda)\rightarrow 0$ in the low-energy and long-wavelength limit of $\Lambda\rightarrow0$, then the gauge fluctuations become increasingly weak. The corresponding interaction between two fermions becomes too weak to bind them together, and the elementary excitations in the spin system are spin-$1/2$ fermionic excitations called spinons. This phenomenon is called deconfinement, and the ground state is a filled Fermi sea of spinons. By contrast, if $g(\Lambda)\rightarrow \infty$ as $\Lambda\rightarrow0$, then two spinons will always be confined together to form a magnon. This phenomenon is called confinement. In this case, the mean-field QSL ground state breaks down into a spin-ordered state because of the strong gauge fluctuations, and magnon excitations are recovered in this ordered state. It is not exactly clear which kinds of mean-field QSL states are stable against gauge fluctuation. It is generally believed that $Z_{2}$ QSL states are stable because $Z_{2}$ (Ising) gauge theories are deconfining [@Fradkin79], whereas $SU(2)$ QSL states are unstable because of the presence of large gauge fluctuations. The case of $U(1)$ QSL states is more nontrivial. The $SU(2)$ gauge group and the corresponding gauge fields are compact in spin liquid states. To reflect the compactness of the $U(1)$ gauge group, one must replace the electromagnetic field tensor $F_{\mu\nu}^2$ with $2(1-\cos F_{\mu\nu})$. This periodic version of $U(1)$ gauge theory is called compact $U(1)$ gauge theory. A pure compact $U(1)$ gauge theory always gives rise to confinement in two dimensions [@Polyakov75; @Polyakov77], but whether deconfinement is possible in the presence of a matter field is an open question. Herbut *et al.* have argued that the theory is always confining in the presence of a Fermi surface [@Herbut03HSSM] or nodal fermions [@Herbut03HS]. Their conclusion depends on an approximate effective action for the gauge field obtained by integrating out the fermions to the lowest order. However, this approximation is questionable for gapless fermions. Indeed, Hermele *et al.* [@Hermele04] proved that when the spin index is generalized to $N$ flavors, deconfinement arises in the case of $2N$ 2-component Dirac fermions coupled to complex $U(1)$ gauge fields for sufficiently large $N$, thus providing a counter example to confinement. Further renormalization group analysis for compact quantum electrodynamics in $2+1$D shows that deconfinement occurs when $N>N_c=36/\pi^3\simeq 1.161$, where $N$ is the number of fermion replicas. This implies that a $U(1)$ spin liquid is stable at the physical value of $N=2$ [@Nogueira05]. Moreover, by mapping the spinon Fermi surface in $2+1$D to an infinite set of (1+1)-dimensional chiral fermions, Lee [@LeeSS08] argued that an instanton has an infinite scaling dimension for any $N>0$. Therefore, the QSL phase is stable against instantons, and the noncompact $U(1)$ gauge theory is a good low-energy description. We note that mechanisms other than confinement arising from gauge fluctuations may also lead to the instability of $U(1)$ QSLs, such as Amperean pairing [@LeeLee07] and spin-triplet pairing [@Galitski07] between spinons. A non-trivial prediction of the $U(1)$ gauge theory for spin liquids is that it leads to charge excitations with a soft gap [@NgLee07], which can be detected by means of their AC conductivities $\sigma(\omega)$. It has been predicted that $\sigma(\omega)\sim\omega^{\alpha}$ in these spin liquid states, with $\alpha\sim3.33$ in a non-relativistic spin liquid and $\alpha=2$ in a Dirac fermion spin liquid [@Potter2013]. It is expected that this soft gap and the related charge fluctuations will manifest themselves most clearly when the system is near the metal-insulator transition (see the next subsection). Because charge fluctuations will manifest themselves near the metal-insulator transition, spin liquids in “weak" Mott insulators become an interesting topic [@Senthil08; @Podolsky09; @Grover10] for investigation. To study the effect of charge fluctuations near the metal-insulator transition, Lee and Lee [@Lee05] began with the Hubbard model and developed a $U(1)$ gauge theory with the help of the slave-rotor representation [@Florens04]. A number of physical phenomena, including transport properties [@Nave07] and Kondo effect [@Ribeiro11], have been studied using this framework. Charge fluctuations correspond to higher-order spin ring-exchange terms in terms of the spin Hamiltonian [@Misguich98; @Yang10]. ### Mott transition: relation between Fermi and spin liquids Zhou and Ng [@ZhouNg13] proposed a different way to understand $U(1)$ spin liquids near the Mott transition. They proposed that spin liquids near the Mott transition can be regarded as “Fermi liquids" with a constraint imposed on the current operator. For isotropic systems, they observed that the charge current carried by quasi-particles, $$\mathbf{J}=\frac{m}{m^{\ast }}(1+{\frac{F_{1}^{s}}{d}})\mathbf{J}^{(0)}, \label{Je}$$ is renormalized by the Landau parameter $F_1^s$ in Fermi liquid theory, but the thermal current, $$\mathbf{J_Q}=\frac{m}{m^{\ast}}\mathbf{J_Q}^{(0)}, \label{Jq}$$ is not, where $\mathbf{J}^{(0)}$ and $\mathbf{J_q}^{(0)}$ are the charge and thermal currents, respectively, carried by the corresponding non-interacting fermions and $d$ is the number of dimensions of the system. For systems with Galilean invariance, the charge current carried by quasi-particles is not renormalized, and $\frac{m^{\ast}}{m}=1+{\frac{F_{1}^{s}}{d}}$ [@Baym]. However, this is not valid in general for electrons in crystals, where Galilean invariance is lost. In this case, $\frac{m^{\ast }}{m}\neq1+{\frac{F_{1}^{s}}{d}}$, and the charge current carried by quasi-particles is renormalized through quasi-particle interaction. In the special case in which $1+F_{1}^{s}/d\rightarrow 0$ while $\frac{m^{\ast }}{m}$ remains finite, $\mathbf{J}\rightarrow 0$, suggesting that the fermionic system is in a special state wherein spin-$1/2$ quasi-particles do not carry charge due to interaction but still carry entropy. This is exactly what one expects for spinons in QSL states. These authors noted that the limit of $1+F_{1}^{s}/d\rightarrow 0$ is a singular point in Fermi liquid theory and that higher-order $\mathrm{q}$- and $\omega$-dependent terms should be included in the Landau interaction to ensure that finite results are obtained when calculating physical response functions. Expanding at small $\mathrm{q}$ and $\omega$, they obtained $$\frac{1+F_{1}^{s}(\mathrm{q},\omega )/d}{N(0)}\sim \alpha-\beta \omega ^{2}+\gamma_t q^{2}_t+\gamma_l q^2_l, \label{f1}$$ where $q_t\sim\nabla\times$ and $q_l\sim\nabla$ are associated with the transverse (curl) and longitudinal (gradient) parts, respectively, of the small-$\vec{q}$ expansion. In a QSL state, $\alpha=0$. They found that to ensure that the system is in an incompressible (insulator) state, it is necessary to have $\gamma_l=0$. To show that this phenomenology actually describes fermionic spin liquids with $U(1)$ gauge fluctuations, Zhou and Ng [@ZhouNg13] considered a Landau Fermi liquid with interaction parameters of $F_{0}^{s}(q)$ and $F_{1}^{s}(q)$ only. The long-wavelength and low-energy dynamics of the Fermi liquid are described by the following effective Lagrangian: $$L_{\text{eff}}=\sum_{\mathrm{k},\sigma }\left[ c_{\mathrm{k}\sigma}^{\dagger }(i\frac{\partial }{\partial t}-\xi _{\mathrm{k}})c_{\mathrm{k}% \sigma }-H^{\prime }(c^{\dag },c)\right] , \label{leff}$$ where $c_{\mathrm{k}\sigma }^{\dagger }(c_{\mathrm{k}\sigma })$ is the spin-$\sigma $ fermion creation (annihilation) operator with momentum $\mathrm{k}$ and $$H^{\prime }(c^{\dag },c)=\frac{1}{2N(0)}\sum_{q}\left[ \frac{F_{1}^{s}(q)}{v_{F}^{2}}\mathbf{j}(q)\cdot \mathbf{j}(-q)+F_{0}^{s}(q)n(q)n(-q)\right] \label{h'eff}$$ describes the current-current and density-density interactions between quasi-particles [@Leggett65; @Larkin64], where $q=(\mathrm{q},\omega )$ and $v_{F}=\hbar k_{F}/m^{\ast }$ is the Fermi velocity. The current and density interactions can be decoupled by introducing fictitious gauge potentials $\mathbf{a}$ and $\varphi $ (Hubbard-Stratonovich transformation) as follows: $$H^{\prime }(c^{\dag },c)\rightarrow \sum_{q}\left[ \mathbf{j}\cdot \mathbf{a}+n\varphi -{\frac{1}{2}}\left( \frac{n}{m^{\ast }}\frac{d}{F_{1}^{s}(q)}\mathbf{% a}^{2}+\frac{N(0)}{F_{0}^{s}(q)}\varphi ^{2}\right) \right] , \label{h'eff2}$$ where $n$ is the fermion density. The equality $d(n/m^{\ast })=N(0)v_{F}^{2}$ was used in formulating Eq. . The Lagrangian presented in Eq.  and can be rewritten in the standard form of $U(1)$ gauge theory by noting that in this representation, the fermion current is given by $$\mathbf{j}=\frac{-i}{2m^{\ast }}\sum_{\sigma }\left[ \psi _{\sigma}^{\dag }\nabla \psi _{\sigma }-(\nabla \psi _{\sigma }^{\dagger })\psi _{\sigma }% \right] -\frac{n}{m^{\ast }}\mathbf{a},$$ where $\psi _{\sigma }(\mathrm{r})=\int e^{-i\mathrm{k\cdot r}}c_{\mathrm{k}\sigma }$ is the Fourier transform of $c_{\mathrm{k}\sigma }$. The Lagrangian can be written as \[lspin\] $$L=\sum_{\sigma }\int d^{d}\mathrm{r}\left[ \psi _{\sigma }^{\dagger }(i\frac{\partial }{\partial t}-\varphi )\psi _{\sigma }-H(\psi _{\sigma }^{\dagger },\psi _{\sigma })\right] +L(\varphi ,\mathbf{a}), \label{leff2}$$where $$H(\psi _{\sigma }^{\dagger },\psi _{\sigma })=\frac{1}{2m^{\ast }}|(\nabla -i% \mathbf{a})\psi _{\sigma }|^{2} \label{heff}$$and $$L(\varphi ,\mathbf{a})={\frac{1}{2}}\int d^{d}\mathrm{r}\left[ \frac{n}{% m^{\ast }}(1+\frac{d}{F_{1}^{s}})\mathbf{a}^{2}+\frac{N(0)}{F_{0}^{s}}% \varphi ^{2}\right] .$$ Using Eq. , they find that in the small-$q$ limit, the transverse part of $L(\varphi ,\mathbf{a})$ in the spin liquid state is given by $$L_{t}(\varphi ,\mathbf{a})=-\frac{n}{2m^{\ast }}\int d^{d}\mathrm{r}\left[ \beta (\frac{\partial \mathbf{a}}{\partial t})^{2}-\gamma _{t}(\nabla \times \mathbf{a})^{2}\right] . \label{lt}$$ The Lagrangian as expressed in Eq.  together with Eq.  is the standard Lagrangian used to describe QSLs with $U(1)$ gauge fluctuations. The analysis can be rather straightforwardly generalized to a $U(1)$ spin liquid with Dirac fermion dispersion. The appearance of a soft charge gap in $U(1)$ spin liquids can be understood from the phenomenological form of $F_{1}^s(q,\omega)$  as expressed in Eq. , which suggests that the quasi-particles carry vanishing charges only in the limit of $q,\omega\rightarrow0$. The appearance of a non-vanishing $\beta$ in leads to an AC conductivity $\sigma(\omega)$ with a power-law form. This picture is very different from theories of spin liquid states that start from simple spin models in which charge fluctuations are absent at all energy scales and suggests that charge fluctuations are important in regions near the Mott transition. We note that charge fluctuations can be (partially) incorporated into the spin models through ring-exchange terms. The close relationship between Fermi liquids and spin liquid states suggests an alternative picture of the Mott metal-insulator transition with respect to that put forward by Brinkman and Rice [@BrinkmanRice], who proposed that a metal-insulator (Mott) transition is characterized by a diverging effective mass ${\frac{m^{\ast}}{m}} \to \infty $ and an inverse compressibility $\kappa \to 0$ at the Mott transition point, with a correspondingly vanishing quasi-particle renormalization weight $Z\sim {\frac{m}{m^{\ast}}}\to 0$. The diverging effective mass and vanishing quasi-particle weight imply that the Fermi liquid state is destroyed at the Mott transition and that the Mott insulator state is distinct from the Fermi liquid state on the metal side. The phenomenology described here suggests an alternative picture in which the Fermi surface is not destroyed, but the Landau quasi-particles are converted into spinons $(1+{\frac{F_{1}^{s}}{d}})\rightarrow0)$ at the Mott transition. In particular, the effective mass $m^*/m$ may not diverge at the metal-insulator transition, although $Z\rightarrow0$. A schematic phase diagram for the Mott (metal-QSL) transition is presented in Fig. \[phases\] for a generic Hubbard-type Hamiltonian with a hopping integral $t$ and an on-site Coulomb repulsion $U$. The system is driven into a Mott insulator state at zero temperature at $U=U_{c}$, where $1+F_{1}^{s}(U>U_{c})/d=0$. This picture suggests that a $U(1)$ spin liquid state is likely to exist in an insulator close to the Mott transition. ![ [@ZhouNg13] (a) Schematic zero-temperature phase diagram for the Mott transition. $% U$ is the strength of the Hubbard interaction, and $t$ is the hopping integral. The electron quasi-particle weight and the quasi-particle charge current $\sim 1+F_1^s/d$ vanish at the critical point, whereas the effective mass remains finite. (b) Schematic phase diagram showing finite-temperature crossovers and possible instability toward gapped phases at lower temperatures. There exists a (finite-temperature) critical region around $U_c$ where the phenomenology is not applicable.[]{data-label="phases"}](phases-QSLFL.eps){width="6.4cm"} The point $1+F_{1}^{s}/d=0$ is a critical point in Fermi liquid theory called the Pomeranchuk point. The Fermi surface is unstable with respect to deformation when $1+F_{1}^{s}/d<0$. The criticality of this point implies that the QSLs obtained in this way are marginally stable because of large critical fluctuations. A similar conclusion can be drawn from $U(1)$ gauge theory by analyzing the $U(1)$ gauge fluctuations. As a result, QSLs with large Fermi surfaces are, in general, susceptible to the formation of other, more stable QSLs at lower temperatures, such as $Z_{2}$ QSLs or valence-bond solid (VBS) states that gap out part of or the entire Fermi surface. This is indicated schematically in the phase diagram shown in Fig. \[phases\](b), where the system is driven into a gapped QSL at low temperatures of $T<T_{c}(U)$ on the insulating side. The nature of the low-temperature QSLs depends on the microscopic details of the system and cannot be determined based on the above phenomenological considerations. $Z_2$ spin liquid states ------------------------ An example of a $Z_2$ spin liquid state was first constructed by Wen [@Wen91] for a $J_1-J_2$ Heisenberg model on a square lattice, where $J_1$ and $J_2$ are the nearest neighbor and next nearest neighbor Heisenberg interactions, respectively. Wen considered the mean-field ansatz $$u_{i,i+\hat{\mu}}=\left( \begin{array}{cc} \chi & 0 \\ 0 & -\chi % \end{array}\right)$$ where $\hat{\mu}=\hat{x},\hat{y}$, and $$u_{i,i\pm\hat{x}+\hat{y}}=u_{i,i\mp\hat{x}-\hat{y}}=\left( \begin{array}{cc} 0 & \Delta_0\pm i\Delta_1 \\ \Delta_0\mp i\Delta_1 & 0 % \end{array}\right) ,$$ where $\chi$, $\Delta_0$ and $\Delta_1$ are nonzero real numbers; $a_0^{2,3}=0$; and $a_0^1\neq0$. It is easy to check $P(C)$ for two loops: $C_1=i\rightarrow i+\hat{x}\rightarrow i+\hat{x}+\hat{y}\rightarrow i$ and $C_2=i\rightarrow i+\hat{y}\rightarrow i+\hat{y}-\hat{x}\rightarrow i$. We obtain $$P(C_1)=\chi^2\left(\Delta_0\tau^1+\Delta_1\tau^2\right)$$ and $$P(C_2)=-\chi^2\left(\Delta_0\tau^1-\Delta_1\tau^2\right),$$ which clearly demonstrates that $\mathbf{B}(C_1)\neq\mathbf{B}(C_2)$ and that the spin liquid state described above is a $Z_2$ spin liquid state. The mean-field ground state describes a half-filled spinon band with a band dispersion given by $E_{\pm}(\mathbf{k})=\pm\sqrt{\varepsilon_1(\vec{k})^2+\varepsilon_2(\vec{k})^2+\varepsilon_3(\vec{k})^2}$, where $$\begin{aligned} \varepsilon_1(\vec{k}) & = & 2J_1\chi(\cos(k_x)+\cos(k_y)), \\ \nonumber \varepsilon_2(\vec{k}) & = & 2J_2\Delta_0(\cos(k_x+k_y)+\cos(k_x-k_y))+a_0^1, \\ \nonumber \varepsilon_3(\vec{k}) & = & 2J_2\Delta_1(\cos(k_x+k_y)-\cos(k_x-k_y)). \end{aligned}$$ Note that the spinon spectrum is fully gapped. Many other examples of $Z_2$ spin liquid states have been constructed in the literature. For instance, a nodal gapped $Z_2$ spin liquid state was proposed by Balents, Fisher and Nayak [@Balents98] and by Senthil and Fisher [@Senthil00]. The corresponding mean-field ansatz includes nearest neighbor and next nearest neighbor hopping as well as d-wave pairing on nearest neighbor bonds on the square lattice: $$u_{i,i+\hat{x}}=\left( \begin{array}{cc} \chi_1 & \Delta \\ \Delta & -\chi_1 % \end{array}\right),$$ $$u_{i,i+\hat{y}}=\left( \begin{array}{cc} \chi_1 & -\Delta \\ -\Delta & -\chi_1 % \end{array}\right),$$ and $$u_{i,i\pm\hat{x}\pm\hat{y}}=\left( \begin{array}{cc} \chi_2 & 0 \\ 0 & -\chi_2 % \end{array}\right) ,$$ where $\chi_1$, $\chi_2$, and $\Delta$ are nonzero real numbers; $a_0^{1,2}=0$; and $a_0^3\neq0$. The spinon dispersion is given by $E_{\pm}(\mathbf{k})=\pm\sqrt{\varepsilon(\vec{k})^2+\Delta(\vec{k})^2}$, where $$\begin{aligned} \varepsilon(\vec{k}) & = & 2J_1\chi_1(\cos(k_x)+\cos(k_y)) \\ \nonumber & & +2J_2\chi_2(\cos(k_x+k_y)+\cos(k_x-k_y))+a_0^3, \\ \nonumber \Delta(\vec{k}) & = & 2J_1\Delta(\cos(k_x)-\cos(k_y))+a_0^3, \end{aligned}$$ and is found to be gapless at four $\vec{k}$ points with a linear dispersion. Thus, this spin liquid is a $Z_2$ nodal spin liquid. We reiterate that $Z_2$ spin liquid states are expected to be the most stable because the $SU(2)$ gauge fields are gapped and the fermionic spins are interacting only through short-range interactions. It has been observed by Wen [@Wen91] that in addition to spinons, a soliton-type excitation exists in a $Z_2$ spin liquid. This excitation is nothing but a $\pi$ flux in the $Z_2$ gauge field, called a “$Z_2$ vortex". This $Z_2$ vortex can be described by a new mean-field ansatz, $$\tilde{u}_{ij}=u_{ij}\Theta_{ij},$$ where $\Theta_{ij}=\pm 1$ generates a $\pi$ flux on a lattice. One possible choice of $\Theta_{ij}$ is illustrated in Fig. \[Z2Vortex\], where $\Theta_{ij}=-1$ on the bonds cut by the dashed line and $\Theta_{ij}=1$ on the other bonds. An interesting consequence of such a $Z_2$ vortex is that the statistics of a spinon can be changed from bosonic to fermionic and vice versa if it is bound to a vortex. Therefore, $Z_2$ spin liquids may contain charge-neutral spin-$1/2$ spinons with both bosonic and fermionic statistics [@Ng99]. The dynamics of $Z_2$ vortices can give rise to interesting physical consequences [@Ng99; @Qi09]. ![ A $Z_2$ vortex created by flipping the signs of the $u_{ij}$ on the bonds cut by the dashed line (indicated by thick lines).[]{data-label="Z2Vortex"}](Z2Vortex.eps){width="5.4cm"} It is worth noting that the $J_1-J_2$ model on a square lattice has been well studied. The lowest-energy $Z_2$ spin liquid state is a nodal spin liquid with four Dirac points [@Capriotti2001; @Hu2013], labeled as Z2Azz13 in the projected group symmetry classification scheme [@Wen02-PSG], which we discuss in section \[sec:PSG\]. This nodal $Z_2$ spin liquid is energetically competitive with calculations performed using the DMRG [@Jiang2012; @Gong2014] and PEPS [@Wang13] approaches. [*Relation to superconductivity:*]{} RVB theory were developed not only for QSLs but also for high-T$_c$ superconductivity [@Anderson1987]. It is generally believed that $Z_2$ spin liquid states may become superconductors upon doping [@LeeRMP06]. The superconducting state inherits novel properties from its QSL parent, and new phenomena may also emerge. For instance, it has been proposed that doping a kagome system can give rise to an exotic superconductor with an $hc/4e$-quantized flux (as opposed to the usual $hc/2e$ quantization) [@Ko09]. Numerical realization of Gutzwiller projection: variational Monte Carlo method and some results {#GutzwillerP} ----------------------------------------------------------------------------------------------- The theories of QSL states rely heavily on the reliability of Gutzwiller-projected wavefunctions. In this subsection, we discuss how Gutzwiller projection is performed numerically in practice and how the physical observables can be evaluated using a Monte Carlo method for a given projected wavefunction $\left\vert \Psi _{RVB}\right\rangle =P_{G}\left\vert \Psi _{MF}\right\rangle $. Two types of mean-field ansatz are frequently used in constructing QSL states. The first one contains only (fermionic) spinon hopping terms $\chi$, and the mean-field ground state is a half-filled Fermi sea. The second one includes both hopping terms and pairing terms $\Delta$, and the mean-field ground state is a BCS-type state with a fermion energy gap. These two types of wavefunctions describe $U(1)$ and $Z_{2}$ spin liquid states, respectively, with the proper choice of hopping and pairing parameters. For a given spin Hamiltonian, we can determine these hopping and pairing parameters by optimizing the ground-state energy. Therefore, this approach is called the variational Monte Carlo (VMC) method. For a projected Fermi sea state, the mean-field ground-state wavefunction on a lattice with $N$ sites can be constructed by filling the $N$ lowest states in the mean-field band:$$\left\vert \Psi _{FS}\right\rangle =\prod_{\sigma }\prod_{k=1}^{N/2}\psi _{k\sigma }^{\dagger }\left\vert 0\right\rangle ,$$ where $\sigma =\uparrow ,\downarrow $ is the spin index and the states are sorted in order of ascending energy, $E_{1}\leq \cdots \leq E_{N/2}<E_{F}$. $\psi_{k\sigma }^{\dagger }$ creates an eigenstate in the mean-field band and can be expressed as$$\psi _{k\sigma }^{\dagger }=\sum_{i}a_{k}\left( i\right) c_{i\sigma }^{\dagger },$$ where each value of $i$ denotes a site and $c_{i\sigma }^{\dagger }$ is a local fermion creation operator. The eigenstate wavefunction $a_{k}\left( i\right) $ does not depend on the spin index $\sigma $ for spin-singlet states because of the spin rotational symmetry. More explicitly,$$\left\vert \Psi _{FS}\right\rangle =\prod_{\sigma}\prod_{i=1}^{N/2}\left( \sum_{j=1}^{N}a_{i}\left( j\right) c_{j\sigma }^{\dag }\right) \left\vert 0\right\rangle , \label{PsiFock}$$ and the Gutzwiller-projected wavefunction can be written in terms of the product of three factors:$$\begin{aligned} P_{G}\left\vert \Psi _{FS}\right\rangle &=&\sum_{\left\{ \sigma _{i}\right\} }\text{sgn}\left\{ i_{1},\cdots ,i_{N/2},j_{1},\cdots ,j_{N/2}\right\} \notag \\ &&\times \det \left[ A\left( i_{1},\cdots ,i_{N/2}\right) \right] \notag \\ &&\times \det \left[ A\left( j_{1},\cdots ,j_{N/2}\right) \right] \left\vert \sigma _{1},\cdots ,\sigma _{N}\right\rangle , \label{PgFock}\end{aligned}$$where $\left\vert \sigma _{1},\cdots ,\sigma _{N}\right\rangle $ is a state in the Ising basis with $N/2$ up spins located at sites $i_{1},\cdots ,i_{N/2}$ and $N/2$ down spins located at sites $j_{1},\cdots ,j_{N/2}$; sgn$\left\{ i_{1},\cdots ,i_{N/2},j_{1},\cdots ,j_{N/2}\right\} $ is the sign of the permutation $P=\left\{ i_{1},\cdots ,i_{N/2},j_{1},\cdots ,j_{N/2}\right\} $; and $A\left( i_{1},\cdots ,i_{N/2}\right) $ is an $N/2\times N/2$ matrix given by$$A\left( i_{1},\cdots ,i_{N/2}\right) =\left( \begin{array}{ccc} a_{1}\left( i_{1}\right) & \cdots & a_{1}\left( i_{N/2}\right) \\ \cdots & \ddots & \cdots \\ a_{N/2}\left( i_{1}\right) & \cdots & a_{N/2}\left( i_{N/2}\right)% \end{array}% \right) . \label{matrixA}$$ A BCS-type mean-field ground state with spin-singlet pairing can be written as$$\left\vert \Psi _{BCS}\right\rangle =e^{\frac{1}{2}\sum_{i,j}W_{ij}(c_{i% \uparrow }^{\dag }c_{j\downarrow }^{\dag }-c_{i\downarrow }^{\dag }c_{j\uparrow }^{\dag })}\left\vert 0\right\rangle, \label{Psipair}$$ where $i$ and $j$ are site indices and $W_{ij}=W_{ji}$ for fermionic spin-singlet pairing. For a system with lattice translational symmetry, $W_{ij}$ can be written explicitly as$$W_{ij}=-\sum_{\mathbf{k}}\frac{v_{\mathbf{k}}}{u_{\mathbf{k}}}e^{-i\mathbf{k}% \cdot (\mathbf{R}_{i}-\mathbf{R}_{j})},$$ where $u_{\mathbf{k}}$ and $v_{\mathbf{k}}$ are given in the BCS form. In the more general situation in which lattice translational symmetry is lost, the $W_{ij}$s are determined from the Bogoliubov-de Gennes equations. Gutzwiller projection retains only states with a number of electrons equal to the number of lattice sites and removes all terms with more than one electron per site, i.e., $$\left\vert \Psi _{RVB}\right\rangle =P_{G}\left( \sum\nolimits_{i<j}W_{ij}c_{i\uparrow }^{\dag }c_{j\downarrow }^{\dag }\right) ^{N/2}\left\vert 0\right\rangle . \label{Pgpair0}$$In the spin representation, the projected BCS state can be written as$$\begin{aligned} P_{G}\left\vert \Psi _{BCS}\right\rangle &=&\sum_{\left\{ \sigma _{i}\right\} }\text{sgn}\left( i_{1},\cdots ,i_{N/2},j_{1},\cdots ,j_{N/2}\right) \notag \\ &&\times \det \left[ w\left( i_{1},\cdots ,i_{N/2},j_{1},\cdots ,j_{N/2}\right) \right] \notag \\ &&\times \left\vert \sigma _{1},\cdots ,\sigma _{N}\right\rangle , \label{Pgpair}\end{aligned}$$where $\left\vert \sigma _{1},\cdots ,\sigma _{N}\right\rangle $ is a state in the Ising basis with $N/2$ up spins located at sites $i_{1},\cdots ,i_{N/2}$ and $N/2$ down spins located at sites $j_{1},\cdots ,j_{N/2}$ and $% w\left( i_{1},\cdots ,i_{N/2},j_{1},\cdots ,j_{N/2}\right) $ is an $% N/2\times N/2$ matrix given by$$w\left( i_{1},\cdots ,i_{N/2},j_{1},\cdots ,j_{N/2}\right) =\left( \begin{array}{ccc} W_{i_{1}j_{1}} & \cdots & W_{i_{1}j_{N/2}} \\ \cdots & \ddots & \cdots \\ W_{i_{N/2}j_{1}} & \cdots & W_{i_{N/2}j_{N/2}}% \end{array}% \right) . \label{matrixw}$$ A key observation regarding these two projected wavefunctions, Eqs. and , is that both of them can be written as a determinant or as a product of two determinants. This allows us to evaluate a projected wavefunction numerically. For a large system, the number of degrees of freedom increases exponentially with the system size. In this case, the Monte Carlo method is applied to evaluate the energy, magnetization and spin correlation for these projected wavefunctions [@Horsch83; @Gros89]. Below, we briefly describe how the MC method works. Those who are interested in the details may refer to Gros[@Gros89]. The expectation value of an operator $\Theta $ in a system with the spin wavefunction $|\Psi\rangle $ can be written as $$\langle \Theta \rangle =\frac{\langle \Psi |\Theta |\Psi \rangle }{\langle \Psi |\Psi \rangle }=\sum_{\alpha ,\beta }\langle \alpha |\Theta |\beta \rangle \frac{\langle \Psi |\alpha \rangle \langle \beta |\Psi \rangle }{% \langle \Psi |\Psi \rangle }, \label{MC1}$$ where the spin configurations $|\alpha \rangle $ and $|\beta\rangle $ are states in the Ising basis with $N/2$ up spins and $N/2$ down spins. This sort of expectation value is recognized to be amenable to a Monte Carlo (MC) evaluation [@Horsch83]. The expectation value expression given in Eq. (\[MC1\]) can be rewritten as $$\begin{aligned} \langle \Theta \rangle &=&\sum_{\alpha }\left( \sum_{\beta }\frac{\langle \alpha |\Theta |\beta \rangle \langle \beta |\Psi \rangle }{\langle \alpha |\Psi \rangle }\right) \frac{|\langle \alpha |\Psi \rangle |^{2}}{\langle \Psi |\Psi \rangle } \notag \\ &=&\sum_{\alpha }f(\alpha )\rho (\alpha ), \label{MC2}\end{aligned}$$with $$\begin{aligned} f(\alpha ) &=&\sum_{\beta }\frac{\langle \alpha |\Theta |\beta \rangle \langle \beta |\Psi \rangle }{\langle \alpha |\Psi \rangle }, \\ \rho (\alpha ) &=&\frac{|\langle \alpha |\Psi \rangle |^{2}}{\langle \Psi |\Psi \rangle }.\end{aligned}$$It follows that $$\rho (\alpha )\geqslant 0,\sum_{\alpha }\rho (\alpha )=1.$$Note that for a “local operator" $\Theta$ (e.g., $\Theta=\vec{S}_i\cdot\vec{S}_j$) and a given spin configuration $|\alpha \rangle$, only a limited number of “neighbor" configurations $|\beta\rangle$ give rise to a nonvanishing $\langle \alpha |\Theta|\beta \rangle$. As noted by Horsch and Kaplan [@Horsch83], the computation time for the ratio $\frac{\langle \beta |\Psi \rangle }{\langle \alpha |\Psi \rangle }$ is of $O(N^2)$. Therefore, $\langle \Theta \rangle $ can be evaluated by means of a random walk through spin configuration space with weight $\rho (\alpha )$. As in the standard MC method, the probability $T(\alpha \rightarrow \alpha ^{\prime })$ of transitioning from one configuration $\alpha $ to another configuration $% \alpha ^{\prime }$ can be chosen as follows:$$T(\alpha \rightarrow \alpha ^{\prime })=\left\{ \begin{array}{cc} 1, & \rho (\alpha ^{\prime })\ge \rho (\alpha ), \\ \frac{\rho (\alpha ^{\prime })}{\rho (\alpha )}, & \rho (\alpha ^{\prime })<\rho (\alpha ).% \end{array}% \right.$$The new configuration $\alpha ^{\prime }$ is accepted with probability $% T(\alpha \rightarrow \alpha ^{\prime })$. Because $\langle \alpha |\Psi \rangle $ is either a determinant or a product of two determinants, the computation time for $\langle\alpha |\Psi \rangle $ is of $O(N^{3})$. The computational resource consumption for the MC weight factor $T(\alpha \rightarrow \alpha ^{\prime })$ is not too high, and consequently, this MC method is feasible for Gutzwiller projection. Moreover, the computation time of the ratio $T(\alpha \rightarrow \alpha ^{\prime })$ can be reduced to $O(N^{2})$ if the corresponding matrix $A(\alpha ^{\prime })$ or $w(\alpha ^{\prime })$ in Eq.  (\[matrixA\]) or (\[matrixw\]) differs from $A(\alpha )$ or $w(\alpha )$ by only one row or column. This can be achieved by properly choosing the spin update procedure, e.g., the interchange of two opposite spins. This algorithm was first introduced by Ceperley *et al.* for the MC evaluation of a fermionic trial wavefunction [@Ceperley77]. As a variational method, the VMC method not only yields an upper bound on the ground-state energy for a spin Hamiltonian but also provides detailed information on the trial ground state. This information is useful for understanding the nature of the ground-state wavefunction. In the remainder of this subsection, we discuss some numerical results regarding Gutzwiller-projected wavefunctions on one- and two-dimensional frustrated lattices. ### One-dimensional lattice One-dimensional systems usually serve as benchmarks for comparison because exact solutions are often available. It turns out that $P_{G}\left\vert \Psi _{FS}\right\rangle $, which is gauge equivalent to $P_{G}\left\vert \Psi _{BCS}\right\rangle $ in one dimension, is an excellent trial wavefunction for the ground state of the one-dimensional Heisenberg model. The energy for $P_{G}\left\vert \Psi _{FS}\right\rangle $ is higher than that of the exact ground state by only $0.2\%$ [@Yokoyama87; @Gros87; @Gebhard87]. The spin-spin correlation decays following a power law at large distances, $\langle \vec{S}_{i}\cdot \vec{S}_{i+r}\rangle \sim \frac{(-1)^{r}}{|r|}$, consistent with the results obtained through bosonization [@Luther75]. Indeed, it has been shown that this Gutzwiller-projected wavefunction is the exact ground state of the Haldane-Shastry model [@Haldane88; @Shastry88], $$H_{H-S}=\frac{J}{2}\sum_{i=1}^{N}\sum_{r=1}^{N-1}\frac{1}{\sin ^{2}(\pi r/N)}% \vec{S}_{i}\cdot \vec{S}_{i+r},$$ which describes an AFM Heisenberg chain with long-range coupling (a periodic version of $1/r^{2}$ exchange). Excited states with $S_{z}=m=(N_{\uparrow }-N_{\downarrow })/2$ can also be constructed, where $N_{\uparrow }$ and $N_{\downarrow}$ are the numbers of up and down spins, respectively, in the wavefunction. The lowest-energy state in the subspace with $S_{z}=m$ is given by$$P_{G}\left\vert \Psi _{m}\right\rangle =P_{G}\prod_{|k|\leq k_{F\uparrow }}\psi _{k\uparrow }^{\dagger }\prod_{|k|\leq k_{F\downarrow }}\psi _{k\uparrow }^{\dagger }\left\vert 0\right\rangle , \label{PGPsim}$$where $k_{F\sigma }=\pi (N_{\sigma }-1)/N=\pi (N_{\sigma }-1)/(N_{\uparrow }+N_{\downarrow })$. With the help of this trial wavefunction, the spin susceptibility $\chi $ can be calculated [@Gros87]. It is found that $% \chi $ is close to the value obtained from the exact solution [@Griffiths64]. The numerical results are summarized in Table \[tab:Gutz1D\]. $\langle \vec{S}_{i}\cdot \vec{S}_{i+1}\rangle $ $\chi $ ------------ -------------------------------------------------- ---------------------------- Gutzwiller $-0.442118$ [@Gebhard87] $0.058\pm 0.008$ [@Gros87] Exact $-0.443147$ [@Lieb68] $0.0506$ [@Griffiths64] : [@Gros89] Comparison of ground-state energy and spin susceptibility in one dimension. The first row shows the results for the projected Fermi sea. The second row shows the results for the exact ground state of the Heisenberg model.[]{data-label="tab:Gutz1D"} ### Triangular lattice Historically, the AFM spin-$1/2$ Heisenberg Hamiltonian on a triangular lattice was the first model to be proposed for the microscopic realization of a spin liquid ground state [@Fazekas74]. However, the minimum-energy configuration for the classical Heisenberg model on a triangular lattice is well known to be the 120$^{\circ }$ Néel state. There has been a long-standing debate regarding whether the frustration together with quantum fluctuations could destroy the long-range 120$^{\circ }$ Néel order, leading to a spin liquid state. Many trial wavefunctions have been proposed as the ground state of the nearest neighbor Heisenberg model on a triangular lattice, including a chiral spin liquid state [@Kalmeyer87] and 120$^{\circ }$-Néel-order states with quantum mechanical corrections [@Huse88; @Sindzingre94]. In 1999, Capriotti *et al.* [@Capriotti99] utilized the Green’s function Monte Carlo (GFMC) method with the stochastic reconfiguration technique to obtain the state of the model with the lowest energy (to our knowledge; the ground state energy per site is $0.5458\pm 0.0001 $), which exhibits 120$^{\circ }$ long-range Néel order. More recently, the three-sublattice 120$^{\circ }$-Néel-order has been further confirmed by DMRG [@White07]. It thus seemed that for a triangular lattice, the possibility of a spin liquid state had been ruled out. However, the story continues. It was found that a four-spin ring exchange stabilizes the projected Fermi sea state against a long-range AFM state [@Motrunich05]. Because multi-spin ring exchange reflects the charge fluctuations in the vicinity of the Mott transition, this result provides theoretical support for the search for spin liquid states in a Mott insulating state close to the metal-insulator transition. The model Hamiltonian that contains both nearest neighbor Heisenberg exchange and four-spin ring exchange is $${H}_{\mathrm{ring}}=J\sum_{\begin{picture}(17,10)(-2,-2) \put (0,0) {\line (1,0) {12}} \put (0,0) {\circle*{5}} \put (12,0) {\circle*{5}} \end{picture}% }P_{12}+J_{ring}\sum_{\begin{picture}(26,15)(-2,-2) \put (0,0) {\line (1,0) {12}} \put (6,10) {\line (1,0) {12}} \put (0,0) {\line (3,5) {6}} \put (12,0) {\line (3,5) {6}} \put (6,10) {\circle*{5}} \put (18,10) {\circle*{5}} \put (0,0) {\circle*{5}} \put (12,0) {\circle*{5}} \end{picture}}\left( P_{1234}+P_{1234}^{\dagger }\right) ~, \label{Hring}$$ where $P_{12}=2\vec{S}_{1}\cdot \vec{S}_{2}+\frac{1}{2}$ interchanges the two spins at site $1$ and site $2$ and the four-spin exchange operators satisfy the following relations: $P_{1234}^{\dagger }=P_{4321}$ and $% P_{1234}+P_{4321}=P_{12}P_{34}+P_{14}P_{23}-P_{13}P_{24}+P_{13}+P_{24}-1$. ![ [@Motrunich05] Variational phase diagram for the Hamiltonian presented in (\[Hring\]).[]{data-label="ring-exchange-phases"}](ring-exchange-phases.eps){width="8.4cm"} By comparing the trial energies of the AF-ordered states proposed by Huse and Elser [@Huse88] with those of various fermionic spin liquid states, Motrunich found that the ring-exchange term favors a spin liquid ground state over the AFM-ordered state [@Motrunich05]. The results are summarized in Fig. \[ring-exchange-phases\]. For small ring exchange, i.e., $J_{ring}/J\lesssim 0.14$, the ordered states are of lower energy. However, for $J_{ring}/J\gtrsim 0.14$, spin liquid states are energetically favored. For larger values of $J_{ring}/J\gtrsim 0.3-0.35$, the optimal spin liquid state is the projected Fermi sea state. In the intermediate regime, optimized wavefunctions with extended anisotropic $s$-wave, $d_{x^{2}-y^{2}}$, and $ d_{x^{2}-y^{2}}+id_{xy}$ spinon pairings have similar energies. Recently, a novel $Z_2$ spin liquid state on a triangular lattice was proposed, where the paired fermionic spinons preserve all symmetries of the system and the system has a gapless excitation spectrum with quadratic bands that touch at $q=0$. It was shown through the VMC method that this $Z_2$ spin liquid state has a highly competitive energy when $J_{ring}/J$ is realistically large [@Mishmash13]. ### Kagome lattice Unlike the case of a triangular lattice, the classical Heisenberg model on a kagome lattice has an infinite number of degenerate ground states that are connected to one another by continuous “local" distortions of the spin configuration [@Villain80]. This property holds on any lattice with corner-sharing units, such as checkerboard, kagome, and pyrochlore lattices [@Moessner98]. For instance, on a kagome lattice formed by corner-sharing triangles, the nearest neighbor Heisenberg Hamiltonian can be written as the sum of the squares of the total spins $\vec{S% }_{\bigtriangleup }=\vec{S}_{1}+\vec{S}_{2}+\vec{S}_{3}$ of individual triangles that share only one vertex: $$H=J\sum_{\bigtriangleup }(\vec{S}_{\bigtriangleup })^{2}.$$ Classical ground states are obtained whenever $\vec{S}_{\bigtriangleup }=0$. This triangle rule fixes the relative orientations of the three classical spins of a triangle at 120$^{\circ }$ from each other in a plane, but it does not fix the relative orientation of the plane of one triad with respect to the planes of the triads on neighboring triangles. These degrees of freedom lead to a continuous local degeneracy of the ground states. Note that this degeneracy exists even if we restrict ourselves to coplanar spin states. Two of the simplest examples [@Sachdev92] are the three sublattice planar states shown in Fig. \[Kagome-classical-GS\] for the $\mathrm{q}=0$ and $\sqrt{3\times }% \sqrt{3}$ ordered states. ![ Two classical planar Néel states ($\mathrm{{q}=0}$ and $% \protect\sqrt{3}\times \protect\sqrt{3}$) on a kagome lattice. A, B and C specify three coplanar spin orientations with intersection angles of 120$^{\circ }$.[]{data-label="Kagome-classical-GS"}](Kagome.eps){width="6.4cm"} The large classical ground-state degeneracy must be lifted by quantum fluctuations. The nature of the ground state for the quantum model is highly speculative because of the enormous degeneracy in the classical model. Many arguments have been presented in the literature regarding what kind of ground state is favored, and this issue is still under debate [@BookDiep]. In the following, we discuss the $U(1)$ QSL state, which is one of the promising candidates for the ground state of a spin-$1/2$ Heisenberg antiferromagnet on a kagome lattice. Inspired by neutron scattering experiments on herbertsmithite, ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$, Ran *et al.* constructed a series of variational wavefunctions of $U(1)$ spin liquids on a kagome lattice [@Ran07]. The corresponding mean-field ansatz involves only fermionic spinon hopping on nearest neighbor bonds:$$H_{MF}=J\sum_{\langle ij\rangle \sigma }(\chi _{ij}f_{j\sigma }^{\dagger }f_{i\sigma }+h.c.),$$ where the complex field $\chi _{ij}$ lives on the links between two neighboring sites. For a kagome lattice, the mean-field states are characterized by the $U(1)$ gauge fluxes through the triangles and hexagons. Large-$N$ expansion suggests several candidate mean-field states [@Marston91; @Hastings00]: (i) VBS states, which break translation symmetry; (ii) a spin liquid state (SL-$[\frac{\pi }{2},0]$) with a flux of $+\pi /2$ through each triangle on the kagome lattice and zero flux through the hexagons, which is a chiral spin liquid state that breaks time-reversal symmetry; (iii) a spin liquid state (SL-$[\pm \frac{\pi }{2},0]$) with staggered $\pi /2$ fluxes through the triangles ($+\pi /2$ through up triangles and $-\pi /2$ through down triangles) and zero flux through the hexagons; (iv) a spin liquid state (SL-$[ \frac{\pi }{2}% ,\pi ]$) with a flux of $+\pi /2$ flux through each triangle and a flux of $\pi $ through each hexagon; (v) a uniform RVB spin liquid state (SL-$[0,0]$) with zero flux through both triangles and hexagons, which has a spinon Fermi surface; and (vi) a $U(1)$-Dirac spin liquid state (SL-$[0,\pi ]$) with zero flux through the triangles and a flux of $\pi$ through each hexagon, which has four flavors of two-component Dirac fermions. By performing VMC calculations on $8\times 8\times 3$ and $12\times 12\times 3$ lattices, Ran *et al.* [@Ran07] found that the $U(1)$-Dirac spin liquid state (SL-$[0,\pi ]$) has the lowest energy among states (i)-(vi) listed above after Gutzwiller projection, with a ground-state energy of $-0.429J$ per site. Note that there is no tunable parameter in this $U(1)$-Dirac spin liquid state. This energy is remarkably favorable because the value is very close to the exact diagonalization result when extrapolated to the thermodynamic limit. A comparison among the ground-state energies determined using this VMC method and other numerical methods is presented in Table \[tab:Kagome\]. The authors also found that the $U(1)$-Dirac spin liquid state is stable against VBS ordering and chiral spin liquid states with fluxes of $\theta $ through the triangles and $(\pi -2\theta )$ through the hexagons. The spin correlation functions exhibit algebraic decay with distance because of the Dirac nodes in the spinon spectrum. Method Energy per site ------------------------------ ------------------------ Exact diagonalization $-0.43$ [@Waldtmann98] Coupled cluster method $-0.4252$ [@Farnell01] Spin-wave variational method $-0.419$ [@Arrachea04] VMC method $-0.429$ [@Ran07] : Comparison of the ground-state energies (in units of $J$) determined using different methods for the nearest neighbor Heisenberg model on a kagome lattice. In the VMC method, the $U(1)$-Dirac spin liquid state (SL-$[0,\pi ]$) is used.[]{data-label="tab:Kagome"} We note that exact diagonalization [@Leung93; @Lecheminant97; @Waldtmann98; @Mila98] and DMRG calculations [@Jiang08; @Yan11; @Jiang12; @Depenbrock12] strongly indicate the existence of a spin gap and seem to rule out the $U(1)$-Dirac spin liquid scenario. However, this disagreement may be a finite-size effect. The applicability of exact diagonalization is limited to very small lattices of up to 36 sites, and the maximum cylinder circumference used in the DMRG approach is only 17 lattice spacings. Very recently, through the combination of the Lanczos algorithm for projected fermionic wavefunctions with the Green’s function Monte Carlo technique, Iqbal, Becca, Sorella, and Poilblanc [@Iqbal13; @Iqbal14] found that the gapless $U(1)$-Dirac spin liquid is competitive with gapped $Z_2$ spin liquids. By performing a finite-size extrapolation of the ground-state energy, these authors obtained an energy per site of $E/J = -0.4365(2)$, which is within three error bars of the estimates obtained using the DMRG method. In summary, the $U(1)$-Dirac spin liquid state has proven to be a good candidate for describing a critical phase on a kagome lattice. Classification of spin liquid states: quantum orders and projective symmetry groups {#sec:PSG} ----------------------------------------------------------------------------------- The use of Gutzwiller-projected wavefunctions can be made more systematic by using a powerful approach based on classifying spin liquid states according to their symmetry properties. For classical systems, it was observed by Landau that symmetry is a universal property shared by all macroscopic states within the same phase, irrespective of microscopic details. Consequently, the symmetry (or broken symmetry) associated with classical order parameters serves as a powerful tool for characterizing different classical phases. This approach can be generalized to quantum spin systems described by Gutzwiller-projected wavefunctions, with additional constraints. For spin liquid states described by Gutzwiller-projected wavefunctions, one might expect that the quantum phases could be classified according to the symmetry properties of the mean-field ansatz $\left( u_{ij},a_{0}^{l}\tau ^{l}\right)$. However, the usual classical symmetry group (SG) is insufficient for classifying these states for two reasons: (i) Because of the gauge redundancy, different mean-field descriptions exist for the same QSL state. For instance, the uniform RVB state and the zero-flux state correspond to the same spin state, and the $d$-wave RVB state on a square lattice is also the $\pi$-flux state. (ii) QSL states may have inherent (phase) structures contained in the mean-field ansatz $\left( u_{ij},a_{0}^{l}\tau ^{l}\right) $ that cannot be fully distinguished based on the SG constructed for classical systems. To address this issue, X.G. Wen proposed a new mathematical object called a projective symmetry group (PSG) [@Wen02-PSG], which generalizes Landau’s approach and has now become an important tool in studying QSLs and the quantum phase transitions between different QSL states. Wen proposed that the symmetry of the mean-field ansatz $\left(u_{ij},a_{0}^{l}\tau ^{l}\right) $ is a universal property and serves as a kind of “quantum number" that can be used to characterize quantum orders in QSLs. The macroscopic properties of the ansatz are characterized by its projective symmetry group (PSG). An element of a PSG is a combined operation consisting of a symmetry transformation $U$ followed by a local gauge transformation $G_{U}(i)$. The PSG of a given mean-field ansatz consists of all combined operations that leave the ansatz unchanged, i.e., $$PSG\equiv \{G_{U}|G_{U}U(u_{ij})=u_{ij},G_{U}(i)\in SU(2)\}, \label{PSG0}$$ where $U(u_{ij})=\tilde{u}_{ij}\equiv u_{U(i),U(j)}$, $G_{U}U(u_{ij})\equiv G_{U}(i)\tilde{u}_{ij}G_{U}^{\dagger }(j)$, $U$ generates the symmetry transformation (SG), and $G_{U}$ is the associated gauge transformation. From this definition, it is easy to see that $$SG\equiv {PSG\over IGG}.$$ The PSGs of two mean-field ansatze related by a gauge transformation $W$ are obviously also related. From $WG_{U}U(u_{ij})=W(u_{ij})$, where $W(u_{ij})\equiv W_{i}u_{ij}W_{j}^{\dagger } $, we obtain $WG_{U}UW^{-1}W(u_{ij})=W(u_{ij})$. Therefore, if $G_{U}U$ belongs to the PSG of the mean-field ansatz $u_{ij}$, then $WG_{U}UW^{-1}$ belongs to the PSG of the gauge-transformed ansatz $W(u_{ij})$. We see that the gauge transformation $G_{U}$ associated with the transformation $U$ changes in the following way under an $SU(2)$ gauge transformation $W$: $$G_{U}(i)\rightarrow W(i)G_{U}(i)W(U(i))^{\dagger } . \label{E-PSG}$$Wen proposed that mean-field ansatze with different PSGs belong to different classes of QSL states, just as classical states with different SGs belong to different classical phases. As examples, we consider the PSGs of the zero-flux state given in Eq.  and the $\pi$-flux state given in Eq.  on a square lattice. For illustration, let us consider the PSG associated with translational symmetry. First, we consider the zero-flux state. The mean-field ansatz given in Eq.  is invariant under the translation transformations $T_{x}(i\rightarrow i+\hat{x})$ and $T_{y}(i\rightarrow i+\hat{y})$ and the gauge transformation $G(\theta)=e^{i\theta\tau^3}$. The elements of the PSG have the form $G_UU$; $G_U=\pm G(\theta)$, and $U=(T_x)^n(T_y)^m$, where $n$ and $m$ are arbitrary integers. The $\pi$-flux state is different. The mean-field ansatz given in Eq.  breaks translational symmetry in the $x$ direction because of the odd number of lattice sites. Thus, we naively expect that the PSG should consist of elements $G_UU$ with $G_U=\pm G(\theta)$ and $U=(T_x)^{2n}(T_y)^m$. However, this is incorrect because the two mean-field ansatze $$\begin{aligned} \chi _{i,i+\hat{\mu}} &=&\left\{ \begin{array}{cc} \chi , & \mu=x \\ i\chi(-1)^{i_x}, & \mu=y \end{array}\right. \end{aligned}$$ and $$\begin{aligned} \chi _{i,i+\hat{\mu}} &=&\left\{ \begin{array}{cc} \chi , & \mu=x \\ i\chi(-1)^{i_x+1}, & \mu=y \end{array}\right. \end{aligned}$$ are actually related by a gauge transformation $W_i=(-1)^{i_y}\mathbf{\tau}^0$ and correspond to the same physical spin state. As a result, the transformations $G_{U'}U'$ with $G_{U'}=\pm G(\theta)(-1)^{i_y}\mathbf{\tau}^0$ and $U'=(T_x)^{2n+1}(T_y)^m$ are also elements of the PSG for the $\pi$-flux state. The zero-flux state and the $\pi$-flux state have different PSGs and therefore belong to different classes of $U(1)$ QSL states. More generally, other lattice symmetry operations (reflections and rotations), such as the parity transformations $P_{xy}\left( (i_{x},i_{y})\rightarrow (i_{y},i_{x})\right) $ and $P_{x\bar{y}}\left( (i_{x},i_{y})\rightarrow(-i_{y},-i_{x})\right) $ on a square lattice, the spin rotation transformation and the time-reversal transformation, are also considered when constructing PSGs, in addition to translations. The spin rotational symmetry of spin liquid states requires the mean-field ansatz to take the form: $$\begin{aligned} u_{ij} &=&i\rho _{ij}W_{ij}, \notag \\ \rho _{ij} &=&\text{real number}, \notag \\ W_{ij} &\in &SU(2). \label{spin-rotation}\end{aligned}$$ We end with a brief discussion of an issue related to techniques for the classification of PSGs. For any two given symmetry transformations, their corresponding PSG elements must satisfy certain algebraic relations determined by the symmetry transformations. Solving these equations allows us to construct a PSG of a type called an *algebraic PSG*. The name algebraic PSG is introduced to distinguish such PSGs from the invariant PSGs defined above. Any invariant PSG is an algebraic PSG; however, an algebraic PSG is not necessarily an invariant PSG unless there exists an ansatz such that the algebraic PSG is the total symmetry group of that ansatz. To provide an example, we again consider translations. The two translation elements $T_{x}$ and $T_{y}$ satisfy the following relation: $$T_{x}T_{y}T_{x}^{-1}T_{y}^{-1}=1.$$ From the definition of a PSG, we find that the two PSG elements $G_{x}T_{x} $ and $G_{y}T_{y}$ must satisfy the algebraic relation $$\begin{aligned} &&G_{x}T_{x}G_{y}T_{y}(G_{x}T_{x})^{-1}(G_{y}T_{y})^{-1} \notag \\ &=&G_{x}T_{x}G_{y}T_{y}T_{x}^{-1}G_{x}^{-1}T_{y}^{-1}G_{y}^{-1} \notag \\ &=&G_{x}\left( i\right) G_{y}\left( i-\hat{x}\right) G_{x}^{-1}\left( i-\hat{% y}\right) G_{y}^{-1}\left( i\right) \in \mathcal{G}, \label{TxTy}\end{aligned}$$ where we denote the IGG by $\mathcal{G}$. Each solution ($G_xT_x, G_yT_y$) of equation (\[TxTy\]) is an algebraic PSG for $T_{x}$ and $T_{y}$. By adding other symmetry transformations, we can find and classify all algebraic PSGs associated with a given symmetry group. Because an invariant PSG is always an algebraic PSG, we can check whether an algebraic PSG is an invariant PSG by constructing an explicit ansatz $u_{ij}$. If an algebraic PSG supports an ansatz $u_{ij}$ with no additional symmetries, then it is an invariant PSG. Through this method, we can classify symmetric spin liquids in terms of PSGs. In reference [@Wen02-PSG], Wen utilized PSGs to classify QSL states with spin rotational symmetry, time-reversal symmetry and all lattice symmetries on a square lattice. Later, the PSG classification approach for symmetric QSLs was applied to triangular [@ZhouWen02], star [@Choy09], and kagome [@Lu11] lattices. The PSG classification scheme can also be generalized to bosonic QSL states [@Wang06; @WangPSG10] and to QSL states that break spin rotational symmetry and/or time-reversal symmetry [@KouWen09; @Bieri16]. Beyond RVB approaches ===================== There are many reasons to go beyond the simple RVB approach for $S=1/2$ spin systems, for example, the discovery of a plausible spin liquid state in a spin $S=1$ system [@HDZhou2011] and the rise in interest in Mott insulators in systems with strong spin-orbit coupling where rotational symmetry is broken and the ground state cannot be a pure spin singlet[@Jackeli09]. What is the nature of the spin liquid states in these systems? More importantly, we are interested in the possibility of exotic spin liquid states beyond the RVB description, where the elementary excitations may possess exotic properties beyond the simple spinon picture. We introduce some of these developments in this section. We start by introducing the generalization of the RVB approach to spin systems with strong spin-orbit coupling and to $S>1/2$ spin systems in sections IV.1 and IV.2, followed by the introduction of matrix product states and projected entangled pair states in section IV.3, which are completely different ways of constructing spin wavefunctions compared with the RVB approach. We end this section with an introduction to the Kitaev honeycomb model, which represents yet another different approach to constructing spin wavefunctions in a system with strong spin-orbit coupling with exotic properties beyond the simple spinon picture. RVB and its generalization to spin systems with strong spin-orbit coupling -------------------------------------------------------------------------- Strong spin-orbit coupling may cause interesting experimental consequences that are absent in systems with spin rotational symmetry. An example suggested by Zhou et al. [@Zhou08] is presented here, in which strong spin-orbit coupling in Ir atoms is used to explain the anomalous behavior of the Wilson ratio observed in Na$_4$Ir$_3$O$_8$, which was experimentally proposed [@Na4Ir3O8] as the first candidate for a 3D QSL on a hyperkagome lattice with fermionic spinons. Although the Curie-Weiss constant is estimated to be as large as $\theta _{W}\sim 650$ K in Na$_4$Ir$_3$O$_8$, indicating strong AFM coupling, there is no observed thermodynamic and magnetic anomaly indicative of long-range spin ordering down to $2$ K. The specific heat ratio $\gamma=C_V/T$ shows a rather sharp peak at a temperature of $T_c\sim20$ K, indicating the existence of a phase transition or crossover at $T_c$. By contrast, the spin susceptibility $\chi\left( T\right) $ is nearly independent of temperature for all temperatures $T\ll \theta_W$. Using the experimental values of the spin susceptibility $\chi$ and the specific heat ratio $\gamma $ at the specific heat peak at $\sim 20$ K, for $T>T_c$, the Wilson ratio $R_{W}=\pi^{2}k_{B}^{2}\chi /3\mu _{B}^{2}\gamma $ of the material is 0.88, which is very close to that of a Fermi gas where $R_{W}$ is unity. Therefore, for a wide range of temperatures $T_{c}<T<\theta_W$, the system seems to behave as a Fermi liquid of spinons. Below $T_{c}$, the specific heat decreases to zero as $C_V\sim T^{2}$, suggesting a line nodal gap in the low-lying quasi-particle spectrum. However, this picture needs to be reconciled with the observation that the spin susceptibility $\chi $ remains almost constant, resulting in an anomalously large Wilson ratio of $R_W\gg 1$ at temperatures of $T<T_c$. The spins in Na$_4$Ir$_3$O$_8$ originate from the low-spin $5d^{5}$ Ir$^{4+}$ ions, which form a 3D network in the form of a corner-sharing hyperkagome lattice. Chen and Balents [@ChenBalents08] suggested that because of the large atomic number, the spin-orbit coupling in Ir atoms is expected to be strong. In the following section, we explain the anomalous Wilson ratio based on a modified RVB spin liquid picture in which both spin-singlet and spin-triplet pairings exist in the spin-pairing wavefunction. Based on the experimental observations discussed above, Zhou et al. [@Zhou08] proposed that a simple spinon hopping Hamiltonian $H_0$ determines the physics of the spin liquid state at $T>T_c$, where there exists a finite spinon Fermi surface, and that a spinon pairing gap characterized by $H_{pair}$ opens up at $T<T_c$. The power-law behavior $C_{V}\propto T^{2}$ that is observed at low temperatures of $T<T_{c}$ indicates that the gap has line nodes on the Fermi surfaces. To determine the pairing symmetry, Zhou et al. noted that a group theoretical analysis indicates that a spin-triplet pairing state on a cubic lattice can create only full or point nodal gaps [@Sigrist91], which seems to imply singlet pairing. However, because of the broken inversion symmetry on a hyperkagome lattice [@crystallography], the spin-singlet and spin-triplet pairing states are, in general, mixed together in the presence of spin-orbit coupling [@Gorkov01; @Frigeri03]. [^4] In terms of the $d$-vector, the gap function $\Delta _{\alpha \beta }(\mathbf{k})$ ($\alpha,\beta=\uparrow,\downarrow$) has the general matrix form [@Leggett75], $$\Delta (\mathbf{k})=i\left( d_{0}\left( \mathbf{k}\right) \sigma _{0}+\mathbf{d}\left( \mathbf{k}\right) \cdot \mathbf{\sigma }\right) \sigma _{y}, \label{delta}$$ and the spinon pairing must be singlet or a singlet-with-triplet admixture because of spin-orbit coupling in order to have line nodes [@Zhou08]. We now consider the spin susceptibility of such mixed states. Zhou et al. showed that if both singlet and triplet pairings are present and the spin-orbit scattering is much weaker than the pairing gap $\Delta$, then the $k$-dependent electronic contribution to the spin susceptibility is given by $${\chi _{ii}(\mathbf{k})\over \chi_{N}(\mathbf{k})}= 1-\frac{d_{0}d_{0}^{\ast }+d_{i}^{\ast }d_{i}}{d_{0}d_{0}^{\ast } +\mathbf{d\cdot d}^{\ast }}+ \frac{d_{0}d_{0}^{\ast}+d_{i}^{\ast }d_{i}}{d_{0}d_{0}^{\ast }+\mathbf{d\cdot d}^{\ast }}Y(\mathbf{k};T),$$ where $i=x,y,z$; $\chi_{N}$ is the normal state contribution at $\Delta=0$; and $Y(\mathbf{k};T)$ is the $k$-dependent Yosida function [@Leggett75]. Under the assumption that the $d$-vector is pinned by the lattice, for a polycrystalline sample, one must average over all spatial directions, resulting in $$\label{chip} {\frac{\chi _{s}}{\chi _{N}}}=\frac{2}{3}-\frac{2}{3}\frac{\left\vert d_{0}\right\vert ^{2}} {\left\vert d_{0}\right\vert^{2}+\left\vert \mathbf{d}\right\vert ^{2}}+(\frac{1}{3}+\frac{2}{3}\frac{ \left\vert d_{0}\right\vert ^{2}}{\left\vert d_{0}\right\vert^{2}+\left\vert \mathbf{d}\right\vert ^{2}})Y(T),$$ where $Y\left( T\right)$ is the (spatially averaged) Yosida function, which vanishes at zero temperature; $\chi _{s}$ is the spin susceptibility below $T_{c}$; and $\chi _{N}$ is the Pauli spin susceptibility in the normal state. Therefore, $\chi _{s}/\chi _{N}$ reduces to $\frac{2}{3}-\frac{2}{3}\frac{\left\vert d_{0}\right\vert ^{2}}{\left\vert d_{0}\right\vert^{2}+ \left\vert \mathbf{d}\right\vert ^{2}}$ at zero temperature. If the spin-triplet pairing dominates, then $\chi _{s}/\chi_{N}\rightarrow \frac{2}{3}$, whereas if the spin-singlet pairing dominates, then $\chi _{s}/\chi _{N}\rightarrow 0$. However, neither of these cases is observed in experiments; instead, $\chi$ changes only negligibly below $T_c$ [@Na4Ir3O8]. This suggests that strong spin-orbit coupling is needed to explain the absence of a marked change in $\chi$ below $T_c\sim 20$ K. It is well known that in conventional BCS singlet superconductors, the Knight shift, which is proportional to the Pauli paramagnetic susceptibility, changes very little below $T_c$ for heavy elements such as Sn and Hg [@Adroes59]. It is understood that this is caused by the destruction of spin conservation due to the spin-orbit coupling. A clear explanation was presented by Anderson [@Anderson59] using the notion of time-reversed pairing states. We first consider the imaginary part of the spin response function $\chi^{\prime\prime} (q, \omega)$. If the total spin is conserved, then the dynamics are diffusive and $\chi^{\prime\prime} (q, \omega)$ will have a central peak in $\omega$ space with a width of $Dq^2$, which goes to zero as $q \rightarrow 0$. Superconductivity gaps out all low-frequency excitations, thus removing this central peak. By the Kramers-Kronig relation, the real part $\chi^\prime (q=0, \omega=0)$ vanishes in the superconducting ground state. In the presence of spin-orbit coupling, the total spin is not conserved but rather decays with a lifetime $\tau_s$. In this case, $\chi^{\prime\prime}(q=0,\omega)$ has a central peak with a width of ${1\over \tau_s}$. The superconducting gap (formed by a pair of time-reversal states) $\Delta$ cuts a hole in $\chi^{\prime\prime}(\omega)$ for $\omega < \Delta$ but leaves the $\omega \gg \Delta$ region intact, consistent with the physical expectation that the high-frequency region should be unaffected by pairing. By the Kramers-Kronig relation, $\chi^\prime$ will be reduced, but if the spin-orbit coupling is sufficiently strong that $${1\over \tau_s} \gg \Delta, \label{delta-tau}$$ then the reduction will be small, i.e., $${\chi_s \over \chi_N} = 1-{\cal O} (\Delta \tau_s) .$$ Eq. (\[delta-tau\]) is the strong spin-orbit coupling condition that is required to have very little change in the spin susceptibility below $T_c$. We emphasize that the criterion for discriminating strong from weak spin-orbit coupling that is given by Eq.  (\[delta-tau\]) is completely different from the usual criterion, which compares the spin-orbit energy, $\lambda$, with the splitting of the $t_{2g}$ levels, $E_3$ [@ChenBalents08]. Another way to explain the large Wilson ratio observed in Na$_4$Ir$_3$O$_8$ was provided by Chen and Kim [@ChenKim13], in which strong spin-orbit coupling is still essential. From a theoretical perspective, the PSG classification scheme has been applied to classify the spin liquid states on a kagome lattice with the Dzyaloshinskii-Moriya (DM) interaction [@Dodds13]. More recently, to test the validity of the RVB approach in constructing wavefunctions for spin systems with strong spin-orbit coupling, Sze, Zhou and Ng [@Sze2016] applied the Gutzwiller-projected wavefunction of fermion pairing states to study the $S=1/2$ anisotropic Heisenberg (XXZ) chain $$\label{xxz} H=J_z\sum_{i}S^z_iS^z_{i+1}+J_{\perp}\sum_i\left(S^x_iS^x_{i+1}+S^y_iS^y_{i+1}\right),$$ where $J_{\perp},J_z>0$. This model can be mapped to the isotropic (XXX) Heisenberg model with the Dzyaloshinskii-Moriya (DM) interaction, $$\sum_i\mathbf{D}\cdot(\mathbf{S}_i\times\mathbf{S}_{i+1}),$$ in one dimension with open boundary conditions through the transformation $U=\exp(-i\sum_{n}{n\theta\over 2}S_n^z)$ with $\cos\theta=J_z/J_{\perp}$ and $D = J_{\perp}\sin\theta$, where $U^{\dagger}H_{XXZ}U=H_{J+DM}$, with $H_J$ denoting the isotropic Heisenberg model with interaction $J$. Trial mean-field wavefunctions with the general pairing $$\Delta (\mathbf{k})=i\left( d_{0}\left( \mathbf{k}\right) \sigma _{0}+\mathbf{d}\left( \mathbf{k}\right) \cdot \mathbf{\sigma }\right) \sigma _{y}$$ are being considered for the construction of the corresponding Gutzwiller-projected wavefunctions. The trial ground-state wavefunctions have the best energy when the $d$-vector has the form $d_0=0$ and $\mathbf{d}(k)=d_z\hat{z}=i\Delta\sin k$ for $J_z>J_{\perp}$ (Ising regime), whereas the preferred form is $d_0=0$ and $\mathbf{d}(k)=d_y\hat{y}=\Delta\sin k$ for $J_z<J_{\perp}$ (planar regime). The overlap between the trial ground-state wavefunction and the exact ground-state wavefunction obtained through exact diagonalization is better than $95\%$ in all cases that have been considered. Notably, the pairing state with $\mathbf{d}(k)=d_y\hat{y}=\Delta\sin k$ does not conserve $S_z^{tot}$ and is not considered in the classification scheme used in reference [@Dodds13]. RVB approach to $S>1/2$ systems ------------------------------- Historically, the search for spin liquid states has been focused on spin $1/2$ systems because such systems have the strongest quantum mechanical fluctuation effects (see section II) when the unfrustrated Heisenberg model is considered. The situation is different when we consider spin systems with frustrated interactions [@Chandra88]. In this case, it is not obvious whether a spin liquid state is more likely to exist in systems of lower spin. In fact, it has recently been found that gapless spin liquid states may exist in a two-dimensional spin-$1$ compound Ba$_3$NiSb$_2$O$_9$ under high pressure [@Cheng12]. In this subsection, we examine how we can construct spin liquid states for $S>1/2$ systems by generalizing the RVB approach developed for $S=1/2$ systems. It should be noted that there are multiple possible methods of generalization. For example, Greiter and Thomale [@Greiter09] constructed a chiral spin liquid state using a fractional quantum Hall wavefunction, whereas Xu [*et al.*]{} [@Xu12] constructed a spin liquid state for an $S=1$ system by representing a spin of $1$ as the sum of two $S=1/2$ spins. Liu, Zhou, and Ng [@LZN10a; @LZN10b] have developed an alternative approach in which a spin $S$ is represented by $2S+1$ fermions. In the following section, we consider this last approach, and we demonstrate the existence of fundamental differences between half-odd-integer spin and integer spin systems in this approach. We begin with the fermion representation of general spins. To generalize the fermion representation of $S=1/2$ spins to an arbitrary spin $S$, Liu, Zhou and Ng [@LZN10a; @LZN10b] introduce $2S+1$ species of fermionic operators $c_m$ that satisfy anti-commutation relations, $$\begin{aligned} \label{Fermi} \{c_m,c^\dagger_n\}=\delta_{mn}, \end{aligned}$$ where $m,n=S,S-1,\cdots,-S$. The spin operator can be expressed in terms of these operators as follows: $$\begin{aligned} \hat {\mathbf S}=C^\dagger {\mathbf I}C, \end{aligned}$$ where $C=(c_S, c_{S-1},\cdots,c_{-S})^T$ and $I^a$ $(a=x,y,z)$ is a $(2S+1)\times(2S+1)$ matrix whose matrix elements are given by $$I^a_{mn}=\langle S,m|S^a|S,n\rangle.$$ It is straightforward to show that the resulting spin operator $\hat{\mathbf{S}}$ satisfies the $SU(2)$ angular momentum algebra. Under a rotational operation, $C$ is a spin-$S$ “spinor" transforming as $C_m\to D^{S}_{mn}C_{n}$ and $\hat{\mathbf S}$ is a vector transforming as $S^a\to R_{ab}S^{b}$; here, $D^S$ is the $2S+1$-dimensional irreducible representation of the $SU(2)$ group generated by $\mathbf I$, and $R$ is the adjoint representation. As in the $S=1/2$ case, a constraint that there must be only one fermion per site is needed to project the fermionic system into the proper Hilbert space representing spins, i.e., $$\label{contr1} (\hat N_i-N_f)|\mathrm{phy}\rangle=0,$$ where $i$ is the site index and $N_f=1$ (the particle picture, one fermion per site). Alternatively, it is straightforward to show that the constraint $N_f=2S$ (the hole picture, one hole per site) equivalently represents a spin. The $N_f=1$ representation can be mapped to the $N_f=2S$ representation [*via*]{} a [*particle-hole transformation*]{}. For $S=1/2$, the particle picture and the hole picture are identical, reflecting an intrinsic particle-hole symmetry of the underlying Hilbert space, which is absent for $S\ge1$. Following Affleck, Zou, Hsu and Anderson [@AZHA88], Liu, Zhou and Ng [@LZN10b] introduce another “spinor" $\bar C=(c^\dagger_{-S}, -c^\dagger_{-S+1}, c^\dagger_{-S+2},\cdots,(-1)^{2S}c^\dagger_{S})^T$, whose components can be written as $\bar C_m=(-1)^{S-m}c^\dagger_{-m}$, where the index $m$ runs from $S$ to $-S$ as for $C$. Upon combining $C$ and $\bar C$ into a $(2S+1)\times2$ matrix $\psi=(C,\bar C)$, it is straightforward to see that the spin operators can be re-expressed as $$\begin{aligned} \label{spin} \hat {\mathbf S}=\frac{1}{2}\mathrm{Tr}(\psi^\dagger {\mathbf I}\psi)\end{aligned}$$ and that the constraint can be expressed as $$\begin{aligned} \label{contr2} \label{newcontr1}\mathrm{Tr}(\psi\sigma_z\psi^\dagger)=2S+1-2N_f=\pm(2S-1),\end{aligned}$$ where the $+$ sign implies $N_f=1$ and the $-$ sign implies $N_f=2S$. We now examine the internal symmetry group associated with the redundancy in the fermion representation. The internal symmetry group is different for integer and half-odd-integer spins; it is $U(1)\bar\otimes Z_2=\{e^{i\sigma_z\theta}, \sigma_x e^{i\sigma_z\theta}=e^{-i\sigma_z\theta}\sigma_x;\theta\in \mathbb{R}\}$ for the former and $SU(2)$ for the latter. The reason for this difference can be qualitatively understood as follows: Note that $C$ and $\bar C$ are not independent. The operators in the internal symmetry group “mix" the two fermion operators in the same row of $C$ and $\bar C$, i.e., $c_{S}$ and $c_{-S}^\dagger$. For integer spins, $c_0$ and $(-1)^Sc_0^\dagger$ will be “mixed". For the relation $\{c_0,c_0^\dagger\}=1$ to remain invariant, there are only two possible methods of “mixing": one is a $U(1)$ transformation, and the other is interchanging the two operators. These operations form the $U(1)\bar\otimes Z_2$ group. For half-odd-integer spins, the pair $(c_0,(-1)^Sc_0^\dagger)$ does not exist, and the symmetry group is the maximum $SU(2)$ group. Thus, the difference between integer and half-odd-integer spins is a fundamental property of the fermion representation. Now let us see how the constraint expressed in Eq.  transforms under the symmetry groups. For $S=1/2$, constraint given in Eq.  is invariant under the transformation $\psi\to\psi W$ because the right-hand side vanishes (as a result of the particle-hole symmetry of the Hilbert space). For integer spins, if $W=e^{i\sigma_z\theta}$, then $W\sigma_z W^\dagger=\sigma_z$, and Eq.  is invariant. If $W=\sigma_xe^{i\sigma_z\theta}$, then $W\sigma_z W^\dagger=-\sigma_z$, meaning that the “particle" picture ($+$ sign in Eq. ) and the “hole" picture ($-$ sign in Eq. ) are transformed into each other. For a half-odd-integer spin with $S\ge 3/2$, $W\in SU(2)$ is a rotation, and we may extend the constraint into a vector form in a manner similar to the $S=1/2$ case, such that Eq.  becomes $$\begin{aligned} \label{veccontr} \mathrm{Tr}(\psi\vec\sigma\psi^\dagger)=(0,0,\pm(2S-1))^T. \end{aligned}$$ Under the group transformation $\psi\to\psi W$, $$\begin{aligned} \label{veccontr2} \mathrm{Tr}(\psi\vec\sigma\psi^\dagger)\rightarrow(R^{-1})(0,0,\pm(2S-1))^T, \end{aligned}$$ where $W\sigma^aW^\dagger=R_{ab}\sigma^b$, $a,b=x,y,z$, i.e., $R$ is a 3 by 3 matrix representing a 3D rotation. The transformed constraint represents a new Hilbert subspace, which is still a $(2N+1)$-dimensional irreducible representation of the spin $SU(2)$ algebra. Any measurable physical quantity, such as the spin $\mathbf S$, remains unchanged in this new Hilbert space. Therefore, for half-odd-integer spins ($S\ge 3/2$), there exist infinitely many ways of imposing the constraint that gives rise to a Hilbert subspace representing a spin. However, for integer spins, there exist only two possible constraint representations. The fermion representation can be used to construct mean-field Hamiltonians for spin models with arbitrary spins after the spin-spin interaction is written down in terms of fermion operators. For the spin-$1/2$ case, the Heisenberg interaction can be written as (see section III) $$\begin{aligned} \label{S1/2} \hat{\mathbf{S}}_i\cdot\hat {\mathbf{S}}_j&=&-\frac{1}{8} \mathrm{Tr}:(\psi_i^\dagger\psi_j \psi_j^\dagger\psi_i):\nonumber \\&=&-\frac{1}{4}:(\chi_{ij}^\dagger\chi_{ij}+\Delta_{ij}^\dagger\Delta_{ij}):, \end{aligned}$$ where $$\chi_{ij}=C_i^\dagger C_j, ~\quad \Delta_{ij}=\bar C_i^\dagger C_j.$$ The definitions of $\chi_{ij}$ and $\Delta_{ij}$ in the above form can be extended to arbitrary spins. The only difference is that for an integer spin, $\chi_{ji}=\chi_{ij}^\dagger$ and $\Delta_{ji}=-\Delta_{ij}$, whereas for a half-odd-integer spin, $\chi_{ji}=\chi_{ij}^\dagger$ and $\Delta_{ji}=\Delta_{ij}$. The parity of the pairing term $\Delta_{ij}$ differs for integer and half-odd-integer spins [@LZN10b]. For $S=1$, it can be shown, after some straightforward algebra, that the Hamiltonian can be written as [@LZN10b] $$\begin{aligned} \label{H1} \hat{\mathbf{S}}_i\cdot\hat{\mathbf{S}}_j&=&-\frac{1}{2}\mathrm{Tr}:(\psi_i^\dagger\psi_j\psi_j^\dagger\psi_i):\nonumber\\ &=&-:(\chi_{ij}^\dagger\chi_{ij} +\Delta_{ij}^\dagger\Delta_{ij}):.\end{aligned}$$ However, for $S>1$, we cannot write the spin-spin interaction $\hat{\mathbf{S}}_i\cdot\hat{\mathbf{S}}_j$ in terms of $\chi_{ij}$ and $\Delta_{ij}$ alone. In the case of $S=3/2$, triplet hopping and pairing terms must be introduced to represent the Heisenberg interaction. Generally speaking, quintet and higher multipolar hopping and pairing operators are needed to represent the Heisenberg Hamiltonian when $S$ becomes larger [@LZN10b]. In the following, we restrict ourselves to $S=1$ systems. In this case, the mean-field Hamiltonians are BCS-type Hamiltonians, as in the case of $S=1/2$ spins. The physical spin wavefunction can be obtained by applying Gutzwiller projection to the mean-field ground state. There are two major differences between $S=1$ and $S=1/2$ spin systems: (1) Because of the different internal symmetry group ($U(1)\bar\otimes Z_2$), $S=1$ spin liquid states are of either the $U(1)$ or $Z_2$ type. There are no $SU(2)$ spin liquid states for integer spin systems in the fermionic construction. Therefore, we expect that in general, spin liquid states for integer spin systems, if they exist, are more stable against gauge fluctuations. (2) The difference in parity of the pairing terms leads to interesting possibilities for obtaining topological spin liquid states in $S=1$ systems that are not easy to realize in $S=1/2$ systems [@LZN10a; @Bieri12]. This difference leads to the existence of a Haldane phase in the bilinear-biquadratic Heisenberg spin chain in the fermionic description [@LZN12]. Finally, we note the existence of a fundamental difference in the excitation spectrum of an $S=1$ spin system compared with that of an $S=1/2$ system, under the assumption that the ground states are spin singlets. For an integer spin system, we can form spin-singlet states in a lattice with either an even or an odd number of lattice sites $N$, as long as $N>1$, whereas for a half-odd-integer spin system, spin-singlet states can be formed only in a lattice with an even number of sites. In the RVB approach, angular momentum $L=1$ excitations of the system are formed by Gutzwiller projecting the excited states in BCS theory, i.e., by breaking a pair of spin singlets in the BCS ground state. The resulting excited state consists of two excited spinons, which are $S=1/2$ objects for spin $1/2$ systems but are $S=1$ objects for spin $1$ systems. In an $S=1$ spin liquid, these two $S=1$ spinons together form an $L=1$ excitation. There is, however another method of forming an $L=1$ excitation in a spin-$1$ spin liquid. Beginning from a lattice system with $N$ sites, we may form an $L=1$ excitation by rearranging the spins such that the system is a product of spin-singlet ground states for $N-1$ of the sites plus a single spin-$1$ spinon. This excitation is a non-perturbative, topological excitation that cannot be achieved by simply Gutzwiller projecting a BCS excited state in the RVB construction. It has been demonstrated in reference [@LZN14] that the construction of these two kinds of excitations gives rise to the so-called one-magnon and two-magnon excitation spectra in the Haldane phase of the $S=1$ bilinear-biquadratic Heisenberg model. Similar construction approaches are not possible for $S=1/2$ systems. Matrix product state (MPS) and projected entangled pair state (PEPS) -------------------------------------------------------------------- In this subsection, we discuss two approaches to spin liquid states that have completely different starting points from those of the RVB, or Gutzwiller-projected mean-field theory, approach we discussed in section III. We begin with matrix product states (MPSs) and projected entangled pair states (PEPSs), which represent another popular class of variational wavefunctions that are currently being applied to spin systems. Translationally invariant MPSs in spin chains were first constructed and studied by Fannes, Nachtergaele and Werner [@Fannes92] as an extension of the AKLT state [@AKLT]; in this context, the authors called them [*finitely correlated states*]{}. The term MPS was coined by Klümper, Schadschneider and Zittartz [@Klumper93], who extended the AKLT state in a different way. Later, Östlund and Rommer [@Ostlund95] realized that the state resulting from DMRG [@White92] can be written as an MPS. This approach is very successful for one-dimensional systems and can be generalized to systems of two (or more) dimensions. First, let us consider the quantum wavefunction of a one-dimensional spin system that is translationally invariant with a local Hamiltonian $H$. The wavefunction can be generally expressed as $$\label{mps1} |\Psi\rangle=\sum_{s_1,s_2,\cdots,S_N}\phi(s_1,s_2,\cdots,s_N)|s_1,s_2,\cdots,s_N\rangle,$$ where $|s_1,s_2,\cdots,s_N\rangle$ represents a spin configuration with spins $s_i$ on sites $i=1,2,\cdots,N$ and $\phi(s_1,s_2,\cdots,s_N)$ is the amplitude of the spin configuration in the quantum state $|\Psi\rangle$. Because of the spin-spin interaction, spin configurations at far away sites are generally correlated, and we cannot write $\phi(s_1,s_2,\cdots,s_N)=\phi_0(s_1)\phi_0(s_2)\cdots\phi_0(s_N)$ in general. The MPS approach is a powerful method of constructing wavefunctions with non-local quantum correlations. The trick is to extend the direct-product wavefunction $\phi(s_1,s_2,\cdots,s_N)=\phi_0(s_1)\phi_0(s_2)\cdots\phi_0(s_N)$ to matrix products. More explicitly, we associate a matrix $A^{s}$ with each spin state $s$; then, the wavefunction amplitude $\phi(s_1,s_2,\cdots,s_N)$ can be written as $$\label{mps2} \phi(s_1,s_2,\cdots,s_N)={\rm Tr}\{A^{s_1}[1]A^{s_2}[2]\cdots A^{s_N}[N]\},$$ where the trace is used to impose the periodic boundary condition. As an example, we consider an $S=1/2$ two-spin system and choose $A^{\uparrow}=\sigma_z$ and $A^{\downarrow}=\sigma_x$, where the $\sigma$s are Pauli matrices. It is easy to see that in this case, $\phi(\uparrow,\uparrow)=\phi(\downarrow,\downarrow)\neq0$ and $\phi(\uparrow,\downarrow)=\phi(\downarrow,\uparrow)=0$. A different choice of $A^{\uparrow}=\sigma_+$ and $A^{\downarrow}=\sigma_-$ yields $\phi(\uparrow,\downarrow)=\phi(\downarrow,\uparrow)\neq0$ and $\phi(\uparrow,\uparrow)=\phi(\downarrow,\downarrow)=0$. The correlation between the different spin states on the two sites is determined by the matrix $A^s$ that is chosen to link the sites. Extending the construction to more than two sites, one sees that the choice of the matrices $A^{\sigma}$ determines the quantum entanglement structure of the wavefunction. When the MPSs are treated as variational wavefunctions, one may determine the number of variational parameters in the wavefunctions by means of a simple counting argument. The number of parameters $P$ appearing in an MPS wavefunction in the form of Eq.  depends on the size of the matrix $A$ and the number of available states $S$ per site. In general, $P\sim S\times M^2$ for an $M\times M$ matrix as long as $P<S^N$, where $N$ is the number of sites in the system. Thus, MPS wavefunctions are generally variational wavefunctions with a large number of built-in variational parameters. As the dimension $M\rightarrow\infty$, MPSs can represent any quantum state of the many-body Hilbert space with arbitrary accuracy. In practice, the low-energy states of gapped local Hamiltonians in one dimension can be efficiently represented by MPSs with a finite value of $M$ [@Hastings07; @Verstraete06b]. The DMRG method [@White92] and its generalizations [@Schollwock05] can be viewed as systematic approaches for constructing MPS variational wavefunctions as the size of the system gradually increases. The MPS construction can be extended in several ways. First, it can be extended to higher dimensions by replacing the matrices $A$ (= rank 2 tensors) with higher-rank tensors $T$. These wavefunctions are presently known as projected entangled pair states (PEPSs) [@VerstraeteCirac04a; @VerstraeteCirac04b]. Second, the local correlation or entanglement between a pair of sites in a PEPS can be generalized to a cluster (or simplex), resulting in states called projected entangled simplex states (PESSs) [@Xie14]. A representative example of a PESS is the simplex solid state proposed by Arovas [@Arovas08]. ### Valence-bond solids and MPSs in one dimension The physics of an MPS or PEPS wavefunction is encoded in the tensors linking neighboring spin states. In general, these link tensors can be optimally constructed using the DMRG approach or tensor-based renormalization methods [@CiracVerstraeteJPA09]. In this subsection, we discuss a simple example of tensors that represent a particular class of spin states called valence-bond solid (VBS) states. To begin, we introduce a well-known example of a valence-bond solid state - the Affleck-Kennedy-Lieb-Tasaki (AKLT) state [@AKLT]. The AKLT state is an example of a VBS state in which only one spin-singlet configuration is allowed in the wavefunction given in Eq. . It is a one-dimensional VBS state constructed for a $S=1$ spin chain, represented pictorially in Fig. \[fig:AKLT\], where each gray bond represents a spin singlet formed by two $S=1/2$ spins, i.e., Eq. . Each lattice site is connected to two other sites by two valence bonds and is occupied by two $S=1/2$ spins. The AKLT wavefunction is formed by projecting the spin-$1/2 \bigotimes 1/2=1 \bigoplus 0$ quartet states into the spin $S=1$ triplet states. This is represented graphically in Fig. \[fig:AKLT\] by the circles, which represent projection operators tying together two $S=1/2$ spins, projecting out the spin $S=0$ or singlet state and preserving only the spin $S=1$ or triplet states. ![ A valence-bond solid construction of the AKLT state.[]{data-label="fig:AKLT"}](AKLT.eps){width="8.0cm"} For every adjacent pair of $S=1$ spins, two of the four constituent $S=1/2$ spins are projected into a state with a total spin of zero by the valence bond. Therefore, the pair of $S=1$ spins is forbidden from existing in a combined spin $S=2$ state. This condition can be realized by considering a Hamiltonian that is a sum of projectors $P_{i,i+1}$ that project the pairs of $S=1$ spins from the $1\bigotimes 1=2\bigoplus 1 \bigoplus 0$ space into the spin $S=2$ subspace, $$\label{haklt} H_{\rm AKLT}=\sum_iP_{i,i+1}.$$ Because the projection operators $P_{i,i+1}$ are positive semi-definite, the ground state satisfies $H_{AKLT}|\Psi_G\rangle=0$ and is simply the AKLT state. The projection operator $P_{i,i+1}$ can be written in terms of spin-$1$ operators as follows [@AKLT]: $$\label{paklt} P_{i,i+1}={1\over3}+{1\over2}(\mathbf{S}_i\cdot\mathbf{S}_{i+1})+{1\over6}(\mathbf{S}_i\cdot\mathbf{S}_{i+1})^2.$$ The AKLT state is important because it is an explicit spin wavefunction that realizes the Haldane phase for integer spins (see section II). In particular, it is easy to see from Fig. \[fig:AKLT\] that an unpaired $S=1/2$ spin will be left at each end of the spin chain, which is a realization of the end state discussed in section II for $S=1$ Heisenberg spin chains. In the following, we show how the AKLT state can be written as an MPS state. The AKLT state can be constructed in two steps. First, we split each site $i$ in the spin-$1$ chain into two sites $i_L$ and $i_R$, thereby forming a spin-$1/2$ chain with $2N$ sites, as in Fig. \[fig:AKLT\] (where $N$ is the number of sites in the parent spin-$1$ chain) and construct a dimerized chain in which the spins at sites $i_R$ and ${i+1}_L$ ($i=1,2,\cdots,N$) are joined by a valence bond (see Eq. ). The singlet bond between sites $i_R$ and ${i+1}_L$ can be written as $$\label{mps5} (i,i+1)=\sum_{\sigma_{i_R},\sigma_{{i+1}_L}}R_{\sigma_{i_R},\sigma_{{i+1}_L}}|\sigma_{i_R}\rangle|\sigma_{{i+1}_{L}}\rangle,$$ where $\sigma=\uparrow,\downarrow$ and the $R_{\sigma\sigma'}$ are the components of a $2\times2$ matrix: $$\mathbf{R} =\left( \begin{array}{cc} 0 & {1\over\sqrt{2}} \\ -{1\over\sqrt{2}} & 0 \end{array}\right). \label{R}$$ In this representation, the wavefunction of the dimerized spin-$1/2$ chain can be written as $$\label{mps6} |\Psi\rangle=\sum_{\sigma_{1_R},\cdots,\sigma_{N_L}}R_{\sigma_{1_R}\sigma_{2_L}}\cdots R_{\sigma_{{N-1}_R}\sigma_{N_L}} |\sigma_{1_R},\cdots,\sigma_{N_L}\rangle.$$ Note that this state is a direct product state of $S=1/2$ RVB singlet pairs with the two end spins ($\sigma_{1_L}$ and $\sigma_{N_R}$) unspecified. Next, we project the two $S=1/2$ spins at sites $i_L$ and $i_R$ to the spin-$1$ states $|1,m\rangle$ ($m=0,\pm1$) with $$\begin{aligned} \label{pro1} |1,1\rangle & = & |\uparrow\uparrow\rangle, \\ \nonumber |1,0\rangle & = & {1\over\sqrt{2}}\left(|\uparrow\downarrow\rangle+|\downarrow\uparrow\rangle\right), \\ \nonumber |1,-1\rangle & = & |\downarrow\downarrow\rangle. \end{aligned}$$ This projection can be expressed in terms of three matrices, $\mathbf{M}^{0,\pm1}$, where $$\label{pro1m} |1,m\rangle=\sum_{\sigma,\sigma'}M^m_{\sigma\sigma'}|\sigma\rangle|\sigma'\rangle$$ with $$\mathbf{M}^1 =\left( \begin{array}{cc} 1 & 0 \\ 0 & 0 \end{array}\right), \label{m1}$$ $$\mathbf{M}^{-1} =\left( \begin{array}{cc} 0 & 0 \\ 0 & 1 \end{array}\right), \label{m-1}$$ and $$\mathbf{M}^0 =\left( \begin{array}{cc} 0 & {1\over\sqrt{2}} \\ {1\over\sqrt{2}} & 0 \end{array}\right). \label{m0}$$ Thus, the AKLT state can be written as $$\label{mpaklt} |\Psi_{\rm AKLT}\rangle=\sum_{s_1,s_2,\cdots,s_N}\phi_{\rm AKLT}(s_1,\cdots,s_N)|s_1,s_2,\cdots,s_N\rangle,$$ where $s_i=0,\pm1$ and $$\begin{aligned} \phi_{\rm AKLT}(s_1,\cdots,s_N) & = & \sum_{\sigma_{1_R},\cdots,\sigma_{N_L}} [M^{s_1}_{\sigma_{1_L}\sigma_{1_R}} R_{\sigma_{1_R}\sigma_{2_L}} \nonumber \\ & & \times M^{s_2}_{\sigma_{2_L}\sigma_{2_R}}\cdots R_{\sigma_{{N-1}_R} \sigma_{N_L}} ] \nonumber \\ & = & [\mathbf{A}^{s_1}\mathbf{A}^{s_2}\cdots\mathbf{A}^{s_N}]_{\sigma_{1_L}\sigma_{N_R}}. \label{mps7} \end{aligned}$$ Here, $\mathbf{A}^{s}=\mathbf{M}^s\mathbf{R}$, and $\sigma_{1_L},\sigma_{N_R}=\uparrow,\downarrow$ correspond to four degenerate ground states on an open chain. Imposing the periodic boundary condition gives rise to a non-degenerate ground state with $$\phi_{\rm AKLT}(s_1,\cdots,s_N) = {\rm Tr} [\mathbf{A}^{s_1}\mathbf{A}^{s_2}\cdots\mathbf{A}^{s_N}]. \label{mps8}$$ ### PEPSs in higher dimensions and beyond The AKLT construction can be extended to construct other types of VBS states and states in higher dimensions. Straightforward examples include $S=2$ VBS states on a square lattice and $S=3/2$ VBS states on a honeycomb lattice [@AKLTCMP]. These states can be written as PEPSs in their respective lattices. ![ Graphical representation of a PEPS in terms of contracted tensors (tensor network). Each box denotes a tensor $T$ with components $T^{s_{ij}}_{lrud}$ at site $ij$, where $l$, $u$, $r$, and $d$ are tensor indices related to left, right, up and down bonds, respectively, linking to their neighbors; the open lines represent the physical spin states $s_{ij}$; and the connected lines represent the contraction of the tensors.[]{data-label="fig:PEPS"}](PEPS.eps){width="7.8cm"} For instance, on a square lattice with a coordination number of 4, a generic PEPS wavefunction can be written in terms of rank 4 tensors as follows: $$\label{peps1} |\Psi \rangle=\sum_{[s_{ij}]}\phi([s_{ij}])|[s_{ij}]\rangle,$$ where $i,j=1,\cdots,N$ for an $N\times N$ system, $[s_{ij}]=(s_{11},\cdots ,s_{1N},s_{21},\cdots ,s_{2N},\cdots,s_{N1},\cdots,s_{NN})$ denotes a spin configuration, and $$\label{peps2} \phi([s_{ij}])={\rm Tr}[T^{s_{11}}\cdots T^{s_{1N}}T^{s_{21}}\cdots T^{s_{NN}}].$$ where, the $T^s$s are rank 4 tensors with components $$T_{lrud}^{s_{ij}},$$ where $s_{ij}$ is the physical spin index; $l$, $r$, $u$, and $d$ represent links connected to the tensors at the left, right, up and down neighboring sites $(i-1,j)$, $(i+1,j)$, $(i,j-1)$, and $(i,j+1)$, respectively; and “Tr" means tensor contraction. The above mathematical expression of tensor contraction is usually represented by diagrams such as that shown in Fig. \[fig:PEPS\] for a square lattice, where connected lines represent the contraction of tensors with the same index and open lines represent the physical spin states $s_{ij}=-S,\cdots,S$. ![ The VBS construction of an $S=2$ AKLT state on a square lattice and the corresponding tensors.[]{data-label="fig:AKLT2D"}](AKLT2D.eps){width="8.0cm"} As an example, a spin $S=2$ AKLT state on a square lattice can be written in PEPS form as shown in Fig. \[fig:AKLT2D\]. The tensors $T^s$ can be obtained using the VBS construction with the tensors $\mathbf{R}$ and $\mathbf{M}^{s}$, as in one dimension. The tensor $\mathbf{R}$ is still defined by Eq. . The tensors $\mathbf{M}^{s}$, $s=0,\pm 1, \pm 2$, project a state consisting of four $S=1/2$ spins in the auxiliary Hilbert space ${1\over2}\bigotimes{1\over2}\bigotimes{1\over2}\bigotimes{1\over2}=2\bigoplus 1\bigoplus 0$ into the physical $S=2$ spin space, whose components are given by $$M^{s}_{\sigma_l \sigma_r \sigma_u \sigma_d}=\langle s|\sigma_l \sigma_r \sigma_u \sigma_d\rangle, \label{M:AKLT2}$$ where $\sigma_l,\sigma_r,\sigma_u,\sigma_d=\uparrow,\downarrow$. The tensor $\mathbf{T}$ is given by $$T^{s}_{\sigma_l \sigma_r \sigma_u \sigma_d}=\sum_{\sigma_{l'},\sigma_{u'}} M^{s}_{\sigma_{l'} \sigma_r \sigma_{u'} \sigma_d} R_{\sigma_l \sigma_{l'}} R_{\sigma_u \sigma_{u'}}. \label{T:AKLT2}$$ The tensor product state constructed from the above $T^{s}$s give rise to the $S=2$ AKLT state on a square lattice. The VBS construction can be further extended by “fractionalizing" the spins in more exotic ways (for example, using the Majorana fermion representation of spins). In this way, we can write the toric code model [@Kitaev03] in the PEPS form as well as the Kitaev honeycomb model [@Kitaev06] (with a residual fermionic degree of freedom at each site; see section \[sec:Kitaev\]). The relation between RVB states and PEPSs has also been exploited to show that some RVB states can be written as PEPSs [@Verstraete06a; @Schuch12; @Wang13; @Poiblanc13]. However, the general relation between RVB states and PEPSs remains unclear. The PEPS construction provides a way to describe entanglement among local spins based on the construction of local pairs, and its application to geometrically frustrated lattices is limited. To overcome this limitation, researchers have extended the pair construction procedure to consider entanglement between more than two sites, say, a cluster or a simplex, to construct projected states. These projected entangled simplex states form the basis for more elaborate numerical approaches [@Xie14]. Combined with numerical techniques (tensor-based renormalization), these tensor-network methods now provide an alternative means of constructing variational wavefunctions. Readers can refer to references [@CiracVerstraeteJPA09; @VCM08; @Orus13] for details. Kitaev honeycomb model and related issues {#sec:Kitaev} ----------------------------------------- It was previously believed that spin rotational symmetry is essential for a QSL state that supports fractional spinon excitations. If the spin rotational symmetry is broken, the system tends to approach an ordered state. Kitaev [@Kitaev06] provided a counterexample to this belief through an unusual, exactly solvable model in two dimensions with strong spin-orbit coupling, which destroys the spin rotational symmetry, but in which deconfined spinons nevertheless exist on top of the QSL ground states. This famous model is now called the Kitaev honeycomb model. In this section, we briefly review the Kitaev honeycomb model to see how exotic ground states and low-energy excitations emerge from this model with broken rotational symmetry. The possibility of the realization of Kitaev-like models in realistic materials is also discussed. Kitaev considered a spin-$1/2$ model on a honeycomb lattice with spin-orbit coupling [@Kitaev06]. He divided all nearest neighbor bonds in the honeycomb lattice into three types, called $x$-links, $y$-links and $z$-links as shown in Fig. \[KitaevHoneycomb\]. The Hamiltonian is given as follows: $$H=-J_{x}\sum_{x\text{-link}}K_{ij}-J_{y}\sum_{y\text{-link}% }K_{ij}-J_{z}\sum_{z\text{-link}}K_{ij}, \label{H-Kitaev}$$where $K_{ij}$ is defined as $$K_{ij}=\left\{ \begin{array}{cc} \sigma _{i}^{x}\sigma _{j}^{x}, & \text{if }(i,j)\text{ is a }x\text{-link,} \\ \sigma _{i}^{y}\sigma _{j}^{y}, & \text{if }(i,j)\text{ is a }y\text{-link,} \\ \sigma _{i}^{z}\sigma _{j}^{z}, & \text{if }(i,j)\text{ is a }z\text{-link.}% \end{array}% \right. \label{Kij}$$ Note the strong anisotropy in the spin-spin couplings $K_{ij}$. ![ Kitaev honeycomb model. $x$, $y$ and $z$ denote three types of links in the honeycomb lattice.[]{data-label="KitaevHoneycomb"}](KitaevHoneycomb.eps){width="8.0cm"} We first consider the following loop operators $W_p$ defined for a hexagonal loop: $$W_{p}\equiv \sigma _{1}^{x}\sigma _{2}^{y}\sigma _{3}^{z}\sigma _{4}^{x}\sigma _{5}^{y}\sigma _{6}^{z}=K_{12}K_{23}K_{34}K_{45}K_{56}K_{61}, \label{W-plquette}$$ where $p$ is used to label the lattice plaquettes (hexagons), as shown in Fig. \[KitaevLoop\]. It is easy to verify that $[W_{p},K_{ij}]=0$; therefore, $% [H,W_{p}]=0$. Hence, the $W_{p}$s serve as good quantum numbers for the Hamiltonian given in Eq. , and the total Hilbert space for spins can be divided into a direct product of sectors that are eigenspaces of $\{W_{p}\}$. However, the eigenvalue problem cannot be completely solved by determining the eigenspaces of $\{W_p\}$. Each $W_{p}$ has only two eigenvalues, $w_{p}=\pm 1$. Each plaquette contains six sites, and each site is shared by three plaquettes. Therefore, the number of plaquettes is given by $m=N/2$, where $N$ is the number of sites. It follows that the dimension of each eigenspace of $\{W_p\}$ is $2^{N}/2^{m}=2^{N/2}$, i.e., splitting the Hilbert space into eigenspaces of $\{W_{p}\}$ cannot solve the eigenvalue problem completely. ![ Loop operator $W_{p}=\protect\sigma _{1}^{x}\protect\sigma _{2}^{y}\sigma _{3}^{z}\protect\sigma _{4}^{x}\protect\sigma _{5}^{y}\sigma _{6}^{z}$ on a lattice plaquette (hexagon).[]{data-label="KitaevLoop"}](KitaevLoop.eps){width="4.0cm"} Kitaev realized that to solve the model Hamiltonian given in Eq. , spins can be written in terms of four Majorana fermions, because a Majorana fermion can be viewed as the real or imaginary part of a complex fermion. To illustrate this approach, we rewrite the complex fermions $f_{\uparrow }$ and $f_{\downarrow }$ in Eq.  (\[slaveparticle\]) in terms of four Majorana fermions $c_{1}$, $c_{2}$, $c_{3}$ and $c_{4}$: $$\begin{array}{cc} f_{\uparrow }=\frac{1}{2}(c_{1}+ic_{2}), & f_{\uparrow }^{\dag }=\frac{1}{2}% (c_{1}-ic_{2}), \\ f_{\downarrow }=\frac{1}{2}(c_{3}+ic_{4}), & f_{\downarrow }^{\dag }=\frac{1% }{2}(c_{3}-ic_{4}),% \end{array} \label{MFCF}$$ where the operators $c_{\alpha }$ ($\alpha =1,2,3,4$) are Hermitian and satisfy$$c_{\alpha }c_{\beta }+c_{\beta }c_{\alpha }=2\delta _{\alpha \beta }. \label{MF1}$$ Thus, the three spin components read $\sigma ^{x}=\frac{i}{2}(c_{1}c_{4}-c_{2}c_{3})$, $\sigma ^{y}=\frac{i}{2}(c_{3}c_{1}-c_{2}c_{4})$, and $\sigma ^{z}=\frac{i}{2}(c_{1}c_{2}-c_{3}c_{4})$. The single-occupancy condition $f_{\uparrow }^{\dag }f_{\uparrow }+f_{\downarrow }^{\dag }f_{\downarrow }=1$ (and $f_{\uparrow }^{\dag }f_{\downarrow}^{\dag}=f_{\uparrow }f_{\downarrow }=0$) becomes$$c_{1}c_{2}+c_{3}c_{4}=c_{1}c_{3}+c_{2}c_{4}=c_{1}c_{4}+c_{3}c_{2}=0, \label{single_occupancyMF}$$ which can be simplified to the single equation $c_{1}c_{2}c_{3}c_{4}=1$. Using these constraints, the spin operators can be written as $\sigma^{x}=ic_{1}c_{4}$, $\sigma ^{y}=-ic_{2}c_{4}$, and $\sigma^{z}=-ic_{3}c_{4}$. Rewriting $b_{x}=c_{1}$, $b_{y}=-c_{2}$, $b_{z}=-c_{3}$ and $c=c_{4}$, we arrive at the Kitaev representation $$\begin{aligned} \sigma ^{x} &=&ib^{x}c, \notag \\ \sigma ^{y} &=&ib^{y}c, \label{Majorana4} \\ \sigma ^{z} &=&ib^{z}c, \notag\end{aligned}$$with the constraint$$D\equiv b^{x}b^{y}b^{z}c=1. \label{Majorana4D}$$ The Majorana representation without constraints is redundant and enlarges the physical spin Hilbert space. Note that $D^{2}=1$ and that $D$ has two eigenvalues, $D=\pm 1$, thereby splitting the local Hilbert space into two sectors. The physical spin Hilbert space corresponds to the sector with all $D_{j}=1$. Therefore, the physical spin wavefunction $\left\vert \Psi _{spin}\right\rangle $ can be obtained from the Majorana fermion wavefunction $\left\vert \Psi _{Majorana}\right\rangle $ through the projection$$\left\vert \Psi _{spin}\right\rangle =\prod_{j}\frac{1+D_{j}}{2}\left\vert \Psi _{Majorana}\right\rangle , \label{ProjectionMF}$$ which retains the $D_{j}\equiv 1$ sector and removes all other sectors in the enlarged Hilbert space. Note that $\frac{1+D_{j}}{2}=n_{j\uparrow}+n_{j\downarrow }-2n_{j\uparrow }n_{j\downarrow }$ and that Eq.  is nothing but the Gutzwiller projection. In addition, note that $D_{j}$ serves as a $Z_{2}$ gauge transformation in the enlarged Hilbert space ($D_{j}b_{j}^{\alpha }D_{j}=-b_{j}^{\alpha}$, $D_{j}c_{j}D_{j}=-c_{j}$) and commutes with the spin operators ($[D_j,\sigma^{\alpha}_j]=0$, $\alpha=x,y,z$) and thus with the Hamiltonian. As a result, the Gutzwiller projection is “trivial" in the sense that $\prod_{j}\frac{1+D_{j}}{2}\left\vert\Psi _{Majorana}\right\rangle$ is an eigenstate of $H$ in the projected Hilbert space as long as $\left\vert\Psi_{Majorana}\right\rangle$ is an eigenstate of $H$ in the “unprojected" Hilbert space and $\prod_{j}\frac{1+D_{j}}{2}\left\vert\Psi _{Majorana}\right\rangle\neq0$. ![ Graphic representation of the four-Majorana-fermion decomposition of the Hamiltonian expressed in Eq. .[]{data-label="KitaevMajoranaF"}](KitaevMajoranaF.eps){width="8.4cm"} In the Majorana fermion representation, $K_{ij}$ in Eq.  becomes$$K_{ij}=-i(ib_{i}^{\alpha }b_{j}^{\alpha })c_{i}c_{j}, \label{Kij2}$$where $\alpha =x,y,z$ depends on the type of link $(ij)$. The operator $% ib_{i}^{\alpha }b_{j}^{\alpha }$ is Hermitian, and we denote it by $\hat{u}% _{ij}=ib_{i}^{\alpha }b_{j}^{\alpha }$. Thus, we may write $$H=\frac{i}{4}\sum_{\langle j,k\rangle }\hat{A}_{jk}c_{j}c_{k}, \label{H-KitaevMF}$$ with$$\hat{A}_{jk}=2J_{\alpha (jk)}\hat{u}_{jk},\, \hat{u}_{jk}=ib_{j}^{\alpha }b_{k}^{\alpha }, \label{Ajk}$$ where $\langle j,k\rangle $ denotes nearest neighbor links on the honeycomb lattice and, by definition, $\hat{u}_{jk}=-\hat{u}_{kj}$ and $\hat{A}_{jk}=-\hat{A}_{kj}$. The Hamiltonian structure in this Majorana fermion representation is shown schematically in Fig. \[KitaevMajoranaF\]. Note that $[H,\hat{u}_{jk}]=0$ and $[\hat{u}_{jk},\hat{u}_{j^{\prime }k^{\prime }}]=0$. The enlarged Hilbert space of Majorana fermions can be decomposed into common eigenspaces of $\{\hat{u}_{jk}\}$ indexed by the corresponding eigenvalues $u_{jk}=\pm 1$. Thus, the Hamiltonian in the invariant subspace indexed by $u=\{u_{jk}\}$ becomes$$H_{u}=\frac{i}{4}\sum_{\langle j,k\rangle }A_{jk}c_{j}c_{k},A_{jk}=2J_{\alpha (jk)}u_{jk}, \label{H-KitaevMFu}$$ where we have replaced $\hat{A}_{jk}$ and $\hat{u}_{jk}$ with their eigenvalues. Note that $u_{jk}\rightarrow-u_{jk}$ upon the $Z_2$ gauge transformation $u_{jk}\rightarrow D_ju_{jk}D_j$, and it is more convenient to classify the eigenstates of $H$ in terms of the gauge-invariant loop operator $W(j_{0},\cdots ,j_{n})=K_{j_{n}j_{n-1}}\cdots K_{j_{1}j_{0}}$, which can be written as$$W(j_{0},\cdots ,j_{n})=\left( \prod_{s=1}^{n}-i\hat{u}_{j_{s}j_{s-1}}\right) c_{n}c_{0}. \label{W-Loop-Majorana}$$The closed-loop operator $W_{p}$ (see Eq. ) is gauge invariant under the $Z_2$ transformation because $c_{n}=c_{0}$, and the gauge-invariant quantities $w=\{w_{p}\}$ can be used instead of $% u=\{u_{jk}\}$ to parameterize the eigenstates, i.e.,$$H_{w}=\frac{i}{4}\sum_{\langle j,k\rangle }A_{jk}c_{j}c_{k}. \label{H-KitaevMFw}$$For a given set of $A_{ij}$ fixed by $\{w_p\}$, the quadratic Hamiltonian as expressed in Eq.  and Eq.  can be diagonalized into the following canonical form:$$H_{canonical}=\frac{i}{2}\sum_{m}\epsilon _{m}c_{m}^{\prime }c_{m}^{\prime \prime }=\sum_{m}\epsilon _{m}\left( f_{m}^{\dagger }f_{m}-\frac{1}{2}% \right) , \label{H-canonical-Majorana}$$where $\epsilon _{m}\geq 0$, $c_{m}^{\prime }$ and $c_{m}^{\prime \prime }$ are normal Majorana modes, and $f_{m}^{\dagger }=\frac{1}{2}(c_{m}^{\prime }-ic_{m}^{\prime \prime })$ and $f_{m}=\frac{1}{2}(c_{m}^{\prime }+ic_{m}^{\prime \prime })$ are the corresponding complex fermion operators. The ground state of the Majorana system has an energy of$$E=-\frac{1}{2}\sum_{m}\epsilon _{m}. \label{Eg-MF}$$ We now discuss the system of Majorana fermions on the honeycomb lattice. First, we note that the global ground-state energy does not depend on the signs of the exchange constants $J_{x}$, $J_{y}$, and $J_{z}$. For instance, if $J_{z}$ is replaced with $-J_{z}$, we can compensate for this sign change by changing the signs of the variables $u_{jk}$ for all $z$-links using the gauge operator $D_{j}$, leaving the values of $A_{jk}$ and $w_{p}$ unchanged. Therefore, as far as solving for the ground-state energy and the excitation spectrum is concerned, the signs of the exchange constants $J$ do not matter. However, such a sign change does affect other measurable physical quantities. Second, it was proven by Lieb [@Lieb94] and numerically investigated by Kitaev himself that the ground state of the Majorana system is achieved when the system is in the vortex-free configuration, namely, $w_{p}=1$ for all plaquettes $p$. In this vortex-free configuration, one can solve for the (fermionic) energy spectrum of the Hamiltonian by directly Fourier transforming Eq.  to obtain$$\epsilon _{\mathbf{q}}=\pm |J_{x}e^{i\mathbf{q}\cdot \mathbf{a}}+J_{y}e^{i% \mathbf{q}\cdot \mathbf{b}}+J_{z}|, \label{Eq-MajoranaF}$$ where $\mathbf{a=(}\frac{1}{2},\frac{\sqrt{3}}{2}\mathbf{)}$ and $\mathbf{b=(-}\frac{1}{2},\frac{\sqrt{3}}{2}\mathbf{)}$ are two basis vectors in the $% xy$ coordinates. The fermionic spectrum may or may not be gapped, depending on whether a solution to the equation $\epsilon _{\mathbf{q}}=0$ exists. $\epsilon _{\mathbf{q}}=0$ has a solution if and only if $|J_{x}|$, $|J_{y}|$, and $|J_{z}|$ satisfy the triangle inequalities:$$|J_{x}|\leq |J_{y}|+|J_{z}|,|J_{y}|\leq |J_{z}|+|J_{x}|,|J_{z}|\leq |J_{x}|+|J_{y}|. \label{Jxyz-phase}$$ ![ Phase diagram of the Kitaev honeycomb model. The triangle is the section of the positive octant ($J_x,J_y,J_z\geq 0$) that lies in the plane $J_x+J_y+J_z=1$. The $A$ phase contains three gapped subphases. The $B$ phase is gapless.[]{data-label="KitaevPhase"}](KitaevPhase.eps){width="6.4cm"} As a result, two phases exist in the system of Majorana fermions on the honeycomb lattice, with the phase diagram shown in Fig. \[KitaevPhase\]. The first phase, called the $A$ phase, is gapped and contains three subphases ($% A_{x}$, $A_{y}$, and $A_{z}$) in the phase diagram. The second, called the $% B$ phase, is gapless. In the $A$ phase, for example, in the $A_{z}$ subphase, the Hamiltonian expressed in Eq. ) can be mapped to the Kitaev toric code model in the limit $|J_{x}|,|J_{y}|\ll |J_{z}|$, and the phase hosts Abelian anyonic excitations. The $B$ phase acquires an energy gap in the presence of a magnetic field. Very interestingly, it hosts stable non-Abelian anyons when the energy gap is opened up by a magnetic field. The $B$ phase is a very attractive state in the context of topological quantum computation. Readers can refer to the recent review article [@NayakRMP08] for details. In addition to the elegant Majorana decomposition method pioneered by Kitaev, other insightful approaches to the Kitaev honeycomb model also exist. For instance, Feng, Zhang and Xiang [@Feng07] and Chen and Nussinov [@ChenNussinov08] found that the original Kitaev honeycomb model can be exactly solved with the help of the Jordan-Wigner transformation. This approach provides a topological characterization of the quantum phase transition from the $A$ phase to the $B$ phase. A nonlocal string order parameter can be defined in one of these two phases [@Feng07; @ChenNussinov08]. In the appropriate dual representations, these string order parameters become local order parameters [*after some singular transformation*]{}, and a description of the phase transition in terms of Landau’s theory of continuous phase transitions becomes applicable [@Feng07]. The Jordan-Wigner transformation also enables a fermionization of the Kitaev honeycomb model, allowing it to be mapped to a $p$-wave-type BCS pairing problem. The spin wavefunction can be obtained from the fermion model, and the anyonic character of the vortex excitations in the gapped phase also has an explicit fermionic construction [@ChenNussinov08]. The Kitaev honeycomb model can also be understood within the framework of fermionic RVB theory. Both confinement-deconfinement transitions from spin liquids to AFM or stripy AF/FM phases and topological quantum phase transitions between gapped and gapless spin liquid phases can be described within the framework of $Z_2$ gauge theory [@Baskaran07; @Mandal11; @Mandal12]. Exact diagonalization has been applied to study the Kitaev honeycomb model on small lattices [@ChenWangDasSarma10]. Perturbative expansion methods have been developed to study the gapped phases of the Kitaev honeycomb model and its generalization [@Schmidt08; @Vidal08; @Dusuel08]. Several papers [@Lee07; @Yu08a; @Yu08b; @Kells09] have noted the existence of an analogy between the $Z_2$ vortices in the Kitaev honeycomb model and the vortices in $p+ip$ superconductors. Enormous efforts have been devoted to searching for exactly solvable generalizations of the Kitaev honeycomb model. It has been proposed that the exact solvability will not be spoiled when the fermion gap is opened for the non-Abelian phase [@Lee07; @Yu08b]. Generalizations to other lattice models and even to three dimensions have also been developed [@Yao07; @Yang07; @Yao09; @Nussinov09; @Wu09; @Ryu09; @Baskaran09; @Tikhonov10; @Yao11; @Lai11]. Non-trivial emergent particles, such as chiral fermions [@Yao07], have been constructed in these exactly solvable lattice models. These developments have significantly advanced our understanding of emergent phenomena based on solvable models in dimensions greater than one. The exotic properties of the Kitaev honeycomb model have motivated researchers to search for realizations of this model in realistic materials. It has been demonstrated by Jackeli and Khaliullin [@Jackeli09] and by Chaloupka, Jackeli and Khaliullin [@Chaloupka10] that a generalization of the Kitaev honeycomb model may indeed arise in layered honeycomb lattice materials in the presence of strong spin-orbit coupling. These authors showed that in certain iridate magnetic insulators (A$_2$IrO$_3$, A=Li, Na), the effective low-energy Hamiltonian for the effective $J_{eff}=1/2$ iridium moments is given by a linear combination of the AFM Heisenberg model ($H_H$) and the Kitaev honeycomb model ($H_K$), $$H = (1 - \alpha)H_H + 2\alpha H_K, \label{HHK}$$ where $\alpha$, expressed in terms of the microscopic parameters, determines the relative strength of the Heisenberg and Kitaev interactions. Interestingly, the Kitaev honeycomb model can also be realized as the exact low-energy effective Hamiltonian of a spin-$1/2$ model with spin rotational and time-reversal symmetries [@Wang10]. The Heisenberg-Kitaev model (\[HHK\]) exhibits a rich phase diagram. Readers who are interested in these developments may refer to, for example, references [@Chaloupka10; @Jiang11; @Reuther11; @Kimchi11; @Singh12; @Price12; @Schaffer12; @Yu13; @Lee14; @Kimchi14] for details. A comprehensive review on this topic has also been published by Nussinov and van den Brink [@Nussinov13]. QSL states in real materials ============================ Experimental studies of interacting spins in geometrically frustrated lattices aim at identifying non-trivial and exotic ground states. Among these ground states, spin liquid states have been sought ever since the proposal of the RVB state [@Anderson73]. This issue has been intensively debated in the context of the spin states behind the high-Tc superconductivity of cuprates. However, before this century, there was no direct observation of spin liquid states. The situation changed in 2003, when an organic Mott insulator with a quasi-triangular lattice was found to exhibit no magnetic ordering even at tens of mK, four orders of magnitude lower than the energy scale of the exchange interactions [@Kanoda03]. The low-temperature state is most likely a form of the sought-after spin liquids. Since then, what can be called spin liquids have been successively reported for quasi-triangular, kagome and hyperkagome lattices. In this section, we review the experimental studies mainly with respect to the magnetic and thermodynamic properties of the materials for which sound experimental data have been accumulated in discussing the presence of spin liquids. Anisotropic triangular lattice systems: $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ and EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ ----------------------------------------------------------------------------------------------------------------------- Both are half-filled band systems with anisotropic triangular lattices, which are isosceles for $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ and three different laterals for EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ [@Kanoda06b; @Kanoda11; @Kato14]. At ambient pressure, they are Mott insulators; however, the spins are not ordered at low temperatures on the order of tens of mK. A noticeable feature of both systems is that they undergo Mott transitions at moderate pressures 0.4 GPa for $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ [[@Komatsu96; @Kurosaki05; @Furukawa15a]]{} and 0.5 GPa for EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ [@Kato06]. (Note that these pressure values indicate pressures applied at room temperature and are reduced by approximately 0.2 GPa at low temperatures.) The temperature-pressure phase diagram of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ is depicted in Fig. \[Kuroaki\_k\_Cu2CN3\_Phase\_Diagram\]. A spin liquid phase resides in proximity to the Mott transition; this feature appears to be a key to the stability of spin liquids and can be closely linked to the metal-insulator transition [@Senthil08; @ZhouNg13]. According to the numerical studies of the anisotropic triangular-lattice Hubbard model, the ground states near to the Mott transition are controversial [[@Morita02; @Kyung06; @Watanabe08; @Tocchio13; @Laubach15]]{}, implying that spin-liquid and magnetic phases are competing very closely and can be easily imbalanced by a tiny perturbation. ![[@Kurosaki05] Temperature-pressure phase diagram of the spin-liquid compound with a quasi-triangular lattice, $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$, which undergoes a Mott transition at moderate pressure.[]{data-label="Kuroaki_k_Cu2CN3_Phase_Diagram"}](Kuroaki_k_Cu2CN3_Phase_Diagram.eps){width="8cm"} i\) $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ is a layered compound, where $\kappa$-(ET)$_{2}$X has a variety of anions X and ET is bis(ethylenedithio)tetrathiafulvalene [@Komatsu96]. $\kappa$-(ET)$_{2}$X is composed of the ET layers with 1/2 hole per ET and the layers of monovalent anions X$^{-}$, which have no contribution to the electronic conduction or magnetism. In the ET layer, strong ET dimers are formed (ET)$_{2}$, each of which accommodates a hole in an anti-bonding orbital of the highest occupied molecular orbital (HOMO) of the ET. As the anti-bonding band is half-filled and the Coulomb repulsive energy is comparable to the band width, the family of $\kappa$-(ET)$_{2}$X is good model system to study Mott physics [@Kino95; @Kanoda97a; @Kanoda97b; @Kanoda06; @Powell11]. The estimates of the transfer integrals between the adjacent anti-bonding orbitals on the isosceles triangular lattices, $t$ and $t^{\prime}$, are in a range of 50 meV, depending on the method of calculation, e.g., either the molecular orbital (MO)-based tight-binding calculation [@Mori84; @Mori99; @Komatsu96] or the first principles calculation [@Nakamura09; @Kandpal09; @Koretsune14]. Nevertheless, one can see that the values have clear systematic variation in terms of anion X, as shown in Fig. \[Kandpal\_Nakamura\_transfer\], where the values of $t$ and $t^{\prime}$ are calculated via the latter method: the $t^{\prime}/t$ value of $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl is 0.75 (the MO-based calculations) and 0.44-0.52 (first principles calculations), while that of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ is 1.06 and 0.80-0.99, respectively, suggestive of high geometrical frustration. ![(a) In-plane structure of the ET layer in $\kappa$-(ET)$_{2}$X. It is modelled to (b) an anisotropic triangular lattice. (c) First principles calculations of transfer integrals in $\kappa$-(ET)$_{2}$X for X=Cu\[N(CN)$_{2}$\]Cl, Cu(NCS)$_{2}$ and Cu$_{2}$(CN)$_{3}$ ; squares [@Nakamura09], circles [@Kandpal09], and triangles [@Koretsune14].[]{data-label="Kandpal_Nakamura_transfer"}](Kandpal_Nakamura_transfer.eps){width="8cm"} The temperature dependence of the spin susceptibility, $\chi$, of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ differs from that of the less frustrated compound $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl, as seen in Fig. \[Shimizu\_sus\] [@Kanoda03]. An abrupt upturn at 27 K in the latter is a manifestation of the antiferromagnetic transition, with a slight spin canting of approximately 0.3 degree [@Miyagawa95]. However, $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ has no anomaly in $\chi(T)$. Its overall behavior features a broad peak, which is reconciled by the triangular-lattice Heisenberg model with an exchange interaction of $J \sim$ 250 K. In contrast to $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl, the magnetic susceptibility of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ may be described by the Heisenberg model because it is situated further from the Mott boundary, while $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl undergoes a Mott transition at a low pressure (25 MPa) as it is about to enter a metallic state [@Lefebvre00; @Kagawa05]. There is no indication of magnetic ordering in the susceptibility of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$, at least down to 2 K, the lowest temperature measured. Furthermore, no Curie-like upturn can be identified; the concentration of Cu$^{2+}$ impurity spins detected by ESR is estimated to be less than 0.01 % for $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ [@Kanoda06]. ![[@Kanoda03] Magnetic susceptibility of poly crystalline $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ and $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl. The core diamagnetic susceptibility has already been subtracted. The solid and dotted lines represent the result of the series expansion of the triangular-lattice Heisenberg model using \[6/6\] and \[7/7\] Padé approximations with $J$ = 250 K. The susceptibility of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ below 30 K is expanded in the inset.[]{data-label="Shimizu_sus"}](Shimizu_sus.eps){width="8cm"} ![[@Kanoda03] $^{1}$H NMR spectra for single crystals of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ and $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl.[]{data-label="Shimizu_1H_spc"}](Shimizu_1H_spc.eps){width="8cm"} The detailed spin states can be examined by performing NMR measurements, which probe the static and dynamical hyperfine fields at the nuclear sites. Fig. \[Shimizu\_1H\_spc\] shows the single-crystal $^{1}$H NMR spectra for the two compounds [@Kanoda03]. A clear line splitting in $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl at 27 K is evidence for commensurate antiferromagnetic ordering, with the moment estimated to be 0.45 $\mu_{\textrm{B}}$ per ET dimer in separate $^{13}$C NMR studies [@Miyagawa04]. However, the spectra for $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ shows neither a distinct broadening nor splitting down to 32 mK, which is four orders of magnitude lower than the $J$ value of 250 K. This indicates the absence of long-range magnetic ordering. The absence of ordering is also corroborated by zero field $\mu$SR experiments [@Pratt11]. The nuclear spin lattice relaxation rate, $1/T_{1}$, which probes the spin dynamics, behaves similarly at the $^{1}$H and $^{13}$C sites. Fig. \[Shimizu\_13C\_T1\] shows $1/T_{1}$ at the $^{13}$C sites, which decreases monotonically with a square-root temperature dependence down to 10 K and exhibits a dip-like anomaly at approximately 6 K [@Kanoda06]. Below 6 K, $1/T_{1}$ levels off down to 1 K or lower, followed by a steep decrease approximated by $T^{3/2}$ at even lower temperatures. The two anomalies at 6 K and 1.0 K are obvious. However, they are not so sharp as to be considered as phase transitions. Due to the large hyperfine coupling of the $^{13}$C sites located in the central part of ET, an electronic inhomogeneity gradually developing on cooling is captured by spectral broadening, which is enhanced at approximately 6 K and saturates below 1 K [@Kanoda06; @Kawamoto06]. The detailed NMR [@Kanoda06] and $\mu$SR [@Pratt11] measurements point to the field-induced emergence of staggered-like moments, which is distinct from the conventional magnetic order. A separate $\mu$SR study [@Goto12] suggests a phase separation. The degree of inhomogeneity in the $^{13}$C relaxation curve, which is characterized by the deviation of the exponent in the stretched exponential fitting of the relaxation curve (see Inset of Fig. \[Shimizu\_13C\_T1\]), increases below 5-6 K [@Kanoda06]. The $^{1}$H relaxation curve also starts to bend at the much lower temperatures, e.g., below 0.4 K, and fits to a roughly equally weighed sum of two exponential functions, the 1/$T_{1}$’s of which are proportional to $T$ and $T^{2}$. No appreciable field dependence of the $^{13}$C relaxation rate is observed between 2 T and 8 T. There is no experimental indication of a finite excitation gap in any of the magnetic measurements. ![[@Kanoda06] $^{13}$C nuclear spin-lattice relaxation rate for a single crystal of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$. The open triangles and circles represent the relaxation rates of two separated lines coming from two non-equivalent carbon sites in an ET. At low temperatures below 5 K, the two lines merge and are not distinguished. The inset shows the exponent in the stretched exponential fitting to the relaxation curves of the whole spectra, whose relaxation rates are plotted using closed diamonds.[]{data-label="Shimizu_13C_T1"}](Shimizu_13C_T1.eps){width="8cm"} Thermodynamic investigations were conducted by means of the specific heat and thermal conductivity measurements. Fig. \[SYamashita\_specific\_heat\] shows the specific heat for $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ and several Mott insulators with antiferromagnetic spin ordering [@SYamashita]. For all of the antiferromagnetic materials, the electronic specific heat coefficient, $\gamma$, is vanishing, as expected for insulators. For the $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ spin liquid system, however, the extrapolation of the $C/T$ vs. $T^{2}$ line to absolute zero yields $\gamma$=12$\sim$15 mJ/K$^{2}$mol. The linearity holds down to 0.3 K, below which a nuclear Schottky contribution overwhelms the electronic contribution to $C$. The finite value despite the Mott insulating state is a marked feature of spin liquids and suggests fermionic excitations in the spin degrees of freedom. Interestingly, the low-temperature susceptibility and the $\gamma$ value give the Wilson ratio on the order of unity. A spinon Fermi sea is an intriguing model for this phenomenon [@Motrunich05]. However, neither the $U(1)$ spin liquid, where $C$ follows $T^{2/3}$ scaling, nor the $Z_{2}$ spin liquid, where $C$ is gapped, reconciles the observed features in their original forms. Randomness may be an optional parameter to modify the temperature dependence. Another interesting feature is the field-insensitivity up to 8 T, which appears incompatible with the $U(1)$ spin liquid states with Dirac cones. ![[@SYamashita] Low-temperature specific heat $C_p$ of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ for several magnetic fields up to 8 T in $C_p/T$ versus $T^{2}$ plots. Those of antiferromagnetic insulators $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl, deuterated $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Br and $\beta$-(ET)$_{2}$ICl$_{2}$ are also plotted for comparison.[]{data-label="SYamashita_specific_heat"}](SYamashita_specific_heat.eps){width="8cm"} Thermal transport measurements result in somewhat controversial consequences [@MYamashita09]. The thermal conductivity divided by the temperature tends to vanish with decreasing temperature, as shown in Fig. \[MYamashita\_TC\]. The gap, if one is present, is estimated to be 0.43 K, which is quite small compared with the exchange energy of 250 K. The extremely small gap may indicate a gapped $Z_{2}$ spin liquid located near a quantum critical point. The discrepancy between the thermal transport and NMR and specific heat data remains an open issue. It may be attributed to the Anderson localization of spinons. ![[@MYamashita09] Low-temperature thermal conductivity $\kappa$ of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ (samples A and B) in $\kappa /T$ versus $T^{2}$ plots. Sample A was investigated at 10 T applied perpendicular to the basal plane, as well as at 0 T.[]{data-label="MYamashita_TC"}](MYamashita_TC.eps){width="8cm"} The 6-K anomaly in the NMR spectrum and relaxation rate also manifests itself in the specific heat [@SYamashita] and thermal conductivity [@MYamashita09] as a hump and a shoulder, respectively, indicating that the anomaly is thermodynamic, as well as magnetic. However, the thermal expansion coefficient shows a cusp [@Manna2010] and the ultrasonic velocity shows a dip-like minimum, signifying lattice softening at approximately 6 K [@Poirier14]. In view of these results, this anomaly is likely associated with spin-lattice coupling. Instabilities of the spinon Fermi surfaces (e.g., [@Lee05; @Galitski07; @Grover10; @YZhou2011]) are among the possible origins of the anomaly. Although the spin liquid is insulating, anomalous charge dynamics are suggested for the low-energy optical and dielectric responses. The optical gap for $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ is much smaller than that for $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl, although the former system is situated further from the Mott transition than the latter [@Ketzmarki06]. It is proposed that gapless spinons are responsible for low-energy optical absorption inside the Mott gap [@NgLee07]. The dielectric [@Abdel-Jawad10], microwave [@Poirier12] and terahertz [@Itoh13] responses are enhanced at low temperatures. The possible charge-imbalance excitations within the dimer are theoretically proposed [@Hotta10; @Naka10; @Dayal11]. Relaxor-like dielectric, transport and optical properties are discussed in terms of coupling with disordered anion layers [@Pinteric14; @Dressel16]. ii\) EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ This compound is a member of the A\[(Pd(dmit)$_{2}$\]$_{2}$ family of materials, which contain a variety of monovalent cations such as A$^{+}$=Et$_{x}$Me$_{4-x}$Z$^{+}$ (Et = C$_{2}$H$_{5}$, Me= CH$_{3}$, Z =N, P, As, Sb, and $x$ = 0, 1, 2), where dmit is 1,3-dithiole-2-thione-4,5-dithiolate [@Kato14] A\[(Pd(dmit)$_{2}$\]$_{2}$ is a layered system composed of conducting Pd(dmit)$_{2}$ layers and insulating A layers. In the conducting layers, Pd(dmit)$_{2}$ is strongly dimerized as in $\kappa$-(ET)$_{2}$X, whereas the \[Pd(dmit)$_{2}$\]$_{2}$ dimer accepts an electron from cation A$^{+}$ instead of the hole in ET$_{2}^{+}$. A prominent feature of the A\[(Pd(dmit)$_{2}$\]$_{2}$ family is that the transfer integrals of the three laterals in the triangular lattice can be finely tuned via chemical substitution of A$^{+}$=Et$_{x}$Me$_{4-x}$Z$^{+}$ (Kato, 2014). Their first principles calculations are shown in Fig. \[Tsumuraya\_band\_cals\] [@Tsumuraya13] The spin liquid material EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ is in a region where the three transfer integrals are equalized. As expected, the materials situated outside of this region have antiferromagnetic ground states. The alloying of the boundary materials offers the chance to study possible critical regions between spin liquids and ordered states [@Kato14]. There is a charge-ordered material near the spin liquid, suggesting that the charge cannot always be assumed to be separate degrees of freedom from the spin physics. ![[@Tsumuraya13] First principles calculations of band width $W$ (a) and transfer integrals (b) in A\[(Pd(dmit)$_{2}$\]$_{2}$ for various cations, A. The Pd(dmit)$_{2}$ layers are modeled to triangular lattices characterized by transfer integrals, $t_{\textrm{B}}$, $t_{\textrm{S}}$ and $t_{r}$. $t_{3}$ is the interlayer transfer integral. AF, QSL and CO stand for antiferromagnet, quantum spin liquid and charge-ordered insulator.[]{data-label="Tsumuraya_band_cals"}](Tsumuraya_band_cals.eps){width="8cm"} Below, we review the properties of EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ and other related materials. The magnetic susceptibility of EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ shows a broad peak at approximately 50 K and points to a finite value in the low-temperature limit without any anomaly down to 2K, as shown in Fig. \[Kato\_sus\] [@Kato14; @Kanoda11], which is reminiscent of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$. The fitting of the triangular lattice Heisenberg model to the data yields an exchange interaction of 220 K to 280 K, which is nearly the same as for $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$. Also shown are the susceptibilities of antiferromagnetic and charge-ordered insulators, which exhibit small kink signaling of magnetic ordering and a sudden decrease indicative of a spin gapful state, respectively, despite their similar behaviors at high temperatures [@Tamura02]. This indicates that the diversity in the ground states is an outcome of low-energy physics, while the same diversity is not distinguished at high energy scales. ![[@Kato14] Magnetic susceptibility of an antiferromagnet Me$_{4}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$, a spin liquid EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ and a charge-ordered insulator Et$_{2}$Me$_{2}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$. The core diamagnetic susceptibility has already been subtracted.[]{data-label="Kato_sus"}](Kato_sus.eps){width="8.4cm"} ![[@Itou10] $^{13}$C nuclear spin-lattice relaxation rate $1/T_{1}$ of EtMe$_{3}$Sb\[Pd(dmit)$_{2}$\]$_{2}$. Inset shows the $1/T_{1}T$ versus $T$ plots. The circles indicate the values determined from the stretched-exponential fitting to the relaxation curves and the squares denote the values determined from the initial decay slopes of the relaxation curves.[]{data-label="Itou_13C_T1"}](Itou_13C_T1.eps){width="8.4cm"} The $^{13}$C NMR captures no signature of magnetic ordering down to 20 mK, although a slight broadening equivalent to the broadening for $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ is observed at low temperatures [@Itou10]. The temperature dependence of the $^{13}$C nuclear spin-lattice relaxation rate is shown in Fig. \[Itou\_13C\_T1\] [@Itou10]. It exhibits a non-monotonic temperature dependence. At low temperatures below 1 K, it follows a $T^{2}$ dependence, suggesting no finite gap. However, the power of 2 implies a complicated nodal gap, which is not obviously consistent with the finite susceptibility value and the thermodynamic measurements described below. Furthermore, $1/T_{1}$ forms a shoulder or a kink at approximately 1 K and becomes moderate in temperature dependence above 1 K. The kink temperature increases for higher magnetic fields or frequencies. The relaxation curve becomes a non-single exponential curve below 10 K but reverses below 1 K, indicating that the inhomogeneity increases below 10 K [@Itou10; @Itou11]. The reversal at 1 K can be an indication of either a recovery in the homogeneity below 1 K or the microscopic nature of the inhomogeneity, which is subject to spin-diffusion averaging of the heterogeneous relaxation time that is longer at lower temperatures. The 1-K relaxation-rate anomaly in Et$_{2}$Me$_{2}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ may be compared to the broad anomaly around nearly the same temperature for $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$. However, they appear different with respect to field (or frequency) dependence and spatial scale of inhomogeneity. The thermodynamic measurements are indicative of fermionic low-energy excitations. Fig. \[SYamashita\_specific\_heat2\] shows the temperature dependence of the specific heat [@SYamashita11]. The linearity of $C/T$ versus $T^{2}$ in EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ is extrapolated to a zero Kelvin to give a finite value of $\gamma$, whereas other Mott insulators appear to have vanishing $\gamma$, as expected for conventional insulators. There is no field dependence in $C/T$ in EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ up to 8 T, as in $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$. The thermal conductivity results are consistent with the specific heat results, as seen in Fig. \[MYamashita\_TC2\], where the low-temperature $\kappa /T$ value for EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$ is as high as 0.2 WK$^{-2}$m in the zero-Kelvin limit, implying the presence of gapless thermal transporters with fermionic statistics [@MYamashita10]. The mean free path for thermal transport is estimated to be of the order of 1 $\mu$m. $\kappa$ is enhanced by the application of a magnetic field above a threshold value, suggesting that the gapped excitations coexist with the gapless excitations [@MYamashita10]. ![[@SYamashita11] Low-temperature specific heat $C_p$ of EtMe$_{3}$Sb\[Pd(dmit)$_{2}$\]$_{2}$ for several magnetic fields up to 10 T in $C_p/T$ versus $T^{2}$ plots. The data of other insulating systems, i.e., Et$_{2}$Me$_{2}$As\[(Pd(dmit)$_{2}$\]$_{2}$, EtMe$_{3}$As\[(Pd(dmit)$_{2}$\]$_{2}$ and EtMe$_{3}$P\[(Pd(dmit)$_{2}$\]$_{2}$, are also plotted for comparison. A large upturn below 1 K is probably attributable to the rotational tunneling of Me groups. The low-temperature data are expanded in the inset.[]{data-label="SYamashita_specific_heat2"}](SYamashita_specific_heat2.eps){width="8cm"} ![[@MYamashita10] Low-temperature thermal conductivity $\kappa$ of EtMe$_{3}$Sb\[Pd(dmit)$_{2}$\]$_{2}$ (dmit-131) in $\kappa /T$ versus $T^{2}$ and $\kappa /T$ versus $T$ (inset) plots. The data of other insulators, i.e., Et$_{2}$Me$_{2}$Sb\[Pd(dmit)$_{2}$\]$_{2}$ (dmit-221, non-magnetic) and $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$, are also plotted for comparison.[]{data-label="MYamashita_TC2"}](MYamashita_TC2.eps){width="8cm"} Kagome-lattice system: ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$ --------------------------------------------------- The kagome lattice is constructed by using corner-sharing triangles in contrast to the edge-sharing in the triangular lattices, as shown in Fig. \[Kagome-classical-GS\]. Thus, the spin states in the kagome lattice have larger degeneracy than those in the triangular lattices, leading to high potential for hosting a spin liquid. Actually, the theoretical perspective of seeking a spin liquid is more promising for the kagome-lattice Heisenberg model than for the triangular lattice [@Sachdev92; @Lecheminant97; @Mila98; @Misguich04]. Among several candidates for the kagome spin systems, we select a spin-1/2 system, ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$, which is known as herbertsmithite, whose magnetism has been extensively investigated. This is a member of a family of materials with variable compositions, i.e., Zn$_{x}$Cu$_{4-x}$(OH)$_{6}$Cl$_{2}$ ($0<x<1$ ). As an end material, Cu$_{4}$(OH)$_{6}$Cl$_{2}$ has a distorted pyrochlore lattice of $S=1/2$ Cu$^{2+}$ spins, whereas the other end material, ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$, has a two-dimensional ($a-b$ plane) perfect kagome-lattice of Cu$^{2+}$ spins separated by different crystallographic sites occupied by Zn$^{2+}$ [@Shores05]. The structural symmetry changes across $x=0.33$, above which Cu$^{2+}$ partially occupies the Zn sites in addition to the kagome lattice. There is an argument for the mixture of Zn in the kagome sites in ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$. Magnetic susceptibility [@Bert07] and specific heat [@Vries08] suggest that approximately $6\%$ of the kagome sites are replaced by non-magnetic Zn. The same amount of Cu is assumed to invade the nominal Zn sites. Thus, significant efforts have been made to extract the intrinsic properties of the kagome lattice from the experimental data. ![[@Mendels07] Temperature variation of the spin-frozen fraction determined by muon spin rotation experiments for Zn$_{x}$Cu$_{4-x}$(OH)$_{6}$Cl$_{2}$. Inset shows the $x$- dependence of the spin-frozen fraction at a low temperature.[]{data-label="Mendels_SFf"}](Mendels_SFf.eps){width="8cm"} Experimental evidence for the absence of magnetic ordering in ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$ can be obtained from $\mu$SR experiments [@Mendels07]. The relaxation profile shows no internal field down to 50 mK. The experiments for a wide range of $x$ found that the absence of an internal field was persistent in a certain range below $x=1$ (see Fig. \[Mendels\_SFf\]) [@Mendels07]. The magnetic susceptibility exhibits a Curie-Weiss behavior at high temperatures above 100 K, as shown in Fig. \[Helton\_sus\] [@Kagome07]. The Weiss temperature is $\sim$300 K, which implies an antiferromagnetic exchange interaction of $J=17$ meV. The dc and ac magnetic susceptibility indicates no magnetic ordering down to 0.1 K and 0.05 K, respectively, which is four orders of magnitude lower than $J$ [@Kagome07]. The susceptibility increases progressively at lower temperatures. Two mechanisms are possible. First, impurities from Cu/Zn inter-site mixing can give a Curie-like upturn. Second, Dzyaloshinsky-Moriya interactions may be present between the adjacent sites with broken inversion symmetry, as in the kagome lattice [@Rigol07]. The high-field magnetization measurements suggest that the former is mainly responsible for the increasing susceptibility [@Bert07]. ![[@Kagome07] Temperature dependence of the inverse magnetic susceptibility $\chi ^{-1}$ of ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$. The line denotes a Curie-Weiss fit. Inset: ac susceptibility (at 654 Hz) at low temperatures.[]{data-label="Helton_sus"}](Helton_sus.eps){width="8cm"} ![[@Olariu08] $^{17}$O NMR shift of two lines (M and D) decomposed from the observed spectra for a powder of ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$. The M and D lines are considered to come from the oxygen sites depicted in the inset. The red curve represents the trace of a half of the value of the M line. The sketch in the lower left corner illustrates the environment of a Zn substituted on the Cu kagome plane, and thick lines represent Cu-Cu dimers.[]{data-label="Olariu_17O_NMR_shift"}](Olariu_17O_NMR_shift.eps){width="8cm"} NMR, which probes magnetism in a site-selective manner, was informative particularly for this material because the analysis of spectra allows one to distinguish the intrinsic magnetism from the extrinsic one. The NMR spectra at $^{35}$Cl and $^{17}$O sites are broad [@Imai08; @Olariu08], reflecting the inhomogeneous local fields, supposedly due to the Zu/Cu mixture. However, the smallest shift value in the broad $^{35}$Cl spectrum follows a Curie-Weiss law down to 25 K, followed by a decrease at lower temperatures [@Imai08]. This is considered to indicate intrinsic magnetism for the kagome lattice [@Imai08]. The $^{17}$O probes the kagome sites more preferentially than the nominal Zn sites due to larger hyperfine coupling with the kagome sites [@Olariu08]. The $^{17}$O NMR spectra were decomposed into two components. One is from the $^{17}$O sites coordinated by two Cu$^{2+}$ ions, while the other is from the $^{17}$O sites coordinated by a Cu$^{2+}$ and a Zu$^{2+}$ in the kagome plane. The relative fraction of the two components was consistent with a 6 % Zn admixture. The NMR shifts of the respective components, as shown in Fig. \[Olariu\_17O\_NMR\_shift\], are considered to be local susceptibilities at Cu sites with and without Zn$^{2+}$ at the neighboring sites [@Olariu08]. Both decrease below 50 K and saturate to finites values, indicating the gapless nature of the spin excitations. The low-temperature decrease in the shift at the Cu site with a Zn neighbor is in contrast to the enhancement commonly observed in the neighborhood of non-magnetic impurities [@Olariu08]. This behavior also suggests that the Curie-like upturn in the bulk susceptibility at low temperatures is not from the kagome plane. For the NMR relaxation rate, all of the O, Cl and Cu nuclear spins exhibit power-laws against temperature down to 0.47 K for O and 2 K or lower for Cl and Cu, indicating a gapless spin liquid (see Fig. \[Imai\_Olariu\_NMR\_T1\]) [@Imai08; @Olariu08]. Although the power somewhat depends on the nuclear site, the relaxation profile is overall nuclear site-insensitive, which is filtered by the nuclear site-specific form-factor determined by its location relative to the kagome lattice, suggesting non-dispersive spin dynamics. Otherwise, the temperature profile of the relaxation rate would be site-dependent [@Olariu08]. This feature is potentially relevant to the spinon excitation with the continuum. More recently, NMR experiments performed at low temperatures have revealed an anomaly in the relaxation rate at a temperature depending on the applied field, which may signify field-induced spin freezing [@Jeong11]. Very recently, a $^{17}$O NMR experiment performed with a single crystal has found different features from those observed so far in the powder samples [@Fu15]. According to the analysis of NMR spectra, there is no significant contamination of Zn in the Cu sites within the kagome plane, and the Knight shift shows appreciable temperature- and field-dependences, suggesting a spin gap of the order of 10 K, as shown in FIG \[Fu\_Imai\_NMR\_2015\], in contradiction with the consequences of the earlier NMR and neutron (see below) experiments. ![[@Olariu08] $^{17}$O, $^{63}$Cu and $^{35}$Cl nuclear spin-lattice relaxation rates $1/T_{1}$ for a powder of ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$. Inset shows $^{17}$O $1/T_{1}$ versus $1/T$ plots.[]{data-label="Imai_Olariu_NMR_T1"}](Imai_Olariu_NMR_T1.eps){width="8cm"} ![[@Fu15] (a) Temperature dependence of $^{17}$O Knight shift and (b) the field dependence of the spin gap deduced from the Knight shift for a single-crystal ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$.[]{data-label="Fu_Imai_NMR_2015"}](Fu_Imai_NMR_2015.eps){width="8cm"} ![[@Kagome07] (a) Specific heat $C$ of ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$ in various applied fields. Inset shows $C$ over a wider temperature range in applied fields of 0 T (square) and 14 T (star). (b) $C$ in a zero field at low temperatures. The lines represent power law fits.[]{data-label="Helton_specific_heat"}](Helton_specific_heat.eps){width="8cm"} The low-temperature specific heat was investigated under external magnetic fields [@Kagome07; @Vries08]. As shown in Fig. \[Helton\_specific\_heat\](a) [@Kagome07], there is an enormous field dependence. The temperature dependence at a zero field is approximated by a power law $T^{\alpha}$ with $\alpha$ unity or smaller (see Fig. \[Helton\_specific\_heat\](b)). The broad peak present even at a zero field is shifted to higher temperatures under higher fields. Assuming that the field-dependent peak is a Schottky contribution associated with a field-induced gap, the data for different fields were analyzed in detail to reveal the intrinsic specific heat of the kagome lattice [@Vries08]. The deduced Schottky component is consistent with Zeeman splitting of the $6\%$ Cu impurities in the Zn sites at higher fields, and the intrinsic $C/T$ follows a power law $T^{\alpha}$ with $\alpha=1.3$ as the best estimate, suggesting gapless excitations [@Kagome07; @Vries08; @Shaginyan11]. Neutron-scattering experiments, which are capable of profiling spin excitations with respect to momentum and energy transfers, are available for herbertsmithite. One of the key issues of elementary excitations in spin liquids is the possible fractionalization of $S=1$ spin excitations into $S=1/2$ spinons, which could manifest themselves as a continuum in the spin excitation spectrum, *i.e.*, dynamic structure factor $S(\textbf{Q}, \omega )$, where **Q** and $\omega$ are momentum transfer and energy transfer divided by $\hbar$, respectively. Such a continuum is observed in a highly anisotropic triangular-lattice system, Cs$_{2}$CuCl$_{4}$, ($J^{\prime}/J \sim$ 3 and $J^{\prime}\sim$ 0.34 meV in Fig. 17), although it undergoes a magnetic transition into a spin-spiral order at 0.62 K [@Coldea01; @Coldea03]. Several features signifying the continuum are found via neutron experiments of herbertsmithite, which were first performed for polycrystalline or powder samples. The inelastic scattering experiments exhibit no excitation gap at least down to 0.1 meV, which corresponds to $\sim J/170$, and insensitivity of the scattering strength to **Q**, indicating gapless and local natures of spin fluctuations [@Kagome07]. Furthermore, the scattering intensity is only weakly dependent on $\omega$ up to 25 meV and temperature up to 120 K and shifts toward lower **Q** as temperature is increased [@Vries09]. Some of the results are displayed in Fig. \[deVries\_Neutron\]. All these features are suggestive of a continuum in spin excitations and the persistence of the short-range nature of spin correlations even at low temperatures. Recent experiments on a large single crystal have succeeded in capturing the continuum nature, as seen in the green area in Fig. \[Han\_Neutron\]. The momentum profile of the excitation intensity (dynamic structure factor), $S(\textbf{Q}, \omega )$, is approximately reproduced by the simulated structure factor of uncorrelated dimer-singlets, which indicates to the short-ranged spin correlations at least down to 1.6 K [@Han12]. The short-range nature that persists even at low temperatures, as suggested by the powder experiments as well, is generally in favor of a gapped state, whereas there is no indication of a spin gap down to 0.25 meV at any **Q** values in the excitation spectra [@Han12]. It is puzzling that spin dynamic correlation exhibits short-range RVB nature while the spectrum is gapless. One possibility is that the Herbertsmithite is in a $Z_2$ spin liquid in close proximity to a critical point to the $U(1)$ Dirac liquid, as indicated by some recent numerical works [@LiTao16], although the true ground state of the isotropic Heisenberg model on a kagome lattice is still under debate [@Iqbal16comment]. ![[@Vries09] (a) Instantaneous magnetic correlations at 4 K and 10 K for a time scale corresponding to approximately 6.5 meV. The solid lines are a guide to the eye. (b) The $Q$ dependence in the dynamic correlations with the energy integration interval indicated in the legend. The dotted line in panel (a) and (b) is the structure factor for dimer-like AF correlations. The dashed line, a single-ion contribution corresponding to the 6$\%$ antisite spins in this system, is added. (c) The energy and temperature dependence at $Q$=1.3 Å$^{-1}$. D7, IN4 and MARI in the legends stand for the types of spectrometers used.[]{data-label="deVries_Neutron"}](deVries_Neutron.eps){width="8cm"} ![[@Han12] Contour plot of dynamical structure factor, $S_{mag} (\textbf{Q}, \omega )$, integrated over 1$\leq \hbar \omega \leq$ 9 meV for a single-crystal ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$ at 1.6 K. The intense scattering is extended in a green-colored region, without peaking at any specific points.[]{data-label="Han_Neutron"}](Han_Neutron.eps){width="8cm"} Hyperkagome-lattice system: Na$_{4}$Ir$_{3}$O$_{8}$ --------------------------------------------------- The hyperkagome lattice is a three-dimensional network of corner-sharing triangular lattices. In Na$_{4}$Ir$_{3}$O$_{8}$, the Ir$^{4+}$ ion with $5d^{5}$ electrons likely takes on a low-spin state. These ions locate on the corners, forming a S=1/2 hyperkagome lattice [@Na4Ir3O8]. The resistivity of the ceramic sample is 10 Ohmcm at room temperature. The samples are semiconducting, with a charge transport gap of 500 K, implying the proximity of this system to the Mott transition, which is different from the kagome materials reviewed above [@Na4Ir3O8]. A connection between the spin liquid and the metal-insulator transition, similar to the case of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$, is shown [@Podolsky09]. A distinct feature of Na$_{4}$Ir$_{3}$O$_{8}$ among spin liquid candidates is its large spin-orbit coupling, which introduces additional interest to the physics of spin liquids [@ChenBalents08; @Zhou08]. Several theoretical studies propose that Na$_{4}$Ir$_{3}$O$_{8}$ is a 3D QSL with fermionic spinons [@Zhou08; @Lawler08]. Fig. \[Okamoto\_sus\_specific\_heat\](a) shows the magnetic susceptibility of Na$_{4}$Ir$_{3}$O$_{8}$, which weakly increases with decreasing temperature, as characterized by the Curie-Weiss temperature of -650 K [@Na4Ir3O8]. This implies an antiferromagnetic interaction of hundreds of Kelvin. There is no clear indication of magnetic ordering at least down to 2 K, whereas a small anomaly reminiscent of spin glass observed in the magnetization history against the field/temperature variation is attributed to a tiny fraction of the total spins [@Na4Ir3O8]. ![[@Na4Ir3O8] (a) Temperature dependence of the inverse magnetic susceptibility $\chi^{-1}$ of polycrystalline Na$_{4}$Ir$_{3}$O$_{8}$ under 1 T. Inset shows magnetic susceptibility $\chi$ in various fields up to 5 T; for clarity, the curves are shifted by 3, 2, and 1 $\times$ 10$^{-4}$ emu/mol Ir for 0.01, 0.1, and 1 T data, respectively. (b) Magnetic specific heat $C_{m}$ divided by temperature $T$ of polycrystalline Na$_{4}$Ir$_{3}$O$_{8}$. To estimate Cm, data for Na$_{4}$Sn$_{3}$O$_{8}$ is used as a reference of the lattice contribution. Inset shows $C_{m}/T$ versus $T$ in various fields up to 12 T. (c) Magnetic entropy. []{data-label="Okamoto_sus_specific_heat"}](Okamoto_sus_specific_heat.eps){width="8cm"} ![[@Singh13] Thermal conductivity $\kappa$ of Na$_{4}$Ir$_{3}$O$_{8}$ in $\kappa /T$ versus $T^{2}$ plots for magnetic fields of 0 T and 5 T. Inset shows the low-temperature part of the data.[]{data-label="Singh_thermal_conductivity"}](Singh_thermal_conductivity.eps){width="8cm"} The electronic (magnetic) contribution to the specific heat of Na$_{4}$Ir$_{3}$O$_{8}$, as shown in Fig. \[Okamoto\_sus\_specific\_heat\](b), has a broad peak at 20 K. However, no anomaly signifying magnetic ordering is apparent [@Na4Ir3O8]. The magnetic entropy estimated by integrating the $C/T$ in Fig. \[Okamoto\_sus\_specific\_heat\](b) reaches $70-80\%$ of $R\ln 2$ ($=5.7$ J/molK) at 100 K, a much lower temperature than the Weiss temperature of $\sim$600 K, which features frustrated magnetism. The $C/T$ is characterized by a curious $T^{2}$ dependence at the lowest temperatures. The $\gamma$ term, when present, appears on the order of 1 mJ/K$^{2}$mol Ir. Recent experiments extended down to 500 m K have found that $C_{\textrm{m}}/T$ is well approximated by a form of $\gamma + \beta T^{2.4}$ with $\gamma =$2.5 mJ/K$^{2}$molIr [@Singh13]. As seen in the inset of Fig. \[Okamoto\_sus\_specific\_heat\](b), the applied field has no influence on the specific heat, at least up to 12 T. The temperature dependence of thermal conductivity is shown in Fig. \[Singh\_thermal\_conductivity\] [@Singh13]. At low temperatures down to 75 mK, $\kappa/T$ is linear in $T^{2}$. The $\kappa /T$ value extrapolated to $T=0$ is $6.3 \times 10^{-2}$ mW/K$^{2}$m, which is a vanishingly small value, compared with the value of EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$, 0.2 W/K$^2$m in Fig. \[MYamashita\_TC2\]. The suppression of the $\kappa /T$ value by the extrinsic grain-boundary effect is not ruled out [@Singh13]. The feature that $\gamma$ is diminished and $\kappa /T$ is vanishing at low temperatures, while both are sizable at high temperatures of the order of Kelvin, appears to be in accordance with a theoretical picture of spinon Fermi surfaces that undergo a pairing instability at low temperatures [@Zhou08]. In this context, the magnetic susceptibility, remaining large even at low temperatures, can be due to the large spin-orbit interactions of Ir [@Zhou08]. The substitution of non-magnetic Ti$^{4+}$ ions at Ir sites will give rise to a Curie-like tail in the spin susceptibility curve [@Na4Ir3O8], similar to Zn substitution for Cu in high-Tc cuprates, indicating an RVB spin background. The scaling analysis of magnetic Gruneisen parameters is suggestive of the proximity of Na$_{4}$Ir$_{3}$O$_{8}$ to a zero-field quantum critical point [@Singh13]. Very recent $\mu$SR [@Dally14] and NMR experiments [@Shockley15] have found some indications that are not in accordance with the above claims. Both probes detected the emergence of local fields signifying the freezing of moments at low temperatures, as shown in Fig. \[Shockley\_NMR\]. The muons are revealed to sense an inhomogeneous local field of electronic origin that appears at 6 K, where the irreversibility in magnetization occurs, and levels off to 70 G on average, which may correspond to 0.5 $\mu_{\textrm{B}}$ on Ir. It is suggested, however, that the spin correlation is short-ranged (of the order of one unit-cell) and quasi-static in that the slow dynamics captured by the relaxation rate persist down to 20 mK. The quasi-static nature is also seen in the S=1 triangular-lattice system, NiGa$_{2}$S$_{4}$ [@Nakatsuji05; @MacLaughlin08]. $^{17}$O and $^{23}$Na NMR lines show broadening, which is roughly scaled to the $\mu$SR results at low temperatures, as seen in Fig. \[Shockley\_NMR\]; the moment is estimated at 0.27 $\mu_{\textrm{B}}$ on Ir. The NMR line profile also suggests inhomogeneous spin freezing and slow dynamics persisting down to low temperatures although the temperature dependence of the relaxation rates on the muon and $^{23}$Na differ. Noticeably, the $^{23}$Na relaxation rate exhibits a peak indicative of the critical slowing down at approximately 7.5 K despite no anomaly in specific heat. The nature and origin of these anomalous properties are not clear at present; however, it is likely that disorder plays a vital role in this system, which can host configurationally degenerate phases with fluctuating order [@Dally14]. Considering that muon, $^{17}$O and $^{23}$Na captured the behavior of the majority of spins in the sample, the disorder effect, if any, is such that it is not restricted to finite areas but extended over the system, being reminiscent of the quantum Griffiths effect given the inhomogeneity and slow dynamics. ![[@Shockley15]The line width (FWHM) of Gaussian-broadened $^{17}$O and $^{23}$Na NMR spectra and the mean value of the distributed local fields detected based on $\mu$SR [@Dally14]. For the NMR line width, its deviation from the value at 15 K is plotted. Inset: $^{23}$Na spectra at 78.937 MHz for 7 T (empty circles) and 45.046 MHz for 4 T (solid line) with the horizontal axis shifted by 3.005 T at 1.3 K. The blue line shows the expected powder pattern of the spectrum, with every Ir-site carrying the same moment. []{data-label="Shockley_NMR"}](Shockley_NMR.eps){width="8cm"} **Material** Triangular, $\kappa$**-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$** Triangular **M\[Pd(dmit)$_{2}$\]$_{2}$** Kagome **ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$** Hyper-Kagome, **Na$_{4}$Ir$_{3}$O$_{8}$** -------------------------- ------------------------------------------------------- ------------------------------------------ -------------------------------------------------------- ------------------------------------------- **Susceptibility** A broad peak at 60 K, A broad peak at 50 K, Curie-Weiss at high-T Curie-Weiss Finite at 2 K, $J=$250 K Finite at 2 K, $J=220 \sim$ $\Theta_{\textrm{W}}=$ -300 K, $J=$230 K, $\Theta_{\textrm{W}}=$ -650 K (\*1) 280 K (\*7) Upturn at low-T due to impurity sites (\*19, \*20) (\*11, \*12) **Specific heat** Gapless, Gapless, Gapless, Gapless, $\gamma =$15 mJ/K$^{2}$ mol, $\gamma$=20 mJ/K$^{2}$mol, $C \sim T^{\alpha}$ $C \sim T^{2}$ (\*19), Field-independent Field-independent $\alpha = 1.3$ at high fields $C \sim \gamma T + \beta T^{2.4}$, (\*2) (\*8) (\*13) $\gamma =2$ mJ/K$^{2}$mol (\*20), Field-independent (\*21, \*22) **Thermal conductivity** Gapped; $\Delta =$ 0.46 K (\*3) Gapless; finite $\kappa/T$ (\*9) Vanishingly small $\kappa /T$ (\*22) **NMR shift** Not precisely resolved Not precisely resolved High-T $^{17}$O shift — scales to (\*4) (\*10) Broad peak at at 50-60 K $\chi_{\textrm{bulk}}$ in 100 K - 300 K for $^{17}$O (\*14,\*15), but levels off below at 25-50 K for $^{35}$Cl (\*16) 80 K (\*23) Low-T $^{17}$O,$^{23}$Na- gapless :finite value inhomogeneous line (\*14) broadening at low-$T$ gapped : $\Delta \sim$ 10 K (\*15) (\*23) **NMR $1/T_{1}$** Inhomogeneous $1/T_{1}$, Inhomogeneous $1/T_{1}$, $1/T_{1} \sim T^{\alpha}$ $^{23}$Na $1/T_{1}$– a peak Power law, Power law, $\alpha \sim 0.73$ for $^{17}$O (\*14) formation typical of $^{1}$H $1/T_{1};\sim T$ / $\sim T^{2}$ at $^{13}$C $1/T^{2}$ at $<$ 0.5 K $\alpha \sim$ 0.5 for $^{63}$O (\*16) critical slowing down at $T< 0.3 $K (stretched exponential) 7.5 K (two components) (\*1), (\*10) Field-induced spin (\*23) $^{13}$C $1/T_{1};\sim1/T^{1.5}$ at $T< 0.2$ K freezing (\*17) (stretched exponential) (\*4) **$\mu$SR** No internal field at 0 T No internal field at 0 T Emergence of distributed (\*5,\*6) (\*18) local fields below 6 K Quasi-static short-ranged spin freezing with slow dynamics (\*24) **Neutron** Powders $\sim$ gapless ($<$0.1 meV) (\*11,\*19) Single crystal $\sim$ gapless ($<$0.25 meV) (\*20) Continuum in dynamic structure factor (\*11,\*19,\*20) **References** \*1 Shimizu *et al.*, 2003, \*7 Kato, 2014, \*11 Helton *et al.*, 2007, \*21 Okamoto *et al.*, 2007 \*2 Yamashita *et al.*, 2008 \*8 Yamashita *et al.*, 2011, \*12 Bert *et al.*, 2007, \*22 Singh *et al.*, 2013 \*3 Yamashita *et al.*, 2009, \*9 Yamashita *et al.*, 2010, \*13 de Vries *et al.*, 2008, \*23 Shockley *et al.*, 2015 \*4 Shimizu *et al.*, 2006, \*10 Itou *et al.*, 2010 \*14 Olariu *et al.*, 2008, \*24 Dally *et al.*, 2014 \*5 Pratt *et al.*, 2011, \*15 Fu *et al.*, 2015 \*6 Nakajima *et al.*, 2012 \*16 Imai *et al.*, 2008 \*17 Jeong *et al.*, 2011, \*18 Mendels *et al.*, 2007 \*19 de Vries *et al.*, 2009 \*20 Han *et al.*, 2012 : Spin liquid materials summary \[Spin\_liquid\_summary\] Experimental summary -------------------- Due to intensive experimental studies, unconventional thermodynamic and magnetic properties that evoke spin liquids have been found in several materials with anisotropic triangular lattices, kagome lattices and hyperkagome lattices as seen above. These materials exhibit no indications of conventional magnetic ordering. Their magnetic and thermodynamic properties are summarized in Table \[Spin\_liquid\_summary\]. It appears that the gapless nature is a property that a class of frustrated lattices constructed with triangles possesses, although the thermal conductivity of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ suggested a tiny excitation gap three orders of magnitude smaller than $J$. A recent NMR work on herbertsmithite insists on gapped spin excitations, and anomalous quasi-static spin freezing has recently been revealed by $\mu$SR and NMR studies of the hyperkagome system. This feature and the successful observation of fractionalized excitations in a kagome lattice [@Han12] tempt ones to think about spinons as promising elementary excitations in spin liquids. How to detect the spinon Fermi surfaces, if they exist, is a focus—- smoking-gun experiments are awaited. As seen in Table \[Spin\_liquid\_summary\], several experimental characteristics are seemingly inconsistent within given materials; understanding the apparently contradicting data in a consistent way requires clarification of the nature of the spin states. One of the key issues may be the randomness present in real materials. In particular, it has long been recognized that the effect of inevitable Zn/Cu admixtures in herbertsmithite has to be separated from the intrinsic magnetism. More recently, the issue of inhomogeneous quasi-static spin correlation with slow dynamics in the hyperkagome-lattice system has emerged as a consequence of disorder. Theoretically, it was proposed that as randomness is intensified, the 120-degree Neel order in the triangular-lattice Heisenberg model is changed to a sort of random singlets but not spin glass state. It is intriguing that randomness appears to enhance the quantum nature because the singlet is a purely quantum state [@Watanabe14; @Shimokawa2015]. In the case of kagome lattices, it was theoretically suggested that disorder could lead to a valence-bond glass state[@Singh10] or a gapless spin liquid state [@Shimokawa2015; @Kawamura2014]. Furthermore, a recent NMR experiment on an organic Mott insulator, i.e., $\kappa$-(ET)$_{2}$Cu\[N(CN)$_{2}$\]Cl, found that the antiferromagnetic ordering in the pristine crystal, when irradiated by X-rays, disappears. Spin freezing, spin gap and critical slowing down are not observed, but gapless spin excitations emerge, suggesting a novel role of disorder that brings forth a QSL from a classical ordered state [@Furukawa15b]. Whether the randomness is fatal or vital to the physics of a QSL is a non-trivial issue to be resolved. The development of new materials, although not addressed in this article, is under way. Among them is a new type of hydrogen-bonded $\kappa$-H$_{3}$(Cat-EDT-TTF)$_{2}$ with a triangular lattice of one-dimensional anisotropy [@Isono13] and $\kappa$-(ET)$_{2}$Ag$_{2}$(CN)$_{3}$, an analogue of $\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ [@GSaito14]. Another compound with hyperkagome lattice structure, i.e., PbCuTe$_2$O$_6$, with Curie-Weiss temperature $\theta=-22$K is also proposed to be a spin liquid candidate [@Koteswararao14; @Khuntia16]. The entanglement of additional degrees of freedom with quantum spins may be another direction for future studies; e.g., Ba$_{3}$CuSb$_{2}$O$_{9}$ is proposed to host a spin-orbital coupled liquid state [@HDZhou2011; @Nakatsuji12]. It should be emphasized that the identification of QSL experimentally is a very important and challenging task. As a “featureless" Mott insulator, there exists no simple magnetic order for identifying QSL states, and so far, there exists only indirect experimental evidence for mobile fermionic spinons in some candidate compounds as discussed above. To remedy this situation, theorists have proposed new experiments to identify QSLs through identifying nontrivial properties of spinons and gauge fields. For example, power law AC conductivity inside the Mott gap has been noted [@NgLee07]. A giant-magnetoresistance-like experiment was proposed to measure mobile spinons through oscillatory coupling between two ferromagnets via a QSL spacer [@Norman2009]. The thermal Hall effect in insulating quantum magnets was proposed as a probe for the thermal transport of spinons, where different responses were used to distinguish between magnon- and spinon- transports [@KNL2010]. Raman scattering was proposed as a signature to probe the $U(1)$ QSL state [@Ko10]. It was also proposed that the spinon life time and mass as well as gauge fluctuations can be measured through a sound attenuation experiment [@YZhou2011], and neutron scattering can be used to detect scalar spin chirality fluctuations in the kagome system [@Lee13]. Low energy electron spectral functions were evaluated for future ARPES experiments [@Tang13] and it was proposed that spin current flow through a metal-QSL-metal junction can be used to distinguish different QSLs [@Chen13]. More recently, it was suggested that there exists a long-life surface plasmon mode propagating along the interface between a linear medium and a QSL with spinon Fermi surface at frequencies above the charge gap, which can be detected by the widely used Kretschmann-Raether three-layer configuration [@Ma15]. However, there exists an important discrepancy between existing experiments and theories in some of the above experiments. 1\) Specific heat: Using the one-loop calculation supplemented by scaling analysis [@LeeNagaosa92; @Polchinski94], it is found that the strong coupling between the $U(1)$ gauge field and spinon Fermi surface leads to $T^{2/3}$ correction to the temperature dependence of specific heat in $U(1)$ gauge theory. This predicted $T^{2/3}$ behavior has never been observed in experiments. Instead, linear, Fermi-liquid-like specific heat is found to exists in a wide range of temperatures in both organic materials ($\kappa$-ET and dmit). Some theories exsit that try to explain this missing singular $T^{2/3}$ specific heat. For instance, $Z_4$ and $Z_2$ spin liquid states with a spinon Fermi surface have been proposed [@Barkeshli13] as well as $Z_2$ spin liquid states with quadratic touched spinon bands [@Mishmash13]. However, all these proposals require fine-tuned parameters. A more natural way of explaining existing experiments is still missing. 2\) Thermal Hall effect: Katsura, Nagaosa and Lee [@KNL2010] have theoretically investigated the thermal Hall effect induced by the external magnetic field in a $U(1)$ spin liquid with a spinon Fermi surface and have predicted measurable electronic contributions. Their predicted sizable thermal Hall effect have never been observed in an experiment on dmit compounds [@MYamashita10]. This contradiction between experiment and theory remains unsolved, although an explanation that depends on fine-tuned parameters has been proposed [@Mishmash13]. 3\) Power law AC conductivity: A power law AC conductivity inside the Mott gap was proposed by Ng and Lee [@NgLee07]. Indeed, power law behavior $\sigma(\omega)\sim \omega^{\alpha}$ has been observed in both $\kappa$-ET [@Elsasser12] and Herbertsmithite [@Pilon13]. However, the power $\alpha$ observed in both compounds is smaller than predicted value, indicating that there exist more in-gap electronic excitations than those predicted in the $U(1)$ gauge theory. Thus, despite all the theoretical efforts, the understanding and finding of realistic “smoking gun" evidence for QSLs remains the greatest challenge in the study of QSLs. Summary ======= In this article, we provide a pedagogical introduction to the subject of QSLs and review the current status of the field. We first discuss the semi-classical approach to simple quantum antiferromagnets. We explain how it leads to the Haldane conjecture in one dimension and why it fails for frustrated spin models. We then focus on spin-$1/2$ systems with spin rotational symmetry and introduce the RVB concept and the slave-particle plus Gutzwiller-projected wavefunction approaches. We explain the technical difficulties associated with these approaches and why slave-particle approaches naturally lead to gauge theories for spin liquid states. The natures of $SU(2)$, $U(1)$ and $Z_2$ spin liquid states are explained, and the extensions of the approach to systems with spin-orbit coupling and $S>1/2$ systems are introduced. We explain that because of the intrinsic limitations of the analytical slave-particle approach, many alternative approaches to spin liquid states have been developed, both numerically and analytically. These approaches complement each other and often lead to exotic possibilities not covered by the simple fermionic slave-particle approach. The experimental side of the story is also introduced with a review of the properties of several candidate spin liquid materials, including anisotropic triangular lattice systems ($\kappa$-(ET)$_{2}$Cu$_{2}$(CN)$_{3}$ and EtMe$_{3}$Sb\[(Pd(dmit)$_{2}$\]$_{2}$), kagome lattice systems (ZnCu$_{3}$(OH)$_{6}$Cl$_{2}$) and hyperkagome lattice systems (Na$_{4}$Ir$_{3}$O$_{8}$). We note several outstanding difficulties with attempts to explain experimental results using existing theories. [*These difficulties indicate that the field of QSLs is still wide open and immature and that important physics may still be missing in our present understanding of QSLs*]{}. While keeping the article at an introductory level, we are not able to cover many important developments in the study of spin liquid states, and many technical details have been neglected, both theoretically and experimentally. For example, the important techniques of renormalization groups and conformal field theory are not addressed in this article. We also do not discuss in detail the many developments related to MPSs and/or PEPSs and the corresponding numerical DMRG technique, the understanding of spin systems with broken rotational symmetry following the discovery of the Kitaev state, and the spin liquid physics of $S>1/2$ systems. The role of topology in spin liquid states is not touched upon except as it is relevant to examples of spin liquid states. These are rapidly evolving areas in which new discoveries are expected. In the following section, we outline a few other topics that are neglected in this article but either have played important historical roles in the development of the field of QSLs or shed light on future research: [*Quantum dimer models:*]{} Quantum dimer models (QDMs) are a class of models defined in the Hilbert space of nearest neighbor valence bond (or dimer) coverings over a lattice instead of the spin Hilbert space [@Rokhsar88]. QDMs can be obtained in certain large-$N$ limits of $SU(N)$ or $Sp(N)$ antiferromagnets [@Read89b] and provide a simplified description of RVB states. This simplification allows researcher to proceed further in analytical treatments because of the close relations that arise to classical dimer problems, Ising models and Z$_2$ gauge theory [@Kasteleyn61; @Fisher61; @Kasteleyn63; @Moessner02; @Misguich02a; @Moessner03]. However, by construction, QMDs focus on the dynamics in the spin-singlet subspace and ignore spin-triplet excitations. Therefore, they are not directly relevant to spin systems in which the magnetic excitations are gapless. An advantage of QDMs is that some QDMs are exactly solvable [@Misguich02a; @Yao12]. Thus, many issues related to QSLs that are difficult to address, such as spinon deconfinement, $Z_2$ vortices and topological order, can be addressed explicitly in QDMs. Interestingly, some spin-$1/2$ Hamiltonians give rise to sRVB ground states defined in the dimer Hilbert space when the relationship between the spin and dimer configurations is properly chosen [@Fujimoto05; @Seidel09; @Cano10]. Readers who are interested in further details on QDMs can refer to Chapter 5.5 in reference [@BookDiep] and Chapter 17 in reference [@BookLacroix]. [*Chiral spin liquids:*]{} QSL states that break the parity (P) and time-reversal (T) symmetries while conserving the spin rotational symmetry have been proposed by Kalmeyer and Laughlin [@Kalmeyer87; @Kalmeyer89]. These states are called chiral spin liquids. Kalmeyer and Laughlin proposed that some frustrated Heisenberg antiferromagnets in 2D can be described by bosonic fractional quantum Hall wavefunctions. Soon afterward, Wen, Wilczek and Zee [@Wen89] introduced a generic method of describing chiral spin liquids. They suggested that chiral spin states can be characterized in terms of the spin chirality $E_{123}=\vec{S}_1\cdot(\vec{S}_2\times\vec{S}_3)$, defined for three different spins, $\vec{S}_1$, $\vec{S}_2$ and $\vec{S}_3$. The expectation value of the spin chirality in fermionic RVB theory is given by $\langle E_{123}\rangle={1\over 2}\rm{Im} \langle\chi_{12}\chi_{23}\chi_{31}\rangle$, where the $\chi_{ij}$ are the short-range order parameters defined in Eq. . Exactly solvable Hamiltonians hosting both gapful chiral spin liquid states [@Laughlin89; @Yao07; @Schroeter07; @Thomale09] and gapless chiral spin liquids [@Chua11] have been found. There is also numerical evidence for chiral spin liquids on some 2D frustrated lattices [@Sorella03; @nielsen13; @bauer13; @He14; @He14a; @Gong14; @Gong15; @Zhu15]. It has been suggested that the statistics of spinons in these chiral spin liquid states can be non-Abelian; see, e.g., [@Yao07; @Greiter09]. [*Characterizing spin liquid states numerically:*]{} Because of rapid advancements in the power of numerical approaches to spin models, the characterization of spin liquid states for specific spin models from numerical data has become a rapidly evolving field. In addition to the MPS and/or PEPS approach and the corresponding numerical DMRG technique, Tang and Sandvik developed a quantum Monte Carlo method of characterizing spinon size and confinement length in quantum spin systems, which allows the spinon confinement-deconfinement issue to be studied numerically [@TangSandvik13]. Another important achievement is the use of entanglement entropy to characterize QSL states. Readers may consult reference [@Grover13] for a brief review. To conclude, the field of QSLs is still wide open, both theoretically and experimentally. The major difficulty in understanding QSLs is that they are intrinsically strongly correlated systems, for which no perturbative approach is available. Theorists have been using all of the available tools as well as inventing new theoretical tools to understand QSLs with the hope that novel emerging phenomena not covered by perturbative approaches can be uncovered. Thus far, there have been a few successes, and new experimental discoveries and theoretical ideas are rapidly emerging. However, a basic mathematical framework that can be used to understand QSLs systematically is still lacking. We expect that more new physics will be discovered in QSLs, posing a challenge to both theorists and experimentalists to construct a basic framework for the understanding of QSLs. Y.Z. and T.K.N. thank Patrick A. Lee, Zheng-Xin Liu, Naoto Nagaosa, Shaojin Qin, Zhaobin Su, Hong-Hao Tu, Tao Xiang, Xiao-Gang Wen, Zheng-Yu Weng, Guang-Ming Zhang, and Fu-Chun Zhang for their close collaboration on related issues over the years. K.K. is grateful to K. Miyagawa, Y. Shimizu, Y. Kurosaki, H. Hashiba, H. Kobashi, H. Kasahara, T. Furukawa, M. Maesato, G. Saito, F. Pratt, and M. Poirier for their collaboration on the topic of spin liquids. We benefited greatly from discussions with our colleagues Yan Chen, Yin-Chen He, Bruce Normand, Fa Wang, Cenke Xu, and Hong Yao. Y.Z. is supported by the National Key R&D Program of China (No.2016YFA0300202), the National Basic Research Program of China under Grant No. 2014CB921201 and by the National Natural Science Foundation of China under Grant No. 11374256. He also wishes to acknowledge the hospitality of the Max Planck Institute for the Physics of Complex Systems in Dresden, where this review article was finalized. K.K. is partially supported by JSPS KAKENHI under Grant Nos. 20110002, 25220709, and 24654101, by the US National Science Foundation under Grant No. PHYS-1066293, and by the hospitality of the Aspen Center for Physics. T.K.N. acknowledges support from HKRGC through Grant No. 603913. Path integral for a single spin =============================== We consider the path integral for a single spin $\mathbf{S}$ in a magnetic field $\mathbf{B}$ ($H=\mathbf{S}\cdot\mathbf{B}$) in the coherent state representation. Spin coherent states are defined as $$\mathbf{\hat{S}}|\mathbf{n}\rangle=S\mathbf{n}|\mathbf{n}\rangle,$$ where $\mathbf{\hat{S}}$ is the spin operator. The path integral can be derived by using the identity operator \[app1\] $$\mathbf{I}=\left({2S+1\over4\pi}\right)\int d^3n\delta(\mathbf{n}^2-1)|\mathbf{n}\rangle\langle\mathbf{n}|=\int D{\mathbf{n}}|\mathbf{n}\rangle\langle\mathbf{n}|$$ and the corresponding inner product $$\langle\mathbf{n}_1|\mathbf{n}_2\rangle=e^{iS\Phi(\mathbf{n}_1,\mathbf{n}_2,\mathbf{n}_0)}\left({1+\mathbf{n}_1\cdot\mathbf{n}_2\over2}\right)^S,$$ where $\mathbf{n}_0$ is a fixed unit vector and is usually chosen to be $\mathbf{n}_0=\hat{z}$, $\Phi(\mathbf{n}_1,\mathbf{n}_2,\mathbf{n}_0)$ is the area of the spherical triangle with vertices $\mathbf{n}_1$, $\mathbf{n}_2$, and $\mathbf{n}_0$, and $S\Phi$ is the Berry’s phase acquired by a particle traveling through a loop formed by the edges of the spherical triangle. The partition function $Z=e^{-\beta H}$ can be written as a path integral using the standard procedure: $$\begin{aligned} \label{app2} Z & = & \lim_{N_t\rightarrow\infty, \delta t\rightarrow 0}\left(e^{-\delta tH}\right)^{N_t} \\ \nonumber & = & \lim_{N_t\rightarrow\infty, \delta t\rightarrow 0}\left(\Pi_{j=1}^{N_t}\int D\mathbf{n}_j\right)\left(\Pi_{j=1}^{N_t} \langle\mathbf{n}_j|e^{-i\delta tH}|\mathbf{n}_{j+1}\rangle\right), \end{aligned}$$ with the periodic boundary condition $|\mathbf{n}(0)\rangle=|\mathbf{n}(\beta)\rangle$. In the limit $\delta t\rightarrow0$, we may approximate $$\begin{aligned} \label{app3} \langle\mathbf{n}_j|e^{-i\delta tH}|\mathbf{n}_{j+1}\rangle & \sim & \langle\mathbf{n}_j|\mathbf{n}_{j+1}\rangle-\delta t\langle\mathbf{n}_j|H|\mathbf{n}_{j+1}\rangle \\ \nonumber & \sim & \langle\mathbf{n}_j|\mathbf{n}_{j+1}\rangle(1-\delta t {\langle\mathbf{n}_j|H|\mathbf{n}_{j+1}\rangle\over\langle\mathbf{n}_j|\mathbf{n}_{j+1}\rangle}) \\ \nonumber & \sim & e^{iS\Phi(\mathbf{n}_j,\mathbf{n}_{j+1},\mathbf{n}_0)}({1+\mathbf{n}_j\cdot\mathbf{n}_{j+1}\over2})^S \\ \nonumber & & \times(1-\delta t S\mathbf{B}\cdot\mathbf{n}_t), \end{aligned}$$ which is valid to the first order in $\delta t$. In deriving the last equality in Eq. , we have made use of the result $\langle\mathbf{n}|\hat{\mathbf{S}}=\langle\mathbf{n}|\mathbf{n}$. Furthermore, we note that $$\begin{aligned} \label{app4} ({1+\mathbf{n}_j\cdot\mathbf{n}_{j+1}\over2})^S & \sim & e^{S\ln(1+{\delta t\over2}\mathbf{n}(t)\cdot\partial_t\mathbf{n}(t))_{t=t_j}} \\ \nonumber & \sim & e^{S\delta t\partial_t[\mathbf{n}(t)]^2}=e^{(0)} \end{aligned}$$ to leading order in $\delta t$. Therefore, $$\begin{aligned} \label{app5} \langle\mathbf{n}_j|e^{-i\delta tH}|\mathbf{n}_{j+1}\rangle & \sim & e^{iS\Phi(\mathbf{n}_j,\mathbf{n}_{j+1},\mathbf{n}_0)-\delta t S\mathbf{B}\cdot\mathbf{n}_t} \end{aligned}$$ and $$\begin{aligned} \label{app6} Z & \sim & \int\mathbf{D}\mathbf{n}(t)e^{iS\Omega(\mathbf{n}(t))-S\int^{\beta}_0 dt\mathbf{B}\cdot\mathbf{n}(t)}, \end{aligned}$$ where $\int\mathbf{D}\mathbf{n}(t)=\lim_{N_t\rightarrow\infty, \delta t\rightarrow 0}\left(\Pi_{j=1}^{N_t}\int D\mathbf{n}_j\right)$ and $$\Omega(\mathbf{n}(t))=\sum_j\Phi(\mathbf{n}_j,\mathbf{n}_{j+1},\mathbf{n}_0)$$ is the total area on the surface of the unit sphere covered by the (closed) path swept out by the spin $\mathbf{n}(t)$ from $t=0$ to $t=\beta$. The classical action of the system in real time is given by \[app7\] $$S_{cl}=S\Omega(\mathbf{n}(t))-S\int^{T}_0 dt\mathbf{B}\cdot\mathbf{n}(t),$$ and the classical equation of motion ${\delta S_{cl}\over\delta'\mathbf{n}(t)}=0$ leads to the Euler equation of motion $$\mathbf{n}\times\left((\mathbf{n}\times\partial_t\mathbf{n})-\mathbf{B}\right)=0,$$ where we have used the result that a small variation $\delta\mathbf{n}$ leads to a change in $\Omega(C[\mathbf{n}])$ that is given by $$\delta\Omega[\mathbf{n(t)})=\int_0^{\beta}dt\delta\mathbf{n}(t)\cdot(\mathbf{n}(t)\times\partial_t\mathbf{n}(t)).$$ [^1]: For $S=1$ chains, the $S=1/2$ end states are protected by a weaker $Z_2\times Z_2$ symmetry [@Chen11; @Chen11a]. [^2]: For historical reasons, the fermion representation is also called the slave-boson representation, and the Schwinger boson representation is also called the slave-fermion representation. In the context of doped Mott insulators, one can decompose the electron annihilation operator as $c_{i\sigma}=h_i^{\dagger} f_{i\sigma}$, where $f_{i\sigma}$ carries a charge-neutral spin and $h_i^{\dagger}$ is the (spinless) hole creation operator. If the spinon operator $f_{i\sigma}$ is fermionic, then the charge carrier ($h_i^{\dagger}$) is a “slave boson", whereas if the spinon operator is bosonic, then the charge carrier is a “slave fermion". [^3]: This criterion for fractionalization works only in dimensions $d>1$. In one dimension, gauge fields are always confining, while spinons appear in energy spectrum as the gapless spin-$1/2$ excitations of the quantum antiferromagnet Heisenberg model [@Mudry94a; @Mudry94b]. [^4]: In general, for a many-spin system in which spin rotational symmetry is broken, the spin $S=0$ state(s) will mix with spin $S\geq1$ states even in the presence of spatial inversion symmetry. The only exception is the two-spin system, in which inversion symmetry provides a good quantum number that separates the spin-singlet state from the spin-triplet states. Because the RVB approach begins from mean-field spin wavefunctions that are superpositions of two-spin pairing states, broken inversion symmetry is needed for the construction of mixed spin-singlet and spin-triplet states.
{ "pile_set_name": "ArXiv" }
--- abstract: | [**Abstract**]{} – Kinetics of homogeneous nucleation and growth of copper precipitates under electron irradiation of Fe$_{1-x}$Cu$_x$ alloys at concentrations $x$ from 0.06 to 0.4at.% and temperatures $T$ from 290 to 450$^{\circ}$C is studied using the kinetic Monte Carlo (KMC) simulations and the statistical theory of nucleation (STN). The conventional assumption about the similarity of mechanisms of precipitation under electron irradiation and during thermal aging is adopted. The earlier-developed $ab \ initio$ model of interactions in Fe-Cu alloys is used for both the KMC simulations and the STN calculations. Values of the nucleation barrier $F_c$ and the prefactor $J_0$ in the Zeldovich-Volmer relation for the nucleation rate $J$ are calculated for a number of concentrations and temperatures. For the dilute alloys with $x\leq 0.2\%$, the STN and the KMC results for the nucleation barrier $F_c$ do virtually coincide with each other, which seems to confirm a high reliability of the STN for this problem. The STN calculations are also used to estimate the temperature dependencies of concentrations which correspond to the homogeneous or the heterogeneous precipitation limit, $x_l^{hom}(T)$ and $x_l^{het}(T)$, and both dependencies are found to be rather sharp. precipitation, dilute Fe-Cu alloys, kinetic Monte Carlo simulations, statistical theory of nucleation author: - | V.G. Vaks$^{\rm a,b}$$^{\ast}$ [^1], F. Soisson$^{\rm c}$, and I.A. Zhuravlev$^{\rm a}$\ $^{\rm a}$[*[National Research Center “Kurchatov Institute”, 123182 Moscow, Russia]{}*]{}; $^{\rm b}$[*[Moscow Institute of Physics and Technology (State University), 117303 Moscow, Russia]{}*]{}; $^{\rm c}$[*CEA, DEN, Service de Recherches de Métallurgie Physique, 91191 Gif-sur-Yvette, France*]{}\ title: 'Studies of homogeneous precipitation in very dilute iron-coper alloys using kinetic Monte Carlo simulations and statistical theory of nucleation' --- Introduction ============ Studies of copper precipitation in irradiated Fe-Cu-based alloys at temperatures $T\sim 290-300^{\circ}$C and low copper concentrations $x$ between $x$=0.05-0.06% and $x$=0.2-0.3% (in atomic percents, here and below) attract great attention, first of all in connection with the problem of hardening and embrittlement of the nuclear reactor pressure vessels (RPVs), see, e. g. [@Akamatsu-95]-[@Meslin-10b]. Numerous experiments combined with some simulations showed that for the usual conditions of electron irradiation, that is, when the radiation-induced clusters of vacancies and self-interstitial atoms are not formed and the heterogeneous precipitation on them is basically absent, SANS and ATP measurements at $x\lesssim 0.1\%$ typically do not reveal any copper-rich precipitates [@Akamatsu-95; @Mathon-97; @Radiguet-07], even though the alloys remain to be strongly supersaturated. The absence of a significant homogeneous copper precipitation in these electron-irradiated Fe-Cu-based alloys is commonly explained “by the very low volume fraction of probably very small precipitates,... and ... one can infer that the limit of observable copper precipitation at 290 $^{\circ}$C is located between $x$=0.09 and 0.08%” [@Mathon-97], or “between $x$=0.1 and 0.2%” [@Auger-00]. In particular, cluster dynamics simulations of homogeneous precipitation at $x$=0.088%, described in [@Radiguet-07] predict “a very low density ($5\cdot 10^{-16}$ m$^{-3}$) of very big ($R\simeq 30$ nm) copper precipitates”. Under neutron or ion irradiations when copper precipitation can occur heterogeneously at clusters of vacancies or self-interstitial atoms, such precipitation at $T=290^{\circ}$C was observed for $x\gtrsim 0.08$ [@Radiguet-07; @Meslin-10a], but not for $x\lesssim 0.06\%$ [@Miller-06; @Miller-09]. Therefore, theoretical studies of copper precipitation in very dilute Fe-Cu-based alloys near the concentration or temperature limits of such precipitation, $x_l^{hom}(T)$ for the homogeneous nucleation and $x_l^{het}(T)$ for the heterogeneous nucleation, seem to be interesting from both fundamental and applied standpoints, even for binary Fe-Cu alloys which can be considered as model alloys for the RPV steels [@Akamatsu-95; @Mathon-97; @Duparc-02; @Radiguet-07]. Such theoretical studies can be particularly interesting if they can sufficiently reliably predict the temperature dependencies of these precipitation limits at elevated $T>290^{\circ}$C for which experimental estimates are either uncertain or absent. A consistent $ab \ initio$ model for studies of precipitation kinetics in Fe-Cu alloys has been recently developed by Soisson and Fu [@SF-07], and detailed kinetic Monte Carlo (KMC) simulations of precipitation in a number of Fe-Cu alloys at different $T$ and $x$ based on this model yield a good agreement with available experimental data [@SF-07; @KVZ-13]. Therefore, this model and KMC methods developed in [@SF-07] seem to be prospective to study the above-discussed problem of copper precipitation in dilute Fe-Cu alloys. We will also use the statistical theory of nucleation and growth of isolated precipitates (to be abbreviated as STN) developed by Dobretsov and Vaks [@DV-98a; @DV-98b]. This theory provides microscopic expressions for both the nucleation barrier (activation barrier for the formation of a critical embryo) $F_c$ and the prefactor $J_0$ in the phenomenological Zeldovich and Volmer formula for the nucleation rate $J$ (the number of supercritical embryos formed in unit volume per unit time) [@Lif-Pit-79] valid for the initial, steady-state nucleation stage of precipitation: $$J=J_0\exp (-\beta F_c) \label{J-J_0}$$ where $\beta =1/T$ is the reciprocal temperature. The STN corresponds to a number of refinements of earlier models of nucleation suggested by Cahn-Hilliard [@Cahn-Hil-59] and Langer [@Langer-69]. It is based on the generalized Gibbs distribution approach [@Vaks-04] and enables one to quantitatively calculate values of $F_c$ and $J_0$ for the microscopic alloy model chosen, particularly in the case of high nucleation barriers, $\beta F_c\gg 1$, which just corresponds to the dilute alloys under consideration. In description of growth of supercritical embryos we also use the phenomenological equation of such growth suggested in the classical theory of nucleation [@Lif-Pit-79] and access validity of this equation by comparison with our KMC simulations. In all our simulations and calculations we consider only homogeneous precipitation and accept the conventional assumption (discussed, in particular, by Mathon et al. [@Mathon-97]) that for the moderate irradiation intensity (i.e. when the ratio of ballistic jump frequency to thermally activated jump frequency is small), the main effect of irradiation is to increase the point defect concentrations, and hence the Fe and Cu diffusion coefficients. It means that the precipitation kinetics under electron irradiation is the same as that during thermal aging at the same temperature, except for an acceleration factor $A_{irr}$ defined by the relation: $$A_{irr}=D^{irr}_{Cu}/D^{th}_{Cu}\label{D_Cu}$$ where $D^{th}_{Cu}$ and $D^{irr}_{Cu}$ are the copper diffusivities during thermal aging and under irradiation, respectively. As evolution of microstructure under precipitation is determined by the diffusion of copper, Eq. (\[D\_Cu\]) implies an analogous relation between the evolution times $t_{irr}$ and $t_{th}$ under irradiation and during thermal aging: $$t_{irr}=t_{th}/A_{irr}. \label{t_irr}$$ Then the precipitation kinetics under electron irradiation can be described by the KMC codes developed by Soisson and Fu [@SF-07] for thermal aging with replacing the thermal aging time $t_{th}$ by the “scaled” time $t_{irr}$ according to Eq. (\[t\_irr\]). This approach is accepted throughout this work. In more detail, scaling relation (\[D\_Cu\]) and estimates of the acceleration factor $A_{irr}$ are discussed in Sec. 4. At the same time, the main results and conclusions of this work do not depend on $A_{irr}$ values as this scaling factor is canceled in many important kinetic characteristics. In Sec. 2 we present some results of calculations of structure of critical embryos and parameters $F_c$ and $J_0$ in (\[J-J\_0\]) for dilute Fe-Cu alloys obtained using the STN [@DV-98a; @DV-98b]. In Sec. 3 we describe our KMC simulations of nucleation and growth of copper precipitates in dilute Fe-Cu alloys. In Sec. 4 we present an estimate of the acceleration factor $A_{irr}$ based on the rate theory models [@Sizmann-78] and find this estimate to reasonably agree with those obtained using the available experimental data [@Mathon-97] and our KMC simulations. In Sec. 5 we use the phenomenological equation for growth of a supercritical embryo [@Lif-Pit-79] and our KMC simulations to estimate some important kinetic parameter which enters into the STN expression for the prefactor $J_0$ in Eq. (\[J-J\_0\]). In Sec. 6 we show that the nucleation barrier $F_c$ values calculated using the STN agree well with those estimated in our KMC simulations, particularly at large $\beta F_c\gg 1$ which correspond to the concentrations and temperatures near the above-mentioned precipitation limits. In Sec. 7 we use the STN calculations combined with some plausible physical assumptions to estimate the temperature dependence of both the homogeneous and the heterogeneous precipitation limit, $x_l^{hom}(T)$ and $x_l^{het}(T)$, for practically interesting temperatures $T$ between 290 and 450$^{\circ}$C. Our main conclusions are summarized in Properties of critical embryos in dilute Fe-Cu alloys calculated using the statistical theory of nucleation =========================================================================================================== In this section we use the STN described in Refs. [@DV-98a; @DV-98b] (to be referred to as I and II) to calculate some thermodynamic, structural and kinetic characteristics of critical embryos. For these calculations we use the $ab$ $initio$ model of Fe-Cu alloys developed by Soisson and Fu and described in detail in Ref. [@SF-07]. Here we only note that this model uses the following values of the binding energy between two copper atoms and between a copper atom and a vacancy, $E^{bn}_{\rm CuCu}$ and $E^{bn}_{v{\rm Cu}}$, for the $n$-th neighbors (in eV): $$\begin{aligned} &&E^{b1}_{\rm CuCu}=0.121-0.182T,\qquad E^{b2}_{\rm CuCu}=0.021-0.091T,\nonumber\\ &&E^{b1}_{v{\rm Cu}}=0.126,\qquad E^{b2}_{v{\rm Cu}}=0.139. \label{E^b-SF}\end{aligned}$$ The high values of $E^{bn}_{\rm CuCu}$ in Eqs. (\[E\^b-SF\]) correspond to strong thermodynamic driving forces for precipitation, while a strong attraction between vacancy and copper atoms leads to the strong trapping of vacancies by copper precipitates discussed in detail in [@SF-07]. First we consider the structural and thermodynamic characteristics of embryos which are described in the STN in terms of the occupation number $n_i$ for each lattice site $i$. The operator $n_i$ is unity when a copper atom is at site $i$ and zero otherwise, while probabilities of various distributions $\{n_i\}$ are described by the distribution function $P\{n_i\}$ given by Eq. (I-3): $$P\{n_i\}=\exp [\beta(\Omega+\mu\sum_in_i-H)]\,.\label{P}$$ Here $H$ is the configurational Hamiltonian supposed to be pairwise, $\mu$ is the chemical potential, and the grand canonical potential $\Omega$ is determined by the normalization condition: $$H=\sum_{i>j}v_{ij}n_in_j,\qquad \Omega =-T\ln {\rm Tr}\exp [\beta(\sum_i\lambda_in_i-H)]\label{H_Omega}$$ where $v_{ij}$ are the configurational interactions, while symbol “Tr” means summation over all configurations $\{n_i\}$. The free energy $F$ of an alloy with embryo is determined by relations $$F=\Omega+\mu\sum_ic_i\, ,\qquad \mu=\partial F/\partial c_i={\rm constant}\label{F-mu}$$ where $c_i=\langle n_i \rangle = {\rm Tr}(n_iP)$ is the mean occupation number, or the local concentration. The free energy $F=F\{c_i\}$ in (\[F-mu\]) was calculated using several statistical approximations described in detail in I and II: mean-field, “mean-field with fluctuations”, pair-cluster, and “pair-cluster-with fluctuations”. The latter was found to be more consistent and accurate than others, thus below we use only the “pair-cluster-with fluctuations” approximation. In tables 1 and 2 and Fig. 1 we present some characteristics of critical embryos. To illustrate both concentration and temperature dependencies of these characteristics, we consider two series of Fe$_{1-x}$Cu$_x$ alloys: those with the same temperature $T=290^{\circ}$C but different concentrations $x$ varying from $0.06$ to $0.4$% (Table 1), and those with the same concentration $x=0.2\%$ but different temperatures $T$ varying from $290$ to 390$^{\circ}$C (Table 2). In these tables, $N_c$ is the total number of copper atoms within the embryo, and $R_c$ is its effective radius defined as that of the sphere having the same volume as $N_c$ copper atoms in the BCC lattice of $\alpha$-iron with the lattice constant $a_0$=0.288 nm: $$R_c=a_0(3N_c/8\pi)^{1/3}=0.142\,N_c^{1/3}\,{\rm nm}. \label{R_c}$$ Note that this critical radius (as well as the precipitate radius $R$ in Eq. (\[R-N\]) below) characterizes the total number of atoms but not the “geometrical” size of precipitate, in particular, not its mean squared radius $\langle r_i^2\rangle^{1/2}$ defined by Eq. (II-10). Therefore, for small and “loose” precipitates, such as those shown by lines B and C in Fig. 1, these radii can notably differ from “geometrical” ones. However, quantities $R_c$ and $R$ in (\[R\_c\]) and (\[R-N\]) are convenient to describe the precipitation kinetics. Quantities $\Delta\Omega_0$ and $\Delta\Omega_1$ in tables 1 and 2 are the zero-order and the first-order terms in fluctuative contributions to the total nucleation barrier $F_c=\Delta\Omega_0+\Delta\Omega_1$ (defined as the difference between grand canonical potentials of two alloy states, that with the embryo and that without the embryo). These two terms correspond to the iterative treatment of some exact relation of thermodynamics of nonuniform systems given by Eq. (I-9) which relates the free energy $F=F\{c_i\}$ to the correlator $K_{ij}$ of fluctuations of site occupations $n_i$: $K_{ij}=\langle (n_i-c_i)(n_j-c_j)\rangle$. As discussed in II and below, the accuracy of such iterative treatment of fluctuations for dilute alloys considered is usually rather high. Fig. 1 illustrates variations of structure of critical embryos with concentration or temperature. The results presented in tables 1, 2 and Fig.1 clearly illustrate very sharp variations of both structure and thermodynamics of critical embryos with concentration $x$ and temperature $T$. In particular, at $T=290^{\circ}$C, the decrease of concentration from $x=0.4$ to $x=0.06\%$ leads to the increase of the embryo size $N_c$ and the reduced nucleation barrier $\beta F_c$ by about three times: from $N_c\simeq 8$ to $N_c\simeq 25$, and from $\beta F_c\simeq 12$ to $\beta F_c\simeq 35$. According to the Zeldovich-Volmer relation (\[J-J\_0\]), the latter implies decreasing the nucleation rate by about ten orders of magnitude (variation with $x$ of the prefactor $J_0$ in (\[J-J\_0\]) will be shown to be negligible). Table 2 illustrates a similar sharp increase of $N_c$ and $\beta F_c$ under elevating temperature between 290 and $390^{\circ}$C. Fig. 1 shows that this sharp rise of sizes and nucleation barriers is accompanied by notable changes in the structure of the embryo: its boundary at high $\beta F_c\gtrsim 30$ is much less diffuse than that at moderate $\beta F_c\lesssim 20$. Discussing the concentration profiles shown in Fig. 1 we note that these profiles (as well as those shown in Fig. 1 in [@KVZ-13] and Fig. 2 in [@KSSV-11]) usually correspond to rather diffuse interfaces. At the same time, KMC simulations show that only very small clusters can have a diffuse interface, while for $R\gtrsim 0.3$ nm (that is, $N\gtrsim 10$ copper atoms) one observes almost “pure” copper clusters with a sharp interface, see, e.g., Fig. 8 in [@SF-07]. However, one should take into account that the profiles shown in Fig. 1 (and other similar figures) correspond to statistical averaging over all orientations of a cluster which is typically rather anisotropic, and this averaging leads to diffuse interfaces; it is illustrated, in particular, by Fig. 8 in [@SF-07]. Let us also note that in phenomenological treatments of critical embryos based on the Cahn-Hilliard continuum approach [@Nagano-06; @Philippe-11], the embryo interfaces are usually notably more diffuse than those obtained in the microscopic STN used. In particular, rather sharp interfaces shown by curves A and D in Fig. 1 can hardly be obtained in the phenomenological treatments. Let us now discuss the prefactor $J_0$ in Eq. (\[J-J\_0\]). In the STN, it is given by Eq. (II-3): $$J_0=\left(\beta |\gamma_0|/2\pi\right)^{1/2}{\cal N} D_{\bf R}({\bf u})D_{aa}.\label{J_0}$$ Here the first three factors (discussed in detail in I) have the “thermodynamic” origin. They describe dependencies of the free energy and the distribution function of the embryo on the “critical” variable $a$ characterizing its size, and on three variables $\bf u$ characterizing tits position $\bf R$. The last factor $D_{aa}$ is the generalized diffusivity in the $c_i$-space that corresponds to an increase of $a$, i. e. to growth of the embryo. This factor is defined by Eq. (I-55) which expresses $D_{aa}$ as a certain linear combination of generalized mobilities $M_{ij}$ which describe the temporal evolution of mean occupations $c_i$ via Eq. (I-50): $$dc_i/dt=\sum_j[M_{ij}-\delta_{ij}\sum_kM_{ik}]\beta\partial F/\partial c_j.\label{dc_i/dt}$$ The $M_{ij}$ values can be calculated using some microscopic models, for example, those employed in II. In particular, for the ideal solution (or in the dilute alloy limit which corresponds to small $c_i$), the mobility $M_{ij}=M_{ij}^0$ is given by Eq. (II-8) with $u_{ik}=0$: $$M_{ij}^0=\delta_{ij,nn}\gamma_{nn}^0[c_i(1-c_i)c_j(1-c_j)]^{1/2}.\label{M_ij^0}$$ Here $\delta_{ij,nn}$ is the Kroneker symbol equal to unity when sites $i$ and $j$ are the nearest neighbors and zero otherwise, while quantity $\gamma_{nn}^0$ has the meaning of the mean rate of exchanges between neighboring copper and iron atoms. It is related to the copper diffusivity $D_0$ in a pure iron by the following relation (see, e. g., Eqs. (72) and (74) in [@VZh-12]): $$\gamma_{nn}^0=D_0/a_0^2\label{gamma_nn^0}$$ where $a_0$ is the BCC iron lattice constant used in Eq. (\[R\_c\]). At the same time, mobilities $M_{ij}$ can be considered as phenomenological parameters in the general Onsager-type equations (\[dc\_i/dt\]), and for weakly nonuniform alloys under consideration they can be supposed to be nonzero only for neighboring sites $i$ and $j$, just as in Eq. (\[M\_ij\^0\]). For our problem, the phenomenological model of $M_{ij}$ can be considered as adequate if it properly describes the diffusion of solute (copper) atoms in processes of nucleation and growth of embryos. In Sec. 5 we show that growth of a supercritical embryo in dilute Fe-Cu alloys can be described well using some generalization of the phenomenological diffusion equation suggested in the classical theory of nucleation [@Lif-Pit-79] given by Eq. (\[dR/dt\]) below. This generalized diffusion equation differs from those for ideal or dilute solutions mainly by the value of the effective diffusivity $D_{eff}$ which takes into account the acceleration of precipitation kinetics due to the high mobility of small copper clusters noted in Refs. [@SF-07; @Jourdan-10] and discussed below. The $D_{eff}$ values for dilute Fe-Cu alloys considered are estimated in Sec. 5 using kinetic Monte Carlo simulations. Therefore, we suggest that the adequate phenomenological description of precipitation in Fe-Cu alloys under consideration can be obtained if we assume for mobilities $M_{ij}$ in (\[dc\_i/dt\]) the same simplest form as that in Eq. (\[M\_ij\^0\]), $$M_{ij}=\delta_{ij,nn}\gamma_{nn}[c_i(1-c_i)c_j(1-c_j)]^{1/2}\,, \label{M_ij}$$ but assume the effective exchange rate $\gamma_{nn}$ in (\[M\_ij\]) to be proportional not to the diffusivity $D_0$ in a pure iron as in Eq. (\[gamma\_nn\^0\]), but to the phenomenological effective diffusivity $D_{eff}$ mentioned above: $$\gamma_{nn}=D_{eff}/a_0^2.\label{gamma_nn}$$ Validity of this assumption will be checked by comparison with the KMC simulation results presented in Secs. 5 and 6. In Table 3 we present values of various factors in the expression (\[J\_0\]) for the prefactor $J_0$ in Eq. (\[J-J\_0\]) for some alloys Fe-Cu. Quantity $\gamma_0$ in this table is the derivative of the free energy $F$ with respect to the effective embryo size $a$: $\gamma_0=\partial^2F/\partial a^2$. As the critical embryo corresponds to the saddle-point of $F$ in the $c_i$-space with respect to this size (I, II), $\gamma_0$ values in table 3 are negative. ${\cal N}$ is the normalizing constant in the embryo size distribution function; $D_{aa}$ is the generalized diffusivity discussed above, and $D_{\bf R}({\bf u})$ is some geometrical factor which for a large embryo is proportional to its surface. For quantities $D_{aa}$, $D_{\bf R}({\bf u})$, and the total prefactor $J_0$ we present their “reduced”, dimensionless values: $D_{aa}/\gamma_{nn}$, $v_aD_{\bf R}({\bf u})$, and $\tilde{J_0}=v_aJ_0/\gamma_{nn}$. Note that the embryo size $N_c$ decreases with $x$ and increases with $T$, as tables 1 and 2 show. Table 3 shows that both concentration and temperature dependencies of quantities ${\cal N}$ and $D_{aa}/\gamma_{nn}$ are rather weak. On the contrary, the geometrical factor $D_{\bf R}({\bf u})$ notably increases with the embryo size $N_c$, while $\beta|\gamma_0|$ somewhat decreases with $N_c$. The total reduced prefactor $\tilde{J_0}= v_aJ_0/\gamma_{nn}$ notably increases with $N_c$, by about 20 or 5 times for the total $x$ or $T$ intervals shown in table 3. However, in the total nucleation rate $J$ given by Eq. (\[J-J\_0\]), these variations of $J_0$ are negligible as compared to the above-mentioned huge changes of the activation factor $\exp\,(-\beta F_c)$ for these intervals of $x$ or $T$. Now we note that a very sharp decrease of this activation factor with $x$ between $x\sim 0.2$ and $x\sim 0.08\%$ seen in table 1 evidently correlates with the above-mentioned experimental estimates of the homogeneous precipitation limit at $T=290^{\circ}$C given in Refs. [@Mathon-97; @Auger-00; @Radiguet-07]. It seems natural to suggest that the position $x_l^{hom}$ of this limit is mainly determined by the activation factor $\exp (-\beta F_c)$, or by the reduced nucleation barrier $\beta F_c$ which at $x=x_l^{hom}$ reaches a certain high value $(\beta F_c)_l$. For example, if we accept for the experimental $x_l^{hom}$=$x_l$ at $290^{\circ}{\rm C}$ the estimate from Ref. [@Mathon-97], $x_l^{exp}\sim 0.08$-$0.09\%$, we have: $(\beta F_c)_l\simeq 28$, while taking this estimate from Ref. [@Auger-00], $x_l^{exp}\sim 0.15\%$, we obtain: $(\beta F_c)_l\simeq 23$. In more detail, estimates of the precipitation limits $x_l$ are discussed below in Sec. 6. Kinetic Monte Carlo simulations of nucleation and growth of precipitates in dilute Fe-Cu alloys =============================================================================================== As mentioned in Sec. 1, our KMC simulations use the “scaling” assumptions (\[D\_Cu\]) and (\[t\_irr\]) which enable us to relate the homogeneous precipitation in irradiated alloys to that during thermal aging for which the KMC codes developed by Soisson and Fu [@SF-07] can be employed. In these simulations we consider the following Fe$_{1-x}$Cu$_x$ alloys: (A) those at $T=290^{\circ}$ C and $x$ equal to 0.088, 0.116, 0.144, 0.177, 0.2, 0.3 and 0.4%, and (B) those at $x=0.2\%$ and temperatures $T$ equal to 290, 305, 320, 335, 350, 365, 380 and 390$^{\circ}$C. We use simulation volume $V_s=L^3$ with periodic boundary conditions and, to check the statistical reliability of results, we usually employ several different values of the simulation size $L$: $$L_1=128a_0,\quad L_2=160a_0, \quad L_3=192a_0.\label{L_n}$$ Before to describe the results, we mention a characteristic feature of such simulations for some low concentrations $x$ or high temperatures $T$: at such $x$ and $T$, we observe only one precipitate within simulation volume, that is, the total number $N_p$ of precipitates within simulation box is unity (while at further lowering $x$ or elevating $T$ we observe no precipitates for the simulation time). In our simulations it was the case for the series (A) alloys with $T=290^{\circ}$C at $x\leq 0.144\%$, and for the series (B) alloys with $x=0.2\%$, at $T\geq 335^{\circ}$C. In all such cases, the relation $N_p=1$ preserves under increasing simulation volume (replacing $L_1$ by $L_2$ or $L_3$), though the incubation time $t_{inc}$ (that which precedes to the formation of the embryo [@SM-00]) somewhat decreases when $V_s$ increases. The presence of changes of physical characteristics (such as the precipitate density $d_p=N_p/V_s$ or the incubation time $t_{inc}$) under variations of simulation volume evidently indicates on the statistical unreliability of simulation results for these characteristics. At the same time, the growth of a precipitate after its formation seems to be described by such simulations quite properly. Therefore, even though the simulations with $N_p=1$ can not be used to study the evolution of precipitate density $d_p(t)$, they will be widely used for studies of precipitate growth described in Sec. 5. When the number of precipitates within simulation box significantly exceeds unity, the simulations become statistically reliable and can be used to study the precipitate density $d_p(t)$. In figures 2 and 3 we show the dependencies $d_p(t)$ for all our simulations for which the final number of precipitates within simulation box is not too small: $N_p\geq 3$. The simulation time on abscissa axis is given in the “thermal aging” values $t_{th}$ which in our model are related to the observed time $t_{irr}$ by Eq. (\[t\_irr\]). The acceleration factor $A_{irr}$ in this relation and comparison with the available data [@Mathon-97] about precipitation in an Fe-0.3%Cu alloy under electron irradiation are discussed below in Sec. 4. In the lower part of frames 2(e) and 2(f) we also show temporal dependencies of the mean precipitate radius $R_m$ for two simulations with $x=0.3\%$. For other our simulations with $N_p\gtrsim 10$ these dependencies are similar and describe usually a smooth increase of $R_m$ by about 20-30%. The dependencies $d_p(t_{th})$ presented in Figs. 2 and 3 are approximately linear. Therefore, if we define the nucleation rate $J=J_{\rm KMC}$ for these simulations as the ratio $$J_{\rm KMC}=d_p(\tau)/\tau\label{J_KMC}$$ where $\tau=(t_{th}-t_{th}^{inc})$ is the evolution time counted off the incubation time $t_{th}^{inc}$, this nucleation rate for each simulation remains approximately constant, particularly for the simulations with not small $N_p$ where fluctuations are not too strong. Hence these our simulations seem to describe mainly the first stage of precipitation, the steady-state nucleation, for which the constant value of the nucleation rate $J$ is characteristic. The dashed line in each frame of Figs. 2 and 3 shows our estimate of the $J_{\rm KMC}$ value in Eq. (\[J\_KMC\]) assuming it to be constant. For this estimate we tried to use the maximum broad interval of time $t_{th}$ for which the temporal dependence $d_p(t_{th})$ is close to linear. In spite of the evident scatter of KMC results presented in Figs 2 and 3, particularly for simulations with not large $N_p$, the linear temporal dependence $d_p(\tau_{th})$ can usually be followed rather clearly . Therefore, our estimates of $J_{\rm KMC}$ values in Eq. (\[J\_KMC\]) seem to be sufficiently definite. At the same time, the Zeldovich-Volmer relation (\[J-J\_0\]) is valid just for the steady-state nucleation stage. Therefore, we can use our estimates of $J_{\rm KMC}$ to estimate the parameters $F_c$ and $J_0$ in (\[J-J\_0\]) putting: $$J_0\exp (-\beta F_c)=J_{\rm KMC}.\label{J_KMC-F_c}$$ As discussed in Sec. 2, variations of the activation factor $\exp (-\beta F_c)$ with concentration or temperature affect the total nucleation rate $J$ much stronger than those of the prefactor $J_0$. Therefore, for the given $J_{\rm KMC}$, the activation factor $\exp (-\beta F_c)$ can be determined from Eq. (\[J\_KMC-F\_c\]) rather accurately even if the prefactor $J_0$ is estimated not too precisely, in particular, when $J_0$ is estimated from the STN calculations described in Sec. 2 which include model assumptions (\[M\_ij\]) and (\[gamma\_nn\]). Substituting for the prefactor $J_0$ its expression via quantities $\tilde{J}_0$, $v_a$ and $\gamma_{nn}$ given in the last line of table 3, and using Eq. (\[gamma\_nn\]) which relates $\gamma_{nn}$ to the effective diffusivity $D_{eff}$ mentioned in Sec. 2, we can re-write relation (\[J\_KMC-F\_c\]) as the equation for the reduced nucleation barrier $\beta F_c$: $$(\beta F_c)_{\rm KMC}=\ln \Big(2\,D_{eff}\tilde{J}_0/a_0^5 J_{\rm KMC}\Big)\label{F_c-KMC}$$ where index “KMC” in the left-hand side indicates that this expression for $\beta F_c$ is based on the KMC simulations. Values of the reduced prefactor $\tilde{J}_0$ can be taken from our STN calculations illustrated by the last line of table 3. Hence to find the reduced nucleation barrier $\beta F_c$ from Eq. (\[F\_c-KMC\]), we need the effective diffusivity $D_{eff}$. Estimates of $D_{eff}$ based on our KMC simulations of growth of precipitates are described below in Sec. 5. Estimate of acceleration precipitation factor $A_{irr}$ for electron irradiation of dilute iron-copper alloys ============================================================================================================= The concentrations of point defects (vacancies and interstitials) under permanent irradiation may frequently exceed their equilibrium values by several orders of magnitudes [@Sizmann-78]. In such conditions diffusive phase transformations, such as precipitation or ordering, are strongly accelerated. For the case of electron irradiation of Fe-Cu alloys this has been observed, in particular, by Mathon et al. [@Mathon-97] and by Radiguet et al. [@Radiguet-07]. If other irradiation effects (such as ballistic mixing or radiation induced segregation) can be neglected, it is natural to expect that the precipitation kinetics under irradiation is the same as that during thermal aging at the same temperature, except for the strong acceleration of this kinetics described by the factor $A_{irr}$ in Eqs. (\[D\_Cu\]) and (\[t\_irr\]). In this section we estimate this acceleration factor for dilute Fe-Cu alloys under typical conditions of electron irradiation [@Mathon-97] and compare this estimate with both the experimental and KMC simulation results. For dilute alloys under consideration, the copper diffusivity under thermal aging in Eq. (\[D\_Cu\]) is given by the conventional expression [@LeClaire-78]: $$D_{Cu}^{th} = \alpha _v c_v^{eq} D_v\label{D_Cu^th}.$$ Here $c_v^{eq}$ is the equilibrium vacancy concentration which is expressed via the enthalpy and entropy of vacancy formation, $H_v^{for}$ and $S_v^{for}$, as: $c_v^{eq}=\exp(S_v^{for}-\beta H_v^{for})$, and $D_v$ is the vacancy diffusion coefficient in a pure iron which is related to the vacancy migration enthalpy and entropy, $ H_v^{mig}$ and $S_v^{mig}$, as follows: $$D_v=a_0^2 \nu _0 \exp(S_v^{mig}-\beta H_v^{mig})\label{D_v}$$ where $a_0$ is the BCC lattice constant and $\nu_0$ is the attempt frequency supposed to have the order of the Debye frequency. The coefficient $\alpha_v$ in (\[D\_Cu\^th\]) describes vacancy-solute correlations, and it is commonly written as [@LeClaire-78]: $$\alpha_v = f_2 \exp \left({\beta G_{Cu-v}^{bin} }\right)\label{alpha_v}$$ where $f_2$ is the impurity correlation factor, and $G_{Cu-v}^{bin}$ is the copper-vacancy binding energy. These parameters may be obtained from experimental measurements, or from [*ab initio*]{} calculations [@SF-07]. Similarly, under irradiation one may write: $$D_{Cu}^{irr} = \alpha _v c_v^{irr} D_v + \alpha _i c_i^{irr} D_i\label{D_irr} ,$$ where $c_v^{irr}$ and $c_i^{irr}$ are the vacancy and the interstitial concentrations under irradiation. It seems natural to assume that the copper-vacancy correlation effects and binding are not modified by irradiation. Detailed information on the interstitial diffusion (needed for calculations of $\alpha _i$ and $D_i$) is usually more difficult to obtain. However, when the point defect concentrations reach their steady-state value, one can show that $c_v^{irr} D_v =c_i^{irr} D_i $ [@Sizmann-78]. If $\alpha_v$ and $\alpha _i$ are of the same order of magnitude, one finally gets $D_{Cu}^{irr} \simeq 2\alpha _v c_v^{irr} D_v$, which corresponds to the following acceleration factor in (\[D\_Cu\]) and (\[t\_irr\]): $$A_{irr} = 2\,(c_v^{irr}/ c_v^{eq})\,.\label{A_irr-FS}$$ To compare our KMC simulations during thermal aging and the experimental kinetics of precipitation under irradiation, the key point is therefore to get a reliable estimation of the vacancy point defect concentrations under irradiation. According to the rate theory models [@Sizmann-78], if point defects created by irradiation disappear by mutual recombination or by annihilation at dislocations (which are assumed to be the dominant point defect sinks), the evolution of their concentrations in a pure metal is described by the following equations: $$\begin{aligned} &&{dc_v\over dt} = K - Rc_i c_v - \rho_d D_v (c_v - c_v^{eq} )\nonumber\\ &&{dc_i\over dt} = K - Rc_i c_v - \rho_d D_i (c_i - c_i^{eq} ) \simeq K - Rc_i c_v - \rho_d D_i c_i\,.\label{eq:dcdt}\end{aligned}$$ Here $K$ is the rate of formation of Frenkel pairs under electron irradiation (in dpa/s), $\rho_d$ is the dislocation density, and $R$ is the vacancy-interstitial recombination rate: $$R = 4\pi r_{vi}(D_v + D_i )/v_a\label{R-recombin}$$ where $r_{vi}$ is the recombination radius, and $v_a=a_0^3/2$ is the atomic volume. For pure iron (for which lattice constant is $a_0$=0.288 nm), most of the point defect properties have been estimated by [*ab initio*]{} calculations. In particular, for the vacancy formation parameters we have: $H_v^{for} = 2.18\;{\rm{eV}}$ [@SF-07] and $S_v^{for} = 4.08$ [@Nichols-78], which gives: $c_v^{eq} \simeq 1.8 \times 10^{ - 18} $ at 290$^{\circ}$C. The migration enthalpies are: $H_v^{mig} = 0.68$ eV [@SF-07] and $H_i^{mig} = 0.34$ eV [@Lucas-09]. The Debye frequency is $\nu _0 = 10^{13} {\rm{s}}^{{\rm{-1}}}$, and with these parameters a vacancy migration entropy $S_v^{mig} \simeq 2.2$ is required to get the experimental pre-exponential factor for self-diffusion of iron (see Refs. [@SF-07; @Fu-04] for details). Lacking both experimental and [*ab initio*]{} estimations, we will assume the same value for the migration entropy of the interstitials. The irradiation conditions of ref. [@Mathon-97] are: $K = 2 \times 10^{-9}$ dpa/s at 290$^{\circ}$C, with dislocation densities between $\rho _d=10^8\; \rm cm^{-2}$ and $\rho _d=10^{11}\; \rm cm^{-2}$. The evolution of point defect concentrations, obtained by numerical integration of Eq. (\[eq:dcdt\]) with these two dislocations densities, is given on Fig. 4. In both cases, the steady-state values are reached very rapidly (after less than $10^{-1}$ s), with vacancy concentrations $c_v^{irr} \simeq 3.5 \times 10^{-11}$ and $3.6 \times 10^{ - 14} $. The acceleration factor is therefore $A_{irr} \simeq 4 \times 10^4 $ for the highest dislocation density and $A_{irr} \simeq 3.9 \times 10^7 $ for the lowest. In Ref. [@Mathon-97], a precipitate density of $0.9\times 10^{23} \rm m^{-3}$ is observed under electron irradiation of Fe-0.3%Cu alloy after approximately $8.3\times10^4$ s. Our KMC simulations of precipitation during thermal aging shown in Fig. 2(f) predict a similar density after $1.5\times 10^9$ s. This corresponds to an acceleration by a factor $A_{irr}\sim 1.8\times 10^4$ under irradiation. Considering the strong approximations used in the rate theory model, and the uncertainties on the point defect properties, the agreement between our estimate of $A_{irr}$ and the combination of experimental and KMC results seems to be reasonable. A more detailed comparison of our simulations with experiments by Mathon et al. [@Mathon-97] is hindered by a low resolution in measurements of precipitate sizes $R$ by SANS method used in [@Mathon-97]: $\Delta R\sim 0.5$ nm. It can explain quantitative disagreements between the results reported in [@Mathon-97] and those shown in frames 2(e) and 2(f). The maximum value of the precipitate density $d_p(t)$ reported in [@Mathon-97], $d_{max}\sim 1.65\times 10^{23}1/{\rm m}^3$, is by about 1.5 times lower than the final values $d_p(t)$ shown in Figs. 2(e) and 2(f), even though the maximum value $d_{max}$ is probably not reached yet in these simulations. Hence, our simulated $d_p$ exceed those reported in [@Mathon-97] by at least several times. At the same time, values of the mean precipitate radius for the nucleation stage (which corresponds to $d_p(t)<d_{max}$) reported in [@Mathon-97], $R_{m,M}^n\sim 1.1$—$1.65$ nm, are by about 2-3 times higher than the analogous values $R_m^n\sim 0.3-0.5$ nm observed in both experiments [@Goodman-73; @Kampmann-86] and simulations [@SF-07; @KVZ-13] for other Fe-Cu alloys, while our $R_m$ in frames 2(e) and 2(f) are close to these usual $R_m^n$. Therefore, it seems probable that because of the above-mentioned low resolution $\Delta R\sim 0.5$ nm, values of the precipitate density $d_p$ reported in [@Mathon-97] are underestimated, while those of the mean precipitate size $R_m$ are overestimated. More accurate measurements of $d_p$ and $R_m$ in dilute Fe-Cu alloys under electron irradiation are evidently needed for a quantitative comparison with our simulations. Estimates of effective diffusivity of copper atoms for growth of a precipitate using kinetic Monte Carlo simulations ==================================================================================================================== In the phenomenological theory of growth of a supercritical embryo developed for a spherical embryo of a radius $R$ with a sharp edge, the temporal dependence $R(t)$ (for the rigid lattice alloy model used) is described by the following equation [@Lif-Pit-79]: $$dR/dt=D_{eff}(R-R_c)(c-c_b)/R^2. \label{dR/dt}$$ Here $D_{eff}$ is the effective diffusivity of a solute atom (in our case, of a copper atom) in an alloy; $R_c$ is the critical radius; $c=\,x$ is the copper concentration far from the embryo; and $c_b$ is the binodal concentration (solubility limit) for temperature under consideration. In more realistic descriptions, in particular, in the STN or KMC simulations described in sections 2 and 3, both critical and supercritical embryos have not a sharp edge but a diffuse surface illustrated by Fig. 1. More important, in the phenomenological derivation [@Lif-Pit-79] of Eq. (\[dR/dt\]), the diffusivity $D_{eff}$ was supposed to be constant independent of the precipitate size $N$ and its mobility $m(N)$. At the same time, the recent KMC studies [@SF-07; @Jourdan-10] have shown that the small precipitates under consideration which contain $N\lesssim$100 copper atoms are highly mobile, being much more mobile than individual copper atoms. It leads to a great acceleration of precipitation kinetics which qualitatively corresponds to an increase of the effective diffusivity $D_{eff}$ in Eq. (\[dR/dt\]) with respect to the diffusivity of an individual copper atom. In spite of all these simplifications, the phenomenological equation (\[dR/dt\]) is commonly believed to realistically describe growth of supercritical embryos [@Lif-Pit-79]. Accepting this point of view, we first present the explicit solution of Eq. (\[dR/dt\]) for $R(t)$. Then we use our KMC simulations to access validity of Eq. (\[dR/dt\]) and to estimate the parameter $D_{eff}$ in this equation for dilute Fe-Cu alloys under consideration. First we note that employing the simplest sharp-edge model of an embryo in usual derivations of Eq. (\[dR/dt\]) seems to be unessential, and this equation (at large $N_c\gg 1$ considered) appears to be valid for any realistic description of growth of an embryo, including that used in KMC simulations. The radius $R(t)$ in Eq. (\[dR/dt\]) mainly characterizes the total number $N$ of solute atoms within the embryo, and for any precipitate this radius can be expressed via $N$ analogously to Eq. (\[R\_c\]) for the critical embryo: $$R=a_0(3N/8\pi)^{1/3}=0.142\,N^{1/3}\,{\rm nm}. \label{R-N}$$ The right-hand-side of Eq. (\[dR/dt\]) contains two basic factors (in addition to the diffusivity $D_{eff}$ and a geometrical factor $1/R^2$) which naturally describe the driving force for growth of the embryo: the factor $(c-c_b)$ that characterizes supersaturation of a metastable alloy, and the factor $(R-R_c)$ that describes vanishing of this force at $R=R_c$. Therefore, the phenomenological equation (\[dR/dt\]) with the generalization (\[R-N\]) seems to be rather plausible, and comparison to our KMC simulations given below seems to confirm its validity, at least for the $R$ and $t$ intervals studied. First we present the explicit solution $R(t)$ of Eq. (\[dR/dt\]). As discussed in detail in [@Lif-Pit-79], the embryo can be considered as “supercritical” only when its size $R$ exceeds some value $(R_c+\Delta)$ where $\Delta$ determines the scale of critical fluctuations of sizes near $R_c$. Therefore, the time $t_{f}$ when the supercritical embryo has been eventually formed is defined by the relation: $$R(t_{f})=(R_c+\Delta).\label{t_fs}$$ For the sharp-edge model of the embryo used in [@Lif-Pit-79], the fluctuation width $\Delta$ is expressed via the interfacial energy $\sigma$ as: $\Delta=(T/8\pi\sigma)^{1/2}$. To generalize this estimate to the case of real embryos with diffuse interfaces, we can express $\sigma$ via the critical radius $R_c$ and the nucleation barrier $F_c$. It yields: $$\Delta= R_c\alpha,\qquad \alpha=(2T/3F_c)^{1/2}.\label{alpha}$$ For large embryos under consideration with $\beta F_c\gg 1$, values of $\alpha$ are small; in particular, for the Fe-Cu alloys listed in tables 1 and 2, we have: $\alpha\sim 0.15-0.2$. Integrating Eq. (\[dR/dt\]) from the initial time $t_{f}$ defined by Eq. (\[t\_fs\]) to the arbitrary time $t$ when $R(t)=R$, we obtain: $$(R-R_c-\Delta)(R+3R_c+\Delta)/2+R_c^2\ln\,[(R-R_c)/\Delta]= D_{eff}(c-c_b)\tau\label{R(t)}$$ where $\tau=(t-t_{f})$ is the total time of growth of a supercritical embryo. If we describe this growth by a reduced variable $y=(R-R_c-\Delta)/R_c$, Eq. (\[R(t)\]) takes a universal form which contains only the reduced dimensionless time $\xi$: $$\begin{aligned} &&y(2+\alpha+y/2)+\ln(1+y/\alpha)=\xi\nonumber,\\ &&\xi=D_{eff}(c-c_b)\tau/R_c^2.\label{y(t)}\end{aligned}$$ At large $R\gg R_c$, the dependence $R(\tau)$ given by Eq. (\[R(t)\]) or (\[y(t)\]) takes the form $$R(\tau)=[2D_{eff}(c-c_b)\tau]^{1/2}\label{R_diff-tau}$$ which describes the diffusion-controlled growth of a spherical embryo, see, e. g., [@Martin-76a]. Now we compare the description of growth of an embryo by the phenomenological equation (\[R(t)\]) or (\[y(t)\]) with that given by our KMC simulations. First, we note that for our model of irradiated alloys based on Eqs. (\[D\_Cu\]) and (\[t\_irr\]), the reduced time $\xi$ in Eq. (\[y(t)\]) does not include the irradiation acceleration factor $A_{irr}$, just as the right-hand side of Eq. (\[F\_c-KMC\]), and this reduced time is explicitly expressed via the thermal aging time $\tau=\tau_{th}= (t_{th}-t_{th}^{f})$ used in our KMC simulations. Therefore, growth of a supercritical embryo can be described, from one side, by the function $R(\xi)$ determined by Eq. (\[R(t)\]) or (\[y(t)\]). From the other side, such growth can be followed in our KMC simulations for any precipitate chosen, which yields the dependence $R=R_{\rm KMC}(t_{th})$ with $R$ defined by Eq. (\[R-N\]). Equating these two quantities, $$R(\xi)=R_{\rm KMC}(t_{th}),\label{R_xi_KMC}$$ with $\xi$ given by Eq. (\[y(t)\]) for $\tau=\tau_{th}$, we obtain the equality which contains only one unknown parameter $D_{eff}$ at all times $\tau_{th}>0$. One can expect that the proper choice of this single parameter can provide a good accuracy for obeying equality (\[R\_xi\_KMC\]) at all $\tau_{th}$ considered only if the phenomenological equation (\[dR/dt\]) holds true for these $\tau_{th}$. Therefore, the check of validity of relation (\[R\_xi\_KMC\]) enables us, first, to access reliability of the phenomenological equations (\[dR/dt\]) and (\[R(t)\]) for description of growth of supercritical embryos and, second, to estimate the effective diffusivity $D_{eff}$ in these equations for dilute Fe-Cu alloys. It is convenient to characterize the effective diffusivity $D_{eff}(x,T)$ by its ratio $A_c$ to the appropriate dilute alloy diffusivity $D(0,T)=D_0(T)$ writing $D_{eff}$ as $$D_{eff}(x,T)=A_c\times D_0(T).\label{A_c}$$ The parameter $A_c=A_c(x,T)$ in (\[A\_c\]) characterizes the effective acceleration of growth of precipitates being mainly due to the above-mentioned high mobility of small copper clusters. To differ it from the irradiation acceleration factor $A_{irr}$ discussed in Sec. 4, $A_c$ will be called “the acceleration diffusion parameter”. For the dilute alloy diffusivity $D_0(T)$ of a copper atom in a pure iron at $T=290-450^{\circ}$C we will use the conventional Arrhenius-type expression $$D_0(T)=A\exp\,(-\beta Q)\label{D_0}$$ with the $A$ and $Q$ values suggested by Soisson and Fu [@SF-07] basing on combinations of their $ab \ initio$ calculations and empirical estimates: $$A=0.97\cdot 10^{-4} \ {\rm m^2/s}\,,\qquad Q=2.67 \ {\rm eV}. \label{SF-parameters}$$ In Figs. 5 and 6 we show the phenomenological functions $R(\xi)$ defined by Eqs. (\[R(t)\]) and (\[y(t)\]) together with the dependencies $R_{\rm KMC}(t_{th})$ observed in our KMC simulations. Each curve $R_{\rm KMC}(t_{th})$ in these figures starts from the time of formation of the precipitate chosen observed in the KMC simulation, while the curve $R(\xi)=R[\xi(t_{th})]$ starts from the time $t_{th}^{f}$ of formation of the supercritical embryo defined by Eq. (\[t\_fs\]). Each frame in these figures corresponds to some of our simulations; for other simulations (not shown in Figs. 5 and 6), the results are similar. The parameter $A_c$ and the time $t_{th}^{f}$ used to draw the curve $R(\xi)$ in each frame have been estimated from the best fit of this $R(\xi)$ to the $R_{\rm KMC}(t_{th})$ curve at this frame. In Fig. 7 we show the concentration and temperature dependencies of the acceleration diffusion parameter $A_c$ obtained in these KMC estimates. Let us discuss the results presented in Figs. 5-7. First, we note that some of irregularities in dependencies $R_{\rm KMC}(t_{th})$ seen in Figs. 5 and 6 can be related not to the real statistical fluctuations of sizes but to the methodical errors due to too long temporal intervals between sequent savings of KMC data. This seems to be the case, in particular, for initial stages of simulations shown in frames 5(a), 6(c) and 6(d), and also for some other frames, e. g., 6(a). At the same time, for the later stages of growth when the precipitate becomes supercritical, $R\gtrsim (R_c+\Delta)$, these methodical distortions seem to be less significant. Second, Figs. 5 and 6 clearly illustrate the strong fluctuations of sizes in the “critical” region $R\lesssim (R_c+\Delta)$ mentioned in the discussion of Eqs. (\[t\_fs\]) and (\[alpha\]). Such fluctuations are pronounced, in particular, in frames 5(b), 5(c), 6(c), 6(e), and 6(f). Third and most important, Figs. 5 and 6 show that the description of growth of precipitates by the phenomenological equations (\[dR/dt\]) and (\[R(t)\]) seems to agree well with the KMC simulations, at least up to $R\gtrsim 1.5 R_c$, that is, under increase of the number of copper atoms in the growing embryo by about four times. Figs. 5 and 6 also clearly illustrate the effects of cluster mobility and their direct coagulation mentioned above. In particular, sudden increases (“jumps”) of $R(t)$ seen in Fig. 5(d) at $t_{th}=10.6\times 10^{9}$ s, as well as in Fig. 5(f) at $t_{th}=1.5\times 10^{9}$ and $t_{th}=1.75\times 10^{9}$ s, occur simultaneously with the disappearance (evidently, due to coagulation) of one of supercritical clusters in Figs. 2(b) and 2(f). Similarly, jumps of $R(t)$ in Fig. 6(a) at $t_{th}=6.56\times 10^{9}$ s and in Fig. 6(b) at $t_{th}=1.46\times 10^{9}$ s occur simultaneously with the disappearance of one supercritical cluster in Figs. 3(b) and 3(d), respectively. The smaller jumps in $R(t)$ seen in Figs. 5(d), 5(e), 6(a) and 6(b) can correspond to coagulation of subcritical clusters which are not registered in Figs. 2 and 3. At the same time, for simulations with $N_p=1$ shown in Figs. 5(a)-5(c) and 6(d)-6(f), for which only one supercritical cluster is present in the simulation box, the large and distinct jumps in $R(t)$ are not clearly seen. Inspection of atomic distributions for these simulations shows that growth of this supercritical cluster is mainly realized via its fast diffusion among almost immobile individual copper atoms which are sometimes “swept” and absorbed by this mobile cluster. Some subcritical clusters containing several copper atoms are also observed in these simulations, and they seem to diffuse notably slower than the supercritical cluster, in a qualitative agreement with Fig. 9 in [@SF-07]. The resulting acceleration diffusion parameter $A_c$ for dilute Fe-Cu alloys is shown in Fig. 7. It is rather high: $A_c\gtrsim 200$, having the same scale as the parameters of acceleration of precipitation in Fe-1.34%Cu alloys at $T=500^{\circ}$C due to the high mobility of copper clusters studied in Refs. [@SF-07] and [@Jourdan-10]. Fig. 7 also shows that the concentration and temperature dependencies of this acceleration are rather sharp. Fig. 7(a) illustrates the notable increase of $A_c$ with increase of copper concentration $x$; it can be explained by a probable enhancement of density of mobile copper clusters (both supercritical and subcritical) under increase of the copper content in an alloy. Fig. 7(b) shows a significant decrease of $A_c$ under elevating temperature $T$. This also seems natural, as elevating $T$ should lead to the weakening of the copper-vacancy binding proportional to the Mayer functions $[\exp\,(\beta E^{bn}_{v{\rm Cu}})-1]$ with $E^{bn}_{v{\rm Cu}}$ from (\[E\^b-SF\]) [@KSSV-11]. Hence the strong trapping of vacancies by copper clusters (which is the physical origin of their high mobility [@SF-07]) should weaken. Comparison of nucleation barriers found using the STN calculations and the KMC simulations ========================================================================================== Estimates of the nucleation rate $J_{\rm KMC}$ described in Sec. 3 and illustrated by Figs. 2 and 3, combined with the estimates of the effective diffusivity $D_{eff}$ discussed in Sec. 5 and illustrated by Figs. 5-7, enable us to find the reduced nucleation barriers $\beta F_c$ in Eq. (\[F\_c-KMC\]) basing on the kinetic Monte Carlo simulations. This equation (\[F\_c-KMC\]) also shows that for the large $\beta F_c$ considered which vary with $x$ and $T$ very sharply, possible errors of a relative order of unity in our estimates of parameters $\tilde{J}_0$, $J_{\rm KMC}$ and $D_{eff}$ make no significant effect on the $\beta F_c$ values obtained. In table 4 we present the $(\beta F_c)_{\rm KMC}$ values estimated for all simulations shown in Figs. 2 and 3, together with the analogous $(\beta F_c)_{\rm STN}$ values calculated in Sec. 2 with the use of the statistical theory of nucleation [@DV-98a; @DV-98b]. In comparison of these KMC and STN results we should remember that the STN assumes the size and the nucleation barrier of the embryo to be large: $N_c\gg 1,\ \beta F_c\gg 1$. Therefore, the accuracy of this theory and its agreement with the KMC simulations should improve when the size of the embryo increases. The results presented in tables 1, 2 and 4 agree with these considerations. We see that at low concentrations $x\leq 0.2\%$ when both $N_c$ and $\beta F_c$ are sufficiently large: $N_c\gtrsim 12$, $\beta F_c\gtrsim 20$, the agreement between KMC and STN results is virtually perfect: for all nine simulations with $x\leq 0.2\%$ shown in table 4, differences between the $(\beta F_c)_{\rm KMC}$ and the $(\beta F_c)_{\rm STN}$ values have the order of a percent. At the same time, at higher $x$ equal to 0.3 or 0.4%, the $N_c$ and $\beta F_c$ values notably decrease, and differences between $(\beta F_c)_{\rm KMC}$ and $(\beta F_c)_{\rm STN}$ increase, which may reflect the lowering of accuracy of the STN. Therefore, the results presented in table 4 enable us to make an important conclusion that at high values $\beta F_c\gtrsim 20$, the STN-based calculations of nucleation barriers are highly reliable. At the same time, as mentioned in Sec. 2, the homogeneous precipitation limit in dilute Fe-Cu alloys corresponds just to the values $\beta F_c> 20$. Therefore, the nucleation barriers near the homogeneous precipitation limits in these alloys can be reliably calculated using the STN. Temperature dependencies of precipitation limits in dilute iron-copper alloys ============================================================================= Calculations of nucleation barriers described in Secs. 2 and 6 can be used for tentative estimates of the homogeneous precipitation limit $x_l^{hom}(T)$ at different temperatures $T$. As mentioned in Sec. 2, it seems natural to suggest that the position of this limit is mainly determined by the value of the activation factor $\exp (-\beta F_c)$, that is, by the reduced nucleation barrier $\beta F_c$ which at $x=x_l^{hom}$ takes a certain high value $(\beta F_c)_l$. It seems also natural to assume that these precipitation limits at different temperatures $T$ correspond to the similar values of the activation factor $\exp (-\beta F_c)$. It implies the following relation: $$(\beta F_c)_l\equiv\beta F_c\Big[x_l^{hom}(T),T\Big]\simeq C_{hom}\label{F_cl-hom}$$ where $C_{hom}$ is a constant independent of temperature, and thus it can be estimated using the experimental value $x_l^{hom}(290^{\circ}{\rm C})\equiv x_l^{exp}$. For example, using estimates of $x_l^{exp}$ mentioned in Sec. 2, $x_l^{exp}\sim 0.08-0.09\%$ from [@Mathon-97] or $x_l^{exp}\sim 0.15\%$ from [@Auger-00], we obtain: $C_{hom}\simeq 28$ or $C_{hom}\simeq 23 $. Then equation (\[F\_cl-hom\]) with the function $F_c(x,T)$ calculated using the STN enable us to find the $x_l^{hom}(T)$ value. A high accuracy of the STN calculations of nucleation barriers combined with the above-mentioned physical considerations allow us to expect that such “semi-empirical” estimates of the homogeneous precipitation limit for temperatures $T$ of practical interest can be sufficiently reliable. One can also try to extend this approach to the case of the heterogeneous precipitation of copper on clusters of irradiation-induced point defects, vacancies and self-interstitial atoms (that is, to the case of neutron or ion irradiation [@Akamatsu-95]-[@Meslin-10b]), to estimate the appropriate heterogeneous precipitation limit $x_l^{het}$, which is more interesting for applications. Even though kinetic paths of the heterogeneous and the homogeneous precipitation differ from each other (see, e. g., [@Martin-76b]), the main physical origin for suppressing precipitation at low solute concentrations $x$ seems to be the same for both processes. In both cases, formation of a precipitate needs overcoming the energetic barrier due to the surface loss in the free energy $F$ which is not compensated by the volume gain in $F$ until the precipitate volume becomes sufficiently large. Differences between these two processes are usually related mainly to some geometrical factors which depend on the structure of the heterogeneity [@Martin-76b]. Therefore, one may expect that the heterogeneous precipitation limit $x_l^{het}$ can also be estimated from a phenomenological relation similar to (\[F\_cl-hom\]), but the “limiting” value $F_{cl}$ in this relation should be higher as the nucleation barrier for the heterogeneous nucleation, $F_c^{het}$, is lower than that for the homogeneous nucleation, $F_c^{hom}$. However, if we assume that the difference between $F_c^{het}$ and $F_c^{hom}$ has mainly a geometrical origin (which is the case for the simplest models of heterogeneous nucleation [@Martin-76b]), then the ratio $F_c^{het}/F_c^{hom}$ can be supposed to weakly vary with temperature. Then the temperature dependence of the heterogeneous precipitation limit $x_l^{het}(T)$ can be estimated from a semi-empirical relation similar to (\[F\_cl-hom\]): $$\beta F_c\Big[x_l^{het}(T),T\Big]\simeq C_{het}\label{F_cl-het}$$ where $F_c(x,T)$ is again the STN calculated nucleation barrier for the homogeneous nucleation, while the constant $C_{het}$ is estimated using the experimental value $x_l^{het}(290^{\circ}{\rm C})\equiv x_{l,exp}^{het}$. For numerical estimates we use the value $x_{l,exp}^{het}\simeq 0.06\%$ given in Refs. [@Miller-06] and [@Miller-09], which yields: $ C_{het}\simeq 35$. In figure 8 we show positions of the homogeneous and the heterogenous precipitation limits in the $(x,T)$ plane, $T_l^{hom}(x)$ and $T_l^{het}(x)$, calculated for dilute Fe-Cu alloys using Eqs. (\[F\_cl-hom\]) and (\[F\_cl-het\]). For comparison, in Fig. 8 we also show the binodal curve $T_b(x)$ (the copper solubility limit) estimated in [@SF-07] from experimental data and $ab \ initio$ calculations. The results in Fig. 8 are presented for the temperature interval between 290 and 450$^{\circ}$C used for annealing of RPVs [@Auger-00; @Miller-09]. The value $C_{hom}=27.6$ used in Fig. 8 corresponds to $x_l^{hom}(290^{\circ}{\rm C})=0.088\%$ for which Radiguet et al. [@Radiguet-07] did not observe any homogeneous precipitation, while the constant $C_{het}=35$ used in Fig. 8 corresponds to the estimate $x_l^{het}= 0.06\%$ given by Miller et al. [@Miller-06; @Miller-09]. Our calculations also show that a slight variation of the constant $C_{hom}$ or $C_{het}$ leads to an “almost rigid” shift of the curve $T_l^{hom}(x)$ or $T_l^{het}(x)$ in Fig. 7. For example, when this variation of $C_{hom}$ leads to the shift $\delta x_l^{hom}$=0.05% to the right at $290^{\circ}$C, the analogous shift at $450^{\circ}$C is almost the same: $\delta x_l^{hom}(450^{\circ}{\rm C})\simeq 0.06\%$. The most interesting general feature seen in Fig. 8 seems to be a rather sharp temperature dependence of the precipitation limits, particularly for the heterogeneous precipitation. For example, the $x_l^{het}$ value at $T=450^{\circ}$C exceeds that at $T=290^{\circ}$C by almost five times, while the $x_l^{hom}$ value increases for this temperature interval by about 3.5 times. These qualitative conclusions can be useful, in particular, for the interpretation of microstructural observations related to the annealing of RPVs [@Miller-09]. Note also that the variations of both precipitation limits, $x_l^{het}(T)$ and $x_l^{hom}(T)$, with temperature for the temperature interval considered are much stronger than those for the solubility limit $x_b(T)$. To conclude this section, we again note that our equations (\[F\_cl-hom\]) and (\[F\_cl-het\]) for temperature dependencies $x_l^{het}(T)$ and $x_l^{hom}(T)$ are of course tentative and have not been proved formally. At the same time, physical considerations about the dominant role of the activation factor $\exp\,(-\beta F_c)$ in suppressing precipitation at low $x$ used in our derivation seem to be rather plausible, particularly for the homogeneous precipitation. In the derivation of Eq. (\[F\_cl-het\]) for the heterogeneous precipitation, we also used an assumption about a weak temperature dependence of the ratio $F_c^{het}/F_c^{hom}$ for the temperature interval considered which may seem to be less evident. Therefore, the accuracy of Eq. (\[F\_cl-het\]) for $x_l^{het}(T)$ can, generally, be lower than the accuracy of Eq. (\[F\_cl-hom\]) for $x_l^{hom}(T)$. However, one may expect that the main qualitative features of both dependencies, $T_l^{hom}(x)$ and $T_l^{het}(x)$, shown in Fig. 8 are correctly described by the simple model used. Conclusions =========== Let us summarize the main results of this work. We study kinetics of homogeneous nucleation and growth of copper precipitates under electron irradiation of iron-copper alloys at low concentrations $x=0.06-0.4$ at.% and temperatures $T=290-450^{\circ}$C used for service of a number of nuclear reactor pressure vessels [@Akamatsu-95]-[@Meslin-10b]. The earlier-described kinetic Monte Carlo (KMC) modeling [@SF-07] and the statistical theory of nucleation (STN) [@DV-98a; @DV-98b] are used. The $ab \ initio$ model of interatomic interactions which describes well the available data about precipitation in Fe-Cu alloys at different $x$ and $T$ [@SF-07; @KVZ-13] is used for both the KMC simulations and the STN calculations. The conventional assumption [@Mathon-97] about the similarity of mechanisms of precipitation under electron irradiation and under thermal aging is also adopted. Then precipitation under electron irradiation can be described by the KMC codes developed by Soisson and Fu [@SF-07] for studies of precipitation during thermal aging but with the acceleration of kinetics under irradiation characterized by the acceleration factor $A_{irr}$ in Eqs. (\[D\_Cu\]) and (\[t\_irr\]). We estimate this acceleration factor for dilute iron-copper alloys considered, and our estimate reasonably agrees with the available experimental data [@Mathon-97]. Our STN-based calculations of the nucleation barrier $F_c$ in the Zeldovich-Volmer relation (\[J-J\_0\]) for the nucleation rate $J$ show that this nucleation barrier varies with $x$ and $T$ very sharply. Thus the concentration and temperature dependencies of the nucleation rate are mainly determined by the variations of the activation factor $\exp\,(-\beta F_c)$ with $x$ or $T$. We also found that at $T=290^{\circ}$C, the interval of concentrations $x$ for which this activation factor starts to fall off very rapidly just corresponds to the interval of positions of the homogeneous precipitation limit $x_l^{hom}$ estimated in experiments [@Mathon-97; @Auger-00; @Radiguet-07]. Our KMC simulations for the dilute alloys considered describe mainly the very initial stage of precipitation, that of the steady-state nucleation characterized by the constant nucleation rate $J$. We also use these KMC simulations to study kinetics of growth of a supercritical embryo and confirm the validity of the phenomenological equation of the classical theory of nucleation which describes this growth [@Lif-Pit-79]. Our simulations also enable us to estimate the effective diffusivity $D_{eff}$ which enters into the STN expression for the nucleation rate $J$. The $D_{eff}$ values are found to exceed the dilute alloy values $D_0(T)$ by two-three orders of magnitude. This effective acceleration of diffusion seems to be mainly due to the high mobility of small copper clusters found earlier for the precipitation in Fe-1.34%Cu alloy at $T=500^{\circ}$C [@SF-07; @Jourdan-10]. The KMC estimates of the nucleation rate $J$ and the effective diffusivity $D_{eff}$ described above enable us to determine the nucleation barrier $F_c$ using Eq. (\[F\_c-KMC\]) based on the KMC simulations. The resulting values of the reduced nucleation barrier $(\beta F_c)_{\rm KMC}$ are compared with the $(\beta F_c)_{\rm STN}$ values calculated using the statistical theory of nucleation [@DV-98a; @DV-98b]. We find that for the dilute alloys with $x\leq 0.2\%$, the STN and the KMC results for the nucleation barrier $F_c$ coincide within about a percent. It seems to confirm a high reliability of the STN for this problem. Making a plausible assumption that the position of the homogeneous nucleation limit $x_l^{hom}$ is mainly determined by the value of the activation factor $\exp\,[-\beta F_c(x,T)]$ which at $x=x_l^{hom}$ takes an approximately same value $\exp\,(-\beta F_c)_l$ for all temperatures $T$ considered, we use our STN-based calculations of reduced nucleation barriers $\beta F_c$ to estimate the temperature dependence $x_l^{hom}(T)$. For temperatures between 290 and 450$^{\circ}$C, this dependence is found to be rather sharp, much sharper than that for the solubility limit $x_{sol}(T)$. Making an additional assumption about a weak temperature dependence of the ratio of nucleation barriers for the heterogeneous and the homogeneous nucleation, $F_c^{het}/F_c^{hom}$, we also estimate the temperature dependence of the heterogeneous precipitation limit $x_l^{het}(T)$ which corresponds to the neutron or ion irradiations. The dependence $x_l^{het}(T)$ is found to be still more sharp than that for the homogeneous precipitation. In spite of the evidently tentative character of these estimates, one can expect that the main qualitative features of both dependencies, $x_l^{hom}(T)$ and $x_l^{het}(T)$, are described by this model properly. Acknowledgements {#acknowledgements .unnumbered} ================ We are very grateful to V. Yu. Dobretsov for his help in the computations using the STN, as well as to A. Barbu and K. Yu. Khromov, for fruitful discussions. The work was supported by the Russian Fund of Basic Research (grant No. 12-02-00093); by the fund for support of leading scientific schools of Russia (grant No. NS-215.2012.2); and by the program of Russian university scientific potential development (grant No. 2.1.1/4540). [10]{} M. Akamatsu, J. C. Van Duysen, P. Pareige, P Auger, J. Nucl. Mater. [**225**]{}, 192 (1995). M.H. Mathon, A. Barbu, F. Dunstetter, F. Maury, N. Lorenzelli, and C.H. de Novion, J. Nucl. Mater. [**245**]{}, 224 (1997). P. Auger, P. Pareige, S. Weltzel, J.-C. Van Duysen, J. Nucl. Mater. [**280**]{}, 331 (2000). A. Harduin Duparc, C. Mingeon, N. Smetniansky-de-Grande, A. Barbu, J. Nucl. Mater. [**399**]{}, 143 (2002). M.K. Miller, M.A. Sokolov, R.K. Nanstad, K. F. Russell, J. Nucl. Mater. [**351**]{}, 187 (2006). B. Radiguet, A. Barbu, P. Pareige, J. Nucl. Mater. [**360**]{}, 104 (2007). M.K. Miller, A.A. Chernobaeva. Y.I. Shtrombakh, K. F. Russell, R.K. Nanstad, D.Y. Erak, O.O. Zabuzov, J. Nucl. Mater. [**385**]{}, 615 (2009). E. Meslin, B. Radiguet, P. Pareige, A. Barbu, J. Nucl. Mater. [**399**]{}, 137 (2010). E. Meslin, M. Lambrecht, et al., J. Nucl. Mater. [**406**]{}, 73 (2010). F. Soisson and C.-C. Fu, Phys. Rev. B [**76**]{}, 214102 (2007). K.Yu. Khromov, V.G. Vaks and I.A. Zhuravlev, JETP [**116**]{}, 236 (2013). V.Yu. Dobretsov and V.G. Vaks, J. Phys.: Condensed Matter [**10**]{}, 2261 (1998). V.Yu. Dobretsov and V.G. Vaks, J. Phys.: Condensed Matter [**10**]{}, 2275 (1998). E. M. Lifshits, L. P. Pitaevsky, [*Physical kinetics*]{} (Moscow, Nauka, 1979), J. W. Cahn and J. E. Hilliard, J. Chem. Phys. [**31**]{}, 88 ( 1959) J. S. Langer, Ann. Phys. (N. Y.) [**54**]{}, 228 (1969) V. G. Vaks, Phys. Reports [**391**]{}, 157-242 (2004). K.Yu. Khromov, F. Soisson, A.Yu. Stroev and V.G. Vaks, JETP [**112**]{}, 415 (2011). T. Nagano and M. Enomoto, Scripta Mater. [**55**]{}, 223 (2006). T. Philippe and D. Blavette, J. Chem. Phys. [**135**]{}, 134508 (2011). T. Jourdan, F. Soisson, E. Clouet, A. Barbu, Acta Mat. [**58**]{}, 3400 (2010). R. Sizmann, J. Nucl. Mater. [**69-70**]{}, 386-412 (1978). V. G. Vaks, I. A. Zhuravlev, JETP [**115**]{}, 635 (2012). F. Soisson and G. Martin, Phys. Rev. B [**62**]{}, 203 (2000). A. D. Le Claire, J. Nucl. Mater. [**69-70**]{}, 70 (1978). F. A. Nichols, J. Nucl. Mater. [**75**]{}, 32-41 (1978). G. Lucas and R. Shaublin, Nucl. Instrum. Meth. B [**267**]{}, 3009 (2009). C. C. Fu, F. Willaime and P. Ordejon, Phys. Rev. Lett. [**92**]{}, 175503 (2004). S.R. Goodman, S.S. Brenner and J.R. Low, Metall. Trans. [**4**]{}, 2363, 2371 (1973). R. Kampmann and R. Wagner, in: [*Atomic Transport and defects in Metals by Neutron Scattering*]{}, ed. by C. Janot et al., Springer, New York (1986) p.73. J. W. Martin, R. D. Doherty [*Stability of microstructure in metallic systems*]{}, Cambridge Univ. Press (1976), Sec. 2.3.1. J. W. Martin, R. D. Doherty [*Stability of microstructure in metallic systems*]{}, Cambridge Univ. Press (1976), Sec. 2.2.3. ------------------------- ------- ------- ------- ------- ------- ------- ------- $x$, % $T$, $^{\circ}$C $L$ $L_1$ $L_3$ $L_1$ $L_2$ $L_3$ $L_1$ $L_2$ $(\beta F_c)_{\rm KMC}$ 20.1 20.0 21.0 22.1 21.5 22.1 21.9 $(\beta F_c)_{\rm STN}$ 19.8 19.8 20.8 20.8 20.8 21.6 21.6 ------------------------- ------- ------- ------- ------- ------- ------- ------- [^1]: $^\ast$Corresponding author. Email: [email protected]
{ "pile_set_name": "ArXiv" }
--- abstract: 'The exact description of many-body quantum systems represents one of the major challenges in modern physics, because it requires an amount of computational resources that scales exponentially with the size of the system. Simulating the evolution of a state, or even storing its description, rapidly becomes intractable for exact classical algorithms. Recently, machine learning techniques, in the form of restricted Boltzmann machines, have been proposed as a way to efficiently represent certain quantum states with applications in state tomography and ground state estimation. Here, we introduce a practically usable deep architecture for representing and sampling from probability distributions of quantum states. Our representation is based on variational auto-encoders, a type of generative model in the form of a neural network. We show that this model is able to learn efficient representations of states that are easy to simulate classically and can compress states that are not classically tractable. Specifically, we consider the learnability of a class of quantum states introduced by Fefferman and Umans. Such states are provably hard to sample for classical computers, but not for quantum ones, under plausible computational complexity assumptions. The good level of compression achieved for hard states suggests these methods can be suitable for characterizing states of the size expected in first generation quantum hardware.' author: - Andrea Rocchetto - Edward Grant - Sergii Strelchuk - Giuseppe Carleo - Simone Severini bibliography: - 'bibliography\_qVAE.bib' title: Learning hard quantum distributions with variational autoencoders --- Introduction {#introduction .unnumbered} ============ One of the most fundamental tenets of quantum physics is that the physical state of a many-body quantum system is fully specified by a high-dimensional function of the quantum numbers, the wave-function. As the size of the system grows the number of parameters required for its description scales exponentially in the number of its constituents. This complexity is a severe fundamental bottleneck in the numerical simulation of interacting quantum systems. Nonetheless, several approximate methods can handle the exponential complexity of the wave function in special cases. For example, *quantum Monte Carlo* methods (QMC), allow to sample exactly from many-body states free of sign problem [@nightingale1998quantum; @gubernatis2016quantum; @suzuki1993quantum], and *Tensor Network approaches* (TN), very efficiently represent low-dimensional states satisfying the area law for entanglement [@verstraete2008matrix; @orus2014practical]. Recently, machine learning methods have been introduced to tackle a variety of tasks in quantum information processing that involve the manipulation of quantum states. These techniques offer greater flexibility and, potentially, better performance, with respect to the methods traditionally used. Research efforts have focused on representing quantum states in terms of *restricted Boltzmann machines* (RBMs). The RBM representation of the wave function, introduced by Carleo and Troyer [@carleo2017solving], has been successfully applied to a variety of physical problems, ranging from strongly correlated spins [@carleo2017solving; @deng_quantum_2017], and fermions [@nomura_restricted-boltzmann-machine_2017] to topological phases of matter [@deng_exact_2016; @glasser_neural_2017; @kaubruegger_chiral_2017]. Particularly relevant to our purposes is the work by Torlai *et al.* [@torlai2017many] that makes use of RBMs to perform *quantum state tomography* of states whose evolution can be simulated in polynomial time using classical methods (*e.g.* *matrix product states* (MPS) [@perez2006matrix]). Although it is remarkable that RBMs can learn an efficient representation of this class of states without any explicitly programmed instruction, it remains unclear how the model behaves on states where no efficient classical description is available. Theoretical analysis of the representational power of RBMs has been conducted in a series of works [@gao2017efficient; @chen_equivalence_2017; @huang_neural_2017; @deng_quantum_2017; @clark_unifying_2017]. Gao and Duan, in particular, showed that RBMs cannot efficiently encode every quantum state [@gao2017efficient]. They proved that Deep Boltzmann Machines (DBMs) with complex weights, a multilayer variant of RBMs, can efficiently represent most physical states. Although this result is of great theoretical interest the practical application of complex-valued DBMs in the context of unsupervised learning has not yet been demonstrated due to a lack of efficient methods to sample efficiently from DBMs when the weights are complex-valued. The absence of practically usable deep architectures remains an important limitation of current neural network based learning methods for quantum systems. Indeed, several research efforts on neural networks [@mhaskar2016learning; @telgarsky2016benefits; @eldan2016power] have shown that depth significantly improves the representational capability of networks for some classes of functions (such as compositional functions). In this Paper, we address several open questions with neural network quantum states. First, we study how the depth of the network affects the ability to compress quantum many-body states. This task is achieved upon introduction of a deep neural network architecture for encoding probability distribution of quantum states, based on *variational autoencoders* (VAEs) [@kingma2013auto]. We benchmark the performance of deep networks on states where no efficient classical description is known, finding that depth systematically improves the quality of the reconstruction for states that are computationally tractable and for hard states that can be efficiently constructed with a quantum computer. Surprisingly, the same does not apply for hard states that cannot be efficiently constructed by means of a quantum process. Here, depth does not improve the reconstruction accuracy. Second, we show that VAEs can learn efficient representations of computationally tractable states and can reduce the number of parameters required to represent an hard quantum state up to a factor $5$. This improvement makes VAE states a promising tool for the characterization of early quantum devices that are expected to have a number of qubits that is slightly larger than what can be efficiently simulated using existing methods [@boixo2016characterizing]. Encoding quantum probability distributions with VAEs {#sec:qVAE .unnumbered} ---------------------------------------------------- Variational autoencoders (VAEs), introduced by Kingma and Welling in 2013 [@kingma2013auto], are generative models based on layered neural networks. Given a set of i.i.d. data points $X=\{x^{(i)}\}$, where $x^{(i)}\in\mathbb{R}^{n}$, generated from some distribution $p_{\theta}(x^{(i)}|z)$ over Gaussian distributed latent variables $z$ and model parameters $\theta$, finding the posterior density $p_{\theta}(z|x^{(i)})$ is often intractable. VAEs allow for approximating the true posterior distribution, with a tractable approximate model $q_{\phi}(z|x^{(i)})$, with parameters $\phi$, and provide an efficient procedure to sample efficiently from $p_{\theta}(x^{(i)}|z)$. The procedure does not employ Monte Carlo methods. As shown in Fig. \[fig:VAE\_schematic\] a VAE is composed of three main components. The *encoder* that is used to project the input in the *latent space* and the *decoder* that is used to reconstruct the input from the latent representation. Once the network is trained the encoder can be dropped and, by generating samples in the latent space, it is possible to sample according to the original distribution. In graph theoretic terms, the graph representing a network with a given number of layers is a *blow up* of a directed path on the same number of vertices. Such a graph is obtained by replacing each vertex of the path with an independent set of arbitrary but fixed size. The independent sets are then connected to form complete bipartite graphs. The model is trained by minimizing over $\theta$ and $\phi$ the cost function: $$\begin{gathered} J(\theta,\phi,x^{(i)})=-\mathds{E}_{z\sim{q_{\phi}(z|x^{(i)})}}[\log p_{\theta}(x^{(i)}|z)]\\ +D_{\mathrm{KL}}(q_{\phi}(z|x^{(i)})||p_{\theta}(z))).\label{eq:jtheta}\end{gathered}$$ The first term (reconstruction loss) $-\mathds{E}_{z\sim{q_{\phi}(z|x^{(i)})}}[\log p_{\theta}(x^{(i)}|z)]$ is the expected negative log-likelihood of the $i$-th data-point and favors choices of $\theta$ and $\phi$ that lead to more faithful reconstructions of the input. The second term (regularization loss) $D_{\mathrm{KL}}(q_{\phi}(z|x^{(i)})||p_{\theta}(z)))$ is the Kullback-Leibler divergence between the encoder’s distribution $q_{\phi}(z|x^{(i)})$ and the Gaussian prior on $z$. A full treatment and derivations of the variational objective are given in [@kingma2013auto]. VAEs can be used to encode the probability distribution associated to a quantum state. Let us consider an $n$-qubit quantum state $\ket{\psi}$, with respect to a basis $\{\ket{b_{i}}\}_{i=1,\dots,2^{n}}$. We can write the probability distribution corresponding to $\ket{\psi}$ as $p(b_{i})=|\langle b_{i}|\psi\rangle|^{2}$. If we consider the computational basis, we can write $\ket{\psi}=\sum_{i=1}^{2^{n}}\psi_{i}\ket{i}$, where each basis element corresponds to an $n$-bit string. A VAE can be trained to generate basis elements $\ket{i}$ according to the probability $p(i)=|\langle i|\psi\rangle|^{2}=|\psi_{i}|^{2}$. We note that, in principle, it is possible to encode a full quantum state (phase included) in a VAE. This requires samples taken from more than one basis and a network structure that can distinguish among the different inputs. The development of VAE encodings for full quantum states will be left to future work. We approximate the true posterior distribution across measurement outcomes in the latent space $z$ with a multivariate Gaussian, having diagonal covariance structure, zero mean and unit standard deviation. The training set consists of a set of basis elements generated according to the distribution associated with a quantum state. Following training, the variables $z$ are sampled from a multivariate Gaussian and used as the input to the decoder. By taking samples from this Gaussian as input, the decoder is able to generate strings corresponding to measurement outcomes that closely follow the distribution of measurement outcomes used to train the network. ![\[fig:VAE\_schematic\]**Encoding quantum probability distributions with VAEs.** A VAE can be used to encode and then generate samples according to the probability distribution of a quantum state. Each dot corresponds to a neuron and neurons are arranged in layers. Input (top), latent, and output (bottom) layers contain $n$ neurons. The number of neurons in the other layers is a function of the compression and the depth. Layers are fully connected with each other with no intra layer connectivity. The network has three main components: the encoder (blue neurons), the latent space (green), and the decoder (red). Each edge of the network is labelled by a weight $\theta$. The total number of weights $m$ in the decoder corresponds to the number of parameters used to represent a quantum state. The network can approximate quantum states using $m<2^{n}$ parameters. The model is trained using a dataset consisting of basis elements drawn according to the probability distribution of a quantum state. Elements of the basis are presented to the input layer on top of the encoder and, during the training phase, the weights of the network are optimized in order to reconstruct the same basis element in the output layer.](VAE_net_schematic.pdf) Hard and easy quantum states {#sec:hard .unnumbered} ---------------------------- In this section we introduce a method to classify quantum states based on the hardness of sampling their probability distribution in a given basis. This will be used to assess the power of deep neural network models at representing many-body wave-functions. We now proceed to define two concepts that will be frequently used throughout the paper and form the basis of our classification method: *reconstruction accuracy* and *compression*. Let $\rho$ and $\sigma$ be $n$–qubit quantum states. We say that $\sigma$ is a good representation of $\rho$ if the fidelity $F = \mathrm{Tr}(\sqrt{\rho^{1/2}\sigma \rho^{1/2}}) \geq 1 - \epsilon$ for an $\epsilon>0$. This accuracy metric cannot be immediately applied to the analysis of VAEs, that can only encode the probability distribution associated to a state. We now show that the fidelity can expressed in terms of the probability distributions over a measurement that maximally distinguishes the two states. Let $E=\{E_i\}$ be a POVM measurement. Then, using a result by Fuchs and Caves [@fuchs1994ensemble] we can write $$\label{eq:fidelity} F = \min_{E} \sum_i \sqrt{\mathrm{Tr}(E_i \rho)\mathrm{Tr}(E_i \sigma)},$$ where the minimum is taken over all possible POVMs. Note that $p(i) = \mathrm{Tr}(E_i \rho)$ and $q(i) = \mathrm{Tr}(E_i \sigma)$ are the probabilities of measuring the state $\rho$ and $\sigma$, respectively, in outcome labelled by $i$ and $\sum_i \sqrt{p(i)q(i)}$ is the Bhattacharyya coefficient between the two distributions. Using Eq. \[eq:fidelity\] we can relate the complexity of a state with the problem of estimating the fidelity $F$. This corresponds to the hardness of sampling the probability distribution $p(i) = \mathrm{Tr}(E'_i \rho)$, where $E'$ minimises Eq. \[eq:fidelity\] (here we assume that sampling from the approximating distribution $q(i)$ is at most as hard as sampling from $p(i)$). Throughout the paper, unless where explicitly mentioned, we will work with states that have only positive, real entries in the computational basis. In this case, it is easy to see that the Bhattacharyya coefficient between the distributions reduces to the fidelity and, hence, measurements in the $Z$ basis minimises Eq \[eq:fidelity\]. We remark that, if it is not possible to find a POVM for which Eq. \[eq:fidelity\] is minimised it is always possible to use the standard formulation of the fidelity as a metric in the context of VAEs. This can be accomplished by making use of $3$ VAEs to encode the state $\sigma$ over $3$ different basis. By using standard tomographic techniques, like maximum likelihood, measurements in a complete basis can be used to reconstruct the full density matrix. In order to connect the above definition of state complexity with VAEs we introduce the *compression factor*. Given an $n$-qubit state that is represented by a VAE with $m$ parameters in the decoder, the compression factor is $C = \frac{m}{2^{n}}$. We say that a state $\rho$ is *exponentially compressible* if there exists a network that approximates $\rho$ with high accuracy using $m = O(\mathrm{poly}(n))$ parameters. Once a network is trained, the cost of generating a sample is proportional to the number of parameters in the network. In this sense the complexity of a state is parametrised by the number of parameters used by a neural network representation. Based on these observation we define *easy states* those that can be represented with high accuracy and exponential compression and *hard states* those that can be represented with high accuracy using at least $O(\mathrm{exp}(n))$ parameters. The last category includes: 1) states that can be efficiently sampled with a quantum computer, but are conjectured to have no classical algorithm to do so; 2) states that cannot be efficiently obtained on a quantum computer starting from some fixed product input state (*e.g.* random states). Under this definition, states that admit an efficient classical description (such as stabilizer states or MPS with low bond dimension) are easy, because we known that $O(\mathrm{poly}(n))$ parameters are sufficient to specify the state. Specifically, for the class of easy states we consider separable states obtained by taking the tensor product of $n$ different $1$-qubit random states. More formally, we consider states of the form $\ket{\tau}=\bigotimes_{i=1}^{n}\ket{r_{i}}$ where $\ket{r_{i}}$ are random $1$-qubit states. These states can be described using only $2n$ parameters. Among the class of hard states of the first kind, we study the learnability of a type of hard distributions introduced in [@fefferman2014power] which can be sampled exactly on a quantum computer. These distributions are conjectured to be hard to approximately sample from classically the existence of an efficient sampler would lead to the collapse of the Polynomial Hierarchy under some natural conjectures described in [@fefferman2014power; @aaronson2011computational]. We discuss how to generate this type of states in the Methods section. Finally, for the second class of hard states, we consider random pure states. These are generated by normalizing a $2^{n}$ dimensional complex vector drawn from the unit sphere according to the Haar measure. Results {#results .unnumbered} ======= The role of depth in compressibility {#sec:depth .unnumbered} ------------------------------------ Classically, depth is known to play a significant role in the representational capability of a neural network. Recent results, such as the ones by Mhaskar, Liao, and Poggio [@mhaskar2016learning], Telgarsky [@telgarsky2016benefits], and Eldan and Shamir [@eldan2016power] showed that some classes of functions can be approximated by deep networks with the same accuracy as shallow networks but with exponentially less parameters. The representational capability of networks that represent quantum states remains largely unexplored. Some of the known results are only based on empirical evidence and sometimes yield to unexpected results. For example, Morningstar and Melko [@morningstar2017deep] showed that shallow networks are more efficient than deep ones when learning the energy distribution of a $2$-dimensional Ising model. In the context of the learnability of quantum states Gao and Duan [@gao2017efficient] proved that DBMs can efficiently represent some states that cannot be efficiently represent by shallow networks (*i.e.* states generated by polynomial depth circuits or $k$-local Hamiltonians with polynomial size gap) [[ using a polynomial number of hidden units]{}]{}. However, there are no known methods to sample efficiently from DBMs when the weights include complex-valued coefficients. We benchmark with numerical simulations the role played by depth in compressing states of different levels of complexities. We focus on three different states: an easy state (the completely separable state discussed in the previous section), a hard state (according to Fefferman and Umans), and a random pure state. Our results are presented in Fig. \[fig:overlap\_v\_layers\]. Here, by keeping the number of parameters in the decoder constant, we determine the reconstruction accuracy of networks with increasing depth. Remarkably, depth affects the reconstruction accuracy of hard quantum states. This might indicate that VAEs are able to capture correlations in hard quantum states. As a sanity check we notice that the network can learn correlations in random product states and that depth does not affect the learnability of random states. Our simulations suggest a further link between neural network and quantum states. This topic has recently received the attention of the community. Specifically, Levine *et al.* [@levine2017deep] demonstrated that convolutional rectifier networks with product pooling can be described as tensor networks. By making use graph theoretic tools they showed that nodes in different layers model correlations across different scales and that adding more nodes to deeper layers of a network can make it better at representing non-local correlations. Efficient compression of physical states {#sec:probing .unnumbered} ---------------------------------------- In this section we focus our attention onto two questions: can VAEs find efficient representations of easy states? What level of compression can we obtain for hard states? Through numerical simulations we show that VAEs can learn to efficiently represent some easy states (that are challenging for standard methods) and achieve good levels of compressions for hard states. Remarkably, our methods allow to compress up to a factor $5$ the hard quantum states introduced in [@fefferman2015power]. We remark that the exponential hardness cannot be overcome for general quantum states and our methods achieve only a factor improvement on the overall complexity. This may nevertheless be sufficient to be used as a characterisation tool where full classical simulation is not feasible. We test the performance of the VAE representation on two classes of states: the hard states that can be constructed efficiently with a quantum computer introduced by Fefferman and Umans [@fefferman2015power] and states that can be generated with a long-range Hamiltonian dynamics, as found for example in experiments with ultra-cold ions [@richerme2014non]. The states generated through this evolution are highly symmetric physical states. However, due to the bond dimension increasing exponentially with the evolution time, these states are particularly challenging for MPS methods. An interesting question is to understand whether neural networks are able to exploit these symmetries and represent these states efficiently. We describe long-range Hamiltonian dynamics in the Methods section. Results are displayed in Fig. \[fig:overlap\_v\_compression\]. For states obtained through Hamiltonian evolution we achieve with almost maximum reconstruction accuracy compression levels of up to $C \approx 10^{-3}$. This corresponds to a number of parameters $m=\mathcal{O}(100)\ll2^{18}$ which implies that the VAE has learned an efficient representation of the state. In the case of hard state we can reach a compression of $0.2$, corresponding to a factor $5$ reduction in the number of parameters required to represent the state. Note that the entanglement properties of hard states [[ are likely to]{}]{} make them hard to compress for tensor network states. For example, [[ if one wanted to compress an $18$ qubits state using MPS (a type of tensor network that is known to be efficiently contractable)]{}]{} we have found that the estimated bond dimension to reconstruct this state is $D \approx 460$. This number is obtained computing the largest bipartite entanglement entropy ($S$), and estimating the bond dimension with $D \approx 2^S$. Considering that an MPS has $D^2$ variational parameters (in the best case), this would yield about $200$ thousands variational parameters required to represent those hard states. The resulting MPS compressing factor is then about $1.23$, a significantly lower figure with respect to the $5$ compression factor obtained with VAEs. [[ We note that this calculation only shows that the entanglement structure of hard states is not well modelled by MPS. Other types of tensor networks might be more amenable to the specific structure of these states but it is unlikely these models will be computationally tractable.]{}]{} Although limited, the levels of compression we achieve for hard states could play a role in experiments aimed at showing quantum supremacy. In this setting a quantum machine with a handful of noisy qubits performs a task that is not reproducible even by the fastest supercomputer. As recently highlighted by Montanaro and Harrow [@harrow2017quantum] one of the key challenges with quantum supremacy experiments is to verify that the quantum machine is behaving as expected. Because quantum computers are conjectured to not be efficiently simulatable, verifying that a quantum machine is performing as expected is a hard problem for classical machines. The paper by Jozsa and Strelchuk [@jozsa2017efficient] provides an introduction to several approaches to verification of quantum computation. Our methods might allow to characterise the result of a computation by reducing the complexity of the problem. [[ Because any verification of quantum supremacy will likely involve a machine with only a few qubits above what can be efficiently classically simulated, even small reductions in the number of parameters of the state might allow to approximate relevant quantities in a computationally tractable way. Potentially, a neural network approach to verification]{}]{} can be accomplished by compressing a trusted initial state into a VAE whose parameters are then evolved according to a set of rules specified by the quantum circuit. By comparing the experimental distribution with the one sampled with the VAE it is then possible to determine whether the device is faulty. [[ We remark that this type of verification protocol would only “approximately verify” the system because of the errors introduced during the compression phase.]{}]{} Discussion {#sec:conclusions .unnumbered} ========== In this work we introduced VAEs, a type of deep, generative, neural network, as way to encode the probability distribution of quantum states. Our methods are completely unsupervised, *i.e.* do not require a labelled training set. By means of numerical simulations we showed that deep networks can represent hard quantum states that can be efficiently obtained by a quantum computer better than shallow ones. On the other hand, for states that are hard and conjectured to be not efficiently producible by quantum computers, depth does not appear to play a role in increasing the reconstruction accuracy. Our results suggest that neural networks are able to capture correlations in states that are provably hard to sample from for classical computers but not for quantum ones. As already pointed out in other works, this might signal that states that can be produced efficiently by a quantum computer have a structure that is well represented by a layered neural network. Through numerical experiments we showed that our methods have two important features. First, they are capable of representing, using fewer parameters, states that that are known to have efficient representation but where other classical approaches struggle. Second, VAEs can compress hard quantum states up to a constant factor. However low, this compression level might enable to approximately verify quantum states of a size expected on near future quantum computers. Presently, our methods allow to encode only the probability distribution of a quantum state. Future research should focus on developing VAE architectures that allow to reconstruct the full set of amplitudes. Other interesting directions involve finding methods to compute the quantum evolution of the parameters of the network and investigating whether the depth of a quantum circuit is related to the optimal depth of a VAE learning its output states. Finally, it is interesting to investigate how information is encoded in the latent layers of the network. Such analysis might provide novel tools to understand the information theoretic properties of a quantum system. Methods {#methods .unnumbered} ======= Numerical experiments {#app:numsim .unnumbered} --------------------- All our networks were trained using the tensorflow r1.3 framework on a single NVIDIA K80 GPU. Training was performed using backpropagation and the Adam optimiser with initial learning rate of $10^{-3}$ [@kingma2014adam]. Leaky rectified linear units (LReLU) function were used on all hidden layers with the leak set to $0.2$ [@maas2013rectifier]. Sigmoid activation functions were used on the final layer. [[ Training involves optimising two objectives: the reconstruction loss and the regularization loss. We used a warm up schedule on the regularisation objective by increasing a weight on the regularisation loss from $0$ to $0.85$ linearly during training  [@sonderby2016ladder]. This turned out to be critical, especially for hard states. A consequence of this approach is that the model does not learn the distribution until close to the end of training irrespective of the number of training iterations. Each network was trained using $50,000$ batches of $1000$ samples each. Each sample consists of a binary string representing a measurement outcome.]{}]{} Following training the state was reconstructed from the VAE decoder by drawing $100(2^{n})$ samples from a multivariate Gaussian with zero mean and unit variance. The samples were decoded by the decoder to generate measurement outcomes in the form of binary strings. The relative frequency of each string was recorded and used to reconstruct the learned distribution which was compared to the true distribution to determine its fidelity. In all experiments the number of nodes in the latent layer is the same as the number of qubits. Using fewer or more nodes in this layer resulted in worse performance. The number of nodes in the hidden layers is determined by the number of layers and the compression $C$ defined by $\frac{m}{2^{n}}$ where $n$ is the number of qubits and $m$ is the number of parameters in the decoder. In all cases the encoder has the same number of hidden layers and nodes in each layer as the decoder. We compress the VAE representation of a quantum state by removing neurons from each hidden layer of the VAE. For small $n$’s achieving a high level of compression caused instabilities in the network (*i.e.* the reconstruction accuracy became more dependent on the weight initialisation). In this respect we note that, by restricting the number of neurons in the penultimate layer, we are effectively constraining the number of possible basis states that can be expressed in the output layer and, as a result, the number of configurations the VAE can sample from. This can be shown noting that the activation functions of the penultimate layer generate a set of linear inequalities that must be simultaneously satisfied. A geometric argument that involves how many regions of an $n$-dimensional space $m$ hyperplanes can separate lead to conclude that, to have full expressive capability, the penultimate layer must include at least $n$ neurons. Similar arguments have been discussed in [@huang1991bounds] for multilayer perceptrons. States that are classically hard to sample from {#app:hardgen .unnumbered} ----------------------------------------------- We study the learnability of a special class of hard states introduced by Fefferman and Umans [@fefferman2015power] which is produced by a certain quantum computational processes which exhibit quantum “supremacy”. The latter is a phenomenon whereby a quantum circuit which consists of quantum gates and measurements on a constant number of qubit lines samples from a particular class of distributions which is known to be hard to sample from on a classical computer modulo some very plausible computational complexity assumptions. To demonstrate quantum supremacy one only requires quantum gates to operate within a certain fidelity without full error-correction. This makes efficient sampling from such distributions feasible to execute on near-term quantum devices and opens the search for possibilities to look for practically-relevant decision problems. To construct a distribution one starts from an encoding function $h:[m]\to\{0,1\}^{N}$. The function $h$ performs an efficient encoding of its argument and is used to construct the following so-called efficiently specifiable polynomial on $n$ variables: $$Q(X_{1},\dots,X_{N})=\sum_{z\in[m]}X_{1}^{h(z)_{1}}\dots X_{N}^{h(z)_{N}},$$ where $h(z)_{i}$ means that we take only the $i$-th bit, and $m$ is an arbitrary integer. In the following, we pick $h$ to be related to the permanent. More specifically, $h:[0,n!-1]\to\{0,1\}^{n^{2}}$ maps the $i$-th permutation (out of $n!$) to a string which encodes its $n\times n$ permutation matrix in a natural way resulting in a $N$-coordinate vector, where $N=n^{2}$. To encode a number $A\in[0,n!-1]$ in terms of its permutation vector we first represent $A$ in factorial number system to get $A'$ obtaining the $N$-coordinate vector which identifies a particular permutation $\sigma$. With the above encoding, our efficiently specifiable polynomial $Q$ will have the form: $$Q(X_{1},\dots,X_{N})=\sum_{z\in[n!-1]}X_{1}^{h(z)_{1}}\dots X_{N}^{h(z)_{N}}.$$ Fix some number $L$ and consider the following set of vectors $y=(y_{1},\ldots,y_{N})\in[0,L-1]^{N}$ (i.e. each $y_{j}$ ranges between $0$ and $L-1$). For each $y$ construct another vector $Z_{y}=(z_{y_{1}},\dots,z_{y_{N}})$ constructed as follows: each $z_{y_{j}}$ corresponds to a complex $L$-ary root of unity raised to power $y_{j}$. For instance, pick $L=4$ and consider $y'=(1,2,3,0,2,3,0,4)$. Then the corresponding vector $Z_{y'}=(w^{1},w^{2},w^{3},w^{0},w^{2},w^{3},w^{0},w^{4})$, where $w=e^{2\pi i/4}$ (for an arbitrary $L$ it will be $e^{2\pi i/L}$). Having defined $Q$ fixed $L$ we are now ready to construct each element of the “hard” distribution ${\cal D}_{Q,L}$: $$\text{Pr}_{{\cal D}_{Q,L}}[y]=\frac{|Q(Z_{y})|^{2}}{L^{N}n!}.$$ A quantum circuit which performs sampling is remarkably easy. It amounts to applying the quantum Fourier transform to a uniform superposition which was transformed by $h$ and measuring in the standard basis (see Theorem 4 of Section 4 of  [@fefferman2015power]). Classical sampling of distributions based on the above efficiently specifiable polynomial is believed to be hard in particular because it contains the permanent problem. Thus, the existence of an efficient classical sampler would imply a collapse of the Polynomial Hierarchy to the third level (see Section 5 and 6 of [@fefferman2015power] for detailed proof). Long-range quantum Hamiltonians {#app:hamil .unnumbered} ------------------------------- The long-range Hamiltonian we consider has the form: $$\left|\Psi(t)\right\rangle =e^{-i\mathcal{H}t}|\Psi(t=0)\rangle,$$ where $$\mathcal{H}=\sum_{i<j}V(i,j)\left(\sigma_{i}^{x}\sigma_{j}^{x}+\sigma_{i}^{y}\sigma_{j}^{y}\right),$$ and $V(i,j)=1/|i-j|^{3/4}$ is a long-range two-body interaction, and the initial state is a fully polarized state is the product state $|\Psi(t=0)\rangle=2^{-n/2}\sum_{i}\ket{i}$. At long propagation times $t\gg1$, the resulting states are highly entangled, and are for example, challenging for MPS-based tomography [@cramer2010efficient]. To assess the ability of VAE to compress highly entangled states, we focus on the task of reconstructing the outcomes of experimental measurements in the computational basis. In particular, we generate samples distributed according to the probability density $|\Psi_{i}(t)|^{2}$, and reconstruct this distribution with our generative, deep models. #### Acknowledgements. {#acknowledgements. .unnumbered} We thank Carlo Ciliberto, Danial Dervovic, Alessandro Davide Ialongo, Joshua Lockhart, and Gillian Marshall for helpful comments and discussions. Andrea Rocchetto is supported by an EPSRC DTP Scholarship and by QinetiQ. Edward Grant is supported by EPSRC [\[]{}EP/P510270/1[\]]{}. Giuseppe Carleo is supported by the European Research Council through the ERC Advanced Grant SIMCOFE, and by the Swiss National Science Foundation through NCCR QSIT. Sergii Strelchuk is supported by a Leverhulme Trust Early Career Fellowship. Simone Severini is supported by The Royal Society, EPSRC and the National Natural Science Foundation of China. #### Contributions. {#contributions. .unnumbered} The concept of using VAEs to encode probability distributions of quantum states was conceived by A.R., E.G., and G.C. The complexity framework was developed by A.R., G.C., and S.St. E.G. wrote the code and performed the simulations with help from S.St. The project was supervised by A.R. and S.Se. The first draft of the manuscript was prepared by A.R. and all authors contributed to the writing of the final version. A.R. and E.G. contributed equally to this work. #### Competing Interests. {#competing-interests. .unnumbered} The authors declare no competing financial interests. #### Data availability statements. {#data-availability-statements. .unnumbered} All data needed to evaluate the conclusions are available from the corresponding author upon reasonable request.
{ "pile_set_name": "ArXiv" }
--- author: - 'Tristan Buckmaster[^1]' - 'Steve Shkoller[^2]' - 'Vlad Vicol[^3]' title: Nonuniqueness of weak solutions to the SQG equation --- [**Abstract.**]{} We prove that weak solutions of the inviscid SQG equations are not unique, thereby answering Open Problem 11 in [@DLSz2012]. Moreover, we also show that weak solutions of the dissipative SQG equation are not unique, even if the fractional dissipation is stronger than the square root of the Laplacian. Introduction {#sec:intro} ============ The two-dimensional surface quasi-geostrophic (SQG) equation is a fundamental example of active scalar transport, and is classically written [@CoMaTa1994] as \[eq:SQG-old\] $$\begin{aligned} &\partial_t \theta + u \cdot \nabla \theta =0 \,, \label{eq:SQG-old-a} \\ &u = \RSZ^\perp \theta := \nabla^\perp \Lambda^{-1} \theta \,, \label{eq:SQG-old-b}\end{aligned}$$ a transport equation for the unknown scalar field $\theta = \theta(x,t)$, where $(x,t)\in \TT^2 \times \RR = [-\pi,\pi]^2 \times \RR$. In , $\Lambda = (-\Delta)^{1/2}$, $\RSZ = (\RSZ_1,\RSZ_2)$ is the vector of Riesz-transforms, $\nabla^\perp = (-\partial_2, \partial_1)$, and $x^\perp = (-x_2 ,x_1)$ for any vector $x = (x_1,x_2)$. We consider solutions of the SQG equation which have zero mean on $\TT^2$, i.e. $\int_{\TT^2} \theta(x,t) dx=0$, a quantity which is conserved in time even for weak solutions. In the context of geophysical fluid dynamics, the variable $\theta$ denotes the temperature (or surface buoyancy function) in a rapidly rotating stratified fluid with uniform potential vorticity [@HePiGaSw1995] and has applications in both meteorological and oceanic flows [@Pe1982]. Mathematically, two-dimensional SQG flows have the potential for finite-time singularity formation [@CoMaTa1994] and possess striking similarities to three-dimensional Euler solutions; in fact, the vector $ \nabla^\perp \theta$ is governed by the same evolution equation as the vorticity of the 3-D Euler flow: $$\begin{aligned} \partial_t (\nabla^\perp \theta) + u \cdot \nabla (\nabla^\perp \theta) = \nabla u \cdot \nabla^\perp \theta. \label{eq:SQG:3D:vorticity}\end{aligned}$$ As such, has been intensively analyzed over the past two decades. While the local existence of smooth solutions in Sobolev spaces $H^s$ with $s>2$, or Hölder spaces $C^{1,\alpha}$ with $\alpha>0$, has been established in [@CoMaTa1994], to date the question of whether a finite-time singularity may develop from smooth initial datum remains open, in analogy to the similar question for the 3-D incompressible Euler system. When is posed on $ \RR^2 \times \RR $ with datum having infinite kinetic energy, a gradient blowup may occur [@CaCo2010], but for datum on $\TT^2$ and with finite energy, it is only known that arbitrarily large growth of high Sobolev norms is possible from arbitrarily small initial datum [@KiNa2012]. The collapsing hyperbolic saddle blowup-scenario from [@CoMaTa1994] was ruled out analytically in [@Co1998; @CoFe2002], and the modern numerical simulations of [@CoLaShTsWu12] were able to resolve the equations past the initially predicted singular time [@CoMaTa1994]. A different blowup scenario via a cascade of filament instabilities of geometrically decreasing spatial and temporal scales was proposed in [@Sc2011]. The first example of a non-steady global in time smooth solution was obtained only very recently [@CaCoGo2016]. SQG conservation laws --------------------- Fundamental to our subsequent analysis is the fact that sufficiently smooth solutions of conserve the square of the $\dot{H}^{- {1/2} }( \mathbb{T} ^2 )$ norm of $\theta$. Upon taking the $L^2$ inner product of with $\Lambda^{-1} \theta$, integrating by parts in the nonlinear term, and using that $u\cdot \nabla \Lambda^{-1} \theta = \nabla^\perp \Lambda^{-1}\theta \cdot \nabla \Lambda^{-1} \theta = 0$, if follows that if $\theta$ is sufficiently smooth ($\theta \in L^{3}_{t,x}(\TT^2 \times \RR)$ is sufficient, cf. [@IsVi2015]), then $$\begin{aligned} \HH(t) := \| \theta(\cdot,t)\|_{\dot{H}^{-1/2}(\TT^2)}^2 = \|\theta_0\|_{\dot{H}^{-1/2}(\TT^2)}^2 \label{eq:Hamiltonian:conserved}\end{aligned}$$ for initial datum $\theta_0 \in \dot{H}^{- 1/2 }(\mathbb{T}^2)$. In fact, the $\dot{H}^{-1/2}$ norm of $\theta$ is the [*Hamiltonian*]{} $\HH$ associated to an action function (we systematically ignore the factor of $1/2$ that is usual present) from which the SQG equation may be derived via an Euler-Poincaré variational principle [@Re1995]. Additionally, due to the pure transport nature of , and the fact that the Lagrangian flow induced by the incompressible vector field $u$ preserves volume, sufficiently smooth solutions of the initial value problem for conserve the $L^p( \mathbb{T}^2)$ norms of $\theta$, so that $$\begin{aligned} \| \theta(\cdot,t)\|_{L^p(\TT^2)} = \|\theta_0\|_{L^p(\TT^2)} \label{eq:Casimir:conserved}\end{aligned}$$ where $\theta_0 \in L^p( \mathbb{T} ^2 )$ is the initial datum, and $1 \leq p \leq \infty$. When $p=2$, from elementary properties of the Riesz transform it follows that the [*kinetic energy*]{} is conserved for smooth solutions $\| u(\cdot,t)\|_{L^2(\TT^2)} = \|u_0\|_{L^2(\TT^2)} $.[^4] As noted in [@Ta2014], formally the conservation laws – are immediate consequences of Noether’s Theorem and the fact that the SQG equation belongs to a general class of active scalar equations satisfied by the vorticity of a generalized two-dimensional Euler equation on a Lie algebra with a specific inner product [@Re1995 Section 2.2] (see also [@Ta2016; @Wa2016; @Con2016] for a more recent account). We address this point of view in more detail in Section \[sec:geodesic\] and Appendix \[appendix:EP\] below, where we also present the momentum equation for the incompressible velocity field $v$ whose vorticity is the function $\theta$ in . While suggests that the problem of finite-time singularities for SQG is similar to that for 3-D Euler, the aforementioned variational point-of-view justifies a direct analogy between the conservation laws for SQG and those for the 2-D Euler equations: plays the role of the conservation of kinetic energy in 2-D Euler, while is analogous to the conservation of the Casimir functions in 2-D Euler. Therefore, we expect that a turbulent SQG solution exhibits a dual cascade of energy, as predicted by the Batchelor-Kraichnan theory for two dimensional Euler flows [@Con1998; @Con2002; @CoTaVi2014]. Motivated by two-dimensional turbulence, we are thus naturally lead to consider weak solutions of the SQG equation. Weak solutions of the SQG equation are not unique ------------------------------------------------- Motivated by with $p=2$ one may define $\theta \in L^2_{\rm loc}(\RR, L^2(\TT^2))$ to be a weak solution of if $$\begin{aligned} \intint_{\RR\times \TT^2} \left( \theta \partial_t \phi + \theta u \cdot \nabla \phi \right) \, dx\, dt = 0 \label{eq:weak:L2}\end{aligned}$$ holds for any smooth test function $\phi\in C^\infty(\TT^2 \times \RR)$, so that holds in the sense of distributions on $\TT^2 \times \RR$. Using this definition, it was established in [@Re1995] that for any $\theta_0$ in $L^2$, there exists a global-in-time weak solution to the Cauchy problem for , with $\theta \in L^\infty([0,\infty), L^2(\TT^2))$. See also  [@PuVa2015] for the global existence of weak solutions to the 3-D quasi-geostrophic system. We note the stark contrast here with 3-D Euler, for which the existence of weak solutions for any $L^2$ initial datum remains open. Moreover, we note that the proof of the existence of global weak solutions to SQG is quite different that the proof of global solutions to 2-D Euler (for which the vorticity is one derivative smoother than the velocity  [@MaBe2002]); in particular, the proof of [@Re1995] relies on a special structure of the nonlinear term in , which arises from the fact that the Fourier multiplier relating $\theta$ to $u$ is an odd function of the frequency. More precisely, $$\begin{aligned} \int_{\TT^2} \theta u\cdot \nabla \phi dx = \int_{\TT^2} \theta \RSZ^\perp \theta \cdot \nabla \phi dx = - \int_{\TT^2} \theta \RSZ^\perp \cdot (\theta \nabla \phi) dx = - \int_{\TT^2} \theta u\cdot \nabla \phi dx - \int_{\TT^2} \theta \left[ \RSZ^\perp \cdot, \nabla \phi\right] \theta dx \label{eq:SQG:magic}\end{aligned}$$ for any smooth test function $\phi$. Here and throughout the paper, we denote by $[A,B]$ the commutator of the operators $A$ and $B$. Since the commutator $\left[ \RSZ^\perp \cdot, \nabla \phi\right]$ is an operator of order $-1$, it maps $\dot{H}^{-1/2}$ into $\dot{H}^{1/2}$ (see Appendix \[sec:commutator\]), so that weak solutions to may be defined for distributions $\theta \in \dot{H}^{-1/2}$. We thus have the following \[def:weak:solution\] A distribution $\theta \in L^2_{\rm loc}(\RR;\dot{H}^{-1/2}(\TT^2))$ is a weak solution of if $$\begin{aligned} \int_{\RR} \langle \RSZ_i^\perp \theta, \partial_t \Lambda^{-1} \phi^i \rangle + \langle \RSZ_j^\perp \theta , \RSZ_i^\perp \Lambda^{-1} \theta \partial_j \phi^i \rangle - \frac 12 \langle \RSZ_i \RSZ_j^\perp \theta, [\Lambda , \phi^i] \RSZ_j^\perp \Lambda^{-1} \theta \rangle \, dt = 0\end{aligned}$$ for any $\phi \in C^\infty_0(\TT^2 \times \RR)$ such that $\div \phi = 0$, where $\langle \cdot , \cdot \rangle$ denotes the $\dot{H}^{-1/2}$-$\dot{H}^{1/2}$ duality pairing. See also Definition \[def:weak:sol:momentum\] for a more intuitive and equivalent formulation of Definition \[def:weak:solution\]. For $\theta \in L^2_{\rm loc}(\RR, L^2(\TT^2))$, Definition \[def:weak:solution\] agrees with that given via . The boundedness of the Calderón commutator $[\Lambda,\phi^i]$ on $\dot{H}^{1/2}$ is used implicitly in Definition \[def:weak:solution\] (cf. Appendix \[sec:commutator\]). Using the cancellation property , it was further shown in [@Ma2008] that for $\theta_0 \in L^p$, with $p>4/3$, there exists a global weak solution $\theta \in L^\infty(\RR,L^p(\TT^2))$ to the Cauchy problem for . The question of uniqueness of these weak solutions has remained, to date, open and has been isolated as a challenging open problem in [@DLSz2012 Problem 11] (see also [@Ma2008; @Ru2011; @AzBe2015; @Co2015]).[^5] One of our main results is that [*below a certain regularity threshold, weak solutions to the SQG equation are not unique*]{}, thereby answering the question posed in [@DLSz2012]. We state this result as the following \[thm:main:inviscid\] Suppose $\ee:\RR\rightarrow \RR^+$ is a smooth function with compact support. Then for every $1/2<\beta<4/5$ and $\sigma< \beta/(2-\beta)$, there exists a weak solution $\theta$, with $\Lambda^{-1}\theta \in C_t^{\sigma}C_x^{\beta}$, satisfying $$\begin{aligned} \HH(t) = \int_{\TT^2}{\left\vert\Lambda^{- {\frac}12}\theta(x,t)\right\vert}^2\,dx=\ee(t)\,.\end{aligned}$$ for $t \in \RR$. Indeed, due to the compact support in time of the function $\theta$ in Theorem \[thm:main:inviscid\], it follows that the trivial solution $\theta \equiv 0$ is not the only weak solution to which vanishes on the complement of a given time interval. Theorem \[thm:main:inviscid\] leaves open the question of whether the exponent $\beta$ can be taken arbitrarily close to $1$, which is the Onsager conjecture for the SQG equation (cf. Conjecture \[conj:Onsager\] below). The proof of Theorem \[thm:main:inviscid\] relies on a modification of the convex integration scheme employed by [@DLSz2012b; @DLSz2013; @BuDLeIsSz2015] to study the Onsager conjecture for the 3-D Euler equations.[^6] It was suggested in [@DLSz2012; @Sh2011; @IsVi2015] that the structure of the SQG nonlinearity is non-amenable to convex integration methods, because the multiplier relating $u$ to $\theta$ is an odd function of frequency. Herein, we overcome this difficulty by rephrasing the equation in terms of a potential velocity $v$ (whose vorticity is the scalar $\theta$, see Section \[sec:geodesic\]), which allows us to apply Fourier analysis techniques to construct nontrivial high-high-low frequency interactions, crucial to the method of convex integration. We discuss these details in Section \[sec:results\] below. Weak solutions of the dissipative SQG equation are not unique ------------------------------------------------------------- Note that while weak solutions of the SQG equation may be defined for $\theta \in L^2_{t} \dot{H}^{-1/2}_x$, the existence of weak solutions obtained in [@Re1995; @Ma2008] requires an initial datum which is more regular (e.g. $\theta_0 \in L^p_x$ for $p>4/3$). One may thus ask a natural question: is it possible for that in a given (low) regularity regime one can both construct weak solutions via compactness arguments (viscosity solutions), and also construct weak solutions via convex integration? In order to answer this question in the positive, we consider the fractionally dissipative SQG system \[eq:dissipative:SQG\] $$\begin{aligned} &\partial_t \theta + u \cdot \nabla \theta + \Lambda ^ \gamma \theta = 0 ,\\ &u = \RSZ^\perp \theta := \nabla^\perp \Lambda^{-1} \theta, \end{aligned}$$ with $\gamma \in (0,2]$. Strong solutions of have been considered extensively. The dissipative SQG equation has a natural scaling symmetry: if $\theta(x,t)$ is a $\TT^2$-periodic solution to the Cauchy problem for with datum $\theta_0(x)$, then $\theta_\lambda(x,t) = \lambda^{\gamma-1} \theta( \lambda x,\lambda^\gamma t)$ is a $\TT_\lambda^2= [-\pi/\lambda,\pi/\lambda]^2$-periodic solution of with initial datum $\theta_{0,\lambda}(x) = \lambda^{\gamma-1} \theta(\lambda x)$. In view of this scaling symmetry, the $L^\infty$ norm is scale invariant for $\gamma=1$.[^7] [^8] Therefore, for $\gamma>1$ becomes semilinear and subcritical, and the global well-posedness of smooth solutions was established in [@CoWu1999]. The critical case $ \gamma =1$ is a quasilinear problem, and the global regularity of smooth solutions for large initial datum was established in [@KiNaVo2007; @CaVa2010], with different proofs given in [@KiNa2010; @CoVi2012; @CoTaVi2015]. For $\gamma<1$, the supercritical case, the question of finite-time singularities from smooth initial datum remains completely open, in analogy to the inviscid equations .[^9] It is known, however, that the global-in-time weak solutions eventually become smooth [@Si2010; @Da2011; @Ki2011; @CZVi2016], which leads us to consider weak solutions of . In analogy to Definition \[def:weak:solution\], we may define a weak solution to as follows. \[def:dissipative:weak:solution\] The distribution $\theta \in L^2_{\rm loc}(\RR;\dot{H}^{-1/2}(\TT^2))$ is a weak solution of if $$\begin{aligned} \int_{\RR} \langle \RSZ_i^\perp \theta, \partial_t \Lambda^{-1} \phi^i \rangle + \langle \RSZ_j^\perp \theta , \RSZ_i^\perp \Lambda^{-1} \theta \partial_j \phi^i \rangle - \frac 12 \langle \RSZ_i \RSZ_j^\perp \theta, [\Lambda , \phi^i] \RSZ_j^\perp \Lambda^{-1} \theta \rangle - \langle \RSZ_i^\perp \theta, \Lambda^{\gamma-1} \phi^i \rangle \, dt = 0\end{aligned}$$ for any $\phi \in C^\infty_0(\TT^2 \times \RR)$ such that $\div \phi = 0$, where $\langle \cdot , \cdot \rangle$ denotes the $\dot{H}^{-1/2}$-$\dot{H}^{1/2}$ duality pairing. Note that Definition \[def:dissipative:weak:solution\] does not require that solutions verify the local energy inequality nor that they possess the additional regularity $\theta \in L^2_{\rm loc}(\RR;\dot{H}^{(\gamma-1)/2}(\TT^2))$; the weak solutions we consider need not be suitable weak solutions (see [@CaVa2010]). In contrast to the inviscid SQG equation, for the dissipative SQG equation it was shown in [@Ma2008] that for any $\gamma>0$ and any $\theta_0 \in \dot{H}^{-1/2}$ there exists a global-in-time weak solution $\theta$ to .[^10] See also [@BaGr2015] for the global existence of weak solutions when $\theta_0 \in L^{1+}$. Using the convex integration scheme developed to prove Theorem \[thm:main:inviscid\], we establish the nonuniqueness of weak solutions to the $\gamma$-dissipative SQG equation , even for a range of values for $ \gamma$ above $1$, the subcritical regime. \[thm:main:dissipative\] Suppose $\ee: \RR \rightarrow \RR^+$ is a smooth function with compact support. Then for every $1/2<\beta<4/5$, $0< \gamma < 2-\beta$ and $\sigma<\beta/(2-\beta)$, there exists a weak solution $\theta$, with $\Lambda^{-1} \theta \in C_t^{\sigma}C_x^{\beta}$, satisfying $$\begin{aligned} \int_{\TT^2}{\left\vert\Lambda^{- {\frac}12}\theta(x,t)\right\vert}^2\,dx=\ee(t)\,\end{aligned}$$ for all $t \in \RR$.[^11] Therefore, for the dissipative SQG equation, convex integration can coexist with weak compactness. This flexibility of the PDE  is both due to the low regularity of the weak solution, and that enforcement of the local energy inequality is not required. To the best of our knowledge, this is the first instance when the convex integration scheme can be employed for an evolution equation arising in fluid dynamics, which is parabolic, and even semi-linear. The main ideas used in the proof of Theorem \[thm:main:dissipative\] are discussed in Section \[sec:results\]. Note that as $\beta \to 1^{-}$, Theorem \[thm:main:dissipative\] holds with $\gamma \to 1^{+}$. This motivates Conjecture \[conj:critical:SQG\] below, which states that for the critical SQG equation ($\gamma=1$) we have a dichotomy of regularity exponents, whereby for $\beta\geq 1$ the energy equality holds, the uniqueness and global regularity of solutions holds; while for $\beta<1$ the uniqueness of weak solutions breaks down and the equation becomes flexible, i.e. amenable to convex integration constructions. Hydrodynamical systems as geodesic equations {#sec:geodesic} -------------------------------------------- At least since the work of Poincaré [@Po1901], it has been well-known that the equations of motion of the finite-dimensional mechanical systems governed by Newtonian mechanics can be interpreted as the geodesic equations of a Riemannian metric on configuration space or Lie group (see, for example, [@Br1946]). The motion of a rigid body, for example, is governed by a left-invariant Riemannian metric on the Lie group $SO(3)$ [@MaRa1999; @KhAr1999]. In his seminal paper, [@Ar1966] showed that infinite-dimensional hydrodynamical systems could also be represented by geodesic equations on the infinite-dimensional group of volume preserving diffeomorphisms $ \mathcal{D} _\mu$. Specifically, Arnold proved that the incompressible Euler equations are geodesics with the respect to the $L^2$-right invariant metric on the Lie group $ \mathcal{D} _\mu$. The Euler-Poincaré variational principle [@MaRa1999; @KhAr1999] asserts that such hydrodynamical geodesic equations can be computed for a rather general metric specified on the associated Lie algebra $ \mathcal{V} $, the space of divergence-free vector fields, with Lie bracket given by $[u,w] = \partial_j u w^j - \partial _j w u^j$. For a positive-definite, self-adjoint operator $A$, we define the metric on $ \mathcal{V} $ by $$\label{metric} (u, w) = \int_{ \mathbb{T} ^2 } Au \cdot w \, dx \,.$$ The metric is then right-translated over the Lie group $ \mathcal{D} _\mu$. As we shall explain Appendix \[appendix:EP\], the geodesic equations associated to this metric are extrema of the action function $$\begin{aligned} \label{eq:action:*} s(u) = \int_{\RR} \int_{\TT^2} A u \cdot u \, dx dt\end{aligned}$$ for incompressible Lie-advected variations $\delta u$ that obey suitable boundary conditions, which gives rise to the following hydrodynamical system: \[eq:hydro\] $$\begin{aligned} \partial_t v + u \cdot \nabla v + ( \nabla u)^T\cdot v & = - \nabla {\widetilde}p, \\ \div u & = 0, \\ v & = A u \,.\end{aligned}$$ Here ${\widetilde}p$ denotes the pressure function, a Lagrange multiplier enforcing the incompressibility of $u$. When the operator $A$ is the identity matrix, then is the incompressible Euler equations (with pressure function $p = {\widetilde}p + {\frac{1}{2}} |u|^2$); however, the operator $A$ can be differential operator, a nonlocal Fourier multiplier, or even a more general operator satisfying the positivity and symmetry conditions noted above.[^12] [^13] As observed in [@Re1995] (see also [@Ta2016; @Wa2016; @Con2016]), if $A= \Lambda ^{-1} $, then is the $\dot{H}^{-1/2}$ metric on $ \mathcal{V} $, and it follows from that $\omega = \nabla^\perp \cdot u$ obeys $$\partial_t \Lambda ^{-1} \omega + u \cdot \Lambda^{-1} \omega =0 \,.$$ A simple computation (see Section \[sec:SQG:momentum\] below) shows that $\theta = - \Lambda^{-1} \omega$ solves the SQG equation . The Onsager conjecture and the uniqueness of weak solutions ----------------------------------------------------------- As discussed in Section \[sec:geodesic\], solutions to the SQG equation are extrema of the action function $s(u)$ in , with $A = \Lambda^{-1}$. Since $s(u)$ does not explicitly depend on $t$, for any such system the corresponding Hamiltonian $$\begin{aligned} \HH(t) = \int_{\TT^2} \Lambda^{-1} u \cdot u \, dx = \int_{\TT^2} \Lambda^{-1} \theta \, \theta \, dx \label{eq:Hamiltonian:2}\end{aligned}$$ is formally conserved (see ). Inspired by the Onsager conjecture for the incompressible Euler equations (see the discussion in Section \[sec:results\] below), a fundamental question arises for the SQG equations: do weak solutions of conserve the Hamiltonian $\HH(t)$? One may conjecture the following dichotomy: \[conj:Onsager\] Let $v = \Lambda^{-1} u = \Lambda^{-1} \RSZ^\perp \theta$. Define $\alpha_O = 1$.[^14] - If $v \in C(\RR;C^\alpha (\TT^2))$ is a weak solution of the SQG equation, with $\alpha > \alpha_O$, then holds on $[0,T]$. - For any $1/2<\alpha <\alpha_O$, there exist infinitely many weak solutions of the SQG equation, with $v \in C ( \RR ; C^\alpha(\TT^2))$, such that fails. The rigid side (a) of this conjecture was resolved in [@IsVi2015], following the classical work of [@CoETi1994], by proving that $\theta \in L^{3}(\RR ; L^3(\TT^2))$ implies the conservation of $\HH$. In this paper we address the flexible side (b) of the conjecture, and prove in Theorem \[thm:main:inviscid\], that (b) holds for $\alpha < 4/5$. The regularity gap, $\alpha \in [4/5,1)$, for nonconservative weak solutions of SQG remains for the same reason that the Onsager conjecture for the 2-D Euler equations remains open, with an open range of values for $ \alpha $ in $[1/5,1/3)$.[^15] This issue is discussed further in Section \[sec:results\] below. Proving part (b) of Conjecture \[conj:Onsager\] for any $\alpha \in [4/5,1)$ appears to be challenging. \[rem:Onsager\] The Euler and SQG equations are particular cases of equations obeyed by extrema of the action functional $s(u)$ in . For Euler $A = {\ensuremath{\mathrm{Id}}}$, while for SQG, $A=\Lambda^{-1}$. In general, let $A$ be any positive-definite self-adjoint operator which is translation invariant and acts on scalar periodic functions with zero mean, such that ${\left\lVertA w\right\rVert}_{L^2(\TT^2)} \approx {\left\lVertw\right\rVert}_{\dot{H}^a(\TT^2)}$ for some $a \in \RR$. Let $u$ be an extremum of the corresponding action functional, such that $v = Au$ obeys the hydrodynamical equation . In analogy with Conjecture \[conj:Onsager\], it is natural to determine the Onsager exponent $\alpha_O$, such that solutions with regularity above $\alpha_O$ conserve the Hamiltonian $\HH = \int_{\TT^2} u \cdot v dx$, while solutions with regularity below $\alpha_O$ do not. Upon rewriting the nonlinear term in as $u \cdot \nabla v - (\nabla v)^T \cdot u = u^\perp (\nabla^\perp \cdot A u)$, taking an inner product with a mollified version of $u$ and integrating over $\TT^2$, an argument similar to [@CoETi1994] shows that for $v = A u \in L^3_t C^{\alpha}_x$, with $\alpha > \alpha_O =: -a + (1+a)/3$, the Hamiltonian is conserved. If indeed this choice of $\alpha_O$ determines an Onsager dichotomy remains to be shown. Naturally, the value of the Onsager exponent $\alpha_O$ discussed in Conjecture \[conj:Onsager\] and Remark \[rem:Onsager\] depends on the precise Banach scale $X^\alpha$ considered. Above, we have only mentioned the scale of Hölder spaces $X^\alpha = C_t C^\alpha_x$. On the other hand the Hamiltonian $\HH$ is quadratic in $u$, and the nonlinear term in is also quadratic in $u$, so that proving the conservation in time of $\HH$ for the Euler [@CoETi1994; @ChCoFrSh2008] and SQG equations [@IsVi2015] only requires control of the solution in the Banach scale $X^\alpha = L^3_t B^{\alpha}_{3,\infty}$, with $\alpha > \alpha_O$. Thus, Conjecture \[conj:Onsager\] may be alternatively posed on this $L^3$-based Banach scale, without changing the value of $\alpha_O$. It is however conceivable that for an Onsager regularity threshold $\alpha_O$ defined in terms of an $L^2$-based Banach scale, such as $X^\alpha = L^2_t \dot{H}^\alpha_x$, the sharp value may be different from the one discussed in Remark \[rem:Onsager\], which is computed in terms of $L^\infty_{t,x}$ or $L^3_{t,x}$. \[rem:Klainerman\] In a recent survey article on the work of J. Nash [@Kl2016], other [*threshold exponents*]{} are discussed for which a dichotomy in the behavior of solutions holds, depending on whether the regularity index of the weak solution is greater than or less than this exponent. For simplicity, fix the Banach scale $X^\alpha = C_t C^\alpha_x$. In analogy to the [*Onsager exponent*]{} $\alpha_O$, we define the following important regularity exponents: the [*Nash exponent*]{} $\alpha_N$ determines whether the nonlinear evolution is flexible or rigid (in the sense that $h$-principles are available); the [*uniqueness exponent*]{} $\alpha_U$ determines the uniqueness of solutions; the [*well-posedness exponent*]{} $\alpha_{WP}$ determines the local well-posedness of the system; and the [*scaling exponent*]{} $\alpha_*$ which determines the space $X^{\alpha_*}$ whose norm is invariant under the natural scaling symmetries of the equation (see Page 11 in [@Kl2016]). For instance, in the case of the Euler equations with the Hölder scale $ C_t C^\alpha_x$ for the velocity field $u$, we have that $\alpha_{WP} = 1$ (cf. [@Ho1933; @BaTi2010; @ElMa2014; @BoLi2015]), $\alpha_U$ is also conjectured to be equal to $1$ (only $\alpha_U \leq 1$ is known), $\alpha_O = 1/3$ (cf. [@CoETi1994; @ChCoFrSh2008; @Is2016]), and $\alpha_* = 0$, and $\alpha_O \leq \alpha_N \leq \alpha_U$ (since the convex integration constructions also prove $h$-principles and nonuniqueness). We note that these exponents are not the same, and one expects them to be linearly ordered $\alpha_* \leq \alpha_O \leq \alpha_N \leq \alpha_U \leq \alpha_{WP}$ (cf. [@Kl2016 Equation (0.7)]). For the inviscid SQG equation, on the Hölder scale $ C_t C^\alpha_x$ for the [*potential velocity field*]{} $v = \Lambda^{-1} u$, this gap between the exponents remains, and one may conjecture that $\alpha_O = 1$, while $\alpha_{WP} = 2$. However, in view of Theorem \[thm:main:dissipative\], for the dissipative SQG equation with $ \gamma =1$ (informally called the [*critical SQG equation*]{}), one may conjecture that all the exponents discussed in Remark \[rem:Klainerman\] are the same, thereby justifying the adjective “critical”. \[conj:critical:SQG\] Consider the dissipative SQG equation  with $\gamma=1$, and fix the Banach scale $X^\alpha = C_t C^\alpha_x$ as a way to measure the regularity of the potential velocity $v = \Lambda^{-1} u = \Lambda^{-1} \RSZ^\perp \theta$. Then $1 = \alpha_* = \alpha_O = \alpha_U = \alpha_{WP}$. That $\alpha_*=1$ follows from the fact that the $L^\infty$ norm is scaling invariant. The fact that $\alpha_{WP}, \alpha_{U} \leq 1$ follows e.g. from [@CaVa2010], while Theorem \[thm:main:dissipative\] shows that $\alpha_O, \alpha_U \geq 4/5$. Establishing the remaining inequalities in Conjecture \[conj:critical:SQG\] remains open. The above conjecture provides the first nonlinear hydrodynamical PDE for which the exponents of Remark \[rem:Klainerman\] are all the same. Outline of the proof ==================== The SQG momentum equation {#sec:SQG:momentum} ------------------------- We shall make use of two velocity fields to describe the SQG equations: we define the [*potential velocity*]{} $$\begin{aligned} \label{eq:pv} v = \Lambda^{-1} u \,,\end{aligned}$$ which is thus one derivative smoother than the SQG [*transport velocity*]{} $u = \RSZ^\perp \theta$. From , it follows that the potential velocity $v$ satisfies \[eq:SQG\] $$\begin{aligned} \partial_t v + u \cdot \nabla v - ( \nabla v)^T\cdot u & = - \nabla p, \label{eq:SQG-a} \\ \div v & = 0, \label{eq:SQG-b} \\ u & = \Lambda v \label{eq:SQG-c} \,,\end{aligned}$$ where $p = {\widetilde}p + u \cdot v$. The SQG momentum equation can be equivalently written as[^16] $$\p_t v + u^\perp \, ( \nabla ^\perp \cdot v) = -\nabla p \,.$$ Upon defining the temperature function $\theta$ as minus the vorticity of the potential velocity: $$\theta = - \nabla ^\perp \cdot v \,,$$ becomes $$\begin{aligned} \label{eq:SQG-p-theta} \partial_t v - \theta u^\perp = - \nabla p.\end{aligned}$$ A direct computation confirms that $\theta$ is indeed a solution of (\[eq:SQG-old\]); taking the scalar product of $\nabla ^\perp$ with , we find that $\partial_t (-\theta) - \nabla^\perp \cdot (\theta u^\perp) = 0$ and hence $\partial_t \theta + u \cdot \nabla \theta = 0$, since $\nabla^\perp \cdot u^\perp = - \nabla \cdot u = 0$. Note that the dissipative SQG equation also can be written as in , by adding $\Lambda^\gamma v$ to the right side of . As we are primarily interested in weak solutions of SQG, we shall need some basic commutator identities. For all test functions $\phi \in C^ \infty ( \mathbb{T} ^2)$, we have that $$\begin{aligned} - \int_{\TT^2} \partial_i v^j \Lambda v^j \partial_i \phi dx & = \int_{\TT} \left( v^j \Lambda \partial_i v^j \partial_i \phi + v \cdot \Lambda v \Delta \phi\right)\, dx \,,\end{aligned}$$ and thus, $$\begin{aligned} - \int_{\TT^2} \partial_i v^j \Lambda v^j \partial_i \phi dx & = \frac 12 \int_{\TT^2} \partial_i v^j \left[ \Lambda, \partial_i \phi \right] v^j\, dx \, + \frac 12 \int_{\TT^2} v \cdot \Lambda v \Delta \phi dx.\end{aligned}$$ This motivates a convenient and equivalent definition of a weak solution, which is clearly equivalent to Definition \[def:weak:solution\] above. \[def:weak:sol:momentum\] We say that $v \in L^2_{\rm loc}(\RR;H^{1/2}(\TT^2))$ is a weak solution of if $$\begin{aligned} \int_{\RR} \langle v^i , \partial_t \phi^i \rangle + \langle \Lambda v^j , v^i \partial_j \phi^i \rangle - \frac 12 \langle \partial_i v^j , [\Lambda , \phi^i] v^j \rangle \, dt = 0\end{aligned}$$ holds for any $\phi \in C_0^\infty(\RR\times \TT^2)$ such that $\div \phi = 0$. Here $\langle \cdot , \cdot \rangle$ denotes the $\dot{H}^{-1/2}$-$\dot{H}^{1/2}$ duality pairing. Note that for smooth $\phi$ the operator $[\Lambda , \phi^i]$ is a zeroth order operator on $v$ (cf. Lemma \[lem:Calderon\] below). Moreover, adding the term $\int_{\RR} \langle v^i, \Lambda^\gamma \phi^i\rangle \, dt$ to Definition \[def:weak:sol:momentum\] gives an equivalent form of Definition \[def:dissipative:weak:solution\]. By taking the divergence of and using , we obtain that $p$ solves $$\label{eq:pressure} - \Delta p = \text{Tr}\left( \nabla v \ \nabla u - \nabla v ^T \ \nabla u \right) - \Delta v \cdot u \,.$$ For weak solutions, we must interpret $p$ as a distribution, and the elliptic equation has the following distributional formulation: for all test functions $\phi \in C^ \infty ( \mathbb{T} ^2)$, $$\label{eq:pressure-weak0} \int_{\TT^2} p \ \Delta \phi \, dx = \int_{ \TT^2} \left( v^i \Lambda v^j \partial_{ij}^2 \phi - \partial_i v^j \Lambda v^j \partial_i \phi \right)\, dx \,.$$ It follows that $$\label{eq:pressure-weak} \int_{\TT^2} p \ \Delta \phi \, dx = \int_{ \TT^2} \left( v^i \Lambda v^j \partial_{ij}^2 \phi + \frac 12 v \cdot \Lambda v \Delta \phi + \partial_i v^j \left[\Lambda \, , \, \partial_i \phi \right] v^j \right)\, dx \,.$$ Therefore, given $v \in L^2(\RR;\dot{H}^{1/2}(\TT^2))$ formula defines $p$ (and therefore also $\nabla p$) as a distribution on $\TT^2$ via $$\begin{aligned} \langle p , \Delta \phi \rangle &= \langle \Lambda v^j , v^i \partial_{ij}^2 \phi \rangle + \frac 12 \langle \Lambda v^i , v^i \Delta \phi \rangle + \langle \partial_i v^j , \left[\Lambda \, , \, \partial_i \phi \right] v^j \rangle.\end{aligned}$$ In particular, for $v \in C^{1/2-}_t C^{4/5-}_{x}$, the $\nabla p$ term in is a well-defined distribution. The main result and the main ideas of the proof {#sec:results} ----------------------------------------------- Employing the potential velocity formulation of SQG, we will prove the following theorem which is easily seen to imply Theorem \[thm:main:inviscid\] and Theorem \[thm:main:dissipative\] (we use the convention that $\gamma=0$ is the inviscid SQG equation): \[thm:main\] Suppose $\ee:[0,T]\rightarrow \RR^+$ is a smooth function with compact support. Then for every $1/2<\beta<4/5$, $0\leq \gamma < 2-\beta$ and $\sigma<\beta/(2-\beta)$, there exists a weak solution $v \in C_t^{\sigma}C_x^{\beta}$ satisfying $$\begin{aligned} \int_{\TT^2}{\left\vert\Lambda^{ {\frac}12}v(x,t)\right\vert}^2\,dx= \int_{\TT^2}\Lambda v(x,t) \cdot v(x,t) \,dx =\ee(t)\,\end{aligned}$$ for all $t \in [0,T]$. The proof will employ a convex integration scheme, similar in style to that presented in [@BuDLeIsSz2015] (cf.  [@DLSz2012b; @DLSz2013]). In [@BuDLeIsSz2015], highly oscillatory Beltrami waves formed the principle building block in the construction. It was noted in [@ChDeSz2012; @Ch2013] that Beltrami waves can be replaced by Beltrami plane waves (see Section \[sec:Beltrami\_plane\_waves\]) in order to prove analogous results for the 2-D Euler equations. Such Beltrami planes waves form a large class of stationary solutions to both the 2-D Euler and the inviscid SQG equation; as such, they will form the principle building block in the construction presented here. As a side remark, we note that it is not difficult to see from the analysis in the present paper that the results in [@BuDLeIsSz2015; @Is2013; @Bu2015; @BuDLeSz2016] for the 3-D Euler equations can be extended to the 2-D setting by replacing Beltrami waves by Beltrami plane waves. The fundamental aim in any convex integration scheme is to introduce high frequency oscillations that self-interact due to the nonlinearity in order to produce low frequency modes that cancel error terms. For SQG this so called high-high-low interaction is highly nontrivial. Indeed, as was already noted in [@IsVi2015], if one works with the usual formulation of the SQG equations in terms of the active scalar $\theta$, and considers a perturbation $\Theta = \sum_k \Theta_k$, with $\Theta_k(x,t) = a_k(t,x) e^{i \lambda k \cdot x}$ and an amplitude $a_k$ which lives at a frequency much smaller than $\lambda$, with $\lambda \gg 1$, then to leading order the corresponding velocity field is given by $U = \sum_k U_k$, where $U_k (x,t)= \RSZ^\perp \Theta_k(x,t) = i k^\perp \Theta_k(x,t) + o(\lambda^{-1})$. This implies that the high-high interactions in the nonlinear term $\div(\Theta U) = {\frac}12 \div(\sum_k \Theta_k U_{-k} + \Theta_{-k} U_k)$ vanish to leading order, since $\Theta_k U_{-k} + \Theta_{-k} U_k = a_k^2 ( i (-k)^\perp + i k^\perp) + o(\lambda^{-1}) = o(\lambda^{-1})$.[^17] To overcome this difficulty we work at the level of the momentum equation for $v$ and employ a bilinear pseudo-product operator (see [@CoMe1978] or Appendix \[app:pseudo:product\] below) to rewrite the nonlinearity $u \cdot \nabla v - ( \nabla v)^T\cdot u$ as the sum of a divergence of a $2$-tensor, and a gradient of a scalar function. Expanding in frequency around our Beltrami plane waves, we will show that the principal term in high-high-low interactions is of the correct form to cancel low frequency modes. This is achieved in Section \[sec:osc\] below. We remark that recently, [@Is2016] proved the full Onsager’s conjecture for the 3-D Euler equations, employing a novel technique involving gluing exact solutions to the Euler equations, along with the use of Mikado flows, introduced in [@DaSz2016] as a replacement to Beltrami waves. Mikado flows have the advantage of satisfying better oscillation-error estimates (see Section \[sec:inductive\_step\] for the definition of the oscillation error in the case of SQG), since they have disjoint spatial support in a thin cylinder. Unfortunately, the construction is inherently three-dimensional as it requires that the Mikado flows do not intersect, which is impossible in 2-D. Finding a suitable replacement for Mikado flows for the case of the SQG equations or the 2-D Euler equations is an interesting open problem. Notation -------- Throughout this paper, we make use of the Einstein summation convention, in which repeated indices are summed from $1$ to $2$. For $s \in \RR$, the [homogeneous]{} Sobolev space norm is $\|u\|^2_{\dot H^{s}( \mathbb{T}^2 )} = \sum_{k \in \mathbb{Z}^2 \setminus \{0\}} |{\widehat}u(k)|^2 |k|^{2s} \,.$ Here it is important that we work with functions of zero mean on $\TT^2$. The fractional Laplacian $\Lambda^s$ may be defined in this context as the Fourier multiplier with symbol $|k|^s$, for all $s\in \RR$. For a function $f: \mathbb{T}^2\times \RR \to \mathbb{R} $, we use the notation $$\|f\|_{C^\beta} \quad \mbox{to denote the space-time norm} \quad \|f\|_{C^0(\RR;C^\beta ( \mathbb{T} ^2 ))}$$ where the spatial Hölder norm is defined as the sum of the $C^0$ norm and the Hölder seminorm $[\cdot]_{C^\beta}$. To distinguish functions with higher regularity in time, we use the norm ${\left\lVertf\right\rVert}_{C_t^{\sigma}C_x^{\beta}} = {\left\lVertf\right\rVert}_{C^\sigma(\RR;C^\beta(\TT^2))}$. That is, $t\mapsto f(x,t)$ has $\sigma$-Hölder regularity in time and $x \mapsto f(x,t)$ has $\beta$-Hölder regularity in space. For $f\colon \TT^2\to \RR$ which is just a function of the space variable $x$, we denote by abuse of notation ${\left\lVertf\right\rVert}_{C^0}$ its $C^0(\TT^2)$ norm. Throughout the manuscript we abuse notation and denote by $f$ the periodic extension to all of $\RR^2$ of a $\TT^2$-periodic function $f$. Consequently, throughout the proof we work with $\RR^2$ convolution kernels and $\RR^2$ Fourier multiplier operators, instead of working with their $\TT^2$, respectively $\ZZ^2$ counterparts, which are obtained via the Poisson summation formula from their $\RR^2$ analogues. See e.g. [@CaZy1954 pp. 256–261], or [@StWe1971 Chapter VII] for the main ideas behind this transference principle. In particular, since we work with functions on $\TT^2$ which have zero mean, the resulting $\RR^2$ functions have support in frequency in the complement of a small neighborhood of the origin. We will use $a \lesssim b$ to denote $a \leq C b$ for a universal constant $C\geq 1$. Moreover, for an integer $N \geq 1$ we will use $D^N$ to denote any spatial derivative $\partial_x^\alpha$, where $|\alpha|=N$. Convex integration scheme {#sec:convex-integration} ========================= We use a convex integration scheme inspired by [@BuDLeIsSz2015]. We shall construct a sequence of solutions $(v_q,p_q,\mathring R_q)$ to the [*relaxed SQG momentum equation*]{} $$\begin{aligned} \partial_t v_q + u_q \cdot \nabla v_q - (\nabla v_q)^T \cdot u_q + \nabla p_q + \Lambda^\gamma v_q &= \div \mathring R_q \\ \div v_q &=0 \\ u_q &= \Lambda v_q\end{aligned}$$ \[eq:relaxed-SQG\] where $\mathring R_q$ is a symmetric trace-free $2 \times 2$ matrix. The goal is to obtain $\mathring R_q \to 0$ as $q\to \infty$ (in a suitable topology), and show that a limiting function $v_q \to v$ exists, and solves . Parameters ---------- We fix $\beta >1/2$ to be the Hölder exponent that we expect to obtain for our weak solution $v$, and write it as $$\begin{aligned} \beta = \frac{4}{5} - \eps\end{aligned}$$ for some $0< \eps \ll 1$. For this $\eps>0$ fixed, we also define $$\begin{aligned} 0 \leq \gamma < 2 - \beta \end{aligned}$$ to be the power of the dissipation in the equation. When $\gamma=0$ it is understood that the equation is inviscid, i.e. that the dissipative term $\Lambda^\gamma$ is absent from the equations. Define the frequency parameter $$\begin{aligned} \lambda_q = \lambda_0^q \end{aligned}$$ for some integer $\lambda_0 \gg 1$ that is sufficiently large integer which is a multiple of $5$. Note thus that the spatial frequency, i.e. wavenumber, parameter $\lambda_q$ is strictly increasing in $q$ and grows exponentially. We also define the amplitude parameter $$\begin{aligned} \delta_q = \lambda_0^2\lambda_q^{-2\beta} \, . \label{eq:delta-q-def}\end{aligned}$$ Inductive assumption {#s:inductive_assump} -------------------- We shall inductively assume that the potential velocity $v_q$ has compact support in frequency, contained inside the ball of radius $2\lambda_q$ and has size $$\begin{aligned} \|v_q\|_{C^1}+\|u_q\|_{C^0} \leq C_0 \delta_q^{1/2}\lambda_q. \label{eq:ind:q:1}\end{aligned}$$ where $C_0\geq 1$ is a universal constant, independent of any of the other parameters in the construction. Similarly, we shall inductively assume that $ \mathring R_q$ has compact support in frequency, inside the ball $\{ \xi \colon |\xi| \leq 4\lambda_q\}$, and has amplitude given by $$\begin{aligned} \| \mathring R_q\|_{C^0} \leq \eps_R \lambda_{q+1} \delta_{q+1} \label{eq:ind:q:2}\end{aligned}$$ holds, where $\ee(t)$ is the prescribed energy profile and $\eps_R$ is a small constant to be chosen precisely in the construction. We also make the inductive assumption that material derivatives for $w_q$ and $\mathring R_q$ are bounded as $$\begin{aligned} {\left\lVert (\partial_t + u_q \cdot \nabla) v_q\right\rVert}_{C^0} &\leq C_0 \lambda_q^2 \delta_q \label{eq:ind:q:3} \\ {\left\lVert (\partial_t + u_q \cdot \nabla) u_q\right\rVert}_{C^0} &\leq C_0 \lambda_q^3 \delta_q \label{eq:ind:q:3:b} \\ {\left\lVert (\partial_t + u_q \cdot \nabla) \mathring R_q\right\rVert}_{C^0} &\leq \lambda_q^2 \delta_q^{1/2} \lambda_{q+1} \delta_{q+1}. \label{eq:ind:q:4}\end{aligned}$$ Here $C_0$ is the same as in . Additionally, we assume that for the given prescribed energy profile $$\label{eq:energy_ind} 0 \leq \ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_q\right\vert}^2~dx\leq \lambda_{q+1}\delta_{q+1}$$ and $$\label{eq:zero_reynolds} \ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_q\right\vert}^2~dx\leq \frac{\lambda_{q+1}\delta_{q+1}}{8}~\Rightarrow \mathring R_q(\cdot,t)\equiv 0\,.$$ Inductive step {#sec:inductive_step} -------------- The convex integration scheme consists of correcting the potential velocity $v_q$ with an increment $w_{q+1}$ and an associated transport velocity increment $\Lambda w_{q+1}$ and obtain new velocity fields $$\begin{aligned} v_{q+1} = w_{q+1} + v_{q} \qquad \mbox{and} \qquad u_{q+1} = \Lambda w_{q+1} + u_q \label{eq:velocity:increment}\end{aligned}$$ such that the following holds: \[prop:main\] Let $\ee:[0,T]\rightarrow \RR^+$ be a given smooth Hamiltonian profile. Then, for sufficiently large $\lambda_0\in5 \NN$, if the pair $(v_q,\mathring R_q)$ satisfy assumptions – specified above then there exists a new pair $(v_{q+1}, \mathring R_{q+1})$ that satisfy these assumptions with $q$ replaced by $q+1$. Moreover, the difference $w_{q+1}=v_{q+1}-v_q$ has frequency support contained in the annulus $\{ \xi \colon \lambda_q/2 \leq |\xi| \leq 2\lambda_q\}$ and has size $$\label{eq:shell_est} {\left\lVertw_{q+1}\right\rVert}_{C^0} \leq C_0 \delta_{q+1}^{{\frac}12}$$ for a fixed universal constant $C_0 \geq 0$. We note here that if $(v_q,\mathring{R}_q)$ solves at step $q$, and the new velocity $v_{q+1}$ is given by , then upon implicitly[^18] defining $\mathring R_{q+1}$ by $$\begin{aligned} \div \mathring R_{q+1} &= \Big( \partial_t w_{q+1} + u_q \cdot \nabla w_{q+1} \Big) \notag\\ &\qquad + \Big( \Lambda w_{q+1} \cdot \nabla v_q - (\nabla v_q)^T \cdot \Lambda w_{q+1} - (\nabla u_q)^T \cdot w_{q+1} \Big) \notag\\ &\qquad + \Lambda^\gamma w_{q+1} \notag \\ &\qquad + \left( \div \mathring R_q + \Lambda w_{q+1} \cdot \nabla w_{q+1} - (\nabla w_{q+1})^T \cdot \Lambda w_{q+1} \right) \notag\\ &\qquad + \nabla {\widetilde}p_{q+1} \notag \\ &=: \div R_{T} + \div R_{N} + \div R_{D} + \div R_{O} + \nabla {\widetilde}p_{q+1} \label{eq:R:new:split} \end{aligned}$$ we have that $(v_{q+1},\mathring{R}_{q+1})$ solves at step $q+1$. In we have split up the Reynolds stress into a Transport, Nash, Dissipation, and Oscillation part, and have denoted by ${\widetilde}p_{q+1}$ a dummy scalar pressure (which is different from the $p_{q+1}$ pressure in equation ). Note that once $w_{q+1}$ is constructed to have frequency support inside the annulus $\{ \lambda_q/2 \leq |\xi|\leq 2 \lambda_q\}$, it follows from and the inductive assumptions on the frequency support of $v_q$ and $\mathring R_q$, that $\mathring R_{q+1}$ has frequency support inside the ball $\{ |\xi|\leq 4 \lambda_{q+1}\}$. The proof of Proposition \[prop:main\] is the main part of the paper, and is achieved in three steps. The first step, achieved in Section \[sec:perturbation\], is to construct the velocity increment $w_{q+1}$ which obeys the estimate , and verify that with this perturbation the bounds and – hold with $q$ replaced with $q+1$. The second step, achieved in Section \[sec:stress\], is to show that the induced Reynolds stress $\mathring{R}_{q+1}$ given by obeys estimates and with $q$ replaced with $q+1$. The third step, achieved in Section \[sec:hamiltonian\], is to show that the new velocity field is sufficiently close to the desired Hamiltonian profile, i.e. that bounds – hold with $q$ replaced with $q+1$. Together, these three steps give the proof of the proposition. Theorem \[thm:main\] is simple consequence of Proposition \[prop:main\], as we show next. Proof of Theorem \[thm:main\] ----------------------------- We start the iteration by setting $(v_{0}, \mathring R_{0})$ to be the trivial zero solution. Then - and follow trivially. Moreover, choosing $\lambda_0$ sufficiently large we can ensure $$\begin{aligned} \ee(t)\leq \lambda_1\delta_1=\lambda_0^{3-2\beta}\,,\end{aligned}$$ and thus holds. Then, we apply Proposition \[prop:main\] iteratively to obtain a sequence $v_q$ converging in $C^{\beta}$ to a weak solution $$v = v_0 + \sum_{q\geq 0} (v_{q+1} - v_q) = v_0 + \sum_{q\geq 0} w_{q+1}$$ of SQG. The convergence in $C^{\beta}$ follows directly from the frequency support of the perturbations $w_{q}$ and the estimate . Moreover, the estimate implies that $$\begin{aligned} \int_{\TT^2}{\left\vert\Lambda^{{\frac}12}v(x,t)\right\vert}^2\,dx=\ee(t)\,.\end{aligned}$$ Note, as a consequence of and , it follows that $$\begin{aligned} {\left\lVert\partial_t v_q\right\rVert}_{C^0}\leq &{\left\lVert(\partial_t+u_q\cdot \nabla)v_q\right\rVert}_{C^0}+{\left\lVertu_q\right\rVert}_{C^0}{\left\lVertv_q\right\rVert}_{C_1}\\ \lesssim & \lambda_q^2\delta_q\,.\end{aligned}$$ Thus, by interpolation, using the decomposition $w_q=v_{q}-v_{q-1}$ we obtain $$\begin{aligned} {\left\lVertw_q\right\rVert}_{C_t^{\sigma}C_x^0}\lesssim &{\left\lVertw_q\right\rVert}_{C_t^{0}C_x^0}^{1-\sigma}{\left\lVertv_q-v_{q-1}\right\rVert}_{C_t^{1}C_x^0}^{\sigma}\\ \lesssim &{\left\lVertw_q\right\rVert}_{C_t^{0}C_x^0}^{1-\sigma}\left({\left\lVertv_q\right\rVert}_{C_t^{1}C_x^0} + {\left\lVertv_{q-1}\right\rVert}_{C_t^{1}C_x^0}\right)^{\sigma}\\ \lesssim & \delta_q^{\frac{1-\sigma}{2}} \left(\lambda_q^2\delta_q\right)^{\sigma}\\ =&\lambda_0^2\lambda_q^{-(1-\sigma)\beta +2\sigma(1-\beta)}\\ =&\lambda_0^2\lambda_q^{-\beta+\sigma(2-\beta)}\,\end{aligned}$$ Hence, if $\sigma<\frac{\beta}{2-\beta}$, then $v_q$ convergences uniformly in ${C_t^{\sigma}C_x^0}$. The velocity perturbation {#sec:perturbation} ========================= Technical preliminaries ----------------------- ### Inverse of the divergence In defining $R_T, R_D, R_O$, and $R_N$, we need to use the fact that any divergence free vector function $f$ with zero mean on $\TT^2$ may be written as a divergence. More precisely: \[def:BB\] Let $f$ be divergence free and with zero mean on $\TT^2$. Then we have $$\begin{aligned} f = \div (\BB f), \quad \mbox{or in components} \quad f^i = \partial_j (\BB f)^{ij}\end{aligned}$$ where $$\begin{aligned} (\BB f)^{ij} := - \partial_j \Lambda^{-2} f^i - \partial_i \Lambda^{-2} f^j.\end{aligned}$$ For $f$ which is not necessarily divergence free, we define $$\begin{aligned} \BB f := \BB \PP f\,,\end{aligned}$$ where $\PP = {\ensuremath{\mathrm{Id}}}+ \RSZ\otimes \RSZ$ is the Leray projector. Lastly, when $f$ does not have zero mean on $\TT^2$, we define $$\begin{aligned} \BB f := \BB \left(f - \frac{1}{|\TT^2|}\int_{\TT^2} f(x) dx \right).\end{aligned}$$ In particular, we have that $\div(\BB f) = \PP f$ is divergence free, and $\BB f$ is a symmetric trace free matrix. Properties of the operator $\BB$ are discussed in Appendix \[sec:BB\] below. ### Beltrami plane waves {#sec:Beltrami_plane_waves} For $k \in \SS^1$, we define $$\begin{aligned} b_k(\xi) = i k^\perp e^{i k \cdot \xi} \qquad \mbox{and} \qquad c_k(\xi) = e^{i k \cdot \xi} \label{eq:bk:ck:def}\end{aligned}$$ where we notice that since $k \in \SS^1$, we have $$\begin{aligned} b_k = \nabla_\xi^\perp c_k \qquad \mbox{and} \qquad c_k = - \nabla_\xi^\perp \cdot b_k.\end{aligned}$$ It is also worth noting here that $b_k$ is an eigenfunction of $\Lambda$ with eigenvalue $1$, that is $$\begin{aligned} \Lambda_\xi b_k(\xi) = b_k(\xi)\,, \label{eq:bk-eigenvalue}\end{aligned}$$ since $k \in \SS^1$. Also it will be sometimes useful to note that since $(k^\perp)^\perp = - k$, we have $$\begin{aligned} (b_k(\xi))^\perp = - i k c_k(\xi) = - \nabla_\xi c_k(\xi). \label{eq:bk-perp}\end{aligned}$$ ### Geometric Lemma For any finite family of vectors $\Omega\subset \SS^1$ and constants $a_k\in \CC$, such that $a_{-k} = \overline{a_k}$, if we set $$\begin{aligned} W(\xi):=\sum_{k\in\Omega}a_k b_k(\xi) \qquad \mbox{and}\qquad V(\xi):=\sum_{k\in\Omega}a_k c_k(\xi)\,,\end{aligned}$$ then we have the following identity $$\label{eq:Beltrami_identity} \begin{split} \div_{\xi}(W\otimes W)=&\frac12 \nabla_{\xi}{\left\vertW\right\vert}^2+(\nabla^{\perp}_{\xi}\cdot W)W^{\perp}\\ =&\frac12 \nabla_{\xi}{\left\vertW\right\vert}^2-V\nabla V\\ =&\frac12 \nabla_{\xi}\left({\left\vertW\right\vert}^2-{\left\vertV\right\vert}^2\right). \end{split}$$ Adopting the notation $W_k(\xi)=a_k b_k(\xi)$, we also note that $$\begin{aligned} \sum_{k\in \Omega} W_k\otimes W_{-k}= \sum_{k\in \Omega} {\left\verta_k\right\vert}^2 k^{\perp}\otimes k^{\perp}\,.\end{aligned}$$ \[lem:split\] Let $B_{\eps}({\ensuremath{\mathrm{Id}}})$ denote the ball of symmetric $2\times 2$ matrices, centered at ${\ensuremath{\mathrm{Id}}}$ of radius $\eps$. We can choose $\eps_{\gamma}>0$ such that there exist disjoint finite subsets $$\Omega_j\subset \SS^1, \qquad j\in \{1, 2\}~,$$ and smooth positive functions $$\gamma_k\in C^{\infty}\left(B_{\eps_{\gamma}} ({\ensuremath{\mathrm{Id}}})\right), \qquad j\in \{1,2\}~, \qquad k\in\Omega_j~,$$ such that - For each $j$ we have $5 \Omega_j\subset \ZZ^2$. - If $k\in \Omega_j$ then $-k\in \Omega_j$ and $\gamma_k = \gamma_{-k}$. - For each $R\in B_{\eps_\gamma} ({\ensuremath{\mathrm{Id}}})$ we have the identity $$\label{e:split} R = \frac{1}{2} \sum_{k\in\Omega_j} \left(\gamma_k(R)\right)^2 (k^{\perp}\otimes k^{\perp}),$$ for all $R\in B_{\eps_{\gamma}}({\ensuremath{\mathrm{Id}}})$. - For $k, k' \in \Omega_j$, with $k+k' \neq 0$, we have that [$|k+k'| \geq \frac 12$]{}. First consider the case of constructing $\Omega_1$. Define $\Omega_1^+=\{k_1,k_2,k_3\}$ where $k_1:=(1,0)$, $k_2:=({\frac}35,{\frac}45)$ and $k_3:=({\frac}35,-{\frac}45)$. With these choices we make the following observations. First, the matrices $k_i^{\perp}\otimes k_i^{\perp}$ for $i=1,2,3$ are linearly independent. Second, we have the identity $$\begin{split} {\ensuremath{\mathrm{Id}}}=&\left(1- \frac{5^2\cdot3^2}{ 4^2\cdot 5^2}\right)k_1\otimes k_1+\frac{5^2}{2\cdot 4^2}k_2\otimes k_2+\frac{5^2}{2\cdot 4^2} k_3\otimes k_3\\ =&\frac{7}{16}k_1\otimes k_1+\frac{25}{32}k_2\otimes k_2+\frac{25}{32} k_3\otimes k_3\\ =&\frac{7}{16}k_1^{\perp}\otimes k_1^{\perp}+\frac{25}{32}k_2^{\perp}\otimes k_2^{\perp}+\frac{25}{32} k_3^{\perp}\otimes k_3^{\perp}. \end{split}$$ Hence, setting $\Omega_1=-\Omega_1^+\cup \Omega_1^+$, and applying the inverse function theorem to construct $\gamma_k$ we obtain properties (a)–(d). Similarly, setting $\Omega_2=\Omega_1^{\perp}= \{k^\perp \colon k \in \Omega_1\}$, we may construct $\gamma_k$ for $k\in \Omega_2$ and obtain properties (a)–(d). ### Time cutoffs and the back-to-labels map We let $0 \leq \chi \leq 1$ be a smooth cutoff function which is identically $1$ on $[1,2]$, vanishes on the complement of $[1/2,4]$, and defines a partition of unity according to $$\begin{aligned} \sum_{j \in \ZZ} \chi^2(t - j) = 1\end{aligned}$$ for all $t \in \RR$. We shall also define $$\begin{aligned} \chi_{j}(t) = \chi(t \tau_{q+1}^{-1} - j) \label{eq:chi:q+1:j:def}\end{aligned}$$ where for ease of notation we suppress the dependence of $\chi_j$ on $q$. For every $j \in \ZZ$, we define the following back-to-labels map $\Phi_j(x,t)$ by solving the transport equation $$\begin{aligned} \left( \partial_t + u_q \cdot \nabla \right) \Phi_j &= 0 \, ,\\ \Phi_j(x,j \tau_{q+1}) &= x \, ,\end{aligned}$$ where we define the time step parameter $\tau_{q+1}$ by $$\begin{aligned} \tau_{q+1}^{-1} = \lambda_{q} \lambda_{q+1} \delta_{q}^{1/4} \delta_{q+1}^{1/4} \, . \label{eq:tau:q}\end{aligned}$$ The motivation for this scaling of $\tau_{q+1}$ comes from balancing the oscillation and the transport error (cf. estimates and below). In particular, we note that $$\begin{aligned} \tau_{q+1} {\left\lVert\nabla u_q\right\rVert}_{C_0} \leq \tau_{q+1} \lambda_q^2 \delta_q^{1/2}=\lambda_0^{-1+\beta/2}, \label{eq:CFL}\end{aligned}$$ so that since $\beta<2$, then for $\lambda_0$ large, on a time interval of length $2 \tau_{q+1}$, the flow $\Phi_j$ induced by $u_q$ does not depart substantially from the identity. ### Leray projector and a frequency localizer Lastly, we define $\PP = {\ensuremath{\mathrm{Id}}}+ \RSZ \otimes \RSZ$ to be the Leray projector, and for $k \in \SS^1$ and $\lambda_{q+1}$ as above we set $$\begin{aligned} \PP_{q+1,k} = \PP P_{\approx k \lambda_{q+1}} \label{eq:PP:q+1:k:def}\end{aligned}$$ where $P_{\approx \lambda_{q+1} k}$ is a zero order Fourier multiplier operator with symbol $ {\widehat}K_{\approx k \lambda_{q+1}} = {\widehat}K_{\approx 1} (\xi/\lambda_{q+1} - k)$. That is, $$\begin{aligned} (P_{\approx k \lambda_{q+1}} f)^{{\widehat}\ }(\xi) = {\widehat}K_{\approx k \lambda_{q+1}}(\xi) {\widehat}f(\xi) = {\widehat}{K}_{\approx 1}\left( \frac{\xi}{\lambda_{q+1}} - k \right) {\widehat}{f}(\xi),\end{aligned}$$ where the function ${\widehat}K_{\approx 1}$ is a smooth bump function supported on the ball $\{ \xi \colon |\xi|\leq \frac{1}{8}\}$, and such that ${\widehat}K_{\approx 1}(\xi) = 1$ on the smaller ball $ \left\{ \xi \colon |\xi| \leq \frac{1}{16} \right\}$. Note in particular that $0 \leq a,b$ we have $$\sup_{\xi \in \RR^2} |\xi|^a |\nabla_{\xi}^b {\widehat}K_{\approx k \lambda_{q+1}} | \leq C_{a,b} \lambda_{q+1}^{a-b}$$ for a suitable constant $C_{a,b}$ that is independent of $\lambda_{q+1}$, and similarly $$\begin{aligned} {\left\lVert |x|^{b} \nabla_x^a K_{\approx k \lambda_{q+1} } \right\rVert}_{L^1_x(\RR^2)} \leq C_{a,b} \lambda_{q+1}^{a-b}\end{aligned}$$ holds for $0 \leq a,b \leq 2$, and the constant $C_{a,b}$ is independent of $\lambda_{q+1}$. We note here that $\PP_{q+1,k}$ is a convolution operator with kernel $K_{q+1,k}(x)$, i.e., for $f$ which is $\TT^2$-periodic we may write $$\PP_{q+1,k} f(x) = \int_{\RR^2} K_{q+1,k}(y) f(x-y) dy\,,$$ with a kernel $K_{q+1,k}$ that obeys $$\begin{aligned} {\left\lVert |x|^{b} \nabla_x^a K_{q+1,k}(x) \right\rVert}_{L^1_x(\RR^2)} \leq C_{a,b} \lambda_{q+1}^{a-b} \label{eq:projection:kernel}\end{aligned}$$ for $0 \leq a,b$, and the constant $C_{a,b}$ is independent of $\lambda_{q+1}$. Here we have implicitly used that $\RSZ \otimes \RSZ = \PP - {\ensuremath{\mathrm{Id}}}$ is a matrix zero order Fourier multiplier, whose symbol is smooth away from the origin. Construction of the perturbation {#sec:w:q+1:construction} -------------------------------- With these notations in hand, we now define the potential velocity perturbation $w_{q+1}$ as $$\begin{aligned} w_{q+1}(x,t) = \sum_{j \in \ZZ,k \in \Omega_{j}} \chi_{j}(t) \PP_{q+1,k} \big( a_{k,j}(x,t) b_k( \lambda_{q+1} \Phi_j(x,t) ) \big)\,, \label{eq:w:q+1:def}\end{aligned}$$ where the functions $a_{k,j}(x,t)$ are to be defined in below, and $\Omega_j = \Omega_1$ if $j$ is odd, while $\Omega_j = \Omega_2$ if $j$ is even. The definition of $\PP_{q+1,k}$ implies that the increment $w_{q+1}$ has compact support in frequency space inside $\{ \xi \colon \lambda_{q+1}/2 \leq |\xi|\leq 2 \lambda_{q+1}\}$, as required in the inductive step. Let $\mathring R_{q,j}$ define the solution to the transport equation: \[eq:R:q:transported\] $$\begin{aligned} \left( \partial_t + u_q \cdot \nabla \right) \mathring R_{q,j} &= 0 \, ,\\ \mathring R_{q,j}(x,j \tau_{q+1}) &= \mathring R_q(x,j \tau_{q+1}) \, ,\end{aligned}$$ and set $$R_{q,j}:= {\lambda_{q+1}} \rho_j{\ensuremath{\mathrm{Id}}}- \mathring R_{q,j} \label{eq:R:q:def}$$ where $$\label{eq:rho_def} \rho(t) := \frac{1}{(2\pi)^2}\min\left(\ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_q\right\vert}^2~dx-\frac{\lambda_{q+ 2}\delta_{q+2}}{2},\, 0\right)~\mbox{and}~ \rho_j = \rho(\tau_{q+1}j)\,.$$ The constants $\rho_j$ are chosen in order to ensure convergence of the Hamiltonian to the desired profile. Note that by the inductive assumption  we have that $$\begin{aligned} \rho_j \leq \delta_{q+1}\,.\end{aligned}$$ Then to conclude our definition of the perturbation $w_{q+1}$, we define $$\begin{aligned} a_{k,j}(x,t):= \begin{cases} \rho_j^{{\frac}12}\gamma_k\left(\frac{ R_{q,j}(x,t)}{{\lambda_{q+1}} \rho_j}\right) &\mbox{if } \rho_j\neq 0 \\ 0 & \mbox{if } \rho_j=0. \end{cases} \label{eq:ak:def}\end{aligned}$$ Note that in order that $a_{k,j}$ is well defined we need to ensure that if $\rho_j\neq 0$ then $$\frac{ R_{q,j}(x,t)}{{\lambda_{q+1}} \rho_j}\in B_{\eps_{\gamma}({\ensuremath{\mathrm{Id}}})}\,.$$ Since $R_{q,j}$ satisfies a transport equation it suffices to prove that $$\label{eq:rho_cond} \frac{{\left\lVert\mathring R_{q}(\cdot,j\tau_{q+1})\right\rVert}_{C^0)}}{{\lambda_{q+1}} \rho_j}\leq \eps_{\gamma}\,.$$ By is suffices to consider the case when $$e(j\tau_{q+1}) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_q(x,j\tau_{q+1})\right\vert}^2~dx\geq \frac{\lambda_{q+1}\delta_{q+1}}{8}\,,$$ which implies (assuming $\lambda_0$ is chosen sufficiently large) $$\rho_j\geq \frac{\lambda_q\delta_{q+1}}{8}.$$ Applying yields $$\begin{aligned} \frac{\|\mathring R_q(\cdot, j\tau_{q+1})\|_{C^0} }{\lambda_{q+1}\rho_j} &\leq 8\eps_R\end{aligned}$$ Then as long as $2\eps_R\leq \eps_{\gamma}$ we obtain . Throughout the paper it will be sometimes convenient to denote $$\begin{aligned} {\widetilde}w_{q+1,j,k} = \chi_{j}(t) a_{k,j}(x,t) b_k( \lambda_{q+1} \Phi_j(x,t) ) \label{eq:tilde:w:q+1:def}\end{aligned}$$ so that in adopting the abuse of notation $\sum_{j ,k}=\sum_{\{j:\rho_j\neq 0\},k \in \Omega_{j}}$, the equation reads $$\begin{aligned} w_{q+1}(x,t) = \sum_{j ,k} \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\,.\end{aligned}$$ We also adopt the notation $$\begin{aligned} \psi_{q+1,j,k}(x,t)= \frac{c_k(\lambda_{q+1} \Phi_j(x,t))}{c_k(\lambda_{q+1} x)} = e^{i \lambda_{q+1} (\Phi_j(x,t) - x) \cdot k} \label{eq:psi:q+1:j:k:def}\end{aligned}$$ so that $$\begin{aligned} b_k( \lambda_{q+1} \Phi_j(x,t) ) = b_k(\lambda_{q+1} x) \psi_{q+1,j,k}(x,t)~.\end{aligned}$$ Bounds on the perturbation -------------------------- \[lem:w:q+1:bounds\] With $w_{q+1}$ as defined in , we have that $$\begin{aligned} {\left\lVertw_{q+1}\right\rVert}_{C^0} &\leq C_0 \delta_{q+1}^{1/2} \label{eq:w:q+1}\\ {\left\lVertv_{q+1}\right\rVert}_{C^1} +{\left\lVertu_{q+1}\right\rVert}_{C^0} &\leq C_0 \delta_{q+1}^{1/2} \lambda_{q+1} \label{eq:v+u:q+1}\\ {\left\lVertD_{t,q} w_{q+1}\right\rVert}_{C^0} &\leq C_0 \tau_{q+1}^{-1} \delta_{q+1}^{1/2} \label{eq:Dt:w:q+1}\\ {\left\lVertD_{t,q+1} v_{q+1}\right\rVert}_{C^0} &\leq C_0 \lambda_{q+1}^2 \delta_{q+1} \label{eq:Dt:v:q+1}\\ {\left\lVertD_{t,q+1} u_{q+1}\right\rVert}_{C^0} &\leq C_0 \lambda_{q+1}^3 \delta_{q+1} \label{eq:Dt:u:q+1}\end{aligned}$$ for a universal constant $C_0 \geq 1$, which is the same as the constant in Section \[s:inductive\_assump\]. In particular, the bounds , , and show that the inductive estimates , , and hold with $q$ replaced with $q+1$. From it follows that ${\left\lVert\PP_{q+1,k}\right\rVert}_{C^0 \to C^0} \lesssim 1$, and hence $$\begin{aligned} {\left\lVertw_{q+1}\right\rVert}_{C^0} &\lesssim \sum_{j,k} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert} \notag\\ &\lesssim \sum_{j,k} {\bf 1}_{\supp (\chi_{j})} {\left\lVerta_{k,j}\right\rVert}_{C^0} \lesssim \sum_{j} {\bf 1}_{\supp (\chi_{j})} \rho_j^{1/2} \lesssim \delta_{q+1}^{1/2}\end{aligned}$$ in view of , and the fact that $\rho_j \lesssim \delta_{q+1}^{1/2}$. Estimate follows from , the frequency support of $w_{q+1}$, and the inductive estimates . In order to estimate the material derivative $D_{t,q} = \partial_t + u_q \cdot\nabla$ of $w_{q+1}$, we note that $D_{t,q} a_{k,j}(x,t) = 0$ since $\mathring R_{q,j}$ obey the transport equation for the vector field $u_q$, and similarly $D_{t,q} b_k(\lambda_{q+1} \Phi_j(x,t)) = 0$ since the phases $\Phi_j$ also obey the same transport equation. We thus have that $$\begin{aligned} D_{t,q} w_{q+1} &= D_{t,q} \left( \sum_{j,k} \PP_{q+1,k} {\widetilde}w_{q+1,j,k} \right) \notag\\ &= \sum_{j,k} \PP_{q+1,k} \big( (\partial_t \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j ) \big) + \sum_{j,k} [D_{t,q},\PP_{q+1,k}] {\widetilde}w_{q+1,j,k} .\end{aligned}$$ Therefore, by appealing to the boundedness on $C^0$ of $\PP_{q+1,k}$, the definition of $\chi_{j}$ in , and the commutator estimate in Corollary \[cor:Dt:commutator\] with $s=0$, we arrive at $$\begin{aligned} {\left\lVertD_{t,q} w_{q+1}\right\rVert}_{C^0} &\lesssim \tau_{q+1}^{-1} \sum_{j,k} {\bf 1}_{\supp (\chi_{j})} {\left\lVerta_{k,j}\right\rVert}_{C^0} + \sum_{j,k} {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} \notag\\ &\lesssim \tau_{q+1}^{-1} \sum_j {\bf 1}_{\supp (\chi_{j})} \rho_j^{1/2} + \lambda_q^2 \delta_q^{1/2} \sum_j {\bf 1}_{\supp (\chi_{j})} \rho_j^{1/2} \notag\\ &\lesssim \left(\tau_{q+1}^{-1} + \lambda_q^2 \delta_q^{1/2} \right) \delta_{q+1}^{1/2} \notag\\ &\lesssim \tau_{q+1}^{-1} \delta_{q+1}^{1/2}\end{aligned}$$ since $\rho_j \lesssim \delta_{q+1} $ and by we have $\tau_{q+1}^{-1} = \lambda_{q} \lambda_{q+1} \delta_{q}^{1/4} \delta_{q+1}^{1/4} \geq \lambda_q^{2} \delta_{q}^{1/2}$, where here we used the fact that $\beta<2$. Using the inductive estimate and the bound established above, since $$\begin{aligned} D_{t,q+1} v_{q+1} = D_{t,q} (v_q + w_{q+1}) + \Lambda w_{q+1} \cdot \nabla (v_q + w_{q+1})\end{aligned}$$ we obtain $$\begin{aligned} {\left\lVertD_{t,q+1} v_{q+1}\right\rVert}_{C^0} \lesssim \lambda_q^2 \delta_q + \tau_{q+1}^{-1} \delta_{q+1}^{1/2} + \lambda_{q+1}^2 \delta_{q+1} \lesssim \lambda_{q+1}^2 \delta_{q+1}.\end{aligned}$$ Similarly, $$\begin{aligned} D_{t,q+1} u_{q+1} = D_{t,q} (u_q + {\widetilde}P_{\approx \lambda_{q+1}} \Lambda w_{q+1}) + \Lambda w_{q+1} \cdot \nabla (u_q + \Lambda w_{q+1} ), \end{aligned}$$ the inductive estimate , and Lemma \[lem:u:grad:commutator\] (with $s=1$ and $\lambda = \lambda_{q+1}$) implies $$\begin{aligned} {\left\lVertD_{t,q+1} u_{q+1}\right\rVert}_{C^0} &\lesssim \lambda_q^3 \delta_q + {\left\lVert{\widetilde}P_{\approx \lambda_{q+1}} \Lambda (D_{t,q} w_{q+1})\right\rVert}_{C^0} + {\left\lVert[u_q \cdot \nabla,{\widetilde}P_{\approx \lambda_{q+1}} \Lambda] w_{q+1}\right\rVert}_{C^0} + \lambda_{q+1}^3 \delta_{q+1}\\ &\lesssim \lambda_q^3 \delta_q + \lambda_{q+1} \tau_{q+1}^{-1} \delta_{q+1}^{1/2} + \lambda_{q+1} \lambda_q^2 \delta_{q}^{1/2} \delta_{q+1}^{1/2} + \lambda_{q+1}^3 \delta_{q+1} \notag\\ &\lesssim \lambda_{q+1}^3 \delta_{q+1}\end{aligned}$$ which concludes the proof of the lemma. In estimating the Reynolds stress error, the following bounds concerning the derivatives of $a_{k,j}$ and $\psi_{q+1,j,k}$ are very useful. \[lem:D:N:a:psi\] With $a_{k,j}$ as defined in and $\psi_{q+1,j,k}$ as defined in , we have that $$\begin{aligned} {\left\lVertD^N a_{k,j}\right\rVert}_{C^0(\supp \chi_{j})} &\lesssim \lambda_{q}^N \delta_{q+1}^{1/2} \label{eq:D:N:ak}\end{aligned}$$ for all $N \geq 0$, and $$\begin{aligned} {\left\lVertD^N \psi_{q+1,j,k}\right\rVert}_{C^0(\supp \chi_{j})} &\lesssim \left(\tau_{q+1} \lambda_{q+1} \lambda_q^2 \delta_q^{1/2}\right)^N = \left(\lambda_q \delta_q^{1/4} \delta_{q+1}^{-1/4}\right)^N \label{eq:D:N:psi:j:k}\end{aligned}$$ for all $N \geq 1$, where the implied constant depends on $N$. Since $R_q$ is supported on frequencies less than $4 \lambda_{q}$, it follows from the chain rule estimate  and the smoothness of the $\gamma_k$ functions that $$\begin{aligned} {\left\lVertD^N a_{k,j}\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} \rho_j^{-1/2} {\left\lVertD^N R_{q,j}\right\rVert}_{C^0} + \rho_j^{1/2-N} \lambda_{q+1}^{-N} {\left\lVertD R_{q+1,j}\right\rVert}_{C^0}^N\\ &\lesssim \delta_{q+1}^{ 1/2} \lambda_q^N + \delta_{q+1}^{1/2} \lambda_{q+1}^{-N} (\lambda_q \lambda_{q+1})^N\end{aligned}$$ where we have also used that $\rho_j \lesssim \delta_{q+1}$. This proves . In order to prove , we appeal to the chain rule estimate  and the transport estimates – and find that $$\begin{aligned} {\left\lVertD^N \psi_{q+1,j,k}\right\rVert}_{C^0(\supp \chi_{j})} &\quad \lesssim \lambda_{q+1} {\left\lVertD^{N-1}\left( D\Phi_j-{\ensuremath{\mathrm{Id}}}\right)\right\rVert}_{C^0(\supp \chi_{j})} + \lambda_{q+1}^{N} {\left\lVertD \Phi_j - {\ensuremath{\mathrm{Id}}}\right\rVert}_{C^0(\supp \chi_{j})}^N \\ &\quad \lesssim \lambda_{q+1} \tau_{q+1} {\left\lVertD^N u_q\right\rVert}_{C^0} e^{C \tau_{q+1} \|D u_q\|_{C^0}} + \lambda_{q+1}^{N} \left( \tau_{q+1} {\left\lVertD u_q\right\rVert}_{C^0} e^{C \tau_{q+1} \|D u_q\|_{C^0}} \right)^N\end{aligned}$$ for a suitable constant $C$. From , $\tau_{q+1} {\left\lVertD u_q\right\rVert}_{C^0} \leq 1$, and thus we conclude that $$\begin{aligned} {\left\lVertD^N \psi_{q+1,j,k}\right\rVert}_{C^0(\supp \chi_{j})} &\lesssim \lambda_{q+1} \tau_{q+1} \lambda_q^{N+1} \delta_q^{1/2} + \lambda_{q+1}^{N} \left( \tau_{q+1} \lambda_q^2 \delta_q^{1/2} \right)^N \notag\\ &\lesssim \lambda_q^N \delta_q^{1/4} \delta_{q+1}^{-1/4} + \lambda_q^N \left(\delta_q^{1/4} \delta_{q+1}^{-1/4}\right)^N \notag\\ &\lesssim \left(\lambda_q \delta_q^{1/4} \delta_{q+1}^{-1/4}\right)^N\end{aligned}$$ since $\delta_{q+1}\leq \delta_q$. This estimate shows that $\psi_{q+1,j,k}$ lives at spatial frequency $\lambda_q \delta_q^{1/4} \delta_{q+1}^{-1/4}$. The Reynolds stress error {#sec:stress} ========================= Transport error --------------- \[lem:transp:error\] For any $\eps>0$, if $\lambda_0$ is sufficiently large then for $R_T$ as defined in , we have that $$\begin{aligned} {\left\lVertR_{T}\right\rVert}_{C^0} &\leq \eps \lambda_{q+2} \delta_{q+2} \\ {\left\lVertD_{t,q} R_T\right\rVert}_{C^0} &\leq \eps \lambda_{q+1}^2 \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2}\,.\end{aligned}$$ ### Amplitude of the Transport error By definition, we have that $$\begin{aligned} R_{T}&=\BB \left( D_{t,q} w_{q+1} \right) \notag \\ &=\BB\left(\sum_{j,k} \PP_{q+1,k} \big( (\partial_t \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j ) \big) + \sum_{j,k} [D_{t,q},\PP_{q+1,k}] {\widetilde}w_{q+1,j,k}\right) \notag\\ &=\BB {\widetilde}P_{\approx \lambda_{q+1}} \left(\sum_{j,k} \PP_{q+1,k} \big( (\partial_t \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j ) \big) + \sum_{j,k} [u_q \cdot \nabla,\PP_{q+1,k}] {\widetilde}w_{q+1,j,k}\right) \end{aligned}$$ where in the last equality we used the compact support of $u_q$ to conclude that $u_q \cdot \nabla \PP_{q+1,k} = {\widetilde}P_{\approx \lambda_{q+1}} \left( u_q \cdot \nabla \PP_{q+1,k}\right)$. Using Lemma \[lem:u:grad:commutator\] we obtain $$\begin{aligned} {\left\lVertR_T\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} \sum_{j,k} {\left\lVert(\partial_t \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j )\right\rVert}_{C^0} + \lambda_{q+1}^{-1} \sum_{j,k} {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVert\chi_j a_{k,j} b_k(\lambda_{q+1} \Phi_j )\right\rVert}_{C^0} \notag\\ &\lesssim \lambda_{q+1}^{-1} \tau_{q+1}^{-1} \delta_{q+1}^{1/2} + \lambda_{q+1}^{-1} \lambda_q^{2} \delta_q^{1/2} \delta_{q+1}^{1/2} \notag\\ &\lesssim \lambda_{q+1}^{-1} \tau_{q+1}^{-1} \delta_{q+1}^{1/2}\notag\\ &= \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4} \notag\\ &= \lambda_0^{-2+\frac{5\beta}{2}}\lambda_{q+2} \delta_{q+2}. \label{eq:sharp:1}\end{aligned}$$ Assuming $\lambda_0$ is sufficiently large and $\beta<{\frac}45$ we obtain our claim. ### Material derivative of the Transport error Using the frequency support of $u_q$ and $w_{q+1}$ we write $$\begin{aligned} D_{t,q}R_{T} &=[D_{t , q}, \BB {\widetilde}P_{\approx_{\lambda_{q+1}}}] D_{t,q} w_{q+1} +\BB {\widetilde}P_{\approx_{\lambda_{q+1}}}D_{t,q}\left(\sum_{j,k} \PP_{q+1,k} \big( (\partial_t \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j ) \big)\right) \notag\\ &\quad +\BB {\widetilde}P_{\approx_{\lambda_{q+1}}}D_{t,q} \left( \sum_{j,k} [D_{t,q},\PP_{q+1,k}] {\widetilde}w_{q+1,j,k}\right) \notag \\ &=T_1+T_2+T_3 \label{eq:RT:Dt}\end{aligned}$$ The first term in is bounded directly using Corollary \[cor:Dt:commutator\] and the bound as $$\begin{aligned} {\left\lVertT_1\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVertD_{t,q} w_{q+1}\right\rVert}_{C^0} \notag\\ &\lesssim \lambda_{q+1}^{-1} \lambda_q^2 \delta_q^{1/2} \tau_{q+1}^{-1} \delta_{q+1}^{1/2} \notag\\ &= \lambda_q^3 \delta_{q+1}^{3/4} \delta_q^{3/4}. \label{eq:T_1_est}\end{aligned}$$ We decompose the second term in as $$\begin{aligned} T_2= &\BB {\widetilde}P_{\approx_{\lambda_{q+1}}} \left(\sum_{j,k}[D_{t,q},\PP_{q+1,k}]\big( (\partial_t \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j ) \big)+\sum_{j,k}\PP_{q+1,k}(\partial_t^2 \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j )\right)\end{aligned}$$ which allows it to be estimated by $$\begin{aligned} {\left\lVertT_2\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} \sum_{j,k} \left( {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVert(\partial_t \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j )\right\rVert}_{C^0} + {\left\lVert(\partial_t^2 \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j )\right\rVert}_{C^0} \right) \notag\\ &\lesssim \lambda_{q+1}^{-1} \left( \lambda_q^2 \delta_{q}^{1/2} \tau_{q+1}^{-1} \delta_{q+1}^{1/2} + \tau_{q+1}^{-2} \delta_{q+1}^{1/2} \right) \notag\\ & \lesssim \lambda_{q+1}^{-1} \tau_{q+1}^{-2} \delta_{q+1}^{1/2} \notag\\ &= \lambda_q^2\lambda_{q+1} \delta_{q}^{1/2} \delta_{q+1}. \label{eq:T_2_est}\end{aligned}$$ Next, $$\begin{aligned} T_3 &=\BB {\widetilde}P_{\approx_{\lambda_{q+1}}} D_{t,q} \left( \sum_{j,k}u_q\cdot \nabla\PP_{q+1,k} {\widetilde}w_{q+1,j,k}-\sum_{j,k}\PP_{q+1,k} \left(u_q\cdot \nabla{\widetilde}w_{q+1,j,k}\right)\right) \notag \\ &=\BB{\widetilde}P_{\approx_{\lambda_{q+1}}} \left(\sum_{j,k}(D_{t,q}u_q)\cdot \nabla\left(\PP_{q+1,k} {\widetilde}w_{q+1,j,k}\right)+\sum_{j,k} u_q\cdot \nabla \left([D_{t,q},\PP_{q+1,k}] {\widetilde}w_{q+1,j,k}\right)\right) \notag \\ &\quad+\BB{\widetilde}P_{\approx_{\lambda_{q+1}}} \left(\sum_{j,k}u_q\cdot \nabla \left(\PP_{q+1,k} \big( (\partial_t \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j ) \big)\right) - \sum_{j,k} \left((u_q \cdot \nabla) u_q \right) \cdot \nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k} \right) \notag \\ &\quad-\BB{\widetilde}P_{\approx_{\lambda_{q+1}}} \left(\sum_{j,k}[D_{t,q},\PP_{q+1,k}] \left(u_q\cdot \nabla{\widetilde}w_{q+1,j,k}\right)+\sum_{j,k}\PP_{q+1,k}\left((D_{t,q}u_q)\cdot \nabla{\widetilde}w_{q+1,j,k}\right)\right) \notag \\ &\quad-\BB{\widetilde}P_{\approx_{\lambda_{q+1}}} \left(\sum_{j,k}\PP_{q+1,k}\left(u_q\cdot \nabla\big( (\partial_t \chi_{j} ) a_{k,j} b_k(\lambda_{q+1} \Phi_j ) \big)\right) - \sum_{j,k} \PP_{q+1,k} \left( \left( (u_q \cdot \nabla) u_q \right) \cdot \nabla {\widetilde}w_{q+1,j,k} \right)\right).\end{aligned}$$ In order to estimate $T_3$ we appeal to the inductive bound , Lemma \[lem:D:N:a:psi\], the commutator estimate in Corollary \[cor:Dt:commutator\] (with $s = 0$ and $\lambda = \lambda_{q+1}$), and to the fact that $\tau_{q+1} \lambda_{q}^2 \delta_q^{1/2} \lesssim 1$ and the estimate , imply $${\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^1} \lesssim {\left\lVerta_{k,j}\right\rVert}_{C^1} + {\left\lVerta_{k,j}\right\rVert}_{C^0} \lambda_{q+1} {\left\lVert\nabla \Phi_j\right\rVert}_{C^0} \lesssim \lambda_q \delta_{q+1}^{1/2} + \lambda_{q+1} \delta_{q+1}^{1/2} \lesssim \lambda_{q+1} \delta_{q+1}^{1/2}.$$ All these yield $$\begin{aligned} {\left\lVertT_3\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} \sum_{j,k} \left( {\left\lVertD_{t,q} u_q\right\rVert}_{C^0} \lambda_{q+1} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} + {\left\lVertu_q\right\rVert}_{C^0} \lambda_{q+1} {\left\lVertu_q\right\rVert}_{C^1} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} \right) \notag\\ &\quad + \lambda_{q+1}^{-1} \sum_{j,k} \left( {\left\lVertu_q\right\rVert}_{C^0} \lambda_{q+1} \tau_{q+1}^{-1} {\left\lVerta_{k,j}\right\rVert}_{C^0(\supp \chi_j)} + {\left\lVertu_q\right\rVert}_{C^0} {\left\lVertu_q\right\rVert}_{C^1} \lambda_{q+1} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} \right) \notag\\ &\quad + \lambda_{q+1}^{-1} \sum_{j,k}\left( {\left\lVertu_q\right\rVert}_{C^1} {\left\lVertu_q\right\rVert}_{C^0} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^1} + {\left\lVertD_{t,q} u_q\right\rVert}_{C^0} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^1}\right) \notag\\ &\quad + \lambda_{q+1}^{-1} \sum_{j,k}\left( {\left\lVertu_q\right\rVert}_{C^0} \tau_{q+1}^{-1} {\left\lVerta_{k,j} b_k(\lambda_{q+1} \Phi_j)\right\rVert}_{C^1} + {\left\lVertu_q\right\rVert}_{C^0} \lambda_{q} {\left\lVertu_q\right\rVert}_{C^1} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^1}\right) \notag \\ &\lesssim \lambda_{q}^3 \delta_q \delta_{q+1}^{1/2} + \lambda_q \delta_q^{1/2} \tau_{q+1}^{-1} \delta_{q+1}^{1/2} \notag\\ &\lesssim \lambda_q \delta_q^{1/2} \tau_{q+1}^{-1} \delta_{q+1}^{1/2} \notag\\ &= \lambda_q^2\lambda_{q+1} \delta_q^{3/4}\delta_{q+1}^{3/4}. \label{eq:T_3_est}\end{aligned}$$ Combining , and we obtain $$\begin{aligned} {\left\lVertD_{t,q}R_T\right\rVert}_{C^0}\lesssim &\lambda_q^3\delta_q^{3/4}\delta_{q+1}^{3/4}+\lambda_q^2\lambda_{q+1} \delta_{q}^{1/2}\delta_{q+1}+\lambda_q^2\lambda_{q+1}\delta_q^{3/4}\delta_{q+1}^{3/4}\\ \lesssim & \lambda_q^2\lambda_{q+1}\delta_q^{3/4}\delta_{q+1}^{3/4}\\ =&\lambda_0^{-3+7\beta/2}\lambda_{q+1}^{2}\ \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2},.\end{aligned}$$ This completes the proof of Lemma \[lem:transp:error\]. Nash error ---------- \[lem:Nash:error\] For any $\eps>0$, if $\lambda_0$ is sufficiently large then for $R_N$ as defined in , we have that $$\begin{aligned} {\left\lVertR_{N}\right\rVert}_{C^0} &\leq \eps \lambda_{q+2} \delta_{q+2} ,\notag \\ {\left\lVertD_{t,q} R_N\right\rVert}_{C^0} &\leq \eps \lambda_{q+1}^2\delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2}\,.\label{eq:Dt_Nash}\end{aligned}$$ ### Amplitude of the Nash error We recall from and the definition of $\BB$ that we may write $$\begin{aligned} R_N &=- \BB \left( (\nabla u_q)^T \cdot w_{q+1} \right) + \BB \left(\Lambda w_{q+1} \cdot \nabla v_q - (\nabla v_q)^T \cdot \Lambda w_{q+1}\right) \notag\\ &= - \BB \left( (\nabla u_q)^T \cdot w_{q+1} \right) + \BB \left((\nabla^\perp \cdot v_q) \Lambda w_{q+1}^\perp \right) \notag\\ &=: N_1 + N_2. \label{eq:RN:def}\end{aligned}$$ The bound on $N_1$ is direct. Recalling the definition of $w_{q+1}$, and recalling that $u_q$ is supported on frequencies $|\xi| \leq 2\lambda_q \leq \lambda_{q+1}/8$, upon applying with $\lambda = \lambda_{q+1}$ (which is inherent in $\PP_{q+1,k}$), we obtain $$\begin{aligned} {\left\lVertN_1\right\rVert}_{C^0} &\lesssim \frac{{\left\lVertu_q\right\rVert}_{C^1} }{\lambda_{q+1}} \sum_{j,k} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} \notag\\ &\lesssim \frac{\lambda_q^2 \delta_q^{1/2} \delta_{q+1}^{1/2}}{\lambda_{q+1}}.\end{aligned}$$ It is convenient to rewrite the term $N_2$, exploring the special structure of the perturbations $w_{q+1}$. We note that since $b_k^\perp(\xi) = i (k^\perp)^\perp e^{i k\cdot \xi} = - i k e^{i k \cdot \xi} = - \nabla_\xi c_k(\xi)$, we may write $$\begin{aligned} \Lambda w_{q+1}^\perp &= \sum_{j,k} \Lambda \PP_{q+1,k} \big(\chi_j a_{k,j} \psi_{q+1,j,k} b_k(\lambda_{q+1} x)^\perp\big) \notag\\ &= -\frac{1}{\lambda_{q+1}} \sum_{j,k} \Lambda \PP_{q+1,k} \big(\chi_j a_{k,j} \psi_{q+1,j,k} \nabla c_k(\lambda_{q+1} x) \big) \notag\\ &= -\frac{1}{\lambda_{q+1}} \nabla \sum_{j,k} \Lambda \PP_{q+1,k} \big(\chi_j a_{k,j} \psi_{q+1,j,k}c_k(\lambda_{q+1} x) \big) + \frac{1}{\lambda_{q+1}} \sum_{j,k} \Lambda \PP_{q+1,k} \big(\chi_j \nabla \big( a_{k,j} \psi_{q+1,j,k} \big) c_k(\lambda_{q+1} x) \big). \end{aligned}$$ Therefore, recalling that $\BB$ has incorporated into it the Leray projector $\PP$, we have that $$\begin{aligned} N_2 &= \frac{1}{\lambda_{q+1}} \BB\left( \nabla (\nabla^\perp \cdot v_q) \sum_{j,k} \Lambda \PP_{q+1,k} \big(\chi_j a_{k,j} c_k(\lambda_{q+1} \Phi_j ) \big) \right)\notag\\ &\quad +\frac{1}{\lambda_{q+1}} \BB\left( (\nabla^\perp \cdot v_q) \sum_{j,k} \Lambda \PP_{q+1,k} \big( \nabla \big(\chi_j a_{k,j} \psi_{q+1,j,k} \big) c_k(\lambda_{q+1} x) \big) \right). \label{eq:N2:rewriting}\end{aligned}$$ In order to bound $N_2$ we again use with $\lambda = \lambda_{q+1}$ and obtain $$\begin{aligned} {\left\lVertN_2\right\rVert}_{C^0} &\lesssim \frac{1}{\lambda_{q+1}^2} {\left\lVertv_q\right\rVert}_{C^2} \sum_{j,k} {\left\lVert\Lambda P_{\approx \lambda_{q+1}} (\chi_j a_{k,j} c_k(\lambda_{q+1} \Phi_j))\right\rVert}_{C^0} \notag\\ &\quad + \frac{1}{\lambda_{q+1}^2} {\left\lVertv_q\right\rVert}_{C^1} \sum_{j,k} {\left\lVert\Lambda P_{\approx \lambda_{q+1}} \Big(\chi_j \nabla a_{k,j} c_k(\lambda_{q+1} \Phi_j) + \chi_j a_{k,j} \nabla \psi_{q+1,j,k} c_k(\lambda_{q+1} x) \Big)\right\rVert}_{C^0} \notag\\ &\lesssim \frac{1}{\lambda_{q+1}^2} \lambda_q^2 \delta_{q}^{1/2} \lambda_{q+1} \delta_{q+1}^{1/2} + \frac{1}{\lambda_{q+1}^2} \lambda_q \delta_{q}^{1/2} \lambda_{q+1} \left(\lambda_q \delta_{q+1}^{1/2} + \delta_{q+1}^{1/2} \lambda_q\delta_q^{1/4}\delta_{q+1}^{-1/4} \right) \\ &\lesssim \frac{\lambda_q^2 \delta_q^{1/2} \delta_{q+1}^{1/2}}{\lambda_{q+1}} + \frac{\lambda_q^2}{\lambda_{q+1}} \delta_{q}^{3/4} \delta_{q+1}^{1/4} \end{aligned}$$ where in the second last inequality we have used Lemma \[lem:D:N:a:psi\]. We see that the first terms in the above bound obeys the same estimate as $N_1$. Then again using $\delta_{q+1}<\delta_q$ we obtain $$\begin{aligned} {\left\lVertR_N\right\rVert}_{C^0}\lesssim \frac{\lambda_q^2}{\lambda_{q+1}} \delta_{q}^{3/4} \delta_{q+1}^{1/4} = \lambda_0^{-3+\frac{7 \beta}{2}}\lambda_{q+2} \delta_{q+2}\end{aligned}$$ Thus we obtain the desired estimate so long as $\beta<\frac 67$ and $\lambda_0$ is sufficiently large. ### Material derivative of the Nash error Recall that $u_q$ has frequency support inside of the ball of radius $2 \lambda_q \leq \lambda_{q+1}/8$, and therefore $$\begin{aligned} N_1= \BB \left( (\nabla u_q)^T \cdot w_{q+1} \right) = \BB {\widetilde}P_{\approx \lambda_{q+1}} \left( (\nabla u_q)^T \cdot w_{q+1} \right)\end{aligned}$$ where we denote by ${\widetilde}P_{\approx \lambda_{q+1}}$ the Fourier multiplier operator whose symbol is supported on frequencies $\{\xi \colon \lambda_{q+1}/4 \leq |\xi| \leq 4 \lambda_{q+1}\}$, and is identically $1$ on the annulus $\{\xi \colon 3 \lambda_{q+1}/8 \leq |\xi| \leq 3 \lambda_{q+1} \}$. Therefore, appealing to Corollary \[cor:Dt:commutator\] (with $s=-1$ and $\lambda = \lambda_{q+1}$), we have that $$\begin{aligned} {\left\lVertD_{t,q} N_1\right\rVert}_{C^0} &\leq {\left\lVert\BB {\widetilde}P_{\approx \lambda_{q+1}} D_{t,q} \left( (\nabla u_q)^T \cdot w_{q+1} \right)\right\rVert}_{C^0} + {\left\lVert[D_{t,q}, \BB {\widetilde}P_{\approx \lambda_{q+1}}] \left( (\nabla u_q)^T \cdot w_{q+1} \right)\right\rVert}_{C^0} \notag\\ &\lesssim \frac{1}{\lambda_{q+1}} {\left\lVertD_{t,q} \left( (\nabla u_q)^T \cdot w_{q+1} \right)\right\rVert}_{C^0} + \frac{1}{\lambda_{q+1}} {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVert(\nabla u_q)^T \cdot w_{q+1}\right\rVert}_{C^0} \notag\\ &\lesssim \frac{1}{\lambda_{q+1}} \left( {\left\lVertD_{t,q} (\nabla u_q)^T\right\rVert}_{C^0} {\left\lVertw_{q+1} \right\rVert}_{C^0} + {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVertD_{t,q} w_{q+1}\right\rVert}_{C^0} + {\left\lVert\nabla u_q\right\rVert}_{C^0}^2 {\left\lVertw_{q+1}\right\rVert}_{C^0} \right).\end{aligned}$$ Using the inductive hypothesis , we have that $$\begin{aligned} {\left\lVertD_{t,q} (\nabla u_q)\right\rVert}_{C_0} &\lesssim {\left\lVertu_q\right\rVert}_{C^1}^2 + {\left\lVertD_{t,q} u_q\right\rVert}_{C^1} \lesssim \lambda_q^4 \delta_q.\end{aligned}$$ From Lemma \[lem:w:q+1:bounds\], we conclude that $$\begin{aligned} {\left\lVertD_{t,q} N_1\right\rVert}_{C^0} &\lesssim \frac{1}{\lambda_{q+1}} \left( \lambda_q^4 \delta_q \delta_{q+1}^{1/2} + \lambda_q^2 \delta_q^{1/2} \tau_{q+1}^{-1} \delta_{q+1}^{1/2} \right) \\ &\lesssim \lambda_{q+1}^{-1} \lambda_q^2 \delta_q^{1/2} \tau_{q+1}^{-1} \delta_{q+1}^{1/2} \\ &= \lambda_q^3\delta_q^{3/4}\delta_{q+1}^{3/4}\\ &= \lambda_0^{-4+\frac{7\beta}{2}}\lambda_{q+1}^2 \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2} \,.\end{aligned}$$ In order to estimate the material derivative of $N_2$, we recall , and as above, using the compact support of $v_q$, we find that $$\begin{aligned} - \lambda_{q+1} D_{t,q} N_2 &= D_{t,q} \BB {\widetilde}P_{\approx \lambda_{q+1}} \left( \nabla (\nabla^\perp \cdot v_q) \sum_{j,k} \Lambda \PP_{q+1,k} \big(\chi_j a_{k,j} c_k(\lambda_{q+1} \Phi_j ) \big) \right)\notag\\ &\quad + D_{t,q} \BB {\widetilde}P_{\approx \lambda_{q+1}} \left( (\nabla^\perp \cdot v_q) \sum_{j,k} \Lambda \PP_{q+1,k} \big(\chi_j \nabla \big( a_{k,j} \psi_{q+1,j,k} \big) c_k(\lambda_{q+1} x) \big) \right) \notag\\ &= \left(\BB {\widetilde}P_{\approx \lambda_{q+1}} D_{t,q} + [D_{t,q} ,\BB {\widetilde}P_{\approx \lambda_{q+1}}] \right) \left( \nabla (\nabla^\perp \cdot v_q) \sum_{j,k} \Lambda \PP_{q+1,k} \big(\chi_j a_{k,j} c_k(\lambda_{q+1} \Phi_j ) \big) \right)\notag\\ &\quad + \left(\BB {\widetilde}P_{\approx \lambda_{q+1}} D_{t,q} + [D_{t,q} ,\BB {\widetilde}P_{\approx \lambda_{q+1}}] \right) \left( (\nabla^\perp \cdot v_q) \sum_{j,k} \Lambda \PP_{q+1,k} \big( \nabla \big(\chi_j a_{k,j} \psi_{q+1,j,k} \big) c_k(\lambda_{q+1} x) \big) \right).\end{aligned}$$ We now appeal to the commutator estimate of Corollary \[cor:Dt:commutator\]: first with $\lambda = \lambda_{q+1}$ and $s=0$ for $[D_{t,q}, \BB {\widetilde}P_{\approx \lambda_{q+1}}]$, second with $\lambda = \lambda_{q+1}$ and $s=1$ for $[D_{t,q}, \Lambda \PP_{q+1,k}]$, and third with $\lambda = \lambda_{q+1}$ and $s=-1$ for $[D_{t,q}, \BB {\widetilde}P_{\approx \lambda_{q+1}}]$. We obtain that $$\begin{aligned} \lambda_{q+1} {\left\lVertD_{t,q} N_2\right\rVert}_{C^0} &\lesssim \sum_{k,j} {\left\lVertD_{t,q} \nabla (\nabla^\perp \cdot v_q)\right\rVert}_{C^0} {\left\lVert\chi_j a_{k,j}\right\rVert}_{C^0} + {\left\lVert\nabla (\nabla^\perp \cdot v_q)\right\rVert}_{C^0} \left( {\left\lVert\chi_j' a_{k,j}\right\rVert}_{C^0} + {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVert\chi_j a_{k,j}\right\rVert}_{C^0} \right) \notag\\ &\quad + {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVert\nabla (\nabla^\perp \cdot v_q)\right\rVert}_{C^0} \sum_{k,j} {\left\lVert\chi_j a_{k,j}\right\rVert}_{C^0} + \sum_{k,j} {\left\lVertD_{t,q} (\nabla^\perp \cdot v_q)\right\rVert}_{C^0} {\left\lVert\chi_j \nabla (a_{k,j} \psi_{q+1,j,k})\right\rVert}_{C^0} \notag\\ &\quad + \sum_{k,j} {\left\lVert\nabla^\perp \cdot v_q\right\rVert}_{C^0} \left( {\left\lVert\chi_j' \nabla (a_{k,j} \psi_{q+1,j,k})\right\rVert}_{C^0} + {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVert\chi_j \nabla(a_{k,j} \psi_{q+1,j,k})\right\rVert}_{C^0} \right) \notag\\ &\quad + {\left\lVert\nabla u_q\right\rVert}_{C^0} \sum_{k,j} {\left\lVert\nabla^\perp v_q\right\rVert}_{C^0} {\left\lVert\chi_j \nabla(a_{k,j} \psi_{q+1,j,k})\right\rVert}_{C^0} \,.\end{aligned}$$ Using the previously established bounds and the inductive estimates yields $$\begin{aligned} {\left\lVertD_{t,q} N_2\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} \lambda_q^4 \delta_q \delta_{q+1}^{1/2} + \lambda_{q+1}^{-1} \lambda_q^2 \delta_q^{1/2} \left( \tau_{q+1}^{-1} \delta_{q+1}^{1/2} + \lambda_{q}^2 \delta_q^{1/2} \delta_{q+1}^{1/2} \right) \notag\\ &\quad + \lambda_{q+1}^{-1} \lambda_q \delta_q^{1/2} \left( \tau_{q+1}^{-1} + \lambda_q^{2} \delta_q^{1/2} \right) \left( \lambda_q \delta_{q+1}^{1/2} + \delta_{q+1}^{1/2} \lambda_{q+1} \tau_{q+1} \lambda_q^2 \delta_{q}^{1/2} \right) \notag \\ &\lesssim \lambda_q^3 \delta_q \delta_{q+1}^{1/2} \,.\end{aligned}$$ Combining the above estimates shows that $$\begin{aligned} {\left\lVertD_{t,q} R_n\right\rVert}\lesssim \lambda_q^3\delta_q^{3/4}\delta_{q+1}^{3/4}+\lambda_q^3 \delta_q \delta_{q+1}^{1/2} \lesssim \lambda_0^{-4+\frac{7\beta}{2}}\lambda_{q+1}^2 \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2}\,.\end{aligned}$$ Therefore holds so long as $\beta< \frac{8}{7}$. Dissipation error ----------------- \[lem:disip:error\] For any $\eps>0$, if $\lambda_0$ and $q$ are sufficiently large then for $R_D$ as defined in , we have that $$\begin{aligned} {\left\lVertR_{D}\right\rVert}_{C^0} &\leq \eps \lambda_{q+2} \delta_{q+2} \\ {\left\lVertD_{t,q} R_D\right\rVert}_{C^0} &\leq \eps \lambda_{q+1}^2 \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2}\end{aligned}$$ ### Amplitude of the dissipation error By definition, $$\begin{aligned} R_D = \BB \Lambda^\gamma w_{q+1} = \BB \Lambda^\gamma {\widetilde}P_{\approx \lambda_{q+1}} w_{q+1}. \label{eq:RD:def}\end{aligned}$$ Therefore, it follows from Bernstein’s inequality for Fourier multipliers [@Le2002] that $$\begin{aligned} {\left\lVertR_D\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{\gamma-1} {\left\lVertw_{q+1}\right\rVert}_{C^0} \notag\\ &\lesssim \lambda_{q+1}^{\gamma-1} \delta_{q+1}^{1/2} \notag\\ &\leq \lambda_{0}^{2\beta -2} \lambda_{q+1}^{\gamma + \beta -2 } \lambda_{q+2} \delta_{q+2}.\end{aligned}$$ Thus we obtain the stated estimate so long as $\gamma < 2-\beta$, $\beta<1$ and $\lambda_0$ is sufficiently large. ### Material derivative of the dissipation error The estimate on the material derivative of the dissipation error follows directly from , the previously established bound for the material derivative of the perturbation, and the commutator estimate of Corollary in which we set $\lambda = \lambda_{q+1}$, and $s= \gamma - 1$, which is the order of the Fourier multiplier operator $\BB \Lambda^\gamma {\widetilde}P_{\approx \lambda_{q+1}}$. We obtain that $$\begin{aligned} {\left\lVertD_{t,q} R_{D}\right\rVert}_{C^0} &\leq {\left\lVert(\BB \Lambda^\gamma {\widetilde}P_{\approx \lambda_{q+1}}) D_{t,q} w_{q+1}\right\rVert}_{C^0} + {\left\lVert \left[ D_{t,q}, \BB \Lambda^\gamma {\widetilde}P_{\approx \lambda_{q+1}}\right]w_{q+1}\right\rVert}_{C^0}\notag\\ &\lesssim \lambda_{q+1}^{\gamma-1} {\left\lVertD_{t,q} w_{q+1}\right\rVert}_{C^0} + \lambda_{q+1}^{\gamma-1} {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVertw_{q+1}\right\rVert}_{C^0} \notag\\ &\lesssim \lambda_{q+1}^{\gamma-1} (\tau_{q+1}^{-1} + \lambda_q^2 \delta_q^{1/2} ) \delta_{q+1}^{1/2} \notag\\ &\lesssim \lambda_q \lambda_{q+1}^\gamma \delta_{q}^{1/4} \delta_{q+1}^{3/4} \notag\\ &= \lambda_0^{\gamma + 7 \beta/2 - 4} \lambda_q^{\gamma+\beta-2}\lambda_{q+1}^{2} \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2}.\end{aligned}$$ Thus again we obtain the stated estimate if $\gamma < 2-\beta$ and $q$ is sufficiently large. Oscillation error {#sec:osc} ----------------- \[lem:osc:error\] For any $\eps>0$, if $\lambda_0$ and $q$ are sufficiently large, then for $R_O$ as defined in , we have that $$\begin{aligned} {\left\lVertR_{O}\right\rVert}_{C^0} &\leq \eps \lambda_{q+2} \delta_{q+2} \,, \\ {\left\lVertD_{t,q} R_O\right\rVert}_{C^0} &\leq \eps \lambda_{q+1}^2 \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2} \,.\end{aligned}$$ ### Decomposition of the oscillation error Recall from that the oscillation error $R_O$ is defined so that the following equality is satisfied: $$\begin{aligned} \div R_O &= \div \mathring R_q + \Lambda w_{q+1} \cdot \nabla w_{q+1} - (\nabla w_{q+1})^T \cdot \Lambda w_{q+1} \notag\\ &= \div \left( \sum_j \chi_j^2 ( \mathring R_{q} - \mathring R_{q,j}) \right) + \div \left( \sum_j \chi_j^2( \mathring R_{q,j} + \rho_j \lambda_{q+1} {\ensuremath{\mathrm{Id}}}) \right) \notag\\ &\qquad + \Big( \Lambda w_{q+1} \cdot \nabla w_{q+1} - (\nabla w_{q+1})^T \cdot \Lambda w_{q+1}\Big) . \label{RO}\end{aligned}$$ Note that in the above formula, as well as throughout this paper, we somewhat abuse notation and write $\sum_{j}$ to mean the summation $\sum_{\{j:\rho_j\neq 0\}} $.[^19] It this then important to note that in view of , the decomposition is valid so long as $$\begin{aligned} \ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_q\right\vert}^2~dx\leq \frac{\lambda_{q+1}\delta_{q+1}}{8}\end{aligned}$$ on the support of $\chi_j$ for $\rho_j=0$. The proof of this fact will be delayed to Lemma \[lem:rho\_diff\], equation . Recalling that $$\begin{aligned} w_{q+1}(x,t) = \sum\limits_{\substack{ \{j:\rho_j\neq 0\} \\ k \in \Omega_{j}}} \PP_{q+1,k} {\widetilde}w_{q+1,j,k} \qquad \mbox{with} \qquad {\widetilde}w_{q+1,j,k} = \chi_{j}(t) a_{k,j}(x,t) b_k( \lambda_{q+1} \Phi_j(x,t) ) \,,\end{aligned}$$ the term $ \Lambda w_{q+1} \cdot \nabla w_{q+1} - (\nabla w_{q+1})^T \cdot \Lambda w_{q+1}$ in has both high and low frequency components, depending whether $k \neq - k'$ or $k=-k'$. We notice that due to the frequency localization induced by $\PP_{q+1,k}$, for $k,k' \in \Omega_j$ with $k+k' \neq 0$, we have that $1/2 \leq |k+k'| \leq 2$, and due to the localization in the angular frequency variable, we obtain that $$\begin{aligned} \Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k} \cdot \nabla \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} = {\widetilde}P_{\approx \lambda_{q+1}} \left( \Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k} \cdot \nabla \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} \right) \label{eq:Lambda:angle:local}\end{aligned}$$ and $$\begin{aligned} (\nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k})^T \otimes \Lambda \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} = {\widetilde}P_{\approx \lambda_{q+1}} \left( (\nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k})^T \otimes \Lambda \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} \right).\end{aligned}$$ We shall thus isolate the high-frequency part of $R_O$ due to the nonlinear interactions in $\Lambda w_{q+1} \cdot \nabla w_{q+1} - (\nabla w_{q+1})^T \cdot \Lambda w_{q+1}$, as $$\begin{aligned} R_{O,{\rm high}} &= \BB {\widetilde}P_{\approx \lambda_{q+1}} \Bigg(\sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big) \cdot \nabla \big(\PP_{q+1,k'} {\widetilde}w_{q+1,j',k'}\big) \Bigg) \notag\\ &\quad - \BB {\widetilde}P_{\approx \lambda_{q+1}} \Bigg(\sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big)^T \cdot \big( \Lambda \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} \big) \Bigg). \label{eq:R:O:high}\end{aligned}$$ Similarly, we need to isolate the low-frequency part of $\Lambda w_{q+1} \cdot \nabla w_{q+1} - (\nabla w_{q+1})^T \cdot \Lambda w_{q+1}$ which occurs when $k+k'=0$. Since $\Omega_1 \cap \Omega_2 = \emptyset$, $k + k' = 0$ implies that $j = j'$, so that upon symmetrizing, we may define the low frequency part of the nonlinear term as $$\begin{aligned} \TTT_{j,k} &= \frac{1}{2}\Bigg( \big(\Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big) \cdot \nabla \big(\PP_{q+1,-k} {\widetilde}w_{q+1,j,-k}\big) +\big(\Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big) \cdot \nabla \big(\PP_{q+1,-k} {\widetilde}w_{q+1,j,-k}\big) \notag \\ &\quad - \big(\nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big)^T \cdot \big( \Lambda \PP_{q+1,-k} {\widetilde}w_{q+1,j,-k} \big) - \big(\nabla \PP_{q+1,-k} {\widetilde}w_{q+1,j,-k}\big)^T \cdot \big( \Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k} \big) \Bigg). \label{eq:TTT:j:k:def}\end{aligned}$$ The challenge now is to obtain the decomposition $$\begin{aligned} {\TTT}_{j,k} = \div ({\mathcal Q}_{j,k}) + \nabla {\mathcal P}_{j,k} \label{eq:R:O:low}\end{aligned}$$ for a suitably defined $2$-tensor ${\mathcal Q}_{j,k}$ which gains one derivative over ${\TTT}_{j,k}$ and obeys good transport estimates, and a scalar function ${\mathcal P}_{j,k}$. This is achieved in Section \[sec:T:j:k:div\], equation below. In turn, the decomposition allows us to write the oscillation stress as $$\begin{aligned} R_O &=\sum_j \chi_j^2 (\mathring R_q - \mathring R_{q,j}) +\left(\sum_j \chi_j^2 \mathring R_{q,j} + \sum_{j,k}\mathring{ \mathcal Q}_{j,k} \right) + R_{O,{\rm high}}\notag\\ &=: R_{O,{\rm approx}} + R_{O,{\rm low}} + R_{O,{\rm high}} \,, \label{eq:RO:decompose}\end{aligned}$$ where we have used the notation $\mathring R_{q,j}$ and $\mathring{ \mathcal Q}_{j,k}$ to denote the traceless parts of $R_{q,j}$ and ${ \mathcal Q}_{j,k}$, respectively. We have also used the fact that $\BB$ already contains the Leray projector, so that it annihilates gradients. ### The definition of ${\mathcal Q}_{j,k}$ {#sec:T:j:k:div} Before explaining how we obtain the $2$-tensor ${\mathcal Q}_{j,k}$ and prior to estimating the three terms in , two technical remarks are in order. First, since $\PP_{q+1,k} b_k(\lambda_{q+1} x) = b_k(\lambda_{q+1} x)$, we may write $$\begin{aligned} \PP_{q+1,k} {\widetilde}w_{q+1,j,k} &= {\widetilde}w_{q+1,j,k} + \chi_{j} \big[\PP_{q+1,k}, a_{k,j} \psi_{q+1,j,k} \big] b_k( \lambda_{q+1} x ) \label{eq:local:w:q+1}\end{aligned}$$ and thus $$\begin{aligned} w_{q+1} &= \sum_{j,k} {\widetilde}w_{q+1,j,k} + \sum_{j,k} \chi_{j} \big[\PP_{q+1,k}, a_{k,j} \psi_{q+1,j,k} \big] b_k( \lambda_{q+1} x ). \label{eq:local:w:q+1:*}\end{aligned}$$ And second, since $\Lambda$ and $\PP_{q+1,k}$ commute, using we may write $$\begin{aligned} \Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k} &= \chi_{j} a_{k,j} \psi_{q+1,j,k} \PP_{q+1,k} \Lambda b_k( \lambda_{q+1} x ) + \chi_{j} [\PP_{q+1,k} \Lambda, a_{k,j} \psi_{q+1,j,k} ] b_k( \lambda_{q+1} x ) \notag \\ &= \lambda_{q+1} {\widetilde}w_{q+1,j,k} + \chi_{j} [\PP_{q+1,k} \Lambda, a_{k,j} \psi_{q+1,j,k} ] b_k( \lambda_{q+1} x ). \label{eq:Lambda:w:q+1}\end{aligned}$$ and thus $$\begin{aligned} \Lambda w_{q+1} = \lambda_{q+1} \sum_{j,k} {\widetilde}w_{q+1} + \sum_{j,k} \chi_{j} [\PP_{q+1,k} \Lambda, a_{k,j} \psi_{q+1,j,k} ] b_k( \lambda_{q+1} x ). \label{eq:Lambda:w:q+1:*}\end{aligned}$$ Let us define the potential vorticity associated to the perturbation $\PP_{q+1,k} {\widetilde}w_{q+1,j,k}$ as $$\begin{aligned} \vartheta_{j,k} = \nabla^{\perp} \cdot \PP_{q+1,k} {\widetilde}w_{q+1,j,k}. \label{eq:vartheta:def}\end{aligned}$$ Using the identity $$\begin{aligned} \Lambda f \cdot \nabla g - \left(\nabla g\right)^T \Lambda f = \Lambda f^\perp (\nabla^\perp \cdot g) = \left( \RSZ \big( \nabla^{\perp}\cdot f\big) \right) (\nabla^{\perp}\cdot g) \label{eq:2D:MAGIC}\end{aligned}$$ which holds for any vector fields $f,g \colon \TT^2 \to \CC^2$ with $\nabla \cdot f = 0$, we may write $\TTT_{j,k}$, as defined in , in the convenient form $$\begin{aligned} \TTT_{j,k}=\frac 12\left((\RSZ \vartheta_{j,k})\vartheta_{j,-k}+\vartheta_{j,k}(\RSZ\vartheta_{j,-k})\right) =: \TTT(\vartheta_{j,k},\vartheta_{j,-k}) \label{eq:TTT:useful}\end{aligned}$$ where as usual $\RSZ = (\RSZ_1, \RSZ_2)$ is the Riesz-transform vector. Our goal next is to rewrite the operator $\TTT$ defined by as a sum of a pressure gradient and a divergence of a $2$-tensor. By an abuse of notation concerning Fourier transforms and Fourier series, we can rewrite $\TTT$ as a bilinear Fourier operator whose $\ell^{\rm th}$ component is given by $$\begin{aligned} 2 \left(\TTT^\ell (f,g)\right)^\wedge (\xi) =& \int_{\RR^2} (\RSZ^\ell f)^\wedge(\xi-\eta){\widehat}g(\eta)~d\eta+ \int_{\RR^2} (\RSZ^\ell g)^\wedge(\eta){\widehat}f(\xi-\eta)~d\eta \\ =&\int_{\RR^2} \frac{i(\xi-\eta)^\ell}{{\left\vert\xi-\eta\right\vert}} {\widehat}f(\xi-\eta) {\widehat}g (\eta)~d\eta+ \int_{\RR^2} \frac{i\eta^\ell}{{\left\vert\eta\right\vert}}{\widehat}g(\eta) {\widehat}f(\xi-\eta)~d\eta \\ =&\int_{\RR^2} \left(\frac{i(\xi-\eta)^\ell}{{\left\vert\xi-\eta\right\vert}} + \frac{i\eta^\ell}{{\left\vert\eta\right\vert}}\right) {\widehat}f(\xi-\eta) {\widehat}g(\eta)~d\eta~.\end{aligned}$$ Rearranging the symbol, we have that $$\begin{aligned} \frac{i(\xi-\eta)^\ell}{{\left\vert\xi-\eta\right\vert}} + \frac{i\eta^\ell}{{\left\vert\eta\right\vert}} =&\frac{i\left((\xi-\eta)^\ell {\left\vert\eta\right\vert} + \eta^\ell {\left\vert\xi-\eta\right\vert} \right)}{{\left\vert\xi-\eta\right\vert} {\left\vert\eta\right\vert}} \notag \\ =&\frac{i\xi^\ell {\left\vert\eta\right\vert}}{{\left\vert\xi-\eta\right\vert} {\left\vert\eta\right\vert}} + \frac{i \eta^\ell \left({\left\vert\xi-\eta\right\vert}-{\left\vert\eta\right\vert}\right) }{{\left\vert\xi-\eta\right\vert} {\left\vert\eta\right\vert}} \notag \\ =&\frac{i\xi^\ell}{{\left\vert\xi-\eta\right\vert}} + \frac{i\eta^\ell}{{\left\vert\xi-\eta\right\vert} {\left\vert\eta\right\vert}} \int_0^1\left(\frac{d}{dr} {\left\vert\eta-r\xi\right\vert}\right)~dr \notag \\ =&\frac{i\xi^\ell}{{\left\vert\xi-\eta\right\vert}} - \frac{i\eta^\ell}{{\left\vert\xi-\eta\right\vert} {\left\vert\eta\right\vert}} \xi^m \int_0^1 \frac{\left(\eta-r\xi\right)^m} {{\left\vert\eta-r\xi\right\vert}}~dr \notag \\ =&\left(i\xi^\ell\right) \frac{1}{{\left\vert\xi-\eta\right\vert} } + \left(i\xi^m\right) \frac{i\eta^\ell}{{\left\vert\eta\right\vert}} \frac{1}{{\left\vert\xi-\eta\right\vert} } s^m(\xi-\eta,\eta) \end{aligned}$$ where we define the symbol $s \colon \RR^2 \times \RR^2 \to \CC$ by $$\begin{aligned} s^{m}(\zeta,\eta) = \int_0^1 \frac{i\left((1-r)\eta-r\zeta\right)^m} {{\left\vert(1-r)\eta-r\zeta\right\vert}}~dr. \label{eq:sm:symbol:def}\end{aligned}$$ An important property of the symbol $s^m$ (which will later be essential to our proof) is that $$\begin{aligned} s^m(-\eta,\eta) = \frac{i \eta^m}{|\eta|} \label{eq:sm:property}\end{aligned}$$ which is the symbol of the Riesz transform $\RSZ^m$. As a result of the above computations, may write $$\label{eq:T_symb} \begin{split} \left( \TTT^\ell(f,g)\right)^\wedge (\xi) =& \frac{i\xi^\ell}{2} \int_{\RR^2} \frac{{\widehat}f(\xi-\eta)}{{\left\vert\xi-\eta\right\vert}} {\widehat}{g}(\eta)~d\eta + \frac{i\xi^m}{2} \int_{\RR^2} s^m(\xi-\eta,\eta) \frac{{\widehat}f(\xi-\eta)}{{\left\vert\xi-\eta\right\vert}} \frac{i\eta^l}{{\left\vert\eta\right\vert}} {\widehat}g(\eta)~d\eta. \end{split}$$ Upon defining the bilinear pseudo-product operator $\SSS^m$ in Fourier space as $$\left(\SSS^m (f,g)\right)^\wedge(\xi):= \int_{\RR^2} s^m(\xi-\eta, \eta) {\widehat}f(\xi-\eta){\widehat}g(\eta) ~d\eta~, \label{eq:SSS:def}$$ we then obtain from the following formula for $\mathcal T$: $$\begin{aligned} \TTT^\ell(f,g) = \frac 12 \partial_\ell (\Lambda^{-1}f g)+ \frac 12 \partial_m (\SSS^m(\Lambda^{-1}f,\RSZ^\ell g))~. \label{eq:T:grad:div}\end{aligned}$$ The representation of the bilinear operator $\SSS^m$ is not very convenient to estimate; instead, we compute the inverse Fourier transform with respect to $\xi$ and rewrite $\SSS^m$ as $$\begin{aligned} \SSS^m(f,g)(x) := \frac{1}{(2\pi)^2}\intint_{\RR^2\times \RR^2} s^m(\zeta,\eta) {\widehat}{f}(\zeta) {\widehat}{g}(\eta) e^{i x \cdot (\zeta + \eta)} d\zeta \, d\eta\, , \label{eq:SSS:def:real}\end{aligned}$$ and upon further computing the inverse Fourier transform with respect to $(\eta,\zeta) \in \RR^4$, we rewrite $\SSS^m$ as follows: $$\begin{aligned} \SSS^m(f,g)(x) = \intint_{\RR^2\times \RR^2} K_{s^m}(x-y,x-z) f(y) g(z) dy dz \,, \label{eq:SSS:def:conv}\end{aligned}$$ where $K_{s^m}$ is the inverse Fourier transform in $\RR^4$ of $s^m$. Thus, another way to view $\SSS^m$ is as a bilinear convolution operator. We refer to [@CoMe1978; @GrTo2002; @MuSh2013] and Appendix \[app:pseudo:product\] for further properties of pseudo-product operators of the type , and for the equivalence of the definitions –. In view of and , we have now defined the tensor ${\mathcal Q}_{j,k}$ and the scalar ${\mathcal P}_{j,k}$ in , namely $$\begin{aligned} \TTT_{j,k} = \frac 12 \nabla \left(\Lambda^{-1} \vartheta_{j,k} \, \vartheta_{j,-k} \right) + \frac 12 \div \left( \SSS\Big( \Lambda^{-1} \vartheta_{j,k}, \RSZ \vartheta_{j,-k} \Big) \right),\end{aligned}$$ so that $$\begin{aligned} \left({\mathcal Q}_{j,k}\right)^{m \ell} &= \frac 12 \SSS^m\left( \Lambda^{-1} \vartheta_{j,k}, \RSZ^\ell \vartheta_{j,-k} \right) \,, \label{eq:Q:j:k:def} \\ {\mathcal P}_{j,k} &= \Lambda^{-1} \vartheta_{j,k} \vartheta_{j,-k} \,, \notag\end{aligned}$$ with the bilinear pseudo-product operator $\SSS^m$ being defined by . ### Canceling the principal part of the $R_{O,{\rm low}}$ stress Before estimating $R_{O,{\rm low}}$, we need to extract the leading order term in the matrices ${\mathcal Q}_{j,k}$ defined by . For this purpose recall cf.  and that $$\begin{aligned} \vartheta_{j,k} &= \nabla^\perp \cdot \left( P_{\approx k\lambda_{q+1} } {\widetilde}w_{q+1,j,k} + (\RSZ \otimes \RSZ) P_{\approx k \lambda_{q+1} } {\widetilde}w_{q+1,j,k} \right) \notag\\ &= P_{\approx k\lambda_{q+1} } \left(\nabla^\perp \cdot {\widetilde}w_{q+1,j,k} \right).\end{aligned}$$ Here ${\widehat}K_{\approx k \lambda_{q+1}}(\xi) = {\widehat}K_{\approx 1}(\xi/\lambda_{q+1} - k)$ is the Fourier symbol of $P_{\approx k \lambda_{q+1}}$. Using the precise definition of ${\widetilde}w_{q+1,j,k}$ in , the definition of $b_k$ and $c_k$ in , and the notation , we obtain $$\begin{aligned} \Lambda^{-1} \vartheta_{j,k} &= \Lambda^{-1} P_{\approx k \lambda_{q+1}} \left(\nabla^\perp \cdot {\widetilde}w_{q+1,j,k} \right) = P_{\approx k \lambda_{q+1}} \RSZ^\perp \cdot {\widetilde}w_{q+1,j,k} \notag\\ &= \chi_j (i k^\perp) \cdot \RSZ^\perp P_{\approx k \lambda_{q+1}} \Big( a_{k,j} \psi_{q+1,j,k} c_k(\lambda_{q+1}x ) \Big) \label{eq:Lambda:-1:vartheta}\end{aligned}$$ and $$\begin{aligned} \RSZ^\ell \vartheta_{j,-k} &= \RSZ^\ell P_{\approx - k \lambda_{q+1}} \left(\nabla^\perp \cdot {\widetilde}w_{q+1,j,-k} \right) \notag\\ &= - \chi_j (i k^\perp) \cdot \nabla^\perp \RSZ^\ell P_{\approx - k \lambda_{q+1}} \Big( a_{k,j} \psi_{q+1,j,-k} c_{-k}(\lambda_{q+1}x ) \Big) . \label{eq:Riesz:vartheta}\end{aligned}$$ We note that since multiplication by $c_k(\lambda_{q+1} x)$ results in a shift by $\lambda_{q+1} k$ in frequency, the Fourier analogues of and are $$\begin{aligned} \left(\Lambda^{-1} \vartheta_{j,k} \right)^\wedge(\xi) &= - \chi_j(t) \frac{k \cdot \xi}{|\xi|} {\widehat}K_{\approx 1}\left(\frac{\xi}{\lambda_{q+1}} - k\right) \left( a_{k,j} \psi_{q+1,j,k}\right)^\wedge (\xi - k\lambda_{q+1}) \label{eq:Lambda:-1:vartheta:F}\end{aligned}$$ and $$\begin{aligned} \left(\RSZ^\ell \vartheta_{j,-k} \right)^\wedge(\xi) &= \chi_j(t) i \xi^\ell \frac{( k \cdot \xi)}{|\xi|} K_{\approx 1}\left(\frac{\xi}{\lambda_{q+1}} + k\right)\left(a_{k,j} \psi_{q+1,j,-k}\right)^\wedge(\xi+ k \lambda_{q+1}). \label{eq:Riesz:vartheta:F}\end{aligned}$$ Inserting – in formula , and recalling –, we obtain that $$\begin{aligned} {\mathcal Q}_{j,k}^{m\ell}(x) &= \frac{\chi_j^2(t)}{2 (2 \pi)^2} \intint_{\RR^2\times\RR^2} (- i s^m(\zeta,\eta)) \frac{k \cdot \zeta}{|\zeta|} {\widehat}K_{\approx 1}\left(\frac{\zeta - k \lambda_{q+1}}{\lambda_{q+1}} \right) \left( a_{k,j} \psi_{q+1,j,k}\right)^\wedge (\zeta - k\lambda_{q+1}) \notag\\ &\qquad \qquad \quad \times \eta^\ell \frac{ k \cdot \eta}{|\eta|} {\widehat}K_{\approx 1}\left(\frac{\eta + k\lambda_{q+1}}{\lambda_{q+1}} \right)\left(a_{k,j} \psi_{q+1,j,-k}\right)^\wedge(\eta+ k \lambda_{q+1}) e^{i x \cdot(\zeta + \eta)} d\zeta d\eta \notag\\ &= \frac{\chi_j^2(t)}{2 (2 \pi)^2} \intint_{\RR^2\times\RR^2} (- i s^m(\zeta + k \lambda_{q+1},\eta - k \lambda_{q+1})) \frac{k \cdot (\zeta + k \lambda_{q+1})}{|\zeta + k \lambda_{q+1}|} {\widehat}K_{\approx 1}\left(\frac{\zeta}{\lambda_{q+1}} \right) \left( a_{k,j} \psi_{q+1,j,k}\right)^\wedge (\zeta) \notag\\ &\qquad \qquad \quad \times (\eta^\ell - k^\ell \lambda_{q+1}) \frac{ k \cdot (\eta - k \lambda_{q+1}) }{|\eta - k \lambda_{q+1}|} {\widehat}K_{\approx 1}\left(\frac{\eta}{\lambda_{q+1}} \right)\left(a_{k,j} \psi_{q+1,j,-k}\right)^\wedge(\eta) e^{i x \cdot(\zeta + \eta)} d\zeta d\eta \notag\\ &= \frac{\chi_j^2}{2} \frac{1}{(2\pi)^2} \int\int_{\RR^2\times \RR^2} M_k^{m\ell}(\zeta,\eta) \left( a_{k,j} \psi_{q+1,j,k}\right)^\wedge (\zeta) \left(a_{k,j} \psi_{q+1,j,-k}\right)^\wedge(\eta) e^{i x \cdot(\zeta + \eta)} d\zeta d\eta \label{eq:Q:j:k:var1}\end{aligned}$$ where in the second to last line we have used the change of variables in $\zeta$ and $\eta$ by shifting with $\pm k \lambda_{q+1}$, and in the last line we have denoted $$\begin{aligned} &M_k^{m\ell}(\zeta,\eta) \notag\\ &= - i s^m(\zeta + k \lambda_{q+1},\eta - k \lambda_{q+1}) \frac{k \cdot (\zeta + k \lambda_{q+1})}{|\zeta + k \lambda_{q+1}|} {\widehat}K_{\approx 1}\left(\frac{\zeta}{\lambda_{q+1}} \right) (\eta^\ell - k^\ell \lambda_{q+1}) \frac{ k \cdot (\eta - k \lambda_{q+1}) }{|\eta - k \lambda_{q+1}|} {\widehat}K_{\approx 1}\left(\frac{\eta}{\lambda_{q+1}} \right) \notag\\ &=: \int_0^1 M^{m\ell}_{k,r}(\zeta,\eta) dr. \label{eq:Q:j:k:var2}\end{aligned}$$ and $$\begin{aligned} M^{m\ell}_{k,r}(\zeta,\eta) &= \frac{\big( (1-r) \eta - r \zeta - k \lambda_{q+1} \big)^m}{| (1-r) \eta - r \zeta - k \lambda_{q+1}|} (\eta^\ell - k^\ell \lambda_{q+1}) \notag\\ &\qquad \qquad \times \frac{k \cdot (\zeta + k \lambda_{q+1})}{|\zeta + k \lambda_{q+1}|} \frac{ k \cdot (\eta - k \lambda_{q+1}) }{|\eta - k \lambda_{q+1}|} {\widehat}K_{\approx 1}\left(\frac{\zeta}{\lambda_{q+1}} \right) {\widehat}K_{\approx 1}\left(\frac{\eta}{\lambda_{q+1}} \right) . \label{eq:Q:j:k:var22}\end{aligned}$$ We observe here that the multiplier $M_{kr}^{m\ell}$ defined in has two important features: the first concerns smoothness and will allow us to establish bounds on the induced bilinear pseudo-product operator, while the second concerns structure, and allows us to define the principal term in ${\mathcal Q}_{j,k}^{m\ell}$ and cancel the leading order term in the oscillation stress $R_{O,{\rm low}}$. First, we note that by , we have that $$\begin{aligned} M_{k,r}^{m\ell}(\zeta,\eta) = \lambda_{q+1} (M_{k,r}^*)^{m\ell}\left(\frac{\zeta}{\lambda_{q+1}}, \frac{\eta}{\lambda_{q+1}}\right) \label{eq:M:k:r:rescale}\end{aligned}$$ where $$\begin{aligned} (M_{k,r}^*)^{m\ell}(\xi_1, \xi_2) = \frac{\big( (1-r) \xi_2 - r \xi_1 - k \big)^m}{| (1-r) \xi_2 - r \xi_1 -k |} (\xi_2^\ell - k^\ell) \frac{k \cdot (\xi_1 + k )}{|\xi_1 + k|} \frac{ k \cdot (\xi_2 - k ) }{|\xi_2 - k|} {\widehat}K_{\approx 1}\left( \xi_1 \right) {\widehat}K_{\approx 1}\left(\xi_2 \right) \label{eq:M:k:r:*:def}\end{aligned}$$ for $\xi_1,\xi_2\in\RR^2$. We notice here that $M_{k,r}^*$ is independent of $\lambda_{q+1}$, and that by the definition of ${\widehat}K_{\approx 1}$, the multiplier $M_{k,r}^*$ is supported on $(\xi_1,\xi_2) \in B_{1/8}(0) \times B_{1/8}(0)$. The latter property ensures that $|\xi_1+ k|\geq 1/2$, $|\xi_2-k| \geq 1/2$, and $|k+ (1-r)\xi_1 - r \xi_2|\geq 1/8$. This ensures that the multiplier $M_{k,r}^*$ is infinitely many times differentiable, with bounds that are uniform in $r \in (0,1)$. Second, we note that from it follows that $$\begin{aligned} M_{k,r}^{m\ell}(0,0) = - \lambda_{q+1} k^m k^\ell \label{eq:MAGIC:*}\end{aligned}$$ whenever $k \in \SS^1$. Moreover, by the definition of the inverse Fourier transform, we have that $$\begin{aligned} &\frac{1}{(2\pi)^2} \int\int_{\RR^2\times \RR^2} \left( a_{k,j} \psi_{q+1,j,k}\right)^\wedge (\zeta) \left(a_{k,j} \psi_{q+1,j,-k}\right)^\wedge(\eta) e^{i x \cdot(\zeta + \eta)} d\zeta d\eta \notag\\ &\qquad = a_{k,j}(x) \psi_{q+1,j,k}(x) a_{k,j}(x) \psi_{q+1,j,-k}(x) \notag\\ &\qquad = a_{k,j}^2(x) \,, \label{eq:Q:j:k:var3}\end{aligned}$$ since by , $\psi_{q+1,j,k}(x) \psi_{q+1,j,-k}(x) = 1$. Therefore, combining –, we decompose ${\mathcal Q}_{j,k}^{m\ell}$ as a principal and commutator term: $$\begin{aligned} {\mathcal Q}_{j,k}^{m\ell} &= - \frac{\lambda_{q+1}}{2} \chi_j^2 (k \otimes k)^{ m \ell} a_{k,j}^2 + {\widetilde}{\mathcal Q}_{j,k}^{m\ell} \notag \\ &= \frac{\lambda_{q+1}}{2} \chi_j^2 \left( k^{\perp}\otimes k^{\perp} - {\ensuremath{\mathrm{Id}}}\right)^{ml} a_{k,j}^2 + {\widetilde}{\mathcal Q}_{j,k}^{m\ell}. \label{eq:Q:j:k:decompose}\end{aligned}$$ where,t since $|k|=1$, ${\ensuremath{\mathrm{Tr}}}(k^\perp \otimes k^\perp) = 1$, and where $$\begin{aligned} {\widetilde}{\mathcal Q}_{j,k}^{m\ell}(x) &= \frac{\chi_j^2}{2(2\pi)^2} \int_0^1 \intint_{\RR^2\times\RR^2} \left( M_{k,r}^{m\ell}(\zeta,\eta) - M_{k,r}^{m\ell}(0,0) \right) \notag\\ &\qquad \qquad \qquad \qquad \times \left( a_{k,j} \psi_{q+1,j,k}\right)^\wedge(\zeta) \left(a_{k,j} \psi_{q+1,j,-k}\right)^\wedge(\eta) e^{i x \cdot(\zeta + \eta)} d\zeta d\eta dr.\end{aligned}$$ Next, we use the mean value theorem together with , and find that $$\begin{aligned} {\widetilde}{\mathcal Q}_{j,k}^{m\ell}(x) &= \frac{\chi_j^2}{2(2\pi)^2} \int_0^1 \!\! \int_0^1 \intint_{\RR^2\times\RR^2}\left( (\zeta \cdot \nabla_\zeta + \eta \cdot \nabla_\eta) M_{k,r}^{m\ell}\right) (\bar r \zeta,\bar r \eta) \notag\\ &\qquad \qquad \qquad \qquad \times \left( a_{k,j} \psi_{q+1,j,k}\right)^{\wedge}(\zeta) \left(a_{k,j} \psi_{q+1,j,-k}\right)^\wedge(\eta) e^{i x \cdot(\zeta + \eta)} d\zeta d\eta d\bar r dr \notag\\ &= \frac{ \chi_j^2}{2(2\pi)^2}\int_0^1 \!\! \int_0^1\intint_{\RR^2\times\RR^2} \left( - i \nabla_{\xi_1} (M_{k,r}^*)^{m\ell}\right) \left(\frac{\bar r \zeta}{\lambda_{q+1}}, \frac{\bar r \eta}{\lambda_{q+1}}\right) \cdot \left(\nabla (a_{k,j} \psi_{q+1,j,k})\right)^\wedge(\zeta) \notag\\ &\qquad \qquad \qquad \qquad \times \left(a_{k,j} \psi_{q+1,j,-k}\right)^\wedge(\eta) e^{i x \cdot(\zeta + \eta)} d\zeta d\eta d\bar r dr \notag\\ &\quad + \frac{ \chi_j^2}{2(2\pi)^2} \int_0^1 \!\! \int_0^1 \intint_{\RR^2\times\RR^2}\left(-i \nabla_{\xi_2} (M_{k,r}^*)^{m\ell}\right) \left(\frac{\bar r \zeta}{\lambda_{q+1}}, \frac{\bar r \eta}{\lambda_{q+1}}\right) \cdot \left(\nabla(a_{k,j} \psi_{q+1,j,-k})\right)^\wedge(\eta) \notag\\ &\qquad \qquad \qquad \qquad \times \left( a_{k,j} \psi_{q+1,j,k}\right)^\wedge(\zeta) e^{i x \cdot(\zeta + \eta)} d\zeta d\eta d\bar r dr \notag\\ &= \left({\widetilde}{\mathcal Q}_{j,k}^{(1)}\right)^{m\ell}(x) + \left({\widetilde}{\mathcal Q}_{j,k}^{(2)}\right)^{m\ell}(x). \label{eq:Q:j:k:decompose:nasty}\end{aligned}$$ Note that ${\widetilde}Q^{(1)}_{j,k}$ and ${\widetilde}Q^{(2)}_{j,k}$ are both bilinear pseudo-product operators, and thus similarly to the equivalence between – we may take the inverse Fourier transform of with respect to the variable $(\zeta,\eta) \in \RR^4$. For $(z_1,z_2) \in \RR^2 \times \RR^2$, we denote the inverse Fourier transforms of the above vectors of multipliers as \[eq:multilinear:kernels\] $$\begin{aligned} ({\mathcal K}_{k,r,\bar r}^{(1)})^{m\ell} (z_1,z_2) &= \frac{\lambda_{q+1}^4}{\bar r^4} \left( - i \nabla_{\xi_1} (M_{k,r}^*)^{m\ell}\right)^{\vee}\left(\frac{\lambda_{q+1} z_1}{\bar r},\frac{\lambda_{q+1} z_2 }{\bar r} \right) , \\ ({\mathcal K}_{k,r,\bar r}^{(2)})^{m\ell} (z_1,z_2) &= \frac{\lambda_{q+1}^4}{\bar r^4} \left( - i \nabla_{\xi_2} (M_{k,r}^*)^{m\ell}\right)^{\vee}\left(\frac{\lambda_{q+1} z_1}{\bar r},\frac{\lambda_{q+1} z_2 }{\bar r} \right).\end{aligned}$$ It follows from basic scaling properties of the Fourier transform that $$\begin{aligned} \left({\widetilde}{\mathcal Q}_{j,k}^{(1)}\right)^{m\ell}(x) &= \frac{\chi_j^2}{2} \int_0^1\!\!\int_0^1 \!\! \intint_{\RR^2\times\RR^2} ({\mathcal K}_{k,r,\bar r}^{(1)})^{m\ell} (x-z_1 , x-z_2) \cdot \nabla (a_{k,j} \psi_{q+1,j,k}) (z_1) \left(a_{k,j} \psi_{q+1,j,-k}\right)(z_2) dz_1 dz_2 d\bar r dr \notag\\ &=: \chi_j^2 {\widetilde}\SSS_k^{(1),m\ell}\big (\nabla (a_{k,j} \psi_{q+1,j,k}), a_{k,j} \psi_{q+1,j,-k} \big) \label{eq:Q:j:k:1:convolution} \\ \left({\widetilde}{\mathcal Q}_{j,k}^{(2)}\right)^{m\ell}(x) &= \frac{\chi_j^2}{2} \int_0^1\!\!\int_0^1 \!\! \intint_{\RR^2\times\RR^2} ({\mathcal K}_{k,r,\bar r}^{(2)})^{m\ell} (x-z_1 , x-z_2) \cdot \nabla (a_{k,j} \psi_{q+1,j,-k}) (z_2) \left(a_{k,j} \psi_{q+1,j,k}\right)(z_1) dz_1 dz_2 d\bar r dr \notag\\ &=: \chi_j^2 {\widetilde}\SSS_k^{(2),m\ell}\big (a_{k,j} \psi_{q+1,j,k} , \nabla (a_{k,j} \psi_{q+1,j,-k})\big) \label{eq:Q:j:k:2:convolution}\end{aligned}$$ Here as usual we have identified the $\TT^2$-periodic functions of $z_1$ and $z_2$ with their periodic extensions to all of $\RR^2$. The precise form of the above kernels appearing in – is not important. The only important property of these kernels, which we will use repeatedly when bounding these bilinear convolution operators, is that for $i \in \{1,2\}$, with the notation $z = (z_1,z_2) \in \RR^2\times \RR^2$, we have that $$\begin{aligned} {\left\lVert z^a \nabla_z^b ({\mathcal K}_{k,r,\bar r}^{(i)})^{m\ell}\right\rVert}_{L^1_{z_1,z_2}(\RR^2 \times\RR^2)} &\leq C_{a,b} \left(\frac{\lambda_{q+1}}{\bar r}\right)^{|b|-|a|} \label{eq:K:k:r:bounds}\end{aligned}$$ uniformly for $r\in (0,1)$, and for all $0\leq |a|,|b|\leq 1$. The bound follows upon rescaling from the fact that the multiplier $M_{k,r}^*$ defined in is in $C_0^\infty( B_{1/8}(0) \times B_{1/8}(0))$, and thus so are $\partial_{\xi_1}M_{k,r}^*$ and $\partial_{\xi_2} M_{k,r}^*$. In summary, the decomposition –, allows us to split the low frequency part of the oscillation error, defined by , as $$\begin{aligned} R_{O,{\rm low}} = \mathring O_1 + \mathring O_2\end{aligned}$$ where the principal term $\mathring O_1$ is the traceless part of $$\begin{aligned} O_1 = \sum_j \chi_j^2 \mathring R_{q,j} + \frac{\lambda_{q+1}}{2} \sum_{j,k} \left( k^{\perp}\otimes k^{\perp} - {\ensuremath{\mathrm{Id}}}\right) \chi_j^2 a_{k,j}^2 \label{eq:O1:def}\end{aligned}$$ while $\mathring O_2$ is the traceless part of the commutator terms, given by $$\begin{aligned} O_2 = \sum_{j,k} {\widetilde}{\mathcal Q}_{j,k}^{(1)} + {\widetilde}{\mathcal Q}_{j,k}^{(2)} = O_{21} + O_{22}. \label{eq:O2:def}\end{aligned}$$ The key observation here is that the $\mathring O_1$ term vanishes. Indeed, by the definition of the $a_k$ in , and of the functions $\gamma_k$ in , we have $$\begin{aligned} \frac{\mathring R_{q,j}}{\lambda_{q+1}} + \frac 12 \sum_{k \in \Omega_j} a_{k,j}^2 ( k^\perp \otimes k^\perp )= \rho_j {\ensuremath{\mathrm{Id}}}\,,\end{aligned}$$ which shows that $\mathring O_1=0$, since the traceless part of a multiple of the identity is the zero matrix. Therefore, we may summarize our computations in this section as $$\begin{aligned} R_{O,{\rm low}} = \mathring O_{21} + \mathring O_{22} \label{eq:RO:low:final}\end{aligned}$$ with $O_{21}$ and $O_{22}$ as defined by –. ### Amplitude of the $R_{O,{\rm approx}}$ stress In order to bound $R_{O,{\rm approx}}$, we recall cf.  that $$\begin{aligned} \left( \mathring R_q - \mathring R_{q,j}\right)(x,j \tau_{q+1}) = 0\end{aligned}$$ where $j \tau_{q+1}$ is the center of the time-support of $\chi_{j}$, and moreover $$\begin{aligned} D_{t,q} \left( \mathring R_q - \mathring R_{q,j}\right) = D_{t,q} \mathring R_q. \label{eq:material:RO:approx}\end{aligned}$$ We may thus appeal to the inductive assumption and the transport estimate  to find that $$\begin{aligned} {\left\lVert \mathring R_q - \mathring R_{q,j}\right\rVert}_{C^0(\supp \chi_j)} &\lesssim \tau_{q+1} \lambda_{q}^2 \delta_{q}^{1/2} \lambda_{q+1} \delta_{q+1} \notag\\ &\lesssim \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4} \,. \label{eq:RO:approx:amplitude}\end{aligned}$$ Upon summing over $j$, we arrive at $$\begin{aligned} \label{eq:R_O_approx_est} {\left\lVertR_{O,{\rm approx}}\right\rVert}_{C^0} \leq \sum_j \chi_j^2 {\left\lVert \mathring R_q - \mathring R_{q,j}\right\rVert}_{C^0(\supp \chi_j)} \leq \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4}\,.\end{aligned}$$ ### Amplitude of the $R_{O,{\rm low}}$ stress The map from matrices to their traceless part is clearly bounded and thus in view of , we need to estimate $O_{21}$ and $O_{22}$. We only show the estimate for $O_{21}$, since the one for $O_{22}$ is identical, upon changing $(1)$ with $(2)$ below. For this purpose we use – and the kernel estimate with $|a|=|b|=0$ to conclude that $$\begin{aligned} {\left\lVertO_{21}\right\rVert}_{C^0} &\lesssim \sum_{j,k} \chi_j^2 {\left\lVert{\widetilde}\SSS_k^{(1),m\ell}\big (\nabla (a_{k,j} \psi_{q+1,j,k}), a_{k,j} \psi_{q+1,j,-k} \big)\right\rVert}_{C^0} \notag\\ &\lesssim \sum_{j,k} \chi_j^2 {\left\lVert\nabla (a_{k,j} \psi_{q+1,k,j})\right\rVert}_{C^0} {\left\lVerta_{k,j} \psi_{q+1,j,-k}\right\rVert}_{C^0} \sup_{r,\bar r\in(0,1)}{\left\lVert ({\mathcal K}_{k,r,\bar r}^{(1)} )^{m\ell}\right\rVert}_{L^1(\RR^2 \times\RR^2)} \notag\\ &\lesssim \delta_{q+1}^{1/2} \left(\lambda_q \delta_{q+1}^{1/2} + \delta_{q+1}^{1/2} \tau_{q+1} \lambda_{q+1} \lambda_q^2 \delta_{q}^{1/2} \right) \notag\\ &\lesssim \delta_{q+1} \tau_{q+1} \lambda_{q+1} \lambda_q^2 \delta_{q}^{1/2} \notag\\ &= \lambda_{q} \delta_q^{1/4} \delta_{q+1}^{3/4} \,. \notag\end{aligned}$$ Therefore $$\label{eq:R_O_low} {\left\lVertR_{O,{\rm low}}\right\rVert}_{C^0} \leq \lambda_{q} \delta_q^{1/4} \delta_{q+1}^{3/4}\,.$$ ### Amplitude of the $R_{O,{\rm high}}$ stress We recall cf.  that \[eq:RO:high:decompose\] $$\begin{aligned} R_{O,{\rm high}} &= O_3 - O_4 \\ O_3 &=\BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big) \cdot \nabla \big(\PP_{q+1,k'} {\widetilde}w_{q+1,j',k'}\big) \\ O_4 &= \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big)^T \cdot \big( \Lambda \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} \big) .\end{aligned}$$ [**The $O_3$ estimate.**]{} Appealing to and , we have that $$\begin{aligned} O_3 &=\BB {\widetilde}P_{\approx \lambda_{q+1}} \lambda_{q+1}\sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \div \left( \big({\widetilde}w_{q+1,j,k}\big) \otimes \big( {\widetilde}w_{q+1,j',k'}\right) \big) \notag\\ &\quad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \div \Bigg(\sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\lambda_{q+1} {\widetilde}w_{q+1,j,k}\big) \otimes \big( \chi_{j'} \big[\PP_{q+1,k'}, a_{k',j'} \psi_{q+1,j',k'} \big] b_{k'}( \lambda_{q+1} x )\big) \Bigg) \notag\\ &\quad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \div \Bigg(\sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\chi_{j} [\PP_{q+1,k} \Lambda, a_{k,j} \psi_{q+1,j,k} ] b_k( \lambda_{q+1} x )\big) \otimes \big(\PP_{q+1,k'} {\widetilde}w_{q+1,j',k'}\big) \Bigg) \notag\\ &=: O_{31} + O_{32} + O_{33}. \label{eq:O3:decompose}\end{aligned}$$ For $O_{31}$, we need to compute carefully the divergence before estimating it. From , $$\begin{aligned} O_{31} =& \lambda_{q+1} \BB {\widetilde}P_{\approx \lambda_{q+1}} \Bigg(\sum_{ \underset{k+k'\neq 0 } {j,j',k,k'} } \chi_{j}\chi_{j'} \left(b_{k'}( \lambda_{q+1} x) \otimes b_{k}( \lambda_{q+1} x)-\frac12 b_{k'}( \lambda_{q+1} x) \cdot b_{k}( \lambda_{q+1} x){\ensuremath{\mathrm{Id}}}\right) \notag \\ &\qquad\qquad \qquad \qquad \times \nabla\left( a_{k,j} \psi_{q+1,j,k} a_{k'} \psi_{q+1,j',k'}\right) \Bigg)\end{aligned}$$ We thus obtain from Lemma \[lem:D:N:a:psi\] that $$\begin{aligned} \|O_{31}\|_{C^0} &\lesssim \sum_{j,j',k,k'} {\left\lVert \chi_{j}\chi_{j'} \nabla\left( a_{k,j} \psi_{q+1,j,k} a_{k'} \psi_{q+1,j',k'}\right)\right\rVert}_{C^0} \notag\\ &\lesssim \lambda_q \delta_{q+1} + \delta_{q+1} \tau_{q+1} \lambda_{q+1} \lambda_q^2 \delta_{q}^{1/2} \notag\\ &\lesssim \lambda_q \delta_{q}^{1/4}\delta_{q+1}^{3/4}\,. \label{eq:sharp:2}\end{aligned}$$ Using the commutator estimate (with $s=0$ and $\lambda = \lambda_{q+1}$; and respectively with $s=1$ and $\lambda = \lambda_{q+1}$) and Lemma \[lem:D:N:a:psi\], we have $$\begin{aligned} {\left\lVertO_{32}\right\rVert}_{C^0} + {\left\lVertO_{33}\right\rVert}_{C^0} &\lesssim \sum_{j,j',k,k'} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} {\left\lVert \nabla( a_{k',j'} \psi_{q+1,j',k'})\right\rVert}_{C^0(\supp(\chi_{j'})} \notag \\ &\lesssim \delta_{q+1}^{1/2} \left( \lambda_q \delta_{q+1}^{1/2} + \delta_{q+1}^{1/2} \tau_{q+1} \lambda_{q+1} \lambda_q^2 \delta_{q}^{1/2} \right) \notag\\ &\lesssim \lambda_q \delta_{q}^{1/4}\delta_{q+1}^{3/4}\,.\end{aligned}$$ Combining the above estimates yields $$\begin{aligned} \label{eq:O_3_bound} {\left\lVertO_3\right\rVert}_{C^0} \leq \lambda_q \delta_{q}^{1/4}\delta_{q+1}^{3/4}.\end{aligned}$$ [**The $O_4$ estimate.**]{} In order to bound the $O_4$ part of the oscillation error, we note that $$\begin{aligned} \BB \left( (\nabla {\widetilde}w_{q+1,j,k})^T \cdot {\widetilde}w_{q+1,j',k'} + (\nabla {\widetilde}w_{q+1,j',k'})^T \cdot {\widetilde}w_{q+1,j,k} \right) = \frac{1}{2} \BB \left( \nabla \left( {\widetilde}w_{q+1,j,k} \cdot {\widetilde}w_{q+1,j',k'} \right) \right)= 0\end{aligned}$$ since $\BB$ contains the Leray projector. Therefore, using and we obtain $$\begin{aligned} O_4 &= \BB {\widetilde}P_{\approx \lambda_{q+1}} \Bigg(\sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \left( \chi_j \nabla \big[ \PP_{q+1,k}, a_{k,j} \psi_{q+1,j,k}] b_k(\lambda_{q+1} x) \right)^T \cdot \left(\lambda_{q+1} {\widetilde}w_{q+1,j',k'} \right) \Bigg) \notag\\ &\quad+ \BB {\widetilde}P_{\approx \lambda_{q+1}} \Bigg(\sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \left( \nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\right)^T \cdot \left( \chi_{j'} \left( \big[ \PP_{q+1,k'} \Lambda, a_{k',j'} \psi_{q+1,j',k'} \big] b_{k'}(\lambda_{q+1} x) \right) \right) \Bigg) \notag\\ &= O_{41} + O_{42}. \label{eq:O4:decompose}\end{aligned}$$ Appealing to the formula $$\begin{aligned} \nabla \big[ \PP_{q+1,k}, a_{k,j} \psi_{q+1,j,k}] b_k(\lambda_{q+1} x) = \big[ \PP_{q+1,k}, \nabla(a_{k,j} \psi_{q+1,j,k})] b_k(\lambda_{q+1} x) + \big[ \PP_{q+1,k}, a_{k,j} \psi_{q+1,j,k}] \nabla b_k(\lambda_{q+1} x) ,\end{aligned}$$ the commutator estimate (with $s= 0$ and $\lambda = \lambda_{q+1}$), and Lemma \[lem:D:N:a:psi\], we obtain the bound $$\begin{aligned} {\left\lVertO_{41}\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} \sum_{j,j',k,k'} \left( \lambda_{q+1}^{-1} {\left\lVert\nabla^2 (a_{k,j} \psi_{q+1,j,k})\right\rVert}_{C^0(\supp \chi_j)} + {\left\lVert\nabla (a_{k,j} \psi_{q+1,j,k})\right\rVert}_{C^0(\supp \chi_j)} \right) \lambda_{q+1} {\left\lVert{\widetilde}w_{q+1,j',k'}\right\rVert}_{C^0} \notag\\ &\lesssim \delta_{q+1}^{1/2} \left(\lambda_{q+1}^{-1} \lambda_q^2 \delta_{q+1}^{1/2} + \lambda_{q+1}^{-1} \tau_{q+1}^2 \lambda_{q+1}^2 \lambda_q^4 \delta_q + \lambda_q \delta_{q+1}^{1/2} + \delta_{q+1}^{1/2} \tau_{q+1} \lambda_{q+1} \lambda_q^2 \delta_q^{1/2} \right) \notag\\ &\lesssim \frac{\lambda_q^2 \delta_q^{1/2} \delta_{q+1}^{1/2}}{\lambda_{q+1}} + \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4} \notag\\ &\lesssim \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4}.\end{aligned}$$ The term $O_{42}$ is bounded similarly, by appealing to the commutator estimate (with $s=1$ and $\lambda = \lambda_{q+1}$): $$\begin{aligned} {\left\lVertO_{42}\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} \sum_{j,j',k,k'} \lambda_{q+1} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} {\left\lVert\nabla ( a_{k',j'} \psi_{q+1,j',k'})\right\rVert}_{C^0 (\supp \chi_{j'})} \notag\\ &\lesssim \delta_{q+1}^{1/2} \left( \lambda_q \delta_{q+1}^{1/2} + \delta_{q+1}^{1/2} \tau_{q+1} \lambda_q^2 \delta_{q}^{1/2} \right) \notag\\ &\leq \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4}\,.\end{aligned}$$ Combining the above estimates, we arrive at the bound $$\begin{aligned} \label{eq:O_4_bound} {\left\lVertO_4\right\rVert}_{C^0} \lesssim \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4}.\end{aligned}$$ ### Bound on $R_{O}$ Combining the estimates , , and yields $$\begin{aligned} {\left\lVertR_O\right\rVert}_{C^0}\lesssim \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4}= \lambda_0^{-2+\frac{5\beta}{2}}\lambda_{q+2} \delta_{q+2}.\end{aligned}$$ Assuming $\lambda_0$ is sufficiently large and $\beta<{\frac}45$ we obtain our claim. ### Material derivative of the $R_{O,{\rm approx}}$ stress We recall cf.  and that $$\begin{aligned} D_{t,q} R_{O,{\rm approx}} &= \sum_j \chi_j^2 D_{t,q}\Big( \mathring R_q - \mathring R_{q,j} \Big) + 2 \sum_j \chi_j \chi_j' \Big( \mathring R_q - \mathring R_{q,j} \Big) \notag\\ &= \sum_j \chi_j^2 \Big(D_{t,q} \mathring R_q\Big) + 2 \sum_j \chi_j \chi_j' \Big( \mathring R_q - \mathring R_{q,j} \Big).\end{aligned}$$ Therefore, using the inductive estimate , the fact that the $\chi_j^2$ form a partition of unity, and the previously established bound , we obtain that $$\begin{aligned} {\left\lVertD_{t,q} R_{O,{\rm approx}}\right\rVert}_{C^0} &\lesssim \lambda_q^2 \delta_q^{1/2} \lambda_{q+1} \delta_{q+1} + \tau_{q+1}^{-1} \left(\tau_{q+1} \lambda_q^2 \delta_q^{1/2} \lambda_{q+1} \delta_{q+1}\right) \notag\\ &\lesssim \lambda_q^2 \delta_q^{1/2} \lambda_{q+1} \delta_{q+1}\,.\end{aligned}$$ ### Material derivative of the $R_{O,{\rm low}}$ stress Recall cf.  that $R_{O,{\rm low}} = \mathring O_{21}+\mathring O_{22}$, terms which are defined in , with –. Trivially, $$\begin{aligned} {\left\lVertD_{t,q} \left(\mathring O_{21}+\mathring O_{22}\right)\right\rVert}_{C^0}\leq {\left\lVertD_{t,q}\left(O_{21}+ O_{22}\right)\right\rVert}_{C^0} \,.\end{aligned}$$ We only show the estimate for $O_{21}$, since the one for $O_{22}$ is identical, upon changing $(1)$ with $(2)$ below. Note that in view of the bilinear convolution operators defining ${{\widetilde}Q}_{j,k}^{(1)}$ has a kernel which obeys the conditions of Lemma \[lem:multi:Dt:commutator\]. Moreover, by construction we have $$D_{t,q}( a_{k,j} \psi_{q+1,j,k} ) = D_{t,q}( a_{k,j} \psi_{q+1,j,-k} ) = 0$$ and thus, using the notation we obtain $$\begin{aligned} &D_{t,q} \left( {\widetilde}\SSS_{k}^{(1),m\ell} ( \nabla (a_{k,j}\psi_{q+1,j,k}), a_{k,j} \psi_{q+1,j,-k}) \right) \notag\\ &\quad = {\widetilde}\SSS_{k}^{(1),m\ell} (D_{t,q}\left( \nabla (a_{k,j}\psi_{q+1,j,k})\right), a_{k,j} \psi_{q+1,j,-k}) + {\widetilde}\SSS_{k}^{(1),m\ell} ( \nabla (a_{k,j}\psi_{q+1,j,k}),D_{t,q}( a_{k,j} \psi_{q+1,j,-k})) \notag\\ &\quad \qquad + \left[ D_{t,q}, {\widetilde}\SSS_{k}^{(1),m\ell} \right](\nabla (a_{k,j}\psi_{q+1,j,k}), a_{k,j} \psi_{q+1,j,-k}) \notag\\ &\quad = {\widetilde}\SSS_{k}^{(1),m\ell} (\nabla u_q \cdot \nabla (a_{k,j}\psi_{q+1,j,k}) , a_{k,j} \psi_{q+1,j,-k}) + \left[ D_{t,q}, {\widetilde}\SSS_{k}^{(1),m\ell} \right](\nabla (a_{k,j}\psi_{q+1,j,k}), a_{k,j} \psi_{q+1,j,-k}).\end{aligned}$$ From and Lemma \[lem:multi:Dt:commutator\] it then follows that $$\begin{aligned} {\left\lVertD_{t,q} O_{21}\right\rVert}_{C^0} &\lesssim \sum_{j,k} {\left\lVert\chi_j\right\rVert}_{C^0} {\left\lVert\chi_j'\right\rVert}_{C^0} {\left\lVert {\widetilde}\SSS_{k}^{(1),m\ell} ( \nabla (a_{k,j}\psi_{q+1,j,k}), a_{k,j} \psi_{q+1,j,-k})\right\rVert}_{C^0} \notag\\ &\qquad + \sum_{j,k} {\left\lVert\chi_j^2\right\rVert}_{C^0} {\left\lVert {\widetilde}\SSS_{k}^{(1),m\ell} (\nabla u_q \cdot \nabla (a_{k,j}\psi_{q+1,j,k}), a_{k,j} \psi_{q+1,j,-k})\right\rVert}_{C^0} \notag\\ &\qquad + \sum_{j,k} {\left\lVert\chi_j^2\right\rVert}_{C^0} {\left\lVert\left[ D_{t,q}, {\widetilde}\SSS_{k}^{(1),m\ell} \right](\nabla (a_{k,j}\psi_{q+1,j,k}), a_{k,j} \psi_{q+1,j,-k})\right\rVert}_{C^0} \notag\\ &\lesssim \sum_{j,k} {\left\lVert\chi_j\right\rVert}_{C^0} {\left\lVert\chi_j'\right\rVert}_{C^0} {\left\lVert \nabla (a_{k,j}\psi_{q+1,j,k})\right\rVert}_{C^0} {\left\lVerta_{k,j} \psi_{q+1,j,-k}\right\rVert}_{C^0} \notag\\ &\qquad + \sum_{j,k} {\left\lVert\chi_j^2\right\rVert}_{C^0} {\left\lVert \nabla u_q\right\rVert}_{C^0} {\left\lVert\nabla (a_{k,j}\psi_{q+1,j,k})\right\rVert}_{C^0} {\left\lVerta_{k,j} \psi_{q+1,j,-k}\right\rVert}_{C^0} \notag\\ &\lesssim (\tau_{q+1}^{-1} + \lambda_q^2 \delta_{q}^{1/2}) (\lambda_q \delta_{q+1}^{1/2} + \delta_{q+1}^{1/2} \tau_{q+1}\lambda_{q+1} \lambda_q^2 \delta_q^{1/2} ) \delta_{q+1}^{1/2} \notag\\ &\lesssim \lambda_q^2\lambda_{q+1}\delta_q^{1/2}\delta_{q+1}\end{aligned}$$ The estimate for $O_{22}$ is similar, and we obtain that $$\begin{aligned} {\left\lVertD_{t,q} R_{O,{\rm low}}\right\rVert}_{C^0} \leq \lambda_q^2\lambda_{q+1}\delta_q^{1/2}\delta_{q+1}\,.\end{aligned}$$ ### Material derivative of the $R_{O,{\rm high}}$ stress Recall cf. , the decomposition of $R_{O,{\rm high}} = O_3 - O_4$. Applying $D_{t,q} = \partial_t + u_q \cdot \nabla$ to the $O_3$ equation, we find that $$\begin{aligned} D_{t,q} O_3 &= \left[D_{t,q},\BB {\widetilde}P_{\approx \lambda_{q+1}}\right] \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big) \cdot \nabla \big(\PP_{q+1,k'} {\widetilde}w_{q+1,j',k'}\big) \notag\\ &\quad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\left[D_{t,q}, \Lambda \PP_{q+1,k}\right] {\widetilde}w_{q+1,j,k}\big) \cdot \nabla \big(\PP_{q+1,k'} {\widetilde}w_{q+1,j',k'}\big) \notag\\ &\quad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big( \Lambda \PP_{q+1,k} (D_{t,q} {\widetilde}w_{q+1,j,k}) \big) \cdot \nabla \big(\PP_{q+1,k'} {\widetilde}w_{q+1,j',k'}\big) \notag\\ &\quad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big( \Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big) \cdot \left[D_{t,q},\nabla \PP_{q+1,k'} \right]\big({\widetilde}w_{q+1,j',k'}\big) \notag\\ &\quad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big( \Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big) \cdot \nabla \PP_{q+1,k'} \big(D_{t,q} {\widetilde}w_{q+1,j',k'}\big) \notag\\ &=: {\widetilde}O_{31} + {\widetilde}O_{32} + {\widetilde}O_{33} + {\widetilde}O_{34} + {\widetilde}O_{35}.\end{aligned}$$ The term ${\widetilde}O_{31}$ is bounded directly using Corollary \[cor:Dt:commutator\], estimate , with $\lambda = \lambda_{q+1}$ and $s=-1$ as $$\begin{aligned} {\left\lVert{\widetilde}O_{31}\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} {\left\lVert\nabla u_q\right\rVert}_{C^0} \sum_{j,j',k,k'} \lambda_{q+1} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} \lambda_{q+1} {\left\lVert{\widetilde}w_{q+1,j',k'}\right\rVert}_{C^0} \notag\\ &\lesssim \lambda_{q+1} \lambda_q^2 \delta_q^{1/2} \delta_{q+1} \,.\end{aligned}$$ Similarly, the terms ${\widetilde}O_{32}$ and ${\widetilde}O_{34}$ are bounded using with $\lambda = \lambda_{q+1}$ and $s= 1$ as $$\begin{aligned} {\left\lVert{\widetilde}O_{32}\right\rVert}_{C^0} +{\left\lVert{\widetilde}O_{34}\right\rVert}_{C^0} &\lesssim \lambda_{q+1}^{-1} \sum_{j,j',k,k'} \lambda_{q+1} {\left\lVert\nabla u_q\right\rVert}_{C^0} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} \lambda_{q+1} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0}\\ &\lesssim \lambda_{q+1} \lambda_q^2 \delta_q^{1/2} \delta_{q+1} \,.\end{aligned}$$ In order to bound ${\widetilde}O_{33}$ and ${\widetilde}O_{35}$ we note that by the construction of $w_{q+1}$ we have that $$\begin{aligned} D_{t,q}{\widetilde}w_{q+1,j,k}(x,t)=\chi'_{j}(t) a_{k,j}(x,t) b_k( \lambda_{q+1} \Phi_j(x,t) ) = \frac{\chi'_j(t)}{\chi_j(t)} {\widetilde}w_{q+1,j,k}(x,t).\end{aligned}$$ This fact allows us to rewrite ${\widetilde}O_{33}$ and ${\widetilde}O_{35}$ as follows: $$\begin{aligned} {\widetilde}O_{33} &= \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \frac{\chi_j'}{\chi_j} \big( \Lambda \PP_{q+1,k} ({\widetilde}w_{q+1,j,k}) \big) \cdot \nabla \big(\PP_{q+1,k'} {\widetilde}w_{q+1,j',k'}\big)\,, \\ {\widetilde}O_{35} &= \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \frac{\chi_{j'}'}{\chi_{j'}} \big( \Lambda \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big) \cdot \nabla \PP_{q+1,k'} \big( {\widetilde}w_{q+1,j',k'}\big)\,,\end{aligned}$$ and upon noting that $$\begin{aligned} {\left\lVert\frac{\chi_j'}{\chi_j}\right\rVert} + {\left\lVert\frac{\chi_{j'}'}{\chi_{j'}}\right\rVert} \lesssim \tau_{q+1}^{-1} \,,\end{aligned}$$ we may use the bounds previously established for $O_3$ to conclude that $$\begin{aligned} {\left\lVert{\widetilde}O_{33}\right\rVert}_{C^0} +{\left\lVert{\widetilde}O_{35}\right\rVert}_{C^0} &\lesssim \tau_{q+1}^{-1} {\left\lVertO_3\right\rVert}_{C^0} \notag\\ &\lesssim \tau_{q+1}^{-1} \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4} \notag\\ &\leq \lambda_{q+1} \lambda_q^2 \delta_q^{1/2} \delta_{q+1} \,.\end{aligned}$$ Bounding the material derivative of $O_4$ is very similar. We first write $D_{t,q} O_4 $ as $$\begin{aligned} D_{t,q} O_4 &= \left[D_{t,q}, \BB {\widetilde}P_{\approx \lambda_{q+1}}\right] \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big)^T \cdot \big( \Lambda \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} \big) \notag\\ &\qquad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big(\big[D_{t,q},\nabla \PP_{q+1,k}\big] {\widetilde}w_{q+1,j,k}\big)^T \cdot \big( \Lambda \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} \big) \notag\\ &\qquad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \frac{\chi_j'}{\chi_j} \big(\nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big)^T \cdot \big( \Lambda \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} \big) \notag\\ &\qquad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \big( \nabla \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\big)^T \cdot \big( \big[ \Lambda \PP_{q+1,k'}, D_{t,q} \big] {\widetilde}w_{q+1,j',k'} \big) \notag\\ &\qquad + \BB {\widetilde}P_{\approx \lambda_{q+1}} \sum_{\underset{k+k'\neq 0 } {j,j',k,k'}} \frac{\chi_{j'}'}{\chi_{j'}} \big(\big[D_{t,q},\nabla \PP_{q+1,k}\big] {\widetilde}w_{q+1,j,k}\big)^T \cdot \big( \Lambda \PP_{q+1,k'} {\widetilde}w_{q+1,j',k'} \big) \,,\end{aligned}$$ and using similar arguments as above, together with the bound previously established on $O_4$, we obtain that $$\begin{aligned} {\left\lVertD_{t,q}O_4\right\rVert}_{C^0} &\lesssim \tau_{q+1}^{-1} {\left\lVertO_4\right\rVert}_{C^0} + \lambda_{q+1}^{-1} {\left\lVert\nabla u_q\right\rVert}_{C^0} \sum_{j,j',k,k'} \lambda_{q+1} {\left\lVert{\widetilde}w_{q+1,j,k}\right\rVert}_{C^0} \lambda_{q+1} {\left\lVert{\widetilde}w_{q+1,j',k'}\right\rVert}_{C^0} \notag\\ &\lesssim \tau_{q+1}^{-1} \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4} + \lambda_{q+1} \delta_{q+1} \lambda_q^2 \delta_q^{1/2} \notag\\ &\lesssim \lambda_q^2 \lambda_{q+1} \delta_q^{1/2}\delta_{q+1} \,,\end{aligned}$$ which concludes the proof of the material derivative estimate for $R_{O,{\rm high}}$. ### Bound on the material derivative of $R_{O}$ Combining all the estimates yields $$\begin{aligned} {\left\lVertD_{t,q} R_{O}\right\rVert}_{C^0}\lesssim \lambda_q^2 \lambda_{q+1} \delta_q^{1/2}\delta_{q+1}= \lambda_0^{-3+3\beta}\lambda_{q+1}^{2}\ \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2}\end{aligned}$$ from which we obtain our desired estimate if $\beta<1$ and $\lambda_0$ is sufficiently large. Concluding bounds on the new Reynolds stress -------------------------------------------- Combining all the estimates above, we now show that for the new Reynolds stress $\mathring{R}_{q+1}$, estimate holds with $q$ replaced by $q+1$ as follows: Assuming $\lambda_0$ is sufficiently large then and hold, and $$\begin{aligned} \| \mathring R_q\|_{C^0} &\leq \eps_R \lambda_{q+1} \delta_{q+1} \,, \label{eq:R_est_final}\\ {\left\lVert (\partial_t + u_q \cdot \nabla) \mathring R_q\right\rVert}_{C^0} &\leq \lambda_q^2 \delta_q^{1/2} \lambda_{q+1} \delta_{q+1}.\label{eq:D_t_R_est_final}\end{aligned}$$ The first estimate follows directly from Lemmas \[lem:transp:error\]–\[lem:osc:error\]. To prove , first note that since ${\left\lVert\Lambda w_{q+1}\right\rVert}_{C^0} \lesssim \lambda_{q+1} \delta_{q+1}^{1/2}$, and $\mathring R_{q+1}$ has compact support in frequency in the ball of radius $4 \lambda_{q+1}$, we obtain from Lemmas \[lem:transp:error\]–\[lem:osc:error\] that if $\lambda_0$ is sufficiently large, then $$\begin{aligned} {\left\lVertD_{t,q+1} \mathring R_{q+1}\right\rVert}_{C^0} &\leq {\left\lVertD_{t,q} \mathring R_{q+1}\right\rVert}_{C^0} + {\left\lVert\Lambda w_{q+1} \cdot \nabla \mathring R_{q+1}\right\rVert}_{C^0} \\ &\leq \frac12 \lambda_{q+1}^2 \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2} + \frac12\lambda_{q+1} \delta_{q+1}^{1/2} \lambda_{q+1} \lambda_{q+2} \delta_{q+2} = \lambda_{q+1}^2 \delta_{q+1}^{1/2} \lambda_{q+2} \delta_{q+2} \,.\end{aligned}$$ The Hamiltonian increment {#sec:hamiltonian} ========================= In this section, we conclude the proof of Proposition \[prop:main\] by showing that and hold with $q$ replaced by $q+1$. We begin by stating some consequences of the inductive estimates in Section \[s:inductive\_assump\]: \[lem:rho\_diff\] If $t$ is in the support of the cut-off function $\chi_j$, then $$\label{eq:rho_diff} {\left\vert\int_{\TT^2}\left({\left\vert\Lambda^{{\frac}12} v_q(x,t)\right\vert}^2-{\left\vert\Lambda^{{\frac}12} v_q(x,\tau_{q+1}j)\right\vert}^2\right)~dx\right\vert}+{\left\vert\ee(t)+\ee(\tau_{q+1}j)\right\vert}\leq \frac{\lambda_{q+ 2}\delta_{q+2}}{16}~.$$ Consequently, if $\rho_j\neq 0$ then on the support of $\chi_j$ $$\label{eq:rho_j_t} \lambda_{q+1}{\left\vert\rho(t)-\rho_j\right\vert}\leq \frac{\lambda_{q+ 2}\delta_{q+2}}{16}\,,$$ and by the definition $$\label{eq:energy_lower} \ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_q\right\vert}^2~dx\geq\frac{7\lambda_{q+ 2}\delta_{q+2}}{16}.$$\[eq:rho\_vanish\] If $\rho_j=0$ then by the definition and $$\begin{aligned} e(j\tau_{q+1}) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_q(x,j\tau_{q+1})\right\vert}^2~dx \leq \frac{9\lambda_{q+ 2}\delta_{q+2}}{16}~\mbox{and}~\mathring R_q(\cdot, t)\equiv 0~.\end{aligned}$$ Using the equation for $v_q$, we have that $$\begin{aligned} {\left\vert\int_{\TT^2}\left({\left\vert\Lambda^{{\frac}12} v_q(x,t)\right\vert}^2-{\left\vert\Lambda^{{\frac}12} v_q(x,\tau_{q+1}j)\right\vert}^2\right)~dx\right\vert}& = 2{\left\vert\int_{\tau_{q+1}j}^t\Lambda^{{\frac}12}v_q \cdot \div \Lambda^{{\frac}12}R_q~dx\right\vert}\\ & \lesssim {\left\vertt-\tau_{q+1} j\right\vert}\lambda_q^2\delta_q^{{\frac}12}\lambda_{q+1}\delta_{q+1}\\ & \lesssim \tau_{q+1}\lambda_q^2\delta_q^{{\frac}12}\delta_{q+1}\lambda_{q+1}\\ & = \lambda_q\delta_q^{1/4}\delta_{q+1}^{3/4} \,,\end{aligned}$$ where in the second to last line, we used that by our hypothesis ${\left\vertt-\tau_{q+1} j\right\vert}\leq 4\tau_{q+1}$. Trivially, we have that $$\begin{aligned} {\left\vert\ee(t)+\ee(\tau_{q+1}j)\right\vert}\lesssim \tau_{q+1}\,.\end{aligned}$$ Thus, $${\left\vert\int_{\TT^2}\left({\left\vert\Lambda^{{\frac}12} v_q(x,t)\right\vert}^2-{\left\vert\Lambda^{{\frac}12} v_q(x,\tau_{q+1}j)\right\vert}^2\right)~dx\right\vert}+{\left\vert\ee(t)+\ee(\tau_{q+1}j)\right\vert}\lesssim \left(\lambda_0^{-2+5\beta/2}+\lambda_0^{-3\beta/2}\lambda_{q+2}^{-3+3\beta}\right)\lambda_{q+2}\delta_{q+2}$$ Hence, if $\beta<\frac 45$, the lemma is proved. We are now in position to prove that and hold with $q$ replaced by $q+1$. If $\rho(j)\neq 0$ and $t$ is in the support of $\chi_j$, then $$\label{eq:energy_inductive_step} \frac{\lambda_{q+ 2}\delta_{q+2}}{4} \leq \ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_{q+1}\right\vert}^2~dx\leq\frac{3\lambda_{q+ 2}\delta_{q+2}}{4}.$$ Otherwise, if $t$ is *not* in the support of the cut-off $\chi_j$ with $\rho(j)\neq 0$, then $$\label{eq:upper_no_perturb} \ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_{q+1}\right\vert}^2~dx\leq\frac{9\lambda_{q+ 2}\delta_{q+2}}{16}~\mbox{and}~\mathring R_{q+1}(\cdot,t)\equiv 0\,.$$ As a consequence of and , it follows that and hold with $q$ replaced by $q+1$. Assume that $t$ is not on the support of any cut-off $\chi_j$ with $\rho_j\neq 0$, then since $w_{q+1}(\cdot, t)\equiv 0$ and $\mathring R_{q+1}(\cdot,t)=\mathring R_q(\cdot,t)$ (see Lemma \[lem:rho\_diff\] above) then $\mathring R_{q+1}(\cdot,t)\equiv 0$. Moreover, we have $$\ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_{q+1}\right\vert}^2~dx=\ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_{q}\right\vert}^2~dx,$$ and thus by we obtain . Now assume $\rho(j)\neq 0$ and $t$ is on the support of $\chi_j$. By computation it follows that $$\begin{aligned} \ee(t)-\int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_{q+1}\right\vert}^2~dx&= \ee(t)-\int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_{q}\right\vert}^2~dx+2\int_{\TT^2}\Lambda^{{\frac}12} v_{q} \cdot \Lambda^{{\frac}12} w_{q+1}~dx-\int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_{q}\right\vert}^2~dx \\&=\ee(t)-\int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_{q}\right\vert}^2~dx-\int_{\TT^2}{\left\vert\Lambda^{{\frac}12} w_{q+1}\right\vert}^2~dx\end{aligned}$$ where we used the disjoint frequency support of $v_q$ and $w_{q+1}$. Utilizing the frequency supports of ${\widetilde}w_{q+1,j,k}$ we obtain $$\begin{aligned} \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} w_{q+1}\right\vert}^2~dx=&\int_{\TT^2} {\left\vert\sum_{k,k',j,j'} \Lambda^{{\frac}12} \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\right\vert}^2~dx\\ =&\sum_{k,j'}\int_{\TT^2} \Lambda^{{\frac}12} \PP_{q+1,k} {\widetilde}w_{q+1,j,k}\cdot \Lambda^{{\frac}12} \PP_{q+1,-k} {\widetilde}w_{q+1,j,-k}~dx \\ =&\sum_{k,j}\int_{\TT^2} \Lambda^{{\frac}12} \PP_{q+1,k} \left( \chi_{j} a_{k,j} b_k (\lambda_{q+1} \Phi_j) \right) \cdot \Lambda^{{\frac}12} \PP_{q+1,-k} \left( \chi_{j} a_{-k,j} b_{-k} (\lambda_{q+1} \Phi_j) \right)~. ~dx \end{aligned}$$ Then, by the definition of $b_k$, estimates and , and the fact that the mean of a high frequency object vanishes, it follows that $$\begin{aligned} \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} w_{q+1}\right\vert}^2~dx &= \int_{\TT^2} w_{q+1} \cdot \Lambda w_{q+1} ~dx \notag\\ &=\sum_{k,j}\int_{\TT^2} \lambda_{q+1} \chi_j^2{\left\verta_k\right\vert}^2~dx \\ &\quad +\sum_{j,k} \int_{\TT^2}\chi_{j} \big[\PP_{q+1,k}, a_{k,j} \psi_{q+1,j,k} \big] b_k( \lambda_{q+1} x )\cdot \Lambda w_{q+1}~dx \\ &\quad + \sum_{j,k} \int_{\TT^2} {\widetilde}w_{q+1,j,k}\cdot \chi_{j} [\PP_{q+1,-k} \Lambda, a_{k,j} \psi_{q+1,j,-k} ] b_{-k}( \lambda_{q+1} x )~dx \\ &=: (2\pi)^2 \lambda_{q+1}\sum_j \chi_j^2 \rho_j+ E_1+E_2.\end{aligned}$$ Using Lemmas \[lem:w:q+1:bounds\], \[lem:D:N:a:psi\], and \[lem:u:grad:commutator\] we may estimate $$\begin{aligned} E_1+E_2 &\lesssim \sum_{j,k} \chi_j {\left\lVert\nabla (a_{k,j} \psi_{q+1,j,k})\right\rVert} \delta_{q+1}^{1/2} \notag\\ &\lesssim \lambda_q\delta_{q+1}+\lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4} \notag\\ &\lesssim \lambda_q \delta_q^{1/4} \delta_{q+1}^{3/4}=\lambda_0^{-2+5\beta/2}\lambda_{q+1}\delta_{q+2}.\end{aligned}$$ Finally, applying , and and $\beta<4/5$ we obtain $$\begin{aligned} \frac{\lambda_{q+ 2}\delta_{q+2}}{4} \leq \ee(t) - \int_{\TT^2}{\left\vert\Lambda^{{\frac}12} v_{q+1}\right\vert}^2~dx\leq\frac{3\lambda_{q+ 2}\delta_{q+2}}{4}\end{aligned}$$ which concludes the proof. Appendix ======== Variational principle for hydrodynamical systems {#appendix:EP} ------------------------------------------------ We provide a derivation of the system . Many models of incompressible hydrodynamical systems can be written as geodesic equations of right-invariant metrics on the Lie group $ \mathcal{D} _\mu$, the group of volume-preserving diffeomorphisms, with group multiplication given by composition on the right. The Lie algebra $ \mathcal{V} $ associated to this group is the vector space of divergence-free vector fields. The Lie bracket on $ \mathcal{V} $ is given by $[u,v] = \partial_j u v^j - \partial _j v u^j$. On $ \mathcal{V} $ we (formally) define the metric $$(u, w) = \int_{ \mathbb{T} ^2} A u \cdot w \, dx \,,$$ where $A$ is a self-adjoint, positive operator. This metric is then right-translated over the Lie group $ \mathcal{D} _\mu$. We then define the Lagrangian function $l$ on the Lie algebra $ \mathcal{V} $ by $$\label{eq:A-lag} l(u) = {\frac{1}{2}} (u, u) = \int_{ \mathbb{T} ^2} A u \cdot u \, dx \,.$$ The well-known Euler-Poincaré variational principle provides a simple procedure for computing the equations of motion associated to the Lagrangian $l$ on the Lie algebra $ \mathcal{V} $. We shall state this as the following proposition, whose proof can be found in Chapter 13 of [@MaRa1999]). \[prop:ep\] With $l: \mathcal{V} \rightarrow{\mathbb R}$ given by , the following are equivalent: - $u(t):=u(t, \cdot )$ is a geodesic curve, solving $$\frac{d}{dt}\frac{\delta l}{\delta u} = -\operatorname{ad}^*_{u} \frac{\delta l}{\delta u},$$ where $\operatorname{ad}^*_u$ is defined by $$(\operatorname{ad}^*_u v, w) = - (v, [u,w]),$$ for $u,v,w$ in $\mathcal{V} $; - the curve $u(t)$ is an extremum of the action function $$s(u) = \int l(u(t)) dt,$$ for variations of the form $$\nonumber \delta u = \partial_t w + [w, u],$$ where $w \in \mathcal{V} $ vanishes at the endpoints $t=0$ and $t=T$. We make use of Proposition \[prop:ep\] [**(b)**]{} to derive . We define the potential velocity $$v = A u,$$ and compute the first variation of $s(u) = \int_0^T l(u) dt$: $$\begin{aligned} \delta s(u) \cdot \delta u &= \int_0^T \int_{ \mathbb{T} ^2 }v \cdot \delta u \, dx dt \\ &= \int_0^T \int_{ \mathbb{T} ^2 } v ^i\, \left( \partial_t w^i + \p_j w^i u^j - \p_ju^i w^j \right) \, dx dt\\ &= -\int_0^T \int_{ \mathbb{T} ^2 } \left( \partial_t v^i + \p_j v^i u^j + \p_iu^j v^j \right) \, w^i \, dx dt \,.\end{aligned}$$ Setting $\delta s(u) \cdot \delta u=0$ for all divergence-free variations $w$, and applying the Hodge decomposition, we find that there exists a pressure function $ {\widetilde}p : \mathbb{T} ^2 \to \mathbb{R} $ such that \[eq:SQG-bad\] $$\begin{aligned} \partial_t v^i + \p_j v^i u^j + \p_iu^j v^j & = -\p_i {\widetilde}p \,, \\ \operatorname{div} u &= 0 \,,\end{aligned}$$ which is the general hydrodynamical system . Transport and composition estimates {#sec:transport} ----------------------------------- In this section we gather some classical estimates for transport equations. For refer the reader to e.g. [@Bu2014 Section 4.3] or [@BuDLeIsSz2015 Proposition D.1] for proofs of these classical facts. In this section we recall some well known results regarding smooth solutions of the *transport equation*: \[eq:transport\] $$\begin{aligned} \partial_t f + u\cdot \nabla f &=g,\\ f|_{t_0}&=f_0,\end{aligned}$$ where $u=u(t,x)$ is a given smooth vector field. We denote the corresponding material derivative by $D_{t} = \partial_t+u\cdot \nabla$. Moreover, define by $\Phi (t, \cdot)$ to be the inverse of the flow associated to the vector field $u$ starting at time $t_0$ as the identity, i.e., $\Phi = X^{-1}$, where $\frac{d}{dt} X(x,t) = u (X(x,t),t)$ and $X (x, t_0 )=x$. Then we have: \[lem:transport\] Assume $t>t_0$. Any solution $f$ of satisfies $$\begin{aligned} {\left\lVertf (t)\right\rVert}_{C^0} &\leq {\left\lVertf_0\right\rVert}_{C^0} + \int_{t_0}^t {\left\lVertg (\tau)\right\rVert}_{C^0}\, d\tau\,,\label{eq:max:prin}\\ {\left\lVertD f(t)\right\rVert}_{C^0} &\leq {\left\lVertD f_0\right\rVert}_{C^0} e^{(t-t_0) {\left\lVertD u\right\rVert}_{C^0}} + \int_{t_0}^t e^{(t-\tau) {\left\lVertD u\right\rVert}_{C^0}} {\left\lVert Dg (\tau)\right\rVert}_{C^0}\, d\tau\,,\label{eq:trans:est:0}\end{aligned}$$ and, more generally, for any $N\geq 2$ there exists a constant $C=C(N)$ so that $$\begin{aligned} {\left\lVert D^N f (t)\right\rVert}_{C^0} & \leq \bigl( {\left\lVertD^N f_0\right\rVert}_{C^0} + C(t-t_0) {\left\lVertD^N u\right\rVert}_{C^0} {\left\lVertD f_0\right\rVert}_{C^0} \bigr)e^{C(t-t_0) {\left\lVertD u\right\rVert}_{C^0}} \notag \\ &\qquad +\int_{t_0}^t e^{C(t-\tau) {\left\lVertD u\right\rVert}_{C^0} }\bigl({\left\lVertD^N g (\tau)\right\rVert}_{C^0} + C (t-\tau ) {\left\lVert D^N v\right\rVert}_{C^0} {\left\lVert D g (\tau)\right\rVert}_{C^0} \bigr)\,d\tau. \label{eq:trans:est:1}\end{aligned}$$ Moreover, $$\begin{aligned} {\left\lVertD\Phi (t) -{\ensuremath{\mathrm{Id}}}\right\rVert}_{C^0} &\leq e^{(t-t_0){\left\lVertD u\right\rVert}_{C^0}}-1 \leq (t-t_0){\left\lVertD u\right\rVert}_{C^0} e^{(t-t_0){\left\lVertD u\right\rVert}_{C^0}}\,, \label{eq:Dphi:near:id}\\ {\left\lVertD^N \Phi (t)\right\rVert}_{C^0} &\leq C(t-t_0) {\left\lVertD^N u\right\rVert}_{C^0} e^{C(t-t_0){\left\lVertD u\right\rVert}_{C^0}}, \label{eq:Dphi:N}\end{aligned}$$ holds for all $N \geq 2$ and a suitable constant $C = C(N)$. In order to take advantage of Lemma \[lem:transport\] we also need to appeal to the following standard composition estimate: \[lem:composition\] Let $\Psi: \Omega \to \RR$ and $u: \RR^d \to \Omega$ be two smooth functions, with $\Omega\subset \RR^D$. Then, for every $N \in \NN\setminus \{0\}$ there is a constant $C$ such that $$\begin{aligned} {\left\lVertD^N \left( \Psi\circ u \right)\right\rVert}_{C^0} &\leq C \left({\left\lVertD \Psi\right\rVert}_{C^0} {\left\lVertD^N u\right\rVert}_{C^0} +\|D\Psi\|_{C^{N-1}} {\left\lVertu\right\rVert}_{C^0}^{N-1} {\left\lVertD^N u\right\rVert}_{C^0} \right)\, . \label{eq:chain:i} \\ {\left\lVertD^N \left( \Psi\circ u \right)\right\rVert}_{C^0} &\leq C \left({\left\lVertD \Psi\right\rVert}_{C^0} {\left\lVertD^N u\right\rVert}_{C^0} +\|D\Psi\|_{C^{N-1}} {\left\lVertD u\right\rVert}_{C^0}^{N} \right)\, . \label{eq:chain}\end{aligned}$$ where $C = C (N,d,D)$. Inverse of the divergence {#sec:BB} ------------------------- In this section we prove a number of estimates for the operator $\BB$ defined in Definition \[def:BB\]. We take advantage of the frequency localization of our perturbation and establish the following lemma. \[lem:BB\] Let $\BB$ be as defined in Definition \[def:BB\]. For $f \colon \TT^2 \to \CC^2$ that is smooth, we have that $$\div (\BB f) = \PP \left(f - \frac{1}{|\TT^2|} \int_{\TT^2} f(x) dx\right),$$ and $\BB f$ is a symmetric, trace-free matrix. Fix $\lambda \geq 1$, and denote by $P_{\approx \lambda}$ a Fourier multiplier operator with symbol that is supported on $\{ \xi \colon \lambda/2 \leq |\xi|\leq 2\lambda\}$ and is identically $1$ on $\{ \xi \colon 3\lambda/4 \leq |\xi| \leq 3\lambda/2\}$. Then, for a smooth functions $f,g \colon \TT^2 \to \CC$, with $\supp ( {\widehat}g(\xi) ) \subset \{\xi \colon |\xi|\leq \lambda/4\}$, we have $$\begin{aligned} {\left\lVert\BB \big( g(x) P_{\approx \lambda} f(x) \big)\right\rVert}_{C^0} &\lesssim \frac{{\left\lVertg\right\rVert}_{C^0} {\left\lVertP_{\approx \lambda} f\right\rVert}_{C^0} }{\lambda} \lesssim \frac{{\left\lVertg\right\rVert}_{C^0} {\left\lVertf\right\rVert}_{C^0} }{\lambda} \label{eq:BB:L:infty} \\ {\left\lVert[ \BB , g(x)] P_{\approx \lambda} f(x) \right\rVert}_{C^0} &\lesssim \frac{{\left\lVertg\right\rVert}_{C^1} {\left\lVertP_{\approx \lambda} f\right\rVert}_{C^0} }{\lambda^2} \lesssim \frac{{\left\lVertg\right\rVert}_{C^1} {\left\lVertf\right\rVert}_{C^0} }{\lambda^2} \label{eq:BB:comm}\end{aligned}$$ for some implicit universal constant $C> 0$. The first assertions follow directly from the definition of $\BB$. In order to prove we note that by the assumption on the frequency support of $g$, we have $$\begin{aligned} g(x) P_{\approx \lambda} f(x) = {\widetilde}P_{\approx \lambda} \big( g(x) P_{\approx \lambda} f(x) \big)\end{aligned}$$ where we have denoted $$\begin{aligned} {\widetilde}P_{\approx \lambda} = P_{\approx 2^{j-2} \lambda} + P_{\approx 2^{j-1} \lambda} + P_{\approx 2^j \lambda} + P_{\approx 2^{j+1} \lambda} .\end{aligned}$$ Thus, by the definition of $\BB$ and applying twice the Bernstein inequality on $L^\infty$, we obtain $$\begin{aligned} {\left\lVert\BB \big( g(x) P_{\approx \lambda} f(x) \big)\right\rVert}_{C^0} \lesssim \frac{1}{\lambda} {\left\lVertg(x) P_{\approx \lambda} f(x) \right\rVert}_{C^0} \lesssim \frac{1}{\lambda} {\left\lVertg\right\rVert}_{C^0} {\left\lVertf\right\rVert}_{C^0}.\end{aligned}$$ Lastly, in order to prove we note that since $\BB$ has components which are Fourier multipliers, we have $$\begin{aligned} [\BB , g] P_{\approx \lambda} f &= \BB \big( g P_{\approx \lambda} f \big) - g \BB P_{\approx \lambda} f \\ &= \BB {\widetilde}P_{\approx \lambda} \big( g P_{\approx \lambda} f \big) - g \BB {\widetilde}P_{\approx \lambda} P_{\approx \lambda} f \notag\\ &= [ \BB {\widetilde}P_{\approx \lambda} , g] P_{\approx \lambda} f.\end{aligned}$$ Denoting by ${\mathcal K}_{\approx\lambda}$ the real convolution kernel corresponding to the Fourier multiplier operator $\BB {\widetilde}P_{\approx \lambda} $ (which is a symbol of order $-1$), one may check that it obeys the bounds $$\begin{aligned} {\left\lVert|x|^b \nabla^a {\mathcal K}_{\approx \lambda} (x)\right\rVert}_{L^1(\RR^2)} \leq C \lambda^{a-b-1} \label{eq:BB:K:aux}\end{aligned}$$ for all $a,b\geq 0$, and some constant $C = C(a,b)$. Therefore, by the mean value theorem $$\begin{aligned} [ \BB {\widetilde}P_{\approx \lambda} , g] P_{\approx \lambda} f(x) &= \int_{\RR^d} (g(y) - g(x)) {\mathcal K}_{\approx \lambda}(x-y) P_{\approx \lambda} f(y) dy \notag\\ &=- \int_{\RR^d} \left(\int_0^1 \nabla g(y - \tau (y-x)) d\tau\right)\cdot (y-x) {\mathcal K}_{\approx \lambda}(x-y) P_{\approx \lambda} f(y) dy\end{aligned}$$ and the kernel estimate with $b=1$ and $a=0$ we arrive at $$\begin{aligned} {\left\lVert[ \BB {\widetilde}P_{\approx \lambda} , g] P_{\approx \lambda} f\right\rVert}_{C^0} &\lesssim {\left\lVertD g\right\rVert}_{C^0} {\left\lVert|x| {\mathcal K} _{\approx \lambda}(x)\right\rVert}_{L^1} {\left\lVertP_{\approx \lambda} f\right\rVert}_{C^0} \notag\\ &\lesssim \lambda^{-2} {\left\lVertg\right\rVert}_{C^1} {\left\lVertf\right\rVert}_{C^0}.\end{aligned}$$ This concludes the proof of the lemma. Calderon commutator {#sec:commutator} ------------------- We recall cf. [@Le2002 Theorem 10.3, Page 99] and [@Ma2008 Lemma 2.2, Page 50]. \[lem:Calderon\] Let $p \in (1,\infty)$ and $\varphi \in W^{1,\infty}(\TT^2)$. Then for any $v \in L^p(\TT^2)$ with zero mean on $\TT^2$ we have that $$\begin{aligned} {\left\lVert [ \Lambda, \varphi] v\right\rVert}_{L^p} \lesssim_p {\left\lVert\varphi\right\rVert}_{W^{1,\infty}} {\left\lVertv\right\rVert}_{L^p} \label{eq:Calderon:Lp}\end{aligned}$$ where the constant implicitly depends on $p$. Moreover, for $s \in [0,1]$, if $\varphi \in W^{2,\infty} (\TT^2)$ and $v \in H^s(\TT^2)$ has zero mean on $\TT^2$, then[^20] $$\begin{aligned} {\left\lVert [ \Lambda, \varphi] v\right\rVert}_{H^s} \lesssim_s {\left\lVert\varphi\right\rVert}_{W^{2,\infty}} {\left\lVertv\right\rVert}_{H^s} \label{eq:Calderon:Hs}\end{aligned}$$ Note that in the aforementioned references the results are stated for functions defined on $\RR^2$, whose Fourier support is at a positive distance from the origin. The same proofs work in the periodic case $\TT^2$, if the functions we consider have zero mean. To see that holds one uses the Poisson summation convention to write the kernel associated to $\Lambda$ as in [@CoCo2004], so that off the diagonal, the singular integral kernel $K$ associated to $[\Lambda,\varphi]$ is given by $$K(x,y) = c \sum_{k\in \ZZ^2}\frac{\varphi(x) - \varphi(y)}{|x-y-k|^3}$$ For $k = 0$ the proof closely follows the $\RR^2$ case, while for $ k\neq 0$ we are dealing with an $L^1$ kernel. Assertion spaces follows from the case $s=0$ (which holds by letting $p=2$ in ), the case $s=1$ (which holds since $\nabla [\Lambda, \varphi] = [\Lambda, \varphi] \nabla + [\Lambda, \nabla \varphi]$ and the bound with $p=2$), and interpolation. We omit further details. Material derivatives and convolution operators ---------------------------------------------- We recall (similarly to [@IsVi2015 Lemma 7.2]) a commutator estimate involving convolution operators and material derivatives. \[lem:u:grad:commutator\] Let $s\in \RR$, $\lambda \geq 1$, and let $T_K$ be an order $s$ convolution operator localized at length scale $\lambda^{-1}$. That is, $T_K$ acts on smooth functions $f$ as $$T_K f(x) = \int_{\RR^2} K(y) f(x-y) dy$$ for some kernel $K \colon \RR^2 \to \RR$ that obeys $$\begin{aligned} {\left\lVert|x|^a \nabla^b K(x)\right\rVert}_{L^1(\RR^2)} \lesssim \lambda^{b-a + s} \label{eq:K:convolution:assumption}\end{aligned}$$ for all $0 \leq a, |b|\leq 1$ and some implicit constants $C = C(a,b)$. Then, for any smooth function $f \colon \TT^2 \to \CC$ and smooth incompressible vector field $u \colon \TT^2 \to \RR^2$ we have $$\begin{aligned} {\left\lVert[ u \cdot \nabla, T_K ] f\right\rVert}_{C^0} \leq \lambda^s {\left\lVert\nabla u\right\rVert}_{C^0} {\left\lVertf\right\rVert}_{C^0}.\end{aligned}$$ Similarly, we have $$\begin{aligned} {\left\lVert[ b, T_K ] f\right\rVert}_{C^0} \leq \lambda^{s-1} {\left\lVert\nabla b\right\rVert}_{C^0} {\left\lVertf\right\rVert}_{C^0} \label{eq:comm:multiplication}\end{aligned}$$ for smooth functions $b, f \colon \TT^2 \to \CC$. The proof is direct, and uses that $\div u=0$. We have $$\begin{aligned} \left| T_K(u\cdot \nabla f)(x) - u(x) \cdot \nabla T_K f (x) \right| &= \left| \int_{\RR^2} \left( u(x) - u(x-y) \right) \cdot \nabla f(x-y) K(y) dy \right| \notag\\ &= \left| \int_{\RR^2} \left( u(x) - u(x-y) \right) \cdot \nabla K(y) f(x-y) dy \right| \notag\\ &\leq {\left\lVert\nabla u\right\rVert}_{C^0} \int_{\RR^2} |f(x-y)| |y| | \nabla K(y)| dy \notag\\ &\leq {\left\lVert\nabla u\right\rVert}_{C^0} {\left\lVertf\right\rVert}_{C^0} {\left\lVert|y| \nabla K(y)\right\rVert}_{L^1}\end{aligned}$$ at which stage we use that the kernel is integrable cf. . For the second assertion, we use the mean value theorem. We have $$\begin{aligned} \left| T_K(b\, f)(x) - b(x) T_K f (x) \right| &= \left| \int_{\RR^2} \left( b(x) - b(x-y) \right) f(x-y) K(y) dy \right| \notag\\ &= \left| \int_{\RR^2} \left( \int_0^1 \nabla b(x - \lambda y) d\lambda \right) \cdot y K(y)f(x-y) dy \right| \notag\\ &\leq {\left\lVert\nabla b\right\rVert}_{C^0} {\left\lVertf\right\rVert}_{C^0} {\left\lVert|y| K(y)\right\rVert}_{L^1}\end{aligned}$$ so that the proof is completed upon using . When $s=0$, examples of such operators $T_K$ are given by zero-order Fourier multiplier operators with frequency support inside a shell at $|\xi| \approx \lambda$. For instance, the kernel associated to the operator $\PP_{q+1,k}$ obeys for $\lambda = \lambda_{q+1}$, and $s=0$. Similarly, the kernel associated to the Fourier multiplier operator $\Lambda \PP_{q+1,k}$ obeys for $\lambda = \lambda_{q+1}$, and $s=1$, while the kernel of $\BB \PP_{q+1,k}$ obeys for $\lambda = \lambda_{q+1}$, and $s=-1$. An immediate consequence of Lemma \[lem:u:grad:commutator\] is: \[cor:Dt:commutator\] Let $\lambda \geq 1$, $s\in \RR$, and let $K$ be a kernel which obeys . Given a smooth divergence free vector field $u \colon \TT^2 \to\RR^2$, we have $$\begin{aligned} {\left\lVert[D_t , T_K ] f \right\rVert}_{C^{0}} \lesssim \lambda^s {\left\lVert\nabla u\right\rVert}_{C^0} {\left\lVertf\right\rVert}_{C^0}. \label{eq:comm:material:convolution}\end{aligned}$$ where as usual we denote $D_t = \partial_t + u \cdot \nabla$. Material derivatives and bilinear convolution operators {#app:pseudo:product} ------------------------------------------------------- The bilinear convolution operators we are interested here arise from pseudo-product operators, as defined by Coifman and Meyer [@CoMe1978]. Equivalently, these are translation invariant multilinear operators, see e.g. [@GrTo2002 Section 6] and [@MuSh2013 Chapter 2.13] for details. Let $\xi = (\xi_1, \xi_2) \in \RR^2 \times \RR^2$. Let $M \colon \RR^2 \times \RR^2 \to \RR$ be a smooth multiplier. For two Schwartz functions $f_1, f_2 \colon \RR^2\to \RR$, define the bilinear pseudo-product operator $S_M (f_1,f_2)$ by $$\begin{aligned} S_M (f_1,f_2)(x) = \frac{1}{(2\pi)^2} \intint_{\RR^2\times \RR^2} M(\xi_1,\xi_2) {\widehat}{f_1}(\xi_1) {\widehat}{f_2}(\xi_2) e^{i x \cdot (\xi_1+\xi_2)} d\xi_1 d\xi_2. \label{eq:pseudo:product}\end{aligned}$$ Equivalently, denoting by $K_M(z_1,z_2)= M^{\vee}(z_1,z_2)$ the inverse Fourier transform of $M$ in $\RR^2 \times \RR^2$, we may write $$\begin{aligned} S_M(f_1,f_2)(x) = \intint_{\RR^2\times\RR^2} K_M(x-y_1,x-y_2) f_1(y_1) f_2(y_2) dy_1 dy_2 \label{eq:pseudo:product:real}\end{aligned}$$ for Schwartz functions $f_1,f_2\colon \RR^2\to \RR$. As opposed to [@CoMe1978; @GrTo2002] which consider kernels of Calderón-Zygmund type, here we only need to consider kernels $K_M$ which obey $$\begin{aligned} {\left\lVert|z|^{a} \partial_z^b K_M(z)\right\rVert}_{L^1(\RR^2\times\RR^2)} \leq C_{a,b} \lambda^{|b|-|a|} \label{eq:K:convolution:assumption:multi}\end{aligned}$$ for some $\lambda \geq 1$, all $0 \leq |a|, |b|\leq 1$ and some constants $C_{a,b}>0$. Examples are the kernels defined in . Therefore, in contrast to [@CoMe1978 Theorem 1], for us the boundedness of $S_M$ from $L^{p_1} \times L^{p_2} \to L^p$, where $1/p_1 + 1/p_2 = 1/p$ is automatic, and even includes the case $p_1=p_2=p = \infty$, which is the only case needed in this paper. Instead, here we are interested in the commutator between $S_M$ and $D_{t} = \partial_t + u\cdot\nabla_x$, where $u\colon \RR^2\to\RR^2$, which is divergence free vector field. We have \[lem:multi:Dt:commutator\] Let $\lambda \geq 1$, let $K_M$ be a kernel which obeys , and let $S_M$ be the corresponding bilinear convolution operator given by . Given a smooth divergence free vector field $u \colon \TT^2 \to\RR^2$, we denote $$\begin{aligned} [D_t, S_M](f_1,f_2) = D_t (S_M(f_1,f_2)) - S_M(D_t f_1,f_2) - S_M(f_1, D_t f_2) . \label{eq:comm:Dt:SM}\end{aligned}$$ Then we have $$\begin{aligned} {\left\lVert[D_t, S_M](f_1,f_2)\right\rVert}_{C^{0}} \leq C {\left\lVert\nabla u\right\rVert}_{C^0} {\left\lVertf_1\right\rVert}_{C^0} {\left\lVertf_2\right\rVert}_{C^0} \label{eq:comm:material:convolution:multi}\end{aligned}$$ for some constant $C>0$, which is independent of $\lambda$. Upon taking a material derivative of and using the product rule, we obtain $$\begin{aligned} &D_{t}(S_M(f_1,f_2))(x) - S_M(D_t f_1,f_2)(x) - S_M(f_1,D_t f_2)(x) \notag\\ &= \intint_{\RR^2\times\RR^2} u^j(x) \partial_{x}^j K_M(x-y_1,x-y_2) f(y_1) f(y_2) dy_1 dy_2 \notag\\ &\qquad - \intint_{\RR^2\times\RR^2} K_M(x-y_1,x-y_2) u^j(y_1) \partial_{y_1}^j f(y_1) f(y_2) dy_1 dy_2 \notag\\ &\qquad - \intint_{\RR^2\times\RR^2} K_M(x-y_1,x-y_2) f(y_1) u^j(y_2) \partial_{y_2}^j f(y_2) dy_1 dy_2 \notag\\ &= \intint_{\RR^2\times\RR^2} \left(u^j(x) \partial_{z_1}^j K_M(x-y_1,x-y_2) +u^j(x) \partial_{z_2}^j K_M(x-y_1,x-y_2) \right) f(y_1) f(y_2) dy_1 dy_2 \notag\\ &\qquad + \intint_{\RR^2\times\RR^2} u^j(y_1) \partial_{y_1}^j K_M(x-y_1,x-y_2) f(y_1) f(y_2) dy_1 dy_2 \notag\\ &\qquad + \intint_{\RR^2\times\RR^2} u^j(y_2) \partial_{y_2}^j K_M(x-y_1,x-y_2) f(y_1) f(y_2) dy_1 dy_2 \notag\\ &= \intint_{\RR^2\times\RR^2} \left( \big(u^j(x)-u^j(y_1)\big) \partial_{z_1}^j K_M(x-y_1,x-y_2) + \big(u^j(x) - u^j(y_2) \big) \partial_{z_2}^j K_M(x-y_1,x-y_2) \right) \notag\\ &\qquad \qquad \times f(y_1) f(y_2) dy_1 dy_2 \notag\\ &= \intint_{\RR^2\times\RR^2} \left(\frac{u^j(x)-u^j(y_1)}{x-y_1}\right) (x-y_1)\partial_{z_1}^j K_M(x-y_1,x-y_2) f(y_1) f(y_2) dy_1 dy_2 \notag\\ &\qquad + \intint_{\RR^2\times\RR^2} \left(\frac{u^j(x)-u^j(y_2)}{x-y_2}\right) (x-y_2) \partial_{z_2}^j K_M(x-y_1,x-y_2) f(y_1) f(y_2) dy_1 dy_2.\end{aligned}$$ The lemma now follows from condition with $|a|=|b|=1$. [**Acknowledgments.**]{} TB was supported by the National Science Foundation grant DMS-1600868. SS was supported by the National Science Foundation grant DMS-1301380, and by the Royal Society Wolfson Merit Award. VV was partially supported by the National Science Foundation grant DMS-1514771 and by an Alfred P. Sloan Research Fellowship. [BDLISJ15]{} V. Arnold. Sur la géométrie différentielle des groupes de [L]{}ie de dimension infinie et ses applications à l’hydrodynamique des fluides parfaits. , 16(fasc. 1):319–361, 1966. V.I. Arnold and B.A. Khesin. , volume 125 of [ *Applied Mathematical Sciences*]{}. Springer-Verlag, New York, 1998. J. Azzam and J. Bedrossian. Bounded mean oscillation and the uniqueness of active scalar equations. , 367(5):3095–3118, 2015. H. Bae and R. Granero-Belinch[ó]{}n. Global existence for some transport equations with nonlocal velocity. , 269:197–219, 2015. C. Bardos and E.S. Titi. Loss of smoothness and energy conserving rough weak solutions for the [$3d$]{} [E]{}uler equations. , 3(2):185–197, 2010. J. Bourgain and D. Li. Strong illposedness of the incompressible euler equation in integer c m spaces. , 25(1):1–86, 2015. L. Brillouin. . Dover Publications, New York, N. Y., 1946. T. Buckmaster. . PhD thesis, Universit[ä]{}t Leipzig, 2014. T. Buckmaster. Onsager’s conjecture almost everywhere in time. , 333(3):1175–1198, 2015. T. Buckmaster, C. De Lellis, P. Isett, and L. Sz[é]{}kelyhidi Jr. Anomalous dissipation for [1/5]{}-holder [E]{}uler flows. , 182(1):127–172, 2015. T. Buckmaster, C. De Lellis, and L. Sz[[é]{}]{}kelyhidi, Jr. Dissipative euler flows with onsager-critical spatial regularity. , 69(9):1613–1670, 2016. L.A. Caffarelli and A. Vasseur. Drift diffusion equations with fractional diffusion and the quasi-geostrophic equation. , 171(3):1903–1930, 2010. A.P. Calder[[ó]{}]{}n and A. Zygmund. Singular integrals and periodic functions. , 14:249–271, 1954. A. Castro and D. [C[ó]{}rdoba]{}. Infinite energy solutions of the surface quasi-geostrophic equation. , 225(4):1820–1829, 2010. A. [Castro]{}, D. [C[ó]{}rdoba]{}, and D. Faraco. . , 2016. A. [Castro]{}, D. [C[ó]{}rdoba]{}, and J. [G[ó]{}mez-Serrano]{}. . , 2016. A. Cheskidov, P. Constantin, S. Friedlander, and R. Shvydkoy. Energy conservation and [O]{}nsager’s conjecture for the [E]{}uler equations. , 21(6):1233–1252, 2008. A. Choffrut. -principles for the incompressible [E]{}uler equations. , 210(1):133–163, 2013. A. Choffrut, C. De Lellis, and L. Sz[é]{}kelyhidi Jr. Dissipative continuous Euler flows in two and three dimensions. , 2012. A. Choffrut and L. Sz[é]{}kelyhidi Jr. Weak solutions to the stationary incompressible [E]{}uler equations. , 46(6):4060–4074, 2014. R. Coifman and Y. Meyer. Commutateurs d’intégrales singulières et opérateurs multilinéaires. , 28(3):177–202, 1978. P. Constantin. Scaling exponents for active scalars. , 90(3-4):571–595, 1998. P. Constantin. Energy spectrum of quasigeostrophic turbulence. , 89(18):184501, 2002. P. Constantin. agrangian-[E]{}ulerian methods for uniqueness in hydrodynamic systems. , 287:67–102, 2015. P. Constantin. . CBMS Regional Conference Series in Mathematics, American Mathematical Society, 2016. P. Constantin, W. E, and E.S. Titi. Onsager’s conjecture on the energy conservation for solutions of [E]{}uler’s equation. , 165(1):207–209, 1994. P. Constantin, M.-C. Lai, R. Sharma, Y.-H. Tseng, and J. Wu. New numerical results for the surface quasi-geostrophic equation. , 50(1):1–28, 2012. P. Constantin, A.J. Majda, and E. Tabak. Formation of strong fronts in the [$2$]{}-[D]{} quasigeostrophic thermal active scalar. , 7(6):1495–1533, 1994. P. Constantin, A. Tarfulea, and V. Vicol. Absence of anomalous dissipation of energy in forced two dimensional fluid equations. , 212(3):875–903, 2014. P. Constantin, A. Tarfulea, and V. Vicol. Long time dynamics of forced critical [SQG]{}. , 335(1):93–141, 2015. P. Constantin and V. Vicol. Nonlinear maximum principles for dissipative linear nonlocal operators and applications. , 22(5):1289–1321, 2012. P. Constantin, V. Vicol, and J. Wu. Analyticity of lagrangian trajectories for well posed inviscid incompressible fluid models. , 285:352–393, 2015. P. Constantin and J. Wu. Behavior of solutions of 2[D]{} quasi-geostrophic equations. , 30(5):937–948, 1999. D. C[ó]{}rdoba. Nonexistence of simple hyperbolic blow-up for the quasi-geostrophic equation. , 148(3):1135–1152, 1998. A. C[[ó]{}]{}rdoba and D. C[[ó]{}]{}rdoba. A maximum principle applied to quasi-geostrophic equations. , 249(3):511–528, 2004. D. C[ó]{}rdoba, D. Faraco, and F. Gancedo. Lack of uniqueness for weak solutions of the incompressible porous media equation. , 3:725–746, 2011. D. [C[ó]{}rdoba]{} and C. Fefferman. Growth of solutions for [QG]{} and 2[D]{} [E]{}uler equations. , 15(3):665–670, 2002. D. C[ó]{}rdoba, C. Fefferman, and J.L. Rodrigo. Almost sharp fronts for the surface quasi-geostrophic equation. , 101(9):2687–2691, 2004. D. C[[ó]{}]{}rdoba, M.A. Fontelos, A.M. Mancho, and J.L. Rodrigo. Evidence of singularities for a family of contour dynamics equations. , 102(17):5949–5952, 2005. M. Coti Zelati and V. Vicol. On the global regularity for the supercritical [SQG]{} equation. , 65(2):535–552, 2016. M. Dabkowski. Eventual regularity of the solutions to the supercritical dissipative quasi-geostrophic equation. , 21(1):1–13, 2011. C. De Lellis and L. Sz[[é]{}]{}kelyhidi, Jr. Dissipative [E]{}uler flows and [O]{}nsager’s conjecture. , 05 2012. C. De Lellis and L. Sz[[é]{}]{}kelyhidi, Jr. The [$h$]{}-principle and the equations of fluid dynamics. , 49(3):347–375, 2012. C. De Lellis and L. Sz[[é]{}]{}kelyhidi, Jr. Dissipative continuous [E]{}uler flows. , 193(2):377–407, 2013. C. De Lellis and L. Sz[[é]{}]{}kelyhidi, Jr. High dimensionality and h-principle in PDE. , 2016. S. Daneri and L. Sz[é]{}kelyhidi Jr. Non-uniqueness and $h$-principle for [H]{}ölder-continuous weak solutions of the [E]{}uler equations. , 2016. T.M. Elgindi and N. Masmoudi. Ill-posedness results in critical spaces for some equations arising in hydrodynamics. , 2014. C. Fefferman and J.L. Rodrigo. Analytic sharp fronts for the surface quasi-geostrophic equation. , 303(1):261–288, 2011. C. Fefferman and J.L. Rodrigo. Almost sharp fronts for [SQG]{}: the limit equations. , 313(1):131–153, 2012. F. Gancedo. Existence for the [$\alpha$]{}-patch model and the [QG]{} sharp front in [S]{}obolev spaces. , 217(6):2569–2598, 2008. F. Gancedo and R.M. Strain. Absence of splash singularities for surface quasi-geostrophic sharp fronts and the muskat problem. , 111(2):635–639, 2014. L. Grafakos and R.H. Torres. Multilinear [C]{}alder[ó]{}n–[Z]{}ygmund theory. , 165(1):124–164, 2002. I.M. Held, R.T. Pierrehumbert, S.T. Garner, and K.L. Swanson. Surface quasi-geostrophic dynamics. , 282:1–20, 1995. E. Hölder. über die unbeschränkte [F]{}ortsetzbarkeit einer stetigen ebenen [B]{}ewegung in einer unbegrenzten inkompressiblen [F]{}lüssigkeit. , 37(1):727–738, 1933. D.D. Holm, J.E. Marsden, and T.S. Ratiu. The [E]{}uler-[P]{}oincaré equations and semidirect products with applications to continuum theories. , 137(1):1–81, 1998. P. Isett. Hölder [[Continuous Euler Flows]{}]{} with [[Compact Support]{}]{} in [[Time]{}]{}, 2013. P. Isett. A proof of [O]{}nsager’s conjecture. , 2016. P. Isett and V. Vicol. Hölder continuous solutions of active scalar equations. , 1(1):1–77, 2015. N. Ju. Dissipative 2[D]{} quasi-geostrophic equation: local well-posedness, global regularity and similarity solutions. , 56(1):187–206, 2007. A. Kiselev. Nonlocal maximum principles for active scalars. , 227(5):1806–1826, 2011. A. Kiselev and F. Nazarov. A variation on a theme of [C]{}affarelli and [V]{}asseur. , 370(Kraevye Zadachi Matematicheskoi Fiziki i Smezhnye Voprosy Teorii Funktsii. 40):58–72, 220, 2009. A. Kiselev and F. Nazarov. A simple energy pump for the surface quasi-geostrophic equation. In [*Nonlinear Partial Differential Equations*]{}, pages 175–179. Springer, 2012. A. Kiselev, F. Nazarov, and A. Volberg. Global well-posedness for the critical 2[D]{} dissipative quasi-geostrophic equation. , 167(3):445–453, 2007. S. Klainerman. On [N]{}ash’s unique contribution to analysis in just three of his papers. , 2017. P.G. Lemari[[é]{}]{}-Rieusset. , volume 431 of [*Chapman & Hall/CRC Research Notes in Mathematics*]{}. Chapman & Hall/CRC, Boca Raton, FL, 2002. A.J. Majda and A.L. Bertozzi. , volume 27 of [*Cambridge Texts in Applied Mathematics*]{}. Cambridge University Press, Cambridge, 2002. F. Marchand. Existence and regularity of weak solutions to the quasi-geostrophic equations in the spaces [$L^p$]{} or [$\dot H^{-1/2}$]{}. , 277(1):45–67, 2008. J.E. Marsden and T.S. Ratiu. , volume 17 of [*Texts in Applied Mathematics*]{}. Springer-Verlag, New York, second edition, 1999. A basic exposition of classical mechanical systems. J.E. Marsden, T.S. Ratiu, and S. Shkoller. The geometry and analysis of the averaged [E]{}uler equations and a new diffeomorphism group. , 10(3):582–599, 2000. H. Miura. Dissipative quasi-geostrophic equation for large initial data in the critical [S]{}obolev space. , 267(1):141–157, 2006. C. Muscalu and W. Schlag. . Cambridge University Press, 2013. J. Pedlosky. . Springer Verlag, 1982. H. Poincar[é]{}. Sur une forme nouvelle des [é]{}quations de la m[é]{}canique. , 132:369–371, 1901. M. Puel and A.F. Vasseur. Global weak solutions to the inviscid 3[D]{} [Q]{}uasi-[G]{}eostrophic equation. , 339(3):1063–1082, 2015. S.G. Resnick. . PhD thesis, The University of Chicago, 1995. J.L. Rodrigo. On the evolution of sharp fronts for the quasi-geostrophic equation. , 58(6):821–866, 2005. W. [Rusin]{}. Logarithmic spikes of gradients and uniqueness of weak solutions to a class of active scalar equations. , 2011. R.K. Scott. A scenario for finite-time singularity in the quasigeostrophic model. , 687:492–502, 2011. R. Shvydkoy. Convex integration for a class of active scalar equations. , 24(4):1159–1174, 2011. L. Silvestre. Eventual regularization for the slightly supercritical quasi-geostrophic equation. , 27(2):693–704, 2010. L. Sz[é]{}kelyhidi Jr. Relaxation of the incompressible porous media equation. , 45(3):491–509, 2012. E.M. Stein and G. Weiss. . Princeton University Press, Princeton, N.J., 1971. Princeton Mathematical Series, No. 32. T. Tao. Conserved quantities for the surface quasi-geostrophic equation. 2014. T. Tao. Finite time blowup for [L]{}agrangian modifications of the three-dimensional [E]{}uler equation. , 2016. P. [Washabaugh]{}. . , 222(3):1269–1284, 2016. J. Wu. Global solutions of the 2[D]{} dissipative quasi-geostrophic equation in [B]{}esov spaces. , 36(3):1014–1030 (electronic), 2004/05. [^1]: Courant Institute of Mathematical Sciences, New York University, New York, NY 10012, USA. [<[email protected]>]{}. [^2]: Department of Mathematics, University of California, Davis, CA 95616, USA. [<[email protected]>]{}. [^3]: Department of Mathematics, Princeton University, Princeton, NJ 08544, USA. [<[email protected]>]{}. [^4]: In addition to , writing the SQG equation in Lagrangian coordinates has further geometric advantages. For instance, in [@CoViWu2015] it is shown that the solutions $\theta \in C^0_t C^{1,\alpha}_x$ have Lagrangian trajectories which are real-analytic functions of time. [^5]: One may also consider another class of weak solutions, the so-called patch-solutions, or sharp-fronts [@Ro2005; @Ga2008]. These solutions are given by $\theta = {\bf 1}_{\Omega(t)}$, where $\Omega(t)$ is a compact, simply-connected domain, with smooth boundary, evolving with the fluid. Although these weak solutions are not smooth as functions on $\RR^2$, $\theta \in L^\infty_{\rm loc}(\RR,L^\infty(\RR^2))$, the boundary $\partial \Omega (t)$ is smooth and thus local-in-time existence and uniqueness of such solutions is known. For such patch solutions, an important question is if $\partial\Omega(t)$ can self-intersect in finite time [@CoFoMaRo2005; @Sc2011; @GaSt2014]. See also [@CoFeRodrigo2004; @FeRo2011; @FeRo2012] for the existence of almost-sharp-fronts. [^6]: See also [@DLSz2012; @DLSz2016] for excellent review papers on the applicability of convex integration techniques in fluid dynamics, and connections to the $h$-principle. [^7]: For strong solutions, this is the strongest norm on which we have an a priori global in time bound: for any $\gamma>0$ we have that $\|\theta(t)\|_{L^\infty(\TT^2)}$ decays exponentially [@CoCo2004; @CoTaVi2015]. In fact, the argument in [@CaVa2010] (see [@CZVi2016]) shows that $L^\infty_t L^2_x \cap L^2_t \dot{H}^{\gamma/2}_x$ [*suitable weak solutions*]{} (weak solutions which obey the local energy inequality) are bounded $L^\infty(\TT^2)$ for positive time. Thus, in the class of suitable weak solutions with finite kinetic energy, the $L^\infty$ norm is the most important for a priori control. [^8]: The Hamiltonian $\HH$ is scaling invariant for $\gamma =3/2$. [^9]: The finite-time blowup cannot occur for sufficiently smooth and small initial datum [@CoCo2004; @Wu2004; @Mi2006; @Ju2007], and as $\gamma \to 1^-$ it also cannot happen for datum that is large, but not exceedingly large [@CZVi2016]. [^10]: This solution additionally obeys the energy inequality ${\left\lVert\theta(t)\right\rVert}_{\dot{H}^{-1/2}}^2 + 2 \int_0^t {\left\lVert\theta(s)\right\rVert}_{\dot{H}^{(\gamma-1)/2}}^2 ds \leq {\left\lVert\theta_0\right\rVert}_{\dot{H}^{-1/2}}^2$ for any $t\geq 0$. [^11]: The restriction $\gamma+\beta<2$ is sharp, in the sense that the $C^0_tC^\beta_x$ norm for $\Lambda^{-1} \theta$ is scale invariant precisely when $\gamma + \beta = 2$. We show here that for datum which is supercritical for the scaling of the equations, parabolic smoothing does not hold. [^12]: When $A= (1- \alpha ^2 \Delta )$, $ \alpha >0$, the metric on $ \mathcal{V} $ is equivalent to the $H^1$-metric, and the system corresponds the well-studied Lagrangian Averaged Euler or Euler-$ \alpha $ equations [@HoMaRa1998]. More generally, if $A= (1- \Delta )^s$, $s \in \mathbb{R} $, then is an $H^s$ metric on $ \mathcal{V} $, and it was shown in [@MaRaSh2000 Section 3.3] that the vorticity $ \omega = \nabla ^\perp \cdot u$ satisfies $\partial_t (1- \Delta )^s \omega + u \cdot \nabla (1- \Delta )^s \omega =0$. [^13]: Letting $A$ be a specially chosen order $-2$ operator which is self-adjoint and positive-definite, [@Ta2016] proved that admits locally in time smooth solutions which blow up in finite time. [^14]: Here the subindex $O$ of $\alpha_O$ stands for Onsager, as was suggested in [@Kl2016]. [^15]: Indeed, the construction in [@BuDLeIsSz2015], combined with ideas from [@Ch2013; @ChSz2014], or from this paper, shows that $h$-principles are available for 2-D Euler with velocity fields in $C_t^\alpha C^\alpha_x$, for any $\alpha < 1/5$. The recent construction in [@Is2016] does not apply in two dimensions, since any two non-parallel infinite lines on the plane intersect. [^16]: There is yet another form of the SQG equations which should play an important role in the analysis of smooth solutions, which we write as $$\p_t u + (u\cdot \nabla )u -\Lambda \left( [ \Lambda^{-1} , u^\perp]\, \nabla ^\perp\cdot u \right) = -\nabla P \,, \qquad \operatorname{div} u = 0 \,.$$ This form of SQG is written as a zeroth-order perturbation of the incompressible Euler equations for sufficiently smooth vector-fields $u$. [^17]: This is the reason why in [@IsVi2015], one may only consider active scalar equations with non-odd constitutive laws (that is, the Fourier multiplier relating $u$ to $\theta$ in is a non-odd function of frequency), extending prior results in [@CoFaGa2011; @Sh2011; @Sz2012] for the incompressible porous media equation. See also the recent work [@CaCoFa2016]. [^18]: Equation only defines $\div \mathring{R}_{q+1}$. The Reynolds stress $\mathring R_{q+1}$ itself is obtained from once we invert the divergence operator, cf. Definition \[def:BB\] below, for the contributions that have large frequency, and we write the low frequency part of the oscillation error in divergence form. [^19]: We also denote by $\sum_{ j,k} $ the double sum $\sum_{\{j:\rho_j\neq 0\}} \sum_{k \in \Omega_j}$, and similarly $\sum_{j,j',k,k'}$ denotes a quadruple sum. [^20]: [The constant in is not sharp. See e.g. [@Ma2008] where the constant is given as $\max\left\{ {\left\lVert\nabla \varphi\right\rVert}_{C^0} , {\left\lVert\varphi\right\rVert}_{\dot{B}^{2}_{2,\infty}} \right\}$.]{}
{ "pile_set_name": "ArXiv" }
--- abstract: | This paper introduces a novel error estimator for the Proper Generalized Decomposition (PGD) approximation of parametrized equations. The estimator is intrinsically random: It builds on concentration inequalities of Gaussian maps and an adjoint problem with random right-hand side, which we approximate using the PGD. The effectivity of this randomized error estimator can be arbitrarily close to unity with high probability, allowing the estimation of the error with respect to any user-defined norm as well as the error in some quantity of interest. The performance of the error estimator is demonstrated and compared with some existing error estimators for the PGD for a parametrized time-harmonic elastodynamics problem and the parametrized equations of linear elasticity with a high-dimensional parameter space.\ [**Keywords:**]{} A posteriori error estimation, parametrized equations, Proper Generalized Decomposition, Monte-Carlo estimator, concentration phenomenon, goal-oriented error estimation author: - 'Kathrin Smetana[^1], Olivier Zahm[^2]' title: 'Randomized residual-based error estimators for the Proper Generalized Decomposition approximation of parametrized problems' --- Introduction {#sec1} ============ Parametrized models are ubiquitous in engineering applications, life sciences, hazard event prediction and many other areas. Depending on the application, the parameters may, for instance, represent the permeability of soil, the Young’s modulus field, the geometry but also uncertainty in the data. In general, the goal is either to analyze the model response with respect to the parameter or to perform real-time simulations for embedded systems. In either case, the complexity of the model is the computational bottleneck, especially if computing the model response for one parameter value requires the solution to a complex system of partial differential equations (PDEs). Model order reduction methods such as the Reduced Basis (RB) method [@HeRoSt16; @QuMaNe16; @Haa17; @RoHuPa08; @VePrRoPa03], the Proper Generalized Decomposition (PGD) [@AmMoChKe06; @AmMoChKe07; @Nouy07], the hierarchical model reduction approach [@VogBab81a; @PerErnVen10; @OhlSme14; @SmeOhl17; @BTOS16], or, more generally, tensor-based methods [@Hac12; @Nou17] are commonly used to build a fast-to-evaluate approximation to the model response. This paper is concerned with the PGD, which has been developed for the approximation of models defined in multidimensional spaces, such as problems in quantum chemistry, financial mathematics or (high-dimensional) parametrized models. For the variational version of the PGD, where an energy functional is minimized, convergence has been proved for the Poisson problem in [@LeBLelMad09]. For a review on the PGD method see for instance [@ChiAmmCue2010; @ChLaCu11; @ChKeLe14] and we refer to [@ChiCue14] for an overview of how to use PGD in different applications such as augmented learning and process optimization. The goal of this paper is to introduce a novel error estimator for the PGD approximation. Error estimation is of crucial importance, in particular to stop the PGD enrichment process as soon as the desired precision is reached. To that end, we extend the randomized residual-based error estimator introduced in [@SmZaPa19] to the PGD. The randomized error estimator features several desirable properties: first, it does not require the estimation of stability constants such as a coercivity or inf-sup constant and the effectivity index is close to unity with prescribed lower and upper bounds at high probability. The effectivity index can thus be selected by the user, balancing computational costs and desired sharpness of the estimator. Moreover, the presented framework yields error estimators with respect to user-defined norms, for instance the $L^2$-norm or the $H^1$-norm; the approach also permits error estimation of linear quantities of interest (QoI). Finally, depending on the desired effectivity the computation of the error estimator is in general only as costly as the computation of the PGD approximation or even significantly less expensive, which makes our error estimator strategy attractive from a computational viewpoint. To build this estimator, we first estimate the norm of the error with a Monte Carlo estimator using Gaussian random vectors whose covariance is chosen according to the desired error measure, e.g., user-defined norms or quantity of interest. Then, we introduce a dual problem with random right-hand side the solution of which allows us to rewrite the error estimator in terms of the residual of the original equation. Approximating the random dual problems with the PGD yields a fast-to-evaluate error estimator that has low marginal cost. Next, we relate our error estimator to other stopping criteria that have been proposed for the PGD solution. In [@AmChHu10] the authors suggest letting an error estimator in some linear QoI steer the enrichment of the PGD approximation. In order to estimate the error they employ an adjoint problem with the linear QoI as a right-hand side and approximate this adjoint problem with the PGD as well. It is important to note that, as observed in [@AmChHu10], this type of error estimator seems to require in general a more accurate solution of the dual than the primal problem. In contrast, the framework we present in this paper often allows using a dual PGD approximation with significantly less terms than the primal approximation. An adaptive strategy for the PGD approximation of the dual problem for the error estimator proposed in [@AmChHu10] was suggested in [@GNLC13] in the context of computational homogenization. In [@Alfetal15] the approach proposed in [@AmChHu10] is extended to problems in nonlinear solids by considering a suitable linearization of the problem. An error estimator for the error between the exact solution of a parabolic PDE and the PGD approximation in a suitable norm is derived in [@LadCha11] based on the constitutive relation error and the Prager-Synge theorem. The discretization error is thus also assessed, and the computational costs of the error estimator depend on the number of elements in the underlying spatial mesh. Relying on the same techniques an error estimator for the error between the exact solution of a parabolic PDE and a PGD approximation in a linear QoI was proposed in [@LadCha12], where the authors used the PGD to approximate the adjoint problem. In [@Moi13] the hypercircle theory introduced by Prager and Synge is invoked to derive an error between the exact solution of the equations of linear elasticity and the PGD approximation. While the estimators in [@LadCha11; @LadCha12; @Moi13] are used to certify the PGD approximation, in [@Chetal17] the authors slightly modify the techniques used in [@LadCha11; @LadCha12] to present an error estimator that is used to drive the construction of the PGD approximation within an adaptive scheme. To that end, the authors assume that also the equilibrated flux can be written in a variable separation form. In contrast to the error estimator we introduce in this paper, the estimators presented in [@LadCha11; @LadCha12; @Moi13; @Chetal17] also take into account the discretization error; for the a posteriori error estimator suggested in this paper this is the subject of a forthcoming work. Moreover, in [@LadCha11; @LadCha12; @Moi13; @Chetal17] the authors consider (parametrized) coercive elliptic and parabolic problems; it seems that the techniques just described have not been used yet to address problems that are merely inf-sup stable. Error estimators based on local error indicators are proposed to drive mesh adaptation within the construction of the PGD approximation in [@Nadetal15]. In [@BFNoZa14], an ideal minimal residual formulation is proposed in order to build optimal PGD approximations. In that work, an error estimator with controlled effectivity is derived from the solution of some adjoint problem. Finally, a PGD approximation of the adjoint problem is used to estimate a QoI in [@KCLP19]. The latter is then employed to constrain the primal approximation, aiming at a (primal) PGD approximation more tailored to the considered QoI [@KCLP19]. The randomized a posteriori error estimator proposed in [@SmZaPa19], which we extend in this paper to the PGD, is inspired by the probabilistic error estimator for the approximation error in the solution of a system of ordinary differential equations introduced in [@CaoPet04; @HoPeSe07]. Here, the authors invoke the small statistical sample method from [@KenLau94], which estimates the norm of a vector by its inner product with a random vector drawn uniformly at random on the unit sphere. We note that randomized methods for error estimation are gaining interest in the reduced order modeling community in particular and in the context of parametrized PDEs generally. For instance in [@BalNou18], random sketching techniques are used within the context of the RB method to speed-up the computation of the dual norm of the residual used as an error indicator. Here, the authors exploit that the residual lies in a low-dimensional subspace, allowing them to invoke arguments based on $\varepsilon$-nets in order to certify the randomized algorithm. In [@JaNoPr16] for the RB method a probabilistic a posteriori error bound for linear scalar-valued QoIs is proposed, where randomization is done by assuming that the parameter is a random variable on the parameter set. In [@zahm2016interpolation], an interpolation of the operator inverse is built via a Frobenius-norm projection and computed efficiently using randomized methods. An error estimator is obtained by measuring the norm of residual multiplied by the interpolation of the operator inverse, used here as a preconditioner. Next, in the context of localized model order reduction a reliable and efficient probabilistic a posteriori error estimator for the operator norm of the difference between a finite-dimensional linear operator and its orthogonal projection onto a reduced space is derived in [@BuhSme18]. Finally in the recent paper [@Eetal19] the authors use concentration inequalities to bound the generalization error in the context of statistical learning for high-dimensional parametrized PDEs in the context of Uncertainty Quantification. The remainder of this article is organized as follows. After introducing the problem setting and recalling the PGD in Section \[sec2\], we introduce an unbiased randomized a posteriori error estimator in Section \[sec3\]. As this error estimator still depends on the high-dimensional solutions of dual problems, we suggest in Section \[sec4\] how to construct dual PGD approximations for this setting efficiently and propose an algorithm where the primal and dual PGD approximations are constructed in an intertwined manner. In Section \[sec5\] we compare the effectivity of the randomized a posteriori error estimator with a stagnation-based error estimator and the dual norm of the residual for a time-harmonic elastodynamics problem and a linear elasticity problem with $20$-dimensional parameter space. We close with some concluding remarks in Section \[sec6\]. Problem setting and the Proper Generalized Decomposition {#sec2} ======================================================== We denote by ${\boldsymbol{\mu}}=({\boldsymbol{\mu}}_1,\hdots,{\boldsymbol{\mu}}_p)$ a vector of $p$ parameters which belongs to a set $\mathcal{P}\subset{\mathbb{R}}^p$ of admissible parameters. Here, we assume that $\mathcal{P}$ admits a product structure $\mathcal{P}={\mathcal{P}}_{1} \times\hdots\times {\mathcal{P}}_{p}$, where ${\mathcal{P}}_{i}\subset{\mathbb{R}}$ is the set of admissible parameters for ${\boldsymbol{\mu}}_i$, $i=1,\hdots,p$. For simplicity we assume that ${\mathcal{P}}_{i}=\{{\boldsymbol{\mu}}_i^{(1)},\hdots,{\boldsymbol{\mu}}_i^{(\#{\mathcal{P}}_{i})}\}$ is a finite set of parameters (obtained for instance as an interpolation grid or by a Monte Carlo sample of some probability measure) so that $\mathcal{P}$ is a grid containing $\#{\mathcal{P}}=\prod_{i=1}^d\#{\mathcal{P}}_{i}$ points. Let $D \subset \mathbb{R}^{d}$ be a bounded spatial domain and let $\mathcal{A}({\boldsymbol{\mu}}): \mathcal{X} \rightarrow \mathcal{X}'$ be a parametrized linear differential operator, where $\mathcal{X}$ is a suitable Sobolev space of functions defined on $D$; $\mathcal{X}'$ denotes the dual space of $\mathcal{X}$. We consider the following problem: for any ${\boldsymbol{\mu}}\in \mathcal{P}$ and $\mathcal{f}({\boldsymbol{\mu}}) \in \mathcal{X}'$, find $\mathcal{u}({\boldsymbol{\mu}}) \in \mathcal{X}$ such that $$\label{eq:PDE} \mathcal{A}({\boldsymbol{\mu}}) \mathcal{u}({\boldsymbol{\mu}}) = \mathcal{f}({\boldsymbol{\mu}}) \qquad \text{in } \mathcal{X}'.$$ We ensure well-posedness of problem by requiring that $$\label{eq:inf-sup} \beta({\boldsymbol{\mu}}):=\inf_{v \in \mathcal{X}} \sup_{w \in \mathcal{X}} \frac{\langle \mathcal{A}({\boldsymbol{\mu}}) v, w \rangle_{\mathcal{X}^{'},\mathcal{X}}}{\|v\|_{\mathcal{X}}\|w\|_{\mathcal{X}}} > 0 \qquad\text{and}\qquad \gamma ({\boldsymbol{\mu}}):=\sup_{v \in \mathcal{X}} \sup_{w \in \mathcal{X}} \frac{\langle \mathcal{A}({\boldsymbol{\mu}}) v, w \rangle_{\mathcal{X}^{'},\mathcal{X}}}{\|v\|_{\mathcal{X}}\|w\|_{\mathcal{X}}} < \infty ,$$ hold for all ${\boldsymbol{\mu}}\in \mathcal{P}$. Here $\langle \cdot , \cdot \rangle_{\mathcal{X}^{'},\mathcal{X}}$ denotes the duality pairing of $\mathcal{X}'$ and $\mathcal{X}$. Next, we introduce a finite dimensional space $\mathcal{X}_{N} := \text{span}\{\psi_1,\hdots,\psi_N\}\subset \mathcal{X}$ defined as the span of $N\gg 1$ basis functions $\psi_1,\hdots,\psi_N$. For any $\boldsymbol{\mu} \in \mathcal{P}$, we define the high fidelity approximation of as the function $\mathcal{u}_{N}(\boldsymbol{\mu}) \in \mathcal{X}_{N}$ that solves the variational problem $$\label{eq:PDE_discrete} \langle \mathcal{A}({\boldsymbol{\mu}}) \mathcal{u}_{N}({\boldsymbol{\mu}}) , w_{N} \rangle_{\mathcal{X}^{'},\mathcal{X}} = \langle\mathcal{f}({\boldsymbol{\mu}}) , w_{N}\rangle_{\mathcal{X}^{'},\mathcal{X}} \qquad \text{ for all } w_{N}\in\mathcal{X}_{N}.$$ Again, we ensure wellposed-ness of problem by requiring the discrete inf-sup/sup-sup conditions $$\label{eq:inf-sup_discrete} \beta_{N}({\boldsymbol{\mu}}) := \inf_{v \in \mathcal{X}_{N}} \sup_{w \in \mathcal{X}_{N}} \frac{\langle \mathcal{A}({\boldsymbol{\mu}}) v, w \rangle_{\mathcal{X}^{'},\mathcal{X}}}{\|v\|_{\mathcal{X}}\|w\|_{\mathcal{X}}} > 0 \qquad\text{and}\qquad \gamma_{N} ({\boldsymbol{\mu}}):=\sup_{v \in \mathcal{X}_{N}} \sup_{w \in \mathcal{X}_{N}} \frac{\langle \mathcal{A}({\boldsymbol{\mu}}) v, w \rangle_{\mathcal{X}^{'},\mathcal{X}}}{\|v\|_{\mathcal{X}}\|w\|_{\mathcal{X}}} < \infty ,$$ to hold for all ${\boldsymbol{\mu}}\in \mathcal{P}$. Expressing $\mathcal{u}_{N}({\boldsymbol{\mu}})$ in the basis $\mathcal{u}_{N}({\boldsymbol{\mu}}) = \sum_{i=1}^{N} u_{i}({\boldsymbol{\mu}}) \psi_{i}$ and defining the vector of coefficient $u({\boldsymbol{\mu}}):=(u_{1}({\boldsymbol{\mu}}),\hdots,u_{N}({\boldsymbol{\mu}})) \in {\mathbb{R}}^{N}$, we rewrite as a parametrized algebraic system of linear equations $$\label{eq:AUB} A({\boldsymbol{\mu}}) u({\boldsymbol{\mu}}) = f({\boldsymbol{\mu}}),$$ where the matrix $A({\boldsymbol{\mu}}) \in {\mathbb{R}}^{N\times N}$ and the vector $f({\boldsymbol{\mu}}) \in {\mathbb{R}}^{N}$ are defined as $A_{ij}({\boldsymbol{\mu}}) = \langle \mathcal{A}({\boldsymbol{\mu}})\psi_{j},\psi_{i} \rangle_{\mathcal{X}^{'},\mathcal{X}}$ and $f_i({\boldsymbol{\mu}}) = \langle \mathcal{f}({\boldsymbol{\mu}}),\psi_{i} \rangle_{\mathcal{X}^{'},\mathcal{X}}$ for any $1\leq i,j\leq N$. Finally, with a slight abuse of notation, we denote by $\|\cdot\|_{\mathcal{X}_N}$ and $\|\cdot\|_{\mathcal{X}_N'}$ the norms defined on the algebraic space ${\mathbb{R}}^N$ such that $$\label{eq:NormsDefinition} \|v\|_{\mathcal{X}_N} := \| \sum_{i=1}^N v_i \psi_i \|_\mathcal{X}, \qquad\text{and}\qquad \|v\|_{\mathcal{X}_N'} := \sup_{w \in \mathcal{X}_{N}} \frac{\langle \sum_{i=1}^N v_i \psi_i, w \rangle_{\mathcal{X}^{'},\mathcal{X}}}{\|w\|_{\mathcal{X}}} \qquad \text{ for any } v\in{\mathbb{R}}^N.$$ These norms can be equivalently defined by $\|v\|_{\mathcal{X}_N}^2 = v^TR_{{\mathcal{X}}}v$ and $\|v\|_{\mathcal{X}_N'}^2 = v^TR_{{\mathcal{X}}}^{-1}v$ for any $v\in{\mathbb{R}}^N$, where $R_{{\mathcal{X}}}\in{\mathbb{R}}^{N\times N}$ is the gram matrix of the basis $(\psi_1,\hdots,\psi_N)$ such that $(R_{{\mathcal{X}}})_{ij}=(\psi_i,\psi_j)_\mathcal{X}$ where $(\cdot,\cdot)_\mathcal{X}$ is the scalar product in ${\mathcal{X}}$. The Proper Generalized Decomposition {#subsec:PGD} ------------------------------------ To avoid the curse of dimensionality we compute an approximation of $u({\boldsymbol{\mu}})=u({\boldsymbol{\mu}}_1,\hdots,{\boldsymbol{\mu}}_p) \in{\mathbb{R}}^N$ using the Proper Generalized Decomposition (PGD). The PGD relies on the separation of the variables ${\boldsymbol{\mu}}_1,\hdots,{\boldsymbol{\mu}}_p$ to build an approximation of the form $$\label{eq:tensor_appr} \widetilde{u}^{M}({\boldsymbol{\mu}}_{1},\hdots ,{\boldsymbol{\mu}}_{p}) = \sum_{m=1}^{M} u^{m}_{\mathcal{X}} ~ u^{m}_{1}({\boldsymbol{\mu}}_1) \cdots u^{m}_{p}({\boldsymbol{\mu}}_p),$$ where the $M$ vectors $u^{m}_{\mathcal{X}} \in{\mathbb{R}}^N$ and the $p\times M$ scalar univariate functions $u^{m}_{i}:{\mathcal{P}}_i\rightarrow{\mathbb{R}}$ need to be constructed. An approximation as in can be interpreted as a low-rank tensor approximation in the canonical format, see [@Nou17]. With this interpretation, $M$ is called the canonical rank of $\widetilde{u}^{M}$ seen as a tensor $\widetilde{u}^{M} = \sum_{m=1}^{M} u^{m}_{\mathcal{X}} \otimes u^{m}_{1} \otimes\hdots\otimes u^{m}_{p}$. Let us note that if the space of spatial functions $\mathcal{X}$ was a tensor product space $\mathcal{X}=\mathcal{X}^1\otimes\hdots\otimes \mathcal{X}^d$, then one could have further separated the spatial variables and obtained approximations like $\widetilde{u}^{M} = \sum_{m=1}^{M} u^{m}_{\mathcal{X}^1} \otimes\hdots\otimes u^{m}_{\mathcal{X}^d} \otimes u^{m}_{1} \otimes\hdots\otimes u^{m}_{p}$. This idea is however not pursued in this paper. The PGD is fundamentally an $\ell^2$-based model order reduction technique in the sense that it aims at constructing $\widetilde{u}^{M}$ such that the root mean squared (RMS) error $$\label{eq:IdealPGDError} \|u-\widetilde{u}^M \| := \sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|u({\boldsymbol{\mu}})-\widetilde{u}^M({\boldsymbol{\mu}})\|_{\mathcal{X}_N}^2 },$$ is controlled. However, minimizing this error over the functions $\widetilde{u}^{M}$ as in is not feasible in practice. We now recall the basic ingredients of the PGD for the computation of $\widetilde{u}^M({\boldsymbol{\mu}})$, referring for instance to [@ChiAmmCue2010; @ChLaCu11; @ChKeLe14] for further details. ### Progressive PGD {#progressive-pgd .unnumbered} The progressive PGD relies on a greedy procedure where we seek the $M$ terms in one after another. After $M-1$ iterations, we build a rank-one correction $u^{M}_{\mathcal{X}} ~ u^{M}_{1}({\boldsymbol{\mu}}_1) \hdots u^{M}_{p}({\boldsymbol{\mu}}_p)$ of the current approximation $\widetilde{u}^{M-1}({\boldsymbol{\mu}})$ by minimizing some error functional described below. Once this correction is computed, we update the approximation $$\label{eq:GreedyPGD} \widetilde{u}^{M}({\boldsymbol{\mu}}) = \widetilde{u}^{M-1}({\boldsymbol{\mu}}) + u^{M}_{\mathcal{X}} ~ u^{M}_{1}({\boldsymbol{\mu}}_1) \cdots u^{M}_{p}({\boldsymbol{\mu}}_p),$$ and we increment $M\leftarrow M+1$. Such a procedure is called *pure greedy* because it never updates the previous iterates. We now describe how to construct the rank-one correction. A commonly used error functional is the mean squared residual norm, which yields the so-called *Minimal Residual PGD* [@BFNoZa14; @FHMM13; @CaEhLe13]. The correction $u^{M}_{\mathcal{X}} ~ u^{M}_{1}({\boldsymbol{\mu}}_1) \cdots u^{M}_{p}({\boldsymbol{\mu}}_p)$ is defined as a solution to $$\label{eq:MinResPGD} \min_{u_{\mathcal{X}}\in{\mathbb{R}}^N} \min_{u_{1}:\mathcal{P}_1\rightarrow{\mathbb{R}}} \cdots \min_{u_{p}:\mathcal{P}_p\rightarrow{\mathbb{R}}} ~ \sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|A({\boldsymbol{\mu}})\Big(\widetilde{u}^{M-1}({\boldsymbol{\mu}}) + u_{\mathcal{X}} u_{1}({\boldsymbol{\mu}}_1) \cdots u_{p}({\boldsymbol{\mu}}_p)\Big) - f({\boldsymbol{\mu}})\|_{2}^2,$$ where $\|\cdot\|_2$ denotes the canonical norm of ${\mathbb{R}}^N$. When the matrix $A({\boldsymbol{\mu}})$ is symmetric definite positive (SPD), we can alternatively use the *Galerkin PGD* which consists in minimizing the mean squared energy $$\label{eq:MinEnergyPGD} \min_{u_{\mathcal{X}}\in{\mathbb{R}}^N} \min_{u_{1}:\mathcal{P}_1\rightarrow{\mathbb{R}}} \cdots \min_{u_{p}:\mathcal{P}_p\rightarrow{\mathbb{R}}} ~ \sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|A({\boldsymbol{\mu}})\Big(\widetilde{u}^{M-1}({\boldsymbol{\mu}}) + u_{\mathcal{X}} u_{1}({\boldsymbol{\mu}}_1) \cdots u_{p}({\boldsymbol{\mu}}_p)\Big) - f({\boldsymbol{\mu}})\|_{A({\boldsymbol{\mu}})^{-1}}^2 ,$$ where the residual is measured with the norm $\|\cdot\|_{A({\boldsymbol{\mu}})^{-1}} = \sqrt{(\cdot)^TA({\boldsymbol{\mu}})^{-1}(\cdot)}$. The Galerkin PGD is usually much more efficient compared to the Minimal Residual PGD, but the latter is much more stable for non SPD operators. Both for the Minimal Residual PGD and the Galerkin PGD Conditions are sufficient to ensure the convergence of the pure greedy algorithm outlined above, see [@BFNoZa14]. A simple and yet efficient approach to solve or is the *alternating least squares* (ALS) method. It relies on the fact that each of the $p+1$ minimization problems is a least squares problem (the cost function being quadratic with respect to each of the $u_\mathcal{X},u_1,u_2,\hdots$) and thus is simple to solve. ALS consists in iteratively solving each of these least squares problems one after the other while keeping the other variables fixed. That is: starting with a random initialization $u_{\mathcal{X}}^{(0)}, u_{1}^{(0)}, \hdots, u_{p}^{(0)}$, we first compute $u_{\mathcal{X}}^{(1)}$ solution to (or to ) with $u_{1}=u_{1}^{(0)}, \hdots, u_{p}=u_{p}^{(0)}$ fixed. Then we compute $u_{1}^{(1)}$ solution to (or to ) with $u_{\mathcal{X}}=u_{\mathcal{X}}^{(1)}, u_{2}=u_{2}^{(0)}, \hdots, u_{p}=u_{p}^{(0)}$ fixed etc. This yields a sequence of updates $u_{\mathcal{X}}^{(1)}\rightarrow u_{1}^{(1)}\rightarrow\cdots\rightarrow u_{p}^{(1)}\rightarrow u_{\mathcal{X}}^{(2)}\rightarrow u_{1}^{(2)} \rightarrow\cdots$ which we stop once the stagnation is reached. In practice, few iterations (three or four) suffice. ### Stopping criterion for the greedy enrichment process {#stopping-criterion-for-the-greedy-enrichment-process .unnumbered} Ideally we want to stop the greedy iteration process in $M$ as soon as the relative RMS error $\|u-\widetilde{u}^{M}\|/\|u\|$ falls below a certain prescribed tolerance, which is why an error estimator for the relative RMS error is desirable. In *stagnation-based error estimators* the unknown solution $u$ is replaced by a reference solution $\widetilde{u}^{M+k}$ for some $k\geq1$ $$\label{eq:StagnationErrorEstimator} \Delta_{M,k}^\text{stag} := \frac{\|\widetilde{u}^{M+k}-\widetilde{u}^{M}\|}{\|\widetilde{u}^{M+k}\|},$$ where the norm $\|\cdot\|$ is defined as in . The estimator is also often called a *hierarchical error estimator*, considering two different PGD approximations: a coarse one $\widetilde{u}^{M}$ and a fine one $\widetilde{u}^{M+k}$. The quality of such an error estimator depends on the convergence rate of the PGD algorithm: a slow PGD convergence would require choosing a larger $k$ to obtain an accurate estimate of the relative RMS error. Needless to say that the error estimator $\Delta_{M,k}^\text{stag}$ is more expensive to compute when $k\gg1$. Proving that $\Delta_{M,k}^\text{stag}$ is an upper bound of the error and bounding the effectivity typically requires satisfying a so-called saturation assumption, i.e., having a uniform bound on the convergence rate of the PGD, a requirement which is hard to ensure in practice. As one example of a *residual-based error estimator* we consider the relative RMS residual norm $$\label{eq:RelativeResNorm} \Delta_{M}^\text{res} := \sqrt{\frac{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|A({\boldsymbol{\mu}})\widetilde{u}^M({\boldsymbol{\mu}})-f({\boldsymbol{\mu}})\|_{\mathcal{X}_N'}^2}{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|f({\boldsymbol{\mu}})\|_{\mathcal{X}_N'}^2 }},$$ where the norm $\|\cdot\|_{\mathcal{X}_N'}$ has been defined in . The advantage of this error estimator is that it comes with the guarantee that $$\label{eq:RelativeResNormEffectivity} \kappa_N^{-1} \frac{\|u-\widetilde{u}^{M}\|}{\|u\|} \leq \Delta_{M}^\text{res} \leq \kappa_N \frac{\|u-\widetilde{u}^{M}\|}{\|u\|} ,$$ for any $M$, where $\kappa_N = \max_{{\boldsymbol{\mu}}\in\mathcal{P}} \gamma_N({\boldsymbol{\mu}})/\beta_N({\boldsymbol{\mu}}) \geq1$ denotes the supremum of the condition number of $A({\boldsymbol{\mu}})$ seen as an operator from $({\mathbb{R}}^N,\|\cdot\|_{\mathcal{X}_N})$ to $({\mathbb{R}}^N,\|\cdot\|_{\mathcal{X}_N'})$. This inequality can be easily proved using Assumption and Definition . Inequality ensures that the effectivity of $\Delta_{M}^\text{res}$ is contained in $[\kappa_N^{-1},\kappa_N]$. Note however that estimating $\kappa_N$ (or a sharp upper bound of it) is difficult in practice as it requires estimating the inf-sup/sup-sup constants $\beta_N({\boldsymbol{\mu}})$ and $\gamma_N({\boldsymbol{\mu}})$, a task that is general rather costly [@Hetal07; @Cetal09; @Hetal10]. Moreover, for ill-conditioned problems as, say, the Helmholtz equation or a time-harmonic elastodynamics problem presented in section \[sec5\], the constant $\beta_N({\boldsymbol{\mu}})$ can be arbitrarily close to zero, yielding a very large $\kappa_N$. In [@AmChHu10] the authors suggest letting an estimator for the error in some linear QoI steer the enrichment of the PGD approximation. In order to estimate the error they employ an adjoint problem with the linear QoI as a right-hand side and approximate this adjoint problem with the PGD as well. In [@LadCha11] the constitutive relation error and the Prager-Synge theorem is employed to derive an error estimator for the error between the exact solution of the PDE and the PGD approximation in a suitable norm. The error estimator requires the construction of equilibrated fluxes and has been proposed to use within an adaptive scheme to construct the PGD approximation in [@Chetal17]. A randomized a posteriori error estimator {#sec3} ========================================= In this paper we aim at estimating the errors $$\label{eq:error} \|u({\boldsymbol{\mu}})-\widetilde{u}^M({\boldsymbol{\mu}})\|_{\Sigma}, \qquad \sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|u({\boldsymbol{\mu}})-\widetilde{u}^M({\boldsymbol{\mu}})\|_{\Sigma}^2}, \qquad\text{or}\qquad \sqrt{\frac{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|u({\boldsymbol{\mu}})-\widetilde{u}^M({\boldsymbol{\mu}})\|_{\Sigma}^2}{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|u({\boldsymbol{\mu}})\|_{\Sigma}^2}},$$ where $\Sigma\in\mathbb{R}^{N\times N}$ is a symmetric positive semi-definite matrix which is chosen by the user according to the desired error measure, see subsection \[subsec:3.1\]. This means that $\|\cdot\|_\Sigma = \sqrt{(\cdot)^T\Sigma(\cdot)}$ is a semi-norm unless $\Sigma$ is SPD, in which case $\|\cdot\|_\Sigma$ is a norm. We show in subsection \[subsec:estimateNorm\] that, at high probability, we can estimate accurately $\|v\|_\Sigma$ for any $v\in{\mathbb{R}}^N$ by means of *random Gaussian maps*. In subsection \[subsec:estimate many vectors\] we use union bound arguments to estimate the errors $\|u({\boldsymbol{\mu}})-\widetilde{u}^M({\boldsymbol{\mu}})\|_{\Sigma}$ and the norms $\|u({\boldsymbol{\mu}})\|_{\Sigma}$ simultaneously for a finite number of parameters ${\boldsymbol{\mu}}$, which allows to estimate all quantities in . Choice of the error measure {#subsec:3.1} --------------------------- If we wish to estimate the error in the natural norm $\|\cdot\|_{\Sigma}=\|\cdot\|_{\mathcal{X}_N}$ defined in , we simply choose $\Sigma=R_{{\mathcal{X}}}$, where we recall that $(R_{{\mathcal{X}}})_{ij}=(\psi_i,\psi_j)_\mathcal{X}$ is the gram matrix of the basis $(\psi_1,\hdots,\psi_N)$. Alternatively, we can consider the error in the $L^{2}(D)$-norm by letting $\Sigma=R_{L^{2}(D)}$ where the matrix $R_{L^{2}(D)}$ is defined by $(R_{L^{2}(D)})_{ij}=(\psi_i,\psi_j)_{L^{2}(D)}$. The presented framework also encompasses estimating the error in some vector-valued quantity of interest defined as a linear function of $u({\boldsymbol{\mu}})$, say $$s({\boldsymbol{\mu}})=L u({\boldsymbol{\mu}}) ~ \in\mathbb{R}^m ,$$ for some extractor $L\in\mathbb{R}^{m \times N}$. In this context we are interested in estimating $\| s({\boldsymbol{\mu}}) - L\,\widetilde u({\boldsymbol{\mu}}) \|_{\mathcal{W}}$ where $\|\cdot\|_{\mathcal{W}}$ is some norm associated with a SPD matrix $R_{\mathcal{W}}\in{\mathbb{R}}^{m\times m}$. For example, one can be interested in computing the approximation error over a subdomain $D_{sub} \subsetneq D$, that is, estimating the error ${\mathcal{u}}({\boldsymbol{\mu}})|_{D_{sub}} - \widetilde {\mathcal{u}}({\boldsymbol{\mu}})|_{D_{sub}}$ measured in the $L^2(D_{sub})$-norm. In this situation, $L$ would extract the relevant entries from the vector ${\widetilde{u}(\boldsymbol{\mu})}$ and $R_{\mathcal{W}}$ would be defined by $(R_{\mathcal{W}})_{ij} = (\psi_i|_{D_{sub}},\psi_j|_{D_{sub}})_{L^{2}(D_{sub})}$. With the choice $\Sigma = L^T R_W L$ we can write $$\begin{aligned} \| u({\boldsymbol{\mu}}) - \widetilde u({\boldsymbol{\mu}}) \|_\Sigma^2 &= (u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}))^T \big( L^T R_W L \big) (u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}})) = \| s({\boldsymbol{\mu}}) - L\widetilde u({\boldsymbol{\mu}}) \|_{\mathcal{W}}^2 ,\end{aligned}$$ so that measuring the error with respect to the norm $\|\cdot\|_\Sigma$ gives the error associated with the QoI. Notice that if $m<N$ the matrix $\Sigma$ is necessarily singular and then $\|\cdot\|_\Sigma$ is a semi-norm. Finally, consider the scalar-valued QoI given by $s({\boldsymbol{\mu}})= l^T u({\boldsymbol{\mu}})$ where $l\in\mathbb{R}^N$. This corresponds to the previous situation with $m=1$ and $L=l^T$. The choice $\Sigma = l\,l^T $ yields $\| u({\boldsymbol{\mu}}) - \widetilde u({\boldsymbol{\mu}}) \|_\Sigma^2= |s({\boldsymbol{\mu}}) - L\widetilde u({\boldsymbol{\mu}}) |$, where $|\cdot|$ denotes the absolute value. Estimating norms using Gaussian maps {#subsec:estimateNorm} ------------------------------------ Random maps have proved themselves to be a very powerful tool to speed up computations in numerical linear algebra. They can also be exploited to estimate the norm of a vector very efficiently. In this subsection, we will present the required basics to first obtain an estimator for the norm of a fixed vector and then for a set of vectors. The latter is for instance of relevance if we want to compute the error . For a thorough introduction on how to estimate certain quantities in numerical linear algebra by random maps we refer to the recent monograph [@Ver18]; we follow subsection 2.2 of [@SmZaPa19] here. First, we will define a suitable random map $\Phi: \mathbb{R}^{N} \rightarrow {\mathbb{R}}^{K}$ with $K\ll N$ that can be used to estimate the (semi-)norm $\|\cdot\|_\Sigma$ by $\|\Phi\cdot\|_2$, where $\|\cdot\|_2$ is the Euclidean norm in $\mathbb{R}^K$. To that end, let $Z\sim\mathcal{N}(0,\Sigma)$ be a zero mean Gaussian random vector in $\mathbb{R}^N$ with covariance matrix $\Sigma\in\mathbb{R}^{N\times N}$. We emphasize that the user first selects the desired error measure which yields the (semi-)norm $\|\cdot\|_\Sigma$ and, subsequently, the covariance matrix is chosen as that very same matrix. Given a vector $v\in\mathbb{R}^N$ we can write $$\|v\|_\Sigma^2 = v^T \Sigma v = v^T \mathbb{E}( ZZ^T )v = \mathbb{E}( (Z^T v)^2 ),$$ where $\mathbb{E}(\cdot)$ denotes the expected value. This means that $(Z^T v)^2$ is an unbiased estimator of $\|v\|_\Sigma^2$. We may then use $K$ independent copies of $Z$, which we denote by $Z_1,\hdots,Z_K$, to define the random map $\Phi\in\mathbb{R}^{K\times N}$ as $\Phi_{(i,:)}=(1/\sqrt{K})Z_i^T$, where $(i,:)$ denotes the $i$-th row. Then, we can write $$\label{eq:defPhi} \| \Phi v \|_2^2 = \frac{1}{K}\sum_{i=1}^K ( Z_i^T v )^2 \qquad \text{for any } v\in \mathbb{R}^N.$$ We see that $\| \Phi v \|_2^2$ is a $K$-sample Monte-Carlo estimator of $\mathbb{E}( (Z^T v )^2 ) = \|v\|_\Sigma^2$. By the independence of the $Z_i$’s, the variance of $\| \Phi v \|_2^2$ is $\operatorname{Var}( \| \Phi v \|_2^2 ) = \frac{1}{K} \operatorname{Var}( (Z^Tv)^2 )$. Increasing $K$ thus allows to reduce the variance of the estimate, and we get a more precise approximation of $\|v\|_\Sigma^2$. However, the variance is not always the best criterion to access the quality of the estimator. Given a vector $v$ and a realization of $\Phi$, how much $\| \Phi v \|_2^2$ underestimates or overestimates $\|v\|_\Sigma^2$? Because $\Phi$ is a random variable, the answer to that question will hold only up to a certain probability which we need to quantify. To that end, we first note that the random variables $(Z_i^T v ) / \|v\|_\Sigma$ for $i=1,\hdots,K$ are independent standard normal random variables so that we have $$\label{eq:chi-squared} \| \Phi v \|_2^2 = \frac{\|v\|_\Sigma^2}{K} \sum_{i=1}^K \Big( \frac{Z_i^T v}{\|v\|_\Sigma}\Big)^2 \sim \frac{\|v\|_\Sigma^2}{K} Q ,$$ where $Q \sim \chi^2(K)$ is a random variable which follows a chi-squared distribution with $K$ degrees of freedom. Denoting by $\mathbb{P}\{A\}$ the probability of an event $A$ and by $\overline{A}$ the complementary event of $A$, the previous relation yields $$\begin{aligned} \mathbb{P}\Big\{ w^{-1}\|v\|_\Sigma \leq \| \Phi v \|_2 \leq w \|v\|_\Sigma \Big\} &=1-\mathbb{P}\big\{ \overline{ Kw^{-2} \leq Q \leq Kw^2 } \big\} ,\end{aligned}$$ for any $w\geq1$. Then for any given (fixed) vector $v\in \mathbb{R}^N$, the probability that a realization of $\| \Phi v \|_2$ lies between $w^{-1}\|v\|_\Sigma$ and $w\|v\|_\Sigma$ is independent of $v$ but also independent of the dimension $N$. The following proposition gives an upper bound for the failure probability $\mathbb{P}\big\{ \overline{ Kw^{-2} \leq Q \leq Kw^2 } \big\} $ in terms of $w$ and $K$. \[prop:Chi2Tail\] Let $Q\sim\chi^2(K)$ be a chi-squared random variable with $K\geq3$ degrees of freedom. For any $w>\sqrt{e}$ we have $$\label{eq:Chi2Tail} \mathbb{P}\big\{ \overline{ Kw^{-2} \leq Q \leq Kw^2 } \big\} \leq \Big( \frac{\sqrt{e}}{w} \Big)^{K} .$$ The proof relies on the fact that we have closed form expressions for the law of $Q\sim\chi^2(K)$; for details see [@SmZaPa19]. Proposition \[prop:Chi2Tail\] shows that the probability $\mathbb{P}\big\{ \overline{ Kw^{-2} \leq Q \leq Kw^2 } \big\}$ decays at least exponentially with respect to $K$, provided $w\geq\sqrt{e}$ and $K\geq3$. Then for any $v\in \mathbb{R}^N$, the relation $$\label{eq:controlError} w^{-1}\|v\|_\Sigma \leq \| \Phi v \|_2 \leq w \|v\|_\Sigma ,$$ holds with a probability greater than $1-(\sqrt{e}/w)^K$. We highlight that this is a non-asymptotic result in the sense that it holds for finite $K$, not relying on the law of large numbers; for further details on those types of non-asymptotic concentration inequalities we refer to [@Ver18]. Moreover, note that by accepting a larger $w$ and thus a larger deviation of the estimator from $\| v \|_{\Sigma}$ fewer samples are required to ensure that the probability of failure $(\sqrt{e}/w)^K$ is small. For instance with $w=2$ and $K=48$, relation holds with a probability larger than $0.9998$, while for $w=4$ we only need $K=11$ samples to obtain the same probability (see Table \[tab:concentration\_set\]). Next, by combining Proposition \[prop:Chi2Tail\] with a union bound argument we can quantify the probability that relation holds simultaneously for any vector in a finite set $\mathcal{M}\subset \mathbb{R}^N$. \[prop:estimate many vectors\] Given a finite collection of vectors $\mathcal{M}=\{v_1,v_2,\hdots, v_{\# \mathcal{M}}\}\subset {{\color{black} \mathbb{R}^N}}$ and a failure probability $0<\delta<1$. Then, for any $w > \sqrt{e}$ and $$\label{eq:ConditionK_FiniteSet} K\geq \min \left\lbrace \frac{\log( \# \mathcal{M}) + \log(\delta^{-1})}{\log(w/\sqrt{e})} , \enspace 3 \right \rbrace$$ we have $$\label{eq:estimate many vectors} \mathbb{P}\Big\{ w^{-1}\|v\|_\Sigma \leq \| \Phi v \|_2 \leq w \|v\|_\Sigma ~,~\forall v\in\mathcal{M} \Big\} \geq 1 - \delta.$$ Table \[tab:concentration\_set\] gives numerical values of $K$ that satisfy depending on $\delta$, $w$ and $\#\mathcal{M}$. For example with $\delta=10^{-4}$ and $w=10$, estimating simultaneously the norm of $10^{9}$ vectors requires only $K=17$ samples. Again, we highlight that this result is independent on the dimension $N$ of the vectors to be estimated. Moreover, Condition allows reducing the number of samples $K$ by increasing $w$[^3]. Note that the computational complexity of the estimator $\| \Phi v \|_2$ and thus the a posteriori error estimator we propose in this paper depends on $K$ (see section \[sec4\]). By relying on Corollary \[prop:estimate many vectors\] the user may thus choose between a less accurate but also less expensive estimator (higher $w$, lower $K$) or a more accurate but also more expensive estimator (lower $w$, higher $K$). [| l | c c c | c | l | c c c |]{} & $w=2$ & $w=4$ & $w=10$ & & & $w=2$ & $w=4$ & $w=10$\ $\#\mathcal{M}=10^0$ & 24 & 6 & 3 & & $\#\mathcal{M}=10^0$ & 48 & 11 & 6\ $\#\mathcal{M}=10^3$ & 60 & 13 & 7 & & $\#\mathcal{M}=10^3$ & 84 & 19 & 9\ $\#\mathcal{M}=10^6$ & 96 & 21 & 11 & & $\#\mathcal{M}=10^6$ & 120 & 26 & 13\ $\#\mathcal{M}=10^9$ & 132& 29 & 15 & & $\#\mathcal{M}=10^9$ & 155 & 34 & 17\ In actual practice we can draw efficiently from $Z\sim\mathcal{N}(0,\Sigma)$ using *any* factorization of the covariance matrix of the form of $\Sigma=U^{T}U$, for instance a Cholesky decomposition or a symmetric matrix square root. It is then sufficient to draw a standard Gaussian vector $\widehat{Z}$ and to compute the matrix-vector product $Z=U^{T}\widehat{Z}$. As pointed-out in [@BalNou18 Remark 2.9], one can take advantage of a potential block structure of $\Sigma$ to build a (non-square) factorization $U$ with a negligible computational cost. Randomized a posteriori error estimator {#subsec:estimate many vectors} --------------------------------------- We apply the methodology described in the previous subsection to derive residual-based randomized a posteriori error estimators. Recall the definition of the random map $\Phi: {\mathbb{R}}^{N} \rightarrow {\mathbb{R}}^{K}$, $\Phi_{(i,:)}=(1/\sqrt{K})Z_i^T$, $i=1,\hdots,K$, where $(i,:)$ denotes the $i$-th row and $Z_1,\hdots Z_K$ are independent copies of $Z\sim\mathcal{N}(0,\Sigma)$. We may then define the estimators $$\label{defeq:error est} \Delta({\boldsymbol{\mu}}) = \| \Phi \big( u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \big) \|_2 = \sqrt{ \frac{1}{K} \sum_{k=1}^{K} \Big(Z_{i}^{T} \big(u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \big) \Big)^{2} } \qquad\text{and}\qquad \Delta = \sqrt{ \frac{1}{\#\mathcal{P}} \sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \Delta({\boldsymbol{\mu}})^{2} }$$ of the errors $\|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}})\|_\Sigma $ and $\sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}})\|_\Sigma^2}$, respectively. As both $\Delta({\boldsymbol{\mu}})$ and $\Delta$ require the high-dimensional solution $u({\boldsymbol{\mu}})$ of problem , these estimators are computationally infeasible in practice. By introducing the residual $$\label{eq:PrimalResidual} r({\boldsymbol{\mu}})= f({\boldsymbol{\mu}}) - A({\boldsymbol{\mu}})\widetilde u({\boldsymbol{\mu}}),$$ associated with Problem \[eq:AUB\] and by exploiting the *error residual relationship* we rewrite the terms $Z_{i}^{T} (u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}}))$ in as $$\label{eq:err res relationship} Z_{i}^{T} (u({\boldsymbol{\mu}}) - \widetilde u({\boldsymbol{\mu}})) = Z_{i}^{T} A({\boldsymbol{\mu}})^{-1} r({\boldsymbol{\mu}}) = (A({\boldsymbol{\mu}})^{-T}Z_{i})^{T}r({\boldsymbol{\mu}}) = Y_i({\boldsymbol{\mu}})^{T}r({\boldsymbol{\mu}}),$$ where $Y_{i}({\boldsymbol{\mu}}) \in \mathbb{R}^N$ are the solutions to the *random dual problems* $$\label{eq:randomDualProblem} A({\boldsymbol{\mu}})^T Y_i({\boldsymbol{\mu}}) = Z_i, \quad 1\leq i \leq K.$$ Because the right hand side in is random, the solutions $Y_1({\boldsymbol{\mu}}),\hdots,Y_K({\boldsymbol{\mu}})$ are random vectors. Using relation the error estimators $\Delta({\boldsymbol{\mu}})$ and $\Delta$ can be rewritten as $$\Delta({\boldsymbol{\mu}}) = \sqrt{ \frac{1}{K} \sum_{k=1}^{K} \Big( Y_i({\boldsymbol{\mu}})^T r({\boldsymbol{\mu}}) \Big)^2} \qquad\text{and}\qquad \Delta = \sqrt{ \frac{1}{\#\mathcal{P}} \sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \Delta({\boldsymbol{\mu}})^{2} }. \label{eq:DualTrick}$$ This shows that $\Delta({\boldsymbol{\mu}})$ can be computed by applying $K$ linear forms to the residual $r({\boldsymbol{\mu}})$. In that sense, $\Delta({\boldsymbol{\mu}})$ can be interpreted as an a posteriori error estimator. However, computing the solutions to is in general as expensive as solving the primal problem . In the next section, we show how to approximate the dual solutions $Y_1({\boldsymbol{\mu}}),\hdots,Y_K({\boldsymbol{\mu}})$ via the PGD in order to obtain a fast-to-evaluate a posteriori error estimator. Before that, a direct application of Corollary \[prop:estimate many vectors\] with $\mathcal{M}=\{u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) : {\boldsymbol{\mu}}\in\mathcal{P}\}$ gives the following control of the quality of $\Delta({\boldsymbol{\mu}})$ and $\Delta$. \[coro:truth est S\] Given $0 < \delta < 1$ and $w > \sqrt{e}$, the condition $$\label{coro:truth est Condition S} K\geq \min \left\lbrace \frac{\log( \# \mathcal{P}) + \log(\delta^{-1})}{\log(w/\sqrt{e})} , \enspace 3 \right \rbrace,$$ is sufficient to ensure $$\mathbb{P}\Big\{ w^{-1} \Delta({\boldsymbol{\mu}}) \leq \|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \|_\Sigma \leq w \Delta({\boldsymbol{\mu}}) ~,~\forall {\boldsymbol{\mu}}\in \mathcal{P} \Big\} \geq 1 - \delta,$$ and therefore $$\mathbb{P}\Big\{ w^{-1} \Delta \leq \sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}})\|_\Sigma^2} \leq w \Delta \Big\} \geq 1 - \delta.$$ It is important to note that Condition depends only on the cardinality of $\mathcal{P}$. This means that $K$ can be determined only knowing the number of parameters for which we need to estimate the error. Due to the logarithmic dependence of $K$ on $\#\mathcal{P}$ Corollary \[coro:truth est S\] tolerates also (very) pessimistic estimates of $\#\mathcal{P}$. We finish this section by noting that, using the same *error residual relationship* trick , the relative errors $\|u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma/\|u({\boldsymbol{\mu}})\|_\Sigma$ or $\sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}})\|_\Sigma^2}/ \sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| u({\boldsymbol{\mu}})\|_\Sigma^2}$ can be estimated, respectively, using $$\label{eq:rel_err_est} \Delta^\text{rel}({\boldsymbol{\mu}}) = \sqrt{\frac{ \frac{1}{K}\sum_{i=1}^K \big( Y_i({\boldsymbol{\mu}})^T r({\boldsymbol{\mu}}) \big)^2 }{\frac{1}{K}\sum_{i=1}^K \big( Y_i({\boldsymbol{\mu}})^T f({\boldsymbol{\mu}}) \big)^2 } } \qquad\text{and}\qquad \Delta^\text{rel} = \sqrt{ \frac{\frac{1}{K\#\mathcal{P}} \sum_{{\boldsymbol{\mu}}\in \mathcal{P}} \sum_{k=1}^{K} \big(Y_i({\boldsymbol{\mu}})^T r({\boldsymbol{\mu}}) \big)^{2}}{ \frac{1}{K\#\mathcal{P}} \sum_{{\boldsymbol{\mu}}\in \mathcal{P}} \sum_{k=1}^{K} \big(Y_i({\boldsymbol{\mu}})^T f({\boldsymbol{\mu}}) \big)^{2} }}.$$ We then obtain a slightly modified version of Corollary \[coro:truth est S\]; the proof can be found in the appendix. \[coro:truth est S rel\] Given $0 < \delta < 1$ and $w > e$, the condition $$\label{coro:truth est Condition S rel} K\geq \min \left\lbrace \frac{2\log( 2 \,\#\mathcal{P}) + 2\log(\delta^{-1})}{\log(w/e)} , \enspace 3 \right \rbrace,$$ is sufficient to ensure $$\mathbb{P}\Big\{ w^{-1} \Delta^\text{rel}({\boldsymbol{\mu}}) \leq \frac{\|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \|_\Sigma}{\|u({\boldsymbol{\mu}})\|_{\Sigma}} \leq w \Delta^\text{rel}({\boldsymbol{\mu}}) ~,~\forall {\boldsymbol{\mu}}\in \mathcal{P} \Big\} \geq 1 - \delta.$$ and also $$\mathbb{P}\Big\{ w^{-1} \Delta^\text{rel} \leq \frac{\sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}})\|_\Sigma^2}}{\sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| u({\boldsymbol{\mu}})\|_\Sigma^2}} \leq w \Delta^\text{rel} \Big\} \geq 1 - \delta.$$ \[rmk:Fdistribution\] If we have $u({\boldsymbol{\mu}})^{T}\Sigma (u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}))=0$ the random variable $(\Delta^\text{rel}({\boldsymbol{\mu}})\|u({\boldsymbol{\mu}})\|_{\Sigma})^{2}/ \| u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma^{2}$ follows an $F$-distribution with degrees of freedom $K$ and $K$. We may then use the cumulative distribution function (cdf) of the $F$-distribution and an union bound argument to get an even more accurate lower bound for $\mathbb{P}\{ w^{-1} \Delta^\text{rel}({\boldsymbol{\mu}}) \leq \frac{\|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \|_\Sigma}{\|u({\boldsymbol{\mu}})\|_{\Sigma}} \leq w \Delta^\text{rel}({\boldsymbol{\mu}}) ~,~\forall {\boldsymbol{\mu}}\in \mathcal{P} \}$; for further details see section \[sec:f-distribution\]. We remark that although in general we do not have $u({\boldsymbol{\mu}})^{T}\Sigma (u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}))=0$, often $u({\boldsymbol{\mu}})$ and the error are nearly $\Sigma$-orthogonal. As a consequence we observe in the numerical experiments in Section \[sec5\] that the empirical probability density function (pdf) of $(\Delta^\text{rel}({\boldsymbol{\mu}})\|u({\boldsymbol{\mu}})\|_{\Sigma})/\| u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma$ follows the square root of the pdf of the $F$ distribution pretty well. When estimating the error in scalar-valued QoIs of the form of $s({\boldsymbol{\mu}})=l^T u({\boldsymbol{\mu}})$, the covariance matrix is $\Sigma=l\,l^T$, see subsection \[subsec:3.1\]. In that case the random vector $Z\sim\mathcal{N}(0,\Sigma)$ follows the same distribution as $X\,l$ where $X\sim\mathcal{N}(0,1)$ is a standard normal random variable (scalar). The random dual problem then becomes $ A({\boldsymbol{\mu}})^TY_i({\boldsymbol{\mu}}) = X_i\, l $ and the solution is $Y_i({\boldsymbol{\mu}}) = X_i \, q({\boldsymbol{\mu}})$ where $q({\boldsymbol{\mu}})$ is the solution of the deterministic dual problem $A({\boldsymbol{\mu}})^T q({\boldsymbol{\mu}}) = l$. Dual problems of this form are commonly encountered for estimating linear quantities of interest, see [@pierce2000adjoint] for a general presentation and [@Haa17; @RoHuPa08; @zahm2017projection; @AmChHu10] for the application in reduced order modeling. Approximation of the random dual solutions via the PGD {#sec4} ====================================================== In order to obtain a fast-to-evaluate a posteriori error estimator we employ the PGD to compute approximations of the solutions of the $K$ dual problems . Given $K$ independent realizations $Z_1,\hdots,Z_K$ of $Z\sim\mathcal{N}(0,\Sigma)$, we denote by $Y({\boldsymbol{\mu}})=[Y_1({\boldsymbol{\mu}}),\hdots,Y_K({\boldsymbol{\mu}})] \in{\mathbb{R}}^{N\times K}$ the horizontal concatenation of the $K$ dual solutions so that $Y({\boldsymbol{\mu}})$ satisfies the parametrized matrix equation $$A({\boldsymbol{\mu}})^TY({\boldsymbol{\mu}})=[Z_1,\hdots,Z_K].$$ We use the PGD to approximate $Y({\boldsymbol{\mu}})$ by $$\label{eq:dual_PGD} \widetilde{Y}^{L}({\boldsymbol{\mu}}_{1},\hdots ,{\boldsymbol{\mu}}_{p}) = \sum_{m=1}^{L} Y^{m}_{\mathcal{X}} \, Y^{m}_{1} ({\boldsymbol{\mu}}_1)\hdots Y^{m}_{p}({\boldsymbol{\mu}}_p),$$ where the matrices $Y^{m}_{\mathcal{X}}\in{\mathbb{R}}^{N\times K}$ and the functions $Y^{m}_{i}:\mathcal{P}_i \rightarrow{\mathbb{R}}$ are built in a greedy way. The PGD algorithm we employ here is essentially the same as the one we used for the primal approximation, except with minor modifications regarding the fact that $Y({\boldsymbol{\mu}})$ is a matrix and not a vector, see subsection \[subsec:comput\_aspects\]. In subsection \[subsec:intertwined\] we present several stopping criteria for the dual greedy enrichment, including an adaptive algorithm which builds the primal and dual PGD approximation simultaneously. By replacing $Y_i({\boldsymbol{\mu}})$ by $\widetilde Y_i({\boldsymbol{\mu}})$, the $i$-th column of $\widetilde Y({\boldsymbol{\mu}})$, we define pointwise fast-to-evaluate a posteriori error estimators as $$\label{eq: def a post est online} \widetilde \Delta({\boldsymbol{\mu}}) := \sqrt{ \frac{1}{K} \sum_{i=1}^{K} (\widetilde Y_{i}({\boldsymbol{\mu}})^{T} r({\boldsymbol{\mu}}))^{2} } \qquad \text{and} \qquad \widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}}) := \sqrt{\frac{ \frac{1}{K}\sum_{i=1}^K \big(\widetilde Y_i({\boldsymbol{\mu}})^T r({\boldsymbol{\mu}}) \big)^2}{\frac{1}{K}\sum_{i=1}^K \big(\widetilde Y_i({\boldsymbol{\mu}})^T f({\boldsymbol{\mu}}) \big)^2 } },$$ and fast-to-evaluate error estimators for the (relative) RMS error as $$\label{eq: def a post est online RMS} \widetilde \Delta = \sqrt{\frac{1}{\#\mathcal{P}} \sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \widetilde \Delta({\boldsymbol{\mu}})^{2} } \qquad\text{and}\qquad \widetilde{\Delta}^\text{rel} := \sqrt{\frac{ \frac{1}{K\#\mathcal{P}} \sum_{{\boldsymbol{\mu}}\in \mathcal{P}} \sum_{i=1}^K \big(\widetilde Y_i({\boldsymbol{\mu}})^T r({\boldsymbol{\mu}}) \big)^2}{\frac{1}{K\#\mathcal{P}} \sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\sum_{i=1}^K \big(\widetilde Y_i({\boldsymbol{\mu}})^T f({\boldsymbol{\mu}}) \big)^2 } }.$$ Similarly to Proposition 3.3 in [@SmZaPa19], the following propositions bound the effectivity indices for the error estimators and . This will be used in subsection \[subsec:intertwined\] to steer the dual PGD approximation. Again, these results do not depend on the condition number of $A({\boldsymbol{\mu}})$ as it is the case for the residual error estimator . \[prop:dualErrorMultiplicative\] Let $0 < \delta < 1$, $w > \sqrt{e}$ and assume $$\label{eq:assumption dualErrorMultiplicative} K\geq \min \left\lbrace \frac{\log( \#\mathcal{P}) + \log(\delta^{-1})}{\log(w/\sqrt{e})} , \enspace 3 \right \rbrace.$$ Then the fast-to-evaluate estimators $\widetilde\Delta(\mu)$ and $\widetilde \Delta$ satisfy $$\label{eq:dualErrorMultiplicative} \mathbb{P}\Big\{ (\alpha_{\infty} w)^{-1} \widetilde\Delta(\mu) \leq \|u(\mu)-\widetilde u(\mu) \|_\Sigma \leq (\alpha_{\infty} w) \,\widetilde\Delta(\mu), \quad \mu \in \mathcal{P} \Big\} \geq 1-\delta$$ and $$\label{eq:dualErrorMultiplicative2} \mathbb{P}\Big\{ (\alpha_{2} w)^{-1} \widetilde\Delta \leq \sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}})\|_\Sigma^2} \leq (\alpha_{2} w) \,\widetilde\Delta \Big\} \geq 1-\delta,$$ where $$\label{eq:alpha} \alpha_{\infty} := \, \max_{\mu \in \mathcal{P}} \left( \max \left\{\frac{\Delta(\mu)}{\widetilde \Delta(\mu)} \,,\, \frac{\widetilde \Delta(\mu)}{\Delta(\mu)} \right\} \right) \geq 1 \qquad \text{ and } \qquad \alpha_{2}:= \, \max \left\{\frac{\Delta}{\widetilde \Delta} \,,\, \frac{\widetilde \Delta}{\Delta} \right\} \geq 1.$$ The proof follows exactly the proof of Proposition 3.3 in [@SmZaPa19] and we include it in the appendix for the sake of completeness. \[prop:dualErrorMultiplicativerel\] Let $0 < \delta < 1$, $w > e$ and assume $$\label{eq:assumption dualErrorMultiplicativerel} K\geq \min \left\lbrace \frac{2\log( 2\,\#\mathcal{P}) + 2\log(\delta^{-1})}{\log(w/e)} , \enspace 3 \right \rbrace,$$ Then the fast-to-evaluate estimators $\widetilde\Delta^\text{rel}({\boldsymbol{\mu}})$ and $\widetilde \Delta^\text{rel}$ satisfy $$\label{eq:dualErrorMultiplicativerel} \mathbb{P}\Big\{ (\alpha_{\infty}^\text{rel} w)^{-1} \widetilde\Delta^\text{rel}({\boldsymbol{\mu}}) \leq \frac{\|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \|_\Sigma }{\|u({\boldsymbol{\mu}})\|_\Sigma } \leq (\alpha_{\infty}^\text{rel} w) \,\widetilde\Delta^\text{rel}({\boldsymbol{\mu}}), \quad {\boldsymbol{\mu}}\in \mathcal{P} \Big\} \geq 1-\delta$$ and $$\label{eq:dualErrorMultiplicativerel2} \mathbb{P}\Big\{ (\alpha_{2}^\text{rel} w)^{-1} \widetilde\Delta^\text{rel} \leq \frac{\sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}})\|_\Sigma^2}}{\sqrt{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| u({\boldsymbol{\mu}})\|_\Sigma^2}} \leq (\alpha_{2}^\text{rel} w) \,\widetilde\Delta^\text{rel} \Big\} \geq 1-\delta,$$ where $$\label{eq:alpharel} \alpha_{\infty}^\text{rel} := \, \max_{{\boldsymbol{\mu}}\in \mathcal{P}} \left( \max \left\{\frac{\Delta^\text{rel}({\boldsymbol{\mu}})}{\widetilde \Delta^\text{rel}({\boldsymbol{\mu}})} \,,\, \frac{\widetilde \Delta^\text{rel}({\boldsymbol{\mu}})}{\Delta^\text{rel}({\boldsymbol{\mu}})} \right\} \right) \geq 1 \qquad \text{ and } \qquad \alpha_{2}^\text{rel}:= \, \max \left\{\frac{\Delta^\text{rel}}{\widetilde \Delta^\text{rel}} \,,\, \frac{\widetilde \Delta^\text{rel}}{\Delta^\text{rel}} \right\} \geq 1.$$ Can be proved completely analogously to Propostion \[prop:dualErrorMultiplicative\]. PGD algorithm for the random dual solution {#subsec:comput_aspects} ------------------------------------------ To construct the PGD approximation $\widetilde{Y}^{L}({\boldsymbol{\mu}})$ , we employ a pure greedy algorithm. All terms in are constructed one after the other: after $L-1$ iterations, the rank-one correction $Y_{\mathcal{X}}^L Y_{1}^L({\boldsymbol{\mu}}_1) \cdots Y_{p}^L({\boldsymbol{\mu}}_p)$ is computed as the solution to $$\label{eq:MinResPGD_4dual} \min_{Y_{\mathcal{X}}\in{\mathbb{R}}^{N\times K}} \, \min_{Y_{1}:\mathcal{P}_1\rightarrow{\mathbb{R}}} \cdots \min_{Y_{p}:\mathcal{P}_p\rightarrow{\mathbb{R}}} ~\sum_{i=1}^K\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|A({\boldsymbol{\mu}})^T\Big(\widetilde{Y}^{L-1}_i({\boldsymbol{\mu}}) + Y_{\mathcal{X},i} Y_{1}({\boldsymbol{\mu}}_1) \cdots Y_{p}({\boldsymbol{\mu}}_p)\Big) - Z_i\|_{2}^2,$$ where $\widetilde{Y}^{L-1}_i({\boldsymbol{\mu}})$ and $Y_{\mathcal{X},i}$ denote the $i$-th columns of the matrices $\widetilde{Y}^{L-1}_i({\boldsymbol{\mu}})$ and $Y_{\mathcal{X}}$ respectively. Formulation corresponds to the *Minimal residual PGD* as it consists in minimizing the sum of the $K$ residuals associated with the random dual problems . If the matrix $A({\boldsymbol{\mu}})$ is SPD, one can also use the *Galerkin PGD* formulation $$\label{eq:MinEnergyPGD_4dual} \min_{Y_{\mathcal{X}}\in{\mathbb{R}}^{N\times K}} \, \min_{Y_{1}:\mathcal{P}_1\rightarrow{\mathbb{R}}} \cdots \min_{Y_{p}:\mathcal{P}_p\rightarrow{\mathbb{R}}} ~\sum_{i=1}^K\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|A({\boldsymbol{\mu}})^T\Big(\widetilde{Y}^{L-1}_i({\boldsymbol{\mu}}) + Y_{\mathcal{X},i} Y_{1}({\boldsymbol{\mu}}_1) \cdots Y_{p}({\boldsymbol{\mu}}_p)\Big) - Z_i\|_{A({\boldsymbol{\mu}})^{-1}}^2 ,$$ where the canonical norm $\|\cdot\|_2$ has been replaced by the energy norm $\|\cdot\|_{A({\boldsymbol{\mu}})^{-1}}$. As explained before, both of these formulations can be solved efficiently using the *alternating least squares*. Comparing or with the formulations used for the primal PGD solution or , the major difference is that the solution is matrix-valued for $K>1$, so that the first minimization problem over $Y_{\mathcal{X}}\in{\mathbb{R}}^{N\times K}$ might be, at a first glance, more expensive to compute compared to the minimum over $u_{\mathcal{X}}\in{\mathbb{R}}^N$ in or . However, the minimization problem over $Y_{\mathcal{X}}\in{\mathbb{R}}^{N\times K}$ possesses a particular structure which we may exploit for computational efficiency. This structure is that each column of $Y_{\mathcal{X}}$ is the solution to a least squares problem with the same operator but with different right-hand side. Indeed, for fixed $Y_1({\boldsymbol{\mu}}_1),\hdots,Y_p({\boldsymbol{\mu}}_p)$, the $i$-th column of $Y_{\mathcal{X}}$ minimizes $$Y_{\mathcal{X},i} \mapsto \sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|A({\boldsymbol{\mu}})^T Y_{\mathcal{X},i} Y_{1}({\boldsymbol{\mu}}_1) \cdots Y_{p}({\boldsymbol{\mu}}_p) - \big(Z_i-A({\boldsymbol{\mu}})^T\widetilde{Y}^{L-1}_i({\boldsymbol{\mu}}) \big)\|_{*}^2,$$ with either $\|\cdot\|_*=\|\cdot\|_2 $ or $\|\cdot\|_*=\|\cdot\|_{A({\boldsymbol{\mu}})^{-1}}$. This means that $Y_{\mathcal{X},i}$ can be computed by solving a linear system of the form $\widetilde A_{\mathcal{X}} Y_{\mathcal{X},i} = \widetilde Z_i$ for some operator $\widetilde A_{\mathcal{X}} \in{\mathbb{R}}^{N\times N}$ and right-hand side $\widetilde Z_i\in{\mathbb{R}}^N$ which are assembled[^4] using $A({\boldsymbol{\mu}})^T$, $Z_i$ and also $Y_{1}({\boldsymbol{\mu}}_1),\hdots,Y_{p}({\boldsymbol{\mu}}_p)$ and $\widetilde{Y}^{L-1}_i({\boldsymbol{\mu}})$. Then, by concatenating the $K$ right-hand sides, the matrix $Y_{\mathcal{X}}$ solution to or can be obtained by solving a matrix equation of the form $$\widetilde A_{\mathcal{X}} Y_{\mathcal{X}} = [\widetilde Z_1,\hdots,\widetilde Z_K].$$ For certain solvers $Y_{\mathcal{X}}$ can be computed at a complexity which hardly depends on $K$. For instance, one can precompute say a sparse LU or Cholesky factorization of $\widetilde A_{\mathcal{X}}$ and use this factorization to solve for the multiple right-hand sides. Thus, when $K\ll N$, the complexity is dominated by the factorization of $\widetilde A_{\mathcal{X}}$ so that the costs for computing the dual PGD approximation is at least comparable to the one for the primal PGD approximation. \[rmk:RankDualVSRankStag\] Thanks to the above remark, it is worth mentioning that the cost for computing the dual approximation $\widetilde Y^L({\boldsymbol{\mu}})$ is essentially the same as for computing a hierarchical estimate $\widetilde u^{M+k}({\boldsymbol{\mu}})$ with $k=L$. Indeed, both $\widetilde Y^L({\boldsymbol{\mu}})$ and $\widetilde u^{M+k}({\boldsymbol{\mu}})$ require the computation of $k=L$ PGD updates (primal or dual) which can be computed at comparable cost. Therefore, in the numerical experiments, we pay particular attention in comparing our randomized estimators or with the stagnation-based error estimator with $k=L$ to ensure a fair comparison. Instead of looking for $\widetilde{Y}^{L}({\boldsymbol{\mu}})$ as in , one could have build on the tensor product structure of ${\mathbb{R}}^{N\times K}={\mathbb{R}}^N \otimes {\mathbb{R}}^K$ and look for approximation on the form of $Y_i({\boldsymbol{\mu}})\approx \sum_{m=1}^{L} Y^{m}_{\mathcal{X}} \,Y^{m}_K(i)\, Y^{m}_{1} ({\boldsymbol{\mu}}_1)\hdots Y^{m}_{p}({\boldsymbol{\mu}}_p)$ where $Y^{m}_{\mathcal{X}}\in{\mathbb{R}}^N$ is a vector (not a matrix) and where $Y^{m}_K:\{1,\hdots,K\}\rightarrow{\mathbb{R}}$. This amounts to consider the column index $i$ as an additional parameter, and thus to increase the order of the problem by one. This option is however not pursued in this paper. Primal-dual intertwined PGD algorithm {#subsec:intertwined} ------------------------------------- In this subsection we discuss how to choose the rank $L$ of the dual approximation $\widetilde Y^L({\boldsymbol{\mu}})$. First, one could choose the rank *based on a given (computational) budget*. Assuming for instance that the budget is tight one could only consider dual PGD approximations of say rank $1$ to $5$. The numerical experiments in section \[sec5\] show that this choice works remarkably well even for high dimensional examples. This choice, however, does not allow controlling the quality of the dual approximation so that neither $\alpha_{2/\infty}$ nor $\alpha_{2/\infty}^\text{rel}$ in Propositions \[prop:dualErrorMultiplicative\] and \[prop:dualErrorMultiplicativerel\] can be bounded. In order to control the error in the dual approximation, we suggest building the primal and dual approximations simultaneously in an intertwined manner as described in Algorithm \[algo:pgf\_greedy\]. We focus here on the relative error estimator $\Delta^\text{ref}$ to simplify the notation, but other errors such as $\Delta$ or $\max_{{\boldsymbol{\mu}}\in {\mathcal{P}}} \Delta({\boldsymbol{\mu}})$ can be considered as well. First, we calculate $K$ in line \[algo:calc\_K\] for given effectivity $\alpha w$ and failure probability $\delta$ via . As $\widetilde \Delta^\text{ref}$ invokes Proposition \[prop:Chi2Tail\] $2\#{\mathcal{P}}$-times in each greedy iteration and the maximal number of greedy iterations is $M_{max}$, we need to apply Proposition \[prop:Chi2Tail\] at most $M_{max}2\#{\mathcal{P}}$-times during Algorithm \[algo:pgf\_greedy\] and thus have to replace $2\#{\mathcal{P}}$ by $M_{max}2\#{\mathcal{P}}$ in in the calculation of $K$. We then draw $K$ independent Gaussian random vectors with mean zero and covariance matrix $\Sigma$. Next, we describe how to build $\widetilde u^M({\boldsymbol{\mu}})$ and $\widetilde Y^L({\boldsymbol{\mu}})$ in an intertwined manner. Within each iteration of Algorithm \[algo:pgf\_greedy\] we first compute the *primal* rank-one correction using or . Algorithm \[algo:pgf\_greedy\] terminates either if $\widetilde \Delta^\text{rel} \leq tol$ or $M=M_{max}$. However, due to the dual PGD approximation it is a priori not clear whether $\widetilde \Delta^\text{rel}$ has an effectivity close to unity and thus estimates the relative error well. An effectivity that is well below one might result in an premature termination of the algorithm, which is clearly undesirable. This is why every time we need to estimate the error, we first assess the quality of $\widetilde \Delta^\text{rel}$ before using it as a stopping criterion and then, if necessary, enrich the dual PGD approximation until the quality of $\widetilde \Delta^\text{rel}$ is satisfactory. \[algo:calc\_K\]\ \ \[algo:while\] \ \ Motivated by Proposition \[prop:dualErrorMultiplicativerel\], the natural way to measure the quality of $\widetilde \Delta^\text{rel}$ is $\alpha_2^\text{rel} = \max\{\Delta^\text{rel}/\widetilde \Delta^\text{rel} ; \widetilde \Delta^\text{rel}/\Delta^\text{rel}\}$. Since the unbiased error estimator $\Delta^\text{rel}$ is not computationally feasible, as the solution $u({\boldsymbol{\mu}})$ is not known in advance, one possibility is to replace $\Delta^\text{rel}$ with the computationally feasible error estimator $$\Delta^\text{rel}_{+k} = \sqrt{ \frac{ \sum_{{\boldsymbol{\mu}}\in \mathcal{P}} \sum_{k=1}^{K} \big(Z_i^T (\widetilde u^{M+k}({\boldsymbol{\mu}})-\widetilde u^{M}({\boldsymbol{\mu}}) )\big)^{2}}{ \sum_{{\boldsymbol{\mu}}\in \mathcal{P}} \sum_{k=1}^{K} \big(Z_i^T \widetilde u^{M+k}({\boldsymbol{\mu}}) \big)^{2} }}.$$ Here, the number of additional increments $k$ are chosen depending on the available computational budget. We also introduce $$\widetilde\Delta^\text{rel}_{+k} = \frac{\sqrt{\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \sum_{i=1}^{K}(\widetilde Y_{i}^L({\boldsymbol{\mu}})^{T} A({\boldsymbol{\mu}})(\widetilde u^{M+k}({\boldsymbol{\mu}})-\widetilde u^{M}({\boldsymbol{\mu}})))^{2}}}{\sqrt{\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\sum_{i=1}^{K}(\widetilde Y_{i}^L({\boldsymbol{\mu}})^{T} A({\boldsymbol{\mu}})\widetilde u^{M+k}({\boldsymbol{\mu}}))^{2}}},$$ which is the error estimator of the *known* increment $\widetilde u^{M+k}({\boldsymbol{\mu}})-\widetilde u^{M}({\boldsymbol{\mu}})$. For larger $k$ we expect $\widetilde u^{M+k}({\boldsymbol{\mu}})-\widetilde u^{M}({\boldsymbol{\mu}})$ to be already a quite good error model justifying the choice $\widetilde\Delta^\text{rel}_{+k} = \widetilde \Delta^\text{rel}$. We recall and emphasize that the stagnation-based or hierarchical error estimator does neither ensure that the respective estimator bounds the error or allows to bound the effectivity unless we have for instance a uniform bound for the convergence rate of the PGD; see subsection \[subsec:PGD\]. Therefore, we use the increments solely as an error model to train the dual PGD approximation such that we can then rely on Proposition \[prop:dualErrorMultiplicativerel\]. We then utilize in line \[algo:while\] $$\label{eq:alpha2k} \alpha_{2,k}^\text{rel} := \max \left\{\frac{\Delta^\text{rel}_{\pm k}}{\widetilde \Delta^\text{rel}_{\pm k}} \,;\, \frac{\widetilde \Delta^\text{rel}_{\pm k}}{\Delta^\text{rel}_{\pm k}} \right\} ,$$ where $$\label{eq:less_exp_err_ind} \Delta^\text{rel}_{-k} = \sqrt{ \frac{ \sum_{{\boldsymbol{\mu}}\in \mathcal{P}} \sum_{k=1}^{K} \big(Z_i^T (\widetilde u^{M}({\boldsymbol{\mu}})-\widetilde u^{M-k}({\boldsymbol{\mu}}) )\big)^{2}}{ \sum_{{\boldsymbol{\mu}}\in \mathcal{P}} \sum_{k=1}^{K} \big(Z_i^T \widetilde u^{M}({\boldsymbol{\mu}}) \big)^{2} }} \quad\text{ and }\quad \widetilde\Delta^\text{rel}_{-k} = \sqrt{ \frac{ \sum_{{\boldsymbol{\mu}}\in \mathcal{P}} \sum_{k=1}^{K} \big(\widetilde Y_i^L({\boldsymbol{\mu}})^T A({\boldsymbol{\mu}})(\widetilde u^{M}({\boldsymbol{\mu}})-\widetilde u^{M-k}({\boldsymbol{\mu}})) \big)^{2}}{ \sum_{{\boldsymbol{\mu}}\in \mathcal{P}} \sum_{k=1}^{K} \big(\widetilde Y_i^L({\boldsymbol{\mu}})^T A({\boldsymbol{\mu}})\widetilde u^{M}({\boldsymbol{\mu}}) \big)^{2} }}.$$ With this choice, one can compute $\alpha_{2,k}^\text{rel}$ with *no additional* computational effort. Apart from that, the motivation behind the definition of $\Delta^\text{rel}_{-k}$ and $\widetilde\Delta^\text{rel}_{-k}$ is that we can estimate the norms of the increments $\widetilde u^{M}({\boldsymbol{\mu}})-\widetilde u^{M-k}({\boldsymbol{\mu}})$ well. Assuming a relatively uniform convergence behavior of the primal PGD approximation, it can thus be conjectured that $\widetilde \Delta^\text{rel}$ then also estimates the norm of $u({\boldsymbol{\mu}}) - \widetilde u^{M}({\boldsymbol{\mu}})$ well; this is demonstrated in section \[sec5\]. Algorithm \[algo:pgf\_greedy\] terminates the enrichment of the dual PGD approximation within each iteration if $\alpha_{2,k}^\text{rel}$ falls below some prescribed value $\alpha>1$. We note that if the error in the primal PGD approximation is already close to the target tolerance $tol$ the primal PGD approximation can become dependent on the respective realizations of $Z_{1},\hdots,Z_{K}$ in the following sense: For some realizations the estimator $\widetilde \Delta^\text{rel}$ lies below $tol$ and thus terminates Algorithm \[algo:pgf\_greedy\] while for others we have $\widetilde \Delta^\text{rel} >tol$ and thus obtain a primal PGD approximation of higher rank. A careful analysis of this effect on the failure probabilities in Proposition \[prop:dualErrorMultiplicativerel\] is however not necessary as by replacing $2\#\mathcal{P}$ by $2M_{max}\#{\mathcal{P}}$ in the calculation of $K$ in line \[algo:calc\_K\] we ensure that the number of samples $K$ is chosen large enough such that we can estimate all possible outcomes. Numerical experiments {#sec5} ===================== We demonstrate the randomized error estimator for two test cases which are derived from the benchmark problem introduced in [@lam2018multifidelity]. The source code is freely available in Matlab$^{\circledR}$ at the address[^5] so that all numerical results presented in this section are entirely reproducible. We consider the system of linear PDEs whose solution $\mathcal{u}:D\rightarrow{\mathbb{R}}^2$ is a vector field satisfying $$\label{eq:WrenchmarkModel} - \text{div}(C:\varepsilon(\mathcal{u})) - k^2 \mathcal{u} = \mathcal{f},$$ where $D\subset{\mathbb{R}}^2$ is a domain which has the shape of a wrench; see Fig. \[fig:Wrench\]. Here, $\varepsilon(\mathcal{u})=\frac{1}{2}(\nabla\mathcal{u}+\nabla\mathcal{u}^T)$ is the strain tensor, $k^2$ the wave number, and $C$ the fourth-order stiffness tensor derived from the plane stress assumption with Young’s modulus $E$ and fixed Poisson coefficient $\nu=0.3$ such that $$\label{eq:Hooke} C:\varepsilon = \frac{E}{1+\nu}\varepsilon + \frac{\nu E}{1-\nu^2}\text{trace}(\varepsilon)I_2.$$ The right hand-side $\mathcal{f}$ and the boundary conditions are represented in Fig. \[fig:Wrench\]. The finite element discretization using piecewise affine basis functions $\psi_1,\hdots,\psi_N$ yields $N=1467$ degrees of freedom. In the remainder of the paper, we only consider estimating errors in the natural Hilbert norm $\|\cdot\|_{\mathcal{X}}$ defined by $$\|\mathcal{u}\|_{\mathcal{X}}^2 = \int_{D} \text{trace}(\nabla \mathcal{u}(x)^T\nabla \mathcal{u}(x)) + \|\mathcal{u}(x)\|_2^2 \,\text{d}x,$$ so that $\Sigma = R_{\mathcal{X}_N}$, as explained section \[sec2\] (see equation ). [0.43]{}           [0.43]{} [0.43]{}           [0.43]{} Non-homogeneous time-harmonic elastodynamics problem ---------------------------------------------------- In this benchmark problem the Young’s modulus is fixed to $E=1$ uniformly over the domain $D$ and the wave number $k^2$ varies between $0.5$ and $1.2$. The scalar parameter ${\boldsymbol{\mu}}=k^2$ varies on the parameter set $\mathcal{P}$ defined as the uniform grid of $[0.5 , 1.2]$ containing $\#\mathcal{P}=500$ points. The parametrized operator derived from is $$\mathcal{A}({\boldsymbol{\mu}}) = - \text{div}(C:\varepsilon(\cdot)) - {\boldsymbol{\mu}}\, (\cdot).$$ The parametrized stiffness matrix $A({\boldsymbol{\mu}})\in{\mathbb{R}}^{N\times N}$, obtained via a FE discretization, admits an affine decomposition $A({\boldsymbol{\mu}}) = A_1 - {\boldsymbol{\mu}}A_2$, where $(A_1)_{ij} = \int_D \varepsilon(\psi_i):C:\varepsilon(\psi_j) \text{d}x$ and $(A_2)_{ij} = \int_D \psi_i^T \psi_j \text{d}x$. As this benchmark problem contains only one scalar parameter, it is convenient to inspect how the method behaves within the parameter set. We observe in Fig. \[fig:Helmholtz\] that the norm of the solution blows up around three parameter values, which indicates resonances of the operator. This shows that for this parameter range, the matrix $A({\boldsymbol{\mu}})$ is not SPD and therefore we use the Minimal Residual PGD formulations and for the primal and dual PGD approximation. ### Estimating the relative error [0.48]{}           [0.48]{} We first compute a PGD approximation $\widetilde{u}^M({\boldsymbol{\mu}})$ with rank $M=10$ and we aim at estimating the relative error $$\label{eq:HelmholtzExactRelativeError} \frac{\|u({\boldsymbol{\mu}}) - \widetilde{u}^{10}({\boldsymbol{\mu}}) \|_{\mathcal{X}_N}}{\|u({\boldsymbol{\mu}}) \|_{\mathcal{X}_N}}$$ for every ${\boldsymbol{\mu}}\in\mathcal{P}$. In Fig. \[fig:Helmholtz\_ErrorsOnParameterSpace\] we show a comparison of the performances of the randomized residual-based error estimator $\widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}})$ defined in , the error estimator based on the dual norm of the residual, and the stagnation-based error estimator defined, respectively, as $$\label{eq:HelmholtzResidualAndStagnationEstimators} \Delta^\text{res}_{M}({\boldsymbol{\mu}}):=\frac{\|A({\boldsymbol{\mu}})\widetilde{u}^{M}({\boldsymbol{\mu}})-f({\boldsymbol{\mu}})\|_{\mathcal{X}_N'}}{\|f({\boldsymbol{\mu}})\|_{\mathcal{X}_N'} } \qquad\text{and}\qquad \Delta^\text{stag}_{M,k}({\boldsymbol{\mu}}):= \frac{\| \widetilde{u}^{M+k}({\boldsymbol{\mu}}) - \widetilde{u}^{M}({\boldsymbol{\mu}}) \|_{\mathcal{X}_N}}{\|\widetilde{u}^{M+k}({\boldsymbol{\mu}}) \|_{\mathcal{X}_N}}.$$ First, we observe in Fig. \[fig:Helmholtz\_ErrorsOnParameterSpace\] that although the residual-based error estimator $\Delta^\text{res}_{10}({\boldsymbol{\mu}})$ follows the overall variations in the error, it underestimates the error significantly over the whole considered parameter range. Following Remark \[rmk:RankDualVSRankStag\], in order to have a fair comparison between the stagnation-based and the randomized error estimator, we set $k=L=\text{rank}(\widetilde Y^L)$ so that the two estimators require the same computational effort in terms of number of PGD corrections that have to be computed. We observe that with either $k=L=3$ (left) or $k=L=10$ (right), the randomized residual-based error estimator is closer to the true relative error compared to the stagnation-based estimator $\Delta^\text{stag}_{10,k}({\boldsymbol{\mu}})$. Notice however that the randomized error estimator $\widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}})$ is converging towards $\Delta^\text{rel}({\boldsymbol{\mu}})$ with $\text{rank}(\widetilde Y)$ which is not the true relative error (see the dashed red lines on Fig. \[fig:Helmholtz\_ErrorsOnParameterSpace\]), whereas $\Delta^\text{stag}_{10,k}({\boldsymbol{\mu}})$ will converge to the true relative error with $k$. Nevertheless, we do observe that for small $k=L$, our random dual is always competitive. In addition, increasing the number of samples $K$ will reduce the variance of both $\widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}})$ and $\Delta^\text{rel}({\boldsymbol{\mu}})$ such that the randomized error estimator will follow the error even more closely (see also Fig. \[fig:Helmholtz\_histOfEffectivities\] to that end). We finally note that Fig. \[fig:Helmholtz\_ErrorsOnParameterSpace\] shows only one (representative) realization of $\widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}})$ and $\Delta^\text{rel}({\boldsymbol{\mu}})$, and the behavior of the effectivity of $\widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}})$ over various realizations will be discussed next. In detail, we first draw $100$ different realizations of $Z=[Z_1,\hdots,Z_K]$ , $Z_i\sim\mathcal{N}(0,\Sigma)$, for $K=1,2,6,20$ and compute the corresponding dual solutions $\widetilde Y^L$ with either $L=\text{rank}(\widetilde Y^L) = 1,4,8,16$. Next, we compute the randomized error estimator $\widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}})$ and calculate the effectivity index $$\label{eq:Helmholtz_ETA} \eta({\boldsymbol{\mu}}) = \frac{ \widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}}) }{\|u({\boldsymbol{\mu}}) - \widetilde{u}^{10}({\boldsymbol{\mu}}) \|_{\mathcal{X}_N}/\|u({\boldsymbol{\mu}})\|_{\mathcal{X}_N}},$$ for all ${\boldsymbol{\mu}}\in\mathcal{P}$. The resulting histograms are shown in Fig. \[fig:Helmholtz\_histOfEffectivities\]. First, we emphasize that especially for $\widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}})$ based on a PGD approximation of the dual problems of rank $8$, $16$, and to some extent even $4$, we clearly see a strong concentration of $\eta({\boldsymbol{\mu}})$ around one as desired. If the number of samples $K$ increases we observe, as expected, that the effectivity of $\widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}})$ is very close to one (see for instance picture with $K=20$ and $\operatorname{rank}(\widetilde Y)=8$ in Fig. \[fig:Helmholtz\_histOfEffectivities\]). We recall that for the primal PGD approximation used for these numerical experiments we chose the rank $M=10$. We thus highlight and infer that by using a dual PGD approximation of lower rank than those of the primal PGD approximation, we can obtain an error estimator with an effectivity index very close to one. Using a dual PGD approximation of rank one and to some extend rank four, we observe in Fig. \[fig:Helmholtz\_histOfEffectivities\] that the mean over all samples of $\eta({\boldsymbol{\mu}})$ is significantly smaller than one (see for instance picture with $K=6$ and $\operatorname{rank}(\widetilde Y)=1$ in Fig. \[fig:Helmholtz\_histOfEffectivities\]). This behavior can be explained by the fact that in these cases the dual PGD approximation is not rich enough to characterize all the features of the error which results in an underestimation of the error. Finally, as explained in Remark \[rmk:Fdistribution\] and more detailed in section \[sec:f-distribution\], $\eta({\boldsymbol{\mu}})$ is distributed as the square root of an $F$-distribution with degress of freedom $(K,K)$, provided some orthogonality condition holds. The solid black line in Fig. \[fig:Helmholtz\_histOfEffectivities\] is the pdf of such a square root of an $F$-distribution with degress of freedom $(K,K)$. In particular, we observe a nice accordance with the histograms showing that, for sufficiently large $L=\text{rank}(\widetilde Y)$, the effectivity indices concentrate around unity with $K$, as predicted by Proposition \[prop:dualErrorMultiplicativerel\]. ### Testing the intertwined algorithm [0.48]{}        [0.48]{} We now illustrate the intertwined algorithm introduced in Subsection \[subsec:intertwined\]. Here, the goal is to monitor the convergence of the PGD algorithm during the iteration process $M=1,2,\hdots$. We thus estimate the relative RMS error $$\frac{\|u-\widetilde{u}^M \|}{\|u\|} \overset{\eqref{eq:IdealPGDError}}{=} \sqrt{\frac{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|u({\boldsymbol{\mu}})-\widetilde{u}^M({\boldsymbol{\mu}})\|_{\mathcal{X}_N}^2 }{ \frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}} \|u({\boldsymbol{\mu}})\|_{\mathcal{X}_N}^2 }},$$ with our relative randomized error estimator $\widetilde{\Delta}^\text{rel}({\boldsymbol{\mu}})$ introduced in with the rank adaptation for the random dual approximation $\widetilde Y^L$. We employ the economical error indicators within the intertwined algorithm \[algo:pgf\_greedy\] with $k=6$. Fig. \[fig:Helmholtz\_intertwined\] shows the performance of our estimator (10 realizations) as well as the increase in $L=\text{rank}(\widetilde Y^L)$ (averaged over the 10 realizations) during the PGD iteration process. We also compare to the relative RMS residual norm $\Delta_{M}^\text{res}$ and to the relative RMS stagnation-based error estimator $\Delta_{M,k}^\text{stag}$ with $k=5$ for which we recall the expression $$\label{eq:HelmholtzResidualAndStagnationEstimators_Intertwined} \Delta_{M}^\text{res} := \sqrt{\frac{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\|A({\boldsymbol{\mu}})\widetilde{u}^{M}({\boldsymbol{\mu}})-f({\boldsymbol{\mu}})\|^2_{\mathcal{X}_N'} }{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\|f({\boldsymbol{\mu}})\|^2_{\mathcal{X}_N'}}} \qquad\text{and}\qquad \Delta_{M,5}^\text{stag} :=\sqrt{\frac{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\| \widetilde{u}^{M+5}({\boldsymbol{\mu}}) - \widetilde{u}^{M}({\boldsymbol{\mu}}) \|^2_{\mathcal{X}_N} }{\frac{1}{\#\mathcal{P}}\sum_{{\boldsymbol{\mu}}\in\mathcal{P}}\|\widetilde{u}^{M+5}({\boldsymbol{\mu}}) \|^2_{\mathcal{X}_N} }}$$ We notice that our estimator is much more stable compared to the stagnation-based error estimator which tends to dramatically underestimate the error during the phases where the PGD is plateauing. Our estimator is also much more accurate compared to $\Delta_{M}^\text{res}$. When comparing the results with $\alpha=5$ (left) and $\alpha=2$ (right), we observe that a large $\alpha$ results in smaller dual rank $L=\text{rank}(\widetilde Y^L)$ but the quality of the error indicator is deteriorated. Reciprocally, a smaller $\alpha$ yields more demanding dual approximations so that $L=\text{rank}(\widetilde Y^L)$ is, on average, larger. Let us notice that a rank explosion of the dual variable $\widetilde Y^L$ is not recommended because the complexity for evaluating $\widetilde \Delta^\text{rel}$ will also explode. This will be further discussed in an upcoming paper. Parametrized linear elasticity with high-dimensional parameter space -------------------------------------------------------------------- [0.43]{}           [0.43]{} In this benchmark problem the wave number is set to $k^2=0$ so that becomes an elliptic PDE. Young’s modulus $E$ is a random field on $D$ that is log-normally distributed so that $\log(E)$ is a Gaussian random field on $D$ with zero-mean and covariance function $cov(x,y) = \sigma_0 \exp( -(\|x-y\|_2/l_0)^2 )$ with $\sigma=0.4$ and $l_0=4$. A realization of the Young’s modulus field is shown in Fig. \[fig:Elliptic\]. Since $E$ is a random field, such a parametrized problem contains infinitely many parameters. We use here a truncated Karhunen Loève decomposition with $p=20$ terms to approximate the random Young’s modulus field by a finite number of parameter ${\boldsymbol{\mu}}_1,\hdots,{\boldsymbol{\mu}}_{20}$, that is $$\label{eq:KLdecomp} E \approx E({\boldsymbol{\mu}}_1,\hdots,{\boldsymbol{\mu}}_{20}) = \exp( \sum_{i=1}^{20} {\boldsymbol{\mu}}_i \sqrt{\sigma_i}\phi_i ),$$ where $(\sigma_i,\phi_i)$ is the $i$-th eigenpair of the covariance function $cov$ and ${\boldsymbol{\mu}}_i$ are independent standard Gaussian random variables. This means that the parameter set is ${\mathbb{R}}^{20}$ and that $\rho$ is the density of a standard normal distribution in dimension $p=20$. In actual practice, we replace the parameter set by a $20$-dimensional grid $\mathcal{P}:=(x_1,\hdots,x_{50})^{20} \subset {\mathbb{R}}^{20}$, where $x_1,\hdots,x_{50}$ is an optimal quantization of the univariate standard normal distribution using $50$ points which are obtained by Lloyd’s algorithm [@du1999centroidal]. Next, we use the Empirical Interpolation Method (EIM) [@BMNP04] using $28$ “magic points” in order to obtain an affine decomposition of the form $$\label{eq:EIM} E({\boldsymbol{\mu}}_1,\hdots,{\boldsymbol{\mu}}_{20}) \approx \sum_{i=1}^{28} E_j \,\zeta^j({\boldsymbol{\mu}}_1,\hdots,{\boldsymbol{\mu}}_{20}).$$ for some spatial functions $E_j:D\rightarrow{\mathbb{R}}$ and parametric functions $\zeta^j:{\mathbb{R}}^{20}\rightarrow{\mathbb{R}}$. With the EIM method, the functions $\zeta^j$ are defined by means of magic point $x^j\in D$ by $\zeta^j({\boldsymbol{\mu}}) = \exp( \sum_{i=1}^{20} {\boldsymbol{\mu}}_i \sqrt{\sigma_i}\phi_i(x^j) ) = \prod_{i=1}^{20}\zeta_i^j({\boldsymbol{\mu}}_i)$, where $\zeta_i^j({\boldsymbol{\mu}}_i)=\exp( {\boldsymbol{\mu}}_i\sqrt{\sigma_i}\phi_i(x^j) )$. Together with the truncated Karhunen Loève decomposition and with the EIM the approximation of the Young’s modulus fields does not exceed $1\%$ of error, and the differential operator is given by $$\mathcal{A}({\boldsymbol{\mu}}) = - \text{div}(C({\boldsymbol{\mu}}):\varepsilon(\cdot)) \approx - \sum_{j=1}^{28} \text{div}(K_j:\varepsilon(\cdot)) ~ \zeta_1^j({\boldsymbol{\mu}}_1)\hdots\zeta_{20}^j({\boldsymbol{\mu}}_{20}) ,$$ where the forth-order stiffness tensors $C_j$ are given as in with $E$ replaced by $E_j$. The parametrized stiffness matrix $A({\boldsymbol{\mu}})$, obtained via a FE discretization, admits an affine decomposition $A({\boldsymbol{\mu}}) = \sum_{i=1}^{28} A_i \zeta_1^j({\boldsymbol{\mu}}_1)\hdots\zeta_{20}^j({\boldsymbol{\mu}}_{20})$, where $A_i\in{\mathbb{R}}^{N\times N}$ is the matrix resulting from the discretization of the continuous PDE operator $\text{div}(K_j:\varepsilon(\cdot))$. As $A({\boldsymbol{\mu}})$ is a parametrized SPD matrix, we use the *Galerkin PGD* formulation and to compute the primal and dual PGD solutions $\widetilde{u}^{M}({\boldsymbol{\mu}})$ and $\widetilde{Y}^{L}({\boldsymbol{\mu}})$, respectively. For this high-dimensional benchmark problem our goal is to monitor the convergence of the PGD. Again, we estimate the relative RMS error $\|u-\widetilde{u}^M \|/\|u\|$ using $\widetilde{\Delta}^\text{rel}_M$ as in where we fix $K=3$ and the rank of the approximate dual $\widetilde Y^L$ as $L=\text{rank}(\widetilde Y^L)=1,3,5$. The results are reported in Fig. \[fig:EllipticMonitorPGD\]. We highlight that, remarkably, a very small dual rank $L$ is enough to obtain a very accurate estimation of the error. In particular, our approach outperforms the stagnation based error estimator, which is very unstable, and also the relative RMS residual norm, which is almost one order of magnitude off. As in the previous subsection, we mention that the evaluation of the estimator $\widetilde\Delta^\text{rel}$ can be numerically expensive due to high primal and dual ranks ($M$ and $L$). Conclusions {#sec6} =========== In this paper we proposed a novel a posteriori error estimator for the PGD, extending the concept introduced in [@SmZaPa19] to this setting. The proposed randomized error estimator does not require the estimation of stability constants, its effectivity is close to unity and lies within an interval chosen by the user at specified high probability. Moreover, the suggested framework allows estimating the error in any norm chosen by the user and the error in some (vector-valued) QoI. To derive the error estimator we make use of the concentration phenomenon of Gaussian maps. Exploiting the error residual relationship and approximating the associated random dual problems via the PGD yields a fast-to-evaluate a posteriori error estimator. To ensure that also the effectivity of this fast-to-evaluate error estimator lies within a prescribed interval at high probability we suggested building the primal and dual PGD approximations in a intertwined manner, where we enrich the dual PGD approximation in each iteration until the error estimator reaches the desired quality. The numerical experiments for a parametrized time-harmonic elastodynamics problem and a parametrized linear elasticity problem with $20$-dimensional parameter space show that very often and even for a very tight effectivity interval it is sufficient to consider a dual PGD approximation that has a significantly lower rank than the primal PGD approximation. Thanks to its favorable computational complexity, the costs for the proposed a posteriori error estimator are about the same or much smaller than the costs for the primal PGD approximation. A numerical comparison with an error estimator defined as the dual norm of the residual and a stagnation-based error estimator (also called hierarchical error estimator) shows that the randomized error estimator provides a much more precise description of the behavior of the error at a comparable computational cost. Proofs ====== Assuming $w > \sqrt{e}$ and $K\geq3$ we can use Proposition \[prop:Chi2Tail\] for any vector $v({\boldsymbol{\mu}}) \in {\mathbb{R}}^{N}$ and a fixed parameter ${\boldsymbol{\mu}}\in \mathcal{P}$ and obtain $$\begin{aligned} \label{eq:aux_prop} \mathbb{P}\Big\{ w^{-1}\|v({\boldsymbol{\mu}})\|_\Sigma \leq \| \Phi v({\boldsymbol{\mu}}) \|_2 \leq w \|v({\boldsymbol{\mu}})\|_\Sigma \Big\} \geq 1-\Big( \frac{\sqrt{e}}{w} \Big)^{K}. \end{aligned}$$ Using the definition of $\Delta^\text{rel}({\boldsymbol{\mu}})$, invoking twice and using a union bound argument the following inequality $$\begin{aligned} \label{eq:aux_coro} \frac{1}{w^{2}} \Delta^\text{rel}({\boldsymbol{\mu}}) = \frac{1}{w^{2}} \frac{\|\Phi (u({\boldsymbol{\mu}}) - \widetilde u({\boldsymbol{\mu}})\|_{2}}{\|\Phi u({\boldsymbol{\mu}}) \|_{2}} \leq \frac{1}{w} \frac{\|u({\boldsymbol{\mu}}) - \widetilde u({\boldsymbol{\mu}})\|_{\Sigma}}{\|\Phi u({\boldsymbol{\mu}}) \|_{2}} \leq \frac{\|u({\boldsymbol{\mu}}) - \widetilde u({\boldsymbol{\mu}})\|_{\Sigma}}{\| u({\boldsymbol{\mu}}) \|_{\Sigma}} \leq w \frac{\|u({\boldsymbol{\mu}}) - \widetilde u({\boldsymbol{\mu}})\|_{\Sigma}}{\|\Phi u({\boldsymbol{\mu}}) \|_{2}} \leq w^{2} \Delta^\text{rel}({\boldsymbol{\mu}})\end{aligned}$$ holds at least with probability $1-2\Big( \frac{\sqrt{e}}{w} \Big)^{K}$. Using again a union bound argument yields $$\begin{aligned} &\mathbb{P}\Big\{ \frac{1}{w^{2}} \Delta^\text{rel}({\boldsymbol{\mu}}) \leq \frac{\|u({\boldsymbol{\mu}}) - \widetilde u({\boldsymbol{\mu}})\|_{\Sigma}}{\| u({\boldsymbol{\mu}}) \|_{\Sigma}} \leq \frac{1}{w^{2}} \Delta^\text{rel}({\boldsymbol{\mu}}) ~,~\forall {\boldsymbol{\mu}}\in\mathcal{P} \Big\} \geq 1-2(\# \mathcal{P}) \Big( \frac{\sqrt{e}}{w} \Big)^{K}. \end{aligned}$$ For a given $0<\delta<1$ condition $ K\geq (\log( 2 \# \mathcal{P}) + \log(\delta^{-1}))/\log(w/\sqrt{e}), $ is equivalent to $1-(2 \# \mathcal{P})( \frac{\sqrt{e}}{w} )^{K} \geq 1-\delta$ and ensures that holds for all ${\boldsymbol{\mu}}\in\mathcal{P}$ with probability larger than $1-\delta$. Setting $\hat{w}=w^{2}$ and renaming $\hat{w}$ as $w$ concludes the proof for $\Delta^\text{rel}({\boldsymbol{\mu}})$; the proof for $\Delta^\text{rel}$ follows similar lines. We provide the proof for $\Delta({\boldsymbol{\mu}})$ noting that the proof for $\Delta$ can be done all the same. By Corollary \[coro:truth est S\], it holds with probability larger than $1-\delta$ that $w^{-1} \Delta({\boldsymbol{\mu}}) \leq \|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \|_\Sigma \leq w \Delta({\boldsymbol{\mu}})$ for all ${\boldsymbol{\mu}}\in\mathcal{P}$. Then with the same probability we have $$\|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \| \leq w \Delta({\boldsymbol{\mu}}) \leq w \left( \sup_{{\boldsymbol{\mu}}'\in\mathcal{P}} \frac{\Delta({\boldsymbol{\mu}}')}{\widetilde \Delta({\boldsymbol{\mu}}')} \right) \widetilde \Delta({\boldsymbol{\mu}}) \overset{\eqref{eq:alpha}}{\leq} (\alpha_{\infty} w) \widetilde \Delta({\boldsymbol{\mu}}),$$ and $$\|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \| \geq w^{-1} \Delta({\boldsymbol{\mu}}) \geq w^{-1} \left( \inf_{{\boldsymbol{\mu}}'\in\mathcal{P}} \frac{\Delta({\boldsymbol{\mu}}')}{\widetilde \Delta({\boldsymbol{\mu}}')} \right) \widetilde \Delta({\boldsymbol{\mu}}) \overset{\eqref{eq:alpha}}{\geq} (\alpha_{\infty} w)^{-1} \widetilde \Delta({\boldsymbol{\mu}}),$$ for any ${\boldsymbol{\mu}}\in\mathcal{P}$, which yields and concludes the proof. Connection of $\Delta^\text{rel}({\boldsymbol{\mu}})$ with the $F$-distribution {#sec:f-distribution} =============================================================================== If we have $u({\boldsymbol{\mu}})^{T}\Sigma (u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}))=0$ we can give an even more precise description of the random variable $(\Delta^\text{rel}({\boldsymbol{\mu}})\|u({\boldsymbol{\mu}})\|_{\Sigma})^{2}/ \| u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma^{2}$ and thus $\mathbb{P}\Big\{ w^{-1} \Delta^\text{rel}({\boldsymbol{\mu}}) \leq \frac{\|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \|_\Sigma}{\|u({\boldsymbol{\mu}})\|_{\Sigma}} \leq w \Delta^\text{rel}({\boldsymbol{\mu}}) ~,~\forall {\boldsymbol{\mu}}\in \mathcal{P} \Big\}$: In fact, under this assumption the random variable $(\Delta^\text{rel}({\boldsymbol{\mu}})\|u({\boldsymbol{\mu}})\|_{\Sigma})^{2}/ \| u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma^{2}$ follows an $F$-distribution with degrees of freedom $K$ and $K$; we also say $(\Delta^\text{rel}({\boldsymbol{\mu}})\|u({\boldsymbol{\mu}})\|_{\Sigma})^{2}/\| u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma^{2} \sim F(K,K)$ as stated in the following lemma. \[lem:f-distribution\] Assume that there holds $u({\boldsymbol{\mu}})^{T}\Sigma (u({\boldsymbol{\mu}}) - {\widetilde{u}(\boldsymbol{\mu})})=0$. Then the random variable $(\Delta^\text{rel}({\boldsymbol{\mu}})\|u({\boldsymbol{\mu}})\|_{\Sigma})^{2}/\| u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma^{2}$ follows an $F$-distribution with parameters $K$ and $K$. We will exploit the fact that in order to show that a random variable $X$ follows an $F$-distribution it is sufficient to show that we have $X = \frac{Q_{1}/K_{1}}{Q_{2}/K_{2}}$, where $Q_{1}$ and $Q_{2}$ are independent random variables that follow a chi-square distribution with $K_{1}$ and $K_{2}$ degrees of freedom, respectively (see for instance [@DeSc12 Definition 9.7.1]). Similar to we have that $$\Delta^\text{rel}({\boldsymbol{\mu}})^{2} \frac{\|u({\boldsymbol{\mu}})\|_{\Sigma}^{2}}{\| u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma^{2}} = \frac{ \frac{1}{K} \sum_{k=1}^{K} \Big(Z_{i}^{T} \big(u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \big) \Big)^{2} }{\frac{1}{K} \sum_{k=1}^{K} \Big(Z_{i}^{T} u({\boldsymbol{\mu}}) \Big)^{2}} \frac{\|u({\boldsymbol{\mu}})\|_{\Sigma}^{2}}{\| u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma^{2}} = \frac{ \frac{1}{K} \sum_{k=1}^{K} \Big(Z_{i}^{T} \frac{\big(u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \big)}{\| u({\boldsymbol{\mu}})- \widetilde u({\boldsymbol{\mu}})\|_\Sigma}\Big)^{2} }{\frac{1}{K} \sum_{k=1}^{K} \Big(Z_{i}^{T} \frac{u({\boldsymbol{\mu}})}{\|u({\boldsymbol{\mu}})\|_{\Sigma}} \Big)^{2} } \sim \frac{\frac{1}{K}Q_{1}}{\frac{1}{K}Q_{2}},$$ where $Q_{1}$ and $Q_{2}$ follow a chi-squared distribution with $K$ degrees of freedom. It thus remains to show that under the above assumption $Q_{1}$ and $Q_{2}$ are independent. To that end, recall that $Z_{1},\hdots,Z_{K}$ are independent copies of $Z\sim \mathcal{N}(0,\Sigma)$. Per our assumption that $\Sigma$ is symmetric and positive semi-definite there exists a symmetric matrix $\Sigma^{1/2}$ that satisfies $\Sigma = \Sigma^{1/2}\Sigma^{1/2}$, exploiting for instance the Schur decomposition of $\Sigma$. Then we have $Z_{i} = \Sigma^{1/2}g_{i}$, where $g_{i}$ are $K$ independent copies of a standard Gaussian random vector. Thanks to the assumption $u({\boldsymbol{\mu}})^{T}\Sigma (u({\boldsymbol{\mu}}) - {\widetilde{u}(\boldsymbol{\mu})})=0$ the vectors $\hat{u}({\boldsymbol{\mu}}):= \Sigma^{1/2}u({\boldsymbol{\mu}})$ and $\hat{e}({\boldsymbol{\mu}}):=\Sigma^{1/2}(u({\boldsymbol{\mu}}) - {\widetilde{u}(\boldsymbol{\mu})})$ satisfy $\hat{u}({\boldsymbol{\mu}})^{T}\hat{e}({\boldsymbol{\mu}})=0$. As a consequence there exists an orthogonal matrix $V({\boldsymbol{\mu}}) \in {\mathbb{R}}^{N\times N}$ that maps $\hat{u}({\boldsymbol{\mu}})$ to $e_{1}$ and $\hat{e}({\boldsymbol{\mu}})$ to $e_{2}$, respectively, where $e_{j}$, $j=1,\hdots,N$ denote the standard basis vectors in ${\mathbb{R}}^{N}$. We can then write for $i=1,\hdots,K$ $$(Z_{i}^{T}u({\boldsymbol{\mu}}))^{2}=([\Sigma^{1/2}g_{i}]^{T} u({\boldsymbol{\mu}}))^{2} = (g_{i}^{T}\Sigma^{1/2}u({\boldsymbol{\mu}}))^{2} = (g_{i}^{T}\hat{u}({\boldsymbol{\mu}}))^{2} = (g_{i}^{T}V({\boldsymbol{\mu}})^{T}V({\boldsymbol{\mu}})\hat{u}({\boldsymbol{\mu}}))^{2} = ([V({\boldsymbol{\mu}})g_{i}]^{T}e_{1})^{2}.$$ As there holds $V({\boldsymbol{\mu}})g_{i} \sim \mathcal{N}(0,I_{N})$, where $I_{N}$ is the identity matrix in ${\mathbb{R}}^{N}$, we have for $\bar{g}_{i} \sim \mathcal{N}(0,I_{N})$ $$(Z_{i}^{T}u({\boldsymbol{\mu}}))^{2} = (\bar{g}_{i}^{T}e_{1})^{2} = (\bar{g}^{1}_{i})^{2} \qquad \text{ and } \qquad [Z_{i}^{T} \big(u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \big)]^{2} = (\bar{g}_{i}^{T}e_{2 })^{2} = (\bar{g}^{2}_{i})^{2}.$$ Here, $\bar{g}^{s}_{i}$ denotes the $s$-th entry of the vector $\bar{g}_{i}$. As the entries of a standard Gaussian random vector are independent, we can thus conclude that the random variables $Q_{1}$ and $Q_{2}$ are independent as well. We may then use the cumulative distribution function (cdf) of the $F$-distribution and an union bound argument to obtain the following result. Assume that there holds $u({\boldsymbol{\mu}})^{T}\Sigma (u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}))=0$. Then, we have $$\mathbb{P}\Big\{ w^{-1} \Delta^\text{rel}({\boldsymbol{\mu}}) \leq \frac{\|u({\boldsymbol{\mu}})-\widetilde u({\boldsymbol{\mu}}) \|_\Sigma}{\|u({\boldsymbol{\mu}})\|_{\Sigma}} \leq w \Delta^\text{rel}({\boldsymbol{\mu}}) ~,~\forall {\boldsymbol{\mu}}\in \mathcal{P} \Big\} \geq 1 - (\#\mathcal{P})\left(I_{\left(\frac{K}{K+w^{2}K}\right)}\left(\frac{K}{2},\frac{K}{2}\right) + 1 - I_{\left(\frac{Kw^{2}}{Kw^{2}+K}\right)}\left(\frac{K}{2},\frac{K}{2}\right)\right),$$ where $I_{x}(a,b)$ is the regularized incomplete beta function defined as $(\int_{0}^{x} t^{a-1}(1-t)^{b-1} \,dt)/(\int_{0}^{1} t^{a-1}(1-t)^{b-1} \,dt)$. Note that the cdf $F(x,K,K)$ of an $F$-distribution with $K$ and $K$ degrees of freedom is given as $F(x,K,K)=I_{\left(\frac{Kx}{Kx+K}\right)}\left(\frac{K}{2},\frac{K}{2}\right)$. [10]{} , [*An error estimator for real-time simulators based on model order reduction*]{}, Adv. Model. and Simul. in Eng. Sci., 2 (2015). , [*An error estimator for separated representations of highly multidimensional models*]{}, Comput. Methods Appl. Mech. Engrg., 199 (2010), pp. 1872–1880. , [*A new family of solvers for some classes of multidimensional partial differential equations encountered in kinetic theory modeling of complex fluids*]{}, J. Non-Newtonian Fluid Mech., 139 (2006), pp. 153–176. height 2pt depth -1.6pt width 23pt, [*A new family of solvers for some classes of multidimensional partial differential equations encountered in kinetic theory modelling of complex fluids: Part ii: Transient simulation using space-time separated representations*]{}, J. Non-Newtonian Fluid Mech., 144 (2007), pp. 98–121. , [*Randomized linear algebra for model reduction. part i: Galerkin methods and error estimation*]{}, tech. rep., arXiv, 1803.02602, 2018. , [*An ’empirical interpolation’ method: application to efficient reduced-basis discretization of partial differential equations*]{}, C. R. Math. Acad. Sci. [P]{}aris [S]{}eries [I]{}, 339 (2004), pp. 667–672. , [*A tensor approximation method based on ideal minimal residual formulations for the solution of high-dimensional problems*]{}, ESAIM Math. Model. Numer. Anal., 48 (2014), pp. 1777–1806. , [*[Problem adapted Hierarchical Model Reduction for the Fokker-Planck equation]{}*]{}, in [Proceedings of ALGORITMY 2016, 20th Conference on Scientific Computing, Vysoke Tatry, Podbanske, Slovakia, March 13-18, 2016]{}, [Proceedings of ALGORITMY 2016, 20th Conference on Scientific Computing, Vysoke Tatry, Podbanske, Slovakia, March 13-18, 2016]{}, 2016, pp. 13–22. , [*Randomized [L]{}ocal [M]{}odel [O]{}rder [R]{}eduction*]{}, SIAM J. Sci. Comput., 40 (2018), pp. A2120–A2151. , [*Greedy algorithms for high-dimensional non-symmetric linear problems*]{}, in C[ANUM]{} 2012, 41e [C]{}ongr[è]{}s [N]{}ational d’[A]{}nalyse [N]{}umérique, vol. 41 of ESAIM Proc., EDP Sci., Les Ulis, 2013, pp. 95–131. , [*A posteriori error estimation and global error control for ordinary differential equations by the adjoint method*]{}, SIAM J. Sci. Comput., 26 (2004), pp. 359–374. , [*A posteriori error estimation and adaptive strategy for pgd model reduction applied to parametrized linear parabolic problems*]{}, Comput. Methods Appl. Mech. Engrg., 327 (2017), pp. 118–146. , [*Improved successive constraint method based a posteriori error estimate for reduced basis approximation of 2[D]{} [M]{}axwell’s problem*]{}, M2AN Math. Model. Numer. Anal., 43 (2009), pp. 1099–1116. , [*Recent advances and new challenges in the use of the proper generalized decomposition for solving multidimensional models*]{}, Arch. Comput. Methods Eng., 17 (2010), pp. 327–350. , [*[PGD-Based Modeling of Materials, Structures and Processes]{}*]{}, Springer International Publishing, 2014. , [*The proper generalized decomposition for advanced numerical simulations. A primer*]{}, SpringerBriefs in Applied Sciences and Technology, Springer, Cham, 2014. , [*A short review on model order reduction based on proper generalized decomposition*]{}, Arch. Comput. Methods Eng., 18 (2011), pp. 395–404. , [*[An elementary proof of the Johnson-Lindenstrauss lemma]{}*]{}, International Computer Science Institute, Technical Report, (1999), pp. 99–006. , [*Probability and statistics. Fourth Edition*]{}, Pearson Education, 2012. , [*Centroidal voronoi tessellations: Applications and algorithms*]{}, SIAM Rev., 41 (1999), pp. 637–676. , [*Variational monte carlo – bridging concepts of machine learning and high-dimensional partial differential equationsbridging concepts of machine learning and high-dimensional partial differential equations*]{}, Adv. Comput. Math., (2019). , [*Numerical strategies for the [G]{}alerkin-proper generalized decomposition method*]{}, Math. Comput. Modelling, 57 (2013), pp. 1694–1702. , [*Tensor-based methods for numerical homogenization from high-resolution images*]{}, Comput. Methods Appl. Mech. Engrg., 254 (2013), pp. 154–169. , [*Reduced basis methods for parametrized [PDE]{}s – a tutorial*]{}, in Model Reduction and Approximation, P. Benner, A. Cohen, M. Ohlberger, and K. Willcox, eds., SIAM, Philadelphia, PA, 2017, pp. 65–136. , [*Tensor spaces and numerical tensor calculus*]{}, Springer, Heidelberg, 2012. , [*Certified reduced basis methods for parametrized partial differential equations*]{}, Springer Briefs in Mathematics, Springer, Cham, 2016. , [*[Error Estimation for Reduced-Order Models of Dynamical Systems]{}*]{}, SIAM Rev., 49 (2007), pp. 277–299. , [*[A natural-norm Successive Constraint Method for inf-sup lower bounds]{}*]{}, Comput. Methods Appl. Mech. Engrg., 199 (2010), pp. 1963 – 1975. , [*A successive constraint linear optimization method for lower bounds of parametric coercivity and inf-sup stability constants*]{}, C. R. Math. Acad. Sci. Paris, 345 (2007), pp. 473–478. , [*Goal-oriented error estimation for the reduced basis method, with application to sensitivity analysis*]{}, J. Sci. Comput., 68 (2016), pp. 21–41. , [*[Extensions of Lipschitz mappings into a Hilbert space]{}*]{}, Contemp. Math., 26 (1984), p. 1. , [*Small-sample statistical condition estimates for general matrix functions*]{}, SIAM J. Sci. Comput., 15 (1994), pp. 36–61. , [*On a goal-oriented version of the proper generalized decomposition method*]{}, J. Sci. Comput., (2019). , [*On the verification of model reduction methods based on the proper generalized decomposition*]{}, Comput. Methods Appl. Mech. Engrg., 200 (2011), pp. 2032–2047. height 2pt depth -1.6pt width 23pt, [*Toward guaranteed PGD-reduced models. Bytes and Science*]{}, CIMNE: Barcelona, 2012, pp. 143–154. , [*Multifidelity dimension reduction via active subspaces*]{}, arXiv preprint arXiv:1809.05567, (2018). , [*Results and questions on a nonlinear approximation approach for solving high-dimensional partial differential equations*]{}, Constr. Approx., 30 (2009), pp. 621–651. , [*A basis for bounding the errors of proper generalised decomposition solutions in solid mechanics*]{}, Internat. J. Numer. Methods Engrg., 94 (2013), pp. 961–984. , [*A separated representation of an error indicator for the mesh refinement process under the proper generalized decomposition framework*]{}, Comput. Mech., 55 (2015), pp. 251–266. , [*A generalized spectral decomposition technique to solve a class of linear stochastic partial differential equations*]{}, Comput. Methods Appl. Mech. Engrg., 196 (2007), pp. 4521–4537. , [*Low-rank methods for high-dimensional approximation and model order reduction*]{}, in Model reduction and approximation, P. Benner, A. Cohen, M. Ohlberger, and K. Willcox, eds., SIAM, Philadelphia, PA, 2017, pp. 171–226. , [*A dimensional reduction approach based on the application of reduced basis methods in the framework of hierarchical model reduction*]{}, SIAM J. Sci. Comput., 36 (2014), pp. A714–A736. , [*Hierarchical local model reduction for elliptic problems: a domain decomposition approach*]{}, Multiscale Model. Simul., 8 (2010), pp. 1102–1127. , [*[Adjoint recovery of superconvergent functionals from PDE approximations]{}*]{}, SIAM Rev., 42 (2000), pp. 247–264. , [*Reduced basis methods for partial differential equations*]{}, vol. 92, Springer, Cham, 2016. , [*Reduced basis approximation and a posteriori error estimation for affinely parametrized elliptic coercive partial differential equations: application to transport and continuum mechanics*]{}, Arch. Comput. Meth. Eng., 15 (2008), pp. 229–275. , [*Hierarchical model reduction of nonlinear partial differential equations based on the adaptive empirical projection method and reduced basis techniques*]{}, ESAIM Math. Model. Numer. Anal., 51 (2017), pp. 641–677. , [*Randomized residual-based error estimators for parametrized equations*]{}, SIAM J. Sci. Comput., 41 (2019), pp. A900–A926. , [*A posteriori error bounds for reduced-basis approximation of parametrized noncoercive and nonlinear elliptic partial differential equations*]{}, 16th AIAA Computational Fluid Dynamics Conference, (2003). , [*High-dimensional probability. An introduction with applications in data science*]{}, vol. 47 of Cambridge Series in Statistical and Probabilistic Mathematics, Cambridge University Press, Cambridge, 2018. , [*On a dimensional reduction method. [I]{}. [T]{}he optimal selection of basis functions*]{}, Math. Comp., 37 (1981), pp. 31–46. , [*Projection-based model order reduction methods for the estimation of vector-valued variables of interest*]{}, SIAM J. Sci. Comput., 39 (2017), pp. A1647–A1674. , [*Interpolation of inverse operators for preconditioning parameter-dependent equations*]{}, SIAM J. Sci. Comput., 38 (2016), pp. A1044–A1074. [^1]: Faculty of Electrical Engineering, Mathematics & Computer Science, University of Twente, Zilverling, P.O. Box 217, 7500 AE Enschede, The Netherlands, `[email protected]` [^2]: Univ. Grenoble Alpes, CNRS, Inria, Grenoble INP, LJK, 38000 Grenoble, France, `[email protected]` [^3]: Note that this is in contrast to the celebrated Johnson-Lindenstrauss lemma [@dasgupta1999elementary; @johnson1984extensions].The latter ensures that for any $0<\varepsilon<1$ and any finite set $\mathcal{M}\subset\mathbb{R}^N$ that contains the zero vector, the condition $K\geq 8\varepsilon^{-2}\log(\#\mathcal{M})$ ensures the existence of a linear map $\Phi: \mathbb{R}^{N} \rightarrow \mathbb{R}^{K}$ such that $ \sqrt{1-\varepsilon} \|v\|_2 \leq \| \Phi v \|_2 \leq \sqrt{1+\varepsilon} \| v \|_2 $ for all $v\in\mathcal{M}$. Increasing $\varepsilon$ does therefore not allow reducing the number of samples $K$; for a more detailed discussion see [@SmZaPa19]. [^4]: For example with the Galerkin PGD formulation , we have $\widetilde A_{\mathcal{X}} = \sum_{{\boldsymbol{\mu}}\in{\mathcal{P}}} A({\boldsymbol{\mu}})^T Y_{1}({\boldsymbol{\mu}}_1)^2 \cdots Y_{p}({\boldsymbol{\mu}}_p)^2$ and $\widetilde Z_i =\sum_{{\boldsymbol{\mu}}\in{\mathcal{P}}} ( Z_i - A({\boldsymbol{\mu}})^T\widetilde{Y}^{L-1}_i({\boldsymbol{\mu}})) Y_{1}({\boldsymbol{\mu}}_1) \cdots Y_{p}({\boldsymbol{\mu}}_p) $. [^5]: <https://gitlab.inria.fr/ozahm/randomdual_pgd.git>
{ "pile_set_name": "ArXiv" }
--- abstract: | We analysed two-dimensional maps of 48 early-type galaxies obtained with the SAURON and OASIS integral-field spectrographs using kinemetry, a generalisation of surface photometry to the higher order moments of the line-of-sight velocity distribution (LOSVD). The maps analysed include: reconstructed image, mean velocity, velocity dispersion, $h_3$ and $h_4$ Gauss-Hermite moments. Kinemetry is a good method to recognise structures otherwise missed by using surface photometry, such as embedded disks and kinematic sub-components. In the SAURON sample, we find that 31% of early-type galaxies are single component systems. 91% of the multi-components systems have two kinematic subcomponents, the rest having three. In addition, 29% of galaxies have kinematically decoupled components, nuclear components with significant kinematic twists. We differentiate between slow and fast rotators using velocity maps only and find that fast rotating galaxies contain disks with a large range in mass fractions to the main body. Specifically, we find that the velocity maps of fast rotators closely resemble those of inclined disks, except in the transition regions between kinematic subcomponents. This deviation is measured with the kinemetric $k_5/k_1$ ratio, which is large and noisy in slow rotators and about 2% in fast rotators. In terms of E/S0 classification, this means that 74% of Es and 92% of S0s have components with disk-like kinematics. We suggest that differences in $k_5/k_1$ values for the fast and slow rotators arise from their different intrinsic structure which is reflected on the velocity maps. For the majority of fast rotators, the kinematic axial ratios are equal to or less than their photometric axial ratios, contrary to what is predicted with isotropic Jeans models viewed at different inclinations. The position angles of fast rotators are constant, while they vary abruptly in slow rotators. Velocity dispersion maps of face-on galaxies have shapes similar to the distribution of light. Velocity dispersion maps of the edge-on fast rotators and all slow rotators show differences which can only be partially explained with isotropic models and, in the case of fast rotators, often require additional cold components. We constructed local (bin-by-bin) $h_3 - V/\sigma$ and $h_4 - V/\sigma$ diagrams from SAURON observations. We confirm the classical anti-correlation of $h_3$ and $V/\sigma$, but we also find that $h_3$ is almost zero in some objects or even weakly correlated with $V/\sigma$. The distribution of $h_4$ for fast and slow rotators is mildly positive on average. In general, fast rotators contain flattened components characterised by a disk-like rotation. The difference between slow and fast rotators is traceable throughout all moments of the LOSVD, with evidence for different intrinsic shapes and orbital contents and, hence, likely different evolutionary paths. author: - | Davor Krajnović,$^1$[^1] R. Bacon,$^2$ Michele Cappellari,$^1$ Roger L. Davies,$^1$ P. T. de Zeeuw,$^{3,9}$ Eric Emsellem,$^2$ Jesús Falcón-Barroso,$^{4}$ Harald Kuntschner,$^5$ Richard M. McDermid,$^6$ Reynier F. Peletier,$^{7}$ Marc Sarzi,$^8$ Remco C. E. van den Bosch$^9$and Glenn van de Ven$^{10}$[^2]\ $^1$Denys Wilkinson Building, University of Oxford, Keble Road, OX1 3RH, UK\ $^2$Université de Lyon, France; Université Lyon 1, F-69007; CRAL, Observatoire de Lyon, F-69230 Saint Genis Laval; CNRS, UMR 5574 ; ENS de Lyon, France\ $^3$European Southern Observatory, Karl-Schwarzschild-Str 2, 85748 Garching, Germany\ $^4$European Space and Technology Centre (ESTEC), Keplerlaan 1, Postbus 299, 2200 AG Noordwijk, The Netherlands\ $^5$Space Telescope European Coordinating Facility, European Southern Observatory, Karl-Schwarzschild-Str 2, 85748 Garching, Germany\ $^6$Gemini Observatory, Northern Operations Center, 670 N. A’ohoku Place,Hilo, Hawaii, 96720, USA\ $^7$Kapteyn Astronomical Institute, Postbus 800, 9700 AV Groningen, The Netherlands\ $^8$Centre for Astrophysics Research, University of Hertfordshire, Hatfield, Herts AL1 09AB\ $^9$Sterrewacht Leiden, Leiden University, Niels Bohrweg 2, 2333 CA Leiden, The Netherlands\ $^{10}$Institute for Advanced Study, Peyton Hall, Princeton, NJ 08544, USA title: 'The SAURON project - XII. Kinematic substructures in early-type galaxies: evidence for disks in fast rotators' --- \[firstpage\] galaxies: elliptical and lenticular - galaxies: kinematics and dynamics - galaxies: structure - galaxies:evolution Introduction {#s:intro} ============ The classification of galaxies both acknowledges the complexity of these celestial objects and attempts to understand their formation and evolution. The Hubble Classification of galaxies [@1936RNeb..........H; @1961hag..book.....S; @1975gaun.book.....S; @1994cag..book.....S] recognises the dichotomy between, broadly speaking, disk and ellliptical galaxies, for historical reasons now often called late- and early-types. The classification works well on the late-type galaxies in particular, dividing the class into a number of subgroups which correlate with properties such as bulge-to-disk ratio, morphology of spiral arms, gas and dust content, to name a few, but it fails to bring a physical insight to our understanding of early-types [@1987IAUS..127..367T], where the classification is based on apparent shape and thus dependant on viewing angles. In an effort to eliminate the unsatisfactory situation @1996ApJ...464L.119K proposed a revision of the Hubble classification. It was based on two discoveries in the 1970s and 1980s, both enabled by an improvement in the technical capabilities of astronomical instruments. A series of papers [@1975ApJ...200..439B; @1977ApJ...218L..43I] showed that bright elliptical galaxies do not rotate as fast as they should, if they were oblate isotropic systems supported by rotation [@1978MNRAS.183..501B], whereas less bright and generally smaller systems, including also bulges of spirals, generally agree with such predictions [@1983ApJ...266...41D; @1982modg.proc..113K; @1982ApJ...256..460K]. A complementary discovery that the fast rotating galaxies are more likely to have disky isophotes, while the slow rotating galaxies have boxy isophotes, linked again the kinematics and shape of galaxies. @1996ApJ...464L.119K changed the uniformity of early-type galaxies to a dichotomy (disky vs boxy, fast vs slow, brighter vs less bright) and linked the whole Hubble sequence from right to left, from Sc–Sb–Sa to S0–E types, where ’rotation decreases in dynamical importance compared to random motions’. This important step forward introduced a readily measurable parameter related to some physical properties. However, the higher order variations in the isophotal shape (diskiness/boxiness) are not measurable at all inclinations regardless of the prominence of the disks [@1990ApJ...362...52R], and finally, they are used to infer the dynamical state of the galaxy. This might be a decent approximation, especially if one assumes that all fast rotating galaxies comprise of spheroidal slow rotating components and disks seen at different inclination [@1999ApJ...513L..25R], but the edge-on observations of spheroidal components in spiral galaxies showed that bulges are rotating fast as well [@1982ApJ...256..460K]. To complicate things further the updated classification continues to distinguish S0s from Es, keeping a viewing angle dependent definition of S0s [@1990ApJ...348...57V]. The choice of using the fourth order Fourier term in the isophotometric analysis for classification is natural, because [ *(i)*]{} it is much easier to take images of galaxies than to measure their kinematics, and [*(ii)*]{} until recently it was not realistic to spectroscopically map their two-dimensional structure. The advent of panoramic integral-field units, such as SAURON [@2001MNRAS.326...23B], is changing the technical possibilities and the field itself; it is now possible to systematically map kinematics of nearby galaxies up to their effective radii. We have observed 72 nearby E, S0 and Sa galaxies as part of the SAURON survey [@2002MNRAS.329..513D hereafter Paper II]. Focusing here on a subsample of 48 early-type galaxies (E, S0), these observations clearly show the previously hinted rich variety of kinematic substructures such as: kinematically decoupled cores, kinematic twists, counter-rotating structures and central disks [@2004MNRAS.352..721E hereafter Paper III]. Analysing the global properties of the SAURON velocity and velocity dispersion maps, @2007MNRAS.379..401E [hereafter Paper IX] were able to separate the early-type galaxies into two physically distinct classes of [*slow*]{} and [*fast*]{} rotators, according to their specific (projected) angular momentum measured within one effective radius, $\lambda_R$. This finding augments the view that led to the revision of the classification, but the SAURON observations provide the crucial quantitive data. Moreover, the results of Paper IX suggest a way to dramatically improve on the Hubble classification and substitute S0s and (misclassified) disky ellipticals with one class of fast rotators. @2007MNRAS.379..418C [hereafter Paper X] addressed again the issue of orbital anisotropy of early-type galaxies. They constructed the ($V/\sigma$, $\epsilon$) diagram [@1978MNRAS.183..501B] using an updated formalism [@2005MNRAS.363..937B], and compared it with the results from general axisymmetric dynamical models for a subsample of these galaxies [@2006MNRAS.366.1126C hereafter Paper IV]. They found that slow and fast rotators are clearly separated on the ($V/\sigma$, $\epsilon$) diagram (unlike Es and S0s), such that slow rotators are round, moderately isotropic and are likely to be somewhat triaxial, while fast rotators appear flattened, span a larger range of anisotropies, but are mostly oblate axisymmetric objects. This finding is in a partial agreement with previous studies which either found round early-type galaxies radially anisotropic [@1991MNRAS.253..710V], moderately radially anisotropic [@2001AJ....121.1936G] or only weakly anisotropic with a range of anisotropies for flattened systems [@2003ApJ...583...92G]. The results of Paper X, however, clearly show that intrinsically flatter galaxies tend to be more anisotropic in the meridional plane. The models also indicate that the fast rotators are often two component systems, having also a flat and rotating, kinematically distinct, disk-like component. Dynamical models are often time consuming and difficult to construct. Ultimately, one would like to be able to classify galaxies by their observable properties only. Is it possible to learn about the intrinsic shapes of the early-type galaxies from observations only? Surface photometry, being but the zeroth moment of the ultimate observable quantity for distant galaxies, the line-of-sight velocity distribution (LOSVD), cannot give the final answer. It is necessary to look at the other moments of the LOSVD: mean velocity, velocity dispersion and higher-order moments, commonly parameterised by Gauss–Hermite coefficients, $h_3$ and $h_4$ [@1993ApJ...407..525V; @1993MNRAS.265..213G], which measure asymmetric and symmetric deviation of the LOSVD from a Gaussian, respectively. Indeed, in the last dozen years several studies investigated higher moments of the LOSVD of early-type galaxies observing them along one or multiple slits . These studies deepened the dichotomy among early-type galaxies showing that fast rotating galaxies with disky isophotes also exhibit an anti-correlation between $h_3$ and $V/\sigma$. This is consistent with these galaxies being made of two components: a bulge and a disk. The symmetric deviations, on the other hand are usually smaller than asymmetric ones, and somewhat positive in general. In addition, the observed higher order moments of the LOSVD can be used to constrain the possible merger scenarios of early-type galaxies and their formation in general . However, observations along one or two slits are often not able to describe the kinematical richness of early-type galaxies. In this work we use [[*kinemetry*]{}]{} [@2006MNRAS.366..787K], a generalisation of surface photometry to all moments of the LOSVD, to study SAURON maps of 48 early-type galaxies. The purpose of this work is to investigate observational clues from resolved two-dimensional kinematics for the origin of the differences between the slow and fast rotators. In Section \[s:dat\] we briefly remind the reader of the SAURON observations and data reduction. Section \[s:met\] describes the methods and definitions used in this work. The main results are presented in Section \[s:res\]. In Section \[s:disc\] we offer an interpretation of the results and we summarise the conclusions in Section \[s:conc\]. In Appendix \[s:psf\] we discuss the influence of seeing on the two-dimensional kinematics and in Appendix \[s:profiles\] we present the radial profiles of the kinemetric coefficients used in this study. Sample and Data {#s:dat} =============== In this work we used the data from the SAURON sample which was designed to be [*representative*]{} of the galaxy populations in the plane of ellipticity, $\epsilon$, versus absolute $B$ band magnitude $M_B$. The sample and its selection details are presented in Paper II. In this study we focus on the 48 galaxies of the SAURON E+S0 sample. SAURON is an integral-field spectrograph with a field-of-view (FoV) of about $33\arcsec \times 41\arcsec$ and $0\farcs94 \times 0\farcs94$ square lenses, mounted at the William Herschel Telescope. Complementing the SAURON large scale FoV, we probed the nuclear regions of a number of galaxies with OASIS, then mounted at Canada-Hawaii-France Telescope, a high spatial resolution integral-field spectrograph, similarly to SAURON based on the TIGER concept . The FoV of OASIS is only $10\arcsec \times 8\arcsec$, but the spatial scale is $0\farcs27 \times 0\farcs27$, fully sampling the seeing-limited point spread function and providing on average a factor of 2 improvement in spatial resolution over SAURON. The spectral resolution of OASIS is, however, about 20% lower than that of SAURON, and only a sub-sample of the SAURON galaxies was observed. In this paper we are investigating the stellar kinematics of early-type galaxies. Paper III and @2006MNRAS.373..906M [hereafter Paper VIII] discuss the extraction of kinematics and construction of maps of the mean velocity $V$, the velocity dispersion $\sigma$, and the Gauss-Hermite moments $h_3$ and $h_4$ in great detail. All maps used in this work are Voronoi binned [@2003MNRAS.342..345C] to the same signal-to-noise ratio. The SAURON kinematic data used here are of the same kinematic extraction as in Paper X with the latest improvement on the template mismatch effects in higher moments of the LOSVD. The SAURON mean velocity maps are repeated in this paper for the sake of clarity, but we encourage the reader to have copies of both Paper III and Paper VIII available for reference on other moments of the LOSVD. Method and definitions {#s:met} ====================== Maps of the moments of the LOSVD offer a wealth of information, but also suffer from complexity. It is difficult, if not impossible, to show error bars for each bin on the map, and the richness of the maps can lead to the useful information being lost in detail. As in the case of imaging, it is necessary to extract the useful information from the maps to profit from their two-dimensional coverage of the objects. In this section we describe the method used to analyse the maps and discuss definitions utilised throughout the paper. Kinemetry {#ss:kin} --------- @2006MNRAS.366..787K presented kinemetry, as a quantitative approach to analysis of maps of kinematic moments. Kinemetry is a generalisation of surface photometry to the higher-order moments of the LOSVD. The moments of the LOSVD have odd or even parity. The surface brightness (zeroth moment) is even, the mean velocity (first moment) is odd, the velocity dispersion (second moment) is even, etc. Kinemetry is based on the assumption that for the odd moments the profile along the ellipse satisfies a simple cosine law, while for the even moments the profile is constant (the same assumption is also used in surface photometry). Kinemetry derives such best-fitting ellipses and analyses the profiles of the moments extracted along these by means of harmonic decomposition. It follows from this that the application of kinemetry on even moments is equivalent to surface photometry resulting in the same coefficients for parameterisation of the structures (e.g position angle, ellipticity and fourth order harmonics). Application of kinemetry on odd maps such as velocity maps[^3] provides radial profiles of the kinematic position angle [[PA]{}$_{kin}$]{}, axis ratio or flattening, [[q]{}$_{kin}$]{}=b/a (where b and a are lengths of minor and major axis, respectively), and odd harmonic terms obtained from the Fourier expansion (since velocity is an odd map, even terms are, in principle, not present, while in practice are very small and usually negligible). In the case of stellar velocity maps, the dominant kinemetry term is $k_1=\sqrt{a_1^2+b_1^2}$, representing the velocity amplitude, where $a_1$ and $b_1$ are the first sine and cosine terms, respectively. The deviations from the assumed simple cosine law are given by the first higher order term that is not fitted, $k_5=\sqrt{a_5^2+b_5^2}$, usually normalised with respect to $k_1$. These four parameters form the basis of our analysis because they quantify the kinematical properties of the observed galaxies: orientation of the map (a projection of the angular momentum), opening angle of the iso-velocity contours, the amplitude of the rotation and the deviation from the assumed azimuthal variation of the velocity map. For the other moments of the LOSVD one could derive similar quantities, depending on the parity of the moment. As will be discussed below, we focus on kinemetry coefficients that describe velocity maps in detail and some specific kinemetry coefficients from the maps of the higher order moments. A detailed description of the method, error analysis and parameters is given in @2006MNRAS.366..787K. Radial profiles {#ss:prof} --------------- Kinemetric radial profiles can be obtained along ellipses of different axial ratios and position angles. At each radius there is the best fitting ellipse, along which a profile of the kinematic moment will have a certain shape: it follows a cosine or it is constant, for odd and even moments, respectively. If this is the case, the higher order Fourier terms are non-existent or at least negligible for such an ellipse. In the case of even moments, the best fitting ellipses describe the underlying iso-contours, like isophotes in the case of surface photometry, or contours of constant velocity dispersion, iso-$\sigma$ contours. In the case of odd moments, this is somewhat more difficult to visualise, but the axial ratio of the best-fitting ellipse is related to the opening angle of the iso-velocity contours: the larger the axial ratio, the more open is the spider diagram of the velocity map. In this study, kinemetry is used for extraction of parameters in the following ways: - We apply kinemetry to SAURON reconstructed images of galaxies, which are obtained by summing the spectra along the spectral direction at each sky position. This is equivalent to low resolution surface photometry on galaxies from the sample. We focus on the photometric position angle [[PA]{}$_{phot}$]{} and photometric axial ratio, related to ellipticity as [[q]{}$_{phot}$]{}=1-$\epsilon$. In this case, kinemetry is used in its [*even*]{} mode, where even harmonics are fitted to the profiles extracted along the best fitting ellipses. - We use kinemetry to derive radial profiles of the four parameters that describe velocity maps: [[PA]{}$_{kin}$]{}, [[q]{}$_{kin}$]{}, $k_1$ and $k_5$. In this case, kinemetry is applied to the maps in its [ *odd*]{} mode, when only odd Fourier harmonics are fit to the profiles extracted along the best fitting ellipses, which are, in general, different from the best fitting ellipse of [*(i)*]{}. In some cases when it is not possible to determine the best-fitting ellipse we run kinemetry on circles (see below). - Kinemetry is applied to velocity dispersion maps, using the [*even*]{} mode as in [*(i)*]{}. In this case, however, the parameters of the ellipses used to extract profiles were fixed to the best fitting values of surface photometry obtained in [ *(i)*]{}. - Maps of Gauss-Hermite coefficients $h_3$ and $h_4$ were also parameterised using kinemetry in [*odd*]{} and [*even*]{} mode, respectively. In both cases, we used the best fitting ellipses from the lowest [*odd*]{} (velocity map) and [*even*]{} moment (reconstructed image), respectively. Before proceeding it is worth explaining in more detail our decision not to use kinemetry to fit the ellipses in some cases. Under [ *(ii)*]{} we mentioned that on some velocity maps it was necessary to run kinemetry on circles. In general, the mean stellar velocity has an odd parity, and its map, in an inclusive triaxial case, will be point-antisymmetric. Certain maps, however, do not follow this rule, having no detectable net rotation, e.g. NGC4486, or the inner part of NGC4550. In the latter case, the zero velocity in the inner part can be explained by the superposition of two counter-rotating stellar components as advocated by @1992ApJ...394L...9R and @1992ApJ...400L...5R, where the mass of the counter-rotating component is about 50% of the total mass (Paper X). In other cases the non rotation could be a result of dominant box orbits which have zero angular momentum. The basic assumption of kinemetry for odd kinematic moments therefore breaks down resulting in velocity maps that appear noisy and one cannot expect reasonable results. In practice, this means that the best-fitting ellipse parameters for maps without net rotation will not be robustly determined (degeneracy in both [[PA]{}$_{kin}$]{}and [[q]{}$_{kin}$]{}) while the higher harmonic terms will be large and meaningless. Specifically, $k_5$ will have high values. We partially alleviate this degeneracy by first running an unconstrained kinemetry fit on stellar velocity maps and identifying maps where $k_5/k_1 > 0.1$ and corresponding radii where it occurs. At these radii we re-run kinemetry, but using circles for extraction of velocity profiles and Fourier analysis. In this way we set the axial ratio [[q]{}$_{kin}$]{}=1 in order to break the degeneracy. Although the $k_5/k_1$ term cannot be directly compared with the $k_5/k_1$ term obtained from a best-fitting ellipse, in this case, if there is any indication of odd parity in the map, we can still determine the local amplitude of rotation $k_1$, and give a good estimate for [[PA]{}$_{kin}$]{}. The other note refers to items [*(iii)*]{} and [*(iv)*]{}. Although, in principle, it would be possible to run kinemetry freely on the velocity dispersion maps, or maps of higher Gauss-Hermite moments (e.g. $h_3$ and $h_4$), the noise in the data is too high to give trustworthy results for the whole sample. By setting the shape of the curve to the best-fitting ellipses of the corresponding lowest odd or even moment, the harmonic terms of kinemetry quantify the differences between these even and odd moments of the LOSVD. ![\[f:diff\] [**Left:**]{} Voronoi binned stellar velocity dispersion maps of NGC2549 (top) and NGC4473 (bottom). Over-plotted lines are isophotes. In the lower right corner of each map are values that correspond to minimum and maximum colours on the colour bar. [**Right:**]{} Radial profiles of $b_2/a_2$ kinemetry terms for NGC2549 (top) and NGC4473 (bottom).](figures/fig1.ps){width="\columnwidth"} An example of expected differences can be visualised comparing the isophotes of the surface brightness and the stellar velocity dispersion maps of NGC2549 and NGC4473 (Fig. \[f:diff\]). On both images isophotes are aligned with the vertical (y) axis of the maps. In the case of NGC2549 the contours of constant velocity dispersion seem to be perpendicular to the isophotes, at least within the central 10, while in the rather unusual case of NGC4473, the high values of velocity dispersion have the same orientation as the isophotes. The physical explanation of these striking differences should be looked for in the internal orbital structure. We postpone this discussion to Section \[ss:shape\]. The noise and irregular shape of iso-$\sigma$ contours decrease the usefulness of fitting for the velocity dispersion contours. Extracting harmonic terms along the isophotes, however, can yield a clear signal of the different shape of these two moments. An extracted velocity dispersion profile in these two cases (for example along the second brightest isophote shown on Fig. \[f:diff\]) will go through two maxima and two minima. The minima and maxima of these two profiles will be out of phase, because along the major axis in NGC2549 there is a decrease in velocity dispersion while in NGC4473 there is an increase. The decomposition of these two profiles will give different amplitudes to the harmonic terms. Specifically, $b_2$ (cosine) term will be the most influenced, because this term is related to the error in the axial ratio [@1987MNRAS.226..747J], and the shapes such as in NGC2549 and NGC4473 will give negative (iso-$\sigma$ rounder than isophotes) and positive (iso-$\sigma$ flatter than isophotes) $b_2$, respectively. An alternative way to visualise the difference between these two maps based on the values of $b_2$ is to consider that a negative $b_2,$ corresponds to a decrease of the $\sigma$ at the major axis of the best fitting ellipse compared to the $\sigma$ measured at the minor axis of the best fitting ellipse (NGC2549). In contrast, a positive $b_2$ corresponds to an increased value of the $\sigma$ at the major axis position of the best fitting ellipse (NGC4473). By monitoring these harmonic terms it is possible to quantify the shape difference between the observed zeroth and second moments of the LOSVD. Definition of structures on stellar velocity maps {#ss:def_comp} ------------------------------------------------- A few kinemetric profiles are able to describe a wealth of information from the maps. Specifically, we wish to use them to highlight the kinematic structures on the maps and to recognise hidden kinematic components. Here we present a set of quantitative criteria for describing features on the stellar velocity maps. Some of the criteria are dependent on the quality of the data and they should be modified if used on maps obtained with other IFUs. The following rules were presented by @2006MNRAS.366..787K and Paper IX, but here we list them for the sake of clarity. A single velocity map can contain a number of kinematic components. Often they are easily recognisable by visual inspection. In a quantitive way we differentiate between: - [*Single Component*]{} (SC) map: having a radially constant or slow varying [[PA]{}$_{kin}$]{}and [[q]{}$_{kin}$]{}profiles. - [*Multiple Component*]{} (MC) map: characterised with an abrupt change in either: $\Delta$[[q]{}$_{kin}$]{}$ > 0.1$, or $\Delta$[[PA]{}$_{kin}$]{}$> 10\degr$, or a double-hump in $k_1$ with a local minimum in between, or a peak in $k_5$ where $k_5/k_1>0.02$. MC maps are clearly more complex than SC maps. The above values for changes to the kinemetric coefficients are used to determine the extent of each subcomponent (components $C_1$, $C_2$ and $C_3$ with radii $R_{12}$ and $R_{23}$ between them). Each subcomponent can be described as being of the following type (limiting values apply for the SAURON dataset): - [*Disc-like rotation*]{} (DR): defined when the higher order harmonic $k_5/k_1 < 0.02$, while the variation of [[q]{}$_{kin}$]{} and [[PA]{}$_{kin}$]{}  is less than 0.1 and $10\degr$, respectively. Note that this name [ *does not*]{} imply that the object is a disk intrinsically . - [*Low-level velocity*]{} (LV): defined when the maximum of $k_1$ is lower than $15 {\>{\rm km}\,{\rm s}^{-1}}$. A special case is [*Central low-level velocity*]{} (CLV) when LV occurs in the central kinematical component on the map. - [*Kinematic misalignment*]{} (KM): defined when the absolute difference between the photometric [[PA]{}$_{phot}$]{} and kinemetric position [[PA]{}$_{kin}$]{} angles is larger than $10\degr$. - [*Kinematic twist*]{} (KT): defined by a smooth variation of the kinematics position angle [[PA]{}$_{kin}$]{} with an amplitude of at least $10\degr$ within the extent of the kinematic component. - [*Kinematically decoupled component*]{} (KDC): if there is an abrupt change in [[PA]{}$_{kin}$]{}, with a difference larger than $20\degr$ between two adjacent components, or if there is an outer LV component (in which case the measurement of [[PA]{}$_{kin}$]{} is uncertain). A special case of KDCs are [*Counter rotating cores*]{} (CRC) where $\Delta $[[PA]{}$_{kin}$]{} between the components is $180\degr$ (within the uncertainties). Most of the above definitions are new, arising from two-dimensional maps which offer a more robust detection of structures. The definition of KDC is, however, similar to the one used in the past , where the two-dimensional coverage enables a determination of the orientation of the kinematic components. It should be noted that classification of a kinematic component as a CLV is strongly dependent on the spatial resolution of the instrument. As will be seen later, higher spatial resolution can change the appearance and therefore the classification of the components. Similarly, it should be stressed that the limiting values used for these definitions are geared towards the SAURON data. The OASIS data, due to different instrumental properties and observing set-up, will have somewhat different limiting values, mostly arising in the higher order Fourier terms. For example, the mean uncertainty of $k_5/k_1$ term for the OASIS sample is 0.033, significantly higher compared to the one for the SAURON sample (0.015). In order to treat consistently the two data sets, we adopt a somewhat more conservative value of 0.04 as the limiting values for $k_5/k_1$ in definition of DR component when estimated from the OASIS data. While abrupt changes in the orientation, axial ratio, or velocity amplitude are intuitively clear as evidences for separate kinematic components, the $k_5/k_1$ as an indicator of components is more complex to comprehend. Still, the simple models of two kinematic components rotating at a given relative orientation give rise to $k_5/k_1$ term in the kinemetric expansion in the region where these components overlap [@2006MNRAS.366..787K]. Since we measure luminosity weighted velocities, the position and extent of the raised $k_5/k_1$ region depends on relative luminosity contributions of the components, marking the transition radii between the components and not their start or end. Furthermore, it is also, necessary, to distinguish between high $k_5/k_1$ due to a super-position of kinematic components (a genuine signal) and high $k_5/k_1$ originating from noisy maps, such as maps with no detectable rotation (e.g. NGC4486) or large bin-to-bin variations (e.g. OASIS map of NGC3379). For our data, when the signal in $k_5$ is $10\%$ of $k_1$, we consider the noise too high and the $k_5/k_1$ ratio not usable for detecting individual components. Seeing and quantification of kinematic components {#ss:skc} ------------------------------------------------- Robust estimates of the number of sub-components in velocity maps and their sizes are influenced by three major factors: data quality, physical properties and seeing. While the data quality is described by measurement uncertainties, and in that sense it is quantifiable to some extent, the other two factors are more complex. By ’physical properties’ we assume physical processes that hide kinematic information from our view, such as specific orientation of the object, dust obscuration or simply the fact that we are measuring luminosity weighted quantities and we might miss kinematic components made up of stars that constitute a low luminosity fraction of the total population. The influence of seeing is particularly relevant for subcomponents in the centres of galaxies. In Appendix \[s:psf\] we tested the dependence of the kinemetric coefficients on representative seeings, for velocity maps viewed at different orientations. This exercise showed that: [*(i)*]{} [[PA]{}$_{kin}$]{} and $k_5/k_1$ are not significantly influenced by the seeing, [*(ii)*]{} the amplitude and, to a minor extent, the shape of $k_1$ are somewhat influenced by the seeing, and [*(iii)*]{} the axial ratio [[q]{}$_{kin}$]{} can be strongly influenced by the seeing (Fig. \[f:prof\_fwhm\]). In addition to these conclusions, the test showed that the inclination of an object is also a factor contributing to the change of the intrinsic [[q]{}$_{kin}$]{}, and to a minor extent, $k_1$ profiles, where higher inclinations are particularly influenced by the seeing effects. In practice, this means that the change in [[q]{}$_{kin}$]{} is a less robust indicator of kinematic components. We found that more robust indicators are abrupt changes in $k_5/k_1$ and [[PA]{}$_{kin}$]{} profiles, double humps in $k_1$ profiles or decrease of $k_1$ amplitude below our detection limit for rotation. We used these as estimates of the sizes of kinematic components. It should be, however, noted that the size of a component is just a luminosity weighted estimate, originating from a super-position of luminosities of individual components, and the component can intrinsically extend beyond that radius. Only detailed dynamical models could give a more robust estimate of the internal orbital structure. Determination of global and average values {#ss:glob} ------------------------------------------ In addition to radial profiles we present in this paper a number of average quantities. Similar luminosity-weighted quantities have already been derived in Papers IX and X: global [[PA]{}$_{kin}$]{}, global [[PA]{}$_{phot}$]{}, average $\epsilon$. In this study we use the velocity maps to determine the luminosity weighted average $\langle$[[PA]{}$_{kin}$]{}$\rangle$, $\langle$[[q]{}$_{kin}$]{}$\rangle$, and $\langle k_5/k_1 \rangle$ for the whole map and for each kinematic component. We also measured the luminosity weighted $\langle$[[PA]{}$_{phot}$]{}$\rangle$, $\langle$[[q]{}$_{phot}$]{}$\rangle$ from the reconstructed images (both global and for each component), $\langle b_2/a_0 \rangle$ from velocity dispersion maps and average values of $h_4$ (measured as the $a_0$ harmonic term) from $h_4$ maps. In practice, we do this following the expression from Paper IX. The mean $\langle G \rangle$ of a quantity $G(R)$ derived from its sampled radial profiles can be approximated with: $$\label{eq:av} \langle G \rangle \sim \frac{ \sum_{k=1}^{N} q(R_k) F(R_k) G(R_k) (R^2_{out,k} - R^2_{in,k}) }{\sum_{k=1}^{N}q(R_k) F(R_k) (R^2_{out,k} - R^2_{in,k})} $$ where $q(R_k)$ and $F(R_k)$ are the axial ratio and the surface brightness of the best fit ellipse, with semi-major axis $R_k$. Eq. (\[eq:av\]) is based on an expression defined in @1999ApJ...517..650R. The uncertainties of these average values are calculated in the standard way as the sum of the quadratic differences between the average value and the value at each position $R_k$. In Paper X the global [[PA]{}$_{kin}$]{} was derived using the formalism from Appendix C in @2006MNRAS.366..787K. This approach differs from the one described here in the sense that it is less sensitive to the kinematic structures in the central region, such as abrupt changes of PA in case of a KDC. That approach is well suited for making global comparisons between [[PA]{}$_{kin}$]{} or [[PA]{}$_{phot}$]{}, such as global kinematic misalignment, when it is required that they are measured on large scales to avoid influence of local perturbations in the nuclear regions (e.g. seeing, dust, bars). In this study, however, we want to compare the radial properties of different moments of the LOSVD and for that reason we use the approach of Paper IX to all measured quantities. Note that for the purpose of the direct comparison we measured both kinematic and photometric quantities on the SAURON data, in contrast with Papers IX and X. Results {#s:res} ======= In this section we present the results of kinemetric analysis of the LOSVD moments maps. We look at the presence of kinematic substructures in velocity maps (Section \[ss:subcomp\]), properties of radial profiles of [[PA]{}$_{kin}$]{}, $k_1$, $\sigma$ (Section \[ss:SRFR\]), comparison between [[q]{}$_{kin}$]{} and [[q]{}$_{phot}$]{} (Section \[ss:flat\]), the shape difference between isophotes and iso-$\sigma$ contours (Section \[ss:shape\]) and properties of $h_3$ and $h_4$ Gauss-Hermite moments (Section \[ss:high\]) with the purpose to investigate the internal structure of SAURON galaxies. Kinemetry probes local characteristics of galaxies, and we wish to link those with the global properties described in Papers IX and X. In this analysis, the most useful is the first moment of the LOSVD, the mean velocity, because it is a moment rich in structure and with the strongest signal. We present the kinemetric profiles of this moment in Appendix \[s:profiles\]. Although kinemetry is performed on other moments of the LOSVD, we discuss the dominant terms only. Substructures on the velocity maps {#ss:subcomp} ---------------------------------- --------- ---------- ------- ---------- ---------- ------- ------- ------- ------------- --------- ------------------------------------------------------------------------------ Galaxy Group $N_C$ $R_{12}$ $R_{23}$ $C_1$ $C_2$ $C_3$ $KM$ Rotator Comment    (1)   (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) NGC0474 MC 2 7 – DR DR – KM(1,2) F KT between $C_1$ and $C_2$ NGC0524 SC 1 – – DR – – – F Possible $C_2$ beyond r=12 NGC0821 SC 1 – – DR – – – F Flat $k_1$ profile NGC1023 SC 1 – – KT – – – F $k_5/k_1< 0.02$, NGC2549 MC 2 13 – DR DR – – F $C_2$: Flat $k_1$ profile NGC2685 SC 1 – – DR – – – F – NGC2695 MC 2 7 – DR DR – – F – NGC2699 MC 2 6 – DR DR – KM(1,-) F $C_2$: Flat $k_1$ profile NGC2768 SC 1 – – – – – – F $r \lesssim 10\arcsec$ rigid body rotation; $k_5/k_1<0.02$ for $r>10\arcsec$ NGC2974 SC 1 – – DR – – – F – NGC3032 MC (CLV) 2 2.5 – LV DR – – F CRC in OASIS map NGC3156 SC 1 – – DR – – – F – NGC3377 SC 1 – – KT – – – F $k_5/k_1 < 0.02$ over the map NGC3379 SC 1 – – DR – – – F – NGC3384 MC 2 10 – DR DR – KM(-,2) F – NGC3414 MC (KDC) 2 10 – DR LV – KM(-,2\*) S CRC NGC3489 MC 2 6 – DR DR – – F - NGC3608 MC (KDC) 2 10 – DR LV – KM(-,2\*) S CRC NGC4150 MC (CLV) 3 3.5 9.5 LV – DR KM(1,-,-) F KDC in OASIS map with $r_{size}=1\farcs5$ NGC4262 MC 2 9 – DR – – KM(1,2) F – NGC4270 MC 2 6 – – – – – F Possible there components NGC4278 MC 2 16 – DR – – KM(-,2) F $C_2$: Decreasing $k_1$ profile NGC4374 SC 1 – – LV – – KM\* S - NGC4382 MC (CLV) 3 2 14.5 LV DR DR KM(1,-,-) F CRC in OASIS map, KT between $C_1$ and $C_2$ NGC4387 MC 2 7 – DR DR – – F Decreasing $k_1$ beyond $r=13\arcsec$ NGC4458 MC (KDC) 2 3 – – LV – KM(-,2\*) S – NGC4459 MC 2 12 – DR DR – KM(1,-) F – NGC4473 MC 2 10 – DR – – – F $C_1$: possible KT. $C_2$: decreasing $k_1$ profile. NGC4477 SC 1 – – DR – – KM F – NGC4486 SC 1 – – LV – – KM\* S – NGC4526 MC 2 11 – DR – – – F $C_2$: decreasing $k_1$ profile NGC4546 MC 2 9 – DR DR – KM(1,-) F $C_2$: Flat $k_1$ profile NGC4550 SC 1 – – LV – – KM\* S Two co-spatial counter-rotating disks not detected NGC4552 MC (KDC) 2 4 – KT – – KM(1,2) S Flat $k_1=15{\>{\rm km}\,{\rm s}^{-1}}$ over the map NGC4564 SC 1 – – DR – – – F – NGC4570 MC 2 8 – DR DR – – F – NGC4621 MC (KDC) 2 4 – DR DR – KM(1,-) F CRC in OASIS with r$\sim1\farcs5$ NGC4660 MC 2 7 – DR DR – – F – NGC5198 MC (KDC) 2 2.5 – – KT – KM(1\*,2\*) S $C_2$: LV between 2.5-10 NGC5308 MC 2 7 – DR DR – – F No signature of $C_2$ in $k_5/k_1$ NGC5813 MC (KDC) 2 12 – DR LV – KM(-,2\*) S – NGC5831 MC (KDC) 2 8 – DR LV – KM(-,2\*) S – NGC5838 MC 2 6 – DR DR – – F No signature of $C_2$ in $k_5/k_1$ NGC5845 MC 2 4.5 – DR DR – – F – NGC5846 SC 1 – – LV – – KM\* S – NGC5982 MC (KDC) 2 3.5 – – LV – KM(1,-) S $C_2$: LV is between 2-8 NGC7332 MC (KDC) 3 3 12 KT DR DR KM(1,-,-) F $C_2$: continuously increasing $k_1$; KDC only in PA change NGC7457 MC (CLV) 2 3 – KT DR – KM(1,2) F $C_2$: continuously increasing $k_1$ --------- ---------- ------- ---------- ---------- ------- ------- ------- ------------- --------- ------------------------------------------------------------------------------ \ Notes: (1) Galaxy identifier; (2) Kinematic galaxy group: see text for details; (3) Number of kinematic components; (4) Transition radius between the first and second components (arcsec); (5) Transition radius between the second and third components (arcsec); (6), (7) and (8) Kinematic group for the first second and third components; (9) Local kinematic misalignment between luminosity-weighted averages of [[PA]{}$_{kin}$]{} and [[PA]{}$_{phot}$]{}: numbers refer to the kinematic component and \* notes that the [[PA]{}$_{kin}$]{} was determined in the region with $k_1$ mostly below $15 {\>{\rm km}\,{\rm s}^{-1}}$; (10) Rotator class: S – slow rotator, F – fast rotator (11) Comment. Looking at the kinemetric profiles of 48 SAURON galaxies (Fig. \[f:prof\]) the following general conclusions can be made: - [[PA]{}$_{kin}$]{} profiles are in general smooth and often constant (e.g. NGC2974) or mildly varying (e.g. NGC474). In some cases there are abrupt changes of up to $180\degr$ within $1-2\arcsec$ (e.g. NGC3608). - Profiles of the axial ratio [[q]{}$_{kin}$]{} are generally smooth and often similar to [[q]{}$_{phot}$]{} profiles (e.g. NGC1023, NGC3384, NGC4570). - There is a variety of $k_1$ profiles, most of them rise and flatten, but some continue to rise, while some drop (e.g. NGC4278, NGC4477 and NGC4546). - Considering the $k_5/k_1$ radial dependence, there are three kinds of objects: those that have the ratio below 0.02 along most of the radius (e.g. NGC2974), those that have the ratio greater than 0.1 along most of the radius (e.g. NGC4374) and those that have the ratio below 0.02 with one (or more) humps above this value (e.g. NGC2549). - Objects with $k_5/k_1 > 0.1$ along a significant part of the profile from SAURON data are all classified as slow rotators in Paper IX (e.g. NGC4486). A further step in understanding the complex velocity maps can be made by applying definitions of kinemetric groups (see Section \[ss:def\_comp\]) to the radial profiles. They are summarised in Table \[t:kin\_prop\]. There are 15 galaxies characterised as SC (31%)[^4], the rest being MC galaxies (69%) of which 10 have KDC (21%) and 4 CLV (8%). Higher resolution observations with OASIS, however, show that all SAURON CLVs are in fact small KDCs and, moreover, CRCs (Paper VIII). This means that there are actually 14 (29%) KDCs in the SAURON sample. Kinematic profiles of the OASIS data also clearly show structures that are partially resolved in the SAURON observations, such as KDC (NGC4621, NGC5198, NGC5982) or co-rotating components which often have larger amplitudes of rotation in the OASIS data, corresponding to the nuclear disks visible on the HST images. The effects of specific nuclear kinematics, related mean ages of the components and possible different formation paths were previously discussed in Paper VIII. ![\[f:k1\_k51\] A relation between the maximum amplitude of rotation and luminosity-weighted average of $k_5/k_1$ ratio, which measures departures from the cosine law for velocity profiles extracted along the best fitting ellipse. All points belong to fast rotators. ](figures/fig2.ps){width="\columnwidth"} In addition to this grouping of the velocity maps, we can also describe the kinematic components. Out of 15 SC galaxies, 8 are DR (53%), 2 are KT (13%) and 4 are LV (27%), while the remaining galaxy (NGC2768) is a rather special case of a solid body rotator (see below). The majority of MC galaxies have only two kinematic components (30 or 91%), but there are some with three kinematic components (3 or 9%). Many of the components are similar in properties. If the inner component ($C_1$) is DR then the second component ($C_2$) is also a DR. If $C_1$ has a more complex kinematic structures (KDC, LV, KT), $C_2$ or $C_3$ will in most cases still be a DR. Exceptions are found in a few cases when $C_2$ is not rotating and can be described as LV. In some cases components may show KT, but their $k_5/k_1$ ratio is less than or equal to 0.02. Counting all galaxies that have at least one component with $k_5/k_1\lesssim0.02$, the number of systems with DR-like characteristics rises to 35 (73%). In terms of E/S0 classification, this means that 74% of Es and 92% of S0s have components with disk-like kinematics Using the resolving power of the OASIS data we can go even further: allowing for large error bars and allowing for considerable uncertainty of component boundaries, virtually every galaxy that shows rotation (including parts of KDCs) has at least one region with $k_5/k_1\lesssim0.02$ (0.04 in OASIS). This can be seen on Fig. \[f:k1\_k51\] that shows luminosity weighted average of $k_5/k_1$ ratio versus the maximum rotational amplitude $k_1$ for all fast rotators measured on the SAURON data. The large uncertainties in some cases reflect the multiple component nature of fast rotators, because $k_5/k_1$ ratio rises in the transition region between components [@2006MNRAS.366..787K]. Notably, both average values and uncertainties rise as the maximum rotation velocity decreases. This suggest a more complex structure (more components, larger difference between components) in galaxies with lower amplitude of rotation. We estimated local kinematic misalignment for each kinematic component. The results show that a total of 25 galaxies (52% of the sample) have some evidence of KM. It should, however, be kept in mind that it is difficult to determine the sense of rotation for LV components. Ignoring galaxies with a LV in the single component or in the second component, we are left with 15 galaxies with kinematic misalignments. Of these, 13 galaxies show misalignment in the first component, and 7 in the second component. Moreover, only 5 galaxies show misalignment in two components, as a global property within the SAURON field of view. As it can be seen from the kinemetric profiles there are special cases for which kinemetry is not able to determine the characteristic parameters robustly. This, in general, occurs when velocity drops below $\sim15{\>{\rm km}\,{\rm s}^{-1}}$ (e.g. NGC4486, NGC4550), and it is not surprising since these maps also often lack odd parity, an expected property of the first moment of the LOSVD. As discussed in @2006MNRAS.366..787K, another exception is the case of solid body rotation for which the iso-contours are parallel with the zero-velocity curve. In the SAURON sample this is seen in NGC2768. Since the velocity iso-contours are parallel the axial-ratio and the position angle are poorly defined and in practice strongly influenced by noise in the data. Determination of the best-fit ellipse parameters for the solid body rotation is degenerate. From these reasons this galaxy should also be considered with care when comparing with kinemetry results for other objects. Fast vs Slow Rotators {#ss:SRFR} --------------------- As stated above, all galaxies with $k_5/k_1 > 0.1$ along a significant part of the profile from SAURON data are classified as slow rotators in Paper IX. To some extent this is expected since slow rotators in general show very little rotation and kinemetry assumptions are not satisfied. The origin of the noise in the maps of slow rotators, which generates large $k_5/k_1$, is likely reflecting a special internal structure, and, in principal, does not come from technical aspects of the observations. Even velocity maps with low amplitude of rotation could show regular spider diagrams (e.g. disks seen at very low inclinations) observed at the same signal-to-noise. The one-to-one relation between slow rotators and objects with large higher-order harmonic terms is significant since the slow/fast rotator classification is based on both velocity and velocity dispersion maps and reflects the internal structure. The OASIS data cover only a small fraction of the effective radius and do not show this relationship. The slow rotators NGC3414, NGC3608, NGC5813 and NGC5982 show considerable rotation because the KDC is covering the full OASIS FoV, while the central regions of the fast rotators NGC2768, NGC3032 and NGC3379 still have small amplitudes of rotation and high (and noisy) $k_5/k_1$ ratios. The velocity maps of the 12 slow rotating galaxies in our sample can be described either as LV or as KDC+LV. In that respect, 4 slow rotators are SC systems (NGC4374, NGC4486, NGC4550 and NGC5846) which do not show any detectable rotation (at SAURON resolution) and the other 8 are MC systems where $C_1$ is a KDC and $C_2$ is a LV. In between these two cases is NGC4552 with a rather constant rotation velocity of $15{\>{\rm km}\,{\rm s}^{-1}}$, the boundary level for LV, and the $C_1$ between a KDC and a large KT. The other three quarters of galaxies in the SAURON sample are fast rotators. Only a third of them are described as SC, but as shown above (Section \[ss:subcomp\]) all fast rotators have components with kinematics that can be described as DR. Moreover, in the case of some slow rotators with small, but not negligible rotation in the centres (KDC), the higher resolution OASIS data were able to ascertain that these components have near to DR properties (within often large uncertainties) Figure \[f:radP\] has three panels highlighting most obvious kinematic properties of slow and fast rotators. The top panel shows radial variations of the kinematic position angles, [[PA]{}$_{kin}$]{}, which are present in various forms, ranging from minor twists in the nuclei, through abrupt jumps at the end of KDCs, to almost random changes with radius. However, only [[PA]{}$_{kin}$]{} of slow rotators are characterised by strong and rapid changes. Fast rotators show remarkably constant [[PA]{}$_{kin}$]{}. If twists are present in [[PA]{}$_{kin}$]{}of fast rotators, they are small in amplitude ($\lesssim30\degr$) and confined to the nuclear region in shapes of physically small KDCs (NGC4150, NGC4382, NGC4621, NGC7332, NGC7457). Slow rotators on the other hand show a much greater amplitude in change of [[PA]{}$_{kin}$]{}. It should be stressed again that determination of [[PA]{}$_{kin}$]{} for slow rotators is much more ill defined than for fast rotators in the sense that if there is no rotation, there is also no orientation of rotation. The abrupt changes in [[PA]{}$_{kin}$]{} are the consequence of this in some cases (NGC4374, NGC4486, NGC4550, NGC5846), and while one could debate the robustness of measured [[PA]{}$_{kin}$]{}, one should acknowledge the different nature of these systems from objects with a constant [[PA]{}$_{kin}$]{}. ![\[f:radP\] Radial profiles of (from top to bottom) [[PA]{}$_{kin}$]{} - $\langle$[[PA]{}$_{kin}$]{}$\rangle$, $k_1$ and $\sigma_0$ from the SAURON data. $\langle$[[PA]{}$_{kin}$]{}$\rangle$ are luminosity weighted averages of [[PA]{}$_{kin}$]{} profiles. $\sigma_0$ profiles are normalised at $R_e$/5. The profiles of slow and fast rotators are coloured in red and black, respectively. In the middle panel dashed blue lines are over-plotted to guide the eye for the cases with specific profiles as mentioned in text.](figures/fig3.ps){width="\columnwidth"} The difference between the slow and fast rotators is most visible in the amplitude of rotation. The middle panel of Figure \[f:radP\], shows the radial profiles of $k_1$ kinemetric terms for 48 early-type galaxies. Most of the profiles cover up to 1 $R_e$ in radius. Slow rotating galaxies can show a non-zero amplitude of rotation in the centres (KDCs), but the amplitude is, in general, not very high and towards the edge of the map it is mostly negligible. The only exception is NGC5982, which approaches two fast rotators with the slowest rotation in the outer regions (NGC4278 and NGC4473). Another characteristic of this plot is the variety of profiles. They include: monotonically rising profiles (e.g. NGC3032), profiles with an initial slow rise which turns to a more rapid one (e.g. NGC524), a rapid rise to a maximum followed by a plateau (e.g. NGC4546), rise to a maximum followed by a decrease (e.g. NGC4526), double hump profiles (e.g. NGC4660), flat profiles (e.g. NGC821) and, in slow rotators, profiles showing a decrease below our detection limit. Keeping in mind that the SAURON sample is not a complete sample, among fast rotators there are 17 (47%) with increasing profile at the edge of the SAURON map, 9 (25%) have flat profiles, 5 (14%) decreasing profiles and 4 galaxies have intermediate (difficult to classify) profiles. Among slow rotators there are 3 (25%) galaxies with increasing profiles at the edge of the map, the rest being flat and below the detection limit. These statistics are influenced by the size of velocity maps and the coverage of kinematics components by kinemetric ellipses. Clearly, larger scale observations would detect the end of rise in amplitude in galaxies that are now observed to have increasing $k_1$. Similarly, it is possible that a decrease in $k_1$ could be followed by an additional increase or a flat profile at large radii. Still, there are two general conclusions for fast rotators: they mostly show increasing velocity profiles at $1 R_e$, where the range of maximum velocity amplitude spans 200 ${\>{\rm km}\,{\rm s}^{-1}}$. On the other hand, slow rotators have velocity amplitude mostly less than $20 {\>{\rm km}\,{\rm s}^{-1}}$ at $1 R_e$. The bottom panel of Fig. \[f:radP\] shows radial velocity dispersion profiles, $\sigma_0$, extracted along the isophotes from the velocity dispersion maps as $a_0$ harmonic terms. All profiles are normalised to their value at $R_e/5$. This highlights the similar general shape of the $\sigma_0$ radial profiles. The only outlier is NGC4550 with an $\sigma_0$ profile which increases with radius. Most of the other profiles, while different in detail, show a general trend of increasing $\sigma_0$ towards the centre and also have a similar shape. A few profiles are consistent with being flat ($\Delta\sigma_0$/$\Delta R \lesssim 30 {\>{\rm km}\,{\rm s}^{-1}}$) over the whole profile (visible only in fast rotators such as NGC7457). If there are any real differences between $\sigma_0$ profiles, they are apparent for radii smaller than $R_e/5$. There are a few exceptions to the general trend: [*i)*]{} profiles with a decrease of more than 5% in the normalised $\sigma_0$ within $R_e/5$ (e.g. NGC4382), [*ii)*]{} profiles that are flat to within 5% inside the $R_e/5$ (e.g. NGC7457), and [*iii)*]{} profiles with a central rise followed by a drop and consecutive rise forming a profile with double maxima (e.g. NGC5813). These cases occur mostly in fast rotators, with a few exceptions in slow rotators. These central plateaus and drops are interesting, because classical theoretical work predicts that, for constant mass-to-light ratio, $r^{1/n}$ light-profiles have velocity dispersion minima in the centres of galaxies , unless they contain central black holes [e.g. @1998ApJ...498..625M]. The central $\sigma$-drops evidently do not occur frequently in real early-type galaxies, but are, perhaps marginally more common to fast rotators. About 10% of the SAURON early-types exhibit the central drop (but additional 20% have flat central profiles), which is much less compared to 46% among the Sa bulges, also observed with SAURON data [@2006MNRAS.369..529F]. Distribution of axial ratios {#ss:flat} ---------------------------- We now compare average photometric and kinematic axial ratios of SAURON galaxies. The axial ratio of a velocity map is related to the opening angle of the iso-velocity contours. In other words, the pinching of the contours in a spider diagram is related to the axial ratio of the best fitting ellipse given by kinemetry. As the kinematic axial ratio of slow rotators is an ill defined quantity (set to 1), in the rest of this section we focus on the average axial ratios of the fast rotators. In Fig. \[f:qkvqp\] we compare values of $\langle$[[q]{}$_{kin}$]{}$\rangle$ and $\langle$[[q]{}$_{phot}$]{}$\rangle$ for fast rotating galaxies. Since the typical seeing for SAURON data ranges up to $2\farcs5$, we exclude the inner $5\arcsec$ of the [[q]{}$_{kin}$]{} profiles from our derivation of the luminosity weighted average values (Appendix \[s:psf\]). The left hand panel shows a one-to-one correlation between the two quantities, although the scatter and uncertainties are large. The right hand panel shows more clearly the amount of scatter in these relations, as measured by the difference $\langle$[[q]{}$_{kin}$]{}$\rangle$ - $\langle$[[q]{}$_{phot}$]{}$\rangle$. The typical variation of the measured average $\Delta$[[q]{}$_{kin}$]{} is 0.1, as shown with vertical guidelines on the right panel. Outside this region there are about dozen galaxies. A few of these have $\langle$[[q]{}$_{kin}$]{}$\rangle$ larger than $\langle$[[q]{}$_{phot}$]{}$\rangle$; their photometric axial ratio is flatter than the kinematic, while the majority of outliers have the kinematic axial ratio flatter than the photometric. Let us consider in more detail only objects at significant distances from the vertical lines (i.e. $\langle$qk$\rangle - \langle$qp$\rangle > 0.15$): NGC821, NGC4270, NGC4473, NGC4621, NGC5308 and NGC5838. If we look at the [[q]{}$_{kin}$]{} profiles of the three galaxies (Fig. \[f:vel\_prof\]) with $\langle$[[q]{}$_{kin}$]{}$\rangle > \langle$[[q]{}$_{phot}$]{}$\rangle$, we can see that some of their kinematic sub-components have axial ratios very similar to the local photometric axial ratios. Notably, in the case of NGC5308 this is the outer component, especially at radii near to the edge of the SAURON FoV. The middle range, where the differences are the largest, is also the region of the transition between the two kinematic components. The mixing of the components changes the measured $\langle$[[q]{}$_{kin}$]{}$\rangle$, but it should be also noted that the $\langle$[[q]{}$_{phot}$]{}$\rangle$ varies over the whole map, becoming flatter and similar to $\langle$[[q]{}$_{kin}$]{}$\rangle$ with radius. This is not the case in NGC4473. Here the photometric axial ratio remains constant, but the kinematic axial ratio changes at larger radii. Again this change occurs in the transition region between the two kinematic components. Detailed dynamical modelling of this galaxy shows that it is made of two counter-rotating stellar components of unequal mass. This object is physically similar to NGC4550, where the main difference is in the mass fraction of the two components (Paper X). In the case of NGC4270 it is the [[q]{}$_{phot}$]{} that steadily changes with radius, while [[q]{}$_{kin}$]{} has mostly high values, but shows abrupt changes between the points. These are related to sometimes rather high values of $k_5/k_1$, which also changes abruptly between the adjacent points, a behaviour originating from the noisy transition region between two kinematic components. Since it is hard to disentangle the noise from the genuine physical signal in [[q]{}$_{kin}$]{}  measurement, and given the boxy appearance on the large scales, we note that this object could be a true and unusual outlier from the relation between [[q]{}$_{phot}$]{} and [[q]{}$_{kin}$]{}. Galaxies with $\langle$[[q]{}$_{kin}$]{}$\rangle<\langle$[[q]{}$_{phot}$]{}$\rangle$ are either single component (NGC821 and NGC4621, if we exclude the CRC in NGC4621 of $\sim 4\arcsec$ in size) or multi component (NGC5838). The HST image of NGC5838 has a prominent nuclear dust disk with the axial ratio between 0.3 – 0.4, constraining its inclination to about $70\degr$. It is possible that the velocity map is dominated by the presence of an associated stellar disk embedded in the galaxy body. NGC821 and NGC4621 were parametrised via the Multi-Gaussian Expansion in Paper IV. Both models required very flat Gaussians to reconstruct the light distribution (in both cases, the smallest axial ratios of the Gaussians of the MGE models were 0.3); these Gaussians are tracing disks embedded in spheroids. If we compare the $\langle$[[q]{}$_{kin}$]{}$\rangle$ with the flattest MGE Gaussians, both galaxies move well within the two vertical lines on right hand panel of Fig. \[f:qkvqp\] (blue lines). As a matter of interest, on the same right hand plot of Fig. \[f:qkvqp\] we overplotted axial ratios for the big KDCs in the sample: NGC3414, NGC3608, NGC5813 and NGC5831. The shape of their kinematics is similar with the distribution of light, except in the case of NGC3608. Its KDC has flatter kinematics than the light, but the flattest MGE Gaussian is comparable with the kinematic axial ratio. This suggest that even some of the sub-components of the slow rotating galaxies have similar properties like fast rotating galaxies. Having in mind that [[q]{}$_{kin}$]{} profiles can vary significantly with the radius, and the average values could be contaminated by the contribution of the transition regions between the components, we conclude that there is a near one-to-one correlation between average kinematic and photometric axial ratios in fast rotating galaxies, with a number of objects having $\langle$[[q]{}$_{kin}$]{}$\rangle < \langle$[[q]{}$_{phot}$]{}$\rangle$, and, hence, having disk-like components more visible in their kinematics than in photometry. Shape differences between velocity dispersion and surface brightness maps {#ss:shape} ------------------------------------------------------------------------- As shown in Section \[ss:prof\], the isophotes are not necessary a good representation of contours of constant velocity dispersion and the deviations are visible in the second cosine term ($b_2$) of the harmonic decomposition of the velocity dispersion profiles extracted along the isophotes. On Fig. \[f:b2a0\] we quantify the differences between isophotes and iso-$\sigma$ contours by plotting the normalised luminosity-weighted second term ($\langle b_2/a_0 \rangle _{\sigma}$) extracted along the isophotes of SAURON galaxies. Slow rotators are shown in red. Focusing on the left hand panel, it is clear that galaxies in the sample span a large range of $\langle b_2/a_0 \rangle _{\sigma}$, both positive and negative. If we exclude NGC4473 and NGC4550, there is a tail of galaxies with negative values up to $\sim-0.1$. Also, it seems that there is a trend of high negative $\langle b_2/a_0 \rangle _{\sigma}$ in flat galaxies: as galaxies become rounder, $\langle b_2/a_0 \rangle _{\sigma}$ tends to zero; above $\langle$[[q]{}$_{phot}$]{}$\rangle=0.8$ galaxies have small absolute value of $\langle b_2/a_0\rangle _{\sigma}$. Slow rotators are relatively round systems (Paper X) with [[q]{}$_{phot}$]{}$> 0.7$ (excluding the special case of NGC4550). Their $\langle b_2/a_0\rangle _{\sigma}$ values are small and mostly positive. Typical measurement error of $\langle b_2/a_0 \rangle _\sigma$ is $\sim 0.03$, and hence, the slow rotators are consistent with having velocity dispersion maps very similar to the the distribution of light, with possibly marginally flatter iso-$\sigma$ contours. Before we turn to the right panel on Fig. \[f:b2a0\], let us go back to Fig. \[f:diff\] and the example of NGC2549 and NGC4473. Their stellar velocity dispersion maps have different shapes from the distributions of light, but they have similar absolute values of $\langle b_2/a_0 \rangle _{\sigma}$; they fall on the opposite sides of Fig. \[f:b2a0\]. There is, however, evidence that in both cases the deviations from the photometry have similar physical origin. NGC2549 shows clear photometric and kinematic evidence for a stellar disk viewed at a high inclination (Fig. \[f:vel\_prof\]). The velocity dispersion map has a very specific ’bow-tie’ shape in the central 10, while outside its amplitude is low everywhere. The ’bow-tie’ shape can be explained with a light-dominating dynamically cold stellar disk which decreases the observed velocity dispersion. Outside of the disk, the bulge dominates and the observed velocity dispersion is again higher. It should be noted that the view of the galaxy at a high inclination is crucial for this effect to be seen. =0.79 =0.79 =0.79 NGC4473 was discussed in the previous section where we stated that it actually contains two sub-components. One of them is flat and counter-rotates with respect to the main galaxy body. This can explain the rise of the velocity dispersion along the major axis, specifically at larger radii where the contribution of the flat component becomes similar to the main body. This is also visible in the mean velocity map, where beyond 10 the velocity starts dropping. The only significant difference between these two examples is in the sense of rotation of the flat sub-components. Co-rotating disks or flattened components, viewed at favourable angles will likely show negative $\langle b_2/a_0 \rangle_\sigma$, while counter-rotating components will contribute to positive $\langle b_2/a_0 \rangle _\sigma$. In other words, co-rotation increases, while counter-rotation decreases the flattening of the iso-$\sigma$ contours. A similar case is NGC4550, the most extreme outlier, which also has two counter-rotating disks. NGC4150 and NGC3032 can be put in the same group: the OASIS measurements resolve their CRC components. It is interesting to note that the strongest $\sigma$-drop galaxies (NGC2768 and NGC4382) are both found on the positive side of the left hand panel of Fig. \[f:b2a0\], admittedly with small $\langle b_2/a_0 \rangle _\sigma$ ($<0.02$), indicating that the shapes of the velocity dispersion maps of $\sigma$-drop objects are somewhat different from the shapes of galaxies with central plateaus in velocity dispersion profiles or bow-tie shape of velocity dispersion maps. The shapes of $\sigma$-drops are more similar to the isophotes than the shape of iso-$\sigma$ contours in the objects with strong near-to edge-on disks such as NGC2549 or NGC3377. If $\sigma$-drops, bow-ties and central plateaus are all caused by disks, than the difference between them might originate from the orientation of the galaxy, where bow-ties are seen at edge-on and $\sigma$-drops at face on orientations. Given the uncertainties, it is hard to argue either way about other galaxies with slightly positive $\langle b_2/a_0 \rangle _\sigma$ values. Other effects, such as the actual size of the components or the presence of dust, may play a role on that level. This should be noted, but we, however, continue by suggesting that the main contributor to the shape difference between the zeroth and the second moment of the LOSVD in early-type galaxies are the embedded disks or flattened fast rotating components seen at different inclination. In this respect, on the right hand panel of Fig. \[f:b2a0\] we plot the absolute value of $\langle b_2/a_0\rangle _\sigma$. Here a trend is clear: from round slow rotators with generally small absolute values to flat fast rotating galaxies with increasing $|\langle b_2/a_0 \rangle _\sigma|$. This sequence is the one of increasing contribution of the ordered rotation towards the total energy budget, but it is, unfortunately, dependent on the viewing angles (NGC524 with the face-on disk in the top left corner of the diagram) and on the luminosity-weighted contribution of the flat component to the total kinematics (CRC in NGC4621 is not detected). Higher order moments of the LOSVD {#ss:high} --------------------------------- The final part of our analysis of the kinematic moments of the LOSVD is devoted to maps of $h_3$ and $h_4$, the Gauss-Hermite moments which measure deviations of the LOSVD from a pure Gaussian. They were introduced because LOSVD profiles are rather non-Gaussian, specifically the contribution of $h_3$, the skewness, along the major-axis of galaxies was found to have large amplitudes [e.g. @1994MNRAS.268..521V]. In a diagram of local $h_3 - V/\sigma$ relation, @1994MNRAS.269..785B plotted the major axis points of their objects which could be separated in two distinct groups made of galaxies with disky isophotes and non-disky isophotes. Both groups showed anti-correlation between $h_3$ and $V/\sigma$, where the points with small $V/\sigma$ values in the centre of the diagram had a steeper slope. It was recognised that the disky early-type galaxies cover a large range in $V/\sigma$ and are responsible for tails in upper left and lower right corner of the diagram, while non-disky early-types with small $V/\sigma$ values showed a somewhat smaller range in $h_3$ values. On the top panel of Fig. \[f:h3\_prof\] we plot the local $h_3 - V/\sigma$ relation for all data points of 48 SAURON galaxies. Red points belong to Voronoi bins of the slow rotators. This plot is somewhat different from previous findings. Paper X showed that slow rotators have small global $V/\sigma$ and this is also reflected in bin by bin values. The black points represent the Voronoi bins of fast rotators, and their distribution is different from what was found before for disky ellipticals. The shape of the distribution of black points can be described as a superposition of two components: one which is anti-correlated with $V/\sigma$ and makes distinct tails in upper left and lower right corner of the diagram (large positive and negative $V/\sigma$ values), and the other that is both consistent with $h_3\sim0$, and shows positive correlation at intermediate $V/\sigma$. Existence of $h_3 \sim 0$ distribution of points could be explained by the fact that, while @1994MNRAS.269..785B plot only the points along the major axis, we plot them all, and the horizontal distribution and correlating tails are the consequence of the minor axis contamination. Using kinemetry we extracted profiles of the dominant harmonic terms from $V$, $\sigma$, and $h_3$ maps along the same best-fitting ellipses to the velocity maps[^5]. We plot the $h_3 - V/\sigma$ profiles on the middle panel of Fig. \[f:h3\_prof\]. These profiles are basically major-axis representation of two-dimensional maps and are more comparable with previous major axis data. The slow rotators (in red) have the smallest $V/\sigma$ values and generally small amplitudes of $h_3$. The two trends for fast rotators from the left-hand panel are still visible. Fast rotators cover a range of $V/\sigma$ values, but some have large and some small $h_3$ amplitudes for a large $V/\sigma$ values. In some specific cases there is a suggestion that $h_3$ changes the sign becoming positive and correlating with $V/\sigma$ (e.g. NGC5308) at larger radii. A confirmation of this can also be seen on the bottom panel of Fig. \[f:h3\_prof\]. Here we plot the relation between luminosity averaged $h_3$ and $(V/\sigma)_e$ obtained within $1 R_e$, as advocated by @2005MNRAS.363..937B [values taken from Paper X]. The shown error bars are not statitcial uncertainties but describe the radial variation of $h_3$ profiles. This plot can be compared to Fig.14b of @1994MNRAS.269..785B, which also shows galaxies with $\langle h_3 \rangle\sim 0 $ for intermediate $V/\sigma_m$, while their empirical fitting relation (the solid line) describes the general trend in our data. =0.495 =0.495 The last kinematic moment to be analysed is given by the $h_4$ Gauss-Hermite coefficient. This moment describes the symmetric departures of the LOSVD from a Gaussian. It should be, however, kept in mind that it is very difficult to measure $h_4$ robustly, since it strongly depends on the effect of template mismatch. Also an inaccurate removal of the continuum will cause spurious $h_4$ values. On the left hand panel of Fig. \[f:h4\] we plot the local $h_4 - V/\sigma$ relation. Slow rotators with small $V/\sigma$ values, dominate the central region. They cover a range of positive $h_4$ values and somewhat extend below zero. Fast rotators fill in a cloud around slow rotators, equally filling negative and positive $V/\sigma$ part of the plot. There is a suggestion for a larger spread in $V/\sigma$ for positive $h_4$ values. Luminosity weighted average $\langle h_4\rangle$ values extracted with kinemetry along the isophotes are presented as a function of luminosity-weighted $V/\sigma$ measured within 1 $R_e$ (Paper X) in right panel of Fig. \[f:h4\]. As $(V/\sigma)_e$ increases, there is a marginal trend of an increased spread in the observed $h_4$ values, and galaxies with negative average values start to appear. This result is similar to @1994MNRAS.269..785B, in the sense that negative $\langle h_4\rangle$ appear for larger $(V/\sigma)_e$, but we do not find as negative values of $\langle h_4\rangle$. The only slow rotator with negative $\langle h_4 \rangle$, however, is the usual outlier, NGC4550, the special case of two counter-rotating disks. Discussion {#s:disc} ========== Velocity maps of slow and fast rotators {#ss:velmapSRFR} --------------------------------------- Radial profiles of kinemetric coefficients show that early-type galaxies are [*(i)*]{} multicomponent systems and [*(ii)*]{} in the majority of the cases contain a kinematic equivalent of a disk-like component. These statements are based on the empirical verification that the assumption of kinemetry holds. The assumption is that the azimuthal profiles, extracted from a velocity map along the best fitting ellipses, can be described with a simple cosine variation. Practically, this means that higher-order harmonic terms are negligible and at our resolution we find that the deviation from the pure cosine law is less than 2%, for about 80% of cases, at least on a part of the map. The multiple components are visible in the abrupt and localised changes of kinemetric coefficients. The central regions often harbour separate kinematic components which co- or counter-rotate with respect to the outer body. In some cases there are more than two components (e.g. NGC4382) or components are similar in size and co-spatial (e.g. NGC4473). Although the axial ratio profile in edge-on systems can change with the seeing effects (see Appendix \[s:psf\]), and in some cases the seeing can alter the map considerably, the kinematic sub-components are usually robust features. There are, however, early-type galaxies for which the deviations from the cosine law exceed 10% across a significant radial range. Such objects are all classified as slow rotators in Paper IX. The breakdown of the kinemetry assumptions is another evidence that these objects are intrinsically different. Certainly, we might not be able to apply kinemetry [*a priori*]{} in its odd version on velocity maps that do not show odd parity (e.g. NGC4486), but in the case of slow rotators that show a detectable level of rotation (e.g. NGC5982), the kinemetric analysis clearly shows differences from the maps of fast rotators (e.g. NGC4387). There are three types of intrinsic structures that will show small mean velocities: [*(i)*]{} face-on disks, [*(ii)*]{} two counter-rotating equal in mass and co-spatial components and [ *(iii)*]{} triaxial structures. [*(i) Face-on Disks.*]{} Low inclination thin disks will still have minimal deviations from the cosine law assumed in kinemetry. Although the amplitude of the rotation in such cases is small, the disk-like rotation (negligible $k_5/k_1$ coefficients) is expected at all projections except in the actual case of $i=0\degr$ with no rotation. Such an example in the sample is NGC524 ($i \sim 19\degr$). Based only on the velocity map, one could naively interpret that NGC4486 is a thin disk at $i=0\degr$, given that it has [[q]{}$_{kin}$]{}$~\sim 1$ and shows no rotation. However, other moments of the LOSVD strongly rule out this geometry, while, in general, other slow-rotators are not round enough to even be considered as such extreme cases. [*(ii) Counter-rotating components.*]{} This case can be outlined with NGC4550 and NGC4473. These galaxies are examples of axisymmetric objects with two counter-rotating and co-spatial components. The first one is classified as a slow rotator, while the other one is a fast rotator. In both cases, however, it is the mass fraction of the components that really decides what is observed. In the case of NGC4550 the masses are nearly equal and the luminosity weighted mean velocity is almost zero in the central region. NGC4550 is a product of a very specific formation process, but a clear example of how a superposition of two fast rotating components can imitate a slow rotator. In NGC4473, on the other hand, one component is more massive and dominates the light in the central $10\arcsec$, where the rotation is clearly disk like. Outside this region, where the counter-rotating component starts to significantly contribute to the total light, the shape of the velocity map changes, the amplitude of the rotation drops and non-zero $k_5/k_1$ coefficients are necessary to describe stellar motions. [*(iii) Triaxial objects.*]{} Observationally, these objects are marked by kinematic twists and kinematic misalignments, which are not present in axisymmetric galaxies. A strong restriction to the shape of the velocity maps of axisymmetric galaxies is that they should not have a radial variation of [[PA]{}$_{kin}$]{} or a misalignment between kinematic and photometric position angles. This, however, is not the case for triaxial galaxies, where the change of position angle can be influenced by a true change of the angular momentum vector, by the orientation of the viewing angles or by the relative dominance of different orbital families [@1991ApJ...383..112F; @1991AJ....102..882S; @2008MNRAS.385..647V]. Paper X found that, globally, kinematic misalignments are present only in slow rotating galaxies. Locally, however, we find both kinematic misalignments and twists in some fast rotating objects as well, but they are mostly confined to central components or clearly related to bars (e.g. NGC1023) or galaxies with shells (e.g. NGC474). We do not find strong kinematic twists typical of the extreme cases of maximum entropy models of triaxial galaxies projected at various viewing angles ([@1991AJ....102..882S]; see also @1994MNRAS.271..924A) for velocity maps of Stäckel triaxial models). Given that our sample is not representative of the luminosity function of local early-type galaxies in the sense that it contains too many massive galaxies, which are also more likely to show extreme features on the velocity maps, it is remarkable that we find that only a few velocity maps are similar to those predicted. Still, the observed velocity maps are divers (e.g. the difference between the maps of slow and fast rotators) and their complexity reflects the difference in their internal structure. The profiles of the relative change of the kinematic position angles from the first panel of Fig. \[f:radP\] suggest that fast and slow rotators have genuinely different intrinsic shapes, fast rotators being mostly axisymmetric and slow rotators weakly triaxial. This is also reflected in $k_5/k_1$ ratio. We suggest that the high values of this ratio in slow rotators, which is in practice caused by the noise in non-rotating velocity maps, has its origin in the internal orbital make up of these galaxies. Weakly triaxial slow rotators contain box orbits, and competing contributions of different tube orbit families, as opposed to more axisymmetric fast rotators with short-axis tubes as the only major orbit family [@1985MNRAS.216..273D][^6].This suggestion is also supported by the analysis of the orbital structure of collisionless merger remnants [@2005MNRAS.360.1185J] as well as by the kinemetric analysis of velocity maps of simulated binary disc merger remnants [@2007MNRAS.376..997J]. Evidence for disks in fast rotators ----------------------------------- Kinematic sub-components with azimuthal profiles that can be fitted with a cosine law are described as having a [*Disk-like rotation*]{} (DR). This does not mean that they are actual disks. It just suggests that velocity profiles of early-type galaxies extracted along the best-fitting ellipse [*resemble*]{} the velocity maps of thin disks in circular motion. The rate of occurrence of DRs is, however, striking. There is no reason why this should be the case in early-type galaxies, which in principle as a class can have a triaxial symmetry and complex motion in different planes. As suggested above, the link between the kinemetry assumption and the structure of fast rotating galaxies has its origin in their internal structure. The results of the dynamical models in Paper X reveal that fast rotators show evidence for a kinematically distinct flattened spheroidal component, suggesting that fast rotators are nearly oblate and contain flattened components. In addition to these dynamically cold components, the stellar populations of fast rotators show evidence for different chemical components. @2006MNRAS.369..497K find that all morphologically flat fast rotators have Mgb line-strength distribution flatter than the isophotes, and associate it with the rotationally supported substructure, which features a higher metallicity and/or an increased Mg/Fe ratio as compared to the galaxy as a whole. These are some of the dynamical and chemical evidences for disk-like components in fast rotators. What is the kinemetric evidence? As mentioned above, we find that velocity maps of fast rotators are mostly described by a simple cosine law, as are velocity maps of thin disks. We also find an almost one-to-one correspondence between the projected shape of the stellar distribution and the shape of the observed kinematic structure in fast rotating galaxies. The connection between the shape and the kinematics is supported by an assumption that rotation influences the shape of the object by flattening it and, for isotropic models, the rotation speed responsible for flattening of the shape is related to the shape of the stellar distribution as $\sim \sqrt{\epsilon}$ [@2008gady.book.....B Section 4.8.2]. In order to investigate further the $\langle $[[q]{}$_{phot}$]{}$\rangle - \langle $[[q]{}$_{kin}$]{}$\rangle$ correlation we constructed two-integral analytic models of early-type galaxies. The isotropic models we used were previously presented in Appendix B of Paper X, to which we refer the reader for more details. The main point of these Jeans models is that we used as templates 6 galaxies, which represent some of the typical types from the SAURON sample. Their light distribution was parameterised in Paper IV by the MGE method, and was used as the basis for the intrinsic density distributions. Observables of each Jeans model were projected at 6 different inclinations: $90\degr$, $80\degr$, $65\degr$, $50\degr$, $35\degr$ and $25\degr$. These models are not meant to reproduce the observed kinematics in detail, but they are self-consistent, and under the assumption of axisymmetry and isotropy, they predict velocity maps and offer an opportunity to study the relation between the shape and kinematics. On the left hand panel of Fig. \[f:jeans\] we show the difference between the luminosity-weighted average values of $\langle$[[q]{}$_{kin}$]{}$\rangle$ and $\langle$[[q]{}$_{phot}$]{}$\rangle$, measured by kinemetry on the model images and velocity maps, in the same way as for the SAURON data in Fig. \[f:qkvqp\] (also excluding the inner $5\arcsec$). Different colours represent Jeans models based on different template galaxies. Each symbol corresponds to a model at different inclination, where the points move from top to bottom with increasing inclination (from $25-90\degr$). It is clear that for small inclination $\langle$[[q]{}$_{kin}$]{}$\rangle$ $\approx$ $\langle$[[q]{}$_{phot}$]{}$\rangle$, but as the models are viewed closer to edge-on there is a trend of increasing differences between the axial ratios and in some cases a trend of larger variation along the profiles represented by larger error bars for progressively more inclined models. Specifically, in all but one marginal case $\langle$[[q]{}$_{kin}$]{}$\rangle>\langle$[[q]{}$_{phot}$]{}$\rangle$: in these models velocity maps are ‘rounder’ than images. The velocity map of the Jeans model of NGC4621, whose MGE parametrisation has the flattest Gaussian, has a flat component along the major axis, which becomes more prominent with increasing inclination, contributing to the radial variation and increasing [[q]{}$_{kin}$]{} with respect to [[q]{}$_{phot}$]{}. The kinematics of this component is the most ’disk-like’ in our Jeans models with tightly pinched iso-velocity contours. For comparison, the model for NGC3377 also has a similar thin MGE component. This component contributes to the disk-like kinematic component confined to the central region, but since it is not as prominent in the total light, such as the one in the NGC4621 MGE model, only traces of the disk-like component can be seen in both photometry and kinematics. The contrast between observed galaxies and isotropic models is significant. The isotropic models predict $\langle$[[q]{}$_{kin}$]{}$\rangle \ge \langle$[[q]{}$_{phot}$]{}$\rangle$ for $i\gtrsim 30\degr$, but the models also show that prominent disk-like photometric features will generate pinched iso-velocity contours and decrease $\langle$[[q]{}$_{kin}$]{}$\rangle$ pushing the galaxies towards the observed trend of, on average, $\langle$[[q]{}$_{kin}$]{}$\rangle \sim \langle$[[q]{}$_{phot}$]{}$\rangle$ or even $\langle$[[q]{}$_{kin}$]{}$\rangle<\langle$[[q]{}$_{phot}$]{}$\rangle$. It seems reasonable to assume that while isotropic models can explain certain features of fast rotators, they cannot describe them as a class of galaxies. It is the embedded flattened components, often visible only in the kinematics that are responsible for the observed differences between the models and the data. This can be also seen by comparing the difference in shape between the isophotes and iso-$\sigma$ contours (Fig. \[f:b2a0\]). We repeated the same exercise with velocity dispersion maps of our isotropic Jeans models. Result are shown on the right panel of Fig. \[f:jeans\]. The trend shown here of a larger absolute $\langle b_2/a_0 \rangle _\sigma$ values with lower $\langle$[[q]{}$_{phot}$]{}$\rangle$ is very similar to the observed trend (hatched region). The test can also explain the shape of the observed trend: large absolute values of $\langle b_2/a_0 \rangle _\sigma$ can be observed when the object contains a significant flattened component and it is observed at larger inclinations. The most affected are again NGC4621 and NGC3377 models. As before, the isotropic models are able to explain part of the observed data, but not the details of the distribution. Specifically, the shape of the iso-$\sigma$ contours of slow rotators ($\langle b_2/a_0 \rangle _\sigma > 0$) is not reproduced well by the isotropic models. Similarly, the spread in $\langle b_2/a_0 \rangle _\sigma$ of fast rotators is also not well reproduced. Clearly, our Jeans models are much simpler than the real galaxies, lacking by construction multiple kinematic and especially counter-rotating components. Comparing the kinemetric analysis of the Jeans models and the observed objects, we find that fast rotating galaxies are more complex than isotropic rotators, presumably containing also flattened kinematically distinct components, which can co- or counter-rotate on top of the non-rotating or isotropically rotating spheroid. The evidence for disks in fast rotators are also present in the ratio of $h_3$ and $V/\sigma$. $h_3$ measures the asymmetric deviations from the Gaussian LOSVD and the anti-correlation of $h_3$ with $V/\sigma$ is taken to show presence of disks. Our results confirm previous findings that early-type galaxies on the whole have asymmetric LOSVDs, but this applies only to fast rotators. We also find that many fast rotators show constant and close to zero $h_3$ profiles, or, in a few cases, show a change from negative to positive values with increasing radius, similar to what is seen in peanut bulges and bars [@2004AJ....127.3192C; @2005ApJ...626..159B] The embedded flattened components in fast rotators are also evident when $h_3 - V/\sigma$ diagram is compared with the results of merger simulations . Specifically, the updated $h_3 - V/\sigma$ diagram in Fig. \[f:h3\_prof\] can be compared with Fig.16 from @2006MNRAS.372..839N, who discuss the influence of dissipational mergers in which embedded disks are formed in merger remnants. Our figure compares rather well with a combination of 1:1 dry and 1:3 wet mergers. This comparison suggest that it is not possible to explain the LOSVD of early-type galaxies with one merging track only, but that slow rotators predominantly originate in major colissionless mergers, while fast rotators are remnants of dissipational mergers. The comparison of our bin-by-bin $h_4 - V/\sigma$ diagram (right-hand panel in Fig. \[f:h4\]) and lower four panels of Fig.16 in @2006MNRAS.372..839N is equally impressive, although their simulations predict somewhat too negative values of $h_4$. Again, the observations are largely consistent with the scenario where slow rotators originate from dry 1:1 mergers, while fast rotators from a combination of dry and wet 1:3 mergers. Combining all the evidence presented in the previous section, we suggest that fast rotators are dominated by disks (e.g. NGC3156, NGC2685). When their light is dominated by the bulge, their kinematics still show strong disk components (e.g. NGC821, NGC4660). In either case, [*fast rotators contain flattened fast rotating components and this dynamical property differentiates them from slow rotators*]{}. We suggest that with increasing specific angular momentum, $\lambda_R$, the relative mass of the embedded disks also increases and contributes more significantly to the total mass. Among the disk-like components in fast rotators there is a range of flattenings reflecting a diversity in possible formation paths which create, preserve and/or thicken disks within spheroids, such as passive fading of spirals and multiple minor dissipational mergers. The change within the internal structure is observationally reflected in the transition between slow and fast rotators, which offers possible anchor points for theoretical models of galaxy evolution. Conclusions {#s:conc} =========== Using kinemetry we analysed two-dimensional maps of 48 early-type galaxies observed with SAURON and velocity maps of a sub-sample observed with OASIS. The analysed maps are: reconstructed image, mean velocity, velocity dispersion, $h_3$ and $h_4$ Gauss-Hermite moments. The reconstructed images and the maps of the mean velocity were analysed along the best fit ellipses. Maps of the higher moments of the LOSVD were analysed either along the isophotes (velocity dispersion, $h_4$) or best-fitting ellipses from velocity maps ($h_3$). We presented the profiles of kinemetric coefficients for velocity maps, being the dominant kinematic moment of the LOSVD and having the highest signal-to-nose ratio. Kinemetry and its kinemetric coefficients allow us [*(i)*]{} to differentiate between slow and fast rotators from velocity maps only and [*(ii)*]{} to indicate that fast rotating galaxies contain disks with a larger range in the mass fractions to the main body. The results of this work can be summarised as follows: - Following the kinemetric analysis of the velocity maps it is possible to distinguish between galaxies with [*Single*]{} and [ *Multiple Components*]{}. Components can be described as having a [ *Disk-like rotation*]{}, [*Low-level velocity*]{}, [*Kinematic misalignment*]{}, [*Kinematic twist*]{} or being [*Kinematically Decoupled Component*]{}. The sorting of galaxies in these groups is based on kinemetric coefficients: [[PA]{}$_{kin}$]{}, [[q]{}$_{kin}$]{}, $k_1$ and $k_5/k_1$. - The majority of early-type galaxies are MC systems (69%) - The total fraction of galaxies with a DR component (including all that have $k_5/k_1<0.02$) is 81%. In terms of S0/E classification, 92% of S0s and 74% of Es have components with disk-like kinematics. - KDCs are found in 29% of galaxies. These KDC are of different sizes, some are not resolved at SAURON resolution and appear as CLVs, but are clearly detectable in the OASIS observations at higher spatial resolution. - Early-type galaxies with constant [[PA]{}$_{kin}$]{}($\Delta PA <10\degr$ outside any central component) are fast rotators. All fast-rotators have a DR component. - Most fast rotators have $\langle$[[q]{}$_{kin}$]{}$\rangle \lesssim \langle$[[q]{}$_{phot}$]{}$\rangle$. Their kinematics often show a structure flatter than the distribution of light. This means that images alone are not sufficient to recognise all fast rotators (e.g. NGC524 or N3379 would be missed) - In face-on galaxies, isophotes are coincident with iso-$\sigma$ contours. In edge-on galaxies, however, there are differences which can be detected with kinemetry. Specifically, the edge-on fast rotators contain dynamically cold components which changes the shape of iso-$\sigma$ contours. - Slow rotating galaxies have low $V/\sigma$ and a range in $h_3$ amplitudes. Fast rotating galaxies have large spread in $h_3$ which anti-correlates with $V/\sigma$. There is, however, a significant number of fast rotating galaxies which have $h_3$ radial profiles that change from strongly negative to near to zero $h_3$ values. Similar behaviour was found in bar galaxies or in remnants of colissionless mergers. - Allowing for large uncertainties in the determination of $h_4$, our data show a trend where slow and fast rotators have similar $h_4$ values, with a weak tendency for an increased spread of $h_4$ in fast rotators, also dependant on $V/\sigma$. These trends can be explained through a combination of dry and wet mergers of both equal and unequal in mass progenitors. - Dissimilarities between slow and fast rotators originate in their different internal structures. Slow rotators are mildly triaxial objects supporting a variety of orbital families. Fast rotators are axisymmetric spheroids with embedded flattened components of different mass fraction ranging from completely disk dominated systems to small central disks visible only in kinematics. [**Acknowledgements**]{}\ DK acknowledges the support from Queen’s College Oxford and hospitality of Centre for Astrophysics Research at the University of Hartfordshire. The SAURON project is made possible through grants 614.13.003, 781.74.203, 614.000.301 and 614.031.015 from NWO and financial contributions from the Institut National des Sciences de l’Univers, the Université Lyon I, the Universities of Durham, Leiden, and Oxford, the Programme National Galaxies, the British Council, PPARC grant ‘Observational Astrophysics at Oxford 2002–2006’ and support from Christ Church Oxford, and the Netherlands Research School for Astronomy NOVA. RLD is grateful for the award of a PPARC Senior Fellowship (PPA/Y/S/1999/00854) and postdoctoral support through PPARC grant PPA/G/S/2000/00729. The PPARC Visitors grant (PPA/V/S/2002/00553) to Oxford also supported this work. GvdV acknowledges support provided by NASA through grant NNG04GL47G and through Hubble Fellowship grant HST-HF-01202.01-A awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS 5-26555. JFB acknowledges support from the Euro3D Research Training Network, funded by the EC under contract HPRN-CT-2002-00305. This paper is based on observations obtained at the William Herschel Telescope, operated by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofísica de Canarias. Based on observations obtained at the Canada-France-Hawaii Telescope (CFHT) which is operated by the National Research Council of Canada, the Institut National des Sciences de l’Univers of the Centre National de la Recherche Scientifique of France, and the University of Hawaii. This project made use of the HyperLeda and NED databases. Part of this work is based on data obtained from the ESO/ST-ECF Science Archive Facility. R., [de Zeeuw]{} P. T., [Hunter]{} C., 1994, , 271, 924 R., [Adam]{} G., [Baranne]{} A., [Courtes]{} G., [Dubet]{} D., [Dubois]{} J. P., [Emsellem]{} E., [Ferruit]{} P., [Georgelin]{} Y., [Monnet]{} G., [Pecontal]{} E., [Rousset]{} A., [Say]{} F., 1995, , 113, 347 R., [Copin]{} Y., [Monnet]{} G., [Miller]{} B. W., [Allington-Smith]{} J. R., [Bureau]{} M., [Carollo]{} C. M., [Davies]{} R. L., [Emsellem]{} E., [Kuntschner]{} H., [Peletier]{} R. F., [Verolme]{} E. K., [de Zeeuw]{} P. T., 2001, , 326, 23 M., 1991, , 249, L9 R., 1988a, , 202, L5 R., 1988b, , 193, L7 R., [Doebereiner]{} S., [Moellenhoff]{} C., 1988, , 74, 385 R., [Saglia]{} R. P., [Gerhard]{} O. E., 1994, , 269, 785 R., [Surma]{} P., [Doebereiner]{} S., [Moellenhoff]{} C., [Madejsky]{} R., 1989, , 217, 35 G. J., [Barnes]{} J. E., 2000, , 316, 315 F., [Capaccioli]{} M., 1975, , 200, 439 J., 1978, , 183, 501 J., 1980, , 190, 873 J., 2005, , 363, 937 J., [Tremaine]{} S., 2008, [Galactic dynamics]{}. Princeton, NJ, Princeton University Press, 2008, second edition M., [Athanassoula]{} E., 2005, , 626, 159 M., 2002, , 333, 400 M., [Bacon]{} R., [Bureau]{} M., [Damen]{} M. C., [Davies]{} R. L., [de Zeeuw]{} P. T., [Emsellem]{} E., [Falc[ó]{}n-Barroso]{} J., [Krajnovi[ć]{}]{} D., [Kuntschner]{} H., [McDermid]{} R. M., [Peletier]{} R. F., [Sarzi]{} M., [van den Bosch]{} R. C. E., [van de Ven]{} G., 2006, , 366, 1126 M., [Copin]{} Y., 2003, , 342, 345 M., [Emsellem]{} E., [Bacon]{} R., [Bureau]{} M., [Davies]{} R. L., [de Zeeuw]{} P. T., [Falc[ó]{}n-Barroso]{} J., [Krajnovi[ć]{}]{} D., [Kuntschner]{} H., [McDermid]{} R. M., [Peletier]{} R. F., [Sarzi]{} M., [van den Bosch]{} R. C. E., [van de Ven]{} G., 2007, , 379, 418 A., [Bureau]{} M., 2004, , 127, 3192 L., [Lanzoni]{} B., 1997, , 321, 724 E. M., [Wegner]{} G., [Saglia]{} R. P., [Thomas]{} J., [Bender]{} R., [Thomas]{} D., 2008, , 175, 462 R. L., [Efstathiou]{} G., [Fall]{} S. M., [Illingworth]{} G., [Schechter]{} P. L., 1983, , 266, 41 P. T., 1985, , 216, 273 P. T., [Bureau]{} M., [Emsellem]{} E., [Bacon]{} R., [Carollo]{} C. M., [Copin]{} Y., [Davies]{} R. L., [Kuntschner]{} H., [Miller]{} B. W., [Monnet]{} G., [Peletier]{} R. F., [Verolme]{} E. K., 2002, , 329, 513 S., 1983, Journal of Astrophysics and Astronomy, 4, 271 E., [Cappellari]{} M., [Krajnovi[ć]{}]{} D., [van de Ven]{} G., [Bacon]{} R., [Bureau]{} M., [Davies]{} R. L., [de Zeeuw]{} P. T., [Falc[ó]{}n-Barroso]{} J., [Kuntschner]{} H., [McDermid]{} R., [Peletier]{} R. F., [Sarzi]{} M., 2007, , 379, 401 E., [Cappellari]{} M., [Peletier]{} R. F., [McDermid]{} R. M., [Geacon]{} R., [Bureau]{} M., [Copin]{} Y., [Davies]{} R. L., [Krajnovi[' c]{}]{} D., [Kuntschner]{} H., [Miller]{} B. W., [de Zeeuw]{} P. T., 2004, , 352, 721 E., [Monnet]{} G., [Bacon]{} R., 1994, , 285, 723 J., [Bacon]{} R., [Bureau]{} M., [Cappellari]{} M., [Davies]{} R. L., [de Zeeuw]{} P. T., [Emsellem]{} E., [Fathi]{} K., [Krajnovi[ć]{}]{} D., [Kuntschner]{} H., [McDermid]{} R. M., [Peletier]{} R. F., [Sarzi]{} M., 2006, , 369, 529 M., [Illingworth]{} G., [de Zeeuw]{} P. T., 1991, , 383, 112 K., [Richstone]{} D., [Tremaine]{} S., [Lauer]{} T. R., [Bender]{} R., [Bower]{} G., [Dressler]{} A., [Faber]{} S. M., [Filippenko]{} A. V., [Green]{} R., [Grillmair]{} C., [Ho]{} L. C., [Kormendy]{} J., [Magorrian]{} J., [Pinkney]{} J., 2003, , 583, 92 O., [Kronawitter]{} A., [Saglia]{} R. P., [Bender]{} R., 2001, , 121, 1936 O. E., 1993, , 265, 213 A. C., [Balcells]{} M., [Olshevsky]{} V. S., 2006, , 372, L78 C., [Davies]{} R. L., [Kuntschner]{} H., [Birkinshaw]{} M., [Bender]{} R., [Saglia]{} R. P., [Baggley]{} G., 2001, , 326, 473 G. K. T., [Forbes]{} D. A., 2006, , 371, 633 L., 1990, , 356, 359 E. P., 1936, Yale University Press G., 1977, , 218, L43 R. I., 1987, , 226, 747 R., [Naab]{} T., [Burkert]{} A., 2005, , 360, 1185 R., [Naab]{} T., [Peletier]{} R. F., [Burkert]{} A., 2007, , 376, 997 W., [Zeilinger]{} W. W., 2000, , 145, 71 J., 1982, in [Martinet]{} L., [Mayor]{} M., eds, Saas-Fee Advanced Course 12: Morphology and Dynamics of Galaxies [Observations of galaxy structure and dynamics]{}. pp 113–288 J., [Bender]{} R., 1996, , 464, L119+ J., [Illingworth]{} G., 1982, , 256, 460 D., [Cappellari]{} M., [de Zeeuw]{} P. T., [Copin]{} Y., 2006, , 366, 787 A., [Saglia]{} R. P., [Gerhard]{} O., [Bender]{} R., 2000, , 144, 53 H., [Emsellem]{} E., [Bacon]{} R., [Bureau]{} M., [Cappellari]{} M., [Davies]{} R. L., [de Zeeuw]{} P. T., [Falc[ó]{}n-Barroso]{} J., [Krajnovi[ć]{}]{} D., [McDermid]{} R. M., [Peletier]{} R. F., [Sarzi]{} M., 2006, , 369, 497 T. R., 1985, , 57, 473 R. M., [Emsellem]{} E., [Shapiro]{} K. L., [Bacon]{} R., [Bureau]{} M., [Cappellari]{} M., [Davies]{} R. L., [de Zeeuw]{} T., [Falc[ó]{}n-Barroso]{} J., [Krajnovi[ć]{}]{} D., [Kuntschner]{} H., [Peletier]{} R. F., [Sarzi]{} M., 2006, , 373, 906 D., [Quinlan]{} G. D., 1998, , 498, 625 T., [Burkert]{} A., 2001, , 555, L91 T., [Jesseit]{} R., [Burkert]{} A., 2006, , 372, 839 R. F., [Davies]{} R. L., [Illingworth]{} G. D., [Davis]{} L. E., [Cawson]{} M., 1990, , 100, 1091 E. E., [de Zeeuw]{} P. T., [van der Marel]{} R. P., [Hunter]{} C., 1995, , 274, 602 H., [White]{} S. D. M., 1990, , 362, 52 H.-W., [Carollo]{} C. M., [Freeman]{} K., 1999, , 513, L25 H.-W., [Franx]{} M., [Fisher]{} D., [Illingworth]{} G., 1992, , 400, L5 V. C., [Graham]{} J. A., [Kenney]{} J. D. P., 1992, , 394, L9 B. S., [Terndrup]{} D. M., [Pogge]{} R. W., [Lauer]{} T. R., 1999, , 517, 650 R. P., [Bertschinger]{} E., [Baggley]{} G., [Burstein]{} D., [Colless]{} M., [Davies]{} R. L., [McMahan]{} Jr. R. K., [Wegner]{} G., 1997, , 109, 79 A., 1961, [The Hubble atlas of galaxies]{}. Washington: Carnegie Institution, 1961 A., [Bedke]{} J., 1994, [The Carnegie atlas of galaxies]{}. Washington, DC: Carnegie Institution of Washington with The Flintridge Foundation, |c1994 A., [Sandage]{} M., [Kristian]{} J., 1975, [Galaxies and the Universe]{}. Galaxies and the Universe F., 1979, , 233, 23 F., 1981, , 86, 662 T. S., 1991, , 102, 882 S. D., 1987, in [de Zeeuw]{} P. T., ed., Structure and Dynamics of Elliptical Galaxies Vol. 127 of IAU Symposium, [Summary - Structure and Dynamics of Elliptical Galaxies]{}. pp 367–+ S., 1990, , 348, 57 R. C. E., [van de Ven]{} G., [Verolme]{} E. K., [Cappellari]{} M., [de Zeeuw]{} P. T., 2008, , 385, 647 P. C., [Allen]{} R. J., 1978, , 16, 103 R. P., 1991, , 253, 710 R. P., 1994, , 270, 271 R. P., [Franx]{} M., 1993, , 407, 525 R. P., [Rix]{} H. W., [Carter]{} D., [Franx]{} M., [White]{} S. D. M., [de Zeeuw]{} T., 1994, , 268, 521 G., [Corsini]{} E. M., [Saglia]{} R. P., [Bender]{} R., [Merkl]{} D., [Thomas]{} D., [Thomas]{} J., [Mehlert]{} D., 2002, , 395, 753 Influence of seeing on the shape of the velocity maps {#s:psf} ===================================================== ![image](figures/figA1.ps){width="\textwidth"} The effects of seeing on galaxy images in the regime of non-adaptive optic assisted observations are now well understood [@1997ApJS..109...79S]. They are specifically important in the nuclear regions where the instrumental effects and the atmospheric seeing redistribute photons sufficiently enough to change the intrinsic ellipticity or position angle and mask structures like concentrated stellar nuclei or double nuclei [@1979ApJ...233...23S; @1981AJ.....86..662S; @1983JApA....4..271D; @1985ApJS...57..473L]. Specifically, luminosity profiles are affected more by the seeing when the ellipticity is low, while the ellipticity profiles are affected more when the ellipticity is high [@1990AJ....100.1091P]. In general, effects of seeing can be detected to quite large radii. ![image](figures/figA2.ps){width="\textwidth"} It is natural to expect the redistribution of photons due to the seeing will also influence the two-dimensional shape (maps) of the higher moments of the LOSVD. Specifically, for the first moment, the mean velocity, the obvious effect of larger seeing is a less steep rise of the velocity in the centre. Additionally, one can expect an increased axial ratio (rounder map), or in other words, a more open spider diagram in the central regions. Except for these qualitative expectations there are no quantitative estimates of the influence of seeing on the shape of velocity maps. Observations of the SAURON sample were taken under a variety of atmospheric conditions, so before analysing the velocity maps, we tested the effects of different PSFs. We constructed a number of model velocity maps using a @1990ApJ...356..359H potential which reasonably well approximates the density of real early-type galaxies. Each model map was constructed from the Hernquist circular velocity profile, projected at a certain inclination to imitate different axial ratios seen in the sample ([[q]{}$_{kin}$]{}), and weighted by the given Hernquist surface brightness profiles. All models have the same scale length $r_0=10\arcsec$. The inclinations ranged from 20 to 75 degrees, where 0 and 90 degrees are face on and edge on viewing angles, respectively. This corresponds to velocity maps with axial ratios [[q]{}$_{kin}$]{} in the range of 0.94 to 0.26. Although in the SAURON sample there are near to edge on galaxies, their velocity maps do not resemble the limiting case of edge on disks ([[q]{}$_{kin}$]{}=0), being considerably less flat. The above range in axial ratios were used to match the values observed in the SAURON sample. Maps were then convolved with a Gaussian kernel of different full-width-half-maximum (FWHM), accounting also for the square pixels [@1995MNRAS.274..602Q] in such a way as to make the model maps similar to observations, especially in the central regions which remain unbinned. The range of FWHM was also matched to the measured seeing range of SAURON observations (Paper III). Figure \[f:maps\_fwhm\] shows an example of model maps for axial ratio of 0.5 and seeing FWHMs of $1\arcsec$, $2\arcsec$ and $3\arcsec$. We ran kinemetry on the convolved model velocity maps and extracted the same parameters as for the observed velocity maps: [[PA]{}$_{kin}$]{}, [[q]{}$_{kin}$]{}, $k_{1}$ and $k_{5}/k_{1}$ harmonics. The results can be seen in Fig. \[f:prof\_fwhm\] and they suggest that: - determination of the position angle is not influenced by the seeing; - the influence of the seeing on the determined axial ratio is larger on maps with intrinsically smaller axial ratio. Also, the bigger the seeing is, the rounder (larger axial ratio) is the observed map. - the influence of the seeing on the rotation curve ($k_1$ term) is larger on maps with intrinsically smaller axial ratios, where the bigger is the seeing, the lower is the maximum velocity. The position of maximum velocity is also pushed towards larger radii. - the influence of the seeing on the higher terms ($k_{5}/k_{1}$) is negligible in most cases, except for intrinsically very flat maps, but even there the signal is of the order of 1% or less and is not detectable with the current accuracy. As expected, the seeing influences the axial ratio of velocity maps (or, in other words, the opening of the iso-velocity contours), and the steepness of the rotation curve, both in the maximum values and the positions of these maxima. The largest effect is for maps with small axial ratio (large opening angles), which can be understood as the consequence of the convolution, similar as in the case of photometry and high ellipticity. In terms of changing of the intrinsic values, it is the axial ratio that is the most affected by seeing. At an intrinsic flattening of 0.77, a bad seeing of $2\farcs5$, will increase the flattening by 0.1. This effect rises to surprising 0.5 difference in measured flattening for an intrinsic flattening of 0.26. This effect is milder for the maximum rotation velocity, where only for velocity maps with the smallest axial ratios and strongest seeings the difference between the intrinsic and seeing convolved values exceeds 10 km/s. Another effect of seeing on the velocity maps is the change of position of the maximum axial ratio and the maximum rotational velocity with increasing seeing. In both cases the trend is clear: larger seeing FWHM causes displacement of the maximum ([[q]{}$_{kin}$]{} or $k_1$) towards larger radii. In summary, seeing can strongly influence the appearance of the velocity maps, both in the shape, value and extent of the features. This is important for interpreting the structures on the maps and, especially, when looking for kinematic subcomponents. Kinemetric profiles {#s:profiles} =================== In this section we present the results of kinemetric analysis of SAURON velocity maps. ![image](figures/figB1a_lr.ps){width="98.00000%"} ![image](figures/figB1c_lr.ps){width="98.00000%"} ![image](figures/figB1e_lr.ps){width="\textwidth"} ![image](figures/figB1g_lr.ps){width="\textwidth"} ![image](figures/figB1i_lr.ps){width="\textwidth"} ![image](figures/figB1k_lr.ps){width="\textwidth"} ![image](figures/figB1m_lr.ps){width="\textwidth"} ![image](figures/figB1o_lr.ps){width="\textwidth"} ![image](figures/figB1q_lr.ps){width="\textwidth"} ![image](figures/figB1s_lr.ps){width="\textwidth"} ![image](figures/figB1u_lr.ps){width="\textwidth"} ![image](figures/figB1w_lr.ps){width="\textwidth"} \[lastpage\] [^1]: E-mail: [email protected] [^2]: Hubble Fellow [^3]: It is customary in the literature to refer to the maps of mean velocity as [*velocity fields*]{}. Sometimes, due to the specific shape of contours of constant velocities, velocity maps are referred to as [*spider diagrams*]{} . Two-dimensional representations of the next moment are, however, usually referred to as velocity dispersion maps. Instead of alternating between [*fields*]{} and [*maps*]{} we choose to call all two-dimensional representations of the LOSVD maps: velocity map, velocity dispersion map, $h_3$ map.... [^4]: Although NGC4550 is made of two counter-rotating and co-spatial disks (Section \[ss:prof\]), it is formally characterised as a SC galaxy due to its low velocity within the SAURON FoV. [^5]: Note that here we re-run kinemetry on $\sigma$ maps along the best fitting ellipses for velocity maps, except in the case of the slow rotators when we used circles as explained in Section \[ss:prof\]. These profiles are somewhat different from the profiles presented in Section \[ss:shape\] which were obtained along the isophotes. [^6]: We should keep in mind that two short-axis tube families with opposite angular momentum in certain cases can produce axisymmetric objects which appear as slow rotators, as mentioned above in the case of NGC4550.
{ "pile_set_name": "ArXiv" }
--- author: - | Michael L. Mihalik and John G. Ratcliffe\ Mathematics Department, Vanderbilt University,\ Nashville TN 37240, USA title: '**On the Rank of a Coxeter Group**' --- Introduction ============ Let $W$ be a Coxeter group with Coxeter generators $S$. The [*rank*]{} of the Coxeter system $(W,S)$ is the cardinality $|S|$ of $S$. The Coxeter system $(W,S)$ has finite rank if and only if $W$ is finitely generated by Theorem 2(iii), Ch. IV, §1 of [@Bourbaki]. If $(W,S)$ has infinite rank, then $|S| = |W|$, since every element of $W$ is represented by a finite product of elements of $S$. Thus if $W$ is not finitely generated, the rank of $(W,S)$ is uniquely determined by $W$. If $W$ is finitely generated, then $W$ may have sets of Coxeter generators $S$ and $S'$ of different ranks. In this paper, we determine the set of all possible ranks for an arbitrary finitely generated Coxeter group $W$. This paper is a continuation of our previous paper with Steven Tschantz [@M-R-T] in which we studied the relationship between two sets $S$ and $S'$ of Coxeter generators of a finitely generated Coxeter group $W$. A [*basic subset*]{} of $S$ is a maximal subset $B$ of $S$ such that $B$ generates an irreducible, noncyclic, finite subgroup of $W$. In [@M-R-T], we proved the Basic Matching Theorem which says that there is a natural bijection (matching) between the basic subsets of $S$ and the basic subsets of $S'$. A basic subset $B$ of $S$ matches a basic subset $B'$ of $S'$ if and only if $[\langle B\rangle,\langle B\rangle]$ is conjugate to $[\langle B'\rangle,\langle B'\rangle]$ in $W$. Usually matching basic subsets generate isomorphic groups, in which case, we say that the [*basic subsets match isomorphically*]{}; however, there are exceptions, due to well known isomorphisms between irreducible and reducible finite Coxeter groups (for instance the dihedral group ${\bf D}_2(6)$ of order 12 and ${\bf A}_1\times {\bf A}_2$). We showed that nonisomorphic matching of basic subsets can be understood by [*blowing up*]{} Coxeter generating sets. This is a procedure to replace a given Coxeter generating set $S$ by a Coxeter generating set $R$ such that $|R| = |S|+1$. In [@M-R-T], we proved that there exists a set of Coxeter generators $S'$ of $W$ such that a basic subset $B$ of $S$ matches a basic subset $B'$ of $S'$ with $|\langle B\rangle| > |\langle B'\rangle|$ if and only if $S$ can be blown up. We proved that $S$ has maximum rank over all sets of Coxeter generators of $W$ if and only if $S$ can not be blown up. In this paper, we study the reverse procedure of [*blowing down*]{} Coxeter generating sets, which was introduced by Mihalik in [@Mihalik]. We first determine necessary and sufficient conditions on $(W,S)$ such that there exists a set of Coxeter generators $S'$ of $W$ such that a basic subset $B$ of $S$ matches a basic subset $B'$ of $S'$ with $|\langle B\rangle| < |\langle B'\rangle|$. We then determine necessary and sufficient conditions on $(W,S)$ such that $W$ has a set of Coxeter generators $S'$ such that $|S'| < |S|$. As an application, we determine the rank spectrum of $W$. Preliminaries ============= A [*Coxeter matrix*]{} is a symmetric matrix $M = (m(s,t))_{s,t\in S}$ with $m(s,t)$ either a positive integer or infinity and $m(s,t) =1$ if and only if $s=t$. A [*Coxeter system*]{} with Coxeter matrix $M = (m(s,t))_{s,t\in S}$ is a pair $(W,S)$ consisting of a group $W$ and a set of generators $S$ for $W$ such that $W$ has the presentation $$W =\langle S \ |\ (st)^{m(s,t)}:\, s,t \in S\ \hbox{and}\ m(s,t)<\infty\rangle$$ If $(W,S)$ is a Coxeter system with Coxeter matrix $M = (m(s,t))_{s,t\in S}$, then the order of $st$ is $m(s,t)$ for each $s,t$ in $S$, and so a Coxeter system determines its Coxeter matrix; moreover, any Coxeter matrix $M = (m(s,t))_{s,t\in S}$ determines a Coxeter system $(W,S)$ where $W$ is defined by the above presentation. If $(W,S)$ is a Coxeter system, then $W$ is called a [*Coxeter group*]{} and $S$ is called a set of [*Coxeter generators*]{} for $W$, and the cardinality of $S$ is called the [*rank*]{} of $(W,S)$. Let $(W,S)$ be a Coxeter system. The [*Coxeter diagram*]{} ([*[C]{}-diagram*]{}) of $(W,S)$ is the labeled undirected graph $\Delta(W,S)$ with vertices $S$ and edges $$\{(s,t) : s, t \in S\ \hbox{and}\ m(s,t) > 2\}$$ such that an edge $(s,t)$ is labeled by $m(s,t)$. A Coxeter system $(W,S)$ is said to be [*irreducible*]{} if its C-diagram is connected. A [*visible subgroup*]{} of $(W,S)$ is a subgroup of $W$ of the form $\langle A\rangle$ for some $A \subseteq S$. A visible subgroup $\langle A\rangle$ of $(W,S)$ is said to be [*irreducible*]{} if $(\langle A\rangle, A)$ is irreducible. A subset $A$ of $S$ is said to be [*irreducible*]{} if $\langle A\rangle$ is irreducible. A subset $A$ of $S$ is said to be a [*component*]{} of $S$ if $A$ is a maximal irreducible subset of $S$ or equivalently if $\Delta(\langle A\rangle, A)$ is a connected component of $\Delta(W,S)$. The [*presentation diagram*]{} ([*[P]{}-diagram*]{}) of $(W,S)$ is the labeled undirected graph $\Gamma(W,S)$ with vertices $S$ and edges $$\{(s,t) : s, t \in S\ \hbox{and}\ m(s,t) < \infty\}$$ such that an edge $(s,t)$ is labeled by $m(s,t)$. We continue with the terminology of [@M-R-T]. In particular, we use Coxeter’s notation on p. 297 of [@Coxeter] for the irreducible spherical simplex reflection groups except that we denote the dihedral group ${\bf D}_2^k$ by ${\bf D}_2(k)$. Subscripts denote the rank of a Coxeter system in Coxeter’s notation. Coxeter’s notation partly agrees with but differs from Bourbaki’s notation on p. 193 of [@Bourbaki]. Coxeter proved that every finite irreducible Coxeter system is isomorphic to exactly one of the Coxeter systems ${\bf A}_n$, $n\geq 1$, ${\bf B}_n$, $n\geq 4$, ${\bf C}_n$, $n\geq 2$, ${\bf D}_2(k)$, $k\geq 5$, ${\bf E}_6$, ${\bf E}_7$, ${\bf E}_8$, ${\bf F}_4$, ${\bf G}_3$, ${\bf G}_4$. See §3 of [@M-R-T] for definitions. For uniformity of notation, we define ${\bf B}_3 = {\bf A}_3$, ${\bf D}_2(3) = {\bf A}_2$ and ${\bf D}_2(4) = {\bf C}_2$. Let $(W,S)$ be a Coxeter system. A [*basic subset*]{} of $S$ is a maximal irreducible subset $B$ of $S$ such that $\langle B\rangle$ is a noncyclic finite group. If $B$ is a basic subset of $S$, we call $B$ a [*base*]{} of $(W,S)$ and $\langle B\rangle$ a [*basic subgroup*]{} of $W$. [(Basic Matching Theorem, Theorem 4.18 [@M-R-T])]{} Let $W$ be a finitely generated Coxeter group with two sets of Coxeter generators $S$ and $S'$. Let $B$ be a base of $(W,S)$. Then there is a unique irreducible subset $B'$ of $S'$ such that $[\langle B\rangle,\langle B\rangle]$ is conjugate to $[\langle B'\rangle,\langle B'\rangle]$ in $W$. Moreover, 1. the set $B'$ is a base of $(W,S')$, and we say that $B$ and $B'$ match, 2. if $|\langle B\rangle|=|\langle B'\rangle|$, then $B$ and $B'$ have the same type and there is an isomorphism $\phi:\langle B\rangle \to \langle B'\rangle$ that restricts to conjugation on $[\langle B\rangle,\langle B\rangle]$ by an element of $W$, and we say that $B$ and $B'$ match isomorphically, 3. if $|\langle B\rangle|<|\langle B'\rangle|$, then either $B$ has type ${\bf B}_{2q+1}$ and $B'$ has type ${\bf C}_{2q+1}$ for some $q\geq 1$ or $B$ has type ${\bf D}_2(2q+1)$ and $B'$ has type ${\bf D}_2(4q+2)$ for some $q\geq 1$. Moreover, there is a monomorphism $\phi:\langle B\rangle \to \langle B'\rangle$ that restricts to conjugation on $[\langle B\rangle,\langle B\rangle]$ by an element of $W$. Blowing Down Coxeter Systems ============================ Let $(W,S)$ be a Coxeter system of finite rank. In this section, we determine necessary and sufficient conditions for a base $B$ of $(W,S)$ to match a base $B'$ of $(W,S')$ with $|\langle B\rangle| < |\langle B'\rangle|$. If a base $B$ of $(W,S)$ matches a base $B'$ of $(W,S')$ with $|\langle B\rangle| < |\langle B'\rangle|$, then either $B$ is of type ${\bf B}_{2q+1}$ and $B'$ is of type ${\bf C}_{2q+1}$ for some $q\geq 1$ or $B$ is of type ${\bf D}_2(2q+1)$ and $B'$ is of type ${\bf D}_2(4q+2)$ for some $q\geq 1$ by the Basic Matching Theorem. If $a\in S$, the [*neighborhood*]{} of $a$ in P-diagram of $(W,S)$ is defined to be the set $N(a) =\{s\in S: m(s,a) < \infty\}$. If $A\subseteq S$, define $$A^\perp=\{s\in S: m(s,a) = 2\ \hbox{for all}\ a\in A\}.$$ The following lemma generalizes Proposition 2 of [@Mihalik]. \[First Lemma\] Let $B =\{x,y\}$ be a base of $(W,S)$ of type ${\bf D}_2(2q+1)$ that matches a base $B'$ of $(W,S')$ of type ${\bf D}_2(4q+2)$ for some $q\geq 1$. Then $N(x)\cap N(y) = B\cup B^\perp$. Suppose $s\in S-B$ with $m(s,x), m(s,y) <\infty$. Let $M\subseteq S$ be a maximal simplex containing $\{s,x,y\}$. Then there is a unique maximal simplex $M'\subseteq S'$ such that $\langle M\rangle$ is conjugate to $\langle M'\rangle$ by Prop. 4.21 of [@M-R-T]. By conjugating $S'$, we may assume that $\langle M\rangle=\langle M'\rangle$. Then $M'$ contains $B'$ and $[\langle B\rangle,\langle B\rangle]$ is conjugate to $[\langle B'\rangle,\langle B'\rangle]$ in $\langle M'\rangle$ by the Basic Matching Theorem. Let $B'=\{a,b\}$. Then $m(s',a)=m(s',b)=2$ for all $s'\in M'-B'$ by Theorem 8.7 of [@M-R-T]. Hence $B'$ is a component of $M'$. Therefore $[\langle B'\rangle,\langle B'\rangle]$ is a normal subgroup of $\langle M'\rangle$. Hence $[\langle B\rangle,\langle B\rangle]$ is a normal subgroup of $\langle M\rangle$. Therefore $s\{x,y\}s =\{x,y\}$ by Lemma 4.17 of [@M-R-T], and so $sxs=x$ and $sys=y$ by the deletion condition. Hence $N(x)\cap N(y) = B\cup B^\perp$. The C-diagram of ${\bf B}_n$, for $n\geq 5$, is a Y-shaped diagram with $n$ vertices $b_1,\ldots, b_n$ and two short arms. We call the endpoints $b_{n-1}$ and $b_n$ of the short arms the [*split ends*]{} of the C-diagram of ${\bf B}_n$. The [*split ends*]{} of ${\bf B}_3$ are the end points $b_2$ and $b_3$ of the C-diagram of ${\bf B}_3 = {\bf A}_3$. The C-diagram of ${\bf C}_n$ is a linear diagram with $n$ vertices $c_1,\ldots, c_n$ and $m(c_i,c_{i+1})=3$ for $i=1,\ldots, n-2$ and $m(c_{n-1},c_n)=4$. Note that $b_i=c_i$ for $i=1,\ldots, n-1$, and $b_n=c_nc_{n-1}c_n$, and $b_{n-1}b_n = (c_{n-1}c_n)^2$. See §3 of [@M-R-T] for details. \[Second Lemma\] Let $\phi:{\bf B}_n \to {\bf C}_n$ be a monomorphism with $n$ odd and $n\geq 3$. Then $\phi$ maps $b_{n-1}b_n$ to a conjugate of $(c_{n-1}c_n)^2$ in ${\bf C}_n$. Now $\phi({\bf B}_n)$ does not contain the center of ${\bf C}_n$, since $Z({\bf B}_n) = \{1\}$. Therefore either $\phi({\bf B}_n)={\bf B}_n$ or $\phi({\bf B}_n)=\theta({\bf B}_n)$ where $\theta$ is the automorphism of ${\bf C}_n$ defined by $\theta(c_i) = -c_i$, for $i=1,\ldots,n-1$ and $\theta(c_n)=c_n$. Now $\theta$ restricts to the identity on $[{\bf C}_n,{\bf C}_n]$, and so by composing $\phi$ with $\theta$ in the latter case, we may assume that $\phi({\bf B}_n)={\bf B}_n$. Every automorphism of ${\bf B}_n$ is inner by Theorem 31 of [@F-H]. Hence $\phi$ restricts to conjugation on $[{\bf B}_n,{\bf B}_n]$ by an element of ${\bf B}_n$. As $b_{n-1}b_n$ is in $[{\bf B}_n,{\bf B}_n]$, and $b_{n-1}b_n=(c_{n-1}c_n)^2$, we conclude that $\phi(b_{n-1}b_n)$ is conjugate to $(c_{n-1}c_n)^2$ in ${\bf C}_n$. Let $W$ be a finitely generated Coxeter group with two sets $S$ and $S'$ of Coxeter generators, and let $A$ be a subset of $S$. Let $\ov A$ be the intersection of all subsets $B$ of $S$ such that $B$ contains $A$ and $\langle B\rangle$ is conjugate to $\langle B'\rangle$ for some $B'\subseteq S'$. Then $\ov A$ is the smallest subset $B$ of $S$ such that $B$ contains $A$ and $\langle B\rangle$ is conjugate to $\langle B'\rangle$ for some $B'\subseteq S'$ by Prop. 4.14 of [@M-R-T]. If $A$ is a spherical simplex, then $\ov A$ is a spherical simplex, since for any maximal spherical simplex $M\subseteq S$ that contains $A$, there exists a unique maximal spherical simplex $M'\subseteq S'$ such that $\langle M\rangle$ is conjugate to $\langle M'\rangle$ by Prop. 4.13 of [@M-R-T]. \[Third Lemma\] Let $B$ be a base of $(W,S)$ of type ${\bf B}_{2q+1}$ that matches a base $B'$ of $(W,S')$ of type ${\bf C}_{2q+1}$ for some $q\geq 1$. Let $x,y$ be the split ends of the C-diagram of $(\langle B\rangle, B)$. Then $\ov{\{x,y\}} =\ov B$ and $N(x)\cap N(y) = B\cup B^\perp$. Let $C = \ov{\{x,y\}}$. Then $C$ is a spherical simplex of $(W,S)$ and $\langle C\rangle$ is conjugate to $\langle C'\rangle$ for some $C'\subseteq S'$. By conjugating $S'$, we may assume that $\langle C\rangle = \langle C'\rangle$. Let $a, b, c$ be the elements of $B'$ such that $m(a,b)=4$ and $m(b,c)=3$. Now $xy\in [\langle B\rangle,\langle B\rangle]$, and so $xy$ is conjugate to $(ab)^2$ by the Basic Matching Theorem and Lemma \[Second Lemma\]. Hence there is a $w\in W$ such that $w(ab)^2w^{-1}\in \langle C'\rangle$. Now $\langle (ab)^2\rangle =[\langle a,b\rangle,\langle a,b\rangle]$. Let $u$ be the shortest element of $\langle C'\rangle w\langle a,b\rangle$. Then $u\{a,b\}u^{-1}\subseteq C'$ by Lemma 4.17 of [@M-R-T]. As $m(a,b)=4$, we deduce that $\{a,b\}\subseteq C'$ by Lemma 4.9 of [@M-R-T]. Hence $B'\subseteq C'$ by Lemma 8.1 of [@M-R-T]. By the Basic Matching Theorem, $B\subseteq C$. Hence $\ov B\subseteq C$. As $\{x,y\}\subset B$, we have $C\subseteq \ov B$. Thus $C = \ov B$. Suppose $s\in S-B$ with $m(s,x), m(s,y)<\infty$. Let $M\subseteq S$ be a maximal simplex containing $\{s,x,y\}$. Then there is a maximal simplex $M'\subseteq S'$ such that $\langle M\rangle$ is conjugate to $\langle M'\rangle$ by Prop. 4.21 of [@M-R-T]. By conjugating $S'$, we may assume that $\langle M\rangle=\langle M'\rangle$. Now $\ov{\{x,y\}}\subseteq M$, and so $B\subseteq M$. Hence $B'\subseteq M'$ by the Basic Matching Theorem. Moreover $m(s',t') = 2$ for all $(s',t')\in (M'-B')\times B'$ by Theorem 8.2 of [@M-R-T]. Hence $B'$ is a component of $M'$. Therefore $[\langle B'\rangle,\langle B'\rangle]$ is a normal subgroup of $\langle M'\rangle$. By the Basic Matching Theorem, $[\langle B\rangle,\langle B\rangle]$ is conjugate to $[\langle B'\rangle,\langle B'\rangle]$ in $\langle M\rangle$. Therefore $[\langle B\rangle,\langle B\rangle]$ is a normal subgroup of $\langle M\rangle$. Then $sBs = B$ by Lemma 4.17 of [@M-R-T], and so $sts=t$ for all $t\in B$ by the deletion condition. Hence $N(x)\cap N(y) = B\cup B^\perp$. The [*odd diagram*]{} of $W$ is the labeled undirected diagram $\Omega(W,S)$ obtained from the P-diagram of $(W,S)$ be deleting the even labeled edges. If $a\in S$, we define ${\rm Odd}(a)$ to be the vertex set of the connected component of $\Omega(W,S)$ containing $a$. By Prop. 3, Ch. IV, §1 of Bourbaki [@Bourbaki], we have that $${\rm Odd}(a) = \{s\in S: s\ \hbox{is conjugate to}\ a\ \hbox{in}\ W\}.$$ We define $${\rm EOdd}(a) = {\rm Odd}(a)\cup \{s\in S: m(s,b)\ \hbox{is even for some}\ b\in {\rm Odd}(a)\}.$$ The next lemma has its genesis in Proposition 3 of [@Mihalik]. \[Fourth Lemma\] Let $B$ be a base of $(W,S)$ that matches a base $B'$ of $(W,S')$ with $|\langle B\rangle| < |\langle B'\rangle|$. Then there exists $r\in \ov{B} - B$ such that $N(r) =B\cup B^\perp$ and ${\rm Odd}(r) = \{r\}$, and if $K$ is the component of $B^\perp$ containing $r$, then $K$ is of type ${\bf A}_1$, ${\bf C}_{2q+1}$, or ${\bf D}_2(4q+2)$ for some $q\geq 1$; moreover, if $K\neq\{r\}$, then $K$ is a basic subset of $S$ and if $K'$ is the basic subset of $S'$ that matches $K$, then $K'$ is a component of $(B')^\perp$ and $K'\cup (K')^\perp = B'\cup(B')^\perp$. Let $C=\ov B$. Then $C$ is a spherical simplex of $(W,S)$ and $\langle C\rangle$ is conjugate to $\langle C'\rangle$ for some $C'\subseteq S'$. By conjugating $S'$, we may assume that $\langle C\rangle = \langle C'\rangle$. Then $C'$ contains $B'$ by the Basic Matching Theorem. Hence $B$ is a proper subset of $C$, since otherwise $\langle B'\rangle \subseteq \langle C'\rangle =\langle C\rangle = \langle B\rangle$ which is not the case, since $|\langle B\rangle|< |\langle B'\rangle|$. Let $M\subseteq S$ be a maximal spherical simplex containing $B$, and let $M'\subseteq S'$ be the maximal spherical simplex such that $\langle M\rangle$ is conjugate to $\langle M'\rangle$. Then $M'$ contains $B'$ by the Basic Matching Theorem. Let $w$ be an element of $W$ such that $w\langle M\rangle w^{-1} = \langle M'\rangle$. Now $B'\subseteq \langle C\rangle \subseteq \langle M\rangle$. Hence $wB'w^{-1} \subset \langle M'\rangle$. Let $u$ be the shortest element of $\langle M'\rangle w\langle B'\rangle$. Then $uB'u^{-1}\subseteq M'$ by Lemma 4.3 of [@M-R-T]. As $B'$ is a base of $(W,S')$, we have that $uB'u^{-1} = B'$ by Lemma 4.10 of [@M-R-T], and so $u$ acts as a graph automorphism on $\langle B'\rangle$. Let $z'$ be the longest element of $\langle B'\rangle$. Then $uz'u^{-1} = z'$. Now $w = xuy$ with $x\in \langle M'\rangle$ and $y\in \langle B'\rangle$. Hence $wz'w^{-1} = z'$, since $z'$ is in the center of $\langle M'\rangle$. Therefore $z'$ is in the center of $\langle M\rangle = w^{-1}\langle M'\rangle w$. As $B\cup B^\perp$ is the union of all the maximal spherical simplices of $(W,S)$ that contain $B$, we deduce the $z'$ is in the center of $B^\perp$. Hence, there are distinct components $K_1,\ldots, K_n$ of $B^\perp$, with nontrivial center, such that $z' = z_1\cdots z_n$ with $z_i$ the longest element of $\langle K_i\rangle$ for each $i=1,\ldots,n$. As $z'\in \langle C\rangle$, we have that $K_i\subseteq C$ for each $i=1,\ldots, n$ by Prop. 7, Ch. IV, §1 of [@Bourbaki], since every reduced form of $z_i$ involves every element of $K_i$ for each $i=1,\ldots,n$. Define a homomorphism $\rho:\langle C'\rangle \to \langle z'\rangle$ as follows. To begin with, define $\rho(s') = 1$ if $s'\in C'-B'$. By the Basic Matching Theorem, $B'$ is of type ${\bf C}_{2p+1}$ or ${\bf D}_2(4p+2)$ for some $p\geq 1$. If $B'$ is of type ${\bf C}_{2p+1}$, let $a'\in B'$ be such that $B'\cap {\rm Odd}(a') =\{a'\}$, and define $\rho(a') = z'$ and $\rho(s') = 1$ for each $s'\in B'-\{a'\}$. Suppose $B'$ is of type ${\bf D}_2(4p+2)$ and $B' = \{a',b'\}$. By Lemma 8.6 of [@M-R-T], one of $a'$ or $b'$, say $a'$, has the property that if $a'\in A'\subseteq S'$ and $\langle A'\rangle$ is conjugate to $\langle A\rangle$ for some $A\subseteq S$, then $B'\subseteq A'$. Define $\rho(a') = z'$ and $\rho(b') =1$. In both cases, $\rho$ is well defined and $\rho(z') = z'$. As $z'=z_1\cdots z_n$, there is an $i$ such that $\rho(z_i) = z'$. By reindexing, we may assume $i=1$. Let $K = K_1$. Then there exists $r\in K$ such that $\rho(r) = z'$. As $r\in \ov B$, we have that $\ov{\{r\}}\subseteq \ov B$. Let $A= \ov{\{r\}}$. Then $\langle A\rangle$ is conjugate in $\langle C'\rangle$ to $\langle A'\rangle$ for some $A' \subseteq C'$ by Prop. 4.14 of [@M-R-T]. Now $\rho(\langle A'\rangle) = \rho(\langle A\rangle) = \langle z'\rangle$. Hence $a' \in A'$. Then $B'\subseteq A'$ by Lemma 8.1 of [@M-R-T] or the choice of $a'$. Hence $B \subseteq A$ by the Basic Matching Theorem, and so $\ov B\subseteq A$. Hence $\ov{\{r\}}= \ov B$. As $K$ is a component of $B^\perp$, we have $B\cup B^\perp \subseteq N(r)$. Suppose $s\in N(r)$. Let $M\subseteq S$ be a maximal spherical simplex containing $\{r,s\}$. Then there is a maximal spherical simplex $M'\subseteq S'$ such that $\langle M\rangle$ is conjugate to $\langle M'\rangle$. Then $\ov{\{r\}}\subseteq M$, and so $B\subseteq M$. Therefore $s\in B\cup B^\perp$, since $B$ is a basic subset of $M$. Thus $N(r) = B\cup B^\perp$. As $[\langle K\rangle, \langle K\rangle]$ is in the kernel of $\rho$, we have that $z_1$ is not in $[\langle K\rangle, \langle K\rangle]$. Therefore $K$ is of type ${\bf A}_1$, ${\bf C}_{2q+1}$, ${\bf D}_2(4q+2)$, ${\bf E}_7$, or ${\bf G}_3$ for some $q\geq 1$. Suppose $K$ is of type ${\bf C}_{2q+1}$ for some $q\geq 1$. Let $a\in K$ be such that $K\cap{\rm Odd}(a) = \{a\}$. Then $a[\langle K\rangle,\langle K\rangle] = z_1[\langle K\rangle,\langle K\rangle]$. As the restriction of $\rho$ to $\langle K\rangle$ factors through $\langle K\rangle/[\langle K\rangle,\langle K\rangle]$, we may assume that $r=a$. Then ${\rm Odd}(r) = \{r\}$. If $K= \{r\}$, then we are done. Suppose $K\neq \{r\}$. Then $K$ is a basic subset of $S$, since $N(r) = B\cup B^\perp$ and $K$ is a component of $B^\perp$. Let $K'$ be the basic subset of $C'$ that matches $K$. Then $K'$ is the basic subset of $S'$ that matches $K$ by the Basic Matching Theorem. Let $M'\subseteq S'$ be a maximal spherical simplex that contains $B'$, and let $M\subseteq S$ be the maximal spherical simplex such that $\langle M\rangle$ is conjugate to $\langle M'\rangle$. Then $M$ contains $B$ by the Basic Matching Theorem. Now $K\subseteq C\subseteq M$ and $K$ is a basic subset of $M$. Therefore $K'$ is a basic subset of $M'$ by the Basic Matching Theorem. Hence $K'$ is a component of $M'$. Therefore $K'$ is a component of $(B')^\perp$. Suppose $s'\in (K')^\perp$. Let $M'\subseteq S'$ be a maximal spherical simplex that contains $K'\cup\{s'\}$, and let $M\subseteq S$ be the maximal spherical simplex such that $\langle M\rangle$ is conjugate to $\langle M'\rangle$. Then $K\subseteq M$ by the Basic Matching Theorem. Hence $M$ contains $r$, and so $B\subseteq \ov{\{r\}}\subseteq M$. Therefore $M'$ contains $B'$ by the Basic Matching Theorem. Hence $s'\in B'\cup (B')^\perp$. Therefore $(K')\cup (K')^\perp = B'\cup(B')^\perp$. If $K$ is of type ${\bf C}_{2q+1}$ or ${\bf D}_2(4q+2)$, we are done. Suppose $K$ is of type ${\bf E}_7$ or ${\bf G}_3$. Then $\rho(s) = z'$ for each $s$ in $K$, since $s$ and $r$ are conjugate in $\langle K\rangle$ for each $s\in K$. Hence $\ov{\{s\}} = \ov B$ and $N(s) =B\cup B^\perp$ for each $s\in K$ by the above argument. Therefore ${\rm Odd}(r) = K$ and ${\rm EOdd}(r) = B\cup B^\perp$. Suppose $n > 1$. By Lemma 28 of [@H-M], there is an automorphism $\theta$ of $W$ such that $\theta(s) = s$ for all $s\in S-K$, and $\theta(s) = sz_2\cdots z_n$ for all $s\in K$. The longest element of $\langle\theta(K)\rangle$ is $z_1\cdots z_n=z'$. Now replace $S$ by $\theta(S)$. Let $\ov{\theta(B)}$ denote $\ov B$ with respect to the Coxeter generators $\theta(S)$ and $S'$. Then $\ov{\theta(B)}\subseteq \theta(C)$, since $\langle \theta(C)\rangle = \langle C\rangle =\langle C'\rangle$. If $\ov{\theta(B)}$ is a proper subset of $\theta(C)$, we return to the start of the proof. As $C$ is finite, we will eventually be done or have $\ov{\theta(B)}=\theta(C)$. Thus we may assume without loss of generality that $n = 1$ and $z' = z_1$. Let $\ell'$ be the longest element of $\langle K'\rangle$. Define a homomorphism $\eta:\langle C'\rangle \to \langle \ell'\rangle$ as follows. Define $\eta(s') = 1$ for all $s'\in C'-K'$ and define $\eta(s') = \ell'$ for all $s' \in K'$. Then $\eta$ is well defined and $\eta(\ell') = \ell'$. By the same argument as above, $\ell'$ is in the center of $B^\perp$ and there is a component $L$ of $B^\perp$ such that $L \subseteq C$ and $L$ has nontrivial center, and if $\ell$ is the longest element of $\langle L\rangle$, then $\eta(\ell) = \ell'$ and $\eta(t) = \ell'$ for some $t\in L$. As $z'\in B'\subseteq C'-K'$, we have that $\eta(z') = 1$, and so $L\neq K$. Let $A = \ov{\{t\}}$. Then $\langle A\rangle$ is conjugate in $\langle C'\rangle$ to $\langle A'\rangle$ for some $A' \subseteq C'$ by Prop. 4.14 of [@M-R-T]. Now $\eta(\langle A'\rangle) = \eta(\langle A\rangle) = \langle \ell'\rangle$. Hence $s'\in A'$ for some $s'\in K'$. Now killing $s'$ in $\langle C'\rangle$ kills $K'$. As $[\langle K\rangle, \langle K\rangle] = [\langle K'\rangle,\langle K'\rangle]$ and $\langle A\rangle$ is conjugate to $\langle A'\rangle$ in $\langle C\rangle$, the group $\langle A\rangle$ contains an element that kills $[\langle K\rangle, \langle K\rangle]$ in $\langle C\rangle$. Therefore $A$ contains an element $s\in K$. Now $\ov{\{s\}} \subseteq A$. As $\ov{\{s\}} = \ov B$, we have $B\subseteq A$. Therefore $\ov{\{t\}} = \ov B$. As before, $N(t) = B\cup B^\perp$. By the same argument as above, $L$ is of type ${\bf A}_1$, ${\bf C}_{2q+1}$, ${\bf D}_2(4q+2)$, ${\bf E}_7$, or ${\bf G}_3$, and if $L$ is of type ${\bf A}_1$, ${\bf C}_{2q+1}$, or ${\bf D}_2(4q+2)$, we are done. Suppose $L$ is of type ${\bf E}_7$ or ${\bf G}_3$. Then by the same argument as above, ${\rm Odd}(t) = L$ and ${\rm EOdd}(t) = B\cup B^\perp$, and we may assume that $\ell = \ell'$. By Lemma 38 of [@H-M], there is an automorphism $\beta$ of $W$ such that $\beta(s) = s$ for all $s\in S-(K\cup L)$, and $\beta(s) = s\ell z'$ for each $s \in K\cup L$. Then $\beta(\ell) = z'$ and $\beta(z') = \ell$. As $\beta$ fixes each element of $S-(K\cup L)$ and $\beta$ leaves $\langle C\rangle$ invariant, we may replace $S$ by $\beta(S)$. Then $K$ is replaced by $\beta(L)$ and $\beta(K)$ is removed as a possibility for replacing $L$, since $\langle \beta(K)\rangle = \langle K'\rangle$. In the above procedure only Coxeter generators of components of $B^\perp$ of type ${\bf E}_7$ or ${\bf G}_3$ are replaced. By repeating this procedure a finite number of times, we can remove the possibility that $L$ is of type ${\bf E}_7$ or ${\bf G}_3$, and we are done. A [*cycle*]{} of $S$ is a sequence $\{c_1,\ldots, c_n\}$ of distinct elements of $S$ so that $m(c_i,c_{i+1}) <\infty$ for $i=1,\ldots,n-1$ and $m(c_n,c_1)<\infty$. A [*chord*]{} of cycle $C =\{c_1,\ldots, c_n\}$ of $S$ is a pair of distinct elements $c_i,c_j$ of $C$ such that $m(c_i,c_j) < \infty$ and $c_i, c_j$ are neither consecutive terms of the cycle nor the end terms of the cycle. The next lemma generalizes Proposition 4 of [@Mihalik]. \[Fifth Lemma\] Let $B$ be a base of $(W,S)$ that matches a base $B'$ of $(W,S')$ with $|\langle B\rangle|<|\langle B'\rangle|$. If $B$ is of type ${\bf D}_2(2q+1)$, let $B =\{x,y\}$. If $B$ is of type ${\bf B}_{2q+1}$, let $\{x,y\}$ be the set of split ends of the C-diagram of $(\langle B\rangle, B)$. Then $\{x,y\}$ is not part of a chord-free cycle of $S$ of length at least 4. On the contrary, suppose $C\subseteq S$ is a chord-free cycle of length at least 4 that contains $\{x,y\}$. We may assume that $|S|$ is as small as possible. By Lemmas \[First Lemma\] and \[Third Lemma\], we have that $\ov{\{x,y\}} = \ov B$ and $N(x)\cap N(y) = B\cup B^\perp$. By Lemma \[Fourth Lemma\], there is an $r\in \ov B-B$ such that $N(r)=B\cup B^\perp$ and ${\rm Odd}(r)=\{r\}$. Now $C\cap (B\cup B^\perp) = \{x,y\}$, since $C$ is cord-free of length at least 4. Let $a \in C-B$. Then $(B\cup B^\perp)-\{r\}$ is an $(a,r)$-separator of $S$, that is, every path from $a$ to $r$ in the P-diagram of $(W,S)$ passes through $(B\cup B^\perp)-\{r\}$. Let $S_0$ be a c-minimal separator (see §6 of [@M-R-T]) of $S$ such that $S_0$ is conjugate to a subset of $(B\cup B^\perp)-\{r\}$. By Lemma 4.9 of [@M-R-T], we have that $S_0=S_1\cup S_2$ with $S_1$ a spherical simplex, $S_2\subseteq S_1^\perp$, and $wS_2w^{-1} \subseteq S$ if and only if $w=1$. Then $S_2\subseteq (B\cup B^\perp)-\{r\}$. By Theorem 6.1 of [@M-R-T], there exists $S_0'\subseteq S'$, a reduced visual graph of groups decomposition $\Lambda$ for $(W,S)$, and a reduced visual graph of groups decomposition $\Lambda'$ for $(W,S')$ such that $\langle S_0\rangle$ is conjugate to $\langle S_0'\rangle$, and the edge groups of $\Lambda$ and $\Lambda'$ are conjugate to $\langle S_0\rangle$, and there is a 1-1 correspondence between the vertices of $\Lambda$ and the vertices of $\Lambda'$ such that each vertex group of $\Lambda$ is conjugate to the corresponding vertex group of $\Lambda'$. Now $r$ is not in an edge group of $\Lambda$, since $r$ is not conjugate to an element of $(B\cup B^\perp)-\{r\}$. Let $V$ be the vertex group of $\Lambda$ that contains $r$. Then $N(r)\subset V$, and so $B\cup B^\perp \subset V$. Now $\{x,y\}$ is not contained in an edge group $E$ of $\Lambda$, since otherwise $r\in \ov{\{x,y\}} \subset E$. We claim that $C\subset V$. On the contrary, suppose $C\not\subset V$. Let $C=\{c_1,\ldots,c_n\}$ with $x = c_1$ and $c_n=y$, and $m(c_i,c_{i+1})<\infty$ for each $i=1,\ldots,n-1$. Let $k$ be the first index such that $c_k\not\in V$ and let $\ell$ be the last index such that $c_\ell\not\in V$. Then $c_{k-1}$ is in an edge group $E$ of $\Lambda$ that is a subgroup of $V$ and $c_{\ell+1}$ is in an edge group $F$ of $\Lambda$ that is a subgroup of $V$. Now $E=F$, since the graph of $\Lambda$ is a tree. As $\{x,y\}\not\subset E$, we have that $\{c_{k-1},c_{\ell+1}\}\neq\{x,y\}$. Now $E$ is conjugate to $\langle S_0\rangle$, and so there exists $S_3\subseteq S_2^\perp$ such that $E = \langle S_3\cup S_2\rangle$ and $S_3$ is conjugate to $S_1$. Hence $S_3$ is a spherical simplex. Now $(C-B)\cap (B\cup B^\perp) = \emptyset$, since $C$ is chord-free. As $S_2\subseteq B\cup B^\perp$, we deduce that $\{c_{k-1},c_{\ell+1}\} \cap S_3 \neq \emptyset$, and so $c_{k-1}$ and $c_{\ell+1}$ are joined by a chord, which is a contradiction. Thus $C\subset V$. By conjugating $S'$, we may assume that $V$ is a vertex group $V'$ of $\Lambda'$. Then $B'\subset V'$ by the Basic Matching Theorem. Now ${\rm rank}(V) < |S|$, and so we have a contradiction to the minimality of $|S|$. [(Blow-Down Lemma)]{}\[Sixth Lemma\] Let $(W,S)$ be a Coxeter system of finite rank, and let $B$ be a base of $(W,S)$ of type ${\bf B}_{2q+1}$ or ${\bf D}_2(2q+1)$ for some $q\geq 1$. If $|B| = 2$, let $B =\{x,y\}$. If $B$ is of type ${\bf B}_{2q+1}$, let $\{x,y\}$ be the set of split ends of the C-diagram of $(\langle B\rangle, B)$. Let $r\in B^\perp$ such that $N(r) = B\cup B^\perp$ and $\{r\}$ is a component of $B^\perp$. Suppose $N(y) = B\cup B^\perp$. Let $\ell$ be the longest element of $\langle B\rangle$, let $a=r\ell$, let $S' = (S-\{r,y\})\cup \{a\}$, and let $B'=(B-\{y\})\cup \{a\}$. Then $S'$ is a set of Coxeter generators for $W$ such that 1. the set $B'$ is a base of $(W,S')$ that matches $B$ with $|\langle B\rangle|<|\langle B'\rangle|$, 2. $(B')^\perp = B^\perp-\{r\}$, 3. the neighborhood of $a$ satisfies $N(a) = B'\cup (B')^\perp$, 4. the basic subsets of $S$ and $S'$ are the same except for $B$ and $B'$. Consider the Coxeter presentation $$W = \langle S\ |\ (st)^{m(s,t)}: s,t \in S\ \hbox{and}\ m(s,t)<\infty\rangle.$$ Let $\ell$ be the longest element of $\langle B\rangle$. Then $\ell^2=1$. Regard $\ell$ as a reduced word in the elements of $B$. Add the generator $a$ and the relation $a=r\ell$ to the above presentation of $W$. Then we can add the relators $a^2$ and $(sa)^2$ and $(as)^2$ for all $s\in B^\perp-\{r\}$. As $\ell s\ell = s$ for all $s\in B-\{x,y\}$, we can add the relators $(sa)^2$ and $(as)^2$ for all $s\in B-\{x,y\}$. Next delete the generator $r$ and the relation $a = r\ell$ and replace $r$ by $a\ell$ in the remaining relators. The relator $r^2$ is replaced by $(a\ell)^2$. We delete the relators $(sa\ell)^2$ and $(a\ell s)^2$ for $s\in B\cup B^\perp-\{r,x,y\}$, since they are equivalent to $(a\ell)^2$. As $\ell x\ell = y$ and $\ell a \ell = a$, we have $xa\ell xa\ell = xa\ell x\ell a\ell\ell = xaya$, and so we can replace $(xa\ell)^2$ by $xaya$. Likewise $(a\ell x)^2$ can be replaced by $ayax$, and $(ya\ell)^2$ can be replaced by $yaxa$, and $(a\ell y)^2$ can be replaced by $axay$. Next we delete the generator $y$ and the relators $xaya$, $ayax$, $ayax$, $axay$, and replace $y$ by $axa$ in the remaining relators. We have eliminated all the relators originally involving $r$ except for $(a\ell)^2$. We delete the relators $(saxa)^2$ and $(axas)^2$ for all $s\in B\cup B^\perp-\{r,x,y\}$, since they are equivalent to $(axa)^2 = ax^2a$. The relator $y^2$ is replaced by $(axa)^2$ which we can delete. Assume first that $B$ is of type ${\bf B}_{2q+1}$. The relators $(xy)^2$ and $(yx)^2$ are replaced by $(xa)^4$ and $(ax)^4$. Let $t\in B$ be such that $m(t,y) = 3$. Then $m(t,x)=3$. Now $(ty)^3 = (taxa)^3 = (atxa)^3 = a(tx)^3a$, and so $(ty)^3$ can be deleted. Likewise $(yt)^3$ can be deleted. The relator $(a\ell)^2$ can be deleted, since it is redundant. Then we obtain a Coxeter presentation for $W$ with Coxeter generators $S'$ and $(\langle B'\rangle, B')$ of type ${\bf C}_{2q+1}$. Now assume that $B$ is of type ${\bf D}_2(2q+1)$. The relators $(xy)^{2q+1}$ and $(yx)^{2q+1}$ are replaced by $(xa)^{4q+2}$ and $(ax)^{4q+2}$. The relator $(a\ell)^2$ can be deleted, since it is redundant. Then we obtain a Coxeter presentation for $W$ with Coxeter generators $S'$ and $(\langle B'\rangle, B')$ of type ${\bf D}_2(4q+2)$. As $a = r\ell$ and $\{r\}$ is a component of $B^\perp$, we have that $B^\perp-\{r\} \subseteq (B')^\perp$. Suppose $s\in (B')^\perp$. Then $s\in S-(B\cup\{r\})$. As $s$ commutes with $a=r\ell$, we have that $s\in N(r) = B\cup B^\perp$ by Lemma 8.3 of [@M-R-T]. Hence $s\in B^\perp-\{r\}$. Therefore $(B')^\perp = B^\perp-\{r\}$. Clearly, we have $B'\cup (B')^\perp \subseteq N(a)$. Suppose $s \in N(a)-B'$. Then $s\in N(r) = B\cup B^\perp$ by Lemma 8.3 of [@M-R-T]. Hence $s \in B^\perp-\{r\} = (B')^\perp$, and so $N(a) = B'\cup (B')^\perp$. Therefore $B'$ is a base of $(W,S')$ and $B'$ is the only base of $(W,S')$ that contains $a$. The base $B'$ matches $B$, since $[\langle B'\rangle, \langle B'\rangle] = [\langle B\rangle,\langle B\rangle]$. As $N(y) = B\cup B^\perp$, we have that $B$ is the only base of $(W,S)$ that contains $y$. Therefore the basic subsets of $S$ and $S'$ are the same except for $B$ and $B'$. Let $B$ and $r\in B^\perp$ be as in the Blow-Down Lemma. We call $r$ a [*sink*]{} for the base $B$. Let $B'$ and $S'$ be as in the Blow-Down Lemma. We say that $(W,S')$ is obtained by [*blowing down*]{} $(W,S)$ along the base $B$. We also say that $B'$ has been obtained by [*blowing down*]{} $B$. The next theorem has its genesis in Proposition 5 of [@Mihalik]. \[Blow-Down Theorem\] [(Blow-Down Theorem)]{} Let $(W,S)$ be a Coxeter system of finite rank, and let $B$ be a base of $(W,S)$ of type ${\bf B}_{2p+1}$ or ${\bf D}_2(2p+1)$ for some $p\geq 1$. If $|B| = 2$, let $B =\{x,y\}$. If $B$ is of type ${\bf B}_{2p+1}$, let $\{x,y\}$ be the set of split ends of the C-diagram of $(\langle B\rangle, B)$. Then $W$ has a set of Coxeter generators $S'$ such that $B$ matches a base $B'$ of $(W,S')$ with $|\langle B\rangle|<|\langle B'\rangle|$ if and only if 1. the neighborhoods of $x$ and $y$ satisfy $N(x)\cap N(y) = B\cup B^\perp$, 2. the set $\{x,y\}$ is not part of a chord-free cycle of $S$ of length at least 4, 3. there exists $r\in B^\perp$ such that $N(r) = B\cup B^\perp$ and ${\rm Odd}(r) = \{r\}$, and if $K$ is the component of $B^\perp$ containing $r$, then $K$ is of type ${\bf A}_1$, ${\bf C}_{2q+1}$, or ${\bf D}_2(4q+2)$ for some $q\geq 1$. Suppose $W$ has a set of Coxeter generators $S'$ such that $B$ matches a base $B'$ of $(W,S')$ with $|\langle B\rangle|<|\langle B'\rangle|$. Then condition (1) follows from Lemmas \[First Lemma\] and \[Third Lemma\], condition (2) follows from Lemma \[Fifth Lemma\], and condition (3) follows from Lemma \[Fourth Lemma\]. Conversely, suppose conditions (1), (2), and (3) are satisfied. Let $S_0 = B\cup B^\perp$, and let $T = S-S_0$. Let $T_x$ be the set of all $t\in T$ such that there is a sequence $t_1,\ldots, t_n$ in $T$ such that $m(x,t_1) < \infty$, $m(t_i,t_{i+1})<\infty$ for each $i=1,\ldots, n-1$, and $t_n=t$. Define $T_y$ similarly. We claim that $T_x\cap T_y = \emptyset$. On the contrary suppose that $T_x\cap T_y \neq \emptyset$. Then there is a cycle $C$ of $S$ such that $C\cap S_0 = \{x,y\}$. Assume that $C$ is as short as possible. Then $C$ is chord-free. By condition (1), we deduce that $C$ has length at least 4, but this contradicts condition (2). Therefore $T_x\cap T_y = \emptyset$. Let $S_1 = S-T_y$ and $S_2 = S_0\cup T_y$. Then $S=S_1\cup S_2$ and $S_1\cap S_2 = S_0$, and $m(a,b) = \infty$ for all $a\in S_1-S_0$ and $b\in S_2-S_0$. Let $\ell$ be the longest element of $\langle B\rangle$. Then $\ell S_0\ell^{-1} = S_0$ and the triple $(S_1,\ell, S_2)$ determines an elementary twist (see §5 of [@M-R-T]) of $(W,S)$ giving a new Coxeter generating set $S_\ast = S_1\cup \ell S_2 \ell^{-1}$. As $\ell y\ell^{-1} = x$, we have $\ell T_y\ell^{-1}\subseteq T_x$ with respect to $S_\ast$, and so by replacing $S$ by $S_\ast$, we may assume $T_y = \emptyset$. Then $N(y) = B\cup B^\perp$. If $K = \{r\}$, then we can blow down $B$. Hence $W$ has a set of Coxeter generators $S'$ such that $B$ matches a base $B'$ of $(W,S')$ with $|\langle B\rangle|<|\langle B'\rangle|$ by Lemma \[Sixth Lemma\]. If $K\neq \{r\}$, we can blow up $S$ along $K$ by Theorems 8.4 and 8.8 of [@M-R-T]. This creates a sink for $B$, which allows us to blow down $B$. Therefore $W$ has a set of Coxeter generators $S'$ such that $B$ matches a base $B'$ of $(W,S')$ with $|\langle B\rangle|<|\langle B'\rangle|$ by Lemma \[Sixth Lemma\]. The proof of the Blown-Down Theorem indicates that we may have to blow up along one base in order to create a sink before we can blow down along another base. For example, the base ${\bf D}_2(3)$ of the Coxeter system ${\bf C}_3\times {\bf D}_2(3)$ can be blown down only after the system is blown up along the base ${\bf C}_3$ to yield the system ${\bf B}_3\times {\bf A}_1\times{\bf D}_2(3)$. Then ${\bf A}_1$ is a sink for the base ${\bf D}_2(3)$, and so now we can blow down ${\bf D}_2(3)$ to obtain the system ${\bf B}_3\times{\bf D}_2(6)$. If we blow up a Coxeter system and then blow down the resulting Coxeter system, the initial and final systems have the same rank. For example, the initial system ${\bf C}_3\times {\bf D}_2(3)$ has the same rank as the final system ${\bf B}_3\times{\bf D}_2(6)$. Contracting Coxeter Systems =========================== In this section, we determine necessary and sufficient conditions on $(W,S)$ such that $W$ has a set of Coxeter generators $S'$ such that $|S'|< |S|$. [(Contracting Theorem)]{} Let $(W,S)$ be a Coxeter system of finite rank. Then $W$ has a set of Coxeter generators $S'$ such that $|S'| < |S|$ if and only if there is a base $B$ of $(W,S)$ of type ${\bf B}_{2p+1}$ or ${\bf D}_2(2p+1)$ for some $p\geq 1$ satisfying conditions (1), (2), (3) of the Blow-Down Theorem with $K = \{r\}$. Suppose there is a base $B$ of $(W,S)$ of type ${\bf B}_{2p+1}$ or ${\bf D}_2(2p+1)$ for some $p\geq 1$ satisfying conditions (1), (2), (3) of the Blow-Down Theorem with $K = \{r\}$. By twisting $(W,S)$ as in the proof of the Blow-Down Theorem, leaving $B\cup B^\perp$ invariant, we may assume that $N(y)=B\cup B^\perp$. Then $W$ has a set of Coxeter generators $S'$ such that $|S'| = |S|-1$ by Lemma \[Sixth Lemma\]. Conversely, suppose $W$ has a set of Coxeter generators $S'$ such that $|S'| < |S|$. We may assume that $S'$ has the maximum possible number of basic subsets that isomorphically match basic subsets of $S$. Now $S$ has a basic subset $B$ that nonisomorphically matches a basic subset $B'$ of $S'$ by the Simplex Matching Theorem (Theorem 7.7 of [@M-R-T]). Then $|\langle B\rangle| \neq |\langle B'\rangle|$ by the Basic Matching Theorem. Assume first that $|\langle B\rangle| < |\langle B'\rangle|$. Then $B$ satisfies conditions (1), (2), (3) of the Blow-Down Theorem. Let $r$ and $K$ be as in Lemma \[Fourth Lemma\]. If $K =\{r\}$, we are done. Suppose that $K$ is of type ${\bf C}_{2q+1}$ or ${\bf D}_2(4q+2)$ for some $q\geq 1$. Then $K$ is a basic subset of $S$. Let $K'$ be the basic subset of $S'$ that matches $K$. We claim that $K$ isomorphically matches $K'$. On the contrary, suppose that $K$ nonisomorphically matches $K'$. By Lemma \[Fourth Lemma\], we have that $K'$ is a component of $(B')^\perp$ and $(K')\cup (K')^\perp = B'\cup(B')^\perp$. By Theorems 8.4-8.8 of [@M-R-T], we can blow up $S'$ along $B'$, and after twisting as in the proof of the Blow-Down Theorem, leaving $(K')\cup (K')^\perp$ invariant, we can blow down $K'$ to obtain a set of Coxeter generators $S''$ such that $|S''| = |S'|$ and $S''$ has two more basic subsets than $S'$ isomorphically matching basic subsets of $S$, which contradicts the choice of $S'$. Thus $K$ isomorphically matches $K'$. Let $z'$ be the longest element of $\langle B'\rangle$. As in the proof of Lemma \[Fourth Lemma\], the element $z'$ is in the center of $B^\perp$. By applying the automorphism $\theta$ of $W$ defined by $\theta(s) = s$ for all $s\in S-\{r\}$ and $\theta(r) = rz_2\cdots z_n$ as in the proof of Lemma \[Fourth Lemma\], if necessary, we may assume that $z'$ is the longest element of $\langle K\rangle$. By Theorems 8.5 and 8.7 of [@M-R-T], the group $W$ has a set of Coxeter generators $S''$ such that $K$ matches a base $K''$ of $(W,S'')$ with $|\langle K\rangle| > |\langle K''\rangle|$. Therefore $W$ has a set of Coxeter generators $S''$ such that $K'$ matches a base $K''$ of $(W,S'')$ with $|\langle K'\rangle| > |\langle K''\rangle|$. Let $\ell'$ be the longest element of $\langle K'\rangle$. If $K'$ is of type ${\bf C}_{2q+1}$, let $a'\in K'$ be such that $K'\cap{\rm Odd}(a') = \{a'\}$. If $K'$ is of type ${\bf D}_2(4q+2)$, let $a'\in K'$ be as in Lemma 8.6 of [@M-R-T]. As in the proof of Lemma \[Fourth Lemma\], define a homomorphism $\eta: \langle C'\rangle \to \langle\ell'\rangle$ as follows. Define $\eta(s') = 1$ for all $s'\in C'-\{a'\}$ and define $\eta(a')=\ell'$. Then $\eta$ is well defined and $\eta(\ell') = \ell'$. By the argument in the proof of Lemma \[Fourth Lemma\], the element $\ell'$ is in the center of $B^\perp$ and there is a component $L$ of $B^\perp$ such that $L\subseteq C$ and $L$ has nontrivial center, and if $\ell$ is the longest element of $\langle L\rangle$, then $\eta(\ell) = \ell'$ and $\eta(t) = \ell'$ for some $t\in L$. Moreover $L$ is of type ${\bf A}_1$, ${\bf C}_{2q+1}$, ${\bf D}_2(4q+2)$, ${\bf E}_7$, or ${\bf G}_3$ for some $q\geq 1$, and if $L$ is of type ${\bf C}_{2q+1}$, then $L\cap{\rm Odd}(t)=\{t\}$. As $\eta(z') = 1$, we have that $L\neq K$. Let $A = \ov{\{t\}}$. As $t\in \ov B$, we have that $\ov{\{t\}} \subseteq \ov B$. Then $\langle A\rangle$ is conjugate in $\langle C'\rangle$ to $\langle A'\rangle$ for some $A' \subseteq C'$ by Prop. 4.14 of [@M-R-T]. Now $\eta(\langle A'\rangle) = \eta(\langle A\rangle) = \langle \ell'\rangle$. Hence $a'\in A'$. Therefore $B'\subseteq A'$ by Lemmas 8.1 and 8.6 of [@M-R-T]. Hence $B\subseteq A$ by the Basic Matching Theorem, and so $\ov B\subseteq A$. Therefore $\ov{\{t\}}=\ov B$. As before, $N(t) = B\cup B^\perp$, and so ${\rm Odd}(t) \subseteq L$. If $L =\{t\}$, we are done, and so we may assume that $L$ is not of type ${\bf A}_1$. As before, by applying an automorphism, we may assume that $\ell = \ell'$. By Lemma 38 of [@H-M], there is an automorphism $\beta$ of $W$ such that $\beta(s) = s$ for all $s\in S-(\{r\}\cup {\rm Odd}(t))$, and $\beta(r) = r\ell z'$, and $\beta(s) = s\ell z'$ for all $s\in {\rm Odd}(t)$. Then $\beta(\ell) = z'$ and $\beta(z') = \ell$. As $\beta$ fixes each element of $S-(K\cup L)$ and $\beta$ leaves $\langle C\rangle$ invariant, we may replace $S$ by $\beta(S)$. Then $K$ is replaced by $\beta(L)$, and $\beta(K)$ is removed as a possibility for replacing $L$, since $\beta(K)$ matches $K'$ and $\langle\beta(K)\rangle$ and $\langle K'\rangle$ have the same longest element. By the argument in Lemma \[Fourth Lemma\], we may assume that $\beta(L)$ is of type ${\bf C}_{2q+1}$ or ${\bf D}_2(4q+2)$. In the above procedure only Coxeter generators of components of $B^\perp$ of type ${\bf C}_{2q+1}$, ${\bf D}_2(4q+2)$, ${\bf E}_7$, or ${\bf G}_3$ are replaced. By repeating this procedure a finite number of times, we can remove the possibility that $L$ is of type ${\bf C}_{2q+1}$, ${\bf D}_2(4q+2)$, ${\bf E}_7$, or ${\bf G}_3$. Then $L=\{t\}$, and we are done. Assume now that $|\langle B\rangle| > |\langle B'\rangle|$. Then $B'$ satisfies conditions (1), (2), (3) of the Blown-Down Theorem. Let $r'$ and $K'$ be as in Lemma \[Fourth Lemma\]. If $K'=\{r'\}$, we can blow down $B'$ to obtain a set of Coxeter generators $S''$ such that $|S''| = |S'|-1$ and $S''$ has one more basic subset than $S'$ isomorphically matching basic subsets of $S$, which contradicts the choice of $S'$. Therefore $K'$ is of type ${\bf C}_{2q+1}$ or ${\bf D}_2(4q+2)$ for some $q\geq 1$. As $N(r') = B'\cup (B')^\perp$, we have that $K'$ is a basic subset of $S'$ and $K'\cup (K')^\perp =B'\cup (B')^\perp$. Let $K$ be the basic subset of $S$ that matches $K'$. Then $K$ isomorphically matches $K'$, since otherwise by Theorems 8.4-8.8 of [@M-R-T], we can blow up $S'$ along $K'$ and then blow down $B'$ to obtain a set of Coxeter generators $S''$ such that $|S''| = |S'|$ and $S''$ has two more basic subsets than $S'$ isomorphically matching basic subsets of $S$, which contradicts the choice of $S'$. Let $z$ be the longest element of $\langle B\rangle$. As in the proof of Lemma \[Fourth Lemma\], the element $z$ is in the center of $(B')^\perp$, and by applying an automorphism, we may assume that $z$ is the longest element of $\langle K'\rangle$. Let $\ell$ be the longest element of $\langle K\rangle$. If $K$ is of type ${\bf C}_{2q+1}$, let $a\in K$ be such that $K\cap{\rm Odd}(a) = \{a\}$. If $K$ is of type ${\bf D}_2(4q+2)$, let $a\in K$ be as in Lemma 8.6 of [@M-R-T]. As in the proof of Lemma \[Fourth Lemma\], define a homomorphism $\eta: \langle C\rangle \to \langle\ell\rangle$ as follows. Define $\eta(s) = 1$ for all $s\in C-\{a\}$ and define $\eta(a)=\ell$. Then $\eta$ is well defined and $\eta(\ell) = \ell$. By the argument in the proof of Lemma \[Fourth Lemma\], the element $\ell$ is in the center of $(B')^\perp$ and there is a component $L'$ of $(B')^\perp$ such that $L'\subseteq C'$ and $L'$ has nontrivial center, and if $\ell'$ is the longest element of $\langle L'\rangle$, then $\eta(\ell') = \ell$ and $\eta(t') = \ell$ for some $t'\in L'$. Moreover $L'$ is of type ${\bf A}_1$, ${\bf C}_{2q+1}$, ${\bf D}_2(4q+2)$, ${\bf E}_7$, or ${\bf G}_3$ for some $q\geq 1$, and if $L'$ is of type ${\bf C}_{2q+1}$, then $L'\cap{\rm Odd}(t')=\{t'\}$. As $\eta(z) = 1$, we have that $L'\neq K'$. Let $A' = \ov{\{t'\}}$. As $t'\in \ov {B'}$, we have that $\ov{\{t'\}} \subseteq \ov{B'}$. Then $\langle A'\rangle$ is conjugate in $\langle C\rangle$ to $\langle A\rangle$ for some $A \subseteq C$ by Prop. 4.14 of [@M-R-T]. Now $\eta(\langle A\rangle) = \eta(\langle A'\rangle) = \langle \ell\rangle$. Hence $a\in A$. Therefore $B\subseteq A$ by Lemmas 8.1 and 8.6 of [@M-R-T]. Hence $B'\subseteq A'$ by the Basic Matching Theorem, and so $\ov{B'}\subseteq A'$. Therefore $\ov{\{t'\}}=\ov{B'}$. As before, $N(t') = B'\cup (B')^\perp$, and so ${\rm Odd}(t')\subseteq L'$. If $L'=\{t'\}$, we derive a contradiction as before. Therefore $L'$ is not of type ${\bf A}_1$. By applying an automorphism, we may assume that $\ell = \ell'$. By Lemma 38 of [@H-M], there is an automorphism $\beta$ of $W$ such that $\beta(s') = s'$ for all $s'\in S'-(\{r'\}\cup{\rm Odd}(t'))$, and $\beta(r') = r'\ell z$, and $\beta(s') = s'\ell z$ for all $s'\in {\rm Odd}(t')$. Then $\beta(\ell) = z$ and $\beta(z) = \ell$. As $\beta$ fixes each element of $S'-(K'\cup L')$ and $\beta$ leaves $\langle C'\rangle$ invariant, we may replace $S'$ by $\beta(S')$. Then $K'$ is replaced by $\beta(L')$, and $\beta(K')$ is removed as a possibility for replacing $L'$, since $\beta(K')$ matches $K$ and $\langle\beta(K')\rangle$ and $\langle K\rangle$ have the same longest element. By the argument in Lemma \[Fourth Lemma\], we may assume that $\beta(L')$ is of type ${\bf C}_{2q+1}$ or ${\bf D}_2(4q+2)$. In the above procedure only Coxeter generators of components of $(B')^\perp$ of type ${\bf C}_{2q+1}$, ${\bf D}_2(4q+2)$, ${\bf E}_7$, or ${\bf G}_3$ are replaced. By repeating this procedure a finite number of times, we can remove the possibility that $L'$ is of type ${\bf C}_{2q+1}$, ${\bf D}_2(4q+2)$, ${\bf E}_7$, or ${\bf G}_3$. Then $L'$ is of type ${\bf A}_1$ and we have a contradiction as before. Thus the case $|\langle B\rangle| > |\langle B'\rangle|$ leads to a contradiction. The Rank Spectrum of a Coxeter Group ==================================== In this section, we describe how to determine the set of all possible ranks of an arbitrary finitely generated Coxeter group $W$ by inspection of any presentation diagram for $W$. Let $(W,S)$ be a Coxeter system of finite rank. Suppose that $S_1,S_2\subseteq S$, with $S=S_1\cup S_2$ and $S_0=S_1\cap S_2$, are such that $m(a,b) = \infty$ for all $a\in S_1-S_0$ and $b\in S_2-S_0$. Let $\ell\in \langle S_0\rangle$ such that $\ell S_0\ell^{-1} = S_0$. The triple $(S_1,\ell,S_2)$ determines an elementary twist of $(W,S)$ giving a new set of Coxeter generators $S_\ast = S_1\cup\ell S_2\ell^{-1}$ for $W$ such that $S_1\cap\ell S_2\ell^{-1} = S_0$. Let $B$ be a base of $(W,S)$ of type ${\bf B}_{2p+1}$ or ${\bf D}_2(2p+1)$ for some $p\geq 1$ that satisfies the conditions (1), (2), (3) of the Blown-Down Theorem with $K=\{r\}$. As $B$ is a simplex, either $B\subseteq S_1$ or $B\subseteq S_2$. If $B\subseteq S_1$, define $B_\ast = B$. If $B\subseteq S_2$, define $B_\ast =\ell B\ell^{-1}$. If $B\subseteq S_0$, then $\ell B\ell^{-1} = B$ by Lemma 4.10 of [@M-R-T], and so $B_\ast$ is well defined. If $r\in S_1$, define $r_\ast = r$. If $r\in S_2$, define $r_\ast = \ell r\ell^{-1}$. If $r\in S_0$, then $\ell r\ell^{-1} = r$ by Lemma 4.8 of [@M-R-T], and so $r_\ast$ is well defined. The set $B_\ast$ is a base of $(W,S_\ast)$ of type ${\bf B}_{2p+1}$ or ${\bf D}_2(2p+1)$ for some $p\geq 1$ that satisfies the conditions (1), (2), (3) of the Blown-Down Theorem with $\{r_\ast\}$ a component of $B_\ast^\perp$. As $B_\ast$ is conjugate to $B$, we deduce that $B_\ast$ is a base of $(W,S_\ast)$ of type ${\bf B}_{2p+1}$ or ${\bf D}_2(2p+1)$ for some $p\geq 1$ by the Basic Matching Theorem. By the Blow-Down Theorem, $W$ has a set of Coxeter generators $S'$ such that $B$ matches a base $B'$ of $(W,S')$ with $|\langle B\rangle| < |\langle B'\rangle|$. Hence $B_\ast$ matches $B'$ with $|\langle B_\ast\rangle| < |\langle B'\rangle|$. Therefore $B_\ast$ satisfies conditions (1) and (2) of the Blow-Down Theorem. As $B\cup\{r\}$ is a simplex, either $B\cup\{r\}\subseteq S_1$ or $B\cup\{r\}\subseteq S_2$. Hence $r_\ast \in (B_\ast)^\perp$. Let $s_\ast\in N(r_\ast)-(B_\ast\cup\{r_\ast\})$. Assume first that $B\cup\{r\}\subseteq S_1$. Then $B_\ast = B$ and $r_\ast = r$. Suppose $s_\ast \in S_1$. Then $s_\ast \in N(r) = B\cup B^\perp$. Hence $s_\ast \in B^\perp$, and so $s_\ast\in (B_\ast)^\perp$ and $m(r_\ast,s_\ast) = 2$. Now suppose that $s_\ast\in \ell S_2\ell^{-1}-S_0$. Then $r\in S_0$ and $r=\ell r\ell^{-1}$ and $s_\ast = \ell s\ell^{-1}$ for some $s\in S_2-S_0$. Hence $s\in N(r) = B\cup B^\perp$. As $B\subset S_1$, we have that $s\in B^\perp$. Therefore $m(r,s) = 2$, and so $m(r_\ast,s_\ast) = 2$. Moreover, $B\subseteq S_0$, since $s\in S_2-S_0$. Hence $B_\ast = \ell B\ell^{-1}$, and so $s_\ast\in (B_\ast)^\perp$. Assume now that $B\cup\{r\}\subseteq S_2$. Then $B_\ast = \ell B\ell^{-1}$ and $r_\ast = \ell r\ell^{-1}$. Suppose $s_\ast \in S_1-S_0$. Then $r_\ast \in S_0$, and so $r_\ast = r$. Hence $s_\ast \in N(r) = B\cup B^\perp$. Now $s_\ast \in B^\perp$, since $s_\ast \in S_1-S_0$. Hence $m(r_\ast,s_\ast) = 2$. Moreover $B\subseteq S_0$, and so $B_\ast = B$. Hence $s_\ast \in (B_\ast)^\perp$. Now suppose $s_\ast\in \ell S_2\ell^{-1}$. Then $s_\ast = \ell s\ell^{-1}$ for some $s\in S_2$. As $1 < m(r_\ast,s_\ast) < \infty$, we have $1 < m(r,s) < \infty$. Hence $s\in N(r) = B\cup B^\perp$. Moreover $s\in B^\perp$, since $s_\ast \not\in B_\ast$. Hence $m(r,s) = 2$, and so $m(r_\ast,s_\ast) =2$. Moreover $s_\ast\in (B_\ast)^\perp$. Thus, in all cases, $N(r_\ast) = B_\ast\cup(B_\ast)^\perp$ and $\{r_\ast\}$ is a component of $(B_\ast)^\perp$. Therefore $B_\ast$ satisfies condition (3) of the Blow-Down Theorem with $K = \{r_\ast\}$. Let $B$ be a base of $(W,S)$ of type ${\bf B}_{2p+1}$ or ${\bf D}_2(2p+1)$ for some $p\geq 1$ satisfying conditions (1), (2), (3) of The Blow-Down Theorem with $K = \{r\}$. We call $r$ a [*sink*]{} for $B$. A element $r$ of $S$ may be a sink for more than one base of $(W,S)$. For example, ${\bf A}_1$ is a sink for the two bases of ${\bf A}_1\times {\bf D}_2(3)\times {\bf D}_2(3)$. A base may have more than one sink. For example, the base of ${\bf A}_1\times {\bf A}_1\times {\bf D}_2(3)$ has two sinks. Let $B$ be a base of $(W,S)$ of type ${\bf B}_{2p+1}$ or ${\bf D}_2(2p+1)$ for some $p\geq 1$ that satisfies the hypothesis of the Blow-Down Lemma with sink $r$. Let $S'$ be the set of Coxeter generators obtained by blowing down $S$ along $B$. Let $C$ be a base of $(W,S)$ of type ${\bf B}_{2q+1}$ or ${\bf D}_2(2q+1)$ for some $q\geq 1$ that satisfies the conditions (1), (2), (3) of the Blow-Down Theorem with sink $s$. If $B\neq C$ and $r\neq s$, then $C$ is a base of $(W,S')$ that satisfies the conditions (1), (2), (3) of the Blow-Down Theorem with sink $s$. Let $B'$ be the base of $(W,S')$ obtained by blowing down $B$. Then $S$ and $S'$ have the same basic subsets except for $B$ and $B'$ by Lemma 3.6. Therefore $C$ is a base of $(W,S')$. By the Blow-Down Theorem, $W$ has a set of Coxeter generators $S''$ such that $C$ matches a base $C''$ of $(W,S'')$ with $|\langle C\rangle| < |\langle C''\rangle|$. Therefore $C$ satisfies the conditions (1) and (2) of the Blow-Down Theorem. As ${\rm Odd}(s) = \{s\}$, we have that $s\not\in B$. Therefore $s\in S'$. Let $a$ be the element of $S'$ that is not in $S$. Then $a = r\ell$ with $\ell$ the longest element of $\langle B\rangle$. If $a\not\in N(s)$, then $N(s) = C\cup C^\perp$ and $\{s\}$ is a component of $C^\perp$ with respect to $S'$, since $s$ is a sink for $C$ with respect to $S$. Suppose $a\in N(s)$. Then $s\in N(r)$ by Lemma 8.3 of [@M-R-T]. Hence $s\in B^\perp$. Therefore $m(a,s) = 2$. As $B\cup\{r\} \subseteq N(s)$, we have that $B\cup\{r\}\subseteq C\cup C^\perp$. As $B\neq C$, we have that $B\subseteq C^\perp$. As ${\rm Odd}(r) = \{r\}$, we have that $r\not\in C$, and so $r\in C^\perp$. Therefore $a \in C^\perp$. Hence $N(s) = C\cup C^\perp$ and $\{s\}$ is a component of $C^\perp$ with respect to $S'$. Thus $C$ satisfies condition (3) of the Blow-Down Theorem, with sink $s$, with respect to $S'$. Let $B$ be a base of $(W,S)$ of type ${\bf C}_{2p+1}$ or ${\bf D}_2(4p+2)$ for some $p\geq 1$ along which $(W,S)$ can be blown up. Let $S'$ be the set of Coxeter generators obtained by blowing up $S$ along $B$. Let $C$ be a base of $(W,S)$ of type ${\bf B}_{2q+1}$ or ${\bf D}_2(2q+1)$ for some $q\geq 1$ that satisfies the conditions (1), (2), (3) of the Blow-Down Theorem with sink $s$. If $B\neq C$, then $C$ is a base of $(W,S')$ that satisfies the conditions (1), (2), (3) of the Blow-Down Theorem with sink $s$. Let $B'$ be the base of $(W,S')$ obtained by blowing up $B$. Then $S$ and $S'$ have the same basic subsets except for $B$ and $B'$ by Theorems 8.4 and 8.8 of [@M-R-T]. Therefore $C$ is a base of $(W,S')$. By the Blow-Down Theorem, $W$ has a set of Coxeter generators $S''$ such that $C$ matches a base $C''$ of $(W,S'')$ with $|\langle C\rangle| < |\langle C''\rangle|$. Therefore $C$ satisfies the conditions (1) and (2) of the Blow-Down Theorem. As ${\rm Odd}(s) = \{s\}$, we have that $s\not\in B$. Therefore $s\in S'$. Let $z$ be the longest element of $\langle B\rangle$. If $B$ is of type ${\bf C}_{2p+1}$, let $a,b,c$ be the elements of $B$ such that $m(a,b) = 4$ and $m(b,c)=3$. Then $N(a)=B\cup B^\perp$ by Theorem 8.2 of [@M-R-T]. If $B$ is of type ${\bf D}_2(4p+2)$, let $B=\{a,b\}$ with $N(a)=B\cup B^\perp$. In either case, let $d =aba$. Then $d$ and $z$ are the elements of $S'$ that are not in $S$ by Theorems 8.4 and 8.8 of [@M-R-T]. If $d$ and $z$ are not elements of $N(s)$, then $N(s) = C\cup C^\perp$ and $\{s\}$ is a component of $C^\perp$ with respect to $S'$, since $s$ is a sink for $C$ with respect to $S$. Suppose $d$ or $z$ is an element of $N(s)$. Then $s\in N(a) = B\cup B^\perp$ by Lemma 8.3 of [@M-R-T]. Hence $s\in B^\perp$. Therefore $m(d,s) = 2$ and $m(z,s)=2$. As $B\subseteq N(s)$, we have that $B\subseteq C\cup C^\perp$. As $B\neq C$, we have that $B\subseteq C^\perp$. Therefore $\{d,z\}\subseteq C^\perp$. Hence $N(s) = C\cup C^\perp$ and $\{s\}$ is a component of $C^\perp$ with respect to $S'$. Thus $C$ satisfies condition (3) of the Blow-Down Theorem, with sink $s$, with respect to $S'$. Let $B$ be a base of $(W,S)$ of type ${\bf C}_{2p+1}$ or ${\bf D}_2(4p+2)$ for some $p\geq 1$ along which $(W,S)$ can be blown up. Let $S'$ be the set of Coxeter generators obtained by blowing up $S$ along $B$. Let $C$ be a base of $(W,S)$ of type ${\bf C}_{2q+1}$ or ${\bf D}_2(4q+2)$ for some $q\geq 1$ along which $(W,S)$ can be blown up. If $B\neq C$, then $C$ is a base of $(W,S')$ along which $(W,S')$ can be blown up. Let $B'$ be the base of $(W,S')$ obtained by blowing up $B$. Then $S$ and $S'$ have the same basic subsets except for $B$ and $B'$ by Theorems 8.4 and 8.8 of [@M-R-T]. Therefore $C$ is a base of $(W,S')$. Let $z$ be the longest element of $\langle B\rangle$. If $B$ is of type ${\bf C}_{2p+1}$, let $a,b,c$ be the elements of $B$ such that $m(a,b) = 4$ and $m(b,c)=3$. Then $N(a)=B\cup B^\perp$ by Theorem 8.2 of [@M-R-T]. If $B$ is of type ${\bf D}_2(4p+2)$, let $B=\{a,b\}$ with $N(a)=B\cup B^\perp$. In either case, let $d =aba$. Then $d$ and $z$ are the elements of $S'$ that are not in $S$ by Theorems 8.4 and 8.8 of [@M-R-T]. Let $v$ be the element of $C$ such that $N(v) = C\cup C^\perp$ as in Theorems 8.5 and 8.7 of [@M-R-T]. If $d$ and $z$ are not elements of $N(v)$, then $N(v) = C\cup C^\perp$ with respect to $S'$. Suppose $d$ or $z$ is an element of $N(v)$. Then $v\in N(a) = B\cup B^\perp$ by Lemma 8.3 of [@M-R-T]. Hence $v\in B^\perp$. As $B\subseteq N(v)$, we have that $B\subseteq C\cup C^\perp$. As $B\neq C$, we have that $B\subseteq C^\perp$. Therefore $\{d,z\}\subseteq C^\perp$. Hence $N(v) = C\cup C^\perp$ with respect to $S'$. Then $(W,S')$ can be blown up along $C$ by Theorems 8.4 and 8.8 in [@M-R-T]. [(Rank Spectrum Theorem)]{} Let $\{B_1,\ldots, B_k\}$, $k\geq 0$, be a maximal set of bases of $(W,S)$ of type ${\bf B}_{2p+1}$ or ${\bf D}_2(2p+1)$ for some $p\geq 1$ that satisfy the conditions of the Blown-Down Theorem with distinct sinks $\{s_1,\ldots, s_k\}$. Let $C_1,\ldots, C_\ell$, $\ell \geq 0$, be the bases of $(W,S)$ of type ${\bf C}_{2q+1}$ or ${\bf D}_2(4q+2)$ for some $q\geq 1$ along which $(W,S)$ can be blown up. Then the set of all possible ranks of $W$ is $\big\{|S|-k, \ldots, |S|+\ell\big\}$. By Lemmas 5.1 and 5.2, we get a sequence of sets of Coxeter generators $S_0,\ldots, S_k$ for $W$ such that $S = S_0$ and $S_i$ is obtained from $S_{i-1}$ by twisting $S_i$, as in Theorem 3.7, and then blowing down along a base conjugate to $B_i$ for each $i=1,\ldots,k$. Then $S_k$ has minimum rank over all sets of Coxeter generators of $W$ by Theorem 4.1 and Lemma 5.3. Hence, the minimum rank of $W$ is $|S|-k$. Let $a_i \in C_i$ be the element of $C_i$, for $i=1,\ldots,\ell$ that is removed in the blowing up process. As $N(a_i) = C_i\cup C_i^\perp$ for each $i$, we have that $C_i$ is the only base of $(W,S)$ that contains $a_i$ for each $i$. By Lemma 5.4, we have a sequence of sets of Coxeter generators $S^{(0)},\ldots, S^{(\ell)}$ for $W$ such that $S = S^{(0)}$ and $S^{(i)}$ is obtained from $S^{(i-1)}$ by blowing up $S^{(i)}$ along $C_i$ for each $i=1,\ldots,\ell$. Then $S^{(\ell)}$ has maximum rank over all sets of Coxeter generators of $W$ by Theorem 9.1 of [@M-R-T]. Hence, the maximum rank of $W$ is $|S|+\ell$. Thus the set of all possible ranks of $W$ is $\big\{|S|-k, \ldots, |S|+\ell\big\}$. The numbers $k$ and $\ell$ in the Rank Spectrum Theorem can be determined by inspecting the presentation diagram of $(W,S)$. For example, $k =\ell = 1$ for the system ${\bf A}_1\times {\bf D}_2(3)\times{\bf D}_2(6)$. [99]{} N. Bourbaki, [*Groupes et algèbres de Lie*]{}, Chapitres 4, 5, et 6, Hermann, Paris, 1968. H.S.M. Coxeter, [*Regular Polytopes*]{}, Dover, New York, 1973. W.N. Franzsen and R.B. Howlett, Automorphisms of nearly finite Coxeter groups, [*Adv. Geom.*]{} [**3**]{} (2003), 301-338. R.B. Howlett and B. Mühlherr, Isomorphisms of Coxeter groups which do not preserve reflections, preprint (2004), 18p. M. Mihalik, The even isomorphism theorem for Coxeter groups, [*Trans. Amer. Math. Soc.*]{} [**359**]{} (2007), 4297-4324. M. Mihalik, J. Ratcliffe, and S. Tschantz, Matching theorems for systems of a finitely generated Coxeter group, [*Algebr. Geom. Topol.*]{} [**7**]{} (2007), 919-956.
{ "pile_set_name": "ArXiv" }
--- author: - Charles Baker title: 'A proof of the parabolic Schauder estimates using Trudinger’s method and the mean value property of the heat equation' --- Introduction ============ One method available to prove the Schauder estimates is Neil Trudinger’s method of mollification ([@Tru]). In the case of second order elliptic equations, the method requires little more than mollification and the solid mean value inequality for subharmonic functions. The method was lated adapted to the parabolic setting by Xu-Jia Wang in [@Wan], however in that presentation Wang uses an auxiliary estimate coming from the fundamental solution of the heat equation and the mean value of property of subsolutions of the heat equations is not used. Our goal in this article is show how the mean value property of subsolutions of the heat equation can be used in a similar fashion as the solid mean value inequality for subharmonic functions in Trudinger’s original elliptic treatment, providing a relatively simple derivation of the interior Schauder estimate for second order parabolic equations. Preliminaries ============= For an open subset $U \in \mathbb{R}^d$, we denote the corresponding open parabolic domain $U \times (0, T) \subset \mathbb{R}^{d+1}$ by $U_T$. We denote a (backwards) parabolic cylinder by $Q_R = B_R \times (t - R^2, t)$ . We often notate a point $(x,t) \in U_T$ by $X$. A (second order parabolic) mollifier is a fixed smooth function $\rho \in C_c^{ \infty }( \mathbb{R}^{d+1})$ with $\iint_{ \mathbb{R}^{d+1} } \rho \, dx \, dt = 1$. For $\tau > 0$ we define the scaled mollifier $$\rho_{\tau}(x,t) := \frac{1}{\tau^{d+2}} \rho\left(\frac{x}{\tau},\frac{t}{\tau^{2}}\right).$$ Let $U \in \mathbb{R}^{d+1}$ and $u \in L^1_{\text{loc}}(U)$. For $0 < \tau < d(X, {\partial}P)$, the mollification of $u$ is given by $$u_{\tau}(x,t) := \frac{1}{\tau^{d+2}} \iint \rho \left(\frac{x-y}{\tau},\frac{t-s}{\tau^{2}} \right) u(y,s) \, dy \, ds$$ and satisifes spt $u_{\tau} \subset U_{\tau}$, where $U_{\tau} = \{ X \in U : d(X, {\partial}U) > \tau \}$. The parabolic distance between two points $X = (x,t)$ and $Y = (y,s)$ is defined to be $$d(X, Y) := \max \{ {\lvert x-y\rvert}, {\lvertt-s\rvert}^{1/2} \}.$$ We use both $\sup$ and ${\lvert \cdot \rvert}_0$ to denote the supremum of a function. The Hölder seminorm is defined by $$[ u ]_{\alpha; U_T} : = \sup_{X \neq Y \in U_T} \frac{ {\lvert u(X) - u(Y) \rvert} }{ d(X,Y)^{\alpha} }$$ and the Hölder norms by $$\begin{aligned} &{\lvert u \rvert}_{2, 1; \, U_T } := \sum_{ k=0 }^{2} {\lvert {\partial}_x^k u \rvert}_{ 0; \, U_T } + {\lvert {\partial}_t u \rvert}_{ 0; \, U_T } \\ &{\lvert u \rvert}_{2, 1, \alpha; \, U_T } := {\lvert u \rvert}_{2, 1; \, U_T } + [ {\partial}_x^{2} u ]_{ \alpha; \, U_T } + [ {\partial}_t u ]_{ \alpha; \, U_T } \\\end{aligned}$$ The set of functions $$\{ u \in C^{2, 1}(U_T) : [u]_{ 2, 1, \alpha; \, U_T } < \infty \}$$ endowed with the norm ${\lvert u \rvert}_{ 2m, 1, \alpha; \, U_T }$ is called a Hölder space. Written out in full the norm is $${\lvert u \rvert}_{ 2, 1, \alpha; \, U_T } := \sum_{ k=0 }^{2} {\lvert {\partial}_x^k u \rvert}_{ 0; \, U_T } + {\lvert {\partial}_t u \rvert}_{ 0; \, U_T }+ [ {\partial}_x^{2} u ]_{ \alpha; \, U_T } + [ {\partial}_t u ]_{ \alpha; \, U_T }.$$ We sometimes use $${\partial}^{2,1} := {\partial}_x^2 + {\partial}_t,$$ e.g, $$[ {\partial}^{2,1}u ]_{\alpha; U_T} = [ {\partial}_x^2 u ]_{\alpha; U_T} + [ {\partial}_t u ]_{\alpha; U_T},$$ to compress notation a little. We begin by establishing some very basic estimates that will be use throughout. We have $u_{\tau} \in C_c^{\infty}$. Let $u \in L^1_{\text{loc}}(U)$. The following estimates hold: $$\begin{aligned} &{\lvert u_{ \tau } \rvert}_{0; \, U_{\tau} } \leq {\lvertu\rvert}_{ 0; \, U_{\tau} } \label{e: Holder est 1} \\ &{\lvert {\partial}_x^i {\partial}_t^j u_{\tau}(x,t) \rvert}_{ 0; \, U_{\tau} } \leq C \tau^{-i-2j} {\lvertu\rvert}_{0; \, U_{\tau} }. \label{e: Holder est 2} \end{aligned}$$ To prove , we have $$\begin{aligned} u_{ \tau }(x,t) &= \frac{1}{ \tau^{d+2} } \iint \rho \left( \frac{x-y}{\tau},\frac{t-s}{ \tau^{2}} \right) u(y,s) \, dy \, ds \\ &\leq {\lvert u \rvert}_{0; \, U_{\tau} } \cdot \frac{1}{ \tau^{d+2} } \iint \rho \left( \frac{x-y}{\tau},\frac{t-s}{ \tau^{2}} \right) \, dy \, ds \\ &= {\lvert u \rvert}_{0; \, U_{\tau} }.\end{aligned}$$ And for : $$\begin{aligned} {\partial}_x^i {\partial}_t^j u_{\tau}(x,t) &= \frac{1}{ \tau^{d+2} } \iint_{U_{\tau}} {\partial}_x^i {\partial}_t^j \rho \left( \frac{x-y}{\tau},\frac{t-s}{ \tau^{2}} \right) u(y,s) \, dy \, ds \\ &\leq C \tau^{-i-2j}{\lvert u \rvert}_{ 0; \, U_{\tau} }.\end{aligned}$$ Let $u \in C^{\alpha}_{\text{loc}}(U)$. The following estimates hold: $$\begin{aligned} &{\lvert u_{\tau}(x,t) - u(x,t) \rvert}_{ 0; \, U_{\tau} }\leq \tau^{\alpha} [u]_{\alpha; \, U_{\tau} } \label{e: Holder est 3} \\ &{\lvert {\partial}_x^i {\partial}_t^j u_{\tau}(x,t) \rvert}_{ 0; \, U_{\tau} } \leq C \tau^{ \alpha -i-2j } [u]_{\alpha; \, U_{\tau} }. \label{e: Holder est 4} \end{aligned}$$ For estimate we have $$\begin{aligned} u_{\tau}(x,t) - u(x,t) &= \frac{1}{ \tau^{d+2} } \iint \rho \left( \frac{x-y}{\tau},\frac{t-s}{ \tau^{2}} \right) (u(y,s) - u(x,t)) \, dy \, ds \\ &\leq \operatorname{osc}_{U_{\tau} } u \\ &\leq \tau^{\alpha} [u]_{\alpha; U_{\tau} }.\end{aligned}$$ To prove the second estimate we have $$\begin{aligned} {\partial}_x^i {\partial}_t^j u_{\tau}(x,t) &= \frac{1}{ \tau^{d+2} } \iint {\partial}_x^i {\partial}_t^j \rho \left( \frac{x-y}{\tau},\frac{t-s}{ \tau^{2}} \right) u(y,s) \, dy \, ds \\ &= \frac{1}{ \tau^{d+2} } \iint {\partial}_x^i {\partial}_t^j \rho \left( \frac{x-y}{\tau},\frac{t-s}{ \tau^{2}} \right) (u(y,s) - u(x,t)) \, dy \, ds \\ &\quad + \frac{u(x,t)}{ \tau^{d+2} } \iint {\partial}_x^i {\partial}_t^j \rho \left( \frac{x-y}{\tau},\frac{t-s}{ \tau^{2}} \right) \, dy \, ds.\end{aligned}$$ The mollifier $\rho$ is has compact support on $U_{\tau}$ and so the last term vanishes by the Divergence Theorem. Continuing, we have $$\begin{aligned} {\partial}_x^i {\partial}_t^j u_{\tau}(x,t) &= \frac{1}{ \tau^{d+2} } \iint_{U_{\tau}} {\partial}_x^i {\partial}_t^j \rho \left( \frac{x-y}{\tau},\frac{t-s}{ \tau^{2}} \right) (u(y,s) - u(x,t)) \, dy \, ds \\ &\leq C \tau^{-i-2j}\operatorname{osc}_{Q_\tau} u \\ &\leq C \tau^{ \alpha -i-2j } [u]_{ \alpha; \, U_{\tau} }. \end{aligned}$$ The interior elliptic Schauder estimate for functions of compact support ======================================================================== To motivate things a little for the parabolic setting, we first briefly show how Trudinger’s method works in the elliptic realm by treating the Poisson equation. The crucial ingredient in Trudinger’s method is the following norm equivalence: Let $U \subset \mathbb{R}^d$ be an open bounded domain and $u \in C^{ \alpha }(U)$, where $\alpha \in (0,1)$. Let $R > 0$ and $\tau_0$ be small constants, both of which will be fixed in the course of the proof. There exists a constant $C = C(d, \alpha)$ such that the norm equivalence $$\frac{1}{C}[u]_{ \alpha; \, B_{R} } \leq \sup_{ 0 < \tau < \tau_0 } \tau^{1-\alpha} {\lvert{\partial}_x u_{\tau}\rvert}_{ 0; \, B_{ R } } \leq C[u]_{ \alpha; \, B_{R} }.$$ is valid. The inequality on the right follows directly from equation (the elliptic version) by choosing the appropriate values for the indices $i$: choosing $i = 1$ (there is no $j$ in the elliptic mollifier) gives $${\lvert {\partial}_x u_{\tau}\rvert}_{ 0; \, B_{ R } } \leq C \tau^{ \alpha - 1 } [u]_{\alpha; \, B_{ R } }.$$ The first inequality requires a little more work. Let $x, y \in U$ satisfy ${\lvertx-y\rvert} < d_x/2$ and let $\tau$ satisfy $0 < \tau <\tau_0 < d_x/2$, where $\tau_0$ will be fixed shortly. For $ {\lvert x - y \rvert} < R = d_x/2 $, by the triangle inequality $$\begin{aligned} {\lvertu(x) - u(y)\rvert} &\leq {\lvertu(x) - u_{\tau}(x)\rvert} + {\lvert u_{\tau}(x) - u_{\tau}(y) \rvert} + {\lvert u_{\tau}(y) - u(y) \rvert} \\ &\leq 2\tau^{\alpha}[u]_{ \alpha; \, B_{ R }} + {\lvert {\partial}_x u_{\tau} \rvert}_{ 0; \, B_{ R } }{\lvertx-y\rvert}.\end{aligned}$$ Set $\tau = \epsilon {\lvertx-y\rvert}$, where $\epsilon < 1/2$. Factoring out and dividing by ${\lvertx-y\rvert}^{\alpha}$ we find $${\lvert u(x) - u(y) \rvert} \leq {\lvertx-y\rvert}^{\alpha} \left( 2\epsilon^{\alpha}[u]_{\alpha; B_R} + \epsilon^{\alpha - 1}\tau^{1-\alpha} {\lvert {\partial}_x u_{\tau} \rvert}_{0; \, B_R} \right).$$ Choose $\epsilon = 4^{-1/\alpha}$ then set $\tau_0 = 4^{-1/\alpha} . \, d_x/2$. Taking the supremum over $\tau \in (0, \tau_0)$ completes the proof. We now derive the interior Schauder estimate for Poisson’s equation. For simplicity, we only consider solutions with compact support in a ball $B_R$. The method extends in the usual way to more general equations and domains by using cutoff functions and Simon’s absorption lemma (see, for example, [@Sim]). Suppose that $u \in C_c^{2, \alpha}(B_R)$ solves $$-a^{ij}(x){\partial}_{ij}u(x) = f(x),$$ where we assume $a^{ij}, f \in C^{\alpha}(B_R)$, and there are constants $\lambda >0$ and $\Lambda < \infty$ such that $\lambda {\lvert\xi\rvert}^2 \leq a^{ij}\xi_i\xi_j \leq \Lambda {\lvert \xi \rvert}^2$. We proceed by the method of freezing coefficients, and accordingly fix a point $x_0 \in \mathbb{R}^d$ a rewrite the above equation equation as $$\begin{aligned} \notag -a^{ij}(x_0){\partial}_{ij}u(x) - f (x_0)&= (a^{ij}(x) - a^{ij}(x_0) ){\partial}_{ij}u(x) + f(x) - f(x_0)\\ &:= g(x) \label{eqn: Poisson 1}\end{aligned}$$ After a coordinate transformation, we can assume without loss of generality that $a^{ij}(x_0) = \delta^{ij}$, that is we can assume is the Poisson equation. Mollify equation to get $$-\Delta u_{\tau} - f(x_0) = g_{\tau}$$ and then differentiate thrice with respect to $x$ to obtain $$-\Delta {\partial}_x^3 u_{\tau} = {\partial}_x^3 g_{\tau}.$$ Using inequality we can estimate $$\begin{aligned} {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, B_R } &\leq C(d) \tau^{-3} {\lvert g \rvert}_{0; \, B_R } \\ &\leq C(d) \tau^{-3} R^{\alpha} \big( [ a ]_{\alpha; \, B_R } {\lvert {\partial}_x^2 u \rvert}_{0; \, B_R } + [ f ]_{\alpha; \, B_R } \big).\end{aligned}$$ Recall the solid mean value inequality for subharmonic functions: If $v$ solves $-\Delta v(x) \leq 0$ on a ball $B_R(x) \subset \mathbb{R}^d$, then $v$ satisfies $$v(x) \leq \frac{ C(d) }{ R^n } \int_{B_R} v(y) \, dy.$$ To apply this inequality to our situation, noting $\Delta {\lvert x \rvert}^2 = 2d$, we have $$-\Delta \left( {\partial}_x^3 u_{\tau} + \frac{ {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, B_R } {\lvert x \rvert}^2 }{ 2n } \right) = -\Delta {\partial}_x^3 u_{\tau} - {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, B_R} \leq 0.$$ Thus the function ${\partial}_x^3 u_{\tau} + {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, B_R } {\lvert x \rvert}^2 / (2d)$ is subharmonic and applying the mean value inequality and estimating we obtain $$\begin{aligned} {\lvert {\partial}_x^3 u_{\tau}(x_0) \rvert} &\leq C(d) \left( \frac{1}{R^d} \left\lvert \int_{B_R} {\partial}_y^3 u_{\tau}(y) \, dy \right\rvert + R^2 {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, B_R } \right) \\ &\leq C(d) \left( \frac1R \operatorname{osc}_{ B_R } {\partial}_x^2 u_{\tau}(x) + \tau^{-3} R^{2+\alpha} \big( [ a ]_{\alpha; B_{R} } {\lvert {\partial}_x^2 u \rvert}_{0; \, B_{R} } + [ f ]_{\alpha; \, B_{R} } \big) \right) \\ &\leq C(d) \Big( R^{ \alpha - 1 } [ {\partial}_x^2 u ]_{ \alpha; \, B_R } + \tau^{-3} R^{2+\alpha} \big( [ a ]_{\alpha; B_{R} } {\lvert {\partial}_x^2 u \rvert}_{0; \, B_{R} } + [ f ]_{\alpha; \, B_{R} } \big) \Big).\end{aligned}$$ We set $R = N\tau$ and return to the original coordinates to find $$\tau^{1-\alpha}{\lvert {\partial}_x^3 u_{\tau}(x_0) \rvert} \leq C(d, \lambda, \Lambda) \left( N^{ \alpha - 1 } [ {\partial}_x^2 u ]_{ \alpha; \, B_R } + N^{2+\alpha} \big( [ a ]_{\alpha; B_{R} } {\lvert {\partial}_x^2 u \rvert}_{0; \, B_{R} } + [ f ]_{\alpha; \, B_{R} } \big) \right),$$ then taking the supremum over $\tau$ and using the norm equivalence we obtain $$[ {\partial}_x^2 u ]_{\alpha; \, B_R } \leq C(d, \lambda, \Lambda, \alpha) \left( N^{ \alpha - 1 } [ {\partial}_x^2 u ]_{ \alpha; \, B_R } + N^{2+\alpha} \big( [ a ]_{\alpha; B_R } {\lvert {\partial}_x^2 u \rvert}_{0; \, B_R } + [ f ]_{\alpha; \, B_R } \big) \right).$$ Choosing $N$ sufficiently large and using the Hölder space interpolation inequality on the right gives the desired estimate, namely $$[ {\partial}_x^2 u ]_{\alpha; \, B_R} \leq C \big([ f ]_{\alpha; \, B_R } + {\lvert u \rvert}_{0; \, B_R } \big),$$ where $C$ depends on $d, \lambda, \Lambda$, and $\alpha$. The interior parabolic Schauder estimate for functions of compact support ========================================================================= Having given a feel for Trudinger’s method, we move on to use this method to derive the Schauder estimates for second order parabolic equations. The crucial equivalence of norms lemma in the parabolic setting is the following: Let $U \subset \mathbb{R}^{d+1}$ be an open bounded domain and $u \in C^{ \alpha }(U)$, where $\alpha \in (0,1)$, and $R > 0$ and $\tau_0$ be small constants, both of which will be fixed in the course of the proof. There exists a constant $C = C(d, \alpha)$ such that the norm equivalence $$\frac{1}{C}[u]_{ \alpha; \, Q_{R} } \leq \sup_{0 < \tau < \tau_0 } \left\{ \tau^{1-\alpha} {\lvert{\partial}_x u_{\tau}\rvert}_{ 0; \, Q_{ R } } + \tau^{2-\alpha}{\lvert{\partial}_t u_{\tau}\rvert}_{0; Q_{R} } \right\} \leq C[u]_{ \alpha; \, Q_{R} }.$$ is valid. The second inequality follows directly from equation by choosing the appropriate values for the indices $i$ and $j$. To prove the spatial part of the second inequality, choosing $i = 1$ and $j = 0$ in estimate gives $${\lvert {\partial}_x u_{\tau}(x,t) \rvert}_{ 0; \, Q_{ R } } \leq C \tau^{ \alpha - 1 } [u]_{\alpha; \, Q_{ R } }.$$ The temporal estimate follows similarly. Moving on to the proof of the first inequality, let $X,Y \in U_T $ be points such that $d(X,Y) < d_X/2$. To simplify notation a little, set $R = d_X/2$; thus $d(X,Y) \in Q_R \subset U_T$. Let $\tau$ satisfy $0 < \tau <\tau_0 < R$, where $\tau_0$ will be fixed shortly. For $d(X,Y) < R$, $$\begin{aligned} {\lvertu(X) - u(Y)\rvert} &\leq {\lvertu(X) - u_{\tau}(X)\rvert} + {\lvert u_{\tau}(Y) - u(Y) \rvert} + {\lvert u_{\tau}(x,t) - u_{\tau}(y,t) \rvert} + {\lvert u_{\tau}(y,t) - u_{\tau}(y,s) \rvert} \\ &\leq 2\tau^{\alpha}[u]_{ \alpha; \, Q_{ R }} + {\lvertx-y\rvert} {\lvert {\partial}_x u_{\tau} \rvert}_{ 0; \, Q_{ R } } + {\lvertt-s\rvert} {\lvert {\partial}_t u_{\tau} \rvert}_{ 0; \, Q_{ R } }.\end{aligned}$$ Set $\tau = \epsilon d(X,Y)$, where $\epsilon < 1/2$. Factoring out $d(X,Y)^{\alpha}$ we have $${\lvertu(X) - u(Y)\rvert} \leq d(X,Y)^{\alpha} \left( 2\epsilon^{\alpha}[u]_{\alpha; \, Q_{ R }} + \epsilon^{\alpha-1}\tau^{1-\alpha}{\lvert {\partial}_x u_{\tau} \rvert}_{ 0; \, Q_{ R } } + \epsilon^{\alpha-2}\tau^{2-\alpha}{\lvert {\partial}_t u_{\tau} \rvert}_{ 0; \, Q_{ R } } \right).$$ Choose $\epsilon = 4^{-1/\alpha}$ then set $\tau_0 = 4^{-1/\alpha} \cdot \, d_X/2$. Taking the supremum over $\tau \in (0, \tau_0)$ completes the proof. We now proceed similarly to Poisson’s equation to derive the Schauder estimate for the nonhomongeneous heat equation. Again for simplicity, we only condisider solutions with compact support in a parabolic cylinder $Q_R$ as the more general estimates follow from this case using cutoff functions and Simon’s adsorption lemma. Suppose that $u \in C_c^{2, \alpha}(Q_R)$ solves $${\partial}_t u(x,t) - a^{ij}(x,t){\partial}_{ij}u(x,t) = f(x,t),$$ where we assume $a^{ij}, f \in C^{\alpha}(Q_R)$, and there are constants $\lambda >0$ and $\Lambda < \infty$ such that $\lambda {\lvert\xi\rvert}^2 \leq a^{ij}\xi_i\xi_j \leq \Lambda {\lvert \xi \rvert}^2$. Again we freeze coefficients at a point $(x_0, t_0) \in Q_R$ and perform a coordinate transformation if necessary to get $$\begin{aligned} {\partial}_t u(x,t) - \Delta u(x,t) - f(x_0, t_0) &= ((a^{ij}(x,t) - a^{ij}(x_0,t_0) ) {\partial}_{ij}u(x,t) + f(x,t) - f(x_0,t_0) \\ &:= g(x,t),\end{aligned}$$ and then mollify the equation to obtain $${\partial}_t u_{\tau}(x,t) - \Delta u_{\tau}(x,t) = g_{\tau}(x,t). \label{eqn: heat subsol 0}$$ Given the form of the norm equivalence, the desired Schauder estimate will follow if we can establish the estimates (for the spatial component of the Schauder estimate) $$\begin{aligned} {\lvert {\partial}_x^3 u_\tau(x, t) \rvert} &\leq C \left( \frac1R \operatorname{osc}_{ Q_R } {\partial}_x^2 u + R^2{\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, Q_R} \right) \label{eqn: heat subsol 1} \\ {\lvert {\partial}_t{\partial}_x^2 u_\tau(x, t) \rvert} &\leq C \left( \frac{1}{R^2} \operatorname{osc}_{ Q_R } {\partial}_x^2 u + R^2{\lvert {\partial}_t{\partial}_x^2 g_{\tau} \rvert}_{0; \, Q_R} \right), \label{eqn: heat subsol 2}\end{aligned}$$ and for the temporal part $$\begin{aligned} {\lvert {\partial}_x{\partial}_t u_\tau(x, t) \rvert} &\leq C \left( \frac1R \operatorname{osc}_{ Q_R } {\partial}_t u + R^2{\lvert {\partial}_x{\partial}_t g_{\tau} \rvert}_{0; \, Q_R} \right) \label{eqn: time subsol 1} \\ {\lvert {\partial}_t^2 u_\tau(x, t) \rvert} &\leq C \left( \frac{1}{R^2} \operatorname{osc}_{ Q_R } {\partial}_t u + R^2{\lvert {\partial}_t^2 g_{\tau} \rvert}_{0; \, Q_R} \right). \label{eqn: time subsol 2}\end{aligned}$$ We show how to obtain estimate as the other three estimates follow in the same way. Recall the mean value property for subsolutions of the heat equation: If $v$ is a subsolution to the heat equation on $\mathbb{R}^{d+1}$, that is if $v$ solves ${\partial}_v - \Delta v \leq 0$, then $v$ satisfies $$v(x, t) \leq \frac{1}{4r^n}\iint_{E(x,t;\,r)} v(y,s) \frac{ {\lvert x - y \rvert}^2 }{ (t-s)^2 } \, dy \, ds$$ for each $E(x,t; r) \subset \mathbb{R}^{d+1}$, where the heat ball $E(x,t;r)$ is the set given by $E(x,t;r) = \{ (y,s) \in \mathbb{R}^{d+1} : {\lvertx-y\rvert}^2 \leq \sqrt{ -2\pi s \log[r^2/(-4\pi s)] }, s \in (t - r^2/(4\pi s), t) \}$. The radius of the heat ball $\sqrt{ -2\pi s \log[r^2/(-4\pi s)] }$ is often denoted by $R_r(s)$. We refer the interested reader to [@Eva] and [@Eck] for more information on the mean value property and heat balls. Let us now show : Differentiate thrice in space. Since ${\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, E}{\lvertx\rvert}^2/(2n)$ is independent of time we see $$\begin{aligned} {\partial}_t \left( {\partial}_x^3 u_{\tau} + {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, E}\frac{ {\lvertx\rvert}^2 }{2n} \right) - \Delta \left( {\partial}_x^3 u_{\tau} + {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, E}\frac{ {\lvertx\rvert}^2 }{2n} \right) &= {\partial}_t({\partial}_x^3 u_{\tau}) - \Delta ({\partial}^3_x u_{\tau}) - {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, E} \\ &= {\partial}_x^3 g_{\tau} - {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, E} \leq 0,\end{aligned}$$ and hence the function ${\partial}_x^3 u_{\tau} + {\lvert {\partial}_x^3 g_{\tau} \rvert}_{0; \, E(x_0, t_0; r) } {\lvert x \rvert}^2 / (2n)$ is subsolution of the heat equation. From the mean value property of subsolutions we have $${\partial}_x^3 u_{\tau}(x,t) \leq \frac{1}{4r^n} \iint_{E(x,t;r)} \left( {\partial}_y^3 u_{\tau}(y,s) + {\lvert{\partial}_x^3 g_{\tau}\rvert}_{0} \frac{ {\lvert y \rvert}^2 }{ 2n } \right) \frac{ {\lvert x - y \rvert}^2 }{ {\lvert t - s \rvert}^2 } \, dy ds. \label{eqn: heat subsol 4}$$ By translating coordinates we can assume without loss of generality that $(x,t) = (0,0)$. Establishing the required estimates involves evaluation of an integral of the form $$\frac{1}{r^n}\int_{ \frac{-r^2}{4\pi} } \frac{ R_r(s)^{\alpha} }{ s^{\beta} } \, ds,$$ where $\alpha$ and $\beta$ are given integers. The constants can be computed explicitly, however we are only interested in the scaling behaviour with respect to the radius $r$ (and that the integral is finite). We compute $$\begin{aligned} \frac{1}{r^n}\int_{ \frac{-r^2}{4\pi} }^0 \frac{ R_r(s)^{\alpha} }{ s^{\beta} } \, ds &= \frac{1}{r^n}\int_{ \frac{-r^2}{4\pi} }^0 \frac{ \big( -2ns \log[ r^2/ (-4\pi s ) ] \big)^{ \alpha/2} }{ -s^{\beta} } \\ &= C(n, \alpha, \beta) r^{-n+\alpha-2\beta+2} \int_{ \frac{1}{4\pi} }^0 t^{\alpha/2 - \beta} \big( \log( 4 \pi t) \big)^{\alpha/2} \,dt \\ &= C(n, \alpha, \beta) r^{-n+\alpha-2\beta+2} \int_0^{\infty} s^{\alpha/2} e^{-\alpha/2-\beta+1} \, ds \\ &= C(n, \alpha, \beta) r^{-n+\alpha-2\beta+2},\end{aligned}$$ where the last lines follows from properties of the Gamma function. We point that the integral is independent of $\alpha$ and $\beta$ when $\alpha=\beta=2$. Returning to , we have $${\partial}_x^3 u_{\tau}(x, t) \leq 4r^{-n} \iint_E {\partial}_y^3 u_{\tau}(y,s) \frac{ {\lverty\rvert}^2 }{ s^2 } \, dy ds + \frac{4r^{-n}}{2n}{\lvert {\partial}_x^3 g_{ \tau } \rvert}_{0; \, E} \iint_E \frac{ {\lverty\rvert}^4 }{ s^2 } \, dy ds. \label{eqn: heat subsol 5}$$ We estimate the first term on the right by $$\begin{aligned} 4r^{-n} \iint_E {\partial}_y^3 u_{\tau}(y,s) \frac{ {\lverty\rvert}^2 }{ s^2 } \, dy ds &\leq Cr^{-n} \int_{ \frac{ -r^2 }{4\pi} }^0 \frac{ R_r(s)^2 }{s^2} \left( \int_{B_{R_r(s)}} {\partial}_y^3 u_{\tau} \, dy \right) ds \\ &\leq Cr^{-n} \int_{ \frac{ -r^2 }{4\pi} }^0 \frac{ R_r(s)^2 }{s^2} \left( \int_{{\partial}B_{R_r(s)}} \operatorname{osc}{\partial}_y^2 u \, dy \right) ds \\ &\leq Cr^{-n} \operatorname{osc}_E {\partial}_x^2 u \int_{ \frac{ -r^2 }{4\pi} }^0 \frac{ R_r(s)^{n+1} }{s^2} ds\\ &\leq \frac{ C(d) }{ r } \operatorname{osc}_E {\partial}_x^2 u.\end{aligned}$$ The second term on the right of can be estimated more simply to give $$4{\lvert {\partial}_x^3 g_{ \tau } \rvert}_{0; \, E} r^{-n} \iint_E \frac{ {\lverty\rvert}^4 }{ s^2 } \, dy ds \leq C(d) r^2 {\lvert {\partial}_x^3 g_{ \tau } \rvert}_{0; \, E}.$$ The other three estimates - follow in a similar way. For example, to derive estimate , we differentiate twice in space and once in time. Using the mean value property of subsolutions of the heat equation we obtain $${\partial}_t {\partial}_x^2 u_{\tau}(x, t) \leq 4r^{-n} \iint_E {\partial}_s {\partial}_y^2 u_{\tau}(y,s) \frac{ {\lverty\rvert}^2 }{ s^2 } \, dy ds + \frac{4r^{-n}}{2n}{\lvert {\partial}_t {\partial}_x^2 g_{ \tau } \rvert}_{0; \, E} \iint_E \frac{ {\lverty\rvert}^4 }{ s^2 } \, dy ds. \label{eqn: heat subsol 5}$$ The first part of the estimate follows by integrating by parts in time: $$\begin{aligned} 4r^{-n} \iint_E {\partial}_t{\partial}_y^2 u_{\tau}(y,s) \frac{ {\lverty\rvert}^2 }{ s^2 } \, dy ds &\leq Cr^{-n} \iint_E \operatorname{osc}{\partial}_y^2 u \frac{ {\lverty\rvert}^2 }{ s^3 } \\ &\leq \frac{ C(d) }{ r^2 } \operatorname{osc}_E {\partial}_x^2 u,\end{aligned}$$ and the second term on right can be estimate in the same way as before to give . The derivation of the parabolic Schauder estimate now continues in the the same way as for the Poisson equation, using the estimates - , the equivalence of norms lemma and the Hölder space interpolation inequality. After completing these calculations we obtain obtain the desired parabolic Schauder estimate: $$[{\partial}^{2,1} u ]_{\alpha; \, Q_R } \leq C \big( [ f ]_{\alpha; \, Q_R }+ {\lvert u \rvert}_{0; \, Q_R } \big), \label{eqn: eqn Schauder}$$ where $C$ depends on $d, \lambda, \Lambda$, and $\alpha$.
{ "pile_set_name": "ArXiv" }
--- abstract: | Transmission of information over a discrete-time memoryless Rician fading channel is considered where neither the receiver nor the transmitter knows the fading coefficients. First the structure of the capacity-achieving input signals is investigated when the input is constrained to have limited peakedness by imposing either a fourth moment or a peak constraint. When the input is subject to second and fourth moment limitations, it is shown that the capacity-achieving input amplitude distribution is discrete with a finite number of mass points in the low-power regime. A similar discrete structure for the optimal amplitude is proven over the entire range when there is only a peak power constraint. The Rician fading with phase-noise channel model, where there is phase uncertainty in the specular component, is analyzed. For this model it is shown that, with only an average power constraint, the capacity-achieving input amplitude is discrete with a finite number of levels. For the classical average power limited Rician fading channel, it is proven that the optimal input amplitude distribution has bounded support.\ *Index Terms*: Fading channels, memoryless fading, Rician fading, phase noise, peak constraints, channel capacity, capacity-achieving input. author: - Mustafa Cenk Gursoy - 'H. Vincent Poor' - Sergio Verdú date: | Dept. of Electrical Engineering\ Princeton University\ Princeton, NJ 08544 title: 'The Noncoherent Rician Fading Channel – Part I : Structure of the Capacity-Achieving Input [^1]' --- [1.5]{} [1.6]{} Introduction {#sec:intro} ============ Recently, the information theoretic analysis of fading channels is receiving much attention. This interest is motivated by the rapid advances in wireless technology and the need to use scarce resources such as bandwidth and power as efficiently as possible under severe fading conditions. Providing the ultimate performance, information theoretic measures such as capacity, spectral efficiency and error exponents can be used as benchmarks to which we can compare the performance of practical communication systems. Furthermore with the recent discovery of codes that operate very close to the Shannon capacity, information theoretic limits have gained practical relevance. Although the capacity and other information-theoretic measures of fading channels were investigated in as early as the 1960’s ([@Kennedy], [@Richters]), it is only recently that many interesting fading channel models have been considered under various practically related input and channel constraints. A significant amount of effort has been expended to study fading channel models where side information about the fading is available at either the receiver or the transmitter or both (see [@Goldsmith; @Varaiya], [@Caire; @Shamai], [@Medard; @Goldsmith], [@Medard]). However, under fast fading conditions noncoherent communications, where neither party knows the fading, often becomes the only available alternative. Richters [@Richters] considered the problem of communicating over an average power limited discrete-time memoryless Rayleigh fading channel without any channel side information. He conjectured that the capacity-achieving amplitude distribution is discrete with a finite number of mass points. Recently, Abou-Faycal *et al.* [@Abou] gave a rigorous proof of Richters’ conjecture. This result shows that when the fading is known by neither the transmitter nor the receiver, the optimal amplitude distribution has a notably different character than that of unfaded Gaussian channels. A similar discrete structure for the optimal input was also shown in [@Shamai] for the pulse amplitude modulated direct detection photon channel when average and peak power limitations are imposed on the intensity of a photon emitting source. Katz and Shamai [@Katz; @Shamai] considered the noncoherent AWGN channel and proved that the optimal input amplitude is discrete with an infinite number of mass points. Lapidoth [@Lapidoth] recently analyzed the effects of phase noise over the AWGN channel, characterizing the high- asymptotics of the channel capacity for a general class of phase noise distributions with memory. An extensive study of the capacity of multi-antenna fading channels at high was conducted in [@Lapidoth; @Moser]. Kennedy [@Kennedy] showed that the infinite bandwidth capacity of fading multipath channels is the same as that of the unfaded Gaussian channel. Although any set of orthogonal signals achieves this capacity for the unfaded Gaussian channel, orthogonal signals that are peaky both in time and frequency are needed in the presence of fading [@Gallager Sec. 8.6]. Indeed, for a general class of fading channels, Verdú [@Verdu] has recently shown that if there are no constraints other than average power, [*flash signaling*]{}, a class of unbounded peak-to-average ratio inputs defined in [@Verdu], is necessary to achieve the capacity as $\textrm{\footnotesize{SNR}} \rightarrow 0$ when the channel realization is unknown at the receiver. Flash signaling can be practically employed in systems where sudden discharge of energy (e.g., using capacitors) is allowed, thus sidestepping the use of RF amplifiers. However, these peaky signals are not feasible in communication systems subject to strict peak-to-average ratio requirements. Furthermore, in some systems CDMA-type white signals which spread their energy over the available bandwidth are used because of their anti-jamming and low probability of intercept capabilities. Hence, it is of interest to investigate the effect upon the capacity of imposing peakedness constraints, especially in the low-power regime. Médard and Gallager [@Medard; @Gallager] considered noncoherent broadband fading channels with no specular component and limited the peakedness of the input signals by imposing a fourth moment constraint. Then they showed that such a constraint forces the mutual information to zero inversely with increasing bandwidth. Since CDMA-type signals spread their energy over the available bandwidth, they satisfy the above fourth moment constraint. Therefore, Médard and Gallager conclude that CDMA-type white signals cannot efficiently utilize fading multipath channels at extremely large bandwidth. Other results on this theme are obtained by Telatar and Tse [@Telatar; @Tse] and Subramanian and Hajek [@Subramanian; @Hajek]. In this paper, we consider the noncoherent discrete-time memoryless Rician fading channel and study the capacity and the optimal input structure. In a companion paper \[Part II\], we further investigate the spectral-efficiency/bit-energy tradeoff in the low-power regime. The organization of the paper is as follows. In Section \[sec:model\], we introduce the Rician fading channel model. In Section \[sec:Rician\], we characterize the structure of the capacity achieving input distribution in the low-power regime when the channel input is constrained to have limited peakedness. In Section \[sec:pnRician\], the average power limited Rician fading with phase-noise channel, where there is phase uncertainty in the specular component, is introduced and the structure of the capacity-achieving input is investigated. Numerical results are given in Section \[sec:Numerical\], while Section \[sec:conclusion\] contains our conclusions. Channel Model {#sec:model} ============= In this paper, we consider the following discrete-time memoryless Rician fading channel model $$\begin{aligned} y_i &=& mx_i + a_ix_i + n_i \label{eq:model}\end{aligned}$$ where $\{a_i\}$ and $\{n_i\}$ are sequences of independent and identically distributed (i.i.d.) circular zero mean complex Gaussian random variables, independent of each other and of the input, with variances $E\{|a_i|^2\} = \gamma^2$ and $E\{|n_i|^2\} = N_0$, $m$ is a deterministic complex constant, $x_i$ is the complex channel input and $y_i$ is the complex channel output. $\{a_i\}$ and $\{n_i\}$ represent sequences of fading coefficients and background noise samples, respectively. The Rician fading channel model is particularly appropriate when there is a direct line-of-sight (LOS) component in addition to the faded component arising from multipath propagation. Moreover, the Rician model includes both the unfaded Gaussian channel and the Rayleigh fading channel as two special cases. Hence, results obtained for this model provide a unifying perspective. In the channel model (\[eq:model\]), fading is assumed to be flat and hence has a multiplicative effect on the channel input. This is a valid assumption if the delay spread of the channel is much smaller than the symbol duration. Moreover, frequency selective fading channels can often be decomposed into parallel non-interacting flat fading subchannels using orthogonal multicarrier techniques. Note also that the fading coefficients assume independent realizations at every symbol period. Under such fast fading conditions, reliable estimation of the fading coefficients may be quite difficult because of the short duration between independent fades. Therefore, we consider the noncoherent scenario where neither the receiver nor the transmitter knows the fading coefficients $\{a_i\}$. Channel Capacity and Optimal Input Distribution {#sec:Rician} =============================================== In this section, we elaborate on the structure of the capacity achieving input distribution for the Rician fading channel when the input has limited peakedness which is achieved by imposing a fourth moment or a peak limitation on the input amplitude. Second and Fourth Moment Limited Input -------------------------------------- We first assume that the input amplitude is subject to second and fourth moment constraints: $$\begin{aligned} E\left\{|x_i|^{2}\right\} &\leq& P_{av} \qquad \,\,\,\forall i \label{eq:constraint1} \\ E\left\{|x_i|^{4}\right\} &\leq& \kappa P_{av}^{2} \qquad \forall i \label{eq:constraint2}\end{aligned}$$ where $P_{av}$ is the average power constraint and $ 1< \kappa < \infty$. When the average power constraint is active, the fourth moment constraint is equivalent to limiting the kurtosis, $\frac{E\left\{|x_i|^4\right\}}{\left(E\left\{|x_i|^2\right\}\right)^2} \leq \kappa$, which is a measure of the peakedness of the input signal. For the Rician channel model (\[eq:model\]), the capacity is the supremum of the input-output mutual information over the set of all input distributions satisfying the constraints (\[eq:constraint1\]) and (\[eq:constraint2\]) [^2] $$\begin{gathered} \label{eq:cap} C = \sup_{\substack{F_{x}(\cdot)\\ E\left\{|x|^{2}\right\} \leq P_{av} \\ E\left\{|x|^{4}\right\}\leq \kappa P_{av}^2 }} \int_{\mathbb{C}}\int_{\mathbb{C}} f_{y|x}(y|x) \log \frac{f_{y|x}(y|x)}{f_y(y)} \, \, {\mathrm{d}}y \, \, {\mathrm{d}}F_x(x)\end{gathered}$$ where the conditional density of the output given the input, $$\begin{gathered} \label{eq:conditional} f_{y|x}(y|x) = \frac{1}{\pi(\gamma^{2}|x|^{2}+N_0)} \exp\left( -\frac{|y-mx|^{2}}{\gamma^{2}|x|^{2}+N_0} \right),\end{gathered}$$ is circular complex Gaussian. Moreover note that $f_y(y) = \int_{\mathbb{C}} f_{y|x}(y|x)\, {\mathrm{d}}F_x(x)$ is the marginal output density, $F_x$ is the distribution function of the input and $\mathbb{C}$ denotes the complex plane. First we have the following preliminary result on the optimal phase distribution. \[prop:optphase\] For the Rician fading channel (\[eq:model\]) and input constraints (\[eq:constraint1\]) and (\[eq:constraint2\]), uniformly distributed phase that is independent of the input amplitude is capacity-achieving [^3]. *Proof*: The result follows readily from the arguments in [@Lapidoth; @Moser Sec. IV.D.6] where the optimality of circularly symmetric input distributions is pointed out in a more general setting. Assume that an input random variable $x$ generates a mutual information, $I_0$. Consider a new input $x_1 = x e^{j\theta}$ where $\theta$ is independent of $x$ and uniformly distributed on $(-\pi, \pi]$. Since the conditional distribution has the property $f_{y|x}(e^{j\theta} y | e^{j\theta} x) = f_{y|x}(y|x)$ for any $\theta$, it can be easily seen that mutual information is invariant under deterministic rotations of the input distribution and hence $I(x_1;y | \theta) = I_0$. By the concavity of the mutual information over input distributions, we have $I(x_1;y) \ge I_0$. Further note that input constraints (\[eq:constraint1\]) and (\[eq:constraint2\]) are also invariant to rotation. Therefore there is no loss in optimality in considering input random variables that has uniformly distributed phase independent of the amplitude. $\square$ Note that if $|m| = 0$, we have a Rayleigh fading channel where the phase cannot be used to convey information when the channel is unknown. However, if there is a line-of-sight component, i.e., $|m|>0$, phase can indeed carry information and by Proposition \[prop:optphase\], the uniform distribution maximizes the transmission rate. With this characterization, we have reduced the optimization problem (\[eq:cap\]) to optimal selection of the distribution function of the input amplitude, $F_{|x|}(\cdot)$, under the constraints $E\{|x|^{2}\} \leq P_{av}$ and $E\{|x|^{4}\} \leq \kappa P_{av}^2$. For the sake of simplification in the notation, we introduce new random variables $R = \frac{1}{N_0}|y|^{2}$ and $r = \frac{\gamma}{\sqrt{N_0}} |x|$. Assuming a uniform input phase that is independent of the input magnitude, the mutual information in nats, after a straightforward transformation, can be expressed as follows: $$\begin{aligned} \label{eq:update2mutual} I(F_r) \stackrel{\text{def}}{=} I(x;y) = -\int_{0}^{\infty} \!\! f_R(R;F_r) \ln f_R(R;F_r) \, {\mathrm{d}}R \! - \! \int_{0}^{\infty} \ln(1 + r^{2}) \, {\mathrm{d}}F_r(r) - 1\end{aligned}$$ where $f_R(R;F_r) = \int_{0}^{\infty} g(R,r) \, {\mathrm{d}}F_r(r)$ is the density function of $R$ with the kernel given by $g(R,r) = \frac{1}{1+r^{2}} \exp \left( -\frac{R+{\sf{K}}r^{2}}{1+r^{2}} \right) I_0 \left( \frac{2\sqrt{{\sf{K}}}r \sqrt{R}}{1 + r^{2}} \right)$. Furthermore, $F_r$ is the distribution function of $r$ and ${\sf{K}} = \frac{|m|^2}{\gamma^2}$ is the Rician factor. Now, the capacity formula can be recast as follows $$\begin{gathered} \label{eq:capacity} C(\alpha, \kappa, {\sf{K}}) = \sup_{ \substack{ F_r \\ E\{r^{2}\} \leq \alpha \\ E\{r^4\} \leq \kappa \alpha^2}} I(F_r)\end{gathered}$$ depending on three parameters, namely, $\alpha = \gamma^2 \frac{P_{av}}{N_0}$, the normalized ; $\kappa$; and ${\sf{K}}$, the Rician factor. In [@gursoy], the existence of an optimal amplitude distribution achieving the supremum in (\[eq:capacity\]) is shown and the following sufficient and necessary condition for an amplitude distribution to be optimal is derived employing the techniques used in [@Abou]. () \[prop:ktc\] For the Rician channel (\[eq:model\]) with input constraints (\[eq:constraint1\]) and (\[eq:constraint2\]), $F_{0}$ is a capacity achieving amplitude distribution if and only if there exist $\lambda_1, \lambda_2 \geq 0$ such that the following is satisfied $$\begin{aligned} \int_0^{\infty}g(R,r) \ln f_R(R,F_0) \, {\mathrm{d}}R + \ln(1+r^2)+ \lambda_{1}(r^{2}-\alpha)+\lambda_{2}(r^{4}-\kappa\alpha^{2})+ C+1 \geq 0 \quad \forall r \geq 0 \label{eq:ktc}\end{aligned}$$ with equality if $r \in E_0$ where $E_0$ is the set of points of increase[^4] of $F_0$. Note that in the above formulation, $\lambda_1$ and $\lambda_2$ are the Lagrange multipliers for the second and fourth moment constraints respectively. Using Proposition \[prop:ktc\], we have the following result on the optimal amplitude distribution. \[theo:Rician\] For the Rician fading channel (\[eq:model\]) with input amplitude constraints (\[eq:constraint1\]) and (\[eq:constraint2\]), if the fourth moment constraint (\[eq:constraint2\]) is active then the capacity-achieving input amplitude distribution is discrete with a finite number of mass points. *Proof:* The result is shown by contradiction. The proof can be summarized as follows: i\) We first contradict the assumption that the optimal distribution has an infinite number of points of increase on a bounded interval. ii\) Next, we contradict the assumption that the optimal distribution has an infinite number of points of increase (mass points) but only finitely many of them on any bounded interval. iii\) Ruling out the above assumptions leaves us with the only possibility that the optimal input has a finite number of mass points. Assume $F_0$ is an optimal amplitude distribution. To prove the theorem, we first find a lower bound on the left hand side (LHS) of the Kuhn-Tucker condition (\[eq:ktc\]). To that end, we first bound $f_R$ as follows, $$\begin{aligned} f_R(R;F_0) &= \int_{0}^{\infty} g(R,r) \, {\mathrm{d}}F_0(r) \nonumber \\ &\geq \int_{0}^{\infty} \frac{1}{1+r^{2}} \exp \left( -\frac{R+{\sf{K}}r^{2}}{1+r^{2}} \right) \, {\mathrm{d}}F_0(r) \\ &\geq \exp(-R)\int_{0}^{\infty} \frac{1}{1+r^{2}} \exp \left( -\frac{{\sf{K}}r^{2}}{1+r^{2}} \right) \, {\mathrm{d}}F_0(r) \\ &= D_{F_0}\exp(-R) \qquad \forall R \geq 0 \label{eq:lowerboundRician}\end{aligned}$$ where $0<D_{F_0}\leq 1$ is a constant depending on $F_0$. The first inequality is obtained from the fact that $I_0(x) \geq 1 \,\,\, \forall x \geq 0$ and the second inequality comes from observing that $e^{-R/(1+r^2)} \geq e^{-R} \,\,\, \forall R,r \geq 0$. Using the lower bound (\[eq:lowerboundRician\]), and noting that $g(R,r)$ is a noncentral chi-square density function in $R$, we have the following bound on the LHS of the Kuhn-Tucker condition (\[eq:ktc\]) $$\begin{aligned} \text{LHS} \geq \ln D_{F_0} - 1 -(1+{\sf{K}})r^2 + \ln(1+r^2)+\lambda_{1}(r^{2}-\alpha)+\lambda_{2}(r^{4}-\kappa\alpha^{2})+1+C \,\,\, \forall r\geq 0. \nonumber\end{aligned}$$ If the fourth moment constraint is active, i.e., $\lambda_2 > 0$, then for all $D_{F_0}>0$ and $\lambda_1 \ge 0$, the above lower bound diverges to infinity as $r \rightarrow \infty$. Now, we establish contradiction in the following assumptions.\ i) Assume that the optimal distribution $F_0$ has an infinite number of points of increase on a bounded interval. Note that this assumption is satisfied by continuous distributions. First we extend the LHS of the Kuhn-Tucker condition (\[eq:ktc\]) to the complex domain. $$\begin{gathered} \label{eq:ktccomplex} \Phi(z) = \int_0^{\infty} g(R,z) \ln f_R(R,F_0) \, {\mathrm{d}}R +\ln(1+z^2)+ \lambda_{1}(z^{2}-\alpha)+\lambda_{2}(z^{4}-\kappa\alpha^{2})+ C+1\end{gathered}$$ $\!\!$where $z \in \mathbb{C}$. By the Differentiation Lemma [@Lang Ch. 12], it is easy to see that $\Phi$ is analytic in the region where $\text{Re}\{1+z^2\}>0$ [@gursoy Appendix C]. This choice of region guarantees the uniform convergence of the integral expression in (\[eq:ktccomplex\]). Note that in this region, by our earlier assumption, $\Phi(z)= 0$ for an infinite number of points having a limit point[^5]. By the Identity Theorem[^6], $\Phi(z) = 0$ in the region where $\text{Re}\{1+z^2\}>0$. Hence the Kuhn-Tucker condition (\[eq:ktc\]) is satisfied with equality for all $r\geq 0$. Clearly, this case is not possible from the above lower bound which diverges to infinity as $r \rightarrow \infty$.\ ii) Next assume that the optimal distribution has an infinite number of mass points but only finitely many of them on any bounded interval. Then the LHS of the Kuhn-Tucker condition should be equal to zero infinitely often as $r \rightarrow \infty$ which is again not possible by the above diverging lower bound.\ Hence the optimal distribution must be discrete with a finite number of mass points and the theorem follows. $\square$ The significance of Theorem \[theo:Rician\] comes from the fact that for the Rician fading channel, any fourth moment constraint with a finite $\kappa$ will eventually be active for sufficiently small because, as observed in [@Verdu Sec. V.E], if there is no such constraint, the required value of $\kappa$ grows without bound as ${\text{\footnotesize{SNR}}}\rightarrow 0$. Therefore, Theorem \[theo:Rician\] establishes the discrete nature of the optimal input in the low-power regime. Furthermore, Theorem \[theo:Rician\] easily specializes to the Rayleigh and the unfaded Gaussian channels. For the unfaded Gaussian channel, if the fourth moment constraint is inactive, it is well known that a Rayleigh distributed amplitude is optimal and it has kurtosis $\kappa = 2$. Therefore the fourth moment constraint being active (i.e., $1 < \kappa < 2$) is also a necessary condition for the discrete nature in that case. In [@Abou], the optimal amplitude for the average-power-limited noncoherent Rayleigh fading channel is shown to be discrete with a finite number of levels over the entire range. Theorem \[theo:Rician\] proves that this discrete character does not change when we have an additional fourth moment constraint. We note that the key property which leads to the proof of Theorem \[theo:Rician\] is that we have a moment constraint higher than the second moment. Hence the result of Theorem \[theo:Rician\] holds in a more general setting where the fourth moment constraint (\[eq:constraint2\]) is replaced by a constraint in the following form: $E\{|x|^{2+\delta}\} \leq M$ for some $\delta > 0$ and $M < \infty$. In a related work, Palanki [@Palanki] has independently shown the discrete character of the optimal input for a general type of fading channels when only moment constraints strictly higher than the second moment are imposed. Peak Power Limited Input ------------------------ In this section, we assume that the input amplitude is subject to only a peak power limitation, $$\begin{gathered} \label{eq:peakconstraint} |x_i|^2 \stackrel{\text{a.s.}}{\le} P \quad \forall i.\end{gathered}$$ Although being more stringent than the fourth moment limitation, peak power constraint is more relevant in practical systems. For instance, efficient use of battery power in portable radio units and linear operation of RF amplifiers employed at the transmitter require peak power limited communication schemes. Since the peak constraint (\[eq:peakconstraint\]) is invariant to rotation of the input, optimality of uniform phase follows from Proposition \[prop:optphase\]. Hence, our primary focus is on obtaining a characterization for the optimal amplitude distribution. Existence of a capacity-achieving amplitude distribution readily follows from the results on the second and fourth moment limited case. Similarly, we define $R = \frac{1}{N_0}|y|^2$ and $r = \frac{\gamma}{\sqrt{N_0}}|x|$. Specializing (\[eq:ktc\]), we easily obtain the following sufficient and necessary condition for an amplitude distribution, $F_0$, to be optimal over the peak-power limited Rician channel: $$\begin{gathered} \label{eq:ktcpeak} \int_0^{\infty}g(R,r) \ln f_R(R,F_0) \, {\mathrm{d}}R + \ln(1+r^2)+ C+1 \geq 0 \quad \forall r \in \left[0, \sqrt{\alpha} \right]\end{gathered}$$ with equality if $r \in E_0$ where $E_0$ is the set of points of increase of $F_0$. Note that $\alpha = \gamma^2 \frac{P}{N_0}$. Next, we state the main result on the optimal amplitude distribution. \[theo:peakRician\] For the Rician fading channel (\[eq:model\]) where the input is subject to only a peak power constraint $|x|^2 \stackrel{\text{a.s.}}{\le} P$, the capacity-achieving amplitude distribution is discrete with a finite number of mass points. *Proof*: Since the input is subject to a peak constraint, the result in this case is established by contradicting the assumption that the optimal input distribution has an infinite number of points of increase on a bounded interval. Assume $F_0$ is an optimal distribution. To prove the theorem, we first find an upper bound on the left-hand-side (LHS) of (\[eq:ktcpeak\]). To achieve this goal, we bound $f_R(\cdot,F_0)$ as follows. $$\begin{aligned} f_R(R;F_0) &= \int_{0}^{\sqrt{\alpha}} \frac{1}{1+r^{2}} \exp \left( -\frac{R+{\sf{K}}r^{2}}{1+r^{2}} \right) I_0 \left( \frac{2\sqrt{{\sf{K}}}r \sqrt{R}}{1 + r^{2}} \right) \, {\mathrm{d}}F_0(r) \nonumber \\ &\le \exp\left( -\frac{R}{1+\alpha} + \sqrt{{\sf{K}}R} \right) \int_{0}^{\sqrt{\alpha}} \frac{1}{1+r^{2}} \exp \left( -\frac{{\sf{K}}r^{2}}{1+r^{2}} \right) \, {\mathrm{d}}F_0(r) \label{eq:f_Rupperbound} \\ &= D_{F_0}\exp\left( -\frac{R}{1+\alpha} + \sqrt{{\sf{K}}R} \right) \label{eq:f_Rupperbound2}\end{aligned}$$ where $0< D_{F_0} \le 1$ is a constant depending on $F_0$. Upper bound (\[eq:f\_Rupperbound\]) is easily verified by observing $ \exp\left(- \frac{R}{1+r^2} \right) \le \exp\left(- \frac{R}{1+\alpha} \right) \,\,\, \forall r \le \sqrt{\alpha}$ and $I_0 \left( \frac{2\sqrt{{\sf{K}}}r \sqrt{R}}{1 + r^{2}} \right) \le I_0(\sqrt{{\sf{K}}R}) \le \exp(\sqrt{{\sf{K}}R}) \,\,\, \forall r \ge 0. $ Using (\[eq:f\_Rupperbound2\]), we have the following upper bound: $$\begin{aligned} \int_0^{\infty} g(R,r) \ln f_R(R;F_0) \, {\mathrm{d}}R &\le \ln D_{F_0} - \frac{1}{1+\alpha}\int_0^{\infty} g(R,r) R \,\, {\mathrm{d}}R + \sqrt{K} \int_0^{\infty} g(R,r) \sqrt{R} \,\, {\mathrm{d}}R \\ &\le \ln D_{F_0} - \frac{1+({\sf{K}}+1)r^2}{1+\alpha} + \sqrt{\sf{K}} \sqrt{1 + (1+{\sf{K}})r^2} \quad \forall r \ge 0. \label{eq:intupperbound}\end{aligned}$$ The upper bound in (\[eq:intupperbound\]) follows from the fact that $g(R,r)$ is a non-central chi-square probability density function in $R$, and $ \int_0^{\infty} g(R,r) R \,\, {\mathrm{d}}R = 1 + (1+{\sf{K}})r^2$ and $\int_0^{\infty} g(R,r) \sqrt{R} \,\, {\mathrm{d}}R \le \sqrt{1 + (1+{\sf{K}})r^2}$ which follows from the concavity of $\sqrt{x}$ and the Jensen’s inequality. From (\[eq:intupperbound\]), we obtain the following upper bound on the left hand side (LHS) of (\[eq:ktcpeak\]): $$\begin{aligned} \text{LHS} \le \ln D_{F_0} - \frac{1+({\sf{K}}+1)r^2}{1+\alpha} + \sqrt{\sf{K}} \sqrt{1 + (1+{\sf{K}})r^2} + \ln(1+r^2)+ C+1 \quad \forall r \ge 0. \label{eq:LHSupperbound}\end{aligned}$$ Using the above upper bound, we show that the following assumption cannot hold true.\ i) Assume that the optimal input distribution $F_0$ has an infinite number of points of increase on a bounded interval. Next we extend the LHS of (\[eq:ktcpeak\]) to the complex domain: $$\begin{gathered} \Phi(z) = \int_0^{\infty}g(R,z) \ln f_R(R,F_0) \, {\mathrm{d}}R + \ln(1+z^2)+ C+1\end{gathered}$$ where $z \in \mathbb{C}$. Since the condition in (\[eq:ktcpeak\]) should be satisfied with equality at the points of increase of the optimal input distribution, by the above assumption, $\Phi(z) = 0$ for an infinite number of points having a limit point. Then by the Identity Theorem [@Knopp], $\Phi(z) = 0$ in the whole region where it is analytic. By the Differentiation Lemma , one can easily verify that $\Phi(z)$ is analytic in the region where $\text{Re}(1+z^2) > 0$ which includes the positive real line. Therefore we conclude that $\Phi(r) = 0 \,\,\, \forall r \ge 0$. Clearly, this is not possible from the upper bound in (\[eq:LHSupperbound\]) which diverges to $-\infty$ as $r \to \infty$ for any finite $\alpha, {\sf{K}} \ge 0$, and $D_{F_0}>0$.\ Reaching a contradiction, we conclude that the optimal distribution must be discrete with a finite number of mass points. $\square$ We note that Theorem \[theo:peakRician\] establishes the discrete structure of the optimal input distribution over the entire range. The proof basically uses the observation that a bounded input induces an output probability density function that decays at least exponentially, and in turn provides a diverging bound on the Kuhn-Tucker condition. We also note recent independent work by Huang and Meyn [@Meyn], where the discrete nature of the optimal input is proven by again showing a diverging bound on the Kuhn-Tucker condition for a general class of channels in which the input is subject only to peak amplitude constraints. Rician Fading Channel with Phase Noise {#sec:pnRician} ====================================== In this section, we deviate from the classical Rician channel model (\[eq:model\]) where the specular component is assumed to be static and consider the following model $$\begin{gathered} \label{eq:pnmodel} y_i = (a_i + me^{j\,\theta_i})x_i + n_i.\end{gathered}$$ where phase noise is introduced in the specular component. Here, {$\theta_i$} is assumed to be a sequence of independent and identically distributed uniform random variables on $[-\pi,\pi)$ and $m$ is a deterministic complex constant. We again consider the noncoherent scenario where $\{a_i\}$ and $\{\theta_i\}$ are known by neither the receiver nor the transmitter. This model is relevant in mobile systems where rapid random changes in the phase of the specular component are not tracked. Moreover, such a model is suitable in cases where there is imperfect receiver side information about the fading magnitude. As another departure from the previous section, here we impose only an average power constraint, $E\{|x|^2\} \leq P_{av}$. The discrete nature of the optimal input amplitude follows immediately from the techniques of Section \[sec:Rician\] when there is an additional higher moment constraint. We immediately realize that the channel output, $y$, is conditionally Gaussian given $x$ and $\theta$, $$\begin{gathered} \label{eq:pncondgauss} f_{y|x,\,\theta}(y|x,\theta) = \frac{1}{\pi(\gamma^2|x|^2+N_0)} \exp\left(-\frac{\left| y - me^{j\theta}x\right|^2}{\gamma^2|x|^2+N_0} \right).\end{gathered}$$ Integrating (\[eq:pncondgauss\]) over uniform $\theta$, we obtain the conditional distribution of the channel output given the input, $$\begin{gathered} \label{eq:pncond} f_{y|x}(y|x) = \frac{1}{\pi(\gamma^2|x|^2+N_0)} \exp\left(-\frac{|y|^2+|m|^2|x|^2}{\gamma^2|x|^2+N_0} \right) I_0\left(\frac{2|m||y||x|}{\gamma^2|x|^2+N_0} \right).\end{gathered}$$ We again introduce the following random variables: $R = \frac{1}{N_0}|y|^2 \,\,,\,\, r = \frac{\gamma}{\sqrt{N_0}}|x|.$ Since the phase information is completely destroyed in the channel (\[eq:pnmodel\]) and the above transformations are one-to-one, we have $ I(x;y) = I(|x|;|y|) = I(r;R). $ Furthermore the conditional distribution of $R$ given $r$ is easily obtained from (\[eq:pncond\]): $$\begin{gathered} f_{R|r}(R|r) = \frac{1}{1+r^2}\, \exp\left(- \frac{R+{\sf{K}}r^2}{1+r^2}\right) I_0\left( \frac{2\sqrt{{\sf{K}}}r\sqrt{R}}{1+r^2}\right)\end{gathered}$$ where ${\sf{K}} = \frac{|m|^2}{\gamma^2}$ is the Rician factor. Similarly as in the previous section, the existence of an optimal amplitude distribution is shown and the following sufficient and necessary condition is given in [@gursoy]. () \[prop:pnktc\] For the Rician channel model (\[eq:pnmodel\]) with an average power constraint $E\{|x|^2\} \leq P_{av}$, $F_{0}$ is a capacity-achieving amplitude distribution if and only if there exists $\lambda \geq 0$ such that the following is satisfied $$\begin{aligned} -D(f_{R|r}||f_R) + \lambda(r^2-\alpha) + C &\geq 0 \quad \forall r \geq 0 \label{eq:pnktc} $$ with equality if $r \in E_0$ where $E_0$ is the set of points of increase of $F_0$. In the above formulation, $D(\cdot||\cdot)$ is the divergence (e.g. [@Cover Section 2.3]), $f_R(R;F_0) = \int_0^\infty f_{R|r}(R|r) \,{\mathrm{d}}F_0(r)$ is the density function of $R$, $\alpha = \gamma^2\text{\footnotesize{SNR}}$, and $C$ is the capacity. The next theorem gives the main result on the structure of the optimal input for the Rician fading channel with phase noise. \[theo:pnRician\] For the Rician fading channel with uniform phase noise (\[eq:pnmodel\]) and average power constraint $E\{|x|^2\} \leq P$, the capacity-achieving input amplitude distribution is discrete with a finite number of mass points. *Proof*: The result is shown by contradiction. Let $F_0$ be an optimal amplitude distribution. The proof can be summarized as follows: i\) We first assume that $F_0$ has an infinite number of points of increase on a bounded interval. The impossibility of this case is shown by contravening the fact that under this assumption the left hand side of the Kuhn-Tucker condition which is extended to the complex domain is identically zero over its region of analyticity. ii\) Then we assume that the optimal distribution is discrete with an infinite number of mass points but only finitely many of them on any bounded interval. This assumption is also ruled out by finding a diverging lower bound on the left hand side of the Kuhn-Tucker condition. iii\) Having eliminated the above assumptions, we are left with the only possibility that the optimal distribution is discrete with a finite number of mass points. **Assumption 1**: Assume that $F_0$ has an infinite number of points of increase on a bounded interval. Then the Kuhn-Tucker condition (\[eq:pnktc\]) is satisfied with equality at an infinite number of points having a limit point. First we extend the left hand side of (\[eq:pnktc\]) to the complex domain: $ \Psi(z) = -D(f_{R|r=z}||f_R) + \lambda(z^2-\alpha) + C,$ where $z \in \mathbb{C}$. Equivalently, using the fact that $f_{R|r}$ is a noncentral chi-square density function in $R$, $$\begin{aligned} \label{eq:KTCcomplex} \Psi(z) = \int_0^\infty f_{R|r}(R|z) \ln f_R(R) \,{\mathrm{d}}R -& \int_0^\infty f_{R|r}(R|z) \ln\left(I_0\left(\frac{2\sqrt{{\sf{K}}}z\sqrt{R}}{1+z^2}\right)\right) {\mathrm{d}}R \nonumber \\ +&\lambda(z^2-\alpha) + \ln(1+z^2) + \frac{2{\sf{K}}z^2}{1+z^2} + C + 1.\end{aligned}$$ By the Differentiation Lemma [@Lang Ch. 12], it is easy to see that $\Psi$ is analytic in the region where $\text{Re}\left\{1+z^2\right\} > 0$ and $\text{Re}\{z\}>0$. The first condition guarantees the uniform convergence of the integrals in (\[eq:KTCcomplex\]) by forcing the integrands to decrease exponentially. Since $I_0$ has zeros on the imaginary axis and the second integral in (\[eq:KTCcomplex\]) involves the logarithm of the Bessel function, with the second condition, $\text{Re}\{z\}>0$, we exclude the imaginary axis from the region of analyticity. Since $\Psi(z) = 0$ for an infinite number of points having a limit point, by the Identity Theorem [@Knopp], $\Psi(z) = 0$ in the whole region where it is analytic. In particular, we have $\Psi(jb+\frac{1}{n}) = 0$ for all $|b|<1$ and $n \in \mathbb{Z}^+$, and hence $ \lim_{n \to \infty} \Psi(jb + \frac{1}{n}) = 0 \quad \forall |b|<1. $ All the terms other than the second term in (\[eq:KTCcomplex\]) are analytic also on the imaginary axis with $|z|<1$ and the limiting expression is obtained by letting $\frac{1}{n} \to 0$ in the arguments of these functions. For the second term in (\[eq:KTCcomplex\]), we need to invoke the Dominated Convergence Theorem [@Rudin] to justify the interchange of limit and integral. An integrable upper bound on the magnitude of the integrand of the second integral is shown in [@gursoy Appendix D]. Therefore, we have $$\begin{aligned} \lim_{n \to \infty} \Psi(jb+\frac{1}{n}) = &\int_0^\infty f_{R|r}(R|jb) \ln f_R(R) \,{\mathrm{d}}R \nonumber \\ -&\int_0^\infty \lim_{n \to \infty} f_{R|r}\left(R\bigg|jb + \frac{1}{n}\right) \ln\left(I_0\left(\frac{2\sqrt{{\sf{K}}}\left(jb + \frac{1}{n}\right)\sqrt{R}}{1+\left(jb + \frac{1}{n}\right)^2}\right)\right) {\mathrm{d}}R \nonumber \\ +&\lambda(-b^2-\alpha) + \ln(1-b^2) - \frac{2{\sf{K}}b^2}{1-b^2} + C + 1 = 0 \quad \forall |b| < 1. \label{eq:pnktcimag}\end{aligned}$$ Note that all the terms other than the second term in (\[eq:pnktcimag\]) are real. Next we show that the second term in (\[eq:pnktcimag\]) has a nonzero imaginary component yielding a contradiction. First we evaluate the limit in the integrand as follows. $$\begin{aligned} &\lim_{n \to \infty} f_{R|r}\left(R\bigg|jb + \frac{1}{n}\right) \ln\left(I_0\left(\frac{2\sqrt{{\sf{K}}}\left(jb + \frac{1}{n}\right)\sqrt{R}}{1+\left(jb + \frac{1}{n}\right)^2}\right)\right) = \nonumber \\ & = \left\{ \begin{array}{ll} 0 & \text{if } I_0\left( \frac{2\sqrt{{\sf{K}}}jb\sqrt{R}}{1-b^2} \right) = 0 \\ \frac{1}{1-b^2} \exp \left(-\frac{R - {\sf{K}}b^2}{1-b^2}\right) I_0\left( \frac{2\sqrt{{\sf{K}}}jb\sqrt{R}}{1-b^2} \right) \ln\left(I_0\left( \frac{2\sqrt{{\sf{K}}}jb\sqrt{R}}{1-b^2} \right)\right) & \text{otherwise}, \end{array} \right. \label{eq:limitingintegrand}\end{aligned}$$ which is obtained easily by observing that all the terms are analytic in the entire complex plane (excluding $z = \pm j$) except the logarithm function which is not analytic at the zeros of the Bessel function. However, as we approach the zeros of the Bessel function, $I_0 \ln I_0 \to 0$ and hence we obtain (\[eq:limitingintegrand\]). Noting that $J_0(z) = I_0(jz)$ where $J_0$ is the zeroth order Bessel function of the first kind, and that the set of zeros of the $I_0$ function along the imaginary axis has measure zero, the second integral in (\[eq:pnktcimag\]) can now be expressed as follows: $$\begin{aligned} \int_0^{\infty} \frac{1}{1-b^2} \exp \left(-\frac{R - {\sf{K}}b^2}{1-b^2}\right) J_0\left( \frac{2\sqrt{{\sf{K}}}b\sqrt{R}}{1-b^2} \right) \ln\left(J_0\left( \frac{2\sqrt{{\sf{K}}}b\sqrt{R}}{1-b^2} \right)\right) {\mathrm{d}}R.\end{aligned}$$ By applying a change of variables $v = \frac{2\sqrt{{\sf{K}}}b\sqrt{R}}{1-b^2}$ and expressing $\ln z = \ln |z| + j \,\text{arg}(z)$, the above integral becomes $$\begin{aligned} \int_0^\infty &\frac{1-b^2}{2{\sf{K}}b^2} \, v \exp\left(-\frac{1-b^2}{4{\sf{K}}b^2}v^2 + \frac{{\sf{K}}b^2}{1-b^2}\right) J_0(v) \ln |J_0(v)| \, {\mathrm{d}}v \nonumber \\ &+ j \pi \int_0^\infty \frac{1-b^2}{2{\sf{K}}b^2} \, v \exp\left(-\frac{1-b^2}{4{\sf{K}}b^2}v^2 + \frac{{\sf{K}}b^2}{1-b^2}\right) J_0(v) h(v) \, {\mathrm{d}}v. \label{eq:pnktcimagterm}\end{aligned}$$ In the above formulation $h(v) = 0$ if $v \in (0, \alpha_1)$ and $h(v) = k$ if $v \in (\alpha_k,\alpha_{k+1})$ where $\{\alpha_k\}$ are the zeros of $J_0$. As noted in [@Nuriyev], the second term in (\[eq:pnktcimagterm\]) arises due to the fact that the logarithm jumps in value by $j \pi$ when a zero is passed. This fact can be observed by considering the argument of $J_0(b - j\epsilon)$ as $\epsilon \to 0$ where $b-j\epsilon$ is in a small neighborhood of $\alpha_k$ such that $J_0(b - j\epsilon) \backsimeq J_0^{\,\,'}(\alpha_k)(b - j\epsilon - \alpha_k)$. Using a similar bound obtained in [@Nuriyev], we have $$\begin{aligned} &\left|\int_0^\infty \frac{1-b^2}{2{\sf{K}}b^2} \, v \exp\left(-\frac{1-b^2}{4{\sf{K}}b^2}v^2 \right) J_0(v) h(v) \, {\mathrm{d}}v \right| \ge \nonumber \\ & \ge \left(|J_0(\beta)| - \frac{2 {\sf{K}} b^2}{\beta (1-b^2)}\right)\exp\left(- \frac{\beta^2(1-b^2)}{4{\sf{K}}b^2} \right) - \left( 1 + \frac{4 {\sf{K}} b^2 }{\pi \alpha_2 (1-b^2)}\right) \exp\left(- \frac{\alpha_2^2(1-b^2)}{4 {\sf{K}} b^2} \right)\end{aligned}$$ $\!\!\!$where $\alpha_2$ is the second smallest zero of $J_0$ on the positive axis and $\beta$ is the positive zero of $J_1$ less than $\alpha_2$. As noted in [@Nuriyev], the above lower bound is positive for small enough values of $b$. In particular, for each ${\sf{K}} > 0$, the above lower bound is $9.16 \times 10^{-5}$ when $b^2 = \frac{1}{2{\sf{K}}+1}<1$. **Assumption 2**: Next we assume that the optimal distribution has an infinite number of mass points but only finitely many of them on any bounded interval. Following an approach similar to the one used in [@Abou], we first bound $f_R$ as follows $$\begin{aligned} f_R(R;F_0) = \int_{0}^{\infty} f_{R|r}(R|r) \, {\mathrm{d}}F_0(r) &= \sum_{i = 0}^{\infty} p_i \, f_{R|r}(R|r_i) \\ &\geq p_i \, f_{R|r}(R|r_i) \\ &\geq p_i \, \frac{1}{1+r_i^2} \exp\left(-\frac{R+{\sf{K}}r_i^2}{1+r_i^2}\right) \qquad \forall i \,\,\, \forall R \geq 0.\end{aligned}$$ where $p_i$ and $r_i$ are the probability and location, respectively, of the $i^{th}$ mass point of $F_0$. We obtain the last inequality by using the fact that $I_0(x) \geq 1 \,\,\, \forall x \geq 0$. Noting that $f_{R|r}(R|r)$ is a noncentral chi square density function in $R$, this bound leads to the following lower bound on the left hand side of (\[eq:pnktc\]) $$\begin{aligned} \label{eq:lowerboundonKTC} \text{LHS} \geq &\ln p_i - \ln (1+r_i^2)- \frac{1+{\sf{K}}r_i^2}{1+r_i^2} - \frac{({\sf{K}}+1)r^2}{1+r_i^2} - \int_0^\infty f_{R|r}(R|r) \ln\left(I_0\left(\frac{2\sqrt{{\sf{K}}}r\sqrt{R}}{1+r^2}\right)\right) {\mathrm{d}}R \nonumber \\ &+\lambda(r^2-\alpha) + \ln(1+r^2) + \frac{2{\sf{K}}r^2}{1+r^2}+C+1\,\,,\,\, \forall i \,\, \forall r \geq 0.\end{aligned}$$ Noting that $ I_0\left(\frac{2\sqrt{{\sf{K}}}r\sqrt{R}}{1+r^2}\right) \leq \exp\left(\frac{2\sqrt{{\sf{K}}}r\sqrt{R}}{1+r^2} \right), $ we have $$\begin{aligned} \int_0^\infty f_{R|r}(R|r) \ln\left(I_0\left(\frac{2\sqrt{{\sf{K}}}r\sqrt{R}}{1+r^2}\right)\right) {\mathrm{d}}R &\leq \frac{2\sqrt{{\sf{K}}}r}{1+r^2} E\{\sqrt{R}|r\} \\ &\le \frac{2\sqrt{{\sf{K}}}r}{1+r^2} \sqrt{E\{R|r\}} \\ & = \frac{2\sqrt{{\sf{K}}}r \sqrt{1 + (1 + {{\sf{K}}})r^2}}{1+r^2} \le 2 \sqrt{2{{\sf{K}}}+{{\sf{K}}}^2}.\end{aligned}$$ Since we have assumed that the optimal distribution has an infinite number of mass points with finitely many of them on any bounded interval, we see that for any $\lambda > 0$, we can choose an $r_i$ sufficiently large such that the lower bound (\[eq:lowerboundonKTC\]) diverges to $\infty$ as $r \rightarrow \infty$. But by our assumption the left hand side of (\[eq:lowerboundonKTC\]) should be equal to zero infinitely often as $r \rightarrow \infty$, which is a contradiction. $\lambda = 0$ implies that the power constraint is ineffective. The impossibility of this case is shown in [@Abou]. Therefore the theorem follows. $\square$ Recent results [@Abou] and [@Katz; @Shamai] have shown the discrete nature of the optimal distribution for the two special cases of the model (\[eq:pnmodel\]): the Rayleigh fading channel and the noncoherent AWGN channel. We have proven the discreteness of the capacity-achieving distribution in a unifying setting where there is both multipath fading and a specular component with random phase. For the noncoherent AWGN channel, Katz and Shamai [@Katz; @Shamai] have shown that the optimal input has an infinite number of mass points. An interesting conclusion of Theorem \[theo:pnRician\] is that the presence of an unknown multipath component induces an optimal distribution with a finite number mass points. It is also of interest to consider the classical average-power-limited Rician fading channel (\[eq:model\]) for which tight upper and lower bounds on the capacity were derived in [@Lapidoth; @Moser]. By Proposition \[prop:optphase\], we know that uniform phase is optimal for this model. Moreover, we have the following partial result on the optimal amplitude distribution which proves the suboptimality of input amplitude distributions with unbounded support such as the Rayleigh distribution. \[theo:Rician2\] For the Rician fading channel (\[eq:model\]) with only an average power limitation $E\{|x|^2\} \leq P_{av}$, the optimal input amplitude distribution has bounded support. *Proof*: Assume $F_0$ is an optimal distribution. We will prove the proposition by contradiction. So we further assume that $F_0$ has unbounded support. With this assumption, for any finite $M \geq 0$, $$\begin{aligned} f_R(R;F_0) &= \int_{0}^{\infty} g(R,r) \, {\mathrm{d}}F_0(r) \nonumber \\ &\geq \int_{0}^{\infty} \frac{1}{1+r^{2}} \exp \left( -\frac{R+Kr^{2}}{1+r^{2}} \right) \, {\mathrm{d}}F_0(r) \\ &\geq \int_{M}^{\infty} \frac{1}{1+r^{2}} \exp \left( -\frac{R+Kr^{2}}{1+r^{2}} \right) \, {\mathrm{d}}F_0(r) \\ &\geq \exp(-\frac{R}{1+M^2}) \int_{M}^{\infty} \frac{1}{1+r^{2}} \exp \left( -\frac{Kr^{2}}{1+r^{2}} \right) \, {\mathrm{d}}F_0(r) \\ &= D_{F_0,M}\exp(-\frac{R}{1+M^2}) \quad , \quad \forall R \geq 0 \,\,\, \forall M \geq 0 \,\, , \label{eq:lowerboundRician2}\end{aligned}$$ where $0 < D_{F_0,M} \leq 1 \quad \forall M \geq 0$ and $ \forall F_0 $. Using (\[eq:lowerboundRician2\]), we obtain the following lower bound on the left hand side of the Kuhn-Tucker condition[^7] $$\begin{gathered} \label{eq:lowerboundRician2int} \text{LHS} \geq \ln D_{F_0,M} - \frac{1}{1+M^2} -\frac{1+K}{1+M^2}r^2 +\ln(1+r^2)+\lambda(r^{2}-\alpha)+1+C \,\,\, , \,\, \forall r\geq 0 \,\,\, \forall M \geq 0. \nonumber\end{gathered}$$ For any $\lambda > 0$[^8] and $D_{F_0,M} > 0$, we can choose $M$ sufficiently large such that the above lower bound diverges to infinity as $r \rightarrow \infty$. However, if the optimal input has unbounded support, the LHS of the Kuhn-Tucker condition should be zero infinitely often as $r \rightarrow \infty$. This constitutes a contradiction and hence the theorem follows. $\square$ Numerical Results {#sec:Numerical} ================= In general, the number of mass points of the optimal discrete distribution and their locations and probabilities depend on the . Analytical expressions for the capacity and the optimal distribution as a function of are unlikely to be feasible. Therefore, we resort to numerical methods to examine this behavior. The numerical algorithm used here is similar to the ones employed in [@Smith] and [@Abou]. In particular, we start with a sufficiently small and maximize the mutual information over the set of two-mass-point discrete distributions satisfying the input constraints. Then we test the maximizing two-mass-point discrete distribution with the Kuhn-Tucker condition. If this distribution satisfies the necessary and sufficient Kuhn-Tucker condition, then it is optimal and the mutual information achieved by it is the capacity. As we increase the , the required number of mass points monotonically increases, and therefore to obtain the optimum distribution we repeat the same procedure for discrete distributions with increasing numbers of mass-points.[^9] For the Rician fading channel (${{\sf{K}}}> 0$) with second and fourth moment input constraints, numerical results indicate that for sufficiently small values, the two-mass-point discrete amplitude distribution $$\begin{gathered} \label{eq:optinput} F(|x|) = \left(1-\frac{1}{\kappa} \right) u(|x|) + \frac{1}{\kappa} u(|x|-\sqrt{\kappa N_0 {\text{\scriptsize{SNR}}}})\end{gathered}$$ is optimal. Note that this distribution does not depend on the Rician factor ${\sf{K}}$. Figure \[fig:KTbeta1\] plots the left hand side of the Kuhn-Tucker condition (\[eq:ktc\]) as a function of $r$ for the distribution $F(r) = 0.9 u(r) + 0.1 u(r - 1/\sqrt{2})$ for the Rician fading channel (${\sf{K}} = 1$) with $\alpha = 0.05$ and $\kappa = 10$. From the figure we see that the Kuhn-Tucker condition is satisfied and the optimal distribution is in the form given by (\[eq:optinput\]). Figures \[fig:locationbeta1\] and \[fig:probbeta1\] plot the magnitude and the probability of the nonzero amplitude respectively as a function of ($N_0 = 1$) for various values of $\kappa$. We immediately notice the significant impact of imposing a fourth moment constraint. When there is no such constraint, the nonzero amplitude migrates away from the origin as $\textrm{\footnotesize{SNR}} \rightarrow 0$ while its probability decreases sufficiently fast to satisfy the average power constraint. This type of input is called *flash signaling* in [@Verdu]. However, as we see from Figures \[fig:locationbeta1\] and \[fig:probbeta1\], if there is a fourth moment constraint with a finite $\kappa$, then the behavior is quite different. The nonzero amplitude approaches the origin as $\textrm{\footnotesize{SNR}} \rightarrow 0$ while its probability is kept constant. In the Rayleigh channel (${{\sf{K}}}= 0$), (\[eq:optinput\]) is still optimal at low ${\text{\footnotesize{SNR}}}$ up to a point after which, as ${\text{\footnotesize{SNR}}}$ is further lowered, the second moment constraint becomes inactive and we observe that the nonzero mass point approaches the origin more slowly while its probability decreases. From Fig. \[fig:capacitybeta1\], which plots the capacity curves as a function for various values of $\kappa$ in the low-power regime, we see that all the curves have the same first derivative at zero . This may suggest that performance in the low-power regime is similar for any finite value of $\kappa$. However, as we shall see in [@part2], the picture radically changes when we investigate the spectral-efficiency/bit-energy tradeoff. For the peak-power limited Rician fading channel (${{\sf{K}}}> 0$), numerical results indicate that for sufficiently low values, the optimal amplitude distribution has a single mass at the peak level $\sqrt{P}$ and hence all the information is carried on the uniform phase. For the Rayleigh channel (${{\sf{K}}}= 0$), an equiprobable two-mass-point distribution where one mass is at the origin and the other mass at the peak level is capacity-achieving in the low-power regime. Fig. \[fig:capK012\] plots the capacity curves for the peak-power limited Rayleigh channel and Rician channels with ${{\sf{K}}}= 1,2$ as a function of the peak . Note that the Rayleigh channel capacity curve has a zero slope at zero . For the Rician fading channel with phase noise (\[eq:pnmodel\]), numerical results illustrate again that a two-mass-point discrete distribution is optimal for sufficiently small values. Figures \[fig:pnlocation\] and \[fig:pnprob\] plot the magnitude and probability, respectively, of the optimal nonzero amplitude for this channel with Rician factors ${\sf{K}} = 0,1,2$. Note that only an average power constraint is imposed here. We observe that flash-signaling-type optimal input, where the nonzero amplitude migrates away from the origin as $\text{\footnotesize{SNR}} \rightarrow 0$ while its probability is decreasing, is required in the low-power regime. For fixed , we also see that the nonzero amplitude is closer to the origin for higher Rician factors . Finally Fig. \[fig:pncapacity\] provides the capacity curves as a function of for Rician factors ${\sf{K}} = 0,1,2$. Conclusion {#sec:conclusion} ========== In this paper, we have analyzed the structure of the capacity-achieving input for the noncoherent Rician fading channel. We have limited the peakiness of the input by imposing a fourth moment or a peak constraint. Using a sufficient and necessary condition, we have proven that when the input is subject to second and fourth moment limitations, the optimal input amplitude is discrete with a finite number of levels in the low-power regime. It turns out that a particular two-mass point distribution that depends only on the and $\kappa$ is asymptotically optimal as ${\text{\footnotesize{SNR}}}\rightarrow 0$. Discreteness of the optimal input amplitude distribution has also been shown for the peak-power limited Rician channel over the entire range. This time, the amplitude distribution with a single mass at the peak level is optimal in the low-power regime for the Rician channel with ${{\sf{K}}}> 0$. We also have analyzed a Rician fading channel model where there is phase noise in the specular component. We have shown that under an average power limitation, the optimal amplitude is discrete with a finite number of levels. For this model, we have provided numerical results for the capacity and the optimal input distribution where we observed that a flash-signaling-type input is required in the low-power regime. We have also proved that the optimal input for the average-power-limited classical Rician channel has bounded support. [0.8]{} [99]{} M. C. Gursoy, H. V. Poor, and S. Verdú, “The capacity of the noncoherent Rician fading channel," *Princeton University Technical Report*, 2002. Available online: http://www.princeton.edu/$\sim$mgursoy. M. C. Gursoy, H. V. Poor, and S. Verdú, “The noncoherent Rician fading channel – Part II : Spectral efficiency in the low-power regime," this issue. S. Verdú, “Spectral efficiency in the wideband regime," *IEEE Trans. Inform. Theory*, vol. 48, pp. 1319-1343, June 2002. R. S. Kennedy, *Fading Dispersive Communication Channels.*  Wiley Interscience, New York, 1969. J. S. Richters,“Communication over fading dispersive channels," Tech. Rep., MIT Research Laboratory of Electronics, Cambridge, MA, Nov. 1967. J. G. Smith,“The information capacity of peak and average power constrained Gaussian channels," *Inform. Contr.*, vol. 18, pp. 203-219, 1971. S. Shamai (Shitz) and I. BarDavid, “The capacity of average and peak-power-limited quadrature Gaussian channels," *IEEE Trans. Inform. Theory*, vol. 41, pp. 1060-1071, July 1995. I. Abou-Faycal, M. D. Trott, and S. Shamai (Shitz), “The capacity of discrete-time memoryless Rayleigh fading channels," *IEEE Trans. Inform. Theory*, vol. 47, pp. 1290-1301, May 2001. M. Katz and S. Shamai (Shitz), “On the capacity-achieving distribution of the discrete-time non-coherent and partially-coherent AWGN channels ," submitted to *IEEE Trans. Inform. Theory*, 2002. See also M. Katz and S. Shamai (Shitz), “On the capacity-achieving distribution of the discrete-time non-coherent additive white Gaussian noise channel ," *Proc. 2002 IEEE Int’l. Symp. Inform. Theory,* Lausanne, Switzerland, June 30 - July 5, 2002 , p. 165. S. Shamai (Shitz), “On the capacity of a pulse amplitude modulated direct detection photon channel," *Proc. IEE*, volume 137, pt. I, pages 424-430, Dec. 1990. A. Das, “Capacity-achieving distributions for non-Gaussian additive noise channels," *Proc. 2000 IEEE Int’l. Symp. Inform. Theory,* Sorrento, Italy, June 25-30, 2000, p. 432. R. Nuriyev and A. Anastasopoulos, “Capacity characterization for the noncoherent block-independent AWGN channel," *Proc. 2003 IEEE Int’l. Symp. Inform. Theory,* Yokohama, Japan, June 29 - July 4, 2003 , p. 373. J. Huang and S. P. Meyn, “Characterization and computation of optimal distributions for channel coding," submitted for publication. Published in abridged form in the proceedings of the 37th Annual Conference on Information Sciences and Systems, Baltimore, Maryland, March 12–14, 2003. T. L. Marzetta and B. M. Hochwald, “Capacity of a mobile multiple-antenna communication link in Rayleigh flat fading," *IEEE Trans. Inform. Theory*, vol. 45, pp. 139-157, Jan. 1999. A. Lapidoth, “Capacity bounds via duality: a phase noise example," *Workshop on Concepts in Information Theory,* Breisach, Germany, June 26-28, pp. 58-61, 2002. See also A. Lapidoth, “On phase noise channels at high SNR," *2002 IEEE Information Theory Workshop*, Bangalore, India, Oct. 20-25, pp. 1-4, 2002. A. Lapidoth and S. M. Moser, “Capacity bounds via duality with applications to multiple-antenna systems on flat-fading channels," *IEEE Trans. Inform. Theory*, vol. 49, pp. 2426-2467, Oct. 2003. M. Médard and R. G. Gallager, “Bandwidth scaling for fading multipath channels," *IEEE Trans. Inform. Theory*, vol. 48, pp. 840-852, Apr. 2002. V. G. Subramanian and B. Hajek, “Broad-band fading channels: signal burstiness and capacity," *IEEE Trans. Inform. Theory*, vol. 48, pp. 809-827, Apr. 2002. I. E. Telatar and D. N. C. Tse, “Capacity and mutual information of wideband multipath fading channels," *IEEE Trans. Inform. Theory*, vol. 46, pp. 1384-1400, July 2000. E. Biglieri, J. Proakis, and S. Shamai (Shitz), “Fading channels: Information-theoretic and communications aspects," *IEEE Trans. Inform. Theory*, vol. 44, pp. 2619-2692, October 1998. R. Palanki, “On the capacity achieving distributions of some fading channels," Fortieth Annual Allerton Conference on Communication, Control, and Computing, Monticello, IL, Oct. 2-4, 2002. A. J. Goldsmith and P. P. Varaiya, “Capacity of fading channels with channel side information," *IEEE Trans. Inform. Theory*, vol. 43, pp. 1986-1992, Nov. 1997. G. Caire and S. Shamai (Shitz), “On the capacity of some channels with channel state information," *IEEE Trans. Inform. Theory*, vol. 45, pp. 2007-2019, Sept. 1999. M. Médard and A. J. Goldsmith, “Capacity of time-varying channels with channel side information," *Proc. 1997 IEEE Int’l. Symp. Inform. Theory,* Ulm, Germany, June 29-July 4, 1997, p. 372. M. Médard, “The effect upon channel capacity in wireless communications of perfect and imperfect knowledge of channel," *IEEE Trans. Inform. Theory*, vol. 46, pp. 933-946, May 2000. R. B. Ash, *Information Theory.*  Dover: New York, 1990. R. G. Gallager, *Information Theory and Reliable Communication.*  Wiley: New York, 1968. T. M. Cover and J. A. Thomas, *Elements of Information Theory.*  Wiley: New York, 1991. W. Rudin, *Principles of Mathematical Analysis.*  McGraw Hill: New York, 1964. K. Knopp, *Theory of Functions.*  Dover: New York, 1945. S. Lang, *Complex Analysis 2nd. Ed.*   Springer-Verlag: New York, 1985. G. N. Watson, *A Treatise on the Theory of Bessel Functions 2nd. Ed.*   Cambridge University Press, 1958. [^1]: This research was supported by the U.S. Army Research Laboratory under contract DAAD 19-01-2-0011. The material in this paper was presented in part at the Fortieth Annual Allerton Conference on Communication, Control, and Computing, Monticello, IL, Oct., 2002 and the Canadian Workshop on Information Theory, Waterloo, Ontario, May 18-21, 2003. [^2]: Since the channel is memoryless, without loss of generality we can drop the index $i$. [^3]: The result holds in wider generality: the feasible set defined by (\[eq:constraint1\]) and (\[eq:constraint2\]) can be replaced by any set of constraints that are imposed only on the input magnitude. [^4]: The set of points of increase of a distribution function $F$ is $\{r: F(r-\epsilon) < F(r + \epsilon) \,\,\, \forall \epsilon >0$}. [^5]: The Bolzano-Weierstrass Theorem [@Rudin] states that every bounded infinite set of real numbers has a limit point. [^6]: The Identity Theorem for analytic functions [@Knopp] states that if two functions are analytic in a region $\mathcal{R}$, and if they coincide in a neighborhood, however small, of a point $z_0$ of $\mathcal{R}$, or only along a path segment, however small, terminating in $z_0$, or also only for an infinite number of distinct points with the limit point $z_0$, then the two functions are equal everywhere in $\mathcal{R}$. [^7]: The Kuhn-Tucker condition for the average-power-limited Rician case is essentially the same as (\[eq:ktc\]) with $\lambda_2 = 0$. [^8]: The impossibility of $\lambda = 0$ is shown in [@Abou]. [^9]: We note that the results in this section are obtained numerically, and hence there is no analytical claim of optimality.
{ "pile_set_name": "ArXiv" }
--- abstract: 'The measurement of the triple Higgs coupling is a key benchmark for the LHC and future colliders. It directly probes the Higgs potential and its fundamental properties in connection to new physics beyond the Standard Model. There exist two phase space regions with an enhanced sensitivity to the Higgs self-coupling, the Higgs pair production threshold and an intermediate top pair threshold. We show how the invariant mass distribution of the Higgs pair offers a systematic way to extract the Higgs self-coupling, focusing on the leading channel $pp\to hh+X\to b\bar b\ \gamma\gamma+X$. We utilize new features of the signal events at higher energies and estimate the potential of a high-energy upgrade of the LHC and a future hadron collider with realistic simulations. We find that the high-energy upgrade of the LHC to 27 TeV would reach a 5$\sigma$ observation with an integrated luminosity of 2.5 ab$^{-1}$. It would have the potential to reach 15% (30%) accuracy at the 68% (95%) confidence level to determine the SM Higgs boson self-coupling. A future 100 TeV collider could improve the self-coupling measurement to better than 5% (10%) at the 68% (95%) confidence level.' author: - Dorival Gonçalves - Tao Han - Felix Kling - Tilman Plehn - Michihisa Takeuchi title: 'Higgs Pair Production at Future Hadron Colliders: From Kinematics to Dynamics' --- Introduction ============ The discovery of the Higgs boson [@higgs; @discovery] at the CERN Large Hadron Collider (LHC) is of monumental significance. The completion of the Standard Model (SM) provides us with a consistent theory valid up to high scales. As a perturbative gauge theory, it allows for precision predictions for essentially all LHC observables. In parallel, experimental advances have turned ATLAS and CMS into the first hadron collider precision experiments in history. In combination, these developments open new avenues to tackle fundamental physics questions at the LHC and future high-energy facilities. On the theory side, we are still lacking an understanding of if and how the Higgs mass, the only dimensionful parameter in the theory, is stabilized against a large new physics scale. The Higgs potential responsible for the electroweak symmetry breaking (EWSB) in the SM is determined by the triple and quartic Higgs self-coupling $\lambda_\text{SM}\approx 1/8$. It is a true self-interaction in the sense that it is not associated with any conserved charge after EWSB. With our ignorance for new physics beyond the SM, the shape of the Higgs potential is deeply linked to the fundamental question of electroweak symmetry breaking in the early universe, allowing for a slow second-order phase transition in the SM or a strong first-order phase transition with a modified Higgs potential. It has been argued that a wide range of modified Higgs potentials, which result in a strong first-order EW phase transition, lead to order-one modifications of $\lambda_\text{SM}$ [@ew_phase]. All of this points to the Higgs self-coupling $\lambda$ as a benchmark measurement for the coming LHC runs, as well as any kind of planned colliders [@Arkani-Hamed:2015vfh]. Higgs pair production $pp\rightarrow hh$ offers a direct path to pin down $\lambda$ at a hadron collider [@hh-orig; @hh-early]. Previous studies show that promising final states from the $hh$ decays are $b\bar{b}\gamma\gamma$ [@hh-gamma; @vernon], $b\bar{b}\tau\tau$ [@hh-tautau-4b; @hh-tautau], $b\bar{b}WW$ [@hh-bbww], $b\bar{b}b\bar{b}$ [@hh-4b], and $4W$ [@hh-ww]. Theoretical studies as well as current analyses point to the $b\bar{b}\gamma\gamma$ decay as the most promising signature at the LHC [@current-gamma]. Combinations with indirect measurements of the self-coupling from quantum effects confirm that Higgs pair production provides the most robust self-coupling measurement [@indirect]. For the high-luminosity LHC (HL-LHC), ATLAS and CMS projections indicate a very modest sensitivity to the Higgs self-coupling [@hl-lhc]. In anticipation to probe new physics beyond the SM, it is customary to parametrize the modification of the self-coupling as $$\begin{aligned} \kappa_\lambda = \frac{\lambda}{\lambda_\text{SM}} \; . %= \kappa_\lambda \; \frac{3 m_h^2}{v} \; .\end{aligned}$$ In the optimistic scenario that we can neglect systematic uncertainties, those studies indicate that the LHC will probe the coupling at 95% confidence level $$\begin{aligned} -0.8 < \kappa_\lambda < 7.7\;.\end{aligned}$$ An issue with those studies is that they are based on the total rate for Higgs production, but neglect a wealth of available information. Including a full kinematic analysis could lead to an improved measurement [@madmax-hh] $$\begin{aligned} -0.2< \kappa_\lambda <2.6 \;,\end{aligned}$$ falling short in precision in comparison to other Higgs property measurements at the LHC, and far from satisfactory in probing the Higgs potential. In this study, we systematically compare the prospects for measuring the Higgs self-coupling at current and higher energy $pp$ colliders. We focus on the two leading proposals for future hadron colliders: 1. the 27 TeV high-energy LHC (HE-LHC) with an integrated luminosity of $15~\iab$, 2. a 100 TeV hadron collider with $30~\iab$, under consideration at CERN (FCC-hh) [@europe_100tev] and in China (SppC) [@china_100tev]. We include state of the art signal and background estimates for the $b\bar{b} \gamma \gamma$ channel, as well as realistic acceptance cuts and efficiencies. While there exist a series of 100 TeV studies of Higgs pair production at different levels of sophistication [@hh-nimatron], we include a 100 TeV analysis to be able to compare with the HE-LHC reach on equal footing. We start with a study of relevant phase space regions using a Neyman-Pearson maximum likelihood approach [@madmax-hh; @madmax]. This allows us to estimate the impact of using simple kinematic distributions on the measurement of the Higgs self-coupling at the different colliders. Furthermore, we can evaluate the maximum significance of extracting the Higgs pair signal and the significance of detecting a modified self-coupling under idealized conditions. In the main part of our paper, we perform a state-of-the-art analysis of Higgs pair production including additional jet radiation and a full set of realistic detector efficiencies. Unlike earlier analyses, we include $b$-jets from Higgs decays even when they become sub-leading in transverses momentum to the additional jet radiation. Our analysis focuses on the di-Higgs invariant mass distribution, both for the extraction of the Higgs pair signal and for the measurement of the Higgs self-coupling. Using a log-likelihood approach on this single kinematic distribution, we show that the Higgs self-coupling can be properly measured not only at a future 100 TeV collider, but also at the 27 TeV HE-LHC. Higgs Pair Signature {#sec:frame} ==================== The leading $hh$ production mechanism in the Standard Model at hadron colliders is depicted by the Feynman diagrams in Fig. \[fig:feyn1\]. Due to the difference of the top quark propagators in the loops, the two diagrams interfere destructively. In Fig. \[fig:xs\_hh\_Ecm\] we show the total rate for $hh$ production as a function of the center of mass energy $\sqrt s$ in TeV, including the next-to-leading order (NLO) corrections [@nlo]. The width of the curve illustrated the theoretical uncertainties around 10% [@nnlo]. At the LHC, the signal rate is the limiting factor for Higgs pair studies. At 14 TeV, the cross section including higher-order corrections is in the range of 0.033 pb [@nnlo], corresponding to at most 100k events with an integrated luminosity of $3~\iab$ at the HL-LHC. Assuming one Higgs decay to tagged bottom quarks, the available rate is reduced to 60k events in the life time of the HL-LHC. The crucial question is what kind of second Higgs decay allows us to effectively trigger the events and to reduce the QCD backgrounds to a manageable level. The leading candidate is the signature [@hh-gamma] $$\begin{aligned} pp \to hh \to b\bar{b} \; \gamma \gamma \; , \label{eq:bbgg}\end{aligned}$$ because of the excellent di-photon mass resolution and the guaranteed trigger. The expected number of signal events in the Standard Model at the HL-LHC is 260. Alternatively, the $b\bar{b} \; \tau \tau$ signature leads to 7.2k events times the tau tagging probability rate squared, and hampered by a significantly worse signal-to-background ratio. ![Representative Feynman diagrams contributing to the leading Higgs pair production process via gluon fusion.[]{data-label="fig:feyn1"}](diagram1 "fig:"){width=".22\textwidth"} ![Representative Feynman diagrams contributing to the leading Higgs pair production process via gluon fusion.[]{data-label="fig:feyn1"}](diagram2 "fig:"){width=".22\textwidth"}\ ![Total cross section for $pp\to hh$ production at NLO as a function of the $pp$ collider energy. The width of the curve reflects the 10% theoretical uncertainty. []{data-label="fig:xs_hh_Ecm"}](sigma_hh){width=".44\textwidth"} Because of the rapidly growing gluon luminosity at higher energies, the $hh$ production cross section increases by about a factor of 4 (40) at 27 (100) TeV. This means that at the HE-LHC with the anticipated integrated luminosity of $15~\iab$ the number of events in the $b\bar{b} \; \gamma \gamma$ channel increases by a factor $4 \times 5 = 20$ to around 5k events. A 100 TeV hadron collider with a projected integrated luminosity of $30~\iab$ features another increase by a factor $10 \times 2=20$, to around 100k expected Higgs pair events in the Standard Model. This estimate shows how the combination of increased energy and increased luminosity slowly turns Higgs pair production into a valid channel for precision measurements. The numbers fundamentally affect our proposed analysis strategy, because the small number of signal and background events suggests a kinematic analysis including as few kinematic distributions as possible. It is possible to improve this situation, for example, using the matrix element technique, as we will discuss below. We generate the signal with <span style="font-variant:small-caps;">MadGraph5</span> [@mg5], accounting for a next-to-leading order (NLO) QCD factor $K_\text{NLO}\sim 1.6$ [@nlo]. In the final state we demand two $b$-tagged jets and two isolated photons with the minimal acceptance and trigger cuts $$\begin{aligned} {5} && p_{T,j}>30~\gev , \quad |\eta_j |<2.5 \; , \notag \\ && p_{T,\gamma}>30~\gev, \quad |\eta_\gamma| <2.5 \; , \notag \\ && \Delta R_{\gamma \gamma, \gamma j, jj} >0.4 \; . \label{eq:base_selections}\end{aligned}$$ The background to our $b\bar{b} \; \gamma \gamma$ signal consists of other Higgs production modes ($t\bar{t}h, Zh$) with $h \to \gamma \gamma$, continuum $b\bar{b}\gamma\gamma$ production, and of multi-jet events with light-flavor jets faking either photons or $b$-jets ($jj\gamma\gamma, b\bar{b}\gamma j$) [@hh-gamma]. The different backgrounds are discussed in detail in Sec. \[sec:ana\]. The proper simulation of efficiencies and fake rates are a key ingredient for a realistic background estimate in this analysis. For the HE-LHC and the future 100 TeV collider we follow the ATLAS projections [@performance]. The efficiency for a tight photon identification can be well parametrized by $$\begin{aligned} \epsilon_{\gamma\to\gamma} = 0.863 - 1.07 \cdot e^{-p_{T,\gamma}/34.8~\gev}\;,\end{aligned}$$ and a jet-to-photon mis-identification rate by $$\begin{aligned} \epsilon_{j\to\gamma} = \begin{cases} 5.3\cdot 10^{-4} \exp \left( -6.5 \left( \dfrac{p_{T,j}}{60.4~\gev} - 1 \right)^2 \right)\;, \notag \\[4mm] 0.88 \cdot 10^{-4} \left[ \exp \left( -\dfrac{p_{T,j}}{943~\gev} \right) +\dfrac{248~\gev}{p_{T,j}}\right] \;, \end{cases}\end{aligned}$$ where the upper form applied to softer jets with ${p_{T,j} <65}$ GeV. This leads to a photon efficiency of about 40% at $p_{T,\gamma}=30$ GeV, saturating around 85% for $p_{T,\gamma}>150$ GeV. Note that the Higgs decay products tend to be soft, $p_{T,\gamma}\sim m_h/2$. For $b$-tagging, we adopt an efficiency with $$\begin{aligned} \epsilon_b =0.7 \; , \end{aligned}$$ associated with mis-tag rates of 15% for charm quarks and 0.3% for light flavors. These flat rates present a conservative estimate from the two dimensional distribution on $(p_{Tj},\eta_j)$ shown in the HL-LHC projections [@madmax-hh]. Encouragingly, the small light flavor fake rate projections result in a strong suppression for the initially dominant $jj\gamma\gamma$ background. Obviously, the final outcome of the analyses would depend on the detector performance for the efficiencies of photon identification and $b$-tagging, as well as the background jet rejection. To have a comprehensive exploration and comparison, we will also examine the other available detector parameters, one from CMS [@Chatrchyan:2012jua] and the other from the CERN Yellow Report [@Mangano:2017tke] for the future collider (FCC), as shown in the Appendix. ![image](Plot_mhh_27TeV){width="32.00000%"} ![image](Plot_pth_27TeV){width="32.00000%"} ![image](Plot_draa_27TeV){width="32.00000%"}\ ![image](Plot_mhh_100TeV){width="32.00000%"} ![image](Plot_pth_100TeV){width="32.00000%"} ![image](Plot_draa_100TeV){width="32.00000%"} The Mother of Distributions {#sec:features} =========================== As depicted in Fig. \[fig:feyn1\], Higgs pair production receives contributions from a triangular loop diagram combined with the Higgs self-coupling and from a box or continuum diagram (plus a crossing diagram), where over most of phase space the box contribution completely dominates the total rate. While we can define a number of kinematic observables describing the continuum backgrounds, the measurement of the Higgs self-coupling relies on a simple $2 \to 2$ process with two independent kinematic variables. Three distinct phase space regions provide valuable information on a modified Higgs self-coupling, both from a large destructive interference between the triangle and box contributions. First, there is the threshold [@hh-early; @hh-ww] in the partonic center of mass energy $$\begin{aligned} m_{hh}^\text{(th)} \approx 2 m_h \; .\end{aligned}$$ In the absence of hard additional jets, the di-Higgs invariant mass is identical to the partonic collider energy $s \equiv m_{hh}^2$. Note that this threshold is below $2m_t$. Based on the effective Higgs–gluon Lagrangian [@low_energy] we can write the corresponding amplitude for Higgs pair production as $$\begin{aligned} \frac{\alpha_s}{12 \pi v} \left( \frac{\kappa_\lambda \lambda_\text{SM}}{s-m_h^2} - \frac{1}{v} \right) \to \frac{\alpha_s}{12 \pi v^2} \left( \kappa_\lambda -1 \right) \stackrel{\text{SM}}{=} 0 \; . \label{eq:higgs_pair}\end{aligned}$$ While the heavy-top approximation is known to give a poor description of the signal kinematics as a whole, it does describe the threshold dependence correctly [@hh-ww]. This indicates that we can search for a deviation of the Higgs self-coupling by looking for an enhancement of the rate at threshold. Second, an enhanced sensitivity to the self-coupling appears as top mass effect. For large positive values of $\lambda$ absorptive imaginary parts lead to a significant dip in the combined rate at the threshold $p_{T,h} \approx 100$ GeV [@hh-tautau] or equivalently [@madmax-hh] $$\begin{aligned} m_{hh}^\text{(abs)} \approx 2 m_t \; .\end{aligned}$$ The sharpest interference dip takes place near $\lambda\approx 2$. For negative values of $\lambda$ the interference becomes constructive. Finally, the triangular and box amplitudes generally have different scaling in the limit [@hh-early; @hh-tautau] $$\begin{aligned} m_{hh}^\text{(high)} \gg m_h, m_t \; .\end{aligned}$$ While the triangle amplitude features an explicit suppression of either $m_h^2/m_{hh}^2$ or $m_t^2/m_{hh}^2$ at high invariant mass, the box diagrams drop more slowly towards the high-energy regime. The impact of all three kinematic features can be quantified statistically and is illustrated in detail in Fig. 5 of Ref. [@madmax-hh]. They clearly indicate that essentially the full information on the Higgs self-coupling can be extracted through a shape analysis of the $m_{hh}$ distribution [@martin]. The practical relevance of the different kinematic regimes has to be estimated including the variation of the signal cross section, the number of expected events at a given collider, and the size of the backgrounds. There exist two similar statistical approaches to answer this problem, the <span style="font-variant:small-caps;">MadMax</span> approach based on the Neyman-Pearson lemma [@madmax] and the <span style="font-variant:small-caps;">MadFisher</span> approach based on information geometry [@madfisher]. While the latter is especially well-suited to estimate the reach for example of precision measurements at the LHC, we employ the former for a simple hypothesis test. The integrated log-likelihood ratio over the full phase space or specific kinematic regimes allows us to estimate the maximum significance with which any multi-variate analysis will be able to extract a signal from backgrounds or distinguish two assumed values of the Higgs self-coupling [@madmax-hh]. Throughout maximum likelihood analysis we limit ourselves to irreducible backgrounds and assume that statistical uncertainties dominate over the relevant phase space regions. Events with soft final states typically contribute little to the search for new particles with weak-scale masses. The exact choice of acceptance cuts in Eq.  and the modeling of $b$-tagging or photon identification efficiencies will have a negligible effect on our results. For our numerical analysis, we account for all backgrounds discussed in Sec. \[sec:frame\], except for the $t\bar{t}h$ channel with its significantly different final state. As part of the detailed background analysis in Sec. \[sec:ana\], we will see that this assumption is justified. The setup is essentially identical to Ref. [@madmax-hh], but now using the cuts and fake rates given in Sec. \[sec:frame\]. In particular, we account for the smearing of the Higgs peak as leading detector effect. The invariant mass distributions are smeared by a Gaussian with width 1.52 GeV for the $\gamma\gamma$ channel [@CMS:2016zjv] and 12.6 GeV for the $bb$ channel [@Vernieri:2014wfa]. The signal rate is adjusted to account for the loss of signal rate through a poor description of the tails of the distributions [@madmax-hh]. This allows us to restrict ourself to the two Higgs mass windows $m_{bb} = 80~...~160$ GeV and $m_{\gamma \gamma} = 120~...~130$ GeV. All other detector effects are left to our actual analysis in Sec. \[sec:ana\]. In Fig. \[fig:madmax\_diff\] we first show the signal and background distributions for three relevant kinematic variables, $m_{hh}$, $p_{T,h}$, and $\Delta R_{\gamma \gamma}$. The transverse momentum distributions of the two Higgs bosons will be identical, so we can measure them either as $p_{T,\gamma \gamma}$ or as $p_{T,bb}$. Both, for $m_{hh}$ and $p_{T,h}$ the QCD backgrounds reside at small values, with similar signal-to-background ratios at the HE-LHC and the 100 TeV collider. The geometric separation of the two photons from the continuum background has to be large to generate an invariant mass around the Higgs mass. ![Higgs pair production cross section (red lines with left vertical axis) and maximum significance (black lines with right vertical axis) for discriminating an anomalous self-coupling $\kappa_\lambda \ne 1$ from the SM, as a function of the modified self-coupling. The results are for the HL-LHC, the HE-LHC, and a future 100 TeV collider, respectively. The HL-LHC results are taken from Ref. [@madmax-hh].[]{data-label="fig:madmax_tot"}](XSZ_vs_Lambda){width="40.00000%"} Also in Fig. \[fig:madmax\_diff\], we show how the significance of extracting an anomalous self-coupling $\kappa_\lambda \ne 1$ depends on these key observables. The alternative hypothesis in this case is the combination of the backgrounds and the signal with $\kappa_\lambda = 1$. In addition to the signal features, the significance is limited by the rapidly dropping backgrounds, covering both of the above-mentioned regions with an enhanced dependence on the triangle diagram. In the absence of background, the significance indeed peaks between the production threshold and the top-mass threshold [@madmax-hh]. The drop towards large values of $m_{hh}$ is a combination of the dominance of the box diagram in the signal and the limited number of expected signal events. The significance with which we can extract modified self-couplings either smaller ($\kappa_\lambda = 0$) or larger ($\kappa_\lambda = 2$) than in the Standard Model shows a similar phase space dependence. The only difference is a slightly harder significance distributions for $\kappa_\lambda = 2$, an effect of the dip at $m_{hh}^\text{(abs)}$. Obviously, we can combine the maximum significance distributions into a global maximum significance accumulated over the full phase space. In Fig. \[fig:madmax\_tot\] we show the idealized, maximum significance with which we can hope for at the HL-LHC, the HE-LHC, and a future 100 TeV collider. The asymmetric behavior for the HL-LHC is a remainder of a degeneracy in the total cross section as a function of the self-coupling, also shown in Fig. \[fig:madmax\_tot\]. A SM-like rate appears when an enhanced triangle diagram overcomes the larger box contribution and flips the sign of the amplitude. Obviously, this degeneracy will be broken by kinematic information, for example the $m_{hh}$ distribution. For the HE-LHC and the 100 TeV collider, the total rate constraint becomes increasingly irrelevant for the measurement of the self-coupling. The expected statistical error bars are narrow and approximately symmetric around on $\kappa_\lambda = 1$. For both future colliders, we can indeed expect a proper measurement of the Higgs self-coupling. Detector-level Analysis {#sec:ana} ======================= ![Representative Feynman diagrams contributing to Higgs pair production via gluon fusion including an ISR jet at hadron colliders.[]{data-label="fig:feyn2"}](diagram3 "fig:"){width=".22\textwidth"} ![Representative Feynman diagrams contributing to Higgs pair production via gluon fusion including an ISR jet at hadron colliders.[]{data-label="fig:feyn2"}](diagram4 "fig:"){width=".22\textwidth"} ![image](ptj2_decomposition){width=".38\textwidth"} ![image](ptj2_decomposition_100TeV){width=".38\textwidth"} Following the analysis path laid out in Sec. \[sec:features\], we now design a detailed analysis strategy to extract the Higgs self-coupling with a focus on the shape of the $m_{hh}$ distribution. Our signal is $$\begin{aligned} pp \to hh + X \to b\bar{b} \; \gamma \gamma + X.\end{aligned}$$ In anticipation of increasing QCD radiation at higher energies, we inclusively allow extra jets in the events from initial state radiation, along with two tagged $b$-jets and two isolated hard photons, passing the acceptance cuts of Eq. . [ll|rrr|rrrrr|r|cc]{} Collider & Process & & $t\bar{t}h$ & $Zh$ & $b\bar{b}\gamma\gamma$ & $jj\gamma\gamma$ & $b\bar{b}\gamma j$ & BG tot. & $S/\sqrt{S+B}_{1\iab}$ & $S/B$\ & & 0 & 1 & 2 & & & & & & & &\ &$\sigma$ \[fb\] & 0.69 & 0.36 & 0.18 & 6.43 & 0.77 & 1.24 pb & 36.6 pb & 506 pb & & &\ & Baseline & 2.87K & 1.57K & 838 & 21.8K & 1.44K & 1.19M & 36M & 1.13M & 38.3M & 0.07 & $4\cdot 10^{-5}$\ &$n_{j} \le 3$, $n_b =2$ & 648 & 356 & 190 & 954 & 389 & 200K & 67.4K & 105K & 374K & 0.15& $1 \cdot 10^{-3}$ \ &$\Delta m_{bb} \le 25$ GeV & 470 & 260 & 140 & 195 & 66 & 43.7K & 10.6K & 25.8K & 80.4K & 0.24 & 0.003\ &$\Delta m_{\gamma\gamma} \le 3$ GeV & 459 & 253 & 136 & 197 & 63 & 1.42K & 505 & 758 & 2.94K & 1.2 & 0.09\ (15 ab$^{-1}$) &$\Delta m_{\gamma\gamma} \le 2$ GeV & 459 & 253 & 136 & 197 & 63 & 957 & 342 & 504 & 2.06K & 1.4 & 0.12\ &$\Delta m_{\gamma\gamma} \le 1$ GeV & 459 & 253 & 136 & 197 & 63 & 485 & 182 & 245 & 1.17K & 1.7 & 0.22\ &$\Delta m_{\gamma\gamma} \le 3$ GeV, $m_{hh}>400$ & 320 & 206 & 120 & 56 & 21 & 324 & 97 & 178 & 676 & 1.8 & 0.30\ &$\Delta m_{\gamma\gamma} \le 2$ GeV, $m_{hh}>400$ & 320 & 206 & 120 & 56 & 21 & 220 & 67 & 122 & 485 &2.0 & 0.42\ &$\Delta m_{\gamma\gamma} \le 1$ GeV, $m_{hh}>400$ & 320 & 206 & 120 & 56 & 21 & 115 & 41 & 61 & 293 & 2.4& 0.70\ &$\sigma$ \[fb\] & 6.95 & 3.72 & 1.97 & 84.8 & 3.76 & 6.21 pb & 126 pb & 3.03 nb & & &\ & Baseline & 51.8K & 29.8K & 16.9K & 535K & 13.1K & 13.6M & 330M & 18.6M & 363M & 0.29 & $8\cdot 10^{-5}$&$n_{j} \le 3$, $n_b =2$ & 9.22K & 5.28K & 3.02K & 18K & 2.84K & 1.79M & 773K & 1.42M & 4.00M & 0.48 & 0.001 &$\Delta m_{bb} \le 25$ GeV & 6.45K & 3.80K & 2.18K & 3.3K & 669 & 361K & 218K & 373K & 956K & 0.71 & 0.004\ &$\Delta m_{\gamma\gamma} \le 3$ GeV & 6.30K & 3.70K & 2.13K & 3.12K & 653 & 8.34K & 6.06K & 8.99K & 27.2K & 3.9 & 0.14\ (30 ab$^{-1}$) &$\Delta m_{\gamma\gamma} \le 2$ GeV & 6.30K & 3.70K & 2.13K & 3.12K & 653 & 5.66K & 4.13K & 5.99K & 19.5K & 4.4 & 0.19\ &$\Delta m_{\gamma\gamma} \le 1$ GeV & 6.30K & 3.70K & 2.13K & 3.12K & 653 & 2.82K & 1.91K & 2.99K & 11.4K & 5.5 & 0.32\ &$\Delta m_{\gamma\gamma} \le 3$ GeV, $m_{hh}>400$ & 4.66K & 3.16K & 1.93K &1.09K & 203 & 1.56K & 1.10K & 1.90K & 5.86K & 6.1 & 0.54\ &$\Delta m_{\gamma\gamma} \le 2$ GeV, $m_{hh}>400$ & 4.66K & 3.16K & 1.93K &1.09K & 203 & 1.04K & 747 & 1.14K & 4.23K& 6.7 & 0.73\ &$\Delta m_{\gamma\gamma} \le 1$ GeV, $m_{hh}>400$ & 4.66K & 3.16K & 1.93K &1.09K & 203 & 523 & 359 &617 & 2.79K & 7.5 & 1.13\ For the detector-level analysis we generate the signal and background samples with <span style="font-variant:small-caps;">MadGraph5</span>+<span style="font-variant:small-caps;">Pythia8</span> [@mg5; @pythia8], including one extra jet using the <span style="font-variant:small-caps;">Mlm</span> scheme [@mlm]. A representative set of Feynman diagrams for the signal is shown in Figs. \[fig:feyn1\] and \[fig:feyn2\]. Higher-order corrections are included through a next-to-leading order $K$-factor 1.6 [@nlo; @nnlo; @hh_madgraph], neglecting possible higher-order effects on the $m_{hh}$ distribution. We normalize the $t\bar{t}h$ and $Zh$ to their respective NLO and NNLO rates 2.8 pb and 2.2 pb at 27 TeV (37 pb and 11 pb at 100 TeV) [@tth_zh]. We also include the full set of detector effects with <span style="font-variant:small-caps;">Delphes3</span> [@delphes], following the HL-LHC projections [@performance]. Jets are defined with the anti-$k_T$ algorithm ${R=0.4}$ via <span style="font-variant:small-caps;">FastJet</span> [@fastjet]. While the $t\bar{t}h$ background is almost irrelevant at the 14 TeV LHC, it becomes increasingly important at higher energies. Obviously, the more complex, high-multiplicity final state offers many handles to tame it. We employ a simple veto on leptons with $$\begin{aligned} p_{T,\ell}>10~\gev\ \text{and}\ |\eta_\ell|<2.5 \; ,\end{aligned}$$ combined with a veto of more than three jets passing Eq. . ![image](mhh_lambda_dependence_1GeV_nj3_rebin_log){width=".4\textwidth"} ![image](mhh_lambda_dependence_1GeV_nj3_rebin_100TeV_log){width=".4\textwidth"} To suppress the initially overwhelming $jj\gamma\gamma$ background, we demand two $b$-tags among the three hardest jets. A crucial observation is that at higher energies, initial state radiation (ISR) often leads to a harder jet than the Higgs decay products, such that either the hardest or second-hardest jet is not a $b$-jet for roughly half of all events. This is illustrated in Fig. \[fig:compo\] as the composition of the second-hardest parton-level jet, requiring that both truth-level $b$-jets pass the selection of Eq. . Thus, the $b$-tagging requirement as the two leading jets should be adjusted accordingly. Based on this observation we account for two patterns of the $p_T$ jets, $(bb,bbj)$ and $(jbb,bjb)$. This increases our signal efficiency by around 50%. Expanding this scheme to even more jets is not effective because it eventually also increases the continuum backgrounds and the $t\bar{t}h$ contributions. The reliability of our Monte Carlo simulation underlying this procedure is guaranteed by the fact that the hardest three jets are generated using multi-jet merging. To control the continuum backgrounds, we require two Higgs mass windows, $$\begin{aligned} |m_{bb}-m_h|<25~\gev, \quad |m_{\gamma\gamma}-m_h|<1~\gev . \label{eq:jreslov}\end{aligned}$$ An obvious way to enhance the Higgs pair signal is to improve the resolution on the reconstructed photons and $b$-jets from the Higgs decays. We adopt the rather conservative resolution for $m_{bb}$ as in Eq. . Any improvement on it in experiments would be greatly helpful for the signal identification and background separation. As for the photon resolution, we illustrate this effect by using three representative values where the $m_{\gamma\gamma}$ distribution is smeared by a Gaussian width of $0.75$, $1.5$, $2.25$ GeV, corresponding to Higgs mass windows $$\begin{aligned} |m_{\gamma\gamma} - m_h| \le 1,2,3~\gev. \label{eq:preslov}\end{aligned}$$ A resolution of $1.5$ GeV has already been achieved at the LHC [@CMS:2016zjv]. The results at this stage of the analysis are illustrated in Table \[tab:cutflow\] with a full cut flow for the two collider energies and assuming $\kappa_\lambda = 0,1,2$. We already find a large background suppression $S/B\sim 0.09~...~0.2$ for the HE-LHC and $0.14~...~0.3$ at a future 100 TeV collider. Requiring ${m_{hh}>400}$ GeV improves it to $S/B\sim 0.3~...~0.7$ or $0.5~...~1.1$, respectively. This is entirely due to the rapidly falling backgrounds as compared to the $hh$ signal,but will be at the expense of the self-coupling determination. The $m_{hh}$ distribution of the signal and the different backgrounds is shown in Fig. \[fig:reso\]. ![image](hh_5sigma_2b3j){width=".4\textwidth"} ![image](hh_5sigma_width){width=".4\textwidth"} The signal-to-background ratio can be strongly improved by a better $m_{\gamma\gamma}$ resolution. As long as most of the $h\to \gamma\gamma$ events are captured by an appropriate $m_{\gamma\gamma}$ window, the contributions from continuum backgrounds can be estimated using the side-band measurements. Going beyond a cut-based analysis for example on $m_{hh}$, we employ a binned log-likelihood analysis based on the CL$_{s}$ method, using the full $m_{hh}$ distribution to extract $\kappa_{\lambda}$ [@read]. The dominant backgrounds feature powerful control regions or ratio measurements like $t\bar{t}h/t\bar{t}Z$ [@nimatron_yt]. Therefore, we neglect their systematic uncertainties. As a starting point, we show the $5\sigma$ determination on the Higgs pair signal strength in the left panel of Fig. \[fig:bound1\], requiring two $b$-tagged jets among the two or three leading jets. We decompose the latter case in two sub-samples $(bb,bbj)$ and $(jbb,bjb)$. We see how exploring the extra-jet emission significantly improves the significance as compared to the standard procedure adopted in the literature. The $5\sigma$ measurement for HE-LHC is pushed from $2.8~\iab$ to below $2.3~\iab$. In the right panel of Fig. \[fig:bound1\] we show the discovery reach for the Higgs pair signal as a function of the luminosity of the HE-LHC and the 100 TeV collider. We assume three di-photon invariant mass resolutions with three Higgs mass windows as in Eq.  for a SM self-coupling $\kappa_{\lambda}=1$. Higgs pair production will be discovered at the HE-LHC with approximately $2.5~...~5~\iab$ and at the 100 TeV collider with $0.2~...~0.3~\iab$ of data, in both cases well below the design luminosity. As commented in the Introduction, there exist physics scenarios that the Higgs self-coupling could be modified at the level of order one deviation from the SM value. The accurate measurement of the Higgs self-coupling via Higgs pair production at future colliders has the best promise to uncover the new physics associated with the Higgs sector. In Fig. \[fig:bound2\], we show the accuracy on this measurement. At the 68% confidence level the triple Higgs coupling can be measured with the precision $$\begin{aligned} \kappa_{\lambda} &\approx 1 \pm 15\% \qquad \text{(HE-LHC, 27~TeV, $15~\iab$)} , \notag \\ \kappa_{\lambda} &\approx 1 \pm 5\% \ \,\qquad \text{(100~TeV, $30~\iab$).} \end{aligned}$$ At the 95% confidence level, $$\begin{aligned} \kappa_{\lambda} &\approx 1 \pm 30\% \qquad \text{(HE-LHC, 27~TeV, $15~\iab$)} , \notag \\ \kappa_{\lambda} &\approx 1 \pm 10\% \qquad \text{(100~TeV, $30~\iab$).} \end{aligned}$$ The way to improve these expected limits towards the mathematically-defined best reach shown in Fig. \[fig:madmax\_tot\] is to exploit more kinematic features and this way also suppress the reducible $t\bar{t}h$ background. ![Confidence level for separating an anomalous Higgs self-coupling hypothesis from the Standard Model $\kappa_{\lambda}=1$.[]{data-label="fig:bound2"}](hh_dlamb){width=".4\textwidth"} To gain some insight on how robust our results are, we have also examined the other available choices of detector parameters, one from CMS [@Chatrchyan:2012jua] and the other from the CERN Yellow Report (YR)[@Mangano:2017tke] for the future collider (FCC). As shown in Fig. \[fig:bound\_appendix\] in the Appendix, we find that the results are quite consistent with each other, with the YR performance being slightly better. This indicates possible room for further improvement. Summary and Outlook {#sec:sum} =================== In this paper, we have explored Higgs pair production as a direct way to measure the Higgs self-coupling, the least-known but arguably the most important fundamental parameter of the Standard Model. We first presented the production cross section for $pp \to hh$ at future high-energy colliders in Fig. \[fig:xs\_hh\_Ecm\], Sec. \[sec:frame\]. We discussed the signal rate for the process with leading sensitivity $pp\to hh\to b\bar b\ \gamma\gamma$, and laid out the event selection criteria in accordance with the experimental acceptance at the LHC. In Sec. \[sec:features\], we discussed the kinematic features of the signal and compared with the backgrounds, as shown in Fig. \[fig:madmax\_diff\]. The key variable is the invariant mass distribution of the Higgs pair that presented distinctive behaviors. We first performed a parton-level analysis that combines the maximum significance distributions into a global maximum significance accumulated over the full phase space, for the HL-LHC, the HE-LHC, and a future 100 TeV collider. For both future colliders we found excellent prospects for kinematics-based determinations of the Higgs self-coupling as shown in Fig. \[fig:madmax\_tot\]. In Sec. \[sec:ana\], we then carried out a search strategy based on a rate combined with kinematic shapes with realistic simulations. The approach is not only more powerful [@hh-ww; @madmax-hh] than a purely rate-based measurement but also more stable against systematic and theoretical uncertainties, provided we account for all bin-to-bin correlations. Our method removes all degeneracies which appear in a rate-based measurement and leads to well-defined symmetric error bars on the modified self-coupling. Higher energy colliders allow for including events with high $m_{hh}$. In such more and more common configurations at high energies, the additional jets from QCD radiation frequently surpass the $b$-jet energy about $m_h/2$, as seen in Fig. \[fig:compo\]. To improve the signal efficiency we included at least three observable jets, fully accounting for QCD jet radiation via the [<span style="font-variant:small-caps;">M</span>LM]{} merging, with possibly softer $b$-jets from Higgs decays. We showed a cut-flow in Table \[tab:cutflow\] to illustrate the staged improvements and to give a comparison for the two future colliders. We further enhance our measured significances, decomposing the samples into two sub-samples $(bb,bbj)$ and $(jbb,bjb)$. Finally, we determined the integrated luminosity needed to reach a $5\sigma$ significance to observe the SM $hh$ signal as shown in Fig. \[fig:bound1\]. We found that the high-energy upgrade of the LHC to 27 TeV would reach a 5$\sigma$ observation of the Higgs pair production with an integrated luminosity of about 2.5 ab$^{-1}$. It would have the potential to reach 15% (30%) accuracy at the 68% (95%) confidence level to determine the SM Higgs boson self-coupling. A future 100 TeV collider could improve the self-coupling measurement to better than 5% (10%) at the 68% (95%) confidence level, as shown in Fig. \[fig:bound2\]. These results roughly agree with the optimal reach shown in Fig. \[fig:madmax\_tot\]. Our conclusions are quite robust against some moderate variations of the detector performances as shown in Fig. \[fig:bound\_appendix\] in the Appendix. In the hope of searching for effects from physics beyond the SM, our results should provide conclusive information weather or not the Higgs-self-interaction is modified to a level of order one. While our conclusions on the determination of Higgs-self-interaction at future hadron colliders are robust and important, there is still room for improvement. Although the final state $b\bar b\ \gamma\gamma$ is believed to be the most sensitive channel because of the background suppression and signal reconstruction, there exist complementary channels such as $gg\to hh \to b\bar b\ \tau^+\tau^-$, $b\bar b\ W^+W^-$, $b\bar b\ b\bar b$, etc. The kinematics-based measurement and the all features related to QCD radiation at higher energies should be equally applicable to all of them. Appendix {#sec:appendix} ======== As explained in the text, we optimize our set of selection cuts primarily to reduce the continuum background, which would be accompanied by large systematic uncertainty, and secondarily to reduce the $t\bar{t}h$ background, which is the largest background component with a Higgs mass peak structure. To achieve the above optimization, we take the photon identification working point with a reasonably efficient jet-fake rejection [@performance], and require the additional jet veto ($n_j \le 3$). We believe our selection is almost optimal, but for completeness, we assess the effects of applying different efficiencies taken in the literature and provide the final sensitivities assuming those numbers. For comparison, we have worked on two different efficiency scenarios found for the CMS projections [@Chatrchyan:2012jua] and in the CERN Yellow Report (YR) [@Mangano:2017tke] for the study of Future Circular Colliders (FCC). ![Comparison of the final confidence level for separating an anomalous Higgs self-coupling hypothesis from the Standard Model $\kappa_{\lambda}=1$ for several efficiency choices. We display the results for 27 TeV (top panel) and 100 TeV (bottom panel)[]{data-label="fig:bound_appendix"}](hh_dlamb_projections_27tev "fig:"){width=".4\textwidth"} ![Comparison of the final confidence level for separating an anomalous Higgs self-coupling hypothesis from the Standard Model $\kappa_{\lambda}=1$ for several efficiency choices. We display the results for 27 TeV (top panel) and 100 TeV (bottom panel)[]{data-label="fig:bound_appendix"}](hh_dlamb_projections_100tev "fig:"){width=".4\textwidth"} We adopt the fitted CMS projections as follows: $$\begin{aligned} &&\epsilon_{\gamma\to\gamma}=0.85,\cr &&\epsilon_{j\to\gamma}=\begin{cases} 0.0113\exp(- \frac{p_T}{26.3~\gev}) \ \ [p_T<100~\gev]\cr 0.0025 \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ [p_T \ge 100~\gev] \end{cases},\cr &&\epsilon_{b\to b}=0.85\tanh\left(\frac{p_T}{400~\gev}\right)\frac{25.0}{1+ p_T/15.9~\gev}, \cr &&\epsilon_{c\to b}=0.25\tanh\left(\frac{p_T}{55.6~\gev}\right)\frac{1}{1+ p_T/769~\gev},\cr &&\epsilon_{j\to b}=0.01.\end{aligned}$$ The efficiency set used in the YR is the following: $$\begin{aligned} &&\epsilon_{\gamma\to\gamma}=0.9, \ \ \ \epsilon_{j\to\gamma}=0.01\exp\left(-\frac{p_T}{30~\gev}\right),\cr &&\epsilon_{b\to b}=0.75,\ \ \ \epsilon_{c\to b}=0.1,\ \ \ \epsilon_{j\to b}=0.01.\end{aligned}$$ Fig. \[fig:bound\_appendix\] shows the comparison among the final results using the three different sets of the efficiencies for 27 TeV (top) and 100 TeV (bottom). The red lines show the final results assuming our adopted efficiencies (from the ATLAS HL-LHC projection study) [@performance], while the green and the blue lines show those assuming the YR and the CMS ones, respectively. Our analysis sensitivity is not much improved by taking the working points with a larger photon efficiency used by these two alternative references, due to the corresponding worse light-jet rejection rate, which enhances the continuum background, especially the $bb\gamma j$ contribution. Note that we devise our analysis with a large $S/B$ by targeting to reduce the continuum background $bb\gamma j$ and leave the main background contributions from $t\bar{t}h$. In this way we achieve $S/B\sim 0.7$ against the corresponding numbers 0.45 (YR), and 0.4 (CMS), respectively, for the 27 TeV analysis. For the 100 TeV analysis, we achieve $S/B\sim 1.1$ against 0.6 (YR) and 0.5 (CMS). Thus, we can provide a more robust estimate against the systematic uncertainty of the continuum background. Additionally, it allows us to have a larger sensitivity from the lower $m_{hh}$ profile, a regime that is more background contaminated and that displays larger effects on the triple Higgs coupling. **Acknowledgment** We would like to thank Michelangelo Mangano and Michele Selvaggi for discussions. This work was supported in part by the U.S. Department of Energy under grant No. DE-FG02- 95ER40896 and by the PITT PACC. DG is supported in part by the U.S. National Science Foundation under the grant PHY-1519175. FK is supported by the U.S. National Science Foundation under the grant PHY-162063. MT is supported in part by the JSPS Grant-in-Aid for Scientific Research Numbers JP16H03991, JP16H02176, 17H05399, and by World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan. TP would like to thank Uli Baur (15 years ago) and Michael Spannowsky for helpful discussions concerning kinematic features of Higgs pair production. [99]{} P. W. Higgs, Phys. Lett.  [**12**]{}, 132 (1964); P. W. Higgs, Phys. Rev. Lett.  [**13**]{}, 508 (1964); F. Englert and R. Brout, Phys. Rev. Lett.  [**13**]{}, 321 (1964). G. Aad [*et al.*]{} \[ATLAS Collaboration\], Phys. Lett. B [**716**]{}, 1 (2012) \[ \[hep-ex\]\]; S. Chatrchyan [*et al.*]{} \[CMS Collaboration\], Phys. Lett. B [**716**]{}, 30 (2012) \[ \[hep-ex\]\]. X. m. Zhang, Phys. Rev. D [**47**]{}, 3065 (1993) \[\]; X. Zhang, B. L. Young and S. K. Lee, Phys. Rev. D [**51**]{}, 5327 (1995) \[\]; C. Grojean, G. Servant and J. D. Wells, Phys. Rev. D [**71**]{}, 036001 (2005) \[\]; S. Kanemura, Y. Okada and E. Senaha, Phys. Lett. B [**606**]{}, 361 (2005) \[\]; A. Noble and M. Perelstein, Phys. Rev. D [**78**]{}, 063518 (2008) \[ \[hep-ph\]\]; K. Fuyuto and E. Senaha, Phys. Rev. D [**90**]{}, no. 1, 015015 (2014) \[ \[hep-ph\]\]; F. P. Huang, P. H. Gu, P. F. Yin, Z. H. Yu and X. Zhang, Phys. Rev. D [**93**]{}, no. 10, 103515 (2016) \[ \[hep-ph\]\]; P. Huang, A. Joglekar, B. Li and C. E. M. Wagner, Phys. Rev. D [**93**]{}, no. 5, 055049 (2016) \[ \[hep-ph\]\]; A. Kobakhidze, L. Wu and J. Yue, JHEP [**1604**]{}, 011 (2016) \[ \[hep-ph\]\]; C. Y. Chen, J. Kozaczuk and I. M. Lewis, JHEP [**1708**]{}, 096 (2017) \[ \[hep-ph\]\]; X. Gan, A. J. Long and L. T. Wang, \[hep-ph\]; Q. H. Cao, F. P. Huang, K. P. Xie and X. Zhang, \[hep-ph\]; B. Jain, S. J. Lee and M. Son, \[hep-ph\]; J. de Vries, M. Postma, J. van de Vis and G. White, \[hep-ph\]; M. Reichert, A. Eichhorn, H. Gies, J. M. Pawlowski, T. Plehn and M. M. Scherer, \[hep-ph\]; M. Carena, Z. Liu and M. Riembau, \[hep-ph\]. N. Arkani-Hamed, T. Han, M. Mangano and L. T. Wang, Phys. Rept.  [**652**]{}, 1 (2016) \[arXiv:1511.06495 \[hep-ph\]\]. O. J. P. Eboli, G. C. Marques, S. F. Novaes and A. A. Natale, Phys. Lett. B [**197**]{}, 269 (1987); D. A. Dicus, C. Kao and S. S. D. Willenbrock, Phys. Lett. B [**203**]{}, 457 (1988); E. W. N. Glover and J. J. van der Bij, Nucl. Phys. B [**309**]{}, 282 (1988). T. Plehn, M. Spira and P. M. Zerwas, Nucl. Phys. B [**479**]{}, 46 (1996) Erratum: \[Nucl. Phys. B [**531**]{}, 655 (1998)\] \[\]; A. Djouadi, W. Kilian, M. Muhlleitner and P. M. Zerwas, Eur. Phys. J. C [**10**]{}, 45 (1999) \[\]; X. Li and M. B. Voloshin, Phys. Rev. D [**89**]{}, no. 1, 013012 (2014) \[ \[hep-ph\]\]. U. Baur, T. Plehn and D. L. Rainwater, Phys. Rev. D [**69**]{} (2004) 053004 \[\]. V. Barger, L. L. Everett, C. B. Jackson, A. D. Peterson and G. Shaughnessy, Phys. Rev. D [**90**]{}, no. 9, 095006 (2014) \[ \[hep-ph\]\]; A. Alves, T. Ghosh and K. Sinha, Phys. Rev. D [**96**]{}, no. 3, 035022 (2017) \[ \[hep-ph\]\]; A. Adhikary, S. Banerjee, R. K. Barman, B. Bhattacherjee and S. Niyogi, \[ \[hep-ph\]\]; J. H. Kim, Y. Sakaki and M. Son, \[ \[hep-ph\]\]. U. Baur, T. Plehn and D. L. Rainwater, Phys. Rev. D [**68**]{}, 033001 (2003) \[\]. M. J. Dolan, C. Englert and M. Spannowsky, JHEP [**1210**]{}, 112 (2012) \[ \[hep-ph\]\]; A. J. Barr, M. J. Dolan, C. Englert and M. Spannowsky, Phys. Lett. B [**728**]{}, 308 (2014) \[ \[hep-ph\]\]. A. Papaefstathiou, L. L. Yang and J. Zurita, Phys. Rev. D [**87**]{}, no. 1, 011301 (2013) \[ \[hep-ph\]\]. D. E. Ferreira de Lima, A. Papaefstathiou and M. Spannowsky, JHEP [**1408**]{}, 030 (2014) \[ \[hep-ph\]\]; D. Wardrope, E. Jansen, N. Konstantinidis, B. Cooper, R. Falla and N. Norjoharuddeen, Eur. Phys. J. C [**75**]{}, no. 5, 219 (2015) \[ \[hep-ph\]\]; J. K. Behr, D. Bortoletto, J. A. Frost, N. P. Hartland, C. Issever and J. Rojo, Eur. Phys. J. C [**76**]{}, no. 7, 386 (2016) \[ \[hep-ph\]\]. U. Baur, T. Plehn and D. L. Rainwater, Phys. Rev. Lett.  [**89**]{}, 151801 (2002) \[\]; U. Baur, T. Plehn and D. L. Rainwater, Phys. Rev. D [**67**]{}, 033003 (2003) \[\]; Q. Li, Z. Li, Q. S. Yan and X. Zhao, Phys. Rev. D [**92**]{}, no. 1, 014015 (2015) \[ \[hep-ph\]\]. ATLAS collaboration, ATLAS-CONF-2016-004. G. Aad [*et al.*]{} \[ATLAS Collaboration\], Phys. Rev. D [**92**]{}, 092004 (2015) \[x \[hep-ex\]\]. F. Goertz, A. Papaefstathiou, L. L. Yang and J. Zurita, JHEP [**1504**]{}, 167 (2015) \[ \[hep-ph\]\]; Q. H. Cao, B. Yan, D. M. Zhang and H. Zhang, Phys. Lett. B [**752**]{}, 285 (2016) \[ \[hep-ph\]\]; M. Gorbahn and U. Haisch, \[hep-ph\]; W. Bizon, M. Gorbahn, U. Haisch and G. Zanderighi, JHEP [**1707**]{}, 083 (2017) \[ \[hep-ph\]\]; G. Degrassi, P. P. Giardino, F. Maltoni and D. Pagani, \[hep-ph\]; S. Di Vita, C. Grojean, G. Panico, M. Riembau and T. Vantalon, JHEP [**1709**]{}, 069 (2017) \[ \[hep-ph\]\]. ATLAS Collaboration, ATL-PHYS-PUB-2017-001. CMS Collaboration, CMS-PAS-FTR-16-002. F. Kling, T. Plehn and P. Schichtel, Phys. Rev. D [**95**]{}, no. 3, 035026 (2017) \[ \[hep-ph\]\]. T. Golling [*et al.*]{}, CERN Yellow Report, no. 3, 441 (2017) \[ \[hep-ph\]\]; R. Contino [*et al.*]{}, CERN Yellow Report, no. 3, 255 (2017) \[ \[hep-ph\]\]. CEPC-SPPC Preliminary Conceptual Design Report, The CEPC-SPPC Study Group, IHEP- CEPC-DR-2015-01 (2015). A. J. Barr, M. J. Dolan, C. Englert, D. E. Ferreira de Lima and M. Spannowsky, JHEP [**1502**]{}, 016 (2015) \[ \[hep-ph\]\]; A. Azatov, R. Contino, G. Panico and M. Son, Phys. Rev. D [**92**]{}, no. 3, 035001 (2015) \[ \[hep-ph\]\]; A. Papaefstathiou, Phys. Rev. D [**91**]{}, no. 11, 113016 (2015) \[ \[hep-ph\]\]; H. J. He, J. Ren and W. Yao, Phys. Rev. D [**93**]{}, no. 1, 015003 (2016) \[ \[hep-ph\]\]; A. Papaefstathiou and K. Sakurai, JHEP [**1602**]{}, 006 (2016) \[ \[hep-ph\]\]; C. Y. Chen, Q. S. Yan, X. Zhao, Y. M. Zhong and Z. Zhao, Phys. Rev. D [**93**]{}, no. 1, 013007 (2016) \[ \[hep-ph\]\]; B. Fuks, J. H. Kim and S. J. Lee, Phys. Rev. D [**93**]{}, no. 3, 035026 (2016); \[ \[hep-ph\]\]; R. Contino [*et al.*]{}, \[hep-ph\]; Q. H. Cao, G. Li, B. Yan, D. M. Zhang and H. Zhang, Phys. Rev. D [**96**]{}, no. 9, 095031 (2017) \[arXiv:1611.09336 \[hep-ph\]\]; S. Banerjee, C. Englert, M. L. Mangano, M. Selvaggi and M. Spannowsky, \[hep-ph\]. C. R. Chen and I. Low, Phys. Rev. D [**90**]{}, no. 1, 013018 (2014) \[arXiv:1405.7040 \[hep-ph\]\]. S. Dawson, A. Ismail and I. Low, Phys. Rev. D [**91**]{}, no. 11, 115008 (2015) \[arXiv:1504.05596 \[hep-ph\]\]. K. Cranmer and T. Plehn, Eur. Phys. J. C [**51**]{}, 415 (2007), \[\]; T. Plehn, P. Schichtel and D. Wiegand, Phys. Rev. D [**89**]{}, no. 5, 054002 (2014) \[ \[hep-ph\]\]. S. Dawson, S. Dittmaier and M. Spira, Phys. Rev. D [**58**]{}, 115012 (1998) \[\]; J. Grigo, J. Hoff, K. Melnikov and M. Steinhauser, Nucl. Phys. B [**875**]{}, 1 (2013) \[ \[hep-ph\]\]; F. Maltoni, E. Vryonidou and M. Zaro, JHEP [**1411**]{}, 079 (2014) \[ \[hep-ph\]\]; S. Borowka, N. Greiner, G. Heinrich, S. P. Jones, M. Kerner, J. Schlenk, U. Schubert and T. Zirke, Phys. Rev. Lett.  [**117**]{}, no. 1, 012001 (2016) \[ \[hep-ph\]\]; S. Borowka, N. Greiner, G. Heinrich, S. P. Jones, M. Kerner, J. Schlenk and T. Zirke, JHEP [**1610**]{}, 107 (2016) \[ \[hep-ph\]\]. D. de Florian and J. Mazzitelli, JHEP [**1509**]{}, 053 (2015) \[ \[hep-ph\]\]; J. Grigo, J. Hoff and M. Steinhauser, Nucl. Phys. B [**900**]{}, 412 (2015) \[ \[hep-ph\]\]; D. de Florian, M. Grazzini, C. Hanga, S. Kallweit, J. M. Lindert, P. Maierhofer, J. Mazzitelli and D. Rathlev, \[hep-ph\]. J. Alwall, M. Herquet, F. Maltoni, O. Mattelaer and T. Stelzer, JHEP [**1106**]{}, 128 (2011) \[ \[hep-ph\]\]; J. Alwall [*et al.*]{}, JHEP [**1407**]{}, 079 (2014) \[ \[hep-ph\]\]. ATLAS Collaboration, ATL-PHYS-PUB-2016-026. S. Chatrchyan [*et al.*]{} \[CMS Collaboration\], JINST [**8**]{}, P04013 (2013) \[arXiv:1211.4462 \[hep-ex\]\]. M. Mangano, CERN Yellow Report CERN 2017-003-M \[arXiv:1710.06353 \[hep-ph\]\]. M. A. Shifman, A. I. Vainshtein, M. B. Voloshin and V. I. Zakharov, Sov. J. Nucl. Phys.  [**30**]{}, 711 (1979) \[Yad. Fiz.  [**30**]{}, 1368 (1979)\]; B. A. Kniehl and M. Spira, Z. Phys. C [**69**]{}, 77 (1995); M. Spira, \[hep-ph\]. for a non-trivial modification of $m_{hh}$ see M. Bauer, M. Carena and A. Carmona, \[hep-ph\]. J. Brehmer, K. Cranmer, F. Kling and T. Plehn, Phys. Rev. D [**95**]{}, no. 7, 073002 (2017) \[ \[hep-ph\]\]; J. Brehmer, F. Kling, T. Plehn and T. M. P. Tait, \[hep-ph\]. CMS Collaboration \[CMS Collaboration\], CMS-PAS-HIG-15-005. C. Vernieri \[CMS Collaboration\], \[hep-ex\]. T. Sjostrand, S. Mrenna and P. Z. Skands, Comput. Phys. Commun.  [**178**]{}, 852 (2008) \[ \[hep-ph\]\]. M. Mangano, “The so-called MLM prescription for ME/PS matching", presented at the Fermilab ME/MC Tuning Workshop, October 4, 2002. R. Frederix, S. Frixione, V. Hirschi, F. Maltoni, O. Mattelaer, P. Torrielli, E. Vryonidou and M. Zaro, Phys. Lett. B [**732**]{}, 142 (2014) \[ \[hep-ph\]\]. O. Brein, R. V. Harlander and T. J. E. Zirke, Comput. Phys. Commun.  [**184**]{}, 998 (2013) \[ \[hep-ph\]\]; F. Demartin, F. Maltoni, K. Mawatari, B. Page and M. Zaro, Eur. Phys. J. C [**74**]{}, no. 9, 3065 (2014) \[ \[hep-ph\]\]. J. de Favereau [*et al.*]{} \[DELPHES 3 Collaboration\], JHEP [**1402**]{}, 057 (2014) \[ \[hep-ex\]\]. M. Cacciari, G. P. Salam and G. Soyez, JHEP [**0804**]{}, 063 (2008) \[ \[hep-ph\]\]; M. Cacciari, G. P. Salam and G. Soyez, Eur. Phys. J. C [**72**]{}, 1896 (2012) \[ \[hep-ph\]\]. A. L. Read, J. Phys. G [**28**]{}, 2693 (2002). M. L. Mangano, T. Plehn, P. Reimitz, T. Schell and H. S. Shao, J. Phys. G [**43**]{}, no. 3, 035001 (2016) \[ \[hep-ph\]\]; for the HE-LHC this analysis is waiting to be done.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We report a detailed investigation of the interplay between size quantization and local scattering centers in graphene nanoribbons, as seen in the local density of states. The spectral signatures, obtained after Fourier transformation of the local density of states, include characteristic peaks that can be related to the transverse modes of the nanoribbon. In armchair ribbons, the Fourier transformed density of states of one of the two inequivalent sublattices takes a form similar to that of a quantum channel in a two-dimensional electron gas, modified according to the differences in bandstructure. After addition of the second sublattice contribution, a characteristic modulation of the pattern due to superposition is obtained, similar to what has been obtained in spectra due to single impurity scattering in large-area graphene. We present analytic results for the electron propagator in armchair nanoribbons in the Dirac approximation, including a single scattering center within a T-matrix formulation. For comparison, we have extended the investigation with numerics obtained with an atomistic recursive Green’s function approach. The spectral signatures of the atomistic approach include the effects of trigonal warping. The impurity induced oscillations in the local density of states are not decaying at large distance in few-mode nanoribbons.' author: - Anders Bergvall - Tomas Löfwander title: Spectral footprints of impurity scattering in graphene nanoribbons --- Introduction ============ In graphene, scattering centers such as impurities, defects, adatoms, and substrate inhomogeneities greatly influence the local electronic properties.[@CastroNeto:2009cl] In some samples, the material quality is so high that a single or a few such scattering centers can influence the whole device. This may degrade device function, but can also be taken advantage of by making various sensing devices.[@Schedin:2007hs; @Brar:2011iq] Great attention has therefore been focused on understanding the influence of scattering on the electronic properties of graphene.[@Peres:2010tn] In this context, the scanning tunneling microscope (STM) is becoming of increasing importance.[@Connolly:2010bc; @Deshpande:2012jg] By utilizing its various modes of operation, the STM can be used to map out topography, local density of states, local charge density, and more. In this way, a variety of properties of graphene have been revealed. A few examples include perturbations in the local density of states around impurities[@Rutter:2007ep; @Mallet:2007fg] or near step-edges in the substrate,[@Xue:2012hd] charge puddle formation caused by molecules trapped between graphene flakes and the SiO$_2$ substrate,[@Zhang:2009ce] and resistance caused by steps[@Ji:2011bla; @Wang:2012tj] in the substrate or multilayer regions[@Giannazzo:2012ib] in epitaxial graphene on silicon-carbide. At the same time, encouraging progress has been achieved with fabrication of graphene nanostructures. Top-down approaches include nano-lithography,[@Han:2007bl] scanning probe methods,[@Biro:2010ez] etching with metal nanoparticles along certain crystal directions,[@Campos:2009cd] and utilization of the transmission electron microscope (TEM) to simultaneously image and sculpture graphene.[@Girit:2009tz] A bottom up approach based on chemical synthesis has also been demonstrated.[@Cai:2010bd] Another approach involves unzipping of carbon nanotubes.[@Li:2008ht] With that method, the theoretically predicted zero-energy (midgap) edge states of nanoribbons with zigzag edges[@Fujita:1996vs; @Nakada:1996us] were directly mapped out by scanning tunneling spectroscopy (STS).[@Tao:2011bk] Theory also predicts that by controlling the width and edges of nanoribbons, a bandgap can be opened up at the Dirac point through quantum confinement (see the review ). With further progress it may soon become possible to study in much greater detail the interplay between quantum confinement and impurity scattering in graphene nanoribbons. Many theoretical studies of graphene nanoribbons have been reported in the literature, see the collection of review articles in Ref. . The effect of impurity scattering and the effects of edge disorder on electron transport have been reported in several numerical works. In an effort to simulate the typical experimental situation, random disorder is included and the scaling behavior of resistivity with length of the ribbon is studied, revealing different transport regimes depending on ribbon width and disorder properties. Here, we go back to the well defined problem of a single impurity in order to study in detail the effects on the FT-LDOS. In this paper we present results for the spectral signatures of a local scattering center in graphene, taking into account quantum confinement in a nanoribbon geometry. This study generalizes the consideration of FT-LDOS of a single impurity in bulk graphene[@PeregBarnea:2008ig; @Bena:2008iw] to the case of nanoribbons. We focus the analytic analysis on armchair ribbons in the Dirac approximation (linearization around the K-points in the graphene bandstructure), for which the wavefunctions and propagators for clean ribbons are known, and solve the impurity problem in a T-matrix formulation. Thereby we obtain the electron propagator for an armchair nanoribbon including the effects of a local scattering center. The Fourier transformed density of states (FT-LDOS) is then obtained and explained in terms of scattering processes of Dirac quasiparticles confined in the ribbon. We extend the analysis to an atomistic tight-binding model of graphene, utilizing a numerical recursive Green’s function approach. The main effect of going beyond the Dirac approximation is trigonal warping, which shows up as a triangular distorsion of the FT-LDOS patterns. For comparison we include an analysis of the FT-LDOS in a quantum ribbon in a two-dimensional electron gas (2DEG). Many features of the FT-LDOS patterns in graphene ribbons can be understood from the somewhat simpler case of a 2DEG, and the new features special for graphene can be highlighted. These include a more complicated bandstructure due to the two inequivalent K-points, trigonal warping, as well as interference effects due to the bipartite lattice of graphene. The outline of the paper is the following. In Section \[Ch\_FTSTS\] we discuss the Fourier transform scanning tunneling spectroscopy method and illustrate the basic scattering processes at play in a nanoribbon. In Section \[Ch\_2DEG\] we present results for the FT-LDOS in a 2DEG quantum channel. In Section \[Ch\_armchair\] we report our results for the FT-LDOS in an armchair graphene nanoribbon within the Dirac approximation and compare with the 2DEG case. In Section \[Ch\_numerics\] we present results of numerical simulations of a tight-binding model, including also zigzag nanoribbons as well as effects of edge disorder on the FT-LDOS. In Section \[Ch\_summary\] we summarize the paper and give some conclusions and an outlook. Most technical results of the analytic analysis have been collected in the Appendices. Fourier transform scanning tunneling spectroscopy {#Ch_FTSTS} ================================================= A scattering center induces a perturbation of the local density of states in its vicinity. For elastic scattering, the impurity scatters electrons between states ${\vec{k}}_1\rightarrow{\vec{k}}_2$ with $\epsilon_{{\vec{k}}_1}=\epsilon_{{\vec{k}}_2}$, i.e. on a contour of constant energy $E$. This leads to interference and a wave pattern in the local density of states near the impurity with wavevectors ${\vec{q}}={\vec{k}}_2-{\vec{k}}_1$. After Fourier transformation of the local density of states $\rho({\vec{r}},E)\rightarrow {\cal N}({\vec{q}},E)$, the wave vectors of the interference pattern are highlighted. The resulting pattern in ${\cal N}({\vec{q}},E)$ can then be used to infer the band dispersion $\epsilon_{{\vec{k}}}$. For instance, this has been done for metal surfaces.[@Anonymous:IEkMmlpc] This method has also become a valuable tool for probing the properties of high-$T_c$ superconductors.[@Balatsky:2006ce] It is worth mentioning that, neglecting electron-electron interactions, the interference patterns in the local density of states discussed above are related to the Friedel oscillations in the electron density $n({\vec{r}})$ through integration over energy including the Fermi-Dirac distribution function, $n({\vec{r}})=-e\int \rho({\vec{r}},E)\,f(E)dE$. By using the STM, the local density of states can be extracted as function of energy by applying a finite voltage between tip and sample, i.e. by employing scanning tunneling spectroscopy (STS). By combining Fourier transformation with STS, the band dispersion can be studied in the vicinity of the Fermi energy. This method has therefor become a valuable spectroscopic tool sometimes called Fourier transform scanning tunneling spectroscopy (FT-STS). In graphene, the Fermi energy itself is tunable by a back gate voltage on the substrate graphene is resting on. Thereby, FT-STS is potentially a valuable tool for studies of graphene. Indeed, experiment reproduce the graphene bandstructure.[@Mallet:2007fg] STS bears similarities with angle-resolved photo-emission (ARPES). STS is ideal for spatially inhomogeneous systems, while ARPES relies on large-area spatially homogeneous samples. Indeed, STS measures the spatially resolved spectral function, i.e. local density of states $\rho({\vec{r}};E)$, while ARPES measures the momentum-space spectral function $A({\vec{k}};E)$. By generalizing STS to FT-STS, i.e. Fourier transforming $\rho({\vec{r}};E)\rightarrow {\cal N}({\vec{q}};E)$, a spectroscopy has been introduced that can be used to study materials, although we should remember that ${\cal N}({\vec{q}};E)$ is not equal to $A({\vec{q}};E)$. One advantage of FT-STS is the possibility to study nanoscale systems with high spatial resolution. In this paper we will investigate the consequences of quantum confinement on impurity scattering in graphene, as seen in FT-STS. ![An electron quasiparticle initially in mode $m$, with longitudinal wave vector $\kappa_m$ may be backscattered by an impurity at ${\vec{r}}_i$ into mode $n$. The local density of states at ${\vec{r}}$ is changed due to interference of the initial and final waves. This leads to an interference pattern around ${\vec{r}}_i$.[]{data-label="Fig_handwave"}](handwave.eps){width="\columnwidth"} In Fig. \[Fig\_handwave\] we display a cartoon of a typical scattering process that contributes to the correction to the local density of states in a quantum ribbon with one impurity. For simplicity we here discuss the situation in a 2DEG quantum channel. Quasiparticles occupying for instance mode $m$, propagating in the positive $y$-direction with wavenumber $\kappa_m$, passes the probing position ${\vec{r}}=(x,y)$, after which they can be backscattered by the impurity at ${\vec{r}}_i$ into mode $n$ with wavenumber $\kappa_n$ and propagate back to the probing position ${\vec{r}}$. In this example we neglect evanescent modes for simplicity. The contribution to the full propagator from this scattering event will be proportional to the free propagators before and after scattering and the potential strength $\gamma$, $$\begin{aligned} \tilde G_{nm}({\vec{r}},{\vec{r}};E) &=& -i\frac{\mu}{\hbar^2} \frac{e^{i\kappa_n|y_i-y|}}{\kappa_n} \chi_n(x)\chi_n(x_i)\nonumber\\ &&\times\gamma \left(-i\frac{\mu}{\hbar^2}\right) \frac{e^{i\kappa_m|y-y_i|}}{\kappa_m} \chi_m(x)\chi_m(x_i),\nonumber\end{aligned}$$ where $\chi_n(x)=\sqrt{2/W}\sin(n\pi x/W)$ is the transverse wavefunction in mode $n\ge 1$, and $\mu$ is the electron effective mass. Taking into account multiple scattering by the impurity, the potential strength $\gamma$ is replaced by a $T$-matrix. When we take the imaginary part of the propagator to get the local density of states, we get spatially oscillating terms $$\begin{aligned} \propto \cos[(\kappa_n+\kappa_m)|y-y_i|],\nonumber\end{aligned}$$ and $$\begin{aligned} \propto \sin[(\kappa_n+\kappa_m)|y-y_i|],\nonumber\end{aligned}$$ since the $T$-matrix is a complex number due to multiple scattering. After Fourier transformation, we find peaks at $q_y=\pm(\kappa_n+\kappa_m)$ and at $q_x$ equal to combinations of $n$ and $m$ times $\pi/W$. Thus, in a FT-STS picture of a quantum channel, there will be a discrete number of peaks that reflect the available modes. We also note that the Friedel oscillations (neglecting electron-electron interactions) will at low temperature oscillate without decay far from the impurity site. To probe a 2DEG quantum channel with an STM in the way described here will be challenging since the channel is typically hidden deep down in a semiconducting heterostructure. Graphene, on the other hand, is 100 % surface and directly accessible. FT-LDOS: ribbon in a 2DEG {#Ch_2DEG} ========================= In this Section we improve the above discussion to the general case of multiple scattering in a multimode 2DEG quantum channel of width $W$ with a single impurity scattering center at ${\vec{r}}_i$. The results of this Section will be referenced in the following Sections on graphene in order to highlight the distinguishing features of confined Dirac quasiparticles. Consider the probability amplitude for an electron in the channel to propagate from one point ${\vec{r}^{\;\prime}}$ to another point ${\vec{r}}$. For free propagation in mode $n$, the amplitude is given by the free propagator (unperturbed Green’s function), $g_n({\vec{r}},{\vec{r}^{\;\prime}};E)$. In the presence of the impurity an electron initially in mode $m$ may be scattered into mode $n$. The effect of such an extra process will modify the propagator by adding a second term $$\tilde{G}_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E) = g_n({\vec{r}},{\vec{r}}_i;E)T({\vec{r}}_i;E)g_m({\vec{r}}_i,{\vec{r}^{\;\prime}};E),$$ so that the new Green’s function will be $$G_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E) = g_n({\vec{r}},{\vec{r}^{\;\prime}},E)\delta_{nm} + \tilde{G}_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E).$$ The factor $T({\vec{r}}_i;E)$, see Eq. (\[eqn:T\_2deg\]), includes multiple scattering by the impurity. The full probability amplitude for propagation from ${\vec{r}^{\;\prime}}$ to ${\vec{r}}$ is given by summing over all mode indices $$G({\vec{r}},{\vec{r}^{\;\prime}};E) = \sum_{nm}G_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E).$$ We may now proceed with the local density of states (LDOS). The correction to the LDOS by impurity scattering can be written as $$\begin{aligned} \tilde{\rho}({\vec{r}};E) &=& -\frac{1}{\pi} \sum_{nm}\text{Im}\,\tilde{G}_{nm}({\vec{r}},{\vec{r}};E)\nonumber\\ &=& -\frac{1}{\pi} \sum_{nm}\mathcal{K}_{nm}(E)\tilde{\rho}^x_{nm}(x;E)\tilde{\rho}^y_{nm}(y;E),\end{aligned}$$ where the expressions for the factors $\mathcal{K}_{nm}(E)$, $\tilde{\rho}^x_{nm}(x)$, and $\tilde{\rho}^y_{nm}(y;E)$ are given in Appendix \[Appendix\_2DEG\]. The Fourier transformed local density of states (FT-LDOS) can now be computed as $${\tilde{\mathcal{N}}}({\vec{q}};E) = -\frac{1}{\pi} \sum_{nm}\mathcal{K}_{nm}(E) {\tilde{\mathcal{N}}}^x_{nm}(q_x){\tilde{\mathcal{N}}}^y_{nm}(q_y;E), \label{Eq_FTLDOS}$$ where $$\label{eqn:Nqx} \begin{split} {\tilde{\mathcal{N}}}^x_{nm}(q_x) &= \sum_{l=-\infty}^\infty \delta\left(\frac{q_x}{\pi}-\frac{l}{W}\right)\int_{-W}^W\frac{dx}{2W}\;e^{-i\frac{\pi}{W}l x}\tilde{\rho}^x_{nm}(x) \\ & = \sum_{l=-\infty}^\infty \delta\left(\frac{q_x}{\pi}-\frac{l}{W}\right){\tilde{\mathcal{N}}}^x_{nm}(l) \end{split}$$ and $${\tilde{\mathcal{N}}}^y_{nm}(q_y;E) = \int_{-\infty}^\infty \frac{dy}{2\pi} e^{-iq_y y}\tilde{\rho}^y_{nm}(y;E).$$ The function $\tilde{\rho}^x_{nm}(x)$, originally defined on the interval $[0,W]$, is extended to $[-W,W]$ and assumed to be even with respect to the origin. Due to the finite width, $2W$, of the integration interval, the spectral $x$-component is fixed to be integer multiples of $\pi/W$. This is a trick to be able to resolve the minimum change of transverse momenta, $\pi/W$, when scattering between two different modes. It is important to realize that both propagating and evanescent modes play a role in this scattering problem. The longitudinal momentum is $\kappa_n = \sqrt{2\mu E/\hbar^2 - (n\pi/W)^2}$, where $\mu$ is the electron mass and $n\ge 1$ is the integer mode index. At the bottom of a subband, $\kappa_n\rightarrow 0$, and the evanescent mode extends far from the impurity and play an important role. On the other hand, for energies far from any subband bottom, the local density of states is only affected by the evanescent mode in a small region near the impurity. In the discussion of the FT-LDOS we can then safely neglect evanescent modes in the sums in Eq. (\[Eq\_FTLDOS\]). The evanescent modes are still taken into account in the scattering processes at the impurity through the T-matrix equation, where intermediate modes can be evanescent, while initial and final modes are propagating. In all of our numerical calculations, we include 10 evanescent modes. Adding even more evanescent modes does not qualitatively change our results. As have been shown, a delta-shaped impurity with a finite number of evanescent modes will model an s-like scatterer.[@2000PhRvB..61.5632B] We can now find the different components of the FT-LDOS to be $$\label{eqn:Nl} {\tilde{\mathcal{N}}}_{nm}^x(l) = \frac{1}{2W}\left(\delta_{l,n-m} + \delta_{-l,n-m} - \delta_{l,n+m} - \delta_{-l,n+m}\right)$$ and $${\tilde{\mathcal{N}}}_{nm}^y(q_y) = \frac{e^{-iy_iq_y}}{2\pi}\left[S^y(\kappa_n + \kappa_m + q_y) + S^y(\kappa_n + \kappa_m - q_y)\right],$$ where $$\label{eq:Sy_2DEG} S^y(a) = \lim_{\epsilon \rightarrow 0^+} \frac{\sigma_p \epsilon - a(1/\gamma + \sigma_e)}{\epsilon^2 + a^2}$$ and where $\sigma_{p/e}$ are positive, $\vec{q}$-independent constants defined in Eq.  (\[eqn:sigmas\_2deg\]). The factor $\mathcal{K}_{nm}(E)$ is given by $$\mathcal{K}_{nm}(E) = \frac{1}{(1/\gamma + \sigma_e(E))^2 + \sigma_p^2(E)} \left(\frac{\mu}{\hbar^2}\right)^2 \frac{\chi_n(x_i) \chi_m(x_i)}{\kappa_n(E) \kappa_m(E)}$$ and depends on the scatterer strength $\gamma$ and the transversal wave functions $\chi_n(x) = \sqrt{2/W}\sin(n\pi x/W)$. Together, these components give rise to a number of selection rules that govern the modification of the FT-LDOS by impurity scattering. To illustrate, we select a narrow channel ($W = 50a_0$, where $a_0$ defines the unit length) and a low energy ($E = 0.2\tau$, where $\tau$ defines the unit energy), such that only a total of three propagating modes are open. The scattering FT-LDOS $|{\tilde{\mathcal{N}}}(\vec{q};E)|$ for the case of the impurity in the middle of the ribbon ($x_i = W/2, y_i = 0$) is displayed in Fig. \[fig:2DEG\_ftldos\_5\](a). Since the scattering is elastic, energy conservation requires that the transverse and longitudinal momenta, both before and after scattering, satisfy the relation $2\mu E/\hbar^2 = k_x^2 + \kappa_n^2(E)$, which is the circle shown in Fig. \[fig:2DEG\_ftldos\_5\](b). In the channel, the transverse momentum is quantized, $k_x \rightarrow k_n = n\pi/W$, and the only allowed momentum values between which the electrons can scatter are indicated by dots and squares on this circle. The FT-LDOS is therefor non-zero only at a few, finite number of $\vec{q}$-points. All of of these points lie inside the dotted circle of radius $2\sqrt{2\mu E/\hbar^2}$ shown in Fig. \[fig:2DEG\_ftldos\_5\](a). The factor $\mathcal{K}_{nm}(E)$ will be non-zero only if the transverse wavefunctions of mode $n$ and $m$ have a finite overlap at the position of the impurity. Since we have positioned the impurity at $x_i = W/2$, $\mathcal{K}_{nm}(E)$ will in this example be non-zero only if $n$ and $m$ are both odd integers since all the wavefunctions with even indices will have a node at $x = x_i$. Thus, modes with even number $n$ are not scattered by the impurity in this example. To understand the exact locations of the $\vec{q}$-points, we start by looking at the case $q_x = 0$ (i.e., $l = 0$). Since all mode indices have to be odd, the term ${\tilde{\mathcal{N}}}^x(l=0)$ will be non-zero only when $n = m$, i.e., when ($n = 1, m = 1$) or when ($n = 3, m = 3$). This tells us that the points along $q_x = 0$ are all due to intraband scattering. The factor ${\tilde{\mathcal{N}}}^y(q_y;E)$ peaks when $q_y = \pm 2|\kappa_1|$ or when $q_y = \pm 2|\kappa_3|$. These are the four points we see along the line $l = 0$. When $q_x = \pi/W$ ($l = 1$), at least one of the indices $n$ and $m$ will be even, and the factor $\mathcal{K}_{nm}(E)$ is zero. This is why we see no bright points along this line. This also happens for $l=3$ and $l=5$. Along the line $q_x = 2\pi/W$, we have that ${\tilde{\mathcal{N}}}^x(l=2)$ is non-zero only when ($n = 1, m = 1$), ($n = 1, m = 3$) or when ($n = 3, m = 1$). The factor ${\tilde{\mathcal{N}}}^y(q_y;E)$ peaks at $q_y = \pm 2|\kappa_1|$ or when $q_y = \pm |\kappa_1 + \kappa_3|$, and we see that we have spots at these locations along $l = 2$ in the figure. At $l = 4$ we must have $(n = 3, m = 1)$ or $(n=1, m=3)$, which tells us that $q_y = \pm|\kappa_3+\kappa_1|$. At $l = 6$, we must have ($n = 3, m = 3$) and $q_y = \pm2|\kappa_3|$. A similar argument can be made for $l < 0$, and we can therefor say exactly which scattering processes contribute to each dark spot in Fig. \[fig:2DEG\_ftldos\_5\](a). If the impurity is not located exactly at the middle of the ribbon, the even subbands will also be part of the scattering process. This is illustrated in Fig. \[fig:2DEG\_ftldos\_numbered\], where we have numbered the subband transitions corresponding to each bright point. In Fig. \[fig:ftldos\_2degtb\], we show the result for a wider ribbon calculated both analytically and by doing a recursive tight-binding simulation. The parameters are adjusted such that both cases have 20 propagating modes open, and we see that the main features of our analytical calculation and the numerical simulation coincide. ![(a) FT-LDOS of a 2DEG ribbon of width $W = 50a_0$ at energy $E = 0.025\tau$, with three open (propagating) modes. (b) Energy contour of the 2DEG dispersion. The symbols indicate the allowed momentum values between which scattering can potentially take place (circles, diamonds and squares corresponds to $n=1$, $n=2$ and $n=3$ respectively). The two arrows in figure (b) illustrates two possible scattering processes that gives rise to the two encircled dots in figure (a). The impurity is placed at $x_i = W/2$ and the $n=2$ subband is not scattered by the impurity because the impurity has been located at a node of the corresponding transverse wavefunction.[]{data-label="fig:2DEG_ftldos_5"}](ftldos_a_2deg_50_002.eps "fig:"){width="0.49\columnwidth"} ![(a) FT-LDOS of a 2DEG ribbon of width $W = 50a_0$ at energy $E = 0.025\tau$, with three open (propagating) modes. (b) Energy contour of the 2DEG dispersion. The symbols indicate the allowed momentum values between which scattering can potentially take place (circles, diamonds and squares corresponds to $n=1$, $n=2$ and $n=3$ respectively). The two arrows in figure (b) illustrates two possible scattering processes that gives rise to the two encircled dots in figure (a). The impurity is placed at $x_i = W/2$ and the $n=2$ subband is not scattered by the impurity because the impurity has been located at a node of the corresponding transverse wavefunction.[]{data-label="fig:2DEG_ftldos_5"}](2deg_50_002_proc.eps "fig:"){width="0.49\columnwidth"} ![The first quadrant of the FT-LDOS of the same ribbon as in Fig. \[fig:2DEG\_ftldos\_5\], but with the impurity placed at $x_i = 2W/7$. The numbers indicate the modes that gives rise to the different points.[]{data-label="fig:2DEG_ftldos_numbered"}](ftldos_a_2deg_50_002_quadrant1.eps){width="\columnwidth"} ![(a) Analytical FT-LDOS of a 2DEG ribbon of width $W = 200a_0$ and energy $E = 0.05\tau$. (b) FT-LDOS taken from a numerical tight-binding simulation of a ribbon of width $W = 200a_0$ and energy $E = -3.95\tau$. The energies are adjusted such that each ribbon has a total of 20 propagating modes open.[]{data-label="fig:ftldos_2degtb"}](ftldos_a_2deg_200_005_full.eps "fig:"){width="0.49\columnwidth"} ![(a) Analytical FT-LDOS of a 2DEG ribbon of width $W = 200a_0$ and energy $E = 0.05\tau$. (b) FT-LDOS taken from a numerical tight-binding simulation of a ribbon of width $W = 200a_0$ and energy $E = -3.95\tau$. The energies are adjusted such that each ribbon has a total of 20 propagating modes open.[]{data-label="fig:ftldos_2degtb"}](ftldos_tb_2deg_200_005_full.eps "fig:"){width="0.49\columnwidth"} FT-LDOS: ribbons of graphene {#Ch_armchair} ============================ The procedure of calculating the effect of a single impurity on the local density of states in a graphene armchair nanoribbon (AGNR) much follows that used for the 2DEG case. Due to the bipartite structure of the graphene honeycomb lattice, the propagator $\tilde{\mathbf{G}}({\vec{r}},{\vec{r}^{\;\prime}};E)$ is a 2x2 matrix in sublattice space, denoted by A and B. We therefor start by finding the impurity contribution to the local density of states on each sublattice. The resulting expressions for the A- and B-sublattice LDOS can be written as $$\begin{split} \tilde{\rho}^{A/B}(\vec{r};E) &= -\frac{1}{\pi}\sum_{dc}\sum_{nm} \text{Im}\;\tilde{G}^{AA/BB}_{dncm}(\vec{r},\vec{r};E) \\ &= -\frac{1}{\pi}\sum_{dc}\sum_{nm} \mathcal{K}^{A/B}_{dncm}(E)\tilde{\rho}^x_{nm}(x;E)\tilde{\rho}^{(A/B)y}_{dncm}(y;E), \end{split}$$ where $\tilde{G}^{AA/BB}_{dncm}({\vec{r}},{\vec{r}};E)$ are the two diagonal components of the propagator matrix $\tilde{\mathbf{G}}_{dncm}({\vec{r}},{\vec{r}};E)$. The summation over the variables $c$ and $d$ are added to account for scattering between different sets of non-equivalent Dirac cone pairs $\vec{K}^{\pm}_c$ and $\vec{K}^{\pm}_d$. A further elaboration on this is found in Appendix \[Appendix\_armchair\], together with derivations of the expressions for $\tilde{\rho}^x_{nm}(x;E)$ and $\tilde{\rho}^{(A/B)y}_{dncm}(y;E)$. As discussed in section \[Ch\_2DEG\], we only need to sum over propagating incoming and final transverse modes, labeled by $m$ and $n$. We compute the FT-LDOS on each sublattice as $$\label{eq:M} \tilde{\mathcal{M}}^{A/B}(\vec{q};E) = -\frac{1}{\pi}\sum_{nm}\tilde{\mathcal{M}}^x_{nm}(q_x;E) \tilde{\mathcal{M}}_{nm}^{(A/B)y}(q_y;E), $$ The total FT-LDOS is found as a superposition of the two sublattices $$\label{eqn:Ntot_agnr} \tilde{\mathcal{M}}(\vec{q};E) = \tilde{\mathcal{M}}^A(\vec{q};E) + e^{-i a_0 q_y}\tilde{\mathcal{M}}^B(\vec{q};E),$$ where the extra phase-shift is introduced since the two sublattices are spatially separated by the carbon-carbon distance $a_0$ in the y-direction. Since the transverse wavefunctions, $\chi_{n}(x) = \sqrt{1/W}\sin(n\pi/Wx)$, in our AGNR only differ from those of the 2DEG by a factor of $1/\sqrt{2}$, we have that $\tilde{\mathcal{M}}^x_{nm}(q_x;E) = \tilde{\mathcal{N}}^x_{nm}(q_x;E)/2$, as defined in Eqns. (\[eqn:Nqx\]) and (\[eqn:Nl\]). The longitudinal FT-LDOS expressions for each sublattice are given by $$\label{eqn:My} \begin{split} \tilde{\mathcal{M}}^{(A/B)y}_{nm}(q_y;E) &= \frac{e^{-iq_yy_i}}{2\pi}\sum_{c=1}^3 \mathcal{K}^{A/B}_{cncm}(E)\left[S^{(A/B)y}(\Delta_{cncm}(E) - q_y)^* + S^{(A/B)y}(\Delta_{cncm}(E) + q_y)\right] \\ &+ \frac{e^{-iq_yy_i}}{2\pi}\sum_{d=1}^2\sum_{c=d+1}^3 \mathcal{K}^{A/B}_{dncm}(E)\left[S^{(A/B)y}(\Delta_{cndm}(E)- q_y)^* + S^{(A/B)y}(\Delta_{dncm}(E)-q_y)^* \right.\\ &\left. \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \; + S^{(A/B)y}(\Delta_{cndm}(E) + q_y) + S^{(A/B)y}(\Delta_{dncm}(E)+q_y) \right], \end{split}$$ where $\Delta_{dncm}(E) = \text{sgn}(E)[\kappa_{dn}(E) + \kappa_{dm}(E)] + K^y_d - K^y_c$. The two $\vec{q}$-independent constants are found to be $$\mathcal{K}^A_{dncm}(E) = \frac{1}{\left[1/\gamma + \sigma_e(E)\right]^2 + \sigma_p^2(E)} \left(\frac{|E|}{v_f^2}\right)^2\frac{\chi_n(x_i)\chi_m(x_i)}{\kappa_{dn}(E)\kappa_{cm}(E)}$$ and $\mathcal{K}^B_{dncm}(E) = -(v_f/|E|)^2 \mathcal{K}^A_{dncm}(E)$. The transverse and longitudinal momenta are now cone set dependent, and changes to $k_{dn} = n\pi/W - K_{dx}$ and $\kappa_{dn}(E) = \sqrt{(E/v_f)^2 - k_{dn}^2}$ respectively. Here, $S^{Ay}(a) = S^y(a)$ is the same function as used in the 2DEG case and defined in Eq. (\[eq:Sy\_2DEG\]), and $S^{By}(a) = f_{dncm}(E) S^{y}(a)$ where $f_{dncm}(E) = [-k_nk_m + \kappa_n(E)\kappa_m(E)] + i\text{sgn}(E)[k_n\kappa_m(E) + k_m\kappa_n(E)]$. The $S$-terms for the AGNR A-lattice have the exact same form as the corresponding terms in the 2DEG, while the B-lattice terms are scaled by a complex mode dependent prefactor. This is a consequence of our choice of impurity potential: an impurity fully localized on one A-atom, see Eq. (\[eq:Vgr\]). In Fig. \[fig:ftldos\_full\], we plot $|\tilde{\mathcal{M}}(\vec{q};E)|$ for a semiconducting AGNR of width $W \approx 50$ nm, where $E=0.4|\tau|$ such that 60 channels are propagating. The positions, outer shapes, and sizes of the circular features (of radius $2E/v_f$) are the same as those found when studying a sharp impurity in bulk graphene, and are due to the graphene bandstructure. In addition to the bulk graphene features, we also see an added rich inner structure due to transverse confinement in the nanoribbon. Each peak corresponds to scattering processes that change the transverse momentum by integer multiples of $\pi/W$, and change the longitudinal momentum such that the arguments of at least one of the many $S^y$-terms in Eq. (\[eqn:My\]) vanishes. ![Analytic FT-LDOS, ${\tilde{\mathcal{M}}}(\vec{q};E)$, of an AGNR having width $W \approx 50$, $E = 0.4|\tau|$, $\gamma = 10|\tau|$ and 60 propagating channels.[]{data-label="fig:ftldos_full"}](ftldos_a_agnr_202_04_full.eps){width="0.5\columnwidth"} A schematic illustration of one such scattering process is shown in Fig. \[fig:agnr\_proc\]. An electron, initially in cone pair $c=1$ and mode $m$, is described by a plain wave with momentum $-\kappa_{1m}$ in the longitudinal direction, and a superposition of two plain waves with momenta $\pm k_{1m}$ in the transverse direction (see the lower red dots). After scattering (within the same cone pair) to mode $n$, the momenta are changed to $\kappa_{1n}$ and $\pm k_{1n}$ in the longitudinal and transverse directions respectively (see the upper green squares). A Fourier transform of the LDOS is proportional to a product of the electron wavefunction before and after the scattering event, where each wavefunction is a linear combination of two transverse parts. The FT-LDOS will therefor be finite at the $\vec{q}$-values corresponding to the four arrows shown in the figure. Here, $q_y = |\kappa_{1n}+\kappa_{1m}|$, and $q_x = (m\pm n)\pi/W$ (solid arrows) or $q_x = -(m\pm n)\pi/W$ (dotted arrows). When scattering to a different cone pair $d\neq c$, we instead have $q_y = |\Delta_{dn1m}|$, see Eq. (\[eqn:My\]). ![Schematic picture of one possible scattering process in AGNR’s. Here, the electron (initially in mode $m$, represented by the red dots), is scattered into mode $n$ (green squares). The FT-LDOS will be finite at the $\vec{q}$-values illustrated by the solid ($q_x = (m\pm n)\pi/W$) and dotted ($q_x = -(m\pm n)\pi/W$) arrows.[]{data-label="fig:agnr_proc"}](agnr_proc.eps){width="0.5\columnwidth"} When we zoom in on the circular feature in the middle \[shown in Fig. \[fig:ftldos\_middle\](a)\] we see that the outer ring of non-vanishing $\vec{q}$-points in $|\tilde{\mathcal{M}}(\vec{q};E)|$ appears to be attenuated compared with what is seen on e.g. the A-lattice alone \[$|\tilde{\mathcal{M}}^A(\vec{q};E)|$ shown in Fig. \[fig:ftldos\_middle\](b)\]. This is due to destructive interference when adding the A- and B-lattice FT-LDOS contribution together, as done in Eq. (\[eqn:Ntot\_agnr\]). The $\vec{q}$-points on the outer circle comes from scattering processes which maximimize the change in momenta while still scattering within the same cone pair, i.e., where $d=c$ and $k_{cm} \rightarrow -k_{dn}$ and vice versa. In this case, we have that $f_{dncm} = (E/v_f)^2$ which tells us that ${\tilde{\mathcal{M}}}^{By}_{nm}(q_y;E) = -{\tilde{\mathcal{M}}}^{Ay}(q_y;E)$ so that when the phase factor $e^{-iq_ya}$ in Eq. (\[eqn:Ntot\_agnr\]) is close to unity, the contributions from the A- and B-lattice will cancel each other out. Similar cancellations may be seen in Fig \[fig:ftldos\_full\], e.g. in the circular features to right and left of the central one. For other processes and $\vec{q}$-values, the interference between the two lattice contributions may not play an important role, or we might have constructive interference instead. ![Zoom-ins of the central circular feature of Fig. \[fig:ftldos\_full\] showing (a) the combined A- and B-lattice contributions and (b) the A-lattice alone. When adding the two lattices together, the outer circle is attenuated by destructive interference between the two lattice contributions, ${\tilde{\mathcal{M}}}^{(A/B)y}(q_y;E)$, to the total FT-LDOS.[]{data-label="fig:ftldos_middle"}](ftldos_a_agnr_202_04_middle_AB.eps "fig:"){width="0.49\columnwidth"} ![Zoom-ins of the central circular feature of Fig. \[fig:ftldos\_full\] showing (a) the combined A- and B-lattice contributions and (b) the A-lattice alone. When adding the two lattices together, the outer circle is attenuated by destructive interference between the two lattice contributions, ${\tilde{\mathcal{M}}}^{(A/B)y}(q_y;E)$, to the total FT-LDOS.[]{data-label="fig:ftldos_middle"}](ftldos_a_agnr_202_04_middle_A.eps "fig:"){width="0.49\columnwidth"} The FT-LDOS is left-right mirror symmetric around the line $q_x = 0$, see Fig. \[fig:ftldos\_full\]. This symmetry appears because for every process adding a component $\vec{q}^+$ in the FT-LDOS, there is another process adding a component $\vec{q}^-$, where $q_x^- = -q_x^+$, see the solid arrows in Fig. \[fig:agnr\_proc2\]. After summation of all such processes, the FT-LDOS acquires the left-right symmetry. As a consequence, the feature centered around $\vec{q}=0$ is always mirror symmetric by the above argument . On the other hand, there is not necessarily a mirror symmetry within the other circular features (i.e. mirror symmetry with respect to the individual cone centers). For metallic AGNRs, the transverse modes are constructed from wavevectors symmetrically positioned with respect to the cone center (plus the metallic mode at the cone center). See, for instance, the two wavefectors $k_{dn}$ and $k_{dn'}=-k_{dn}$ in Fig. \[fig:agnr\_proc2\]. For semiconducting AGNRs, the wavevectors are not symmetrically positioned with respect to the cone center, i.e. $k_{dn'}\neq-k_{dn}$ for any $n'$. Therefor, the inner structure of the circular features centered at finite $\vec{q}$ are symmetric for metallic AGNRs and asymmetric for semiconducting AGNRs. This is illustrated in Fig. \[fig:ftldos\_topright\] for the semiconducting and metallic cases in (a) and (b), respectively. We conclude that by looking at what symmetries there are in the FT-LDOS, one can extract information about whether or not an AGNR is metallic or not. ![Zoom-ins of the north-eastern circular feature in Fig. \[fig:ftldos\_full\]. In (a), the AGNR is semiconducting and the left- and right-hand side is not mirror-symmetric. In (b), the ribbon is made metallic by removing 4 rows of carbon atoms, which restores the left-right symmetry again.[]{data-label="fig:ftldos_topright"}](ftldos_a_agnr_202_04_topright_AB.eps "fig:"){width="0.49\columnwidth"} ![Zoom-ins of the north-eastern circular feature in Fig. \[fig:ftldos\_full\]. In (a), the AGNR is semiconducting and the left- and right-hand side is not mirror-symmetric. In (b), the ribbon is made metallic by removing 4 rows of carbon atoms, which restores the left-right symmetry again.[]{data-label="fig:ftldos_topright"}](ftldos_a_agnr_200_04_topright_AB.eps "fig:"){width="0.49\columnwidth"} ![Schematic picture of scattering processes conserving mirror symmetry with respect to $q_x = 0$ (solid arrows) and with respect to $q_x = \pm K_{dx}$ (solid + dotted arrows).[]{data-label="fig:agnr_proc2"}](agnr_proc2.eps){width="0.49\columnwidth"} Numerical simulations of the tight-binding model {#Ch_numerics} ================================================ For our numerical simulations, we use a tight-binding model described by the Hamiltonian $$H = \sum_{i=1}^N \epsilon_i c^\dagger_i c_i + \sum_{i\neq j}^N \tau_{ij}c^\dagger_i c_j,$$ where $c^\dagger_i$ and $c_i$ are creation and destruction operators for site $i$. The onsite energy of site $i$ is denoted $\epsilon_i$, and the hopping amplitude between sites $j$ and $i$ is denoted $\tau_{ij}$. The number of atoms in the system is denoted $N$. We assume that $\tau_{ij}$ is always zero except when the sites $i$ and $j$ are nearest neighbours. The retarded Green’s function matrix is defined as $$\mathbf{G}(E) = \left[(E + i\eta)\mathbf{1} - \mathbf{H}\right]^{-1},$$ where $\eta$ is a small positive number. Even though the Hamiltonian is sparse, when written down as a matrix in site index space, direct inversion is not a viable alternative when the number of atoms $N$ grows large. Instead of direct matrix inversion, we use our own implementation of a recent algorithm[@Kazymyrenko:2008hk] in which the system atoms are added one-by-one, in a recursive manner. This allows us to save both memory and time, and once we have found all the retarded propagators between the system leads and atom $i$ we can calculate the lesser Green’s function, defined as $$G^<_{ii}(E) = \sum_l f_l(E) \sum_{\alpha_l \beta_l} G_{i\alpha_l}(E) \left[\Sigma^\dagger_l(E) - \Sigma_l(E)\right]_{\alpha_l \beta_l} G^\dagger_{\beta_l i}(E),$$ where $l$ is the lead number ($l = 1, 2$ in the case of a simple ribbon), and $\alpha_l$ and $\beta_l$ are indices running over all atoms belonging to the surface of lead $l$. Here, $f_l(E)$ and $\Sigma_l(E)$, are the distrubution function and the self-energy of lead $l$, respectively. The local density of states on atom $i$ is found from $$\rho_{i}(E) = -\frac{1}{\pi}\text{Im}G^<_{ii}(E),$$ and the FT-LDOS is given by doing a discrete Fourier transform over all system atoms, $$\mathcal{N}(\vec{q};E) = \frac{1}{N} \sum_{i=1}^N e^{-i \vec{r}_i \cdot \vec{q}} \rho_{i}(E),$$ where $\vec{r}_i$ is the real space coordinate vector of atom $i$. In Fig. \[fig:ftldos\_tb\](a), the result of such a tight-binding simulation is shown for a ribbon and setup matching the one used in Fig. \[fig:ftldos\_full\], with a delta-like impurity placed in the middle ($W \approx 50$ nm, 60 propagating channels and $x_i = W/2$). Upon inspection, we notice that the general features are similar compared with our analytical results. Some points, such as the outline of the central circle, are attenuated. The tight-binding ribbon do, however, show clear signs of trigonal warping due to the dispersion not being perfectly linear. In Fig. \[fig:ftldos\_tb\](b), we have moved the impurity to the edge of the ribbon and we notice that the resulting FT-LDOS image is not very different from the one with the impurity in the middle of the ribbon. In Fig. \[fig:ftldos\_tb\](c), we have made the impurity more gaussian shaped (long-range), which leads to suppressed scattering and attenuated features. For bulk graphene, it is well known that a long range impurity can not scatter between valleys. In the FT-LDOS, the features centered at $\vec{q}=\vec{K}_d^{pm}$ are then absent. This is not the case here, since the armchair nanoribbon has only one cone in its band structure.[@Wakabayashi:2009dh] ![Numerical tight-binding FT-LDOS of three AGNRs’ ($N = 810$ atoms in the unit cell) with different impurity configurations. (a) Single impurity, (b) edge impurity, (c) smooth (Gaussian long-range) impurity, where $W \approx 50$ nm, $\gamma = 10|\tau|$, $E = 0.4|\tau|$ and 60 propagating channels.[]{data-label="fig:ftldos_tb"}](ftldos_tb_agnr_202_04_single_AB.eps "fig:"){width="0.4\columnwidth"} ![Numerical tight-binding FT-LDOS of three AGNRs’ ($N = 810$ atoms in the unit cell) with different impurity configurations. (a) Single impurity, (b) edge impurity, (c) smooth (Gaussian long-range) impurity, where $W \approx 50$ nm, $\gamma = 10|\tau|$, $E = 0.4|\tau|$ and 60 propagating channels.[]{data-label="fig:ftldos_tb"}](ftldos_tb_agnr_202_04_edge_AB.eps "fig:"){width="0.4\columnwidth"} ![Numerical tight-binding FT-LDOS of three AGNRs’ ($N = 810$ atoms in the unit cell) with different impurity configurations. (a) Single impurity, (b) edge impurity, (c) smooth (Gaussian long-range) impurity, where $W \approx 50$ nm, $\gamma = 10|\tau|$, $E = 0.4|\tau|$ and 60 propagating channels.[]{data-label="fig:ftldos_tb"}](ftldos_tb_agnr_202_04_smooth_AB.eps "fig:"){width="0.4\columnwidth"} In Fig. \[fig:ftldos\_tbz\], we present results for the FT-LDOS of zigzag graphene nanoribbons (ZGNRs). In this simulation the ribbon has $N = 468$ atoms in its unit cell ($W \approx 50$ nm), $\gamma = 10\tau$ and $E = 0.45|\tau|$. This gives $35$ propagating modes. In Fig. \[fig:ftldos\_tbz\](a), the impurity is located in the middle of the ribbon and we see a pattern very similar to that of the same impurity configuration in an armchair ribbon, but with all features rotated $90$ degrees due to the different ribbon alignment (for ZGNRs’, $k_y$ is quantized instead). The result of a single impurity on the edge is shown in Fig. \[fig:ftldos\_tbz\](b), and in Fig. \[fig:ftldos\_tbz\](c) we show the spectra for a ribbon also having rough edges. In the last figure, Fig. \[fig:ftldos\_tbz\] (d), we have used a gaussian shaped (long-range) impurity, and we here see clearly that inter-valley scattering is now fully supressed. Indeed, since the ZGNR has two cones in its bandstructure, this case is similar to bulk graphene. ![Numerical tight-binding FT-LDOS of four ZGNRs’ ($N = 468$ atoms in the unit cell) with different impurity configurations. (a) Single impurity, (b) edge impurity, (c) rough edges, (d) smooth impurity (Gaussian long-range), where $W \approx 50$ nm, $E = 0.45|\tau|$ $\gamma = 10|\tau|$, and $35$ propagating channels.[]{data-label="fig:ftldos_tbz"}](ftldos_tb_zgnr_234_045_single_AB.eps "fig:"){width="0.4\columnwidth"} ![Numerical tight-binding FT-LDOS of four ZGNRs’ ($N = 468$ atoms in the unit cell) with different impurity configurations. (a) Single impurity, (b) edge impurity, (c) rough edges, (d) smooth impurity (Gaussian long-range), where $W \approx 50$ nm, $E = 0.45|\tau|$ $\gamma = 10|\tau|$, and $35$ propagating channels.[]{data-label="fig:ftldos_tbz"}](ftldos_tb_zgnr_234_045_edge_AB.eps "fig:"){width="0.4\columnwidth"} ![Numerical tight-binding FT-LDOS of four ZGNRs’ ($N = 468$ atoms in the unit cell) with different impurity configurations. (a) Single impurity, (b) edge impurity, (c) rough edges, (d) smooth impurity (Gaussian long-range), where $W \approx 50$ nm, $E = 0.45|\tau|$ $\gamma = 10|\tau|$, and $35$ propagating channels.[]{data-label="fig:ftldos_tbz"}](ftldos_tb_zgnr_234_045_rough_AB.eps "fig:"){width="0.4\columnwidth"} ![Numerical tight-binding FT-LDOS of four ZGNRs’ ($N = 468$ atoms in the unit cell) with different impurity configurations. (a) Single impurity, (b) edge impurity, (c) rough edges, (d) smooth impurity (Gaussian long-range), where $W \approx 50$ nm, $E = 0.45|\tau|$ $\gamma = 10|\tau|$, and $35$ propagating channels.[]{data-label="fig:ftldos_tbz"}](ftldos_tb_zgnr_234_045_smooth_AB.eps "fig:"){width="0.4\columnwidth"} Summary {#Ch_summary} ======= In summary, we have presented results for the FT-LDOS of graphene nanoribbons with local scattering centers. The interplay between size quantization and scattering leads to characteristic peaks that can be related to the transverse modes of the nanoribbon. The main features include ring-like structures, analogous to the case of an infinite 2D graphene sheet with a single scattering center. Inside the ring-like structure, new peaks appear that are related to inter and intra band scattering in the ribbon. We have presented analytic results for the electron propagator in armchair nanoribbons in the Dirac approximation, including a single scattering center within a T-matrix formulation. We have also extended the investigation with numerics obtained with an atomistic recursive Green’s function approach. The spectral signatures of the atomistic approach include the lifting of degeneracies of transverse modes in the Dirac approximation, as well as effects of trigonal warping. The impurity induced oscillations in the local density of states are not decaying at large distance in few-mode nanoribbons. Acknowledgements ================ This work has been supported by SSF, the Swedish Foundation for Strategic Research, and the EU through the FP7 project ConceptGraphene.\ Nanoribbon in a 2DEG {#Appendix_2DEG} ==================== Unperturbed Green’s function ---------------------------- For a 2DEG confined in the x-direction, creating a ribbon of width $W$, the wave functions can be written as $$\phi_n({\vec{r}}) = e^{ik_yy}\chi_n(x),$$ where ${\vec{r}}= (x,y)$ and $n$ is the mode number associated with the transverse eigenfunctions (assuming infinitely high confining walls at $x=0$ and $x=W$) given by $$\chi_n(x) = \sqrt{\frac{2}{W}}\sin(k_nx),$$ with the corresponding eigenenergies $$\epsilon_n(k_y) = \frac{\hbar^2}{2\mu}(k_n^2 + k_y^2).$$ Here, $k_n = n\pi/W$ is the transverse momentum, $k_y$ the longitudinal momentum and $\mu$ the electron mass. Using these wave functions we may construct the free propagator, or Green’s function, of an electron (having energy $E^+ = E + i\eta$, in the limit $\eta \rightarrow 0^+$) between the points ${\vec{r}^{\;\prime}}$ and ${\vec{r}}$, in mode $n$, as $$\begin{split} g_n({\vec{r}},{\vec{r}^{\;\prime}};E) &= \int_{-\infty}^\infty \frac{dk_y}{2\pi} \frac{\phi_n({\vec{r}})\phi_n^*({\vec{r}^{\;\prime}})}{E^+ - \epsilon_n(k_y)} \\ &= \chi_n(x)\chi_n(x^\prime) \Gamma_n(y,y^\prime;E), \end{split}$$ where $$\label{eq:gamma1} \Gamma_n(y,y^\prime;E) = \int_{-\infty}^\infty \frac{dk_y}{2\pi} \frac{e^{ik_y(y-y^\prime)}}{E^+ - \frac{\hbar^2}{2\mu}(k_y^2 + k_n^2)}.$$ This integral can be evaluated using standard contour integration techniques,[@Eco] giving us that $$\Gamma_n(y,y^\prime;E) = \begin{cases} -i\frac{\mu}{\hbar^2}\frac{e^{i\kappa_n(E)|y-y^\prime|}}{\kappa_n(E)} & \text{if} \quad E > E_n, \\ -\frac{\mu}{\hbar^2}\frac{e^{-\kappa_n(E)|y-y^\prime|}}{\kappa_n(E)} & \text{if} \quad E < E_n, \end{cases}$$ and $$\kappa_n(E) = \sqrt{\frac{2\mu}{\hbar^2}|E - E_n|},$$ where $E_n = (\hbar^2/2\mu)k_n^2$. Green’s function, one impurity ------------------------------ We introduce a single impurity modelled by an impurity potential with matrix elements $V_{nm}({\vec{r}})$. The perturbed propagator for an electron going from position ${\vec{r}^{\;\prime}}$ to ${\vec{r}}$, while changing mode from $m$ to $n$, can then be written using the Dyson equation as[@Bagwell:2012te] $$\begin{split} G_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E) &= g_n({\vec{r}},{\vec{r}^{\;\prime}};E)\delta_{nm} \\ & + \underbrace{\sum_l g_n({\vec{r}},{\vec{r}}_i;E) V_{nl}({\vec{r}}_i) G_{lm}({\vec{r}}_i,{\vec{r}^{\;\prime}};E)}_{=\tilde{G}_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E)} \\ & = g_n({\vec{r}},{\vec{r}^{\;\prime}};E)\delta_{nm} + \tilde{G}_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E). \end{split}$$ Here we assume that the impurity is positioned at ${\vec{r}}_i = (x_i,y_i)$ and that its potential is highly localized (delta-function shaped) so that all matrix elements but $V_{nm}({\vec{r}}_i)$ are zero. After introducing $$T_{nm}({\vec{r}}_i;E) = V_{nm}({\vec{r}}_i) + \sum_l V_{nl}({\vec{r}}_i) g_l({\vec{r}}_i,{\vec{r}}_i;E) T_{lm}({\vec{r}}_i;E),$$ the scattering part of the Dyson equation can be rewritten on the T-matrix form $$\label{eq:2DEGT} \tilde{G}_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E) = g_n({\vec{r}},{\vec{r}}_i;E)T_{nm}({\vec{r}}_i;E)g_m({\vec{r}}_i,{\vec{r}^{\;\prime}};E),$$ where $$T_{nm}({\vec{r}}_i;E) = V_{nm}({\vec{r}}_i) + \sum_lV_{nl}({\vec{r}}_i)g_l({\vec{r}}_i,{\vec{r}}_i;E)T_{lm}({\vec{r}}_i;E).$$ Since the impurity is highly localized in position space, we may further assume that it will scatter equally between all different modes $n$ and $m$ and we have that $V_{nm}({\vec{r}}_i) = V({\vec{r}}_i) = \gamma$ where $\gamma$ is the impurity strength. Using this assumption, it follows that $T_{nm}({\vec{r}};E) = T({\vec{r}};E)$ and we find that $$\label{eqn:T_2deg} \begin{split} T({\vec{r}}_i;E) &= V({\vec{r}}_i) + V({\vec{r}}_i)\left[\sum_lg_l({\vec{r}}_i,{\vec{r}}_i;E)\right]T({\vec{r}}_i;E) \\ &= \frac{\gamma}{1 - \gamma \sum_l g_l({\vec{r}}_i,{\vec{r}}_i;E)} \\ &= \frac{1}{1/\gamma + \sigma_e(E) + i\sigma_p(E)}, \end{split}$$ where $$\label{eqn:sigmas_2deg} \sigma_{e/p}(E) = \frac{\mu}{\hbar^2} \sum_{l\in e/p} \frac{\chi^2_l(x_i)}{\kappa_l(E)}$$ and $e$ and $p$ are the sets of all evanescent ($E < E_l$) and propagating ($E > E_l$) modes. Inserting the above expression for $T({\vec{r}}_i;E)$ back into Eq. (\[eq:2DEGT\]) allows us to solve for $\tilde{G}_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E)$ and consequently for $G_{nm}({\vec{r}},{\vec{r}^{\;\prime}};E)$. Fourier transformed density of states ------------------------------------- Once the perturbed propagator is known, the change in the local density of states (LDOS) due to scattering is given by $$\begin{split} \tilde{\rho}({\vec{r}};E) &= -\frac{1}{\pi}\sum_{nm}{\textrm{Im}}\left[\tilde{G}_{nm}({\vec{r}},{\vec{r}};E)\right] \\ &= -\frac{1}{\pi}\sum_{nm} \mathcal{K}_{nm}(E)\tilde{\rho}^x_{nm}(x) \tilde{\rho}^y_{nm}(y;E), \end{split}$$ where $$\mathcal{K}_{nm}(E) = \left(\frac{\mu}{\hbar^2}\right)^2 \frac{1}{(1 + \sigma_e(E))^2 + \sigma_p^2(E)} \frac{\chi_n(x_i)\chi_m(x_i)}{\kappa_n(E)\kappa_m(E)},$$ $$\tilde{\rho}^x_{nm}(x) = \chi_n(x)\chi_m(x)$$ and $$\tilde{\rho}^y_{nm}(y;E) = \begin{cases} -f_{sc}(\kappa_n(E),\kappa_m(E)) & \text{if $n,m \in p$}, \\ f_{cs}(\kappa_n(E),0)e^{-\kappa_m(E)|y-y_i|} & \text{if $n\in p, m\in e$}, \\ f_{cs}(0,\kappa_m(E))e^{-\kappa_n(E)|y-y_i|} & \text{if $n\in e, m\in p$}, \\ -\sigma_p(E) e^{-(\kappa_n(E) + \kappa_m(E))|y-y_i|} & \text{if $n,m \in e$}, \end{cases}$$ where $$\begin{split} f_{cs}(\kappa_1,\kappa_2) &= (1 + \sigma_e)\cos\left[(\kappa_1+\kappa_2)|y-y_i|\right] \\ &+ \sigma_p\sin\left[(\kappa_1+\kappa_2)|y-y_i|\right], \end{split}$$ and $$\begin{split} f_{sc}(\kappa_1,\kappa_2) &= (1 + \sigma_e)\sin\left[(\kappa_1+\kappa_2)|y-y_i|\right] \\ &- \sigma_p\cos\left[(\kappa_1+\kappa_2)|y-y_i|\right]. \end{split}$$ When taking the Fourier transform of the scattering LDOS, we want to be able to resolve differences in $x$-momenta equal to or greater than $\pi/W$ (since this is the separation in $k_x$, or $k_n$, between two adjacent subbands). This requires us to integrate over the interval $[-W,W]$ and we extend the function $\tilde{\rho}^x_{nm}(x)$ such that it is even with respect to the origin. The Fourier transform is then defined as $$\begin{split} {\tilde{\mathcal{N}}}_{nm}(\vec{q};E) &= \mathcal{K}_{nm}(E) \times \\ & \times \sum_{n^\prime = -\infty}^\infty \delta\left(\frac{q_x}{\pi} - \frac{n^\prime}{W}\right) \int_{-W}^W \frac{dx}{2W} e^{-iq_x x}\tilde{\rho}^x_{nm}(x)\times \\ & \times \int_{-\infty}^\infty \frac{dy}{2\pi}e^{-iq_y y}\tilde{\rho}^y_{nm}(y;E), \end{split} \label{eq:fourier}$$ where the comb function fixes $q_x$ to multiples of $\pi/W$. The x-part of the Fourier integral is $$\begin{split} \int_{-W}^W \frac{dx}{2W} e^{-iq_x x} \tilde{\rho}^x_{nm}(x) &= \frac{1}{2W}\left(\delta_{l,-n-m} + \delta_{l,n+m} \right. \\ &\quad \left. - \delta_{l,-n+m} - \delta_{l,n-m}\right), \end{split}$$ independent of if $n$ and $m$ are evanescent or propagating modes. The y-part will depend on mode types. We have already shown what happens when $n,m \in p$. In addition, if $n,m \in e$ we get that $$\int_{-\infty}^\infty \frac{dy}{2\pi} e^{-i q_y y} \tilde{\rho}^y_{nm}(y;E) = -e^{-iq_yy_i}\frac{\sigma_p}{\pi}\frac{\kappa_n(E) + \kappa_m(E)}{q^2_y + (\kappa_n(E)+\kappa_m(E))^2}.$$ If $n\in p, m \in e$ then $$\int_{-\infty}^\infty \frac{dy}{2\pi} e^{-i q_y y} \tilde{\rho}^y_{nm}(y;E) = \frac{e^{-iq_yy_i}}{2\pi}\left[S^y_{pe}(\kappa_n(E)-q_y,\kappa_m(E)) + S^y_{pe}(\kappa_n(E)+q_y,\kappa_m(E))\right],$$ where $$S^y_{pe}(a,b) = \frac{(1/\gamma + \sigma_e(E))b + \sigma_p(E)a}{b^2 + a^2}.$$ If $n\in e, m\in p$ we just need to interchange the $n$ and $m$ in the expression above. $$$$ Armchair graphene nanoribbon {#Appendix_armchair} ============================ In this appendix we first derive an analytic expression for the Green’s function of an armchair nanoribbon with a single impurity. For the geometry, see Fig. \[AC\_fig\](a). We then derive the Fourier transformed density of states. ![(a) The geometry of the armchair nanoribbon. (b) The first Brillouin zone with three sets of Dirac cones.[]{data-label="AC_fig"}](agnr_lattice.eps "fig:"){width="0.49\columnwidth"} ![(a) The geometry of the armchair nanoribbon. (b) The first Brillouin zone with three sets of Dirac cones.[]{data-label="AC_fig"}](agnr_1bz.eps "fig:"){width="0.49\columnwidth"} Unperturbed Green’s function ---------------------------- The first Brillouin zone (1BZ) of graphene contains one pair of inequivalent Dirac cones. It is necessary, however, to include three pairs of cones \[see Fig. \[AC\_fig\](b)\] in order to incorporate all scattering events. The cones are located at $\vec{K}_1^{\pm} = (\pm K_x,0)$, $\vec{K}_2^{\pm} = (\pm K_x/2,K_y)$ and $\vec{K}_3^{\pm} = (\pm K_x/2,-K_y)$, where $K_x = 4\pi/3a$ and $K_y = 2\pi/3a_0$. The distance between two neighbouring atoms is denoted $a_0$, while the lattice constant is denoted $a = \sqrt{3}a_0$. This gives us three sets ($d = 1,2,3$) of wave function spinors,[@Brey:2006cb] $$\vec{\Psi}_{dn}({\vec{r}}) = \begin{pmatrix} \Psi^A_{dn}(y) \\ \Psi^B_{dn}(y) \end{pmatrix} \chi_{dn}(x),$$ where the longitudinal wave function components are $$\Psi^A_{dn}(y) = \lambda \frac{(k_y + ik_{dn})}{\sqrt{k_y^2 + k_{dn}^2}}e^{i(K_{dy} + k_y)y},$$ $$\Psi^B_{dn}(y) = i e^{i(K_{dy} + k_y)y},$$ and the transverse wave function is $$\chi_{dn}(x) = 2C\sin\left[(K_{dx} + k_{dn})x\right].$$ The integer number $n$ labels the quantized transverse momentum $k_{dn} = n\pi/W - K_{dx}$ in cone pair $d$. For each mode $n$, we have positive and negative energy subbands $\epsilon_{dn\lambda}(k_y) = \lambda v_f\sqrt{k_y^2 + (k_{dn})^2}$ labeled by $\lambda=\pm 1$. The Fermi velocity is $v_f = 3a_0|t|/2$, where $t$ is the nearest neighbor tight-binding hopping energy. The wave functions have been normalized through a normalization constant $C = \sqrt{1/4W}$ found from the condition $\int_0^W dx |\chi_{dn}(x)|^2 = 1/2$. Thus, $\chi_{dn}(x) = \chi_n(x) = \sqrt{1/W}\sin(n\pi/Wx)$. The free propagator for band $n$ (in cone pair $d$) is computed as $$\begin{split} \mathbf{g}_{dn}({\vec{r}},{\vec{r}^{\;\prime}}) &= \sum_{\lambda = \pm 1} \int_{-\infty}^\infty \frac{dk_y}{2\pi} \frac{\vec{\Psi}_{dn}({\vec{r}})\vec{\Psi}_{dn}^\dagger({\vec{r}^{\;\prime}})}{E^+ - \epsilon_{dn\lambda}(k_y)} \\ &= \chi_{n}(x)\chi_{n}(x^\prime) \begin{pmatrix} \Gamma_{dn}^{AA}(y,y^\prime;E) & \Gamma_{dn}^{AB}(y,y^\prime;E) \\ \Gamma_{dn}^{BA}(y,y^\prime;E) & \Gamma_{dn}^{BB}(y,y^\prime;E) \end{pmatrix}, \end{split}$$ where $$\Gamma_{dn}^{AA/BB}(y,y^\prime;E) = 2E e^{iK_{dy}(y-y^\prime)}\int_{-\infty}^\infty \frac{dk_y}{2\pi} \frac{e^{ik_y(y-y^\prime)}}{(E^+)^2 - v_f^2(k_y^2 + k_{dn}^2)},$$ and $$\Gamma_{dn}^{AB/BA}(y,y^\prime;E) = \mp 2iv_f e^{iK_{dy}(y-y^\prime)} \int_{-\infty}^\infty \frac{dk_y}{2\pi} \frac{(k_y \pm ik_{dn})e^{ik_y(y-y^\prime)}}{(E^+)^2 - v_f^2(k_y^2 + k_{dn}^2)}.$$ After contour integration, we find the final form to be $$\Gamma_{dn}^{AA/BB}(y,y^\prime;E) = -i\frac{|E|}{v_f^2} e^{iK_{dy}(y-y^\prime)}\frac{e^{i{\text{sgn}\left(E\right)}\kappa_{dn}(E)|y-y^\prime|}}{\kappa_{dn}(E)},$$ and $$\Gamma_{dn}^{AB/BA}(y,y^\prime;E) = -\frac{1}{v_f}e^{iK_{dy}(y-y^\prime)}\left[\frac{i{\text{sgn}\left(E\right)}k_{dn}}{\kappa_{dn}(E)} \pm {\text{sgn}\left(y-y^\prime\right)}\right]e^{i{\text{sgn}\left(E\right)}\kappa_{dn}(E)|y-y^\prime|},$$ where $$\kappa_{dn} = \sqrt{|(E/v_f)^2 - k_{dn}^2|}.$$ In the above formulas, we have assumed that $n$ is a propagating mode (e.g. $|E/v_f| > |k_{dn}$). If mode $n$ is evanescent ($|E/v_F| < |k_{dn}|$), we have to modify the longitudinal momentum so that $\kappa_{dn} \rightarrow i\text{sgn}(E)\kappa_{dn}$. Green’s function, one impurity ------------------------------ For the graphene armchair ribbon, we select an impurity fully localized on the A-sublattice, scattering equally between all modes $n$ and $m$. The matrix elements of the impurity potential then is $$\label{eq:Vgr} \mathbf{V}_{nm}({\vec{r}}_i,{\vec{r}}_i) = \mathbf{V}({\vec{r}}_i,{\vec{r}}_i) = \gamma \begin{pmatrix} 1 & 0 \\ 0 & 0 \end{pmatrix},$$ where $\gamma$ is the impurity strength. The T-matrix equation is written down in analogy to the 2DEG case, but for graphene it acquires a 2x2 matrix structure. For the potential in Eq. (\[eq:Vgr\]), we get $$\begin{split} {\mathbf{T}}({\vec{r}}_i,{\vec{r}}_i;E) &= {\mathbf{V}}({\vec{r}}_i,{\vec{r}}_i) + {\mathbf{V}}({\vec{r}}_i,{\vec{r}}_i)\left[\sum_d\sum_l {\mathbf{g}}_{dl}({\vec{r}}_i,{\vec{r}}_i;E) \right]{\mathbf{T}}({\vec{r}}_i,{\vec{r}}_i) \\ &= \frac{1}{1/\gamma + \sigma_e(E) + i\sigma_p(E)} \begin{pmatrix} 1 & 0 \\ 0 & 0 \end{pmatrix}, \end{split}$$ where $$\sigma_{p}(E) = \frac{|E|}{v_f^2}\sum_d\sum_{l\in p} \frac{\chi_{l}^2(x_i)}{\kappa_{dl}(E)}$$ and $$\sigma_{e}(E) = \frac{E}{v_f^2}\sum_d\sum_{l\in e} \frac{\chi_{l}^2(x_i)}{\kappa_{dl}(E)}.$$ The letters $e$ and $p$ denotes sets of evanescent and propagating modes, respectively. The Dyson equation for the Green’s function can now be written as $${\mathbf{G}}_{dncm}({\vec{r}},{\vec{r}^{\;\prime}}) = {\mathbf{g}}_{dn}({\vec{r}},{\vec{r}^{\;\prime}})\delta_{nm} + {\mathbf{\tilde{G}}}_{dncm}({\vec{r}},{\vec{r}^{\;\prime}})$$ where $${\mathbf{\tilde{G}}}_{dncm}({\vec{r}},{\vec{r}^{\;\prime}};E) = {\mathbf{g}}_{dn}({\vec{r}},{\vec{r}}_i;E){\mathbf{T}}({\vec{r}}_i,{\vec{r}}_i;E){\mathbf{g}}_{cm}({\vec{r}}_i,{\vec{r}^{\;\prime}};E).$$ The scattering part ${\mathbf{\tilde{G}}}_{dncm}({\vec{r}},{\vec{r}^{\;\prime}};E)$ takes the form $$\begin{split} \tilde{{\mathbf{G}}}_{dncm}({\vec{r}},{\vec{r}^{\;\prime}};E) &= \frac{1}{1/\gamma + \sigma_e(E) + i\sigma_p(E)} \begin{pmatrix} g^{AA}_{dn}({\vec{r}},{\vec{r}}_i;E)g^{AA}_{cm}({\vec{r}}_i,{\vec{r}^{\;\prime}};E) & g^{AA}_{dn}({\vec{r}},{\vec{r}}_i;E)g^{AB}_{cm}({\vec{r}}_i,{\vec{r}^{\;\prime}};E) \\ g^{BA}_{dn}({\vec{r}},{\vec{r}}_i;E)g^{AA}_{cm}({\vec{r}}_i,{\vec{r}^{\;\prime}};E) & g^{BA}_{dn}({\vec{r}},{\vec{r}}_i;E)g^{AB}_{cm}({\vec{r}}_i,{\vec{r}^{\;\prime}};E) \end{pmatrix} \\ &= \frac{\chi_{n}(x)\chi_{n}(x_i)\chi_{m}(x_i)\chi_{m}(x^\prime)}{1/\gamma+\sigma_e(E) + i\sigma_p(E)} \begin{pmatrix} \Gamma^{AA}_{dn}(y,y_i;E)\Gamma^{AA}_{cm}(y_i,y^\prime;E) & \Gamma^{AA}_{dn}(y,y_i;E)\Gamma^{AB}_{cm}(y_i,y^\prime;E) \\ \Gamma^{BA}_{dn}(y,y_i;E)\Gamma^{AA}_{cm}(y_i,y^\prime;E) & \Gamma^{BA}_{dn}(y,y_i;E)\Gamma^{AB}_{cm}(y_i,y^\prime;E) \end{pmatrix}. \end{split}$$ For the computation of the local density of states, we need the two diagonal components. Their explicit forms ($n,m \in p$) are $$\begin{split} \tilde{G}^{AA}_{dncm}({\vec{r}},{\vec{r}^{\;\prime}};E) &= -\frac{1}{1/\gamma + \sigma_e(E) + i\sigma_p(E)}\left(\frac{|E|}{v_f^2}\right)^2\chi_{n}(x)\chi_{n}(x_i)\chi_{m}(x_i)\chi_{m}(x^\prime)\times \\ & \times e^{iK_{dy}(y-y_i)}e^{iK_{cy}(y_i - y^\prime)}\frac{e^{i{\text{sgn}\left(E\right)}(\kappa_{dn}(E)|y-y_i| +\kappa_{cm}(E)|y_i-y^\prime|)}}{\kappa_{dn}(E)\kappa_{cm}(E)} \label{GAA} \end{split}$$ and $$\begin{split} \tilde{G}^{BB}_{dncm}({\vec{r}},{\vec{r}^{\;\prime}};E) &= \frac{1}{1/\gamma + \sigma_e(E) + i\sigma_p(E)}\left(\frac{1}{v_f}\right)^2 \chi_{n}(x)\chi_{n}(x_i)\chi_{m}(x_i)\chi_{m}(x^\prime)e^{iK_{dy}(y-y_i)}e^{iK_{cy}(y_i - y^\prime)} \times \\ & \times e^{i{\text{sgn}\left(E\right)}(\kappa_{dn}(E)|y-y_i| +\kappa_{cm}(E)|y_i-y^\prime|)} \left[\frac{i{\text{sgn}\left(E\right)}k_{dn}}{\kappa_{dn}(E)} - {\text{sgn}\left(y-y_i\right)}\right] \left[\frac{i{\text{sgn}\left(E\right)}k_{cm}}{\kappa_{cm}(E)} + {\text{sgn}\left(y_i-y^\prime\right)}\right]. \label{GBB} \end{split}$$ Density of states ----------------- The scattering correction to the local density of states can be computed separately for the two sublattices, and is given by $$\begin{split} \tilde{\rho}_{A/B}({\vec{r}};E) &= -\frac{1}{\pi}\sum_{dc}\sum_{nm}{\textrm{Im}}\left[\tilde{G}_{dncm}^{AA/BB}({\vec{r}},{\vec{r}};E)\right] \\ &= -\frac{1}{\pi}\sum_{dc}\sum_{nm}\mathcal{K}^{A/B}_{dncm}(E)\tilde{\rho}^x_{nm}(x)\tilde{\rho}^{(A/B)y}_{dncm}(y;E),\end{split} \label{LDOS_graphene}$$ where $\tilde{\rho}^x_{nm}(x) = \chi_{n}(x)\chi_{m}(x)$. The A/B sublattice corrections are found by substituting Eq. (\[GAA\]) and Eq. (\[GBB\]) respectively in Eq. (\[LDOS\_graphene\]). The results are very similar (the A correction being almost identical) to the 2DEG case, and for $n,m \in p$ we find that $$\mathcal{K}^A_{dncm}(E) = \frac{1}{(1/\gamma+\sigma_e(E))^2 + \sigma_p^2(E)}\left(\frac{|E|}{v_f^2}\right)^2\frac{\chi_{n}(x_i)\chi_{m}(x_i)}{\kappa_{dn}(E)\kappa_{cm}(E)},$$ $$\mathcal{K}^B_{dncm}(E) = -(v_f/|E|)^2\mathcal{K}^A_{dncm}(E),$$ $$\tilde{\rho}^{Ay}_{dncm}(y;E) = \sigma_p(E)F^c_{dncm}(y-y_i;E) - (1/\gamma +\sigma_e(E))F^s_{dncm}(y-y_i;E),$$ and $$\begin{split} \tilde{\rho}^{By}_{dncm}(y;E) &= \left\{ \sigma_p(E)(-k_nk_m + \kappa_n(E)\kappa_m(E)) \right. \\ & \quad\left. +(1/\gamma+\sigma_e(E)){\text{sgn}\left(E\right)}{\text{sgn}\left(y-y_i\right)}(k_n\kappa_m(E)+k_m\kappa_n(E))\right\}F^c_{dncm}(y-y_i;E) \\ &+ \left\{ \sigma_p(E){\text{sgn}\left(E\right)}{\text{sgn}\left(y-y_i\right)}(k_n \kappa_m(E)+k_m\kappa_n(E))\right. \\ & \quad\left. - (1/\gamma+\sigma_e(E)){\text{sgn}\left(E\right)}{\text{sgn}\left(y-y_i\right)}(k_n\kappa_m(E)+k_m\kappa_n(E))\right\}F^s_{dncm}(y-y_i;E) \end{split}$$ where $$F^c_{dncm}(y;E) = \cos\left[\text{sgn}(E)(\kappa_{dn}(E)+\kappa_{cm}(E))|y| + (K_{dy}-K_{cy})y\right]$$ and $$F^s_{dncm}(y;E) = \sin\left[\text{sgn}(E)(\kappa_{dn}(E)+\kappa_{cm}(E))|y| + (K_{dy}-K_{cy})y\right].$$ The Fourier transform of each component is carried out exactly as for the 2DEG, using Eq. (\[eq:fourier\]), and the results for the AGNR are shown in Section \[Ch\_armchair\]. [32]{}ifxundefined \[1\][ ifx[\#1]{} ]{}ifnum \[1\][ \#1firstoftwo secondoftwo ]{}ifx \[1\][ \#1firstoftwo secondoftwo ]{}““\#1””@noop \[0\][secondoftwo]{}sanitize@url \[0\][‘\ 12‘\$12 ‘&12‘\#12‘12‘\_12‘%12]{}@startlink\[1\]@endlink\[0\]@bib@innerbibempty @noop [****,  ()]{} @noop [****, ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [**** ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [ ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****, ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****, ()]{}, H. Raza, Editor, Springer-Verlag, Berlin Heidelberg 2012. @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} @noop [****,  ()]{} E. N. Economou, [*Green’s Functions in Quantum Physics, 3rd Ed.*]{}, Springer Verlag, Berlin 2006 @noop [****, ()]{} @noop [****,  ()]{}
{ "pile_set_name": "ArXiv" }
--- abstract: 'We show that the $h$-vector of a ladder determinantal ring cogenerated by $M=[u_1\mid v_1]$ is log-concave. Thus we prove an instance of a conjecture of Stanley, resp. Conca and Herzog.' author: - Martin Rubey title: | The $h$-vector of a ladder determinantal ring cogenerated by $2\times 2$ minors is log-concave --- [^1] Introduction ============ A sequence of real numbers ${\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodea_{\arabic{seq}},\stepcounter{seq}a_{\arabic{seq}},\dots ,a_{n}\else$a_{\arabic{seq}}$,~\stepcounter{seq} $a_{\arabic{seq}}$, \dots ,~$a_{n}$\fi} \else{\ifmmodea_{1},a_{1+1},\dots ,a_{n}\else$a_{1}$,~$a_{1+1}$, \dots ,~$a_{n}$\fi} \fi}$ is [*logarithmically concave*]{}, for short [*log-concave*]{}, if $a_{i-1}a_{i+1}\le a_i^2$ for $i\in\{{\if\IsInteger{2}{\setcounter{seq}{2}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n-1}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n-1}$\fi} \else{\ifmmode{2},{2+1},\dots ,{n-1}\else${2}$,~${2+1}$, \dots ,~${n-1}$\fi} \fi}\}$. Numerous sequences arising in combinatorics and algebra have, or seem to have this property. In the paper [@Stanley1989] written in 1989, Richard Stanley collected various results on this topic. (For an update see [@Brenti1994].) There he also stated the following conjecture: \[cnj:Stanley\] Let $R=R_0\oplus R_1\oplus\dots$ be a graded (Noetherian) Cohen-Macaulay (or perhaps Gorenstein) *domain* over a field $K=R_0$, which is generated by $R_1$ and has Krull dimension $d$. Let $H(R,m)=\dim_K R_m$ be the Hilbert function of $R$ and write $$\sum_{m\ge0}H(R,m)x^m=(1-x)^{-d}\sum_{i=0}^s h_i x^i.$$ Then the sequence ${\if\IsInteger{0}{\setcounter{seq}{0}\ifmmodeh_{\arabic{seq}},\stepcounter{seq}h_{\arabic{seq}},\dots ,h_{s}\else$h_{\arabic{seq}}$,~\stepcounter{seq} $h_{\arabic{seq}}$, \dots ,~$h_{s}$\fi} \else{\ifmmodeh_{0},h_{0+1},\dots ,h_{s}\else$h_{0}$,~$h_{0+1}$, \dots ,~$h_{s}$\fi} \fi}$ is log-concave. The sequence ${\if\IsInteger{0}{\setcounter{seq}{0}\ifmmodeh_{\arabic{seq}},\stepcounter{seq}h_{\arabic{seq}},\dots ,h_{s}\else$h_{\arabic{seq}}$,~\stepcounter{seq} $h_{\arabic{seq}}$, \dots ,~$h_{s}$\fi} \else{\ifmmodeh_{0},h_{0+1},\dots ,h_{s}\else$h_{0}$,~$h_{0+1}$, \dots ,~$h_{s}$\fi} \fi}$ is called the [*$h$-vector*]{} of the ring. Orginally the question was to decide whether a given sequence can arise as the $h$-vector of some ring. In this sense the validity of the conjecture would imply that log-concavity was a necessary condition on the $h$-vector. It is now known however [@NiesiRobbiano1992; @Brenti1994] that Stanley’s conjecture is not true in general. Several natural weakenings have been considered, but are still open. For example, Aldo Conca and Jürgen Herzog conjectured that the $h$-vector would be log-concave for the special case where $R$ is a ladder determinantal ring. (Note that ladder determinantal rings are Cohen-Macaulay, as was shown in [@HerzogTrung1992 Corollary 4.10], but not necessarily Gorenstein.) We will prove the conjecture of Conca and Herzog in the simplest case, i.e., where $R$ is a ladder determinantal ring cogenerated by $2\times 2$ minors, see Corollary \[cor:main\]. In the case of ladder determinantal rings the $h$-vector has a nice combinatorial interpretation. This follows from work of Abhyankar and Kulkarni [@Abhyankar1988; @AbhyankarKulkarni1989; @Kulkarni1993; @Kulkarni1996], Bruns, Conca, Herzog, and Trung [@BrunsHerzog1992; @Conca1995; @ConcaHerzog1994; @HerzogTrung1992]. In the following paragraphs, which are taken almost verbatim from [@KrattenthalerRubey2001], we will explain these matters. Ladders, ladder determinantal rings and non-intersecting lattice paths {#s:Defs} ====================================================================== First we have to introduce the notion of a ladder: \[dfn:ladder\] Let ${\ensuremath{\mathbf{X}}}=(x_{i,j})_{0\le i\le b, 0\le j\le a}$ be a $(b+1)\times (a+1)$ matrix of indeterminates. Let ${\ensuremath{\mathbf{Y}}}=(y_{i,j})_{0\le i\le b, 0\le j\le a}$ be another matrix of the same dimensions, with the property that $y_{i,j}\in\{0,x_{i,j}\}$, and if $y_{i,j}=x_{i,j}$ and $y_{i^\prime,j^\prime}=x_{i^\prime,j^\prime}$, where $i\le i^\prime$ and $j\le j^\prime$ then $y_{r,s}=x_{r,s}$ for all $r$ and $s$ with $i\le r\le i^\prime$ and $j\le s\le j^\prime$. Such a matrix ${\ensuremath{\mathbf{Y}}}$ is called a [*ladder*]{}. A [*ladder region*]{} $L$ is a subset of $\mathbb Z^2$ with the property that if $(i,j)$ and $(i^\prime,j^\prime)\in L$, $i\le i^\prime$ and $j\ge j^\prime$ then $(r,s)\in L$ for all $r\in\{{\if\IsInteger{i}{\setcounter{seq}{i}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{i^\prime}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${i^\prime}$\fi} \else{\ifmmode{i},{i+1},\dots ,{i^\prime}\else${i}$,~${i+1}$, \dots ,~${i^\prime}$\fi} \fi}\}$ and $s\in\{{\if\IsInteger{j^\prime}{\setcounter{seq}{j^\prime}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{j}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${j}$\fi} \else{\ifmmode{j^\prime},{j^\prime+1},\dots ,{j}\else${j^\prime}$,~${j^\prime+1}$, \dots ,~${j}$\fi} \fi}\}$. Clearly, a ladder region can be described by two weakly increasing functions ${\underline{L}}$ and ${\overline{L}}$, such that $L$ is exactly the set of points $\{(i,j):{\underline{L}}(i)\le j\le{\overline{L}}(i)\}$. We associate with ${\ensuremath{\mathbf{Y}}}$ a ladder region $L\subset\mathbb Z^2$ via $(j,b-i)\in L$ if and only if $y_{i,j}=x_{i,j}$. In Figure \[fig:latticepaths\].a an example of a ladder with $a=8$ and $b=9$ is shown, the corresponding ladder region is shown in Figure \[fig:latticepaths\].b. Now we can define the ring we are dealing with: \[dfn:ring\] Given a $(b+1)\times (a+1)$ matrix ${\ensuremath{\mathbf{Y}}}$ which is a ladder, fix a “bivector” $M=[{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodeu_{\arabic{seq}},\stepcounter{seq}u_{\arabic{seq}},\dots ,u_{n}\else$u_{\arabic{seq}}$,~\stepcounter{seq} $u_{\arabic{seq}}$, \dots ,~$u_{n}$\fi} \else{\ifmmodeu_{1},u_{1+1},\dots ,u_{n}\else$u_{1}$,~$u_{1+1}$, \dots ,~$u_{n}$\fi} \fi}\mid{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodev_{\arabic{seq}},\stepcounter{seq}v_{\arabic{seq}},\dots ,v_{n}\else$v_{\arabic{seq}}$,~\stepcounter{seq} $v_{\arabic{seq}}$, \dots ,~$v_{n}$\fi} \else{\ifmmodev_{1},v_{1+1},\dots ,v_{n}\else$v_{1}$,~$v_{1+1}$, \dots ,~$v_{n}$\fi} \fi}]$ of integers with $1\le{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodeu_{\arabic{seq}}<\stepcounter{seq}u_{\arabic{seq}}<\dots <u_{n}\else$u_{\arabic{seq}}$<~\stepcounter{seq} $u_{\arabic{seq}}$< \dots <~$u_{n}$\fi} \else{\ifmmodeu_{1}<u_{1+1}<\dots <u_{n}\else$u_{1}$<~$u_{1+1}$< \dots <~$u_{n}$\fi} \fi}\le b+1$ and $1\le{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodev_{\arabic{seq}}<\stepcounter{seq}v_{\arabic{seq}}<\dots <v_{n}\else$v_{\arabic{seq}}$<~\stepcounter{seq} $v_{\arabic{seq}}$< \dots <~$v_{n}$\fi} \else{\ifmmodev_{1}<v_{1+1}<\dots <v_{n}\else$v_{1}$<~$v_{1+1}$< \dots <~$v_{n}$\fi} \fi}\le a+1$. By convention we set $u_{n+1}=b+2$ and $v_{n+1}=a+2$. Let $K[{\ensuremath{\mathbf{Y}}}]$ denote the ring of all polynomials over some field $K$ in the $y_{i,j}$’s, where $0\le i\le b$ and $0\le j\le a$. Furthermore, let $I_M({\ensuremath{\mathbf{Y}}})$ be the ideal in $K[{\ensuremath{\mathbf{Y}}}]$ that is generated by those $t\times t$ minors of ${\ensuremath{\mathbf{Y}}}$ that contain only nonzero entries, whose rows form a subset of the last $u_t-1$ rows *or* whose columns form a subset of the last $v_t-1$ columns, $t\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n+1}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n+1}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n+1}\else${1}$,~${1+1}$, \dots ,~${n+1}$\fi} \fi}\}$. Thus, for $t=n+1$ the rows and columns of minors are unrestricted. The ideal $I_M({\ensuremath{\mathbf{Y}}})$ is called a [*ladder determinantal ideal generated by the minors defined by $M$*]{}. We call $R_M({\ensuremath{\mathbf{Y}}})=K[{\ensuremath{\mathbf{Y}}}]/I_M({\ensuremath{\mathbf{Y}}})$ the [*ladder determinantal ring cogenerated by the minors defined by $M$*]{}, or, in abuse of language, the [*ladder determinantal ring cogenerated by $M$*]{}. Note that we could restrict ourselves to the case $u_1=v_1=1$, because all the elements of ${\ensuremath{\mathbf{Y}}}$ that are in one of the last $u_1-1$ rows or in one of the last $v_1-1$ columns are in the ideal. Next, we introduce the combinatorial objects that will accompany us throughout the rest of this paper: A [*two-rowed array of length $k$*]{} is a pair of strictly increasing sequences of integers, both of length $k$. A two-rowed array $T=\left( \begin{smallmatrix} a_1&a_2&\dots&a_k\\ b_1&b_2&\dots&b_k \end{smallmatrix}\right)$ is [*bounded*]{} by $A=(A_1,A_2)$ and $E=(E_1,E_2)$, if $$\begin{array}{lr!{\le}c!{\le}l} & A_1 &{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodea_{\arabic{seq}}<\stepcounter{seq}a_{\arabic{seq}}<\dots <a_{k}\else$a_{\arabic{seq}}$<~\stepcounter{seq} $a_{\arabic{seq}}$< \dots <~$a_{k}$\fi} \else{\ifmmodea_{1}<a_{1+1}<\dots <a_{k}\else$a_{1}$<~$a_{1+1}$< \dots <~$a_{k}$\fi} \fi}& E_1-1\\ \text{and}& A_2+1&{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodeb_{\arabic{seq}}<\stepcounter{seq}b_{\arabic{seq}}<\dots <b_{k}\else$b_{\arabic{seq}}$<~\stepcounter{seq} $b_{\arabic{seq}}$< \dots <~$b_{k}$\fi} \else{\ifmmodeb_{1}<b_{1+1}<\dots <b_{k}\else$b_{1}$<~$b_{1+1}$< \dots <~$b_{k}$\fi} \fi}& E_2. \end{array}$$ Given *any* subset $L$ of $\mathbb Z^2$, we say that the two-rowed array $T$ is [*in $L$*]{}, if $(a_i,b_i)\in L$ for $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{k}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${k}$\fi} \else{\ifmmode{1},{1+1},\dots ,{k}\else${1}$,~${1+1}$, \dots ,~${k}$\fi} \fi}\}$. By ${{\ensuremath{\mathcal{T}}}}^L_k(A\mapsto E)$ we will denote the set of two-rowed arrays of length $k$, bounded by $A$ and $E$ which are in $L$. The [*total length*]{} of a family of two-rowed arrays is just the sum of the lengths of its members. Let $T_1=\left( \begin{smallmatrix} a_1&a_2&\dots&a_k\\ b_1&b_2&\dots&b_k \end{smallmatrix}\right)$ and $T_2=\left( \begin{smallmatrix} x_1&x_2&\dots&x_l\\ y_1&y_2&\dots&y_l \end{smallmatrix}\right)$ be two-rowed arrays bounded by $A^{(1)}=(A^{(1)}_1,A^{(1)}_2)$ and $E^{(1)}=(E^{(1)}_1,E^{(1)}_2)$ and $A^{(2)}=(A^{(2)}_1,A^{(2)}_2)$ and $E^{(2)}=(E^{(2)}_1,E^{(2)}_2)$ respectively. Set $a_{k+1}=E^{(1)}_1$ and $b_0=A^{(1)}_2$. We say that $T_1$ and $T_2$ [*intersect*]{} if there are indices $I$ and $J$ such that $$\begin{aligned} \tag{$\times$}\label{e:intersect} x_J&\le a_I\\ b_{I-1}&\le y_J \end{aligned}$$ where $1\le I\le k+1$ and $1\le J\le l$. A family of two-rowed arrays is [*non-intersecting*]{} if no two arrays in it intersect. Note that a two-rowed array in ${{\ensuremath{\mathcal{T}}}}^L_k(A\mapsto E)$ can be visualized by a lattice path with east and north steps, that starts in $A$ and terminates in $E$ and has exactly $k$ north-east turns which are all in $L$: Each pair $(a_i,b_i)$ of a two-rowed array $\left( \begin{smallmatrix} a_1&a_2&\dots&a_k\\ b_1&b_2&\dots&b_k \end{smallmatrix}\right)$ then corresponds to a north-east turn of the lattice path. It is easy to see that Condition  holds if and only if the lattice paths corresponding to $T_1$ and $T_2$ intersect. For an example see Figure \[fig:latticepaths\].c, where the three two-rowed arrays $$T^{(1)}= \begin{pmatrix} 2 & 3 \\ 6 & 7 \end{pmatrix}\text{, } T^{(2)}= \begin{pmatrix} 3 & 5\\ 4 & 6 \end{pmatrix}\text{ and } T^{(3)}= \begin{pmatrix} 2 & 4 & 6\\ 1 & 3 & 4 \end{pmatrix}$$ bounded by $A^{(1)}=(0,3)$, $A^{(2)}=(0,2)$, $A^{(3)}=(0,0)$ and $E^{(1)}=(5,9)$, $E^{(2)}=(7,9)$, $E^{(3)}=(8,9)$ are shown as lattice paths. The points of the ladder-region $L$ are drawn as small dots, the circles indicate the start- and endpoints and the big dots indicate the north-east turns. Ø [ccc]{} & &lt;10pt,0pt&gt;:\*+\* & &lt;10pt,0pt&gt;:\*+\* \ \ & & A combinatorial interpretation of the $h$-vector of a ladder determinantal ring =============================================================================== We are now ready to state the theorem which reveals the combinatorial nature of the $h$-vector of $R_M({\ensuremath{\mathbf{Y}}})=K[{\ensuremath{\mathbf{Y}}}]/I_M({\ensuremath{\mathbf{Y}}})$, the ladder determinantal ring cogenerated by $M$. \[thm:comb\] Let ${\ensuremath{\mathbf{Y}}}=(y_{i,j})_{0\le i\le b,\ 0\le j\le a}$ be a ladder and let $M=[{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodeu_{\arabic{seq}},\stepcounter{seq}u_{\arabic{seq}},\dots ,u_{n}\else$u_{\arabic{seq}}$,~\stepcounter{seq} $u_{\arabic{seq}}$, \dots ,~$u_{n}$\fi} \else{\ifmmodeu_{1},u_{1+1},\dots ,u_{n}\else$u_{1}$,~$u_{1+1}$, \dots ,~$u_{n}$\fi} \fi}\mid{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodev_{\arabic{seq}},\stepcounter{seq}v_{\arabic{seq}},\dots ,v_{n}\else$v_{\arabic{seq}}$,~\stepcounter{seq} $v_{\arabic{seq}}$, \dots ,~$v_{n}$\fi} \else{\ifmmodev_{1},v_{1+1},\dots ,v_{n}\else$v_{1}$,~$v_{1+1}$, \dots ,~$v_{n}$\fi} \fi}]$ be a bivector of integers with $1\le{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodeu_{\arabic{seq}}<\stepcounter{seq}u_{\arabic{seq}}<\dots <u_{n}\else$u_{\arabic{seq}}$<~\stepcounter{seq} $u_{\arabic{seq}}$< \dots <~$u_{n}$\fi} \else{\ifmmodeu_{1}<u_{1+1}<\dots <u_{n}\else$u_{1}$<~$u_{1+1}$< \dots <~$u_{n}$\fi} \fi}\le a+1$ and $1\le{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodev_{\arabic{seq}}<\stepcounter{seq}v_{\arabic{seq}}<\dots <v_{n}\else$v_{\arabic{seq}}$<~\stepcounter{seq} $v_{\arabic{seq}}$< \dots <~$v_{n}$\fi} \else{\ifmmodev_{1}<v_{1+1}<\dots <v_{n}\else$v_{1}$<~$v_{1+1}$< \dots <~$v_{n}$\fi} \fi}\le b+1$. For $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n}\else${1}$,~${1+1}$, \dots ,~${n}$\fi} \fi}\}$ let $$\begin{aligned} A^{(i)}&=(0,u_{n+1-i}-1)\\ E^{(i)}&=(a-v_{n+1-i}+1,b). \intertext{Let $L^{(n)}=L$ be the ladder region associated with ${\ensuremath{\mathbf{Y}}}$ and for $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n-1}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n-1}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n-1}\else${1}$,~${1+1}$, \dots ,~${n-1}$\fi} \fi}\}$ let} L^{(i)}&=\{(x,y)\in L^{(i+1)}:x\le E^{(i)}_1, y\ge A^{(i)}_2 \text{ and } (x+1,y-1)\in L^{(i+1)}\}. \intertext{Finally, for $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n}\else${1}$,~${1+1}$, \dots ,~${n}$\fi} \fi}\}$ let} B^{(i)}&=\{(x,y)\in L^{(i)}:(x+1,y-1)\notin L^{(i)}\}. \end{aligned}$$ and let $d$ be the cardinality of $\bigcup_{i=1}^n B^{(i)}$. Then, under the assumption that all of the points $A^{(i)}$ and $E^{(i)}$, $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n}\else${1}$,~${1+1}$, \dots ,~${n}$\fi} \fi}\}$, lie inside the ladder region $L$, the Hilbert series of the ladder determinantal ring $R_M({\ensuremath{\mathbf{Y}}})=K[{\ensuremath{\mathbf{Y}}}]/I_M({\ensuremath{\mathbf{Y}}})$ equals $$\sum _{\ell\ge0}\dim_K R_M({\ensuremath{\mathbf{Y}}})_\ell\,z^\ell =\frac{\sum_{\ell\ge0}\size{{{\ensuremath{\mathcal{T}}}}^L_\ell({\ensuremath{\mathbf{A}}}\mapsto{\ensuremath{\mathbf{E}}})}z^\ell} {(1-z)^d}.$$ Here, $R_M({\ensuremath{\mathbf{Y}}})_\ell$ denotes the homogeneous component of degree $\ell$ in $R_M({\ensuremath{\mathbf{Y}}})$ and $\size{{{\ensuremath{\mathcal{T}}}}^L_\ell({\ensuremath{\mathbf{A}}}\mapsto{\ensuremath{\mathbf{E}}})}$ is the number of non-intersecting families of two-rowed arrays with total length $\ell$, such that the $i$^th^ two-rowed array is bounded by $A^{(i)}$ and $E^{(i)}$ and is in $L^{(i)}\setminus B^{(i)}$ for $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n}\else${1}$,~${1+1}$, \dots ,~${n}$\fi} \fi}\}$. The sets $B^{(i)}$, $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n}\else${1}$,~${1+1}$, \dots ,~${n}$\fi} \fi}\}$ can be visualized as being the lower-right boundary of $L^{(i)}$. Viewed as a path, there are exactly $E^{(i)}_1-A^{(i)}_1+E^{(i)}_2-A^{(i)}_2+1$ lattice points on $B^{(i)}$, but not all of them are necessarily in $L$. However, if $L$ is an upper ladder, that is, $(a,0)\in L$, then this must be the case and we have $$\begin{aligned} d&=\sum_{i=1}^n\left(E^{(i)}_1-A^{(i)}_1+E^{(i)}_2-A^{(i)}_2+1\right)\\ &=\sum_{i=1}^n\left(a-v_{n+1-i}+1 + b -u_{n+1-i}+1+1\right)\\ &=n(a+b+3)-\sum_{i=1}^n\left(u_i+v_i\right),\end{aligned}$$ as in [@KrattenthalerRubey2001]. In Figure \[fig:LadderFace\].a, an example for a ladder region $L$ with $a=8$ and $b=9$ is given. The small dots represent elements of $L$, the circles on the left and on the top of $L$ represent the points $A^{(i)}$ and $E^{(i)}$, $i\in\{1,2,3\}$ that are specified by the minor $M=[1,3,4\mid 1,2,4]$. The dotted lines indicate the lower boundary of $L^{(i)}$. Note that the point $(4,9)$ is not an element of $L$. Therefore, in this example we have $$d=n(a+b+3)-\sum_{i=1}^n\left(u_i+v_i\right)-1=44.$$ [cc]{} &lt;11pt,0pt&gt;:\*+\* & &lt;11pt,0pt&gt;:\*+\* \ \ & $$\begin{xy}<11pt,0pt>:*+\xybox{\Ladder\StartEnd\LowestPaths\Face \PathA}*\frm{-}\end{xy} \quad \begin{xy}<11pt,0pt>:*+\xybox{\Ladder\StartEnd\LowestPaths\Face \PathA\PathB}*\frm{-}\end{xy} \quad \begin{xy}<11pt,0pt>:*+\xybox{\Ladder\StartEnd\LowestPaths\Face \PathA\PathB\PathC}*\frm{-}\end{xy}$$ $$\begin{xy}<11pt,0pt>:*+\xybox{\Ladder\StartEnd\LowestPaths\NETurns \PathNEA\PathNEB\PathNEC}*\frm{-}\end{xy}$$ We will use results of Jürgen Herzog and Ngô Viêt Trung. In Section 4 of [@HerzogTrung1992], ladder determinantal rings are introduced and investigated. We equip the indeterminates $x_{i,j}$, $i\in\{{\if\IsInteger{0}{\setcounter{seq}{0}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{b}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${b}$\fi} \else{\ifmmode{0},{0+1},\dots ,{b}\else${0}$,~${0+1}$, \dots ,~${b}$\fi} \fi}\}$ and $j\in\{{\if\IsInteger{0}{\setcounter{seq}{0}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{a}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${a}$\fi} \else{\ifmmode{0},{0+1},\dots ,{a}\else${0}$,~${0+1}$, \dots ,~${a}$\fi} \fi}\}$ with the following partial order: $$x_{i,j}\le x_{i^\prime,j^\prime}\text{ if }i\ge i^\prime \text{ and }j\le j^\prime.$$ A $t$-antichain in this partial order is a family of elements $x_{r_1,s_1},x_{r_2,s_2},\dots,x_{r_t,s_t}$ such that ${\if\IsInteger{1}{\setcounter{seq}{1}\ifmmoder_{\arabic{seq}}<\stepcounter{seq}r_{\arabic{seq}}<\dots <r_{t}\else$r_{\arabic{seq}}$<~\stepcounter{seq} $r_{\arabic{seq}}$< \dots <~$r_{t}$\fi} \else{\ifmmoder_{1}<r_{1+1}<\dots <r_{t}\else$r_{1}$<~$r_{1+1}$< \dots <~$r_{t}$\fi} \fi}$ and ${\if\IsInteger{1}{\setcounter{seq}{1}\ifmmodes_{\arabic{seq}}<\stepcounter{seq}s_{\arabic{seq}}<\dots <s_{t}\else$s_{\arabic{seq}}$<~\stepcounter{seq} $s_{\arabic{seq}}$< \dots <~$s_{t}$\fi} \else{\ifmmodes_{1}<s_{1+1}<\dots <s_{t}\else$s_{1}$<~$s_{1+1}$< \dots <~$s_{t}$\fi} \fi}$. Thus, a $t$-antichain corresponds to a sequence $(s_1,b-r_1),(s_2,b-r_2),\dots,(s_t,b-r_t)$ of $t$ points in the ladder region associated with ${\ensuremath{\mathbf{Y}}}$, where each point lies strictly south-east of the previous ones. Let $D_t$ be the union of the last $u_t-1$ rows and the last $v_t-1$ columns of ${\ensuremath{\mathbf{Y}}}$. Let $\Delta_M({\ensuremath{\mathbf{Y}}})$ be the simplicial complex whose $k$-dimensional faces are subsets of elements of ${\ensuremath{\mathbf{Y}}}$ of cardinality $k+1$ which do not contain a $t$-antichain in $D_t$ for $t\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n+1}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n+1}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n+1}\else${1}$,~${1+1}$, \dots ,~${n+1}$\fi} \fi}\}$. Let $f_k$ be the number of $k$-dimensional faces of $\Delta_M({\ensuremath{\mathbf{Y}}})$ for $k\ge 0$. Then, Corollary 4.3 of [@HerzogTrung1992] states, that $$\dim_K R_M({\ensuremath{\mathbf{Y}}})_\ell=\sum_{k\ge 0}\binom{\ell-1}{k}f_k.$$ In the following, we will find an expression for the numbers $f_k$ involving certain families of non-intersecting lattice paths. In Figure \[fig:LadderFace\].b, a $10$-dimensional face of $\Delta_{[1,3,4\mid 1,2,4]}({\ensuremath{\mathbf{Y}}})$ is shown, the elements of the face are indicated by bold dots. We will describe a modification of Viennot‘s light and shadow procedure (with the sun in the top-left corner) that produces a family of $n$ non-intersecting lattice paths such that the $i$^th^ path runs from $A^{(i)}=(0,u_{n+1-i})$ to $E^{(i)}=(a-v_{n+1-i},b)$ and has north-east turns only in $L^{(i)}$, for $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n}\else${1}$,~${1+1}$, \dots ,~${n}$\fi} \fi}\}$. Imagine a sun in the top-left corner of the ladder region and a wall along the lower-right border $B^{(1)}$ of $L^{(1)}$. Then each lattice point $(r,s)$ that is either in $B^{(1)}$ or corresponds to an element $x_{s,b-r}$ of the face casts a shadow $\{(x,y):x\ge r, y\le s\}$. The first path starts at $A^{(1)}$, goes along the north-east border of this shadow and terminates in $E^{(1)}$. In the left-most diagram of Figure \[fig:Paths\], this is accomplished for the face shown in Figure \[fig:LadderFace\].b. In the next step, we remove the wall on $B^{(1)}$ and all the elements of the face which correspond to lattice points lying on the first path. Then the procedure is iterated. See Figure \[fig:Paths\] for an example. Let $P$ be the resulting family of non-intersecting lattice paths. Now, for each $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n}\else${1}$,~${1+1}$, \dots ,~${n}$\fi} \fi}\}$, we remove all elements of the face except those which correspond to north-east turns of the $i$^th^ path and do not lie on $B^{(i)}$. In the example, $(5,8)$ is a north-east turn of the second path but lies on $B^{(2)}$, therefore the corresponding element $x_{1,5}$ of the face is removed. On the other hand, $(4,5)$ lies on $B^{(1)}$, but is a nort-east turn of the third path, so the corresponding element $x_{4,4}$ of the face is kept. This set of north-east turns defines another family of non-intersecting lattice paths $P^\prime$ that has the property that the $i$^th^ path has north-east turns only in $L^{(i)}$ for $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n}\else${1}$,~${1+1}$, \dots ,~${n}$\fi} \fi}\}$. We now want to count the number of faces of $\Delta_M({\ensuremath{\mathbf{Y}}})$ that reduce under light and shadow to a given family of lattice paths $P^\prime$ with this property. Clearly, $P^\prime$ can be translated into a family $P$ of non-intersecting lattice paths such that the $i$^th^ path does not go below $B^{(i)}$ for $i\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{n}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${n}$\fi} \else{\ifmmode{1},{1+1},\dots ,{n}\else${1}$,~${1+1}$, \dots ,~${n}$\fi} \fi}\}$. Note that the number of lattice points on such a family $P$ of paths is always equal to $d$, independently of the given face. Thus, if $m$ is the number of north-east turns of $P^\prime$, there are $$\binom{d-m}{k+1-m}$$ families of non-intersecting lattice paths $P$ that reduce to $P^\prime$. Hence, $f_k=\binom{d-m}{k+1-m}\size{{{\ensuremath{\mathcal{T}}}}^L_\ell({\ensuremath{\mathbf{A}}}\mapsto{\ensuremath{\mathbf{E}}})}$ and we obtain $$\begin{aligned} \sum_{\ell\ge0}\dim_K R_M({\ensuremath{\mathbf{Y}}})_\ell\,z^\ell &=\sum_{\ell\ge0}\bigg(\sum_{k\ge0}\binom{\ell-1}{k} f_k\bigg)z^\ell\\ &=\sum_{\ell\ge0}\sum_{k\ge0}\binom{\ell-1}{k} \bigg(\sum_{m=0}^{k+1}\binom{d-m}{k+1-m}\size{{{\ensuremath{\mathcal{T}}}}^L_m({\ensuremath{\mathbf{A}}}\mapsto{\ensuremath{\mathbf{E}}})}\bigg)z^\ell\\ &=\sum_{m\ge0}\size{{{\ensuremath{\mathcal{T}}}}^L_m({\ensuremath{\mathbf{A}}}\mapsto{\ensuremath{\mathbf{E}}})}\sum_{\ell\ge0}z^\ell \sum_{k\ge0}\binom{\ell-1}k\binom{d-m}{d-k-1}, \end{aligned}$$ and if we sum the inner sum by means of the Vandermonde summation (see for example [@GrahamKnuthPatashnik], Section 5.1, (5.27)), $$\begin{aligned} \sum_{\ell\ge0}\dim_K R_M({\ensuremath{\mathbf{Y}}})_\ell\,z^\ell&= \sum_{m\ge0}\size{{{\ensuremath{\mathcal{T}}}}^L_m({\ensuremath{\mathbf{A}}}\mapsto{\ensuremath{\mathbf{E}}})}\sum_{\ell\ge0}z^\ell\binom{d+\ell-m-1}{d-1}\\ &=\frac{\sum_{m\ge0}\size{{{\ensuremath{\mathcal{T}}}}^L_m({\ensuremath{\mathbf{A}}}\mapsto{\ensuremath{\mathbf{E}}})}z^m}{(1-z)^d}. \end{aligned}$$ Log-concavity of the $h$-vector in the case $M=[u_1\mid v_1]$ ============================================================= In this paper we will settle Stanley’s conjecture when $R$ is a ladder determinantal ring cogenerated by $M$, where $M$ is just a pair of integers, i.e., $n=1$. We want to stress, however, that data strongly suggest that Conca and Herzog’s conjecture is also true for arbitrary $n$. By the preceding theorem, in the case we are going to tackle, the sum $\sum_{i=0}^s h_i x^i$ that appears in the conjecture is the generating function $\sum_{k\ge0}\size{{{\ensuremath{\mathcal{T}}}}^L_k(A\mapsto E)}z^k$ of two-rowed arrays bounded by $A$ and $E$ which are in the ladder region $L$. As the bounds $A$ and $E$ will not be of any significance throughout the rest of this paper, we will abbreviate ${{\ensuremath{\mathcal{T}}}}^L_k(A\mapsto E)$ to ${{\ensuremath{\mathcal{T}}}}^L_k$. We will show that the $h$-vector is log-concave by constructing an injection from ${{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$ into ${{\ensuremath{\mathcal{T}}}}^L_k\times{{\ensuremath{\mathcal{T}}}}^L_k$. This injection will involve some cut and paste operations that we now define: Let $A$ and $X$ be two strictly increasing sequences of integers, such that the length of $X$ is the length of $A$ minus two, i.e., $A=(a_1,a_2,\dots,a_{k+1})$ and $X=(x_1,x_2,\dots,x_{k-1})$ for some $k\ge 1$. A [*cutting point of $A$ and $X$*]{} is an index $l\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{k}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${k}$\fi} \else{\ifmmode{1},{1+1},\dots ,{k}\else${1}$,~${1+1}$, \dots ,~${k}$\fi} \fi}\}$ such that $$\begin{aligned} \tag{$*$}\label{dfn:cut} a_l&<x_l,\\ \text{and}\quad x_{l-1}&<a_{l+1}, \end{aligned}$$ where we require the inequalities to be satisfied only if all variables are defined. Hence, $1$ is a cutting point if $a_1<x_1$, and $k$ is a cutting point if $x_{k-1}<a_{k+1}$. The [*image of $A$ and $X$ obtained by cutting at $l$*]{} is $$\begin{array}{cccc@{\big\vert\!\!}c@{\!\!\big\vert}cccccc} a_1&a_2&\dots&\multicolumn{1}{c}{a_{l-1}}&a_l &x_l &x_{l+1} &\dots&x_{k-1} \\[4pt] \hline\\[-8pt] x_1&x_2&\dots&x_{l-1}&\multicolumn{1}{c}{a_{l+1}}&a_{l+2}&\hdotsfor2&a_{k+1} \end{array}$$ Note that both the resulting sequences have length $k$. \[l:possible\] Let $A=(a_1,a_2,\dots,a_{k+1})$ and $X=(x_1,x_2,\dots,x_{k-1})$ be strictly increasing sequences of integers, such that the length of $X$ is the length of $A$ minus two. Then there exists at least one cutting point of $A$ and $X$. If $a_l\ge x_l$ for $l\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{k-1}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${k-1}$\fi} \else{\ifmmode{1},{1+1},\dots ,{k-1}\else${1}$,~${1+1}$, \dots ,~${k-1}$\fi} \fi}\}$ then $a_{k+1}>a_{k-1}\ge x_{k-1}$ and $k$ is a cutting point. Otherwise, let $l$ be minimal such that $a_l<x_l$. If $l=1$ then $1$ is a cutting point. Otherwise, because of the minimality of $l$, we have $a_{l+1}>a_{l-1}\ge x_{l-1}$, thus $l$ is a cutting point. Let $T=(T_1,T_2)\in{{\ensuremath{\mathcal{T}}}}_{k+1}\times{{\ensuremath{\mathcal{T}}}}_{k-1}$ be a pair of two-rowed arrays. Then a [*top cutting point of $T$*]{} is a cutting point of the top rows of $T_1$ and $T_2$ and a [*bottom cutting point of $T$*]{} is a cutting point of the bottom rows of $T_1$ and $T_2$. A pair $(l,m)$, where $l,m\in\{{\if\IsInteger{1}{\setcounter{seq}{1}\ifmmode{\arabic{seq}},\stepcounter{seq}{\arabic{seq}},\dots ,{k}\else${\arabic{seq}}$,~\stepcounter{seq} ${\arabic{seq}}$, \dots ,~${k}$\fi} \else{\ifmmode{1},{1+1},\dots ,{k}\else${1}$,~${1+1}$, \dots ,~${k}$\fi} \fi}\}$, such that $l$ is a top cutting point and $m$ is a bottom cutting point of $T_1$ and $T_2$ is a [*cutting point of $T$*]{}. Cutting the top rows of $T$ at $l$ and the bottom rows at $m$ we obtain the [*image of $T$*]{}. Note that both of the two-rowed arrays in the image have length $k$. More pictorially, if $l<m$, $$\begin{array}{ccccccccccc} a_1&\hdotsfor2&a_l&x_l&\hdotsfor1&x_{m-1}&\hdotsfor2&x_{k-1}\\ b_1&\hdotsfor3&b_{l+1}&\hdotsfor1&b_m&y_m&\dots&y_{k-1}\\[1pt] \hline x_1&\dots&x_{l-1}&a_{l+1}&\hdotsfor1&a_m&\hdotsfor3&a_{k+1}\\ y_1&\hdotsfor2&y_l&\hdotsfor1&y_{m-1}&b_{m+1}&\hdotsfor2&b_{k+1}, \end{array}$$ and similarly if $l\ge m$. For $T=(T_1,T_2)\in{{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$, the pair $(l,m)$ is an [*allowed cutting point of $T$*]{}, if both of the two-rowed arrays in the obtained image are in $L$. In Lemma \[l:allowed\] we will prove that every pair of two-rowed arrays in ${{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$ has at least one allowed cutting point. This motivates the following definition: Let $T=(T_1,T_2)\in{{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$ a pair of two-rowed arrays as before. Consider all allowed cutting points $(\bar l,\bar m)$ of $T$. Select those with $\size{\bar l-\bar m}$ minimal. Among those, let $(l,m)$ be the pair which comes first in the lexicographic order. Then we call $(l,m)$ the [*optimal cutting point*]{} of $T$. Now we are ready to state our main theorem, which implies that Stanley’s conjecture is true, when $R$ is a ladder determinantal ring cogenerated by a pair of integers $M$: \[thm:main\] Let $L$ be a ladder region. Let $T\in{{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$. Define $I(T)$ to be the pair of two-rowed arrays obtained by cutting $T$ at its optimal cutting point. Then $I$ is well-defined and an injection from ${{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$ into ${{\ensuremath{\mathcal{T}}}}^L_k\times{{\ensuremath{\mathcal{T}}}}^L_k$. \[cor:main\] The $h$-vector of the ladder determinantal ring cogenerated by $M=[u_1\mid v_1]$ is log-concave. By Theorem \[thm:comb\], the $h$-vector of this ring is equal to the generating function $\sum_{k\ge0}\size{{{\ensuremath{\mathcal{T}}}}^L_k(A\mapsto E)}z^k$ of two-rowed arrays bounded by $A=(0,u_1-1)$ and $E=(a-v_1+1,b)$ which are in the ladder region $L$. By the preceding theorem, there is an injection from ${{\ensuremath{\mathcal{T}}}}^L_{k+1}(A\mapsto E)\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}(A\mapsto E)$ into ${{\ensuremath{\mathcal{T}}}}^L_k(A\mapsto E)\times{{\ensuremath{\mathcal{T}}}}^L_k(A\mapsto E)$, thus $$\size{{{\ensuremath{\mathcal{T}}}}^L_{k+1}(A\mapsto E)}\cdot\size{{{\ensuremath{\mathcal{T}}}}^L_{k-1}(A\mapsto E)} \le\size{{{\ensuremath{\mathcal{T}}}}^L_k(A\mapsto E)}^2.$$ We will split the proof of Theorem \[thm:main\] in two parts. In Section \[s:welldef\] we show that the mapping $I$ is well-defined, that is, for any pair of two-rowed arrays $T\in{{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$ there is an allowed cutting point. Finally, in Section \[s:inj\], we show that $I$ is indeed an injection. The mapping $I$ is well-defined {#s:welldef} =============================== \[l:allowed\] Let $L$ be a ladder region. Then for every pair of two-rowed arrays in ${{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$ there is an allowed cutting point $(l,m)$. For the proof of this lemma, we have to introduce some more notation: Let $(T_1,T_2)\in{{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$ with $T_1=\left( \begin{smallmatrix} a_1&a_2&\dots&a_{k+1}\\ b_1&b_2&\dots&b_{k+1} \end{smallmatrix}\right)$ and $T_2=\left( \begin{smallmatrix} x_1&x_2&\dots&x_{k-1}\\ y_1&y_2&\dots&y_{k-1} \end{smallmatrix}\right)$. We say that Inequality  holds for an interval $[c,d]$ if $$\begin{aligned} \label{eq:topu}\tag{$\overline{top}$} {\overline{L}}(a_j)&\ge y_{j-1},\\ \intertext{for $j\in[c,d]$. Inequality~\eqref{eq:topl} holds for an interval $[c,d]$ if} \label{eq:topl}\tag{\underline{$top$}} {\underline{L}}(a_j)&\le y_{j-1},\end{aligned}$$ for $j\in[c,d]$. Similarly, Inequality  holds for an interval $[c,d]$ if $$\begin{aligned} \label{eq:bottomu}\tag{$\overline{bottom}$} {\overline{L}}(x_{j-1})&\ge b_j,\\ \intertext{for $j\in[c,d]$. Inequality~\eqref{eq:bottoml} holds for an interval $[c,d]$ if} \label{eq:bottoml}\tag{\underline{$bottom$}} {\underline{L}}(x_{j-1})&\le b_j,\end{aligned}$$ for $j\in[c,d]$, where ${\underline{L}}$ and ${\overline{L}}$ are as in Definition \[dfn:ladder\]. We say that any of these inequalities holds for a cutting point $(l,m)$ if it holds for the interval $[l+1,m]$ if $l<m$ and for the interval $[m+1,l]$ if $m<l$. Clearly, a cutting point $(l,m)$ is allowed if and only if all of these inequalities hold for it. Most of the work is done by the following lemma: \[l:imps\] Let $T=(T_1,T_2)\in{{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$, $T_1=\left( \begin{smallmatrix} a_1&a_2&\dots&a_{k+1}\\ b_1&b_2&\dots&b_{k+1} \end{smallmatrix}\right)$ and $T_2=\left( \begin{smallmatrix} x_1&x_2&\dots&x_{k-1}\\ y_1&y_2&\dots&y_{k-1} \end{smallmatrix}\right)$. Let ${\underline{l}}$ and ${\overline{l}}$ be top cutting points, such that there is no top cutting point in the closed interval $[{\underline{l}}+1,{\overline{l}}-1]$. Similarly, let ${\underline{m}}$ and ${\overline{m}}$ be bottom cutting points, such that there is no bottom cutting point in the closed interval $[{\underline{m}}+1,{\overline{m}}-1]$. Then for both of the intervals $[{\underline{l}}+1,{\overline{l}}]$ and $[{\underline{m}}+1,{\overline{m}}]$, - either or hold, - either or hold, - either or hold, - either or hold. Let $l_{min},l_{max},m_{min}$ and $m_{max}$ be the minimal and maximal top and bottom cutting points. Then we have - and hold for $[2,\max(l_{min},m_{min})]$ and - and hold for $[\min(l_{max},m_{max}),k]$. Suppose that does not hold for the interval $[{\underline{l}}+1,{\overline{l}}]$. We claim that in that case there is an index $j\in[{\underline{l}}+1,{\overline{l}}-1]$ such that $a_j<x_j$: For, by hypothesis there is an index $i\in[{\underline{l}}+1,{\overline{l}}]$ such that ${\overline{L}}(a_i)<y_{i-1}$. We have ${\overline{L}}(a_i)<y_{i-1}\le{\overline{L}}(x_{i-1})$ and because ${\overline{L}}$ is a weakly increasing function, $a_i<x_{i-1}$. It follows that $a_{i-1}<a_i<x_{i-1}<x_i$. Thus, if $i={\overline{l}}$ we choose $j=i-1$, otherwise $j=i$. The same statement is true if does not hold for the interval $[{\underline{l}}+1,{\overline{l}}]$: In this case there must be an index $i\in[{\underline{l}}+1,{\overline{l}}]$ such that ${\underline{L}}(x_{i-1})>b_i$. We conclude that ${\underline{L}}(a_i)\le b_i<{\underline{L}}(x_{i-1})$ and thus $a_i<x_{i-1}$. Next, we will use induction to prove that $$\begin{aligned} \tag{$**$} \label{eq:induction} a_{{l}} &< x_{{l}}\\ \text{and}\quad a_{{l}+1}&\le x_{{l}-1} \end{aligned}$$ for ${l}\in[{\underline{l}}+1,{\overline{l}}-1]$. We will first do an induction on ${l}$ to establish the claim for ${l}\in[j,{\overline{l}}-1]$. We start the induction at ${l}=j$: Above we already found that $a_j<x_j$. Therefore we must have $a_{j+1}\le x_{j-1}$, because otherwise $j$ would satisfy and hence were a top cutting point. Now suppose that holds for a particular ${l}<{\overline{l}}-1$. Then $a_{{l}+1}\le x_{{l}-1}<x_{{l}+1}$, and, because there is no top cutting point at ${l}+1$, we have $a_{{l}+2}\le x_{{l}}$. Similarly, to establish for ${l}\in[{\underline{l}}+1,j]$ we do a reverse induction on ${l}$. Suppose that holds for a particular ${l}>{\underline{l}}+1$. Then $a_{{l}-1}<a_{{l}+1}\le x_{{l}-1}$, and, because there is no top cutting point at ${l}-1$, we have $a_{{l}}\le x_{{l}-2}$. Thus we obtain $$\begin{aligned} &{\overline{L}}(x_{{\overline{l}}-1})\ge{\overline{L}}(x_{{\overline{l}}-2})\ge{\overline{L}}(a_{{\overline{l}}})\ge b_{{\overline{l}}} \text{, and}\\ &{\overline{L}}(x_{{l}-1})\ge{\overline{L}}(a_{{l}+1})\ge b_{{l}+1} \ge b_{{l}}, \end{aligned}$$ which means that holds for the interval $[{\underline{l}}+1,{\overline{l}}]$. Furthermore, $$\begin{aligned} &{\underline{L}}(a_{{\underline{l}}+1})\le{\underline{L}}(a_{{\underline{l}}+2})\le{\underline{L}}(x_{{\underline{l}}})\le y_{{\underline{l}}} \text{, and}\\ &{\underline{L}}(a_{{l}}) \le{\underline{L}}(x_{{l}-2})\le y_{{l}-2} \le y_{{l}-1}, \end{aligned}$$ which means that holds for the interval $[{\underline{l}}+1,{\overline{l}}]$. Next we show that and hold for the interval $[2,l_{min}]$: Assume that either of these inequalities does not hold for the interval $[2,l_{min}]$ and that $[2,l_{min}]$ does not contain a top cutting point except $l_{min}$. Then the above reverse induction implies that $a_1\le a_3<x_1$, which means that $1$ is a top cutting point. Thus, $l_{min}=1$ and the interval $[2,l_{min}]$ is empty. The other assertions are shown in a completely analogous fashion. We are now ready to establish Lemma \[l:allowed\]: Let $T=(T_1,T_2)\in{{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$. By Lemma \[l:possible\] there is at least one cutting point $(l,m)$ of $T$. Let $l_{min},l_{max},m_{min}$ and $m_{max}$ be the minimal and maximal top and bottom cutting points of $T$ as before. If there is an index $j$ which is a top *and* a bottom cutting point of $T$, then – trivially – $(j,j)$ is an allowed cutting point. Otherwise, we have to show that there is a cutting point $(l,m)$ for which , , , and hold. Suppose that this is not the case. For the inductive proof which follows, we have to introduce a convenient indexing scheme for the sequence of top and bottom cutting points. Let $$\begin{aligned} {2} m_{1,0}&=\max\{m:m<l_{min}\text{ and $m$ is a bottom cutting point}\},\\ m_{i,0}&=\max\{m:m<l_{i-1,1}\text{ and $m$ is a bottom cutting point}\}&\quad&\text{for $i>1$,}\\ \text{and}\quad l_{i,0}&=\max\{l:l<m_{i,1}\text{ and $l$ is a top cutting point}\} &\quad&\text{for $i\ge 1$,} \end{aligned}$$ where $m_{i,j+1}$ is the bottom cutting point directly after $m_{i,j}$, and $l_{i,j+1}$ is the top cutting point directly after $l_{i,j}$. Furthermore, we set $l_{0,1}=l_{min}$. More pictorially, we have the following sequence of top and bottom cutting points for $i\ge 1$: $$\dots<m_{i,0}<l_{i-1,1}<l_{i-1,2}<\dots<l_{i,0}<m_{i,1}<m_{i,2} <\dots<m_{i+1,0}<\cdots$$ If $m_{min}>l_{min}$, then $m_{1,0}$ does not exist, of course. Note that there are no bottom cutting points between $l_{i,1}$ and $l_{i+1,0}$, and there are no top cutting points between $m_{i,1}$ and $m_{i+1,0}$. Suppose first that $m_{min}<l_{min}$. By induction on $i$, we will show that and hold for the cutting points $(l_{i-1,1},m_{i,0})$, where $i\ge 1$. By Lemma \[l:imps\] we know that and are satisfied for the cutting point $(l_{min},m_{1,0})$, because $[m_{1,0}+1,l_{min}]\subseteq [2,l_{min}]$. It remains to perform the induction step, which we will divide into five simple steps. *Step 1. and hold for the interval $[m_{i,0}+1,l_{i-1,1}]$.* This is just a restatement of the induction hypothesis, i.e., that and hold for the cutting point $(l_{i-1,1},m_{i,0})$. *Step 2. Either or does not hold for the interval $[m_{i,0}+1,m_{i,1}]$.* Because of Step 1, not both of and can hold for $(l_{i-1,1},m_{i,0})$, lest this was an allowed cutting point. Thus either or does not hold for $[m_{i,0}+1,l_{i-1,0}+1]$. This interval is contained in $[m_{i,0}+1,m_{i,1}]$, thus the inequalities and cannot hold on this interval either. *Step 3. and hold for $[l_{i,0}+1,m_{i,1}]$.* Suppose that does not hold for $[m_{i,0}+1,m_{i,1}]$. Then, by Lemma \[l:imps\] we obtain that and hold for $[m_{i,0}+1,m_{i,1}]$, because this interval contains no bottom cutting points except $m_{i,1}$. The same is true, if does not hold for $[m_{i,0}+1,m_{i,1}]$. Because $[l_{i,0}+1,m_{i,1}]$ is a subset of this interval, and hold for the cutting point $(l_{i,0},m_{i,1})$, or, equivalently, for the interval $[l_{i,0}+1,m_{i,1}]$. *Step 4. Either or does not hold for $[l_{i,0}+1,l_{i,1}]$.* Because of Step 3, not both of and can hold for the cutting point $(l_{i,0},m_{i,1})$, nor for the greater interval $[l_{i,0}+1,l_{i,1}]$. *Step 5. and hold for $[m_{i+1,0}+1,l_{i,1}]$.* The interval $[l_{i,0}+1,l_{i,1}]$ does not contain a top cutting point except $l_{i,1}$, thus by Lemma \[l:imps\] and Step 4 we see that and hold. Finally, because $[m_{i+1,0}+1,l_{i,1}]\subset [l_{i,0}+1,l_{i,1}]$, and hold for the cutting point $(l_{i,1},m_{i+1,0})$. If $l_{max}>m_{max}$, then we encounter a contradiction: Let $r$ be such that $m_{r,0}=m_{max}$. We have just shown that and hold for the cutting point $(l_{r-1,1},m_{r,0})$. Furthermore, by Lemma \[l:imps\], and hold for $[m_{r,0},k]$ and thus also for $(l_{r-1,1},m_{r,0})$. Hence, this would be an allowed cutting point, contradicting our hypothesis. If $l_{max}<m_{max}$, let $r$ be such that $l_{r,0}=l_{max}$. By the induction (Step 3) we find that and hold for the cutting point $(l_{r,0},m_{r,1})$. Again, because of Lemma \[l:imps\], we know that and holds for $[l_{r,0},k]$ and thus also for $(l_{r,0},m_{r,1})$. Hence, we had an allowed cutting point in this case also. The case that $m_1>l_1$ is completely analogous. The mapping $I$ is an injection {#s:inj} =============================== The mapping $I$ defined above is an injection. Suppose that $I(T)=I(T^\prime)$ for $T=(T_1,T_2)$ and $T^\prime=(T_1^\prime,T_2^\prime)$, such that $T$ and $T^\prime$ are elements of ${{\ensuremath{\mathcal{T}}}}^L_{k+1}\times{{\ensuremath{\mathcal{T}}}}^L_{k-1}$. Let $(l,m)$ be the optimal cutting point of $T$, and let $(l^\prime,m^\prime)$ be the optimal cutting point of $T^\prime$. Observe that we can assume $\min(l,m,l^\prime,m^\prime)=1$, because the elements of $T$ and $T^\prime$ with index less than or equal to this minimum retain their position in $I(T)$. Likewise, we can assume that $\max(l,m,l^\prime,m^\prime)=k$. Furthermore, we can assume that $l\le l^\prime$, otherwise we exchange the meaning of $T$ and $T^\prime$. Thus, we have to consider the following twelve situations: $$\begin{array}{>(r<{)\quad} @{1=\:} c !{\le} c !{\le} c !{\le} c @{\:=k}} 1 & l & l^\prime & m & m^\prime\\ 2 & l & l^\prime & m^\prime & m\\ 3 & l & m & l^\prime & m^\prime\\ 4 & l & m & m^\prime & l^\prime\\ 5 & l & m^\prime & l^\prime & m\\ 6 & l & m^\prime & m & l^\prime\\ 7 & m & l & l^\prime & m^\prime\\ 8 & m & l & m^\prime & l^\prime\\ 9 & m & m^\prime & l & l^\prime\\ 10 & m^\prime & l & l^\prime & m\\ 11 & m^\prime & l & m & l^\prime\\ 12 & m^\prime & m & l & l^\prime \end{array}$$ We shall divide these twelve cases into two portions according to whether $l\le m$ or not. In all the other cases the reasoning is very similar. Thus we only print the pairs of two-rowed arrays $T^\prime$ and $\tilde T$ and leave it to the reader to check that $\tilde T$ is in the ladder region. [10]{} Shreeram Abhyankar, *Enumerative combinatorics of [Young]{} tableaux*, Marcel Dekker, New York, 1988. Shreeram Abhyankar and Devadatta M. Kulkarni, *On [Hilbertian]{} ideals*, Linear Algebra and its Applications **116** (1989), 53–79. Francesco Brenti, *Log-concave and unimodal sequences in algebra, combinatorics, and geometry: an update*, Contemporary Mathematics **178** (1994), 71–89. W. Bruns and J[ü]{}rgen Herzog, *On the computation of $a$-invariants*, Manuscripta mathematica **77** (1992), 201–213. A. Conca, *Ladder determinantal rings*, Journal of Pure and Applied Algebra **98** (1995), 119–134. A. Conca and Jürgen Herzog, *On the [Hilbert]{} function of determinantal rings and their canonical module*, Proceedings of the American Mathematical Society **122** (1994), 677–681. R. L. Graham, Donald E. Knuth, and Oren Patashnik, *Concrete mathematics*, Addison-Wesley, Reading, Massachusetts, 1989. J[ü]{}rgen Herzog and Ng[ô]{} Vi[ê]{}t Trung, *Gröbner bases and multiplicity of determinantal and [Pfaffian]{} ideals*, Advances in Mathematics **96** (1992), no. 1, 1–37. Christian Krattenthaler and Martin Rubey, *A determinantal formula for the [Hilbert]{} series of one-sided ladder determinantal rings*, preprint (2001). Devadatta M. Kulkarni, *[Hilbert]{} polynomial of a certain ladder determinantal ideal*, Journal of Algebraic Combinatorics **2** (1993), 57–71. [to3em]{}, *Counting of paths and coefficients of [Hilbert]{} polynomial of a determinantal ideal*, Discrete Mathematics **154** (1996), 141–151. G. Niesi and L. Robbiano, *Disproving [Hibi’s]{} conjecture with projective curves with bad [Hilbert]{} functions*, Computational Algebraic Geometry (Boston) (F. Eyssette and A. Galligo, eds.), Progress in Mathematics, no. 109, Birkh[ä]{}user, 1993, pp. 195–201. Richard P. Stanley, *Log-concave and unimodal sequences in algebra, combinatorics, and geometry*, Annals of the New York Academy of Sciences **576** (1989), 500–535. [^1]: In honour of Miriam Rubey, at the occasion of her second birthday
{ "pile_set_name": "ArXiv" }
--- abstract: 'An ordered structure is called [o-minimalistic]{} if it has all the first-order features of an o-minimal structure. We propose a theory, ${{\text{DCTC}}}$ (Definable Completeness/Type Completeness), that describes many properties of [o-minimalistic]{} structures (dimension theory, monotonicity, Hardy structures, quasi-cell decomposition). Failure of cell decomposition leads to the related notion of a [tame]{} structure, and we give a criterium for an [o-minimalistic]{} structure to be [tame]{}. To any [o-minimalistic]{} structure, we can associate its [Grothendieck ring]{}, which in the non-o-minimal case is a non-trivial invariant. To study this invariant, we identify a third [o-minimalistic]{} property, the Discrete Pigeonhole Principle, which in turn allows us to define discretely valued Euler [characteristic]{}[s]{}.' author: - Hans Schoutens title: 'O-minimalism' --- Introduction ============ O-minimality has been studied extensively (see [@vdDomin] for some of the literature). It also has been generalized in many ways (weak o-minimality, quasi-o-minimality, d-minimality, local o-minimality, o-minimal open cores, etc.) These generalizations attempt to bring into the fold certain ordered structures that fail some of the good finiteness properties of o-minimality, but still behave tamely.[^1] We offer a different perspective in this paper, where our point of departure is the observation that, in contrast to ultra[power]{}s, an ultra[product]{} of o-minimal structures need no longer be o-minimal; let us call it *ultra-o-minimal*. This leads to two natural questions: (i) under which conditions on the o-minimal components is an ultra-o-minimal structure again o-minimal? And (ii), what properties do ultra-o-minimal structures have? We will give one answer to (i) by proving a criterion in terms of Euler [characteristic]{}[s]{} in Theorem \[T:ominul\] below, but we will be mainly concerned with (ii). Given a language $L$ with an order relation, let ${T^{\text{omin}}}:={T^{\text{omin}}}(L)$ be the intersection of the theories of all o-minimal $L$-structures. Models of ${T^{\text{omin}}}$ will be called *[o-minimalistic]{}*; they are precisely the elementary substructures of ultra-o-minimal ones. O-minimalism is in essence a non-standard feature, as any [o-minimalistic]{} expansion of the reals is already o-minimal (Corollary \[C:realsomin\]). In the first half of this paper, we will focus on two elementary properties, *definable completeness* (=every definable subset has an infimum) and *type completeness*[^2] (=every one-sided type of a point, including the ones at infinity, is complete). We denote by ${{\text{DCTC}}}$ these axiom schemes on one-variable definable sets (where the dependence on parameters has to be quantified out to get sentences in the language $L$). I do not know whether ${{\text{DCTC}}}$ is equal to ${T^{\text{omin}}}$, but in §\[s:gr\], I will formulate a third [o-minimalistic]{} (first-order) property, the Discrete Pigeonhole Principle (DPP=any definable, injective map from a discrete set to itself is bijective), which as of yet, I do not know how to derive from ${{\text{DCTC}}}$. In fact, it is not clear if we can axiomatize o-minimalism by first-order conditions on one-variable formulae only (note that DPP is a priori not of this form). But even without knowing fully what the theory ${T^{\text{omin}}}$ of o-minimalism is, we can deduce at least its first-order properties. To give just an example, suppose $\mathcal M$ expands an ordered field. If it is o-minimal, then every definable map is continuous, and in fact, differentiable, at all but finitely many points. This is not an elementary statement, so we have to decide with which first-order concept we should replace finiteness (continuity and differentiability are elementary expressible). As we shall argue below, the right candidate is discreteness (boundedness and closedness then follow automatically). So, for each formula $\varphi(x,y,{\mathbf {z}})$ in $n+2$ variables, we can check, for a given $n$-tuple ${\mathbf {c}}$, whether the set defined by $\varphi(x,y,{\mathbf {c}})$ is the graph of a map $f$, and then we can express that, if $f$ is not differentiable at $x$, it must be at every other point in some neighborhood of $x$ where $f$ is defined. It follows that in an [o-minimalistic]{} expansion of an ordered field, every definable map is differentiable outside a bounded, closed, discrete subset (whence outside some finite subset on any compact interval). In fact, this can be proven already within ${{\text{DCTC}}}$ by mimicking the proof in [@vdDomin Chapt. 7]. So this paper investigates properties of [o-minimalistic]{} structures that are modified versions of the corresponding properties of o-minimal structures. Therefore, whereas most papers on generalizing o-minimality are searching for weakenings that would include certain tamely behaving structures, our hands are tied and we have to obey by the properties of o-minimalism. Thus, to the chagrin of some of my esteemed colleagues, this means that we have to discard the structure $(\mathbb Q,<,+,\mathbb Z)$ as it is not [o-minimalistic]{}, although it is definably complete and locally o-minimal. However, it fails to have the type completeness property at infinity, which forces every discrete set to be bounded. We derive most results already from ${{\text{DCTC}}}$—albeit only in detail for two variables, leaving higher arities to the reader, commenting on it occasionally—such as the Monotonicity Theorem (Theorem \[T:disctu\]), Fiber Dimension Theorem (Corollary \[C:fibdim\]), Quasi-Cell Decomposition (Theorem \[T:plane\]), Hardy structures on germs at infinity (Theorem \[T:Hardy\]), definable simpleness of [o-minimalistic]{} expansions of groups (Theorem \[T:Dedlimgr\]), etc. These are the analogues of the o-minimal concepts except that ‘finite’ has to be replaced by ‘discrete’ (which, as already remarked, always implies bounded and closed). This program, however, does not always pan out. For instance, while decomposing into cells, we seem to run into infinite disjunctions, leading to the notion of a *quasi-cell*, which is only locally a cell. However, there is a large class of definable subsets, called *[tame]{}* subsets, that have a ‘definable’ cell decomposition, that is to say, loosely speaking, they admit a cell decomposition in ‘discretely’ many cells (see §\[s:tame\] for the precise definition). A *[tame]{}* structure is then one in which every definable subset is [tame]{}, and we show that it is always at least a model of ${{\text{DCTC}}}$ (at present I have no examples of a non-[o-minimalistic]{} [tame]{} structure, nor of a non-[tame]{} [o-minimalistic]{} structure, but presumably these are different notions). Any [o-minimalistic]{} expansion of a real closed field by one-variable functions is [tame]{}, or more generally, by functions having only a discrete set of discontinuities. In §\[s:gr\], we study the [Grothendieck ring]{} of an [o-minimalistic]{} structure: it is equal to the ring of integers [if and only if]{} the structure is o-minimal (in which case it corresponds to the Euler [characteristic]{}). In the last section, we give some criteria for an expansion of an [o-minimalistic]{} structure by a set to be again [o-minimalistic]{}. For discrete subsets, this leads to the notion of an *[o-finitistic]{}* set, that is to say, a set enjoying all first-order properties of an arbitrary finite set in an o-minimal structure. This notion is particularly interesting when it comes to classifying definable subsets up to ‘virtual’ isomorphism, that is to say, definable in some [o-minimalistic]{} expansion; the corresponding [Grothendieck ring]{} is then called the *virtual [Grothendieck ring]{}*. However, a priori, the treatment depends on a choice of ‘context’, that is to say, of an ultra-o-minimal elementary extension. Using this technology, we can now associate to each definable, discrete subset of $M$ a (discretely valued) Euler [characteristic]{} defined on its virtual [Grothendieck ring]{}. This allows us to calculate explicitly this virtual [Grothendieck ring]{} in the special case of a [tame]{}, [o-minimalistic]{} expansion of an ordered field admitting a power dominant discrete subset (Corollary \[C:gromintame\]). The last section is an application to the study of analytic sets. In the o-minimal context, (sub)analytic sets are normally understood to be given by analytic functions supported on the unit box (often simply called *restricted analytic functions*), as the corresponding structure ${\mathbb R}_{\text{an}}$ is o-minimal, and admits quantifier elimination in an appropriate language by the seminal work of [@DvdD]. There is a good reason to restrict to compact support, as the global sine function defines ${\mathbb Z}$, and hence can never be part of an o-minimal expansion. Our approach here is to look at subsets of ${\mathbb R}^k$ that can be uniformly approximated on compact sets by ${\mathbb R}_{\text{an}}$-definable subsets. More precisely, we call a subset $X\sub {\mathbb R}^k$ a *Taylor* set, if the ultraproduct over all $n$ of the truncations ${ X_{\downharpoonleft_{n}}}:=\{{\mathbf {x}}\in X| {\left|{\mathbf {x}}\right|}\leq n\}$ is definable in ${{\mathcal R^{\text{an}}}_\natural}$, where the latter structure is obtained as the ultraproduct of the scalings of ${\mathbb R}_{\text{an}}$ by a factor $n$ (that is to say, for each $n$, the expansion of ${\mathbb R}$ by power series converging on $ {\left|{\mathbf {x}}\right|}\leq n$). Any subset definable by a quantifier free formula using convergent power series, whence in particular, any globally analytic variety, is [Taylor]{}. A discrete subset is [Taylor]{} [if and only if]{} it is closed, and any such set satisfies the Discrete Pigeonhole Principle with respect to [Taylor]{} maps. However, we can now also define sets by analytic parameterization, like the spiral with polar coordinates $R=\exp\theta$, for $\theta>0$ (in contrast, the spiral obtained by allowing $\theta$ to be negative as well is not [Taylor]{}!). We use our [o-minimalistic]{} results—which, as noted above, are a priori non-standard in nature—to give a geometric treatment of the class of [Taylor]{} sets: to a [Taylor]{} set $X$, we associate an ${{\mathcal R^{\text{an}}}_\natural}$-definable subset ${{X}\mathstrut_\flat}$, called its *protopower*, given as the ultraproduct of its truncations. We obtain a good dimension theory, a monotonicity theorem, a (partly conjectural, locally finite) cell-decomposition, and a corresponding [Grothendieck ring]{}, all indicative of the tameness of the class of [Taylor]{} sets, albeit not first-order. Notations and conventions {#notations-and-conventions .unnumbered} ------------------------- Definable means definable with parameters, unless stated explicitly otherwise. Throughout this paper, $L$ denotes some language containing a distinguished binary relation symbol $<$ and any $L$-structure $\mathcal M$ will be a dense linear order without endpoints. We introduce two new symbols $-\infty$ and $\infty$, and, given an $L$-structure $\mathcal M$, we let $M_\infty:=M\cup\{\pm\infty\}$. When needed, $\tt U$ denotes some predicate (often unary), and we will write $(\mathcal M,X)$ for the $L(\tt U)$-structure in which $X$ is the interpretation of $\tt U$. We will use the following ISO convention for intervals: *open* ${{\, ]a,b[\, }}$ (which we always assume to be non-empty, that is to say, $a<b$), *closed* ${{\, [a,b]\, }}$ (including the singleton $\{a\}={{\, [a,a]\, }}$), *half-open* ${{\, ]a,b]\, }}$ or ${{\, [a,b[\, }}$, and their infinite variants like ${{\, ]-\infty,a[\, }}$, ${{\, ]-\infty,a]\, }}$, ${{\, ]a,\infty[\, }}$, and ${{\, [a,\infty[\, }}$, with $a,b\in M$. Note that the usage of $\infty$ here is only informal since these are definable subsets in the language without the extra constants $\pm\infty$ by formulae of the form $x<a$, etc.: any interval is definable (with parameters). The union and the intersection of two non-disjoint intervals are again intervals. Note that in $\mathbb Q$ the set of all rational numbers $q$ with $3<q<\pi$ is not an interval, as it is only an infinite conjunction of definable subsets. Given a subset $Y\sub M$ and a point $b\in M$, we will sometimes use notations like $Y_{<b}:=Y\cap {{\, ]-\infty,b[\, }}$. When taking ultraproducts, we rarely ever mention the underlying index set or (non-principal) ultrafilter. We use the notation introduced in [@SchUlBook], denoting ultraproducts with a subscript $\natural$. Thus, we write ${{\mathbb N}_\natural}$, ${{\mathbb Z}_\natural}$, and ${{\mathbb R}_\natural}$ for the ultrapower of the set of natural numbers ${\mathbb N}$, integers ${\mathbb Z}$, and reals ${\mathbb R}$ respectively. On occasion we need the (countable) ultraproduct of the diagonal sequence $(n)_n$ in ${{\mathbb N}_\natural}$, which we denote suggestively by ${{\omega_\natural}}$. O-minimality ============ \[D:omin\] An $L$-structure $\mathcal M$ is called *o-minimal*, if every definable subset $ Y\sub M$ is a finite union of open intervals and points. In other words, any $L$-formula (with parameters) in one free variable is equivalent with a (quantifier free) formula using only $<$ (and parameters). The main feature of o-minimality is that this dearth (up to equivalence) of one-variable formulae puts severe restrictions on its many-variable formulae. Real closed fields are o-minimal; for more examples, we refer to the (vast) literature ([@vdDomin] is a good place to start). \[L:ded\] Given a definable subset $Y\sub M$ in an o-minimal structure $\mathcal M$, the infimum and the supremum of $Y$ exist in $M_\infty$. Follows from the fact that the endpoints of an interval are its infimum and supremum. Recall the convention that the infimum and the supremum of the empty set are respectively $\infty$ and $-\infty$. \[E:nonomin\] Because of quantifier elimination, the structure $(\mathbb Q,<)$ is o-minimal. Expanding this to the ordered field $(\mathbb Q, +,-,\cdot;<;0,1)$, however, destroys this. Indeed, the set of all $x>0$ such that $x^2>2$ has no infimum in $\mathbb Q$ (its infimum in ${\mathbb R}$, of course, is $\sqrt 2$), contradicting Lemma \[L:ded\]. We will always view an ordered structure in its *order topology* in which the basic open subsets are the open (possibly unbounded) intervals. Note that this is always a Hausdorff topology by density. The Cartesian powers $M^n$ are equipped with the product topology, for which the open boxes form a basis, where an *open box* is any product of open intervals. Recall that the *interior* $Y^\circ$ of a subset $Y$ in a topological space $M$ consists of all points $x\in Y$ for which there exists an open $U\sub Y$ with $x\in U$; the *exterior* is the interior of the complement $M \setminus Y$; the *closure* $\bar Y$ is the complement of the exterior; the *frontier* ${{\operatorname}{fr}( Y)}$ is the difference $\bar Y\setminus Y$; and the *boundary* is the difference $\partial Y:=\bar Y\setminus Y^\circ$. In the ordered case, if $Y\sub M$ is definable, say, by a formula $\phi(x)$, then its interior is also definable, namely by the formula $\phi^\circ(x)$ given as the conjunction of $\phi(x)$ and the formula $\exists z_1<x<z_2, \forall y: z_1<y<z_2 {\to}\phi(y)$; and likewise for its closure, exterior, frontier, and boundary. Hence if $\mathcal M$ is o-minimal, then $Y^\circ$ is a finite union of open intervals, and $\partial Y$ is finite. A definable subset $X\sub M^n$ is called *definably connected* if it can not be written as a disjoint union of two open definable subsets of $X$. The image of a definably connected subset under a definable and continuous function is definably connected. Any interval is definably connected. However, this notion does not always agree with the notion of topological connectedness: the interval ${{\, ]3,4[\, }}$ in $\mathbb Q$ is definably connected, but the decomposition in the two disjoint (non-definable) opens $3<x<\pi$ and $\pi<x<4$ shows that it is not topologically connected (in fact, $\mathbb Q$ is totally disconnected as a topological space). Over the reals, by Dedekind completeness, both notions agree. It is an easy exercise that in an o-minimal structure, a subset of $M$ is definably connected [if and only if]{} it is an interval. The theory ${{\text{DCTC}}}$ ============================ Given an $L$-formula $\varphi({\mathbf {z}}, x)$ in $n+1$ variables, we will associate several formulae to it with the intention to express certain properties of o-minimality (here the ${\mathbf {z}}$ will be thought of as parameters). Each formula comes with a dual formula upon reversing the order (or, equivalently, by taking the negation), but since we will not need most of these dual ones, we will omit them here. First, we express the Dedekind closure property from Lemma \[L:ded\], that is to say, we construct a formula $\varphi^{\text{inf}}({\mathbf {z}})$ expressing for each tuple of parameters ${\mathbf {z}}$ that $Y:=\varphi({\mathbf {z}},\mathcal M)$ has an infimum in the $L$-structure $\mathcal M$. We allow here infinity as a value, so we will have to treat this case separately. Let $\psi({\mathbf {z}})$ be the formula saying that for each $y$ there exists $x<y$ such that $\varphi({\mathbf {z}}, x)$, expressing that the infimum of $Y$ is $-\infty$. Then $\varphi^{\text{inf}}$ is the disjunction of $\psi$ and the formula saying that there exists $x$ such that $x\leq y$ for all $y\in Y$ and such that if $t$ is another element satisfying that $t\leq y$ for all $y\in Y$, then $t\leq x$ (in other words, $x=\inf(Y)$). By Lemma \[L:ded\], any o-minimal structure satisfies $$\label{eq:infsup} (\forall{\mathbf {z}})\varphi^{\text{inf}}({\mathbf {z}}).\tag{DC}$$ Next we express the following left limit property of intervals: for any point there is an open interval to its left which is either entirely contained or entirely disjoint from the given interval. In terms of characteristic functions of definable subsets of $M$, this says that any such function has a left limit in each point. More precisely, let $\varphi^-({\mathbf {z}}, x)$ be the formula expressing that there exists $y<x$ such that either ${{\, ]y,x[\, }}$ is contained in or disjoint from $Y$. Again, we must treat infinity separately: let $\varphi^{-\infty}({\mathbf {z}})$ express the fact that there exists $y$ such that ${{\, ]-\infty,y[\, }}$ is either contained in or disjoint from $Y$. Since these properties hold for any type of interval, and are preserved under finite unions, we showed that in any o-minimal structure, we have $$\label{eq:event} (\forall x,{\mathbf {z}})\varphi^-({\mathbf {z}}, x){\wedge}\varphi^{-\infty}({\mathbf {z}}).\tag{TC}$$ \[D:ominic\] For a fixed language with a binary order symbol $<$, we define the theory ${{\text{DCTC}}}$ as the extension of the theory of dense linear orders without endpoints by the two axiom schemes given by (*definable completeness*) and (*type completeness*), with $\varphi$ running over all $L_{n+1}$-formulae. Put differently, in a model $\mathcal M$ of ${{\text{DCTC}}}$, any definable subset $Y\sub M$ has an infimum and its [characteristic]{} function has a left limit at each point. More generally, we call any model of the first-order theory ${T^{\text{omin}}}$ of the class of o-minimal $L$-structures an *[o-minimalistic]{}* structure. In other words, an $L$-structure is [o-minimalistic]{}, if it satisfies every sentence in $L$ that holds in every o-minimal $L$-structure. Since this includes and , an [o-minimalistic]{} structure is a model of ${{\text{DCTC}}}$. The converse is not clear, although we will see below an axiom, , which is [o-minimalistic]{}, but is currently not known to follow from ${{\text{DCTC}}}$. \[L:ominred\] A reduct of an [o-minimalistic]{} structure is again [o-minimalistic]{}. Let $L\sub L'$ be languages, let $\mathcal M'$ be an [o-minimalistic]{} $L'$-structure, and let $\mathcal M:={\left.\mathcal M'\right|_{{L}}}$ be its $L$-reduct. To show that $\mathcal M$ is [o-minimalistic]{}, take a sentence in ${T^{\text{omin}}}_L$ and let $\mathcal N'$ be any o-minimal $L'$-structure. Since its reduct ${\left.\mathcal N'\right|_{{L}}}$ is also o-minimal, $\sigma$ holds in the latter, whence also in $\mathcal N'$ itself. As this holds for all o-minimal $L'$-structures, $\sigma$ also holds in $\mathcal M'$. Since $ \sigma$ only mentions $ L$-symbols, it must therefore already hold in the reduct $\mathcal M$, as we needed to show. We will call an ultraproduct of o-minimal $L$-structures an *ultra-o-minimal* structure. Using a well-known elementarity criterion via ultraproducts, we have: \[C:ulomin\] An $L$-structure is [o-minimalistic]{} [if and only if]{} it is elementary equivalent with (equivalently, an elementary substructure of) an ultra-o-minimal structure. This produces many examples of [o-minimalistic]{} structures which fail to be o-minimal. \[E:ulomin\] Let $L$ be the language of ordered fields together with a unary predicate $\tt U$ For each $n$, let $\mathcal R_n:=({\mathbb R},\{0,1,\dots,n\}$ be the expansion of field ${\mathbb R}$. Since $\{0,1,2,\dots,n\}$ is finite whence definable without the predicate $\tt U$, each $\mathcal R_n$ is o-minimal, and therefore their ultraproduct ${\mathcal R_\natural}$ is [o-minimalistic]{}  by Corollary \[C:ulomin\]. The set $D:=\tt U({\mathcal R_\natural})$ is discrete but not finite, so ${\mathcal R_\natural}$ cannot be o-minimal. Note that $D$ contains ${\mathbb N}$ and that ${{\omega_\natural}}$ is its maximum. In fact, $D=({{\mathbb N}_\natural})_{\leq{{\omega_\natural}}}$. By [@MillIVT Corollary 1.5], the definable completeness axiom  is equivalent with $M$ itself being definably connected, and also with the validity of the *Intermediate Value Theorem* (IVT is the statement that if $f\colon {{\, [a,b]\, }}\to M$ is a definable and continuous function, then $f$ assumes each value between $f(a)$ and $f(b)$). Since an ordered field (with no additional structure) is real closed [if and only if]{} it is o-minimal, [if and only if]{} it satisfies IVT for polynomials, we get: \[C:ominfield\] An [o-minimalistic]{} ordered (pure) field is o-minimal. This, however, tells us nothing about proper expansions of ordered fields. We can reformulate axiom  as follows, justifying its name. Given $Y\sub M$ and $a\in M$, we say that *$a^-$ belongs to $Y$*, if there exists an open interval ${{\, ]b,a[\, }}\sub Y$ (similarly, *$a^+$ belongs to $Y$*, if ${{\, ]a,b[\, }}\sub Y$ for some $b>a$). Thinking of $a^-$ as a partial type (that is to say, consisting of all formulae $b<x<a$ in the variable $x$, where $b$ runs over all elements less than $a$), if $Y$ is defined by a formula $\varphi$, then $a^-$ belongs to $Y$ [if and only if]{}  any realization of the type of $a^-$ in any elementary extension of $\mathcal M$ satisfies $\varphi$. Therefore, says that if $Y$ is definable, then $a^-$ belongs either to $Y$ or to $M\setminus Y$, for any $a\in M$ (and the same assertion for $a^+$), or equivalently, that $a^-$ is a complete type. We extend this terminology to the two types $(-\infty)^+$ and $\infty^-$ in the obvious way. Recall that an $L$-structure $\mathcal M$ is called *locally o-minimal*, if for each definable subset $Y\sub M$ and each $a\in Y$, there exists an open interval $I$ containing $a$ such that $Y\cap I$ is a finite union of intervals. If this is the case, then we may shrink $I$ so that $I\cap Y$ is an interval, from which it is now clear that this is equivalent with at $a$. In other words, we showed the first part of the following result, whereas the second is a well-known consequence whose proof we just reproduce here (for more on local o-minimality, see[@ToffVoz; @VozThesis]): \[P:locomin\] An [o-minimalistic]{} structure $\mathcal M$ is locally o-minimal. Let $K\sub M$ be compact and $Y\sub M$ definable. If either $K$ is open or $Y$ is contained in $ K$, then $K\cap Y$ is a finite union of intervals. Given a definable subset $Y\sub M$, by assumption, we can find in the open case, for each $x\in K$, an open interval $I_x\sub K$ such that $Y\cap I_x$ is an interval. Since $K$ is compact and the $I_x$ cover $K$, there exist finitely many points $x_1,\dots,x_n\in K$ such that $K= I_{x_1}\cup\dots\cup I_{x_n}$ and hence $K\cap Y$ is a finite union of intervals. If $K$ is arbitrary, then we cannot arrange for all $I_x$ to be contained in $K$, and so we only get $K\sub I_{x_1}\cup\dots\cup I_{x_n}$. But since $Y\sub K$, the same conclusion can be drawn. The next corollary improves [@DMS 2.13(3)] as it does not assume any underlying field structure. \[C:realsomin\] If $\mathcal M$ is an [o-minimalistic]{} structure with underlying order that of the reals, then it is o-minimal. Identify $M$ with ${\mathbb R}$, and let $Y\sub {\mathbb R}$ be definable. Depending whether $(-\infty)^+$ or $\infty^-$ belong to $Y$ or not, we may assume after possibly removing one or two unbounded intervals that $Y$ is bounded, whence contained in some closed, bounded interval $K:={{\, [a,b]\, }}$. Hence $Y=Y\cap K$ is a finite union of intervals by Proposition \[P:locomin\]. \[R:realsomin\] From the proof it is clear that we have the following more general result: if an [o-minimalistic]{} structure has the Heine-Borel property, meaning that any closed bounded set is compact, then it is o-minimal. Note that is stronger than local o-minimality, since we also have this condition at $\pm \infty$, which seems to the author an omission in the original definition of [@ToffVoz]. Notwithstanding, since local o-minimality is mostly studied in expansions of ordered fields, the condition at $\infty$ then also holds since we can map the latter to zero by taking reciprocals and apply local o-minimality there. We start with gathering some easy facts on one-variable definable subsets in an [o-minimalistic]{} structure. \[P:Dedlim\] For a definable subset $Y\sub M$ in an [o-minimalistic]{} structure $\mathcal M$ (or, more generally, in any model of ${{\text{DCTC}}}$), we have: 1. \[i:infbound\] The infimum of $Y$ is either infinite or a topological boundary point. 2. \[i:intint\] If $a,b\in \partial Y$ and ${{\, ]a,b[\, }}\cap \partial Y=\emptyset$, then ${{\, ]a,b[\, }}$ is either disjoint from or entirely contained in $Y$. 3. \[i:defconn\] If $Y$ is definably connected, then it is an interval. 4. \[i:dmin\] $Y$ either has a non-empty interior or is discrete. 5. \[i:disc\] If $Y$ is discrete, then it has a minimum and a maximum, and it is closed and bounded. 6. \[i:secbound\] The topological boundary $\partial Y$ is discrete, closed, and bounded. To prove , let $l\in M$ be the infimum of $Y$. By and the previous remark, $l^-$ either belongs to $Y$ or to $M\setminus Y$. The former case is excluded since $l$ is the infimum of $Y$. In particular, $l$ is not an interior point of $Y$. If $l^+$ does not belong to $Y$, then $l$ is an isolated point of $Y$, and hence belongs to the (topological) boundary. In the remaining case, $l$ lies in the closure of $Y$, since some open interval ${{\, ]l,x[\, }}$ lies inside $Y$. To prove , suppose there exists $x\in {{\, ]a,b[\, }}\cap Y$. By , either $x^-$ belongs to $Y$ or to $M\setminus Y$. In the latter case, there exists $z<x$ such that ${{\, ]z,x[\, }}$ is disjoint from $Y$. However, $x$ is then not an interior point of $Y$, whence must belong to its topological boundary, contradiction. So $x^-$ belongs to $Y$, which means that the set of all $z\in{{\, ]a,x[\, }}$ such that ${{\, ]z,x[\, }}\sub Y$ is non-empty. The infimum of this set must be a topological boundary point of $Y$ by , and hence must be equal to $a$, showing that ${{\, ]a,x]\, }}\sub Y$. Arguing the same with $x^+$, then shows that also ${{\, [x,b[\, }}\sub Y$, as we needed to prove. To prove , let $Y\sub M$ be definably connected. Let $l$ and $h$ be its respective infimum and supremum (including the case that these are infinite). The case $l=h$ is trivial, so assume $l<h$. If there were some $x\in{{\, ]l,h[\, }}$ not belonging to $Y$, then $Y$ would be the union of the two definable, non-empty, disjoint open subsets $Y_{<x}$ and $Y_{>x}$, contradiction, Hence, $Y$ is an interval with endpoints $l$ and $h$. To prove , assume $Y$ is not discrete. Hence there exists $a\in Y$ which is not isolated, that is to say, such that any open interval containing $a$ has some other point in common with $Y$. If both $a^-$ and $a^+$ belong to $M\setminus Y$, then there are $x<a<y$ such that ${{\, ]x,a[\, }},{{\, ]a,y[\, }}$ are disjoint from $Y$, contradicting that $a$ is not isolated. Hence, say, $a^-$ belongs to $Y$ and $Y$ has non-empty interior. Assume next that $Y$ is discrete and let $l$ be its infimum (including possibly the case $l=-\infty$). If $l^+$ belongs to $Y$, then ${{\, ]l,z[\, }}\sub Y$ for some $z>l$, contradicting discreteness. So $l^+$ does not belong to $Y$, which forces $l\in Y$. In particular, $l$ is finite, proving the first part of , and in particular, that $Y$ is bounded. To show that $Y$ is closed, suppose it is not. Let $x\notin Y$ be a point in its closure. Since $Y\cup\{ x\}$ is definable but not discrete, it must have interior by , so that some open interval $I$ is contained in $Y\cup\{ x\}$. But then $I\cap Y=I\setminus\{x\}$ is not discrete, contradiction. To see , it suffices by to show that $\partial Y$ is discrete. Let $b\in \partial Y$. We have to show that $b$ is an isolated point of $\partial Y$, and this will clearly hold if $b$ is an isolated point of either $Y$ or $M\setminus Y$. In the remaining case, exactly one from among $b^-$ and $b^+$ belongs to $Y$, say, $b^-$. Hence, there exist $x<b<y$ so that ${{\, ]x,b[\, }}\sub Y$ and ${{\, ]b,y[\, }}\cap Y=\emptyset$. Since any point in ${{\, ]x,b[\, }}$ is interior to $Y$ and any point in ${{\, ]b,y[\, }}$ exterior to $Y$, we get $\partial Y\cap{{\, ]x,y[\, }}=\{b\}$, as we needed to prove. \[R:discrete\] Although any discrete, definable subset $Y$ of an [o-minimalistic]{} structure has a maximum, non-definable subsets of $Y$ need not have a maximum. For instance, the set $D\cap {\mathbb R}=\mathbb N$ from Example \[E:ulomin\] has no maximum, whence cannot be definable. Nonetheless, we can define in general a successor function $\sigma_Y$ on $Y$ by letting $\sigma_Y(b)$ be the minimum of (the definable subset) $Y_{>b}$, for any non-maximal $b$ in $Y$. \[C:ddc\] The theory ${{\text{DCTC}}}$ is equivalent to type completeness in conjunction with discrete definable completeness, that is to say, the weaker version of which only requires definable discrete sets to have an infimum. Let $Y\sub M$ be a definable set in a model $\mathcal M$ of the weaker system from the assertion. Inspecting the argument in the proof of , type completeness already implies that $\partial Y$ is discrete. Hence $\partial Y$ has an infimum $b$, and it is now not hard to show that $b$ is also the infimum of $Y$. Recall the definition of *d-minimality* from [@Milldmin]: any definable subset $Y\sub M$ is the union of an open subset and finitely many discrete subsets (where the later finiteness is uniform in the parameters). Hence and Corollary \[C:ulomin\] immediate yield: \[C:dmin\] Any [o-minimalistic]{} structure is d-minimal. The reals with a unary predicate defining the integer powers of $2$ is d-minimal, but cannot be [o-minimalistic]{}, since the latter set is discrete but not closed. The following result fails in general in d-minimal structures, but holds immediately in [o-minimalistic]{} structures since the union of finitely many closed discrete subsets is again discrete. \[C:uniondisc\] In an [o-minimalistic]{} structure $\mathcal M$, any finite union of definable discrete subsets of $M$ is discrete. Using , and Remark \[R:discrete\], we get immediately the following structure theorem for one-variable definable subsets: \[T:onevar\] In an [o-minimalistic]{} structure, every definable subset is the disjoint union of a closed, bounded, discrete set and (possibly infinitely many) disjoint open intervals. Conversely, if every definable subset $Y\sub M$ of an $L$-structure $\mathcal M$ is a disjoint union of open intervals and a single closed, bounded, discrete set, then $\mathcal M$ is a model of ${{\text{DCTC}}}$. We only need to show the second assertion. Let us show that $M$ is definably connected. Indeed, if $U_1$ and $U_2$ are disjoint definable open sets covering $M$, then this would yield a covering of $M$ by disjoint open intervals. However, considering what the endpoints would be, this can only be the trivial covering, showing that one of the $U_i$ must be empty. By [@MillIVT Corollary 1.5], definable connectedness implies definable completeness. To prove type completeness, let $Y\sub M$ be definable and $b\in \partial Y$ (boundary points are the only points in which it can fail). There is nothing to prove if $b$ is an isolated point of $Y$ or of $M\setminus Y$, so assume it is not. Decompose $Y=U\cup D$ into definable subsets with $U$ a disjoint union of open intervals and $D$ closed, bounded, and discrete. Let ${{\, ]p,q[\, }}$ and ${{\, ]u,v[\, }}$ be the open interval in $U$ immediately to the left and to the right of $b$ respectively. Since $b$ is not isolated but in the boundary, it must be equal to exactly one of $q$ or $u$, that is to say, either $q=b<u$ or $q<u=b$. Say the latter holds, so that $b^+$ belongs to $Y$. Since $D\cup\{b\}$ is discrete, we can find an open interval $I$ containing $b$ such that $I\cap (D\cup\{b\})=\{b\}$. Shrinking $I$ if necessary, we can make it disjoint from ${{\, ]p,q[\, }}$, and hence $I_{<b}\cap Y=\emptyset$, showing that $b^-$ belongs to $M\setminus Y$. We need to verify this also at $b=\pm\infty$, where the same argument works in view of the boundedness of $D$. We conclude this section with an entirely topological characterization of o-minimality and o-minimalism, that is to say, without reference to the order relation. \[T:top\] Let $\mathcal M$ be a (densely ordered) $L$-structure, and let $Y\sub M$ range in the following over all non-empty definable subsets. 1. \[i:omintop\] $\mathcal M$ is o-minimal [if and only if]{} all $\partial Y$ are non-empty and finite. 2. \[i:ominismtop\] $\mathcal M$ is [o-minimalistic]{} [if and only if]{} all $\partial Y$ are non-empty, discrete, and contained in some definable subset which is definably connected but whose complement is not. The direct implication in is clear since $\partial Y$ is discrete and bounded by , whence contained in some closed interval $U:={{\, [a,b]\, }}$, which is clearly definably connected whereas its complement is not. For the converse, to prove , we may assume $Y$ is bounded from below, and then replace $Y$ with its upward closure, that is to say, all $x$ such that $x\geq y$ for some $y\in Y$. By assumption, $\partial Y$ contains at least one element, which necessarily must be the infimum of $Y$. To prove at a point $a\in M$, we may replace $Y$ by $Y_{\leq a}$, and assume $Y_{>a}$ is empty. Let $b$ be the supremum of $Y$. If $b<a$, then $a^-$ must belong to $M\setminus Y$ and we are done. So assume $a=b$, whence $a\in\partial Y$. By discreteness, there exists an open interval $I$ around $a$ such that $I\cap \partial Y=\{a\}$. Let $u\in I_{<a}$ and assume $u\notin Y$. Hence the infimum $v$ of $Y_{>u}$ must belong to $\partial Y$, and hence by the choice of $I$ be equal to $a$. On the other hand, since $a$ is the supremum of $Y$, there must be $u<x<a$ such that $x\in Y$, contradicting the previous observation. Hence ${{\, ]u,a[\, }}\sub Y$, as we needed to show. We also have to verify at $\pm\infty$, and this will follow once we show that $\partial Y$ is bounded. Choose $U$ containing $\partial Y$ such that $U$ is definably connected and $M\setminus U$ is not definably connected. It suffices to show that $U$ is bounded. However, under definable completeness , any definably connected subset is an interval (see the proof of above), and so if the complement of $U$ is not an interval, then $U$ must be a bounded interval. The direct implication in is also immediate, and for the converse, we know already from what we just proved that $\mathcal M$ is [o-minimalistic]{}. By Theorem \[T:onevar\], we can write $Y$ as a disjoint union of open intervals and a discrete set $D$. If two intervals in this decomposition would have a common endpoint which also belongs to $D$, then we replace these three objects by their union, which is then again an open interval. After doing this for all possible endpoints, $\partial Y$ now contains all the endpoints of the remaining intervals as well as points in $D$, and therefore both are finite collections, showing that $Y$ is a finite union of open intervals and points. \[R:Pod\] We could apply the criterion given by in other situations where there is an underlying definable topology (or, at least, with a basis of definable opens). For instance, a field viewed with its Zariski topology satisfies the criterion of [if and only if]{} it is minimal. Since any field is compact with respect to its Zariski topology, discrete sets are finite, and hence a field satisfying the condition in is likewise minimal. One might be tempted to deduce from this Podewski’s conjecture: an elementary extension of a minimal field is again minimal. But this is erroneous, because, unlike the ordered case, neither being discrete nor being definably connected is a first-order condition. This stems from the fact that the collection of basic Zariski open subsets does not belong to a single definable family, unlike the ordered case, so that one cannot quantify over basic open sets. \[C:opencore\] An [o-minimalistic]{} structure with o-minimal open core is itself o-minimal. Recall that the *open core* (see, for instance, [@DMS; @MiSpCore]) of an ordered structure $\mathcal M$ is the reduct in which the definable sets are the definable open subsets of $\mathcal M$. Suppose that $\mathcal M$ is [o-minimalistic]{} and its open core is o-minimal. Let $Y\sub M$ be a definable subset. By , its boundary $\partial Y$ is non-empty and discrete. Since $\partial Y$ is closed, it is definable in the open core, and therefore finite by o-minimality. The result now follows from . This also follows from the description of the one-variable definable subsets in a structure with o-minimal core as the disjoint union of a finite set and of finitely many dense and co-dense subsets of intervals. Definable maps ============== Next we study definable maps, where we call a map $f\colon Y\sub M^n\to M^k$ *definable* if its graph $\Gamma(f)\sub M^{n+k}$ is a definable subset. Note that since its domain $Y\sub M^n$ is the projection of its graph, it too is definable. Similarly, the set $\Gamma^*(f)$ consisting of all $(f(x),x)\in M^{k+n}$ is definable and is called the *reverse graph* of $f$. If $k=n=1$, we speak of *one-variable* maps. \[L:imdisc\] Let $f\colon Y\to M$ be a definable map in an [o-minimalistic]{} structure $\mathcal M$. 1. \[i:imdisc\] If $Y$ is discrete, then so is its image $f(Y)$. 2. \[i:fibdisc\] If $f(Y)$ and each fiber of $f$ is discrete, then so is $Y$. Suppose does not hold, so that $Y\sub M$ is discrete but not $f(Y)$. Let $H$ be the (non-empty, definable, discrete) subset of all $x\in Y$ such that $f(Y_{\geq x})$ is non-discrete, and let $h$ be its maximum. Since $h$ cannot be the maximal element of $Y\sub M^n$ lest $f(Y_{\geq h})$ be a singleton, we can find its successor $\sigma(h)\in Y$ by . But $f(Y_{\geq h})=\{f(h)\}\sqcup f(Y_{\geq \sigma(h)})$, so that neither $f(Y_{\geq \sigma(h)})$ can be discrete by Corollary \[C:uniondisc\], contradicting maximality. Assume next that is false, so that $Y$ is non-discrete, but $Z:=f(Y)$ and all ${{f^{-1}(u)}}$ are discrete. This time, let $H$ be the subset of all $x\in Z$ such that ${{f^{-1}(Z_{\geq x})}}$ is non-discrete, and let $h$ be its maximum. Again $h$ must be non-maximal in $Z$, and so admits an immediate successor $\sigma(h)\in Z$. Since both subsets on the right hand side of $${{f^{-1}(Z_{\geq h})}}={{f^{-1}(h)}}\sqcup {{f^{-1}(Z_{\geq \sigma(h)})}}$$ are discrete, the first by assumption and the second by maximality, so must their union be by Corollary \[C:uniondisc\], contradiction. We will say that a one-variable function $f\colon Y\to M$ has a *jump discontinuity* at a point $c$ if the left and right limit of $f$ at $c$ exist, but are different. \[T:disctu\] The set of discontinuities of a one-variable definable map $f\colon Y\to M$ in an [o-minimalistic]{} structure $\mathcal M$ is discrete, closed, and bounded, and consists entirely of jump discontinuities. Moreover, there is a definable discrete, closed, bounded subset $D\sub Y$ so that in between any two consecutive points of $D\cup\{\pm\infty\}$, the map is *monotone*, that is to say, either strictly increasing, strictly decreasing, or constant. We start with proving that all discontinuities are jump discontinuities, or equivalently, that $f$ has a left limit in each point $a\in Y$. For each $y<a$, let $w(y)$ be the supremum of $f({{\, [y,a[\, }})$ and let $b$ be the infimum of $w(Y_{< a})$. I claim that $b$ is the left limit of $f$ at $a$. To this end, choose $p<b<q$, and we need to show that there is some $x<a$ with $p<f(x)<q$. If $b^+$ does not belong $w(Y_{< a})$, and therefore belongs to its complement, then $b$ is an isolated point of $w(Y_{< a})$, implying that $f$ takes constant value $b$ on some interval ${{\, ]y,a[\, }}$, so that $b$ is indeed the left limit at $a$. In the remaining case, we can find $u>b$ such that ${{\, ]b,u]\, }}\sub w(Y_{< a})$. We may choose $u<q$. In particular, $u=w(y)$ for some $y<a$. Since $b$ is strictly less than the supremum $u=w(y)$, we can find $x\in {{\, [y,a[\, }}$ such that $b< f(x)\leq u$, whence $p<f(x)<q$, as required. Let $C\sub Y$ be the definable subset given as the union of the interior of all fibers, that is to say, $x\in C$ [if and only if]{}  $x$ is an interior point of ${{f^{-1}(f(x))}}$. Being an open set, $C$ is a disjoint union of open intervals, and $f$ is constant, whence continuous on each of these open intervals. Every fiber of the restriction of $f$ to $Y\setminus C$ must have empty interior, whence is discrete by . So upon replacing $f$ by this restriction, we may reduce to the case that $f$ has discrete fibers. There is nothing to show if $Y$ is discrete, and so without loss of generality, we may assume $Y$ is an open interval. For fixed $a\in Y$, let $L_a$ (respectively, $H_a$) be the set of all $x\in Y$ such that $f(x)<f(a)$ (respectively, $f(a)<f(x)$). Since $Y$ is the disjoint union of $L_a$, $H_a$, and ${{f^{-1}(f(a))}}$ with the latter being discrete, $a^-$ must belong to one of the first two sets by Corollary \[C:uniondisc\], and depending on which is the case, we will denote this symbolically by writing respectively $f(a^-)< f(a)$ or $f(a^-)> f(a)$ (with a similar convention for $a^+$). Let $L_-$ (respectively, $H_-$, $L_+$, and $H_+$) be the set of all $a\in Y$ such that $f(a^-)<f(a)$ (respectively, $f(a^-)>f(a)$, $f(a^+)<f(a)$, and $f(a^+)>f(a)$), so that $Y$ is the disjoint union of these four definable subsets. Let $D$ be the union of the topological boundaries of these four sets, a discrete set by Corollary \[C:uniondisc\]. If $b<c$ are consecutive elements in $D$, then ${{\, ]b,c[\, }}$ must belong to exactly one of these four sets by , say, to $L_-$. It is now easy to see that in that case $f$ is strictly increasing on ${{\, ]b,c[\, }}$. This then settles the last assertion. Let $S$ be the (definable) subset of all discontinuities of $f$. To prove that $S$ is discrete, we need to show by that it has empty interior, and this will follow if we can show that any open interval $I\sub Y$ contains a point at which $f$ is continuous. By what we just proved, by shrinking $I$ if necessary, we may assume $f$ is monotone on $I$, say, strictly increasing. Note that $f$ is then in particular injective. By Lemma \[L:imdisc\], the image $f(I)$ contains an open interval $J$. Since $f$ is strictly increasing, ${{f^{-1}(J)}}$ is also an open interval, and $f$ restricts to a bijection between ${{f^{-1}(J)}}$ and $J$. We leave it to the reader to verify that any strictly increasing bijection between intervals is continuous. \[R:lim\] Given a definable map $f$ and a point $a$, we denote its left and right limit simply by $f^-(a)$ or $f^+(a)$ respectively, even if these values are infinite (to be distinguished from the symbol $f(a^-)$ which occurred above in formulae of the form $f(a^-)<f(a)$). Note that we even have this property at $\pm\infty$, so that we can define $f^+(-\infty)$ and $f^-(\infty)$, which we then simply abbreviate as $f(-\infty)$ and $f(\infty)$ respectively. \[C:ctuconn\] In an [o-minimalistic]{} structure $\mathcal M$, a definable map $f\colon I\to M$ with domain an open interval $I$ is continuous [if and only if]{} its graph is definably connected. Let $C$ be the graph of $f$. If $f$ is not continuous, then it has a jump discontinuity at some point $a\in I$ by Theorem \[T:disctu\]. Without loss of generality, we may assume $f^-(a)<f^+(a)$. Let $c$ be some element between these two limits and different from $f(a)$. By definition of one-sided limit, there exist $p<a<q$ such that $f(x)<c$ whenever $p<x<a$, and $f(x)>c$ whenever $a<x<q$. Consider the two open subsets $$\begin{aligned} U_-&:=(I_{<a}\times M)\cup (I_{<q}\times M_{<c}) \\ U_+&:=(I_{>a}\times M)\cup (I_{>p}\times M_{>c}).\end{aligned}$$ It is not hard to check that $C$ is contained in their union but disjoint from their intersection, showing that it is not definably connected. Conversely, assume $f$ is continuous but $C$ is not definably connected, so that there exist definable open subsets $U$ and $U'$ whose union contains $C$ but whose intersection is disjoint from $C$. Since the projections $\pi(C\cap U)$ and $\pi(C\cap U')$ onto the first coordinate are definable subsets partitioning $I$, they must have a common boundary point $b\in I$ by Proposition \[P:Dedlim\]. Since $(b,f(b))$ belongs to either $U$ or $U'$, say, to $U$, there exists a box $J\times J'\sub U$ containing $(b,f(b))$. By continuity, we may assume $f(J)\sub J'$. This implies that $(x,f(x))\in U$, for all $x\in J$, and hence that $J\sub\pi(C\cap U)$, contradicting that $b$ is a boundary point of the latter. \[R:genctu\] Without proof, we claim that the above results extend to arbitrary dimensions: given a definable map $f\colon X\sub M^n\to M^k$, the set of discontinuities of $f$ is nowhere dense in $X$. For instance (with terminology to be defined below), if $X\sub M^2$ has dimension two, for each $a,b\in M$, let $D_a$ and $E_b$ be some discrete sets, as given by Theorem \[T:disctu\], such that between any two consecutive points the respective maps $y\mapsto f(a,y)$ and $x\mapsto f(x,b)$ are continuous and monotone. Let $D$ and $E$ be the respective union of all $\{a\}\times D_a$ and all $E_b\times\{b\}$. By Corollary \[C:fibdim\] and Proposition \[P:nondisc2\] below, both $D$ and $E$ are one-dimensional, closed subsets, and hence $X':=X\setminus(D\cup E)$ is open and dense in $X$. It is now not hard to show that $f$ is continuous on $X'$ (see [@vdDomin Chapt. 3, Lemma 2.16]). Discrete sets ============= We start our analysis of multi-variable definable subsets, with a special emphasis on definable subsets of the *plane* $M^2$ (and address the general case through some sporadic remarks). Since projections play an important role, we introduce some notation. Fix $n$ and let $I\sub \{1,\dots,n\}$ of size ${\left|I\right|}:=e$. We let $\pi_I\colon M^n\to M^e$ be the projection ${(a_1,\dots,a_{n})}\mapsto (a_{i_1},\dots,a_{i_e})$, where $I=\{i_1<i_2<\dots<i_e\}$. When $I$ is a singleton $\{i\}$, we just write $\pi_i$ for the projection onto the $i$-th coordinate. Given a tuple ${\mathbf {a}}={(a_1,\dots,a_{e})}\in M_e$, the ($I$-)*fiber* of $X$ above ${\mathbf {a}}$ is the set $${{X}_{{I}}[{\mathbf {a}}]}:=\pi_{I^c}\left({{\pi_I^{-1}({\mathbf {a}})}}\cap X\right),$$ where $I^c$ is the complement of $I$. In other words, ${{X}_{{I}}[{\mathbf {a}}]}$ is the set of all ${\mathbf {b}}\in M^{n-e}$ such that ${\mathbf {c}}\in X$, where ${\mathbf {c}}$ is obtained from ${\mathbf {b}}$ by inserting $a_{i_k}$ at the $k$-th spot. In case $I$ is of the form $\{1,\dots,e\}$, for some $e$, we omit $I$ from the notation, since the length of the tuple ${\mathbf {a}}$ then determines the projection, and we refer to it as a *principal projection*, with a similarly nomenclature for fibers. Thus, for example, the *principal fiber* ${{X}_{{}}[a]}$ is the set ${{X}_{{1}}[a]}$ of all $n-1$-tuples ${\mathbf {b}}$ such that $(a,{\mathbf {b}})\in X$. Recall that by any definable discrete subset of $M$ is closed and bounded. The same is true in higher dimensions, for which we first prove: \[T:projdisc\] A definable subset $X\sub M^n$ in an [o-minimalistic]{} structure $\mathcal M$ is $X$ is discrete [if and only if]{} all projections $\pi_1(X), \dots, \pi_n(X)$ are discrete. Suppose all projections are discrete and let ${(a_1,\dots,a_{n})}\in X$. Hence we can find open intervals $I_k$, for $k={1,\dots,n}$, such that $I_k\cap \pi_k(X)=\{a_k\}$. The open box $I_1\times \cdots\times I_n$ then intersects $X$ only in the point ${(a_1,\dots,a_{n})}$, proving that $X$ is discrete. To prove the converse, we will induct on $n$, proving simultaneously the following three properties for $X\sub M^n$ discrete: 1. \[i:projdisc\] $\pi_1(X), \dots, \pi_n(X)$ are discrete; 2. \[i:totord\] $X$ with the induced lexicographical ordering has a minimal element; 3. \[i:succdisc\] for this ordering, there exists a definable map $\sigma_X$ of $X$, sending every non-maximal element in $X$ to its immediate successor. All three properties have been established by Proposition \[P:Dedlim\] when $n=1$, so assume they hold for $n-1$. Assume towards a contradiction that $\pi_1(X)$ is not discrete. For each $a\in\pi_1(X)$, the fiber ${{X}_{{}}[a]}$ (that is to say, the set of all ${\mathbf {b}}\in M^{n-1}$ such that $(a,{\mathbf {b}})\in X$), is discrete since $a\times {{X}_{{}}[a]}\sub X$. By the induction hypothesis for , in its lexicographical order, ${{X}_{{}}[a]}$ has a minimum, denoted $f(a)$, yielding a definable map $f\colon \pi_1(X)\to M^{n-1}$ whose graph lies in $X$. By Theorem \[T:disctu\], each composition $\pi_i{\circ}f\colon \pi_1(X)\to M$, for $i={1,\dots,n-1}$, is continuous outside a discrete set. The union of these discrete sets is again discrete by Corollary \[C:uniondisc\], and hence, since $\pi_1(X)$ is assumed non-discrete, there is a common point $a$ at which all $f_i$ are continuous, whence also $f$. By the discreteness of $X$, we can find an open interval $I$ and an open box $U\sub M^{n-1}$ containing respectively $a$ and $f(a)$ such that $(I\times U)\cap X=\{(a,f(a))\}$. By continuity, we can find an open interval $J\sub I$ containing $a$ such that $f(J)\sub U$. However, this means that for any $u\in J$ different from $a$, we have $f(u)\in U$, whence $(u,f(u))\in (I\times U)\cap X=\{(a,f(a))\}$, forcing $u=a$, contradiction. To prove , we now have established that $\pi_1(X)$ is discrete, whence has a minimum $l$. The minimum of $X$ in the lexicographical ordering is then easily seen to be $(l,{\operatorname}{min}({{X}_{{}}[l]}))$. To define $\sigma_X$, let ${\mathbf {a}}=(a,{\mathbf {b}})\in X$. For ${\mathbf {a}}\neq \max(X)$, either ${\mathbf {b}}$ is not the maximum of ${{X}_{{}}[a]}$ and we set $\sigma({\mathbf {a}}):=(a,{\mathbf {b}}')$ where ${\mathbf {b}}':=\sigma_{{{X}_{{}}[a]}}({\mathbf {b}})$; or otherwise, $a$ is not the maximum of $\pi_1(X)$ and we set $\sigma({\mathbf {a}}):=(a',{\operatorname}{min}({{X}_{{}}[a']}))$, where $a':=\sigma_{\pi_1(X)}(a)$. Note that the existence of $a'$ and ${\mathbf {b}}'$ follow from the induction hypothesis on . We leave it to the reader to verify that $\sigma_X$ has the required properties. \[C:closeddiscrete\] Any definable, discrete subset in an [o-minimalistic]{} structure is closed and bounded. Let $X\sub M^n$ be a definable, discrete subset in an [o-minimalistic]{} structure $\mathcal M$. By Theorem \[T:projdisc\], all $\pi_i(X)$ are discrete, whence bounded and closed by . It is now easy to deduce from this that so is then $X$. \[C:imdisc\] In an [o-minimalistic]{} structure, the image under a definable map of a definable discrete set is again discrete. If $\mathcal M$ is [o-minimalistic]{}, and $f\colon X\sub M^n\to M$ is definable with $X$ discrete, then the graph of $f$ is also discrete (as a subset of $M^{n+1}$). Since $f(X)$ is the projection of this graph, it is discrete by Theorem \[T:projdisc\]. \[C:discfib\] A definable subset $X\sub M^n$ in an [o-minimalistic]{} structure $\mathcal M$ is discrete [if and only if]{} for some (equivalently, for all) $I\sub\{1,\dots,n\}$, the projection $\pi_I(X)$ as well as each fiber ${{X}_{{I}}[{\mathbf {a}}]}$ is discrete. The converse implication is easy, and for the direct, suppose $X$ is discrete. We may assume, after renumbering, that $I=\{1,\dots,e\}$. Since each fiber ${{X}_{{I}}[{\mathbf {a}}]}$ is homeomorphic to the subset ${\mathbf {a}}\times {{X}_{{I}}[{\mathbf {a}}]}\sub X$, for ${\mathbf {a}}\in M^e$, and since the latter is discrete, so is the former. On the other hand, each $\pi_i(X)$, for $i\leq e$, is discrete by Theorem \[T:projdisc\], and since these are just the projections of $\pi_I(X)$, the latter is also discrete, by the same theorem. Suppose $\mathcal M$ is an [o-minimalistic]{} expansion of an (Abelian, divisible) ordered group (we will discuss this situation in more detail in §\[s:omingroup\]). We define the absolute value ${\left|a\right|}$ as the maximum of $a$ and $-a$, for any $a\in M$. We call a map $f\colon X\to X$, for $X\sub M$, *contractive*, if $$\label{eq:contract} {\left|f(x)-f(y)\right|}<{\left|x-y\right|},$$ for all $x\neq y\in X$. We say that $f$ is *weakly contractive*, if instead we have only a weak inequality in . Recall that a *fixed point* of $f$ is a point $x\in X$ such that $f(x)=x$. If $f$ is contractive, it can have at most one fixed point. \[T:fixpt\] Let $f\colon D\to D$ be a definable map on a discrete, definable subset $D\sub M$ in an [o-minimalistic]{} expansion $\mathcal M$ of an ordered group. If $f$ is contractive, it has a unique fixed point. If $f$ is weakly contractive, then $f^2$ has a fixed point. We treat both cases simultaneously. Assume $f$ does not have a fixed point. In particular, $f(l)>l$, where $l$ is the minimum of $D$. Hence the set of $x\in D$ such that $x<f(x)$ is non-empty, whence has a maximal element $u$. Clearly, $u<h$, where $h$ is the maximum of $D$, and hence $u$ has an immediate successor $v:=\sigma_D(u)$ by . By maximality, we must have $f(v)<v$. Hence $v\leq f(u)$ and $u\leq f(v)$, and therefore $v-u\leq{\left|f(u)-f(v)\right|}$, leading to a contradiction in the contractive case with , showing that $f$ must have a fixed point, necessarily unique. In the weak contractive case, we must have an equality in the latter inequality, whence also in the former two, that is to say, $f(u)=v$ and $f(v)=u$. Hence, $u$ and $v$ are fixed points of $f^2$. Sets with non-empty interior ============================ Shortly, we will introduce the notion of dimension, and whereas the discrete sets are those with minimal dimension (=zero), the sets with non-empty interior will be those of maximal dimension. Note that the non-empty definable subsets of $M$ are exactly of one of these two types by . \[P:nondisc2\] In an [o-minimalistic]{} structure $\mathcal M$, a definable subset $X\sub M^n$ has non-empty interior [if and only if]{} the set of points $a\in M$ such that the fiber ${{X}_{{}}[a]}$ has non-empty interior is non-discrete. If $X$ has non-empty interior, it contains an open box, and the assertion is clear. For the converse, note that, since we can pick definably the first open interval inside a definable non-discrete subset of $M$ by the properties proven in Proposition \[P:Dedlim\], we may reduce to the case that $\pi(X)$ is an open interval and each fiber ${{X}_{{}}[a]}$ for $a\in \pi(X)$ is an open box, where $\pi\colon M^n\to M$ is the projection onto the first coordinate. The proof for $n>2$ is practically identical to that for $n=2$, and so, for simplicity, we assume $n=2$. Let $l(a)$ and $h(a)$ be respectively the infimum and supremum of $X_a$, so that $l,h\colon \pi(X)\to M_\infty$ are definable maps. The subset of $\pi(X)$ where either function takes an infinite value is definable, whence it or its complement contains an open interval, so that we can either assume that $l$ is either finite everywhere or equal to $-\infty$ everywhere, and a similar dichotomy for $h$. The infinite cases can be treated by a similar argument, so we will only deal with the case that they are both finite (this is a practice we will follow often in our proofs). By Theorem \[T:disctu\], there is a point $a\in\pi(X)$ at which both $l$ and $h$ are continuous. Fix some $c<l(a)<p<q<h(a)<d$, so that by continuity, we can find $u<a<v$ so that $l({{\, ]u,v[\, }})\sub {{\, ]c,p[\, }}$ and $h({{\, ]u,v[\, }})\sub {{\, ]q,d[\, }}$. I claim that ${{\, ]u,v[\, }}\times {{\, ]p,q[\, }}$ is entirely contained in $X$. Indeed, if $u<x<v$ and $p<y<q$, then from $c<l(x)<p<y<q<h(x)<d$, we get $y\in {{X}_{{}}[x]}$, that is to say, $(x,y)\in X$. By a simple inductive argument, we get the following analogue of Corollary \[C:discfib\]: \[C:openfib\] A definable subset $X\sub M^n$ in an [o-minimalistic]{} structure $\mathcal M$ has non-empty interior [if and only if]{} for some (equivalently, for all) $I\sub\{1,\dots,n\}$, the set of points ${\mathbf {a}}$ for which ${{X}_{{I}}[{\mathbf {a}}]}$ has non-empty interior, has non-empty interior. \[C:non-disc2\] In an [o-minimalistic]{} structure $\mathcal M$, a finite union of definable subsets of $M^n$ has non-empty interior [if and only if]{} one of the subsets has non-empty interior. One direction is immediate, and to prove the other we may by induction reduce to the case of two definable subsets $X_1,X_2\sub M^n$ whose union $X:=X_1\cup X_2$ has non-empty interior. We induct on $n$, where the case $n=1$ follows from Corollary \[C:uniondisc\] and . Let $W\sub M$ be the subset of all points $a\in M$ for which the fiber ${{X}_{{}}[a]}\sub M^{n-1}$ has non-empty interior. By Proposition \[P:nondisc2\], the interior of $W$ is non-empty. Since ${{X}_{{}}[a]}={{(X_1)}_{{}}[a]}\cup {{(X_2)}_{{}}[a]}$, our induction hypothesis implies that for $a\in W$, at least one of ${{(X_i)}_{{}}[a]}$ has non-empty interior, in which case we put $a$ in $W_i$. In particular, $W=W_1\cup W_2$ so that at least one of the $W_i$ has non-empty interior, say, $W_1$. By Proposition \[P:nondisc2\], this then implies that $X_1$ has non-empty interior. Planar cells and arcs ===================== For the remainder of our analysis of multi-variable definable sets, apart from separate remarks, we restrict to subsets of the plane, that is to say, of $M^2$. Given an ordered structure $\mathcal M$ (soon to be assumed also [o-minimalistic]{}), let us define a $\emph{$2$-cell}$ in $M^2$ as a definable subset $C$ of the following form: suppose $I$ is an open interval, called the *domain* of the cell, and $f,g\colon I\to M$ are definable, continuous maps such that $f<g$ (meaning that $f(x)<g(x)$ for all $x\in I$). Let $C$ be the subset of all $(x,y)\in M^2$ with $x\in I$ and $f(x)\diamond_1 y\diamond_2g(x)$, where $\diamond_i$ is either no condition or a strict inequality (when we only have at most one inequality, we get an example of an *unbounded* cell; the remaining ones are call *bounded*, and in arguments we often only treat the latter case and leave the former with almost identical arguments to the reader). Any $2$-cell is open. By a *$1$-cell* $C\sub M^2$, we mean either the graph of a continuous definable map $f$ with domain an open interval $I$, or a Cartesian product $x\times I$. We call the former *horizontal* and the latter *vertical*. Finally, by a *$0$-cell*, we mean a point. We may combine all these definitions into a single definition: a cell $C$ is determined by elements $a<b$ and definable, continuous maps $f<g\colon M\to M$, as the set of all pairs $(x,y)$ such that $a\diamond_{1} x\diamond_{2} b$ and $f(x)\diamond_3y\diamond_4g(x)$, where each $\diamond_i$ is either no condition, equality or strict inequality. Moreover, if $C$ is non-empty, then it is a $d$-cell, where $d$ is equal to two minus the number of equality signs among the $\diamond_i$. We sometimes use some suggestive notation like $C(I;f<g)$ to denote, for instance, the cell given by $x\in I$ and $f(x)<y<g(x)$. If $C$ is a cell with domain $I$ and $J\sub I$ is an open interval, then we call $C\cap (J\times M)$ the *restriction* of $C$ to $J$. Any restriction of a cell is again a cell, and so is any principal projection. \[R:gencell\] For higher arity, we likewise define cells inductively: we say that $C\sub M^n$ is a *$d$-cell* if either $C$ is the graph of a definable, continuous function with domain some $d$-cell in $M^{n-1}$, or otherwise, is the region strictly between two such graphs with common domain some $(d-1)$-cell in $M^{n-1}$. As we shall see below in Remark \[R:dimdef\], the $d$ in $d$-cell refers to the dimension of the cell. Arcs ---- Assume again that we work in an [o-minimalistic]{} structure $\mathcal M$. Given a definable subset $X\sub M^n$, a point $P=(a,b)\in M^2$, and a definable map $h\colon Y\to M$ such that $a\in Y$ and $h^-(a)=b$, we will say that *$P^-_h$ belongs to $X$*, if there exists an open interval ${{\, ]u,a[\, }}\sub I$ so that the graph of the restriction of $h$ lies inside $X$. By Theorem \[T:disctu\], we may shrink ${{\, ]u,a[\, }}$ so that $h$ is continuous on that interval, and so we could as well view this as a property of the horizontal $1$-cell $C$ defined by $h$. Note that $P$ lies in the closure of $C$. Moreover, we only need $a$ to lie in the closure of $Y$ to make this work. So, given a $1$-cell $C$ such that $P$ lies in its closure, we say that *$P_C^-$ belongs to $X$* if $P_h^-$ does, where $h$ is the definable, continuous map determining $C$, in case $C$ is a horizontal cell, or if $b^-$ belongs to ${{X}_{{}}[a]}$ in case $C$ is a vertical cell. Of course, we can make a similar definition for $P_h^+$ or $P_C^+$. The following result essentially shows that viewed as a type, $P_C^-$ is complete: \[L:limcell\] Given a definable subset $X\sub M^2$ in an [o-minimalistic]{} structure $\mathcal M$, a $1$-cell $C\sub M^2$, and a point $P$ in the closure of $C$, either $P_C^-$ belongs to $X$ or it belongs to its complement. Let $P=(a,b)$ be in the closure of $C$. If $C$ is a vertical cell, then $P_C^-$ belongs to $X$ [if and only if]{} $b^-$ belongs to ${{X}_{{}}[a]}$, and so we are done in this case by . In the horizontal case, there exists a definable, continuous map $h\colon {{\, ]u,a[\, }}\to M$ whose graph is contained in $C$. By , either $a^-$ belongs to $\pi(X\cap C)$ or to its complement. In the former case, after increasing $u$ if necessary, we have ${{\, ]u,a[\, }}\sub \pi(X\cap C)$, whence $(x,h(x))\in X$ for every $x\in {{\, ]u,a[\, }}$. In the latter case, ${{\, ]u,a[\, }}$ is disjoint from $\pi(X\cap C)$, and hence $(x,h(x))\notin X$ for every $x\in {{\, ]u,a[\, }}$. Using this, it is not hard to show that the following is an equivalence relation (and in particular symmetric): given a point $P\in M^2$ and $1$-cells $V,W\sub M^2$ such that $P$ lies in each of their closures, we say that $V\equiv_{P^-}W$, if $P_V^-$ belongs to $W$. By a *left arc* at $P$, we mean an $(\equiv_{P^-})$-equivalence class of $1$-cells whose closure contains $P$; and a similar definition for $V\equiv_{P^+}W$ and *right arc*. It is now easy to see that $P_V^-$ belongs to some definable subset $X\sub M^2$ [if and only if]{} $P_W^-$ belongs to it, for any $W\equiv_{P^-}V$, so that we may make sense of the expression *$P_\alpha$ belongs to $X$*, for any left (or right) arc $\alpha$ at $P$. There are two unique equivalence classes containing a vertical cell, called respectively the *lower* and *upper vertical arc*; the remaining ones are called *horizontal*. Given two left horizontal arcs $\alpha$ and $\beta$ at $P$, we can find a common domain $I={{\, ]u,a[\, }}$ and definable continuous functions $f$ and $g$ on $I$, such that $\alpha$ and $\beta$ are the respective equivalence classes of the graphs of $f$ and $g$. Let $I_-$, $I_=$ and $I_+$ be the subsets of all $x\in I$ such that $f(x)$ is less than, equal to, or bigger than $g(x)$ respectively. If $\alpha\neq \beta$, then $a$ cannot be in the closure of $I_=$, so that upon shrinking even further, we may assume $I_=$ is empty. Hence $a^-$ belongs either to $I_-$ or $I_+$ and we express this by saying that $\alpha<_{P^-}\beta$ and $\alpha>_{P^-}\beta$ respectively. This yields a well-defined total order relation $<_{P^-}$ on the left horizontal arcs at a point $P$. To include the vertical arcs, we declare the lower one to be smaller than any horizontal left arc and the upper one to be bigger than any. \[P:infarc\] Let $X\sub M^2$ be a definable subset, and $P\in M^2$ a point in an [o-minimalistic]{} structure $\mathcal M$. The set of all left arcs $\alpha$ at $P$ such that $P_\alpha$ belongs to $X$ has an infimum $\beta$ (with respect to the order $<_{P^-}$). If $\beta$ is not vertical, then $P_\beta$ belongs to $\partial X$. Since a point is either interior, exterior or a boundary point, we may upon replacing $X$ by its complement, reduce to the case that $P=(a,b)$ is either interior or a boundary point. In the former case, the lower vertical arc is clearly minimal, so assume $P\in\partial X$. In what follows, $\alpha$ always denotes a left arc at $P$. Consider the set $L_\emptyset$ of all $x<a$ such that ${{X}_{{}}[x]}\cap J$ is empty for some open interval $J$ containing $b$. If $a^-$ belongs to $L_\emptyset$, then no $P_\alpha$ belongs to $X$ so that the upper vertical arc is the minimum. So we may assume that the ${{X}_{{}}[x]}\cap J$ are non-empty for $x$ close to $a$ from the left. If $b^-$ belongs to ${{X}_{{}}[a]}$, then the lower vertical arc is the infimum, so assume $b^-$ belongs to $M\setminus{{{X}_{{}}[a]}}$. Hence we may shrink $J$ so that $J\cap ({{X}_{{}}[a]})_{<b}$ is empty. For each $x<a$, let $f(x)$ be the infimum of ${{X}_{{}}[x]}\cap J$. On a sufficiently small open interval ${{\, ]u,a[\, }}$, the function $f$ is continuous, whence defines a $1$-cell $V$. Since $J\cap ({{X}_{{}}[a]})_{<b}=\emptyset$, the left limit $f^-(a)$ must be equal to $b$, showing that $(a,b)$ lies in the closure of $V$, and hence the equivalence class of $V$ at $P^-$ is a left arc $\beta$. It is now easy to show that $\beta$ is the required infimum, and that it is contained in the boundary $\partial X$. \[C:dimcl\] In an [o-minimalistic]{} structure $\mathcal M$, if $C\sub M^2$ is a definable subset without interior, then so is its closure, that is to say, $C$ is nowhere dense. Suppose $P=(a,b)$ is an interior point of the closure $\bar C$, so that there exists an open box $U\sub \bar C$ containing $P$. By Proposition \[P:nondisc2\], the fibers ${{C}_{{}}[x]}$ for $x$ close to $a$ must be discrete. By Proposition \[P:infarc\], the infimum $\alpha$ of all left arcs at $P$ belonging to $C\setminus {{C}_{{}}[x]}$ exists, and by discreteness of the surrounding fibers, it must be a minimum, whence also belong to $C$. Similarly, the infimum $\beta$ of all left arcs at $P$ belonging to $C$ and strictly bigger than $\alpha$ is also a minimum. Choose an open interval ${{\, ]u,a[\, }}$ such that $\alpha$ and $\beta$ are represented by the respective continuous, definable maps $f,g\colon {{\, ]u,a[\, }}\to M$. Enlarging $u$ if needed, we may assume $f<g$, so that the $2$-cell $S:=C({{\, ]u,a[\, }};f<g)$ is disjoint from $C$. Since $S$ is open and $P$ lies in its closure, $S\cap U$ is non-empty. Since $(S\cap U)\cap C=\emptyset$, no point of $S\cap U$ can lie in the closure of $C$, contradiction. The Hardy structure of an [o-minimalistic]{} structure ------------------------------------------------------ We now extend this to infinity in the obvious way: given two horizontal cells $V$ and $W$ with domain an interval unbounded to the right, we say that $V\equiv_\infty W$ if their restrictions to some interval ${{\, ]u,\infty[\, }}$ are equal. Let ${\mathbf H(\mathcal {M})}$ be the set of all equivalence classes of cells defined on an open interval unbounded to the right. Note that any definable map $f\colon Y\to M$ whose domain is unbounded to the right yields an equivalence class in ${\mathbf H(\mathcal {M})}$, denoted $[f]$, since $f$ is continuous by Theorem \[T:disctu\] on some open interval ${{\, ]u,\infty[\, }}\sub Y$. Given a definable subset $X\sub M^2$, we can say, as before, that *$\infty_\alpha$ belongs to $X$*, if $\infty^-$ belongs to the set of all $x\in Y$ such that $(x,f(x))\in X$, for some $f$ with arc $\alpha$. However, in this case we can do more and make ${\mathbf H(\mathcal {M})}$ into an $L$-structure: if $\underline c$ is a constant symbol, then we interpret it in ${\mathbf H(\mathcal {M})}$ as the class of the constant function with value $c:=\underline c^{\mathcal M}$; if $\underline F$ is an $n$-ary function symbol, and $\alpha_1,\dots,\alpha_n\in {\mathbf H(\mathcal {M})}$, then $\underline F(\alpha_1,\dots,\alpha_n)$ is the class given by the definable map $\underline F(g_1,\dots,g_n)$, where the $g_i$ are definable functions with domain $I:={{\, ]u,\infty[\, }}$ such that $[g_i]=\alpha_i$; if $\underline R$ is an $n$-ary predicate symbol, then $\underline R(\alpha_1,\dots,\alpha_n)$ holds in ${\mathbf H(\mathcal {M})}$ [if and only if]{}  $\infty^-$ belongs to the set of all $x\in I$ such that $\underline R(g_1(x),\dots,g_n(x))$ holds in $\mathcal M$. \[D:Hardy\] We call this $L$-structure on ${\mathbf H(\mathcal {M})}$ the *Hardy structure* of $\mathcal M$. In particular, by the same argument as above, $<$ interprets a total order on ${\mathbf H(\mathcal {M})}$, making it into a densely ordered structure without endpoints (note that the notion of vertical arc makes no sense in this context). By induction on the complexity of formulae, we easily can show: \[L:memberHardy\] Given an [o-minimalistic]{} structure $\mathcal M$, let $\varphi{(x_1,\dots,x_{n})}$ be a formula with parameters from $M$ and let $X\sub M^n$ be the set defined by it. For given arcs $\alpha_1,\dots,\alpha_n\in {\mathbf H(\mathcal {M})}$, we have ${\mathbf H(\mathcal {M})}\models \varphi{(\alpha_1,\dots,\alpha_{n})}$ [if and only if]{}  there is a $u\in M$ such that $(g_1(x),\dots,g_n(x))\in X$, for all $x>u$, where each $g_i$ is some continuous function defined on ${{\, ]u,\infty[\, }}$ representing the arc $\alpha_i$. Since a continuous function with values in a discrete set must be constant, Lemma \[L:memberHardy\] yields: \[C:discHardy\] Given an [o-minimalistic]{} structure $\mathcal M$, if a discrete subset $D\sub ({\mathbf H(\mathcal {M})})^n$ is definable with parameters in $\mathcal M$, then $D\sub M^n$. \[T:Hardy\] Given an [o-minimalistic]{} structure $\mathcal M$, there is a canonical elementary embedding $\mathcal M\to {\mathbf H(\mathcal {M})}$. In particular, ${\mathbf H(\mathcal {M})}$ is also [o-minimalistic]{}. The map $M\to {\mathbf H(\mathcal {M})}$ sending an element $a\in M$ to the class of the corresponding constant function $x\mapsto a$, is easily seen to be an elementary embedding. These two results together show that if $\mathcal M$ is [o-minimalistic]{} but not o-minimal, then $(\mathcal M,{\mathbf H(\mathcal {M})})$ is a Vaughtian pair (see, for instance, [@PillIntro Proposition 9.3]). In particular, o-minimalism has Vaughtian pairs. \[R:proto\] Note that we only really used the properties of ${{\text{DCTC}}}$, so that in the above, we may replace [o-minimalistic]{} by this weaker condition. We can think of ${\mathbf H(\mathcal {M})}$ as a sort of *protoproduct*, in the meaning of a “controlled” subring of an ultraproduct as studied in [@SchUlBook Chapter 9]. Namely, endowing the set $M$ with an ultrafilter containing all right unbounded open intervals, then ${\mathbf H(\mathcal {M})}$ consists of all elements in the ultrapower ${\mathcal M_\natural}$ given by definable maps (whereas an arbitrary element is given by any map). We also can define a *standard part* operator, at least on the subset ${\mathbf H^{\text{fin}}(\mathcal {M})}$ of all *finite* elements, that is to say, the set of all arcs $\alpha$ at infinity represented by some definable, continuous map $f\colon {{\, ]u,\infty[\, }} \to M$ such that $f(\infty)\in M$ (see Remark \[R:lim\] for the definition). Indeed, the value of $f(\infty)$ only depends on $\alpha$, thus yielding a standard part map ${\mathbf H^{\text{fin}}(\mathcal {M})}\to M$. Note, however, that as $\mathcal M$ is not definable in ${\mathbf H(\mathcal {M})}$, neither is ${\mathbf H^{\text{fin}}(\mathcal {M})}$. Planar curves ============= In this section, we fix, without further notice, an [o-minimalistic]{} structure $\mathcal M$. Dimension --------- Let us say that a non-empty definable subset $X\sub M^2$ has *dimension zero* if it is discrete, and *dimension two*, if it has non-empty interior. In the remaining case, we will put ${\operatorname}{dim}(X)=1$ and call $X$ a (generalized) *planar curve*. We will assign to the empty set dimension $-\infty$, in order to make the following formula work (with the usual conventions that $-\infty +n=-\infty$): \[C:fibdim\] Given a definable subset $X\sub M^2$, let $F_e$ be the set of all $a\in M$ for which the fiber ${{X}_{{}}[a]}$ has dimension $e$, for $e=0,1$. Then each $F_e$ is definable and the dimension of $X$ is equal to the maximum of all $e+{\operatorname}{dim}(F_e)$. Being discrete and having interior are definable properties, whence so is being a planar curve, showing that each $F_e$ is definable. The formula then follows by inspecting the various cases by means of Corollary \[C:discfib\] and Proposition \[P:nondisc2\]. \[R:dimdef\] There are several ways of extending this definition to larger arity, and the usual one is to define the dimension of a definable subset $X\sub M^n$ as the largest $d$ such that the image of $X$ under some projection $\pi\colon M^n\to M^d$ has non-empty interior. It follows that a $d$-cell has dimension $d$. \[P:dimulomin\] In an ultra-o-minimal structure $\mathcal M$, a definable set has dimension $e$ [if and only if]{} it is an ultraproduct of $e$-dimensional definable sets. Suppose $\mathcal M$ is the ultraproduct of o-minimal structures $\mathcal M_i$, and let $X=\varphi(\mathcal M)$ be a definable subset. By [Łoś’ Theorem]{}, $X$ is the ultraproduct of the definable sets $X_i:=\varphi(\mathcal M_i)$. The result now follows from the definability of dimension: we leave again the general case to the reader, but the planar case is clear from our definitions. Indeed, both being discrete and having non-empty interior are first-order definable properties, and hence pass through the ultraproduct by [Łoś’ Theorem]{}. For instance, for $e=0$, we have that $X$ is discrete [if and only if]{} almost all $X_i$ are, and in the o-minimal context the latter is equivalent with being finite, that is to say, with having dimension zero. In other words, $X$ is discrete [if and only if]{} it is *ultra-finite* in the terminology of §\[s:ominexp\] below. We may rephrase the previous result as a trichotomy theorem for planar subsets: \[T:triplanar\] In an [o-minimalistic]{} structure $\mathcal M$, a definable planar subset $X\sub M^2$ satisfies exactly one of the following three conditions: 1. \[i:0\] $X$ is discrete, closed, and bounded; 2. \[i:1\] $X$ is nowhere dense, but at least one projection onto a coordinate axis has non-empty interior; 3. \[i:2\] $X$ has non-empty interior. We only need to show that [if and only if]{} $X$ is one-dimensional. The converse is clear from Corollaries \[C:fibdim\] and \[C:dimcl\], and for the direct implication, we must show that a definable subset satisfying cannot be discrete, and this follows from Theorem \[T:projdisc\]. Immediate from the definitions and Corollary \[C:non-disc2\], we have: \[C:dimunion\] The dimension of a union $X_1\cup\dots \cup X_n\sub M^2$ of definable subsets is the maximum of the dimensions of the $X_i$. Nodes ----- Assume $C$ is a planar curve. Let us call a point $P\in C$ a *node*, if for every open box $B$ containing $P$, there is an open subbox $I\times J\sub B$ containing $P$ and some point $x\in I$ such that ${{C}_{{}}[x]}\cap J$ is not a singleton. We denote the set of nodes of $C$ by ${{\operatorname}{Node}( C)}$. We call a node an *edge*, if in the above condition ${{C}_{{}}[x]}\cap J$ can be made empty. By an argument similar to the one proving Corollary \[C:ctuconn\], one shows that a function on an open interval $h$ is continuous [if and only if]{} its graph has no edges (since it is a graph, it cannot have any other type of nodes). Note that the closure of a $1$-cell has at most two edges: indeed, if the cell $C$ is given as the graph of a definable, continuous function $h$ on an interval ${{\, ]a,b[\, }}$, then $\bar C\setminus C$ consists of those points among $(a,h^+(a))$ and $(b,h^-(b))$ that are finite (in the notation of Remark \[R:lim\]), and these are then the edges of $\bar C$. Isolated points are edges, and they form a discrete, closed, and bounded subset. Another special case of an edge is any point lying on an open interval inside a vertical fiber ${{C}_{{}}[a]}$. Let ${{\operatorname}{Vert}( C)}$ be the set of all such edges, called the *vertical component* of $C$. Note that ${{\operatorname}{Vert}( C)}$ is equal to the union of the interiors of all fibers, that is to say, ${{\operatorname}{Vert}( C)}=\bigcup_a({{C}_{{}}[a]})^\circ$, and hence in particular is definable. \[P:discnodes\] The set of nodes of a planar curve is the union of its vertical component and a discrete set. Let $C\sub M^2$ be a planar curve. Replacing $C$ by $C\setminus{{\operatorname}{Vert}( C)}$, we may assume its vertical component is empty. Assume towards a contradiction that $N:={{\operatorname}{Node}( C)}$ is not discrete. Therefore, $\pi(N)$ cannot be discrete by Corollary \[C:discfib\], and hence contains an open interval $I$. For each $x\in I$, let $h(x)$ be the minimal $y\in {{C}_{{}}[x]}$ such that $(x,y)\in N$. By Theorem \[T:disctu\], we may shrink $I$ so that $h$ becomes a continuous function on $I$. In particular, its graph $V$ is a $1$-cell contained in $N$. For each $x\in I$, let $l(x)$ and $u(x)$ be the respective predecessor and successor in ${{C}_{{}}[x]}$ of $h(x)$ (if $h(x)$ is always an extremal element of ${{C}_{{}}[x]}$ then we can adjust the argument accordingly, and so we just assume that $l(x)<h(x)<u(x)$ always exist). Since $(x,h(x))$ is a node and $h$ is continuous, for $y<x$ sufficiently close to $x$, and $J$ an open interval such that $J\cap {{C}_{{}}[x]}=\{h(x)\}$, the intersection $J\cap {{C}_{{}}[y]}$ contains at least one other element besides $h(y)$, necessarily either $l(y)$ or $u(y)$. By , either $l(y)$ belongs to all $J\cap {{C}_{{}}[y]}$, for all $y$ sufficiently close to the left of $x$, or otherwise $u(y)$ does. In particular, for a fixed $x\in I$, we have $h(x)=l^-(x)$ or $h(x)=u^-(x)$. Shrinking $I$ if necessary, then reduces to the case that one of these alternatives happens for every $x\in I$, say, $h(x)=l^-(x)$ for all $x\in I$. Shrinking $I$ even further, we may assume that $l$ is continuous on $I$, and hence $l=h$ on $I$, contradiction. \[L:nonnode\] A point $P$ on a planar curve $C$ is not a node [if and only if]{} there is some open box $B$ containing $P$ such that $C\cap B$ is a horizontal $1$-cell. On the other hand, $P$ is an edge [if and only if]{} it does not belong to any horizontal cell inside $C$. If $P=(a,b)\notin{{\operatorname}{Node}( C)}$, there exist open intervals $I$ and $J$ containing respectively $a$ and $b$ such that ${{C}_{{}}[x]}\cap J$ is a singleton $\{f(x)\}$, for every $x\in I$, and this property is preserved for any subbox of $I\times J$ containing $P$. Hence $f\colon I\to J$ is a definable map with $f(a)=b$. Shrinking $I$ if necessary, we may assume by Theorem \[T:disctu\] that $f$ is continuous on $I$ with a possible exception at $a$. As already observed, $f$ is also continuous at $a$ lest $(a,f(a))$ be a node. Hence the graph of $f$ is a cell equal to $(I\times J)\cap C$. If $P$ is not a node, then by definition, no intersection with an open box around $P$ can be a cell. The second assertion is obvious. \[R:regular\] This allows us to generalize the notion to higher arity: let us say that a point $P$ on a definable subset $X\sub M^n$ is *strongly $e$-regular*, for some $e\leq n$, if there exists an open box $B$ containing $P$ such that $B\cap X$ is an $e$-cell. When $n=2$, a point is strongly $2$-regular [if and only if]{} it is interior, and strongly $0$-regular [if and only if]{} it is isolated. The previous result then says that on a planar curve, a point is strongly $1$-regular [if and only if]{} it is not a node. As with cells, this definition of regularity has a directional bias: nodes are really critical points with respect to projection onto the first coordinate. To break this bias, just taking permutations of the variables does not give enough transformations to turn some point on a curve in a non-nodal position, as for instance the origin on the curve given by $(t,t)$ if $t\leq 0$ and $(-t,-2t)$ if $t\geq 0$. However, if we assume that there is an underlying ordered group (see §\[s:omingroup\] below), then we say that a point $x\in X\sub M^n$ is *$e$-regular*, if after a translation bringing $x$ to the origin $O$, we can find a rotation $\rho$ such that $\rho(O)$ is strongly $e$-regular in $\rho(X)$, where by a *rotation* of $M^n$, we mean a linear map $\rho\colon M^n\to M^n$ given by an invertible matrix of determinant one over $\mathbb Q$ (as we shall see below, any [o-minimalistic]{} expansion of a group is divisible, whence admits a natural structure of a $\mathbb Q$-vector space). \[P:dimcl\] A definable subset $X\sub M^2$ has the same dimension as that of its closure $\bar X$, whereas the dimension of its frontier ${{\operatorname}{fr}( X)}$ is strictly less. If $X$ is discrete, then it is closed by Corollary \[C:closeddiscrete\], and so ${{\operatorname}{fr}( X)}=\emptyset$, proving the assertion in this case. If $X$ has dimension one, then so does $\bar X$ by Corollary \[C:dimcl\]. Let $V:={{\operatorname}{Vert}( C)}$ be the vertical component of $C$ and let $\pi(V)$ be its projection. Since $\pi(V)$ is discrete by Proposition \[P:nondisc2\], the boundary $\partial V$ is equal to the union of all $\partial ({{X}_{{}}[a]})$, whence is discrete by Corollary \[C:discfib\]. Hence, upon removing $V$ from $X$, we may reduce to the case that $X$ has no vertical components. Suppose towards a contradiction that ${{\operatorname}{fr}( X)}$ is a planar curve. By Proposition \[P:discnodes\], the set of nodes on $\bar X$ and on ${{\operatorname}{fr}( X)}$ are both discrete sets, and so, there exists a $P\in {{\operatorname}{fr}( X)}$ which is not a node on ${{\operatorname}{fr}( X)}$ nor on $\bar X$. By Lemma \[L:nonnode\], there exists an open box $B$ containing $P$ such that both $B\cap {{\operatorname}{fr}( X)}$ and $B\cap \bar X$ are cells, and therefore the inclusion $B\cap {{\operatorname}{fr}( X)}\sub B\cap \bar X$ must be an equality. In particular, $B\cap X$ is empty, contradicting that $P$ lies in the closure of $X$. Finally, if $X$ has dimension two, then so must $\bar X$. Let $Y:=X^\circ$ and $Z:=X\setminus Y$. Since $\bar X=\bar Y\cup\bar Z$, we have ${{\operatorname}{fr}( X)}=(\bar Y\setminus X)\cup (\bar Z\setminus X)$, so that it suffices to show that neither of these two differences has interior. The first one, $\bar Y\setminus X$, is equal to $\partial Y$ whence has no interior, being the boundary of an open set. By construction, $Z$ has no interior, and hence by Corollary \[C:dimcl\], neither does its closure. Recall that a *constructible* subset is a finite Boolean combination of open subsets, and hence every one-variable definable subset in an [o-minimalistic]{} structure is constructible. This is still true in higher dimensions: by an easy induction on the dimension, and using that the closure is obtained by adjoining the frontier, Proposition \[P:dimcl\] yields: \[C:const\] In an [o-minimalistic]{} structure $\mathcal M$, any definable subset is constructible. \[C:bddim2\] The boundary of a two-dimensional definable subset of $M^2$ has dimension at most one. Let $X\sub M^2$ has dimension two. Its boundary $\partial X$ is the union of its frontier ${{\operatorname}{fr}( X)}$ and $X\setminus X^\circ$. The former has dimension at most one by Proposition \[P:dimcl\] and the latter has no interior. The result now follows from Corollary \[C:non-disc2\]. Recall that a subset in a topological space is called *codense* if its complement is dense. \[C:codense\] If $Y$ is a codense definable subset of a non-empty definable subset $X\sub M^2$, then ${\operatorname}{dim}(Y)<{\operatorname}{dim}(X)$. If $X$ is discrete, then it is closed by Corollary \[C:closeddiscrete\], and hence its only codense subset is the empty set. If $X$ and $Y$ both have dimension two, then $Y^\circ$ is disjoint from the closure of $X\setminus Y$, contradicting that $Y$ is codense in $X$. So remains the case that $X$ is a curve. If $Y$ is codense in $X$, then it must be contained in the frontier of $X\setminus Y$, and the latter has dimension strictly less than one by Proposition \[P:dimcl\]. Quasi-cells ----------- To obtain a cell decomposition as in the o-minimal case, we must generalize the notion of $1$-cell by the following equivalent conditions: \[L:supercell\] Let $S$ be a union of mutually intersecting $1$-cells in $M^2$. We call $S$ a (horizontal) *$1$-quasi-cell* if it satisfies one of the following equivalent conditions: 1. \[i:nonodes\] $S$ has no nodes; 2. \[i:locfct\] $S$ is the graph of a continuous map $h\colon \pi(S)\to M$ which is *locally definable*, meaning that its restriction to any open interval in its domain is definable. Moreover, $\pi(S)$ is open and convex, and $S$ is a $1$-cell [if and only if]{}  $\pi(S)$ is definable. The implication [ $\Rightarrow$ ]{} is easy, since the graph of a continuous function has no nodes. To show [ $\Rightarrow$ ]{}, suppose $S$ has no nodes, so that in particular, no vertical cell lies inside $S$. Fix $a_1,a_2\in\pi(X)$ and choose $1$-cells $C_1$ and $C_2$ containing $a_1$ and $a_2$ respectively. Let $I_k:=\pi(C_k)$ and let $h_k$ be the definable (continuous) function on $I_k$ whose graph is $C_k$. Let $I:=I_1\cup I_2$. Since $C_1\cap C_2$ is non-empty, so is $I_1\cap I_2$, showing that $I$ is an interval. Let $H$ be the subset of $I_1\cap I_2$ on which $h_1$ and $h_2$ agree, that is to say, $H=\pi(C_1\cap C_2)$. For $a\in H$ with common value $b$, since $(a,b)$ is not a node, there exist open intervals $U$ and $V$ containing respectively $a$ and $b$ such that ${{S}_{{}}[x]}\cap V$ is a singleton, for all $x\in U$. Shrinking $U$ if necessary, continuity allows us to assume that $h_k(U)\sub V$, so that $(x,h_k(x))$ both lie in ${{S}_{{}}[x]}\cap V$ for $x\in U$, whence must be equal. This shows that $U\sub H$, and hence that $H$ is open. Let $a\in \partial H$. Since $H$ is open, $a$ does not belong to $H$ whereas either $a^-$ or $a^+$ does. If $a$ lies in $I_1\cap I_2$, then $a\notin H$ implies $h_1(a)\neq h_2(a)$, and by continuity, this remains the case on some open around $a$, contradicting that either $a^-$ or $a^+$ belongs to $H$. Hence $a\notin I_1\cap I_2$. This means that the only boundary points of $H$ are the endpoints of the interval $I_1\cap I_2$, proving that $h_1$ and $h_2$ agree on this interval. Let $h(x)$ be equal to $h_1(x)$ if $x\in I_1$ and to $h_2(x)$ otherwise. Since the graph of $h$ is then equal to $C_1\cup C_2\sub S$, whence contains no nodes, $h$ is continuous. It is not hard to see that $\pi(S)$ is open and convex. The last assertion then follows since $\pi(S)$ is a disjoint union of open intervals by Theorem \[T:onevar\], whence, being also convex, a single open interval. Although we should also entertain the notion of vertical quasi-cells (see Remark \[R:gensupercell\] below), they do not occur in the analysis of planar subsets. Given a curve $C$ without nodes and a quasi-cell $S\sub C$, we say that $S$ is *optimal* in $C$, if no quasi-cell inside $C$ strictly contains $S$. \[C:supercell\] Any point on a planar curve $C$ without nodes lies on a (uniquely determined) optimal quasi-cell in $C$. In particular, $C$ is a disjoint union of quasi-cells. Fix $P\in C$. By Lemma \[L:nonnode\], there exists a cell $V\sub C$ containing $P$. Let $S$ be the union of all cells inside $C$ containing $P$. Since $S\sub C$, has no nodes, it is a quasi-cell by Lemma \[L:supercell\]. Suppose $S'\sub C$ is a quasi-cell containing $S$ and let $P'\in S'$. By Lemma \[L:supercell\], there exists a cell $V'\sub S'$ containing both $P$ and $P'$. By construction, we then have $V'\sub S$, whence $P'\in S$, showing that $S=S'$ is optimal. If the curve has nodes, then to preserve uniqueness of optimal quasi-cells, we have to amend this definition as follows: for an arbitrary planar curve, a horizontal $1$-quasi-cell $S$ is called *optimal* if $S\sub C$ contains no node of $C$ and is maximal with this property. Finally, we define the notion of a *$2$-quasi-cell* $S\sub M^2$ given as the region between two continuous, locally definable maps defined on an open, convex subset of $M$ (again called the *domain* of the quasi-cell), or an unbounded variant as in the case of $2$-cells. More precisely, let $V\sub M$ be an open convex subset, $f,g\colon V\to M$ continuous and locally definable with $f<g$, then $S$ consist of all $(x,y)$ such that $x\in V$ and $f(x)\diamond y\diamond' g(x)$, with $\diamond,\diamond'$ strict inequality or no condition. By definition of local definability, the restriction of a quasi-cell $S$ to an open interval $I\sub V$, that is to say, $S\cap (I\times M)$ is a cell, and hence every $2$-quasi-cell is the union of $2$-cells and therefore open. Moreover, we can arrange for all these cells in this union to contain a given fixed point of the quasi-cell. \[R:gensupercell\] The definition of an arbitrary $d$-quasi-cell is entirely similar: simply replace in the recursive definition from Remark \[R:gencell\] ‘cell’ by ‘quasi-cell’ and ‘definable, continuous map’ by ‘locally definable, continuous’ map everywhere. Locally definable subsets {#s:locdef} ------------------------- Quasi-cells are particular instances of locally definable subsets, which we now briefly study. In an arbitrary ordered structure $\mathcal M$, we call a subset $X\sub M^n$ *locally definable* if for each point $P\in M_\infty^n$, there exists an open box $B$ containing $P$ such that $B\cap X$ is definable. It is important to include in this definition also the *infinite points* of $M_\infty^n$, that is to say, points with at least one coordinate equal to $\pm\infty$, where, just as an example, we mean by an *open box around an infinite point* like $(0,\infty,-\infty)$ one of the form ${{\, ]u,v[\, }}\times{{\, ]p,\infty[\, }}\times{{\, ]-\infty,q[\, }}$, with $u<0<v$. It is also important to note that the definition applies to all points, not just to those belonging to $X$. In fact, the condition is void if $P$ is either an interior or an exterior point, since then some open box is entirely contained in or entirely disjoint from $X$. So we only need to verify local definability at boundary points and at infinite points. Therefore, any clopen is locally definable. It is not hard to show that a finite Boolean combination of locally definable sets is again locally definable. Moreover, the interior, closure and exterior of a locally definable subset are again locally definable. To see that a $1$-quasi-cell $S$ is locally definable, write it as the graph of a locally definable, continuous map $f\colon V\to M$. We leave the infinite points again to the reader, so fix $P=(a,b)$. If $a$ is exterior to $V$, then $P$ lies in the exterior of $S$, and there is nothing to show. If $a$ lies in the interior of $V$, then there is some open interval $I\sub V$ containing $a$, and $S\cap (I\times M)$ is the graph of $f$ restricted at $I$ by , and the same is true if $a$ is a boundary point of $V$, say on the left, by choosing $I={{\, ]a,q[\, }}\sub V$ for some $q>a$. Using this, it is not hard to see that $2$-quasi-cells are also locally definable. \[P:locdefdisc\] In an [o-minimalistic]{} structure, a discrete set is locally definable [if and only if]{} it is closed and bounded. Let $D\sub M^n$ be discrete. If it is not closed, then there is a $P\in\partial D$ not belonging to $D$. But then the intersection $D\cap B$ with any open box $B$ containing $P$ will have $P$ in its closure, so that $D\cap B$ is not closed, whence cannot be definable by Corollary \[C:closeddiscrete\]. Similarly, if $D$ is not bounded, say, in the first coordinate on the right, then its intersection with any open box of the form ${{\, ]p,\infty[\, }}\times B'$ will still be unbounded, whence not definable by Corollary \[C:closeddiscrete\]. Suppose therefore $D$ is closed and bounded. To check local definability at a boundary point $P$, as it belongs to $D$ by closedness, there is an open box $B$ such that $D\cap B=\{P\}$. To check at an infinite point, we can find an open box around $P$ which is disjoint from $D$. \[C:locdefbd\] In an [o-minimalistic]{} structure $\mathcal M$, the topological boundary of a locally definable subset $Y\sub M$ is a discrete, closed, bounded set. If the locally definable set $\partial Y$ has non-empty interior, it would contain an open interval $I$ and we may shrink this so that $F:=I\cap Y$ is definable. Since $\partial F=I\cap \partial Y=I$, we get a contradiction with . Hence $\partial Y$ has no interior, and so, for $b\in \partial Y$ and an open interval $I$ containing $b$ such that $I\cap \partial Y$ is definable, the latter set, having no interior, must be discrete by , and hence, shrinking $I$ further if necessary, $I\cap \partial Y=\{b\}$. Hence $\partial Y$ is discrete, whence bounded and closed by Proposition \[P:locdefdisc\]. Given an arbitrary ordered structure $\mathcal M$, let ${\mathcal {M}^{\text{loc}}}$ be the structure generated by the locally definable subsets of $\mathcal M$ (formally, we have a language with an $n$-ary predicate $\tt X$ for any locally definable subset $X\sub M^n$, and we interpret $\tt X({\mathcal {M}^{\text{loc}}})$ as the subset $X$). Since the class of locally closed subsets is closed under projection, fibers, and finite Boolean combinations, the definable subsets of ${\mathcal {M}^{\text{loc}}}$ are precisely the locally definable subsets of $\mathcal M$. \[C:omincore\] If $\mathcal M$ is an [o-minimalistic]{} structure, then ${\mathcal {M}^{\text{loc}}}$ is locally o-minimal. Given a one-variable definable subset of ${\mathcal {M}^{\text{loc}}}$, whence a locally definable subset $Y\sub M$, and a point $a\in M_\infty$, we may choose an open interval $I$ around $a$ such that $Y\cap I$ is definable. Since $a^-$ belongs to $Y\cap I$ or to its complement, the same is true with respect to $Y$, proving . Since bounded clopens are locally definable but have no infimum, axiom  usually fails and ${\mathcal {M}^{\text{loc}}}$ is in general not [o-minimalistic]{}. Planar cell decomposition ========================= In o-minimality, *cell decomposition* is the property that we can partition any given definable subset $X$ into a disjoint union of cells. Every point is a $0$-cell but writing $X$ as a union of its points should not qualify as a cell decomposition. Slightly less worse, if $X$ is planar, then each fiber ${{X}_{{}}[a]}$ is a disjoint union of intervals and points, so that we can partition $X$ into points and vertical cells. Of course, in the o-minimal context these pathologies are avoided by demanding the partition be finite. In the [o-minimalistic]{} case, however, we can no longer enforce finiteness, and so to exclude any unwanted partitions, we must impose some weaker restrictions. Moreover, at present, I do not see how to avoid having to use quasi-cells (but see §\[s:tame\] below). Let us introduce the following terminology, which we give only for the planar case (but can easily be extended to larger arity, see Remark \[R:celldec\] below). First we extend the definition of dimension to arbitrary planar subsets (which is not necessarily well-behaved if the subset is not definable) by the same characterization: a non-empty subset $B\sub M^2$ has *dimension* $2$, if it has non-empty interior; *dimension* $1$, if it has empty interior but is non-discrete; and *dimension zero* if it is discrete. We can also define the *local dimension* ${\operatorname}{dim}_P(B)$ of $B$ at a point $P\in M^2$ as the minimal dimension of $B\cap U$ where $U$ runs over all open boxes containing $P$. Note that ${\operatorname}{dim}_P(B)\geq 0$ if $P\in\bar B$. It follows that the dimension of $B$ is the maximum of its local dimensions at all points. It is not hard to see that the dimension of $B$ is the largest $e$ for which it contains an $e$-cell. In particular, a $2$-(quasi-)cell has dimension $2$, whereas a $1$-(quasi-)cell has dimension one. More generally, by Corollaries \[C:uniondisc\] and \[C:non-disc2\] and the local nature of dimension, we showed that: \[L:uniondim\] The dimension of a finite union of $e_i$-quasi-cells is equal to the maximum of the $e_i$. Given a collection $\mathcal B$ of (not necessarily definable) subsets of $M^2$, we say that a definable subset $X\sub M^2$ has a *$\mathcal B$-decomposition*, if there exists a partition $X=\bigsqcup_{i\in I} B_i$ with all $B_i\in\mathcal B$, with the additional property that if $X^{(e)}$ denotes the union of all $e$-dimensional $B_i$ in this partition, then $X^{(e)}$ is definable and has dimension at most $e$, for $e=0,1,2$ (whence of dimension $e$ [if and only if]{} it is non-empty). Put simply, in a decomposition there cannot be too many lower dimensional subsets. By a *cell* decomposition (respectively, a *quasi-cell* decomposition) we mean a $\mathcal B$-decomposition where $\mathcal B$ is the collection of all (quasi-)cells. By Lemma \[L:uniondim\], any finite partition into (quasi-)cells is a cell decomposition (there can be no quasi-cell in a finite decomposition since each subset in the partition is then definable). \[R:celldec\] For higher arities, we define the dimension of a subset $B\sub M^n$ to be the largest $d$ such that it contains a $d$-cell (in case $B$ is not definable, this might be different from the largest $d$ such that the projection of $B$ onto some $M^d$ has non-empty interior, but both are equal in the definable case). The definition of $\mathcal B$-decomposition for a definable set $X\sub M^n$ now easily generalizes: it is a partition of $X$ into sets from $\mathcal B$ such that the union of all sets in this partition of a fixed dimension is a definable subset of dimension at most $d$. \[T:plane\] In an [o-minimalistic]{}  structure $\mathcal M$, any planar definable subset $X\sub M^2$ has a quasi-cell decomposition. There is nothing to show if $X$ has dimension zero. If $X$ is a curve, then ${{\operatorname}{Vert}( X)}$ is a disjoint union of vertical cells (see the proof of Proposition \[P:dimcl\]). So after removing it from $X$, we may assume $X$ has no vertical components. In that case, ${{\operatorname}{Node}( X)}$ is discrete by Proposition \[P:discnodes\], and so after removing it, we may assume $X$ has no nodes, and so we are done by Corollary \[C:supercell\]. So remains the case that $X$ is $2$-dimensional. Let $C:=\partial X$ be its boundary. Since $C$ has dimension at most one by Corollary \[C:bddim2\], and so can be decomposed into disjoint quasi-cells by what we just argued, we may assume, after removing it, that $X$ is moreover open. The projection $\pi(N)$ of the set $N:={{\operatorname}{Node}( C)}$ of all nodes is discrete by Proposition \[P:discnodes\] and Theorem \[T:projdisc\], and therefore $X\cap (\pi(N)\times M)$ can be partitioned into vertical cells. So remains to deal with points $(a,b)\in X$ such that $a\notin \pi(N)$. Since $X$ is open, $b$ is an interior point of ${{X}_{{}}[a]}$. Let $l$ and $h$ be respectively the maximum of $({{C}_{{}}[a]})_{\leq b}$ and the minimum of $({{C}_{{}}[a]})_{\geq b}$, so that the open interval ${{\, ]l,h[\, }}$ lies inside ${{X}_{{}}[a]}$ and contains $b$ (we leave the case that one of these endpoints is infinite to the reader). By choice, neither $(a,l)$ nor $(a,h)$ is a node of $C$, so that by Corollary \[C:supercell\], there are (unique) optimal $1$-quasi-cells $L,H\sub C$ containing $(a,l)$ and $(a,h)$ respectively, say, given as the graphs of locally definable, continuous maps $f\colon V\to M$ and $g\colon W\to M$. Consider all open intervals $I\sub V\cap W$ containing $a$ such that the $2$-cell $C(I;{\left.f\right|_{{I}}}<{\left.g\right|_{{I}}})$ lies entirely inside $X$, and let $Z\sub M$ be the union of all these intervals. Hence $Z$ is open and convex, and $C(Z;{\left.f\right|_{{Z}}}<{\left.g\right|_{{Z}}})$ is an (optimal) $2$-quasi-cell inside $X$, by Lemma \[L:supercell\]. To show that this construction produces a disjoint union of quasi-cells, we need to show that if $(a',b')$ is any point in $S$, then the above procedure yields exactly the same quasi-cell containing $(a',b')$. Indeed, by convexity, we can find an open interval $I\sub V$ containing $a$ and $a'$. Since the intersections of $F$ and $G$ with $I\times M$ are $1$-cells, $C(I;{\left.f\right|_{{I}}}<{\left.g\right|_{{I}}})=S\cap (I\times M)$ is a $2$-cell contained in $X$, whence must lie inside $S$ by construction. To show that this is a decomposition, we induct on the dimension $d$ of $X$, where the case $d=0$ is trivial. In the above, at various stages, we had to remove some subsets of $X$ of dimension strictly less than $d$, and partition each separately. Since each of these finitely many exceptional sets was definable, so is their union and by Lemma \[L:uniondim\], has dimension strictly less than $d$. Hence the complement $X^{(d)}$, consisting of all $d$-quasi-cells in the partition, is also definable. After removing $X^{(d)}$, we are left with a definable subset of strictly less dimension, and so we are done by induction. The proof gives in fact some stronger results, where for the sake of brevity, we will view any point as a $0$-quasi-cell: \[R:plane\] Keeping track of the various (quasi-)cells, we actually showed that we may partition $X$ in quasi-cells $S_i$, such that each $\bar S_i\cap X$ is a disjoint union of $S_i$ and some of the other $S_j$. \[R:border\] For a definable subset $Y\sub M$, define a *left outer point* (respectively, a *right outer point*) as a point $b\in Y$ such that $(\partial Y)_{\leq b}$ (respectively, $(\partial Y)_{\geq b}$) is finite. For a subset $X\sub M^2$, we call $(a,b)$ an *outer point*, if $b$ is an outer point in ${{X}_{{}}[a]}$. We can now choose a partition $X=\bigsqcup_i S_i$ in quasi-cells $S_i$ such that any quasi-cell $S_i$ containing an outer point is a cell. Indeed, for a fixed $n$, let $f_n\colon M\to X$ be defined by letting $f_n(x)$ be the $n$-th left outer point in $\partial ({{X}_{{}}[x]})$ (note that this is a definable function by ). The set where $f_n$ takes finite values is definable, and after removing a discrete set, $f_n$ is continuous on it, whence defines a cell above each open interval in its domain. So, given a (left) outer point $(a,b)$, if $n$ is the cardinality of $\partial ({{X}_{{}}[a]})_{\leq b}$, then $(a,b)$ lies either on the graph of $f_n$ or between the graphs of $f_n$ and $f_{n+1}$, and therefore, on a cell. \[C:compactcell\] In an [o-minimalistic]{} structure $\mathcal M$, any definable subset of $M^2$ lying inside a compact subset admits a finite partition into cells. We induct on the dimension $d$ of the definable subset $X\sub M^2$ contained in the compact set $K$. By Corollary \[C:closeddiscrete\], if $d=0$, then $X$ is closed, whence compact, and discrete, whence finite. So assume $X$ is a curve. Let $D$ be the $a\in M$ for which ${{X}_{{}}[a]}$ has non-empty interior, so that $D$ is a discrete subset of the compact $\pi_1(K)$, whence finite. By Proposition \[P:locomin\], since each ${{X}_{{}}[a]}$ lies in the compact set $ \pi_2(K)$, it is a finite union of intervals, and hence ${{\operatorname}{Vert}( X)}$, being the union of all ${{X}_{{}}[a]}$ for $a\in D$, is a finite union of cells. Removing it, we reduced to the case that $X$ has no vertical components. In particular, each fiber is discrete whence finite, and so is ${{\operatorname}{Node}( X)}$ by Proposition \[P:discnodes\]. Upon removing the latter, we may assume $X$ has no nodes. Apply the quasi-cell decomposition procedure from Theorem \[T:plane\] to $X$, which we will call here the *canonical quasi-cell decomposition*. Since all its fibers are finite, every point is an outer point in the terminology of Remark \[R:border\], and hence belongs to a cell. In other words, the partition does not contain quasi-cells and so is a cell decomposition. Suppose there are infinitely many $1$-cells in this partition. Since each cell inside $K$ must have exactly two boundary points, and $\partial X$ is finite, there must be at least two of these boundary points common to infinitely many cells. But then, by continuity, any vertical line in between these two points would intersect each of these cells in a different point, contradicting that the fibers are all finite. So remains the case that $d=2$. We follow again the procedure, and its notation, from Theorem \[T:plane\] to produce the canonical quasi-cell decomposition of $X$. All the parts of $X$ we will throw out in this procedure are lower dimensional and hence have finite cell decomposition by induction. So we may reduce to the case that $X$ is open and that $(a,b)\in X$ has the property that no node of $\partial X$ lies in the closure of ${{X}_{{}}[a]}$. The definable maps $f\colon V\to M$ and $g\colon W\to M$ as in the proof of Theorem \[T:plane\] are uniquely determined by $(a,b)$ and $X$, and their graphs are part of a cell decomposition of $\partial X$. In particular, by induction, there are only finitely many choices for $f$ and $g$. Since the $2$-cell $C(Z;{\left.f\right|_{{Z}}}<{\left.g\right|_{{Z}}})$ in the canonical cell decomposition containing $(a,b)$ is uniquely determined by the property that $Z$ is the largest open interval in $V\cap W$ such that the $2$-cell lies inside $X$, there are only finitely many possibilities for it. \[R:compactcell\] The analogue of Proposition \[P:locomin\] also holds: if $X\sub M^2$ is definable and $K\sub M^2$ is compact and open, then $X\cap K$ admits a finite cell decomposition. Indeed, choose for each $a\in K$ and open box $U_a$ containing $a$ and contained in $K$. Since each $X\cap U_a$ is definable and contained in $K$, it admits a finite cell decomposition by the previous result. But by compactness, $K$ is the union of finitely many of the $U_a$, proving the claim. Both the Corollary and the Remark extend to higher arities, the detail of which we leave to the reader. Tameness {#s:tame} ======== The quasi-cell decomposition version given by Theorem \[T:plane\] is not very useful in applications. Moreover, the non-definable nature of quasi-cells is a serious obstacle. Perhaps quasi-cells never occur, but in the absence of a proof of this, we make the following definitions, for $\mathcal M$ any ordered structure. Let us call a definable map $c\colon X\to Y$ *pre-cellular*, if every fiber ${{c^{-1}(y)}}$ is a cell. Note that the non-empty fibers of $c$ then constitute a partition of $X$ into cells. Injective maps are pre-cellular, but the resulting partition in cells is clearly not a decomposition if $X$ has positive dimension. To guarantee that we get a cell decomposition, we require moreover that the image of $c$ be discrete, bounded, and closed, and we call such a map simply *cellular*. In particular, we may assume, if we wish to do so, that the cellular map $c\colon X\to D$ is surjective, where $D$ is discrete, bounded, and closed. Assume now that $\mathcal M$ is [o-minimalistic]{} (or, merely a model of ${{\text{DCTC}}}$), and let $c\colon X\to D$ be cellular. The collection $X^{(e)}$ of all fibers ${{c^{-1}(y)}}$ of dimension $e$ is a definable subset, for each $e$, since we can express in a first-order way whether a fiber ${{c^{-1}(y)}}$ has dimension $e$ (for instance, if $X$ is planar, then having interior or being discrete are elementary properties). If $X^{(e)}$ is non-empty, then its dimension is equal to $e$ by Corollary \[C:fibdim\] and the fact that $D$ is discrete, showing that we have indeed a cell decomposition. Note that in particular, the graph of $c$ and $X$ have the same dimension. We have the following converse: given a definable subset $X\sub M^k$ and a cellular map $c\colon X\to Y$ such that $X=\bigsqcup_{y\in c(X)}{{c^{-1}(y)}}$ is a cell decomposition, then $c$ is cellular, that is to say, $c(X)$ is discrete. It suffices to show that each $c(X^{(e)})$ is discrete, for $e\leq k$, and this follows from Corollary \[C:fibdim\]. It is not clear whether every definable subset admits a cell decomposition of this type, and so we make the following definition: \[D:tame\] A definable subset $X\sub M^n$ in an ordered structure $\mathcal M$ is called *[tame]{}* if it is the domain of a cellular map. If every definable subset in $\mathcal M$ is [tame]{}, then we call $\mathcal M$ *[tame]{}*. \[L:tameDed\] Any [tame]{} structure is a model of ${{\text{DCTC}}}$. By Theorem \[T:onevar\], it suffices to show that any definable subset $Y\sub M$ is a disjoint union of open intervals and a single discrete, bounded, and closed subset. Let $c\colon Y\to D$ be cellular, with $D$ discrete, bounded, and closed. Hence, each fiber ${{c^{-1}(a)}}$ must be a one-variable cell, that is to say, either a point or an open interval. Let $E$ be the subset of all $a\in D$ for which ${{c^{-1}(a)}}$ is a point. Hence $Y\setminus{{c^{-1}(E)}}$ is a disjoint union of open intervals, so that upon removing them, we may assume $E=D$, so that $c$ is a bijection. Since $D$ is discrete, ${{c^{-1}}}$ is continuous. Therefore, $Y={\operatorname}{Im}({{c^{-1}}})$ is closed and bounded by [@MillIVT Prop. 1.10] (which we may invoke since holds). Moreover, by the same result, any closed subset is again mapped to a closed subset, showing that ${{c^{-1}}}$, whence also $c$, is a homeomorphism. In particular, $Y$ is discrete, as we needed to show. I do not know whether a [tame]{} structure is always [o-minimalistic]{}; in fact, I do not know whether —an [o-minimalistic]{} property to be defined below in §\[s:gr\]—holds in a [tame]{} structure. We now investigate when an [o-minimalistic]{} structure is [tame]{}. However, since so far we only used the axioms of ${{\text{DCTC}}}$ to derive properties of [o-minimalistic]{} structures, everything works within this weaker theory. So, for the remainder of this section, unless noted elsewhere, $\mathcal M$ is a model of ${{\text{DCTC}}}$. Clearly any cell is [tame]{}. Since a (principal) fiber of a cell is again a cell, the same holds for [tame]{} subsets. Since a principal projection of a cell is a cell, the collection of [tame]{} subsets is closed under principal projections (we will generalize this in Corollary \[C:symm\] below). Any finite cell decomposition is easily seen to be given by a cellular map, and hence in particular, any o-minimal structure is [tame]{}. \[P:strongtame\] Suppose $\mathcal M$ and $\tilde{\mathcal M}$ are elementary equivalent ordered structures. If $\mathcal M$ is [tame]{}, then so is $\tilde{\mathcal M}$. Since both structures have isomorphic ultraproducts, we only need to show that [tame]{}[ness]{} is preserved under elementary substructures and extensions. The former is easy, so assume $\mathcal M$ is a [tame]{} elementary substructure of $\tilde{\mathcal M}$ and let $\tilde X$ be a definable subset in $\tilde M^n$. Since [tame]{}[ness]{} is preserved under fibers, we may assume that $\tilde X$ is definable without parameters, say $\tilde X=\varphi(\tilde{\mathcal M})$. By assumption, there exists a cellular map $c\colon \varphi(\mathcal M)\to D$, that is to say, formulae $\gamma$ and $\delta$, with $\gamma(\mathcal M)$ the graph of a map all of whose fibers are cells of dimension at most $n$ and whose image is the discrete, closed, bounded subset $\delta(\mathcal M)$. Since all this is first-order, it must also hold in $\tilde{\mathcal M}$, so that $\gamma(\tilde{\mathcal M})$ is the graph of a cellular map $\tilde c\colon \tilde X\to \delta(\tilde{\mathcal M})$. By Theorem \[T:Hardy\], the associated Hardy structure of a [tame]{} structure is therefore again [tame]{}. An interesting question is whether any ultraproduct of [tame]{} structures is [tame]{}. If so, then any [o-minimalistic]{} structure is [tame]{}, since it is elementary equivalent by Corollary \[C:ulomin\] with an ultra-o-minimal structure, and the latter would then be [tame]{}, whence so would the former be by Proposition \[P:strongtame\]. A priori, the analogue of Lemma \[L:ominred\] will fail, that is to say, [tame]{}[ness]{} may not be preserved under reducts, as some cellular maps may no longer be definable, but we have: \[P:tamered\] In a reduct of a [tame]{} structure, every definable set admits a cell decomposition. Let $\mathcal M$ be a [tame]{} structure, $\bar{\mathcal M}$ some reduct, and $X$ an $\bar{\mathcal M}$-definable subset. By assumption, there exists an ${\mathcal M}$-definable cellular map $c\colon X\to D$, the fibers of which yield an ${\mathcal M}$-cell decomposition of $X$. We will need to show how we can turn this into an $\bar{\mathcal M}$-cell decomposition. As always, we only treat the planar case, $X\sub M^2$. There is nothing to show if $X$ is discrete, so assume it is a curve. We already argued that its vertical component ${{\operatorname}{Vert}( X)}$ admits a cell decomposition, and so we may remove it. The remaining set of nodes is discrete, and hence may be removed as well, so that we are left with the case that $X$ has no nodes. By Corollary \[C:supercell\], every point of $X$ lies on a unique optimal quasi-cell. Hence if $C:={{c^{-1}(d)}}$ is one of the cells in the above decomposition, then it is contained in a unique $\bar{\mathcal M}$-quasi-cell $S$. As $C$ is then the restriction of $S$ to $I$, it is $\bar{\mathcal M}$-definable by Lemma \[L:supercell\]. If $X$ has dimension two, we may assume it is open after removing its boundary, as we already dealt with curves. By Theorem \[T:plane\], there exists an $\bar{\mathcal M}$-quasi-cell decomposition of $X$. Following that proof, we may assume, after removing all points lying on a vertical line containing a node of $\partial X$, that any quasi-cell $S$ in this decomposition is open, and its boundary consists of quasi-cells of $\partial X$. By the one-dimensional case, the latter decompose into $\bar{\mathcal M}$-cells, whence so does $S$. \[R:tamered\] We have the following puzzling fact that at least one among the following three statements holds: 1. \[i:tamenonomin\] there is a [tame]{} structure which is not [o-minimalistic]{}; 2. \[i:nontameomin\] there is an [o-minimalistic]{} structure which is not [tame]{}; 3. \[i:tamered\] any reduct of a [tame]{} structure is again [tame]{}. Indeed, suppose both and fail. So, by the latter, there is a [tame]{} structure $\mathcal M$ with a non-[tame]{} reduct $\bar{\mathcal M}$, and by the former, $\mathcal M$ is [o-minimalistic]{}, whence so is $\bar{\mathcal M}$ by Lemma \[L:ominred\]. Hence $\bar{\mathcal M}$ is [o-minimalistic]{} but not [tame]{}. Note that implies that ${{\text{DCTC}}}$ is not equal to ${T^{\text{omin}}}{}$, and by the discussion preceding the proposition, implies that [tame]{}[ness]{} is not preserved under ultraproducts. \[L:onetame\] In a model $\mathcal M$ of ${{\text{DCTC}}}$, every one-variable definable subset is [tame]{}. Most proofs involving [tame]{}[ness]{} will require some coding of disjoint unions, and as we will gloss over this issue below, let me do the proof in detail here. For ease of discussion, let us assume $Y\sub M$ is bounded (the unbounded case is only slightly more complicated and left to the reader). Assume $0$ and $1$ are distinct elements in $M$. Define $c\colon Y\to M^2$ by letting $c(y)$ be equal to $(y,0)$, in case $y\in \partial Y$; and equal to $(x,1)$ where $x$ is the maximum of $(\partial Y)_{<y}$, in the remaining case. The fiber ${{c^{-1}(d,e)}}$ is either a point in $\partial Y$ (when $e=0$), or the interval ${{\, ]d,\sigma_{\partial Y}(d)[\, }}\sub Y$. Since its image is contained in $\partial Y\times\{0,1\}$, the map $c$ is cellular by . \[R:onetame\] As it will be of use later, note that by the above argument, we can refine the cell decomposition given by $c$ as follows: for any discrete subset $D$ containing $\partial Y$, we can construct a cellular map $c_D\colon Y\to M^2$ whose cells have endpoints in $D$. To facilitate the coding of certain disjoint unions, we make the following definitions. Given a definable subset $X\sub M^n$, we define its *valence* as the smallest $d$ such that $X\preceq M^d$, that is to say, the smallest $d$ for which there exists an injective, definable map $X\to M^d$ (see the paragraph preceding Theorem \[T:Euldim\] below). We say that *definable discrete subsets are univalent*, if every discrete subset $D\sub M^n$ has valence one. The advantage of the assumption that definable discrete subsets are univalent is that in the definition of definable cell decomposition, we may always take $n=1$. If $\mathcal M$ expands an ordered field, then its definable discrete subsets are univalent. Indeed, by induction, it suffices to show that if $D\sub M^{n+1}$ is discrete, then there is a definable, injective map $g\colon D\to M^n$. The set of lines connecting two points of $D$ is again a discrete set (in the corresponding projective space) and hence we can find a hyperplane which is non-orthogonal to any of these lines. But then the restriction to $D$ of the projection onto this hyperplane is injective. \[P:defcelldiscunion\] Let $\mathcal M$ be a model of ${{\text{DCTC}}}$ whose definable discrete subsets are univalent (e.g., an expansion of an ordered field) and let $g\colon X\to M^n$ be a definable map with discrete image. If every fiber ${{g^{-1}(a)}}$ is [tame]{}, then so is $X$. Let $A:=g(X)$. By assumption, there exists for each $a\in A$, a cellular map $c_a\colon {{g^{-1}(a)}}\to D_a$ with $D_a\sub M$ discrete. Let $D$ be the union of all $\{a\}\times D_a\sub M^{n+1}$, for $a\in A$. It follows from Corollary \[C:discfib\] that $D$ is discrete. Define $c\colon X\to D$ by the rule $c(x)=(g(x),c_{g(x)}(x))$. To see that this is cellular, note that the fiber over a point $(a,d)\in D$ is equal to ${{c_a^{-1}(d)}}$ for $d\in D_a$, whence is by assumption a cell. \[R:defcelldiscunion\] The condition on the univalence of discrete sets can be relaxed: let us say that a subset $X\sub M^k$ is *$e$-[tame]{}*, if there is a cellular map $g\colon X\to M^e$. Hence under the assumption that all definable discrete subsets are univalent, [tame]{} is the same as $1$-[tame]{}. If now every fiber ${{g^{-1}(a)}}$ is merely $e$-[tame]{}, the above proof still goes through to show that $X$ is [tame]{}. It is also not necessary to assume that $\mathcal M$ is a model of ${{\text{DCTC}}}$, provided we impose that the image of $g$ be also closed and bounded. Indeed, one can directly prove that $D$ as in the proof is discrete, closed and bounded, without any appeal to Corollary \[C:discfib\]. \[T:defcellBool\] In a model of ${{\text{DCTC}}}$, the collection of [tame]{} subsets is closed under (finite) Boolean combinations. Let $\mathcal M\models{{\text{DCTC}}}$. We only treat the case that the definable discrete subsets are univalent, where the general case follows again by a more careful analysis, based upon Remark \[R:defcelldiscunion\]. Moreover, we restrict for simplicity to the case of planar subsets, and leave the general case to the reader (by an induction on the arity). Since the complement of a cell $V\sub M^2$ is a finite union of cells, it is [tame]{}. For instance, if $V=C({{\, ]a,b[\, }};f<g)$, then its complement consists of the four $2$-cells ${{\, ]-\infty,a[\, }}\times M$, ${{\, ]b,\infty[\, }}\times M$, $C(I;-\infty<f)$ and $C(I;g<\infty)$, and the four $1$-cells, $a\times M$, $b\times M$ and the graphs of $f$ and $g$. Since any union can be written as a disjoint union by taking complements, an application of Proposition \[P:defcelldiscunion\] then reduces to showing that the intersection of two cells $V_1$ and $V_2$ in $M^2$ is [tame]{}. This is trivial if either one is discrete, whence a singleton. Suppose $V_1$ is a $1$-cell, given by the definable, continuous map $f_1\colon I_1\to M$. Let $Y$ be the subset of all $x\in I_1$ such that $(x,f_1(x))$ belongs to $V_2$. Choose a cellular map $c\colon Y\to D$ (by Lemma \[L:onetame\], or, for higher arities, by induction). Its composition with the (bijective) projection $V_1\cap V_2\to Y$ is then also cellular. Suppose next that $V_i=C(I_i;f_i<g_i)$ are both $2$-cells, assumed once more for simplicity to be bounded. Let $I:=I_1\cap I_2$ and for $x\in I$, let $f(x)$ be the maximum of $f_1(x)$ and $f_2(x)$, and let $g(x)$ be the minimum of $g_1(x)$ and $g_2(x)$. Note that $f$ and $g$ are continuous on $I$. Let $Y$ consist of all $x\in I$ for which $f(x)<g(x)$, and let $c\colon Y\to D$ be cellular. The composition of $c$ with the projection $V_1\cap V_2\to Y$ is again cellular, since its fibers are the cells $C({{c^{-1}(a)}};f<g)$. \[E:nontamenondc\] It is important in this result that the structure is already a model of ${{\text{DCTC}}}$. For instance, let $D$ be the subset of the ultrapower ${{\mathbb R}_\natural}$ of the reals (viewed as an ordered field) consisting of all elements of the form $n$ or ${{\omega_\natural}}-n$, for $n\in{\mathbb N}$. Note that $D$ is closed, bounded, and discrete, and hence [tame]{}. However, $({{\mathbb R}_\natural},D)$ is not [tame]{}, since ${\mathbb N}=D_{<{{\omega_\natural}}/2}$ is definable in it but fails to satisfy , and so, $({{\mathbb R}_\natural},D)$ is not even a model of ${{\text{DCTC}}}$. In particular, in the terminology from below, $D$ is not [o-finitistic]{} by Corollary \[C:ofindisc\]. It is not hard to show that the product of two cells is again a cell. Therefore, the product of two [tame]{} subsets is again [tame]{}. Similarly, the fiber of a cell is again a cell, and hence if $X\sub M^n$ is [tame]{}, then so is each fiber ${{X}_{{}}[{\mathbf {a}}]}$. Together with Theorem \[T:defcellBool\] and the fact that a principal projection of a [tame]{} subset is again [tame]{}, we showed that the collection of [tame]{} subsets determines a first-order structure on $M$ (in the sense of [@vdDomin Chapt. 1, 2.1], with a predicate for every [tame]{} subset of $M^n$). Calling this induced structure on $M$ the *[tame]{} reduct* of $\mathcal M$ and denoting it ${\mathcal {M}^{\text{tame}}}$, is justified by: \[C:tamered\] If $\mathcal M\models{{\text{DCTC}}}$, then ${\mathcal {M}^{\text{tame}}}$ is [tame]{}, whence in particular a model of ${{\text{DCTC}}}$. The definable subsets of ${\mathcal {M}^{\text{tame}}}$ are precisely the [tame]{} definable subsets of $\mathcal M$, so that in particular, $\mathcal M$ and ${\mathcal {M}^{\text{tame}}}$ have the same cells. So remains to show that if $c\colon X\to D$ is cellular in $\mathcal M$, then it is also cellular in ${\mathcal {M}^{\text{tame}}}$. Discrete sets are [tame]{} by definition, and the graph $\Gamma(c)$ of $c$ is [tame]{} by applying Proposition \[P:defcelldiscunion\] to the projection $\Gamma(c)\to D$. In particular, $c$ is ${\mathcal {M}^{\text{tame}}}$-definable, and since its fibers are cells, we are done. The last assertion then follows from Lemma \[L:tameDed\]. \[P:tameredomin\] If $\mathcal M$ is [o-minimalistic]{}, then so is ${\mathcal {M}^{\text{tame}}}$. Let $\bar L$ be the language with predicates for the [tame]{} subsets of $\mathcal M$, so that ${\mathcal {M}^{\text{tame}}}$ is an $\bar L$-structure. Viewing $\mathcal M$ as a structure in the language having a predicate for every definable subset of $\mathcal M$ yields again a [tame]{} structure, since we added no new definable subsets (see Lemma \[L:expdef\] below). Therefore, upon replacing $L$ by the latter language, we assume from the start that $\bar L\sub L$, and the result follows from Lemma \[L:ominred\]. We also have the following joint cell decomposition: \[C:jointtame\] Given [tame]{} subsets $Y_1,\dots,Y_n$ of a [tame]{} subset $X$ in $\mathcal M\models{{\text{DCTC}}}$, there exists a cellular map $c\colon X\to D$, such that for each $i$, the restriction of $c$ to $Y_i$ is also cellular. Since any Boolean combination of [tame]{} subsets is again [tame]{} by Theorem \[T:defcellBool\], we may reduce first to the case that all $Y_i$ are disjoint, and then by induction, that we have a single [tame]{} subset $Y\sub X$. Since $X\setminus Y$ is [tame]{} too, we have cellular maps $d\colon Y\to D$ and $d'\colon X\setminus Y\to D'$, and their disjoint union is then the desired cellular map. We call a definable map *[tame]{}*, if its graph is. Note that its domain then must also be [tame]{}. As already observed in the previous proof, cellular maps are [tame]{}. To characterize [tame]{} maps, we make the following observation/definition: given a cellular map $c\colon X\sub M^n\to D$, for $e\leq n$, let $X^{(e)}_c=X^{(e)}$ be the union of all $e$-dimensional ${{c^{-1}(a)}}$. Since dimension is definable, so is each $X^{(e)}$, and hence the restriction of $c$ to $X^{(e)}$ is also cellular, proving in particular that each $X^{(e)}$ is [tame]{}. \[T:tamemap\] In a model of ${{\text{DCTC}}}$, a definable map $f\colon X\to M^k$ is [tame]{} [if and only if]{}  $X$ is [tame]{}, and the restriction of $f$ to the set of its discontinuities is also [tame]{}. In particular, a definable, continuous map with [tame]{} domain is [tame]{}. If $f$ is [tame]{}, then $f$ is ${\mathcal {M}^{\text{tame}}}$-definable, and hence so is its set of discontinuities $X'$, proving that $X'$ is [tame]{}. Since the graph of ${\left.f\right|_{{X'}}}$ is $\Gamma(f)\cap (X'\times M^k)$, the restricted map is again [tame]{}  by Theorem \[T:defcellBool\]. For the converse, $U:=X\setminus X'$ is [tame]{} by Theorem \[T:defcellBool\], so that we have a cellular map $c\colon U\to D$. Since $f$ is continuous on $U$, the composition of $c$ with the principal projection $\Gamma({\left.f\right|_{{U}}})\to U$ is also cellular, showing that $\Gamma({\left.f\right|_{{U}}})$ is [tame]{}. Since by assumption, the graph of the restriction to $X'$ is [tame]{}, so is $\Gamma(f)=\Gamma({\left.f\right|_{{U}}})\cup \Gamma({\left.f\right|_{{X'}}})$ by Theorem \[T:defcellBool\], showing that $f$ is [tame]{}. A [tame]{} map is ${\mathcal {M}^{\text{tame}}}$-definable, and hence so its its image, proving: \[C:symm\] In a model $\mathcal M$ of ${{\text{DCTC}}}$, if the domain of a definable, continuous map is [tame]{}, then so is its image. More generally, the image of a [tame]{} subset under a [tame]{} map is again [tame]{}. Let us call a definable map $f\colon X\sub M^n\to M^k$ *almost continuous*, if its set of discontinuities is discrete. By Theorem \[T:disctu\], any one-variable definable map in a model of ${{\text{DCTC}}}$ is almost continuous. Given a definable map $f\colon X\to M^k$, let us inductively define $D_i(f)\sub X$, by setting $D_0(f):=X$, and by setting $D_i(f)$, for $i>0$, equal to the set of discontinuities of the restriction of $f$ to $D_{i-1}(f)$. By Remark \[R:genctu\], each $D_i(f)$ has strictly lesser dimension than $D_{i-1}(f)$, and hence $D_n(f)$ is empty for $n$ bigger than the dimension of $X$. Hence $f$ is (almost) continuous if $D_1(f)$ is empty (respectively, discrete). Since the domain of a [tame]{} function is [tame]{}, an easy inductive argument using Theorem \[T:tamemap\] immediately yields: \[C:almostctu\] In a model of ${{\text{DCTC}}}$, an almost continuous (e.g., a one-variable) map with [tame]{} domain is [tame]{}. In particular, a definable map $f$ is [tame]{} [if and only if]{} all $D_i(f)$ are [tame]{}. Let us say that an ordered structure is *almost continuous*, if apart from a binary predicate denoting the order, all other symbols represent almost continuous functions. \[C:tame\] If is $\mathcal M$ an almost continuous model of ${{\text{DCTC}}}$, then $\mathcal M$ is [tame]{}. Since the collection of [tame]{} subsets is closed under Boolean operations, projections, and products by Theorem \[T:defcellBool\], we only have to verify that the ones defined by unnested atomic formulae are [tame]{}. Since by assumption the only predicate is the inequality sign, and the set it defines is a cell, we only have to look at formulae of the form $f(x)=g(x)$ or $f(x)<g(x)$, with $f, g$ function symbols. Since $f$ and $g$ are total functions representing almost continuous maps, their graphs are [tame]{} by Corollary \[C:almostctu\], whence so is their intersection by Theorem \[T:defcellBool\]. The projection of the latter is the set defined by $f(x)=g(x)$, proving that is a [tame]{} subset. Let $F$ and $G$ be the subsets of $M^{n+2}$ of all $(a,f(a),c)$ and all $(a,b,g(a))$ respectively, with $a\in M^n$ and $b,c\in M$. Since these are just products of the respective graphs and $M$, both are [tame]{}, and so is the subset $E$ of all $(a,b,c)\in M^{n+2}$ with $a<b$. Therefore, by another application of Theorem \[T:defcellBool\], the intersection $F\cap G\cap E$ is [tame]{}, and so is its projection, which is just the set defined by the relation $f(a)<g(a)$. \[R:tame\] More generally, by the same argument, if $\mathcal M\models{{\text{DCTC}}}$ is an expansion of a [tame]{} structure by [tame]{} functions and by predicates defining [tame]{} subsets, then $\mathcal M$ itself is [tame]{}. [O-minimalistic]{} expansions of groups {#s:omingroup} ======================================= In this section, we study [o-minimalistic]{} $L$-expansions of (ordered) groups. As in the o-minimal case, they are definably simple, whence in particular divisible and Abelian: \[T:Dedlimgr\] Any [o-minimalistic]{} expansion of a group is divisible and commutative, and has no proper, definable subgroups. Let $G$ be an [o-minimalistic]{} expansion of a group. Since we do not assume $G$ to be commutative, we will use the multiplicative notation (although we will continue to write $x^-$ and $x^+$). To show divisibility, fix some $p\in{\mathbb N}$ and let $f\colon G\to G$ be the $p$-th power map $x\mapsto x^p$. Since $f$ is definable and continuous, its image is definably connected, whence an interval by . Since the $p$-th powers are cofinal, this image interval must be the whole set, proving divisibility. In particular, since ordered groups are torsion-free, any element $c\in G$ has by assumption a unique $p$-th root, denoted $\sqrt[p] c$. We start with showing that every definable subgroup $1\neq H\subsetneq G$ is convex. If $H$ were discrete, it would admit a maximum $b>1$ by . Since then $b<b^2\in H$, we get a contradiction. Let us next show that $H$ cannot have any isolated point. Indeed, by , there are $a<b$ with ${{\, [a,b]\, }}\sub H$. If $h\in H$, then $h{{a^{-1}}}H=H$ contains the closed interval ${{\, [h,h{{a^{-1}}}b]\, }}$, showing that $h$ is not isolated. Let $b$ be the infimum of $(M\setminus H)_{>1}$. To establish convexity, it suffices to show that $H_{>b}$ is empty, so towards a contradiction, suppose not and let $c$ be its infimum. Since $b^+$ belongs to $M\setminus H$, we must have $c>b$ and ${{\, ]b,c[\, }}$ is disjoint from $H$. Since neither $1$ nor $c$ can be isolated by our previous observation, $1<b$ and $c^+$ belongs to $H$. In particular, ${{\, [1,b[\, }}$ and ${{\, ]c,e[\, }}$, for some $e>c$, are contained in $H$. Choose $d>1$ strictly smaller than $b$ and $e{{c^{-1}}}$. Since $d$ is in ${{\, ]1,b[\, }}$, it belongs to $H$, and since $c<dc<e$, so does $dc$. Hence $c\in H$. Since $\sqrt b$ is less than $b$, it too belongs to $H$, whence so does $b$. Hence $c{{\, [1,b]\, }}={{\, [c,cb]\, }}\sub H$. Since $b<b^2\in H$, we must have $c\leq b^2$. Now, take any $u\in {{\, ]b,c[\, }}$. Hence $u\notin H$, but, on the other hand, from $b^2<bu<bc$, we get $c<bu<bc$, showing that $bu$ whence also $u$ belongs to $H$, contradiction. So $H$ is convex, whence an interval with endpoints $b'$ and $b$, the respective infimum and supremum of $H$. Since $H\neq G$, we must have $b\in G$. Since $\sqrt b<b<b^2$, the former lies in $H$, whence so do $b$ and $b^2$, contradiction. The commutativity of $G$ follows now immediately by observing that any centralizer, being definable and non-trivial, must be the whole group. \[R:ordgr\] Since an [o-minimalistic]{} structure has the IVT, the fact that an [o-minimalistic]{} expansion of a group is Abelian and divisible then already follows from [@MillIVT Proposition 2.2]. From now on, we will write [o-minimalistic]{} expansions of groups additively. \[L:Dedlimdisc\] Let $\mathcal M$ be an [o-minimalistic]{}  expansion of a group. For any definable subset $Y\sub M$, the topological boundary $\partial Y$ is uniformly discrete, that is to say, there exists $a>0$ such that $a\leq y-x$ for any pair of topological boundary points $x<y$. By , the topological boundary $\partial Y$ is discrete. Consider the the definable map $\partial Y\to M\colon b\mapsto \sigma(b)-b$, where $\sigma=\sigma_{\partial Y}$ is the successor function of $\partial Y$ given by . By Lemma \[L:imdisc\], its image is also discrete, whence has a minimum $a>0$ by , satisfying therefore the required property. \[R:spread\] It follows from the above proof, that there is a maximal $a$ so that $a\leq y-x$ for all $x<y\in\partial Y$, which we call the *spread* $a_{\text{spr}}(Y)$ (and, in particular, there is some $b\in \partial Y$ so that $a+b\in\partial Y$). \[T:ominth\] Every Archimedean, [o-minimalistic]{} expansion of a group is o-minimal. Let $\mathcal M$ be an [o-minimalistic]{} expansion of a group and let $Y\sub M$ be definable. By , it suffices to show that $\partial Y$ is finite, and so, upon replacing $Y$ by its boundary, we may already assume that $Y$ is discrete. Towards a contradiction, suppose $Y$ is not finite and let $a:=a_{\text{spr}}(Y)>0$ be its spread. Let $H$ be the (non-empty) set of all $x\in M$ such that $Y_{<x}$ is infinite. In particular, $H$ is (the upper part of) a Dedekind cut. Since $a>0$ and $M$ is Archimedean, $H$ is strictly contained in $H-\frac a2$. Hence, their difference must contain infinitely many, whence at least two elements $x<y$ belonging to $Y$. However, this leads to the contradiction $y-x<a$. Recall that a first-order structure $\mathcal M$ is said to have *definable Skolem functions*, if for every definable map $f\colon X\to Y$, there is a *definable section* $g\colon f(X)\to X$, where the latter means that $g$ is definable and $f{\circ}g$ is the identity. Assume now that $\mathcal M$ is an expansion of an ordered structure. Let $X\sub M^3$ be the subset of all $(a,b,x)$ such that either $a<x<b$, or $x<a=b$, or $b<a<x$. Note that every open interval occurs exactly once as a fiber of $X$ under the principal projection $\pi\colon M^3\to M^2$. A special case of definable Skolem functions is for this projection to have a definable section ${\beta({\cdot})}$, in which case we say that $\mathcal M$ has *definable barycenters* (with the understanding that we assign ${\beta({M})}:=c$ to be some point $c\in M$ fixed once and for all). Instead of writing ${\beta({a,b})}$, we may also write ${\beta({I})}$ where $I$ is the open interval ${{\, ]a,b[\, }}$. Any expansion of a divisible, ordered Abelian group has definable barycenters, namely let ${\beta({I})}:=(a+b)/2$ be the midpoint of the bounded open interval $I={{\, ]a,b[\, }}$, and let it be equal to respectively $a-c$ and $a+c$ if $I$ is equal to ${{\, ]-\infty,a[\, }}$ and ${{\, ]a,\infty[\, }}$ respectively, where $c$ is a fixed positive element. \[L:defskol\] If an [o-minimalistic]{} structure has definable barycenters (e.g., an expansion of an ordered group), then it has definable Skolem functions. Since a section of a definable map $f\colon X\sub M^n\to Y\sub M^m$ is also a section of the projection of its graph onto $M^m$, we may reduce to the case of a projection of a definable subset. By induction on $m$, we then easily reduce to the case that $f={\left.\pi\right|_{{X}}}$ where $\pi\colon M^{n+1}\to M^n$ and $X\sub M^{n+1}$. For ${\mathbf {a}}\in \pi(X)$, let $l\in M_\infty$ be the infimum of the fiber ${{X}_{{}}[{\mathbf {a}}]}$. If $l\in {{X}_{{}}[{\mathbf {a}}]}$, we put $g({{\mathbf {a}}}):=({\mathbf {a}},l)$. In the remaining case, $l^+$ belongs to ${{X}_{{}}[{\mathbf {a}}]}$, and hence the infimum $b$ of $(M\setminus {{X}_{{}}[{\mathbf {a}}]})_{>l}$ is a boundary point strictly bigger than $l$. In particular, $I:={{\, ]l,b[\, }}$ lies inside ${{X}_{{}}[{\mathbf {a}}]}$, whence so does ${\beta({I})}$. Putting $g({\mathbf {a}}):=({\mathbf {a}},l)$, it is now easy to verify that $g$ is a definable section of ${\left.\pi\right|_{{X}}}$. Paths ----- The following definition can be made in any ordered structure $\mathcal M$. By a *path* in $M^2$, we mean the image $\Gamma=g(I)$ of a closed interval $I={{\, [a,b]\, }}$ under a definable, continuous, injective map $g\colon I\to M^2$. We call $g(a)$ and $g(b)$ the *endpoints* (provided $a$ and/or $b$ are finite); we refer to $\Gamma$ minus its endpoints as an *open path*. A note of caution: an open path $\Gamma$ therefore is the image of an open interval ${{\, ]a,b[\, }}$ under a definable, continuous, injective map $g$ with the additional property that its *endpoints* $g(a^+)$ and $g(b^-)$ are different and do not lie on $\Gamma$. In an [o-minimalistic]{} structure, paths, open or closed, are one-dimensional, and by continuity, definably connected. By Corollary \[C:symm\], every path, being the projection of a graph, is [tame]{}. To obtain some further properties, we need to assume some additional structure: let us fix an [o-minimalistic]{} expansion $\mathcal M$ of a group. Most results are proven in the same way as in the o-minimal case (see [@vdDomin Chapt. 6, §1]). We will write ${\left|x\right|}$ for the maximum of all ${\left|x_i\right|}$ if $x={(x_1,\dots,x_{n})}$. Clearly, a path is a planar curve, and in fact it is smooth in the following sense (see Remark \[R:regular\] for the definition of regular point): \[P:pathsmooth\] In an [o-minimalistic]{} expansion of a group, any point on an open path is $1$-regular. Let $\Gamma=g(I)$ be a path, where $g\colon I\to M^2$. Recall that, after applying a translation, the origin $O\in \Gamma$ is $1$-regular if $\rho(O)$ is not a node on $\rho(\Gamma)$ for some translation given by an invertible $2\times2$-matrix over $\mathbb Q$. Write $g(t)=(g_1(t),g_2(t))$. By the Monotonicity Theorem \[T:disctu\], we may subdivide $I$ in open intervals and a discrete subset $D$ so that $g_1$ and $g_2$ are monotone on each interval. The image of these intervals is then easily seen to consist of non-nodes. So remains the points of the form $g(d)$ with $d\in D$. Checking all cases, one sees that there always exists a rotation $\rho$ such that $\rho {\circ}g$ is strictly increasing in either component on some open interval ${{\, ]u,d[\, }}$ with $u<d$. One then easily checks that $\rho(g(d))$ is not a node on $\rho(\Gamma)$. \[L:pathsel\] Given a definable subset $X$ and a point $x$ in its frontier ${{\operatorname}{fr}( X)}$, there exists an open path in $X$ with one of its endpoints equal to $x$. Let $H$ be the set of all ${\left|x-y\right|}$, for $y\in X$. Since $x\in \bar X$, the type $0^+$ belongs to $H$, so that ${{\, ]0,u[\, }}\sub H$ for some $u>0$. Hence for every $0<t<u$, there exists $y_t\in X$ such that ${\left|x-y_t\right|}=t$. By Lemma \[L:defskol\], there exists therefore a definable map $g\colon {{\, ]0,u[\, }}\to X$ such that ${\left|x-g(t)\right|}=t$, for all $t\in {{\, ]0,u[\, }}$. By Theorem \[T:disctu\], by taking $u$ sufficiently small, we may assume $g$ is continuous, so that its image is an open path. By construction, $g(0^+)=x\notin X$, proving the assertion. \[T:closbdd\] In an [o-minimalistic]{} expansion of a group, the image of a closed and bounded definable subset under a definable, continuous map is closed and bounded. We restrict our proof once more to $n=2$, so that we have a definable, continuous map $f\colon X\sub M^2\to M^2$ with $X$ bounded and closed. If $f(X)$ were unbounded, we could find, for each $t\in M$, some $x_t\in X$ with ${\left|f(x_t)\right|}>t$. By Lemma \[L:defskol\], we get a definable map $p\colon M\to X$ such that ${\left|f(p(t))\right|}>t$, for all $t\in M$. Since $X$ is bounded and closed, the Monotonicity Theorem \[T:disctu\] applied to each component of $p$ shows that $z:=p(\infty^-)$ belongs to $X$. By continuity, $f(z)=(f{\circ}p)(\infty^-)\in M^2$, which is impossible since ${\left|f(p(t))\right|}>t$, for all $t\in M$. Since $f$ is continuous, it is [tame]{} by Theorem \[T:tamemap\], whence ${\mathcal {M}^{\text{tame}}}$-definable. Therefore, so is its reverse graph $\Gamma^*(f)$, proving that the latter is a [tame]{} subset of $ M^2$. Hence, there exists a cellular map $c\colon\Gamma^*(f)\to D$ with $D$ discrete. Since $\Gamma^*(f)$ is closed, it is the union of the closures $\overline{{{c^{-1}(a)}}}$ for $a\in D$. By a similar use of Lemma \[L:pathsel\] as in the proof of [@vdDomin Chapt. 6, Lemma 1.7], each projection $\pi(\overline{{{c^{-1}(a)}}})$ is closed. Moreover, it is not hard to see that each $x\in M^2$ is contained in an open box $U$ intersecting at most three among the $\pi(\overline{{{c^{-1}(a)}}})$, and so their union, that is to say, $f(X)=\pi(\Gamma^*(f))$, is closed by Lemma \[L:closunion\] below. \[L:closunion\] In a metric space, given a locally finite collection of closed subsets, that is to say, such that each point admits an open neighborhood which intersects only finitely many of its members, then their union is again closed. Let $X$ be the union of closed subsets $X_i$ with the given property and let $x$ be a point in the closure of $X$. Let $x_n\in X$ be a sequence converging to $x$ and let $U$ be an open containing $x$ intersecting only finitely many $X_i$. In particular, one of these $X_i$ then contains a subsequence of $x_n$. Since $X_i$ is closed, $x\in X_i\sub X$. Without the boundedness assumption, Theorem \[T:closbdd\] even fails for the closure of a cell under projection: consider the graph of $1/x$ in ${\mathbb R}_{>0}$, which is closed but whose projection is open. Theorem \[T:closbdd\] has the usual corollaries (see [@vdDomin Chapt. 6, §1]): let $f\colon X\to M^n$ be a definable, continuous map with closed, bounded domain $X$, then (i) a subset of $f(X)$ is closed [if and only if]{} its preimage in $X$ is; (ii) a definable map $f(X)\to M^n$ is continuous [if and only if]{} its composition with $f$ is; (iii) if $n=1$, then $f$ attains its maximum and minimum; (iv) if $f$ is injective, then it is a homeomorphism onto its image $f(X)$ (in particular, any such map is open). The [Grothendieck ring]{} of an [o-minimalistic]{} structure {#s:gr} ============================================================ Given any first-order structure $\mathcal M$ in a language $L$, we define its *[Grothendieck ring]{}* ${{\mathbf {Gr}(\mathcal {M})}}$ as follows. Given two formulae $\varphi(x)$ and $\psi(y)$ in $L(M)$ (that is to say, allowing parameters), with $x={(x_1,\dots,x_{n})}$ and $y={(y_1,\dots,y_{m})}$, we say that $\varphi$ and $\psi$ are *$\mathcal M$-definably isomorphic* (or simply *definably isomorphic* if $\mathcal M$ is understood), if there exists a definable bijection $f\colon \varphi(\mathcal M)\to \psi(\mathcal M)$. Let ${{\mathbf {Gr}(\mathcal {M})}}$ be the quotient of the free Abelian group generated by ${\mathcal M}$-definable isomorphism classes ${{\langle \varphi\rangle}}$ of formulae $\varphi\in L(M)$ modulo the *scissor relations* $$\label{eq:sciss} {{\langle \varphi\rangle}}+ {{\langle \psi\rangle}}-{{\langle \varphi{\wedge}\psi\rangle}}-{{\langle \varphi{\vee}\psi\rangle}}\tag{sciss}$$ where $\varphi,\psi$ range over all pairs of formulae in the same free variables. See for instance [@KraEul; @KraSca] for more details. We will write ${{[ \varphi]}}$ or ${{[ Y]}}$ for the image of the formula $\varphi$, or the set $Y$ defined by it, in ${{\mathbf {Gr}(\mathcal {M})}}$. Since we can always replace a definable subset with a definable copy that is disjoint from it, the scissor relations can be simplified, by only requiring them for disjoint unions: ${{[ X\sqcup Y]}}={{[ X]}}+{{[ Y]}}$. In particular, combining all terms with a positive sign as well as all terms with a negative sign by taking disjoint unions, we see that every element in the [Grothendieck ring]{} is of the form ${{[ X]}}-{{[ Y]}}$, for some definable subsets $X$ and $Y$. To make ${{\mathbf {Gr}(\mathcal {M})}}$ into a ring, we define the product of two classes ${{[ \varphi]}}$ and ${{[ \psi]}}$ as the class of the *product* $\varphi(x){\wedge}\psi (y)$ where $x$ and $y$ are disjoint sets of variables. One checks that this is well-defined and that the class of a point is the unit for multiplication, therefore denoted $1$. Note that in terms of definable subsets, the product corresponds to the Cartesian product and the scissor relation to the usual inclusion/exclusion relation. Variants are obtained by restricting the class of formulae/definable subsets. For our purposes, we will only do this for discrete subsets. Call a formula *discrete* if it defines a discrete subset. In an [o-minimalistic]{} structure, discrete formulae are closed under Boolean combinations and products, and if two discrete definable subsets are definably isomorphic, then the graph of this isomorphism is also given by a discrete formula. Therefore, the [Grothendieck ring]{} on discrete formulae is well-defined and since definably discrete is equivalent with having dimension zero, we will denote this by ${{\mathbf {Gr}_{0}(\mathcal {M})}}$. We have a canonical [homomorphism]{} ${{\mathbf {Gr}_{0}(\mathcal {M})}}\to {{\mathbf {Gr}(\mathcal {M})}}$ with image the subring generated by classes of discrete formulae. The following is useful when dealing with [Grothendieck ring]{}[s]{}: \[L:grequal\] Two definable subsets $X$ and $Y$ in a first-order structure $\mathcal M$ have the same class in ${{\mathbf {Gr}(\mathcal {M})}}$ [if and only if]{} there exists a definable subset $Z$ such that $X\sqcup Z$ and $Y\sqcup Z$ are definably isomorphic. One direction is immediately, for if $X\sqcup Z$ and $Y\sqcup Z$ are definably isomorphic, then ${{[ X]}}+{{[ Z]}}={{[ X\sqcup Z]}}={{[ Y\sqcup Z]}}={{[ Y]}}+{{[ Z]}}$ in ${{\mathbf {Gr}(\mathcal {M})}}$, from which it follows ${{[ X]}}={{[ Y]}}$. Conversely, if ${{[ X]}}={{[ Y]}}$, then by definition of scissor relations, there exist mutually disjoint, definable subsets $A_i,B_i,C_i,D_i\sub M^{n_i}$ such that $${{\langle X\rangle}}+\sum_i{{\langle A_i\rangle}}+ {{\langle B_i\rangle}}-{{\langle A_i\sqcup B_i\rangle}}= {{\langle Y\rangle}}+\sum_i{{\langle C_i\rangle}}+ {{\langle D_i\rangle}}-{{\langle C_i\sqcup D_i\rangle}}$$ in the free Abelian group on isomorphism classes. Bringing the terms with negative signs to the other side, we get an expression in which each term on the left hand side must also occur on the right hand side, that is to say, the collection of all isomorphism classes $\{{{\langle X\rangle}}, {{\langle A_i\rangle}}, {{\langle B_i\rangle}}, {{\langle C_i\sqcup D_i\rangle}}\}$ is the same as the collection of all isomorphism classes $\{{{\langle Y\rangle}}, {{\langle C_i\rangle}}, {{\langle D_i\rangle}}, {{\langle A_i\sqcup B_i\rangle}}\}$. By properties of disjoint union, we therefore get ${{\langle X\sqcup Z\rangle}}={{\langle Y\sqcup Z\rangle}}$, where $Z$ is the disjoint union of all definable subsets $A_i,B_i,C_i,D_i$. If $\mathcal M$ is an expansion of an ordered, divisible Abelian group, then we have the following classes of open intervals. If $I={{\, ]a,b[\, }}$, then $I$ is definably isomorphic to ${{\, ]0,b-a[\, }}$ via the translation $x\mapsto x-a$. Moreover, ${{\, ]0,a[\, }}$ is definably isomorphic to ${{\, ]0,2a[\, }}$ via the map $x\mapsto 2x$. Hence the class $\mathbbm i$ of ${{\, ]2,a[\, }}$ is by equal to the sum of the classes of ${{\, ]0,a[\, }}$, $\{a\}$, and ${{\, ]a,2a[\, }}$. In other words, $\mathbbm i=2\mathbbm i+1$, whence $\mathbbm i=-1$ (the additive inverse of $1$). Let $\mathbbm h$ be the class of the unbounded interval ${{\, ]0,\infty[\, }}$. By translation and/or the involution $x\mapsto -x$, any half unbounded interval is definably isomorphic with ${{\, ]0,\infty[\, }}$. Finally, we put ${\mathbbm L}:={{[ M]}}$ (the so-called *Lefschetz class*). Since $M$ is the disjoint union of ${{\, ]-\infty,0[\, }}$, $\{0\}$, and ${{\, ]0,\infty[\, }}$, we get $$\label{eq:eulint} {\mathbbm L}=2 \mathbbm h+1.\tag{lef}$$ If $M$ is moreover an ordered field, then taking the reciprocal makes ${{\, ]0,1[\, }}$ and ${{\, ]1,\infty[\, }}$ definably isomorphic, so that $\mathbbm h=\mathbbm i=-1$, and hence also ${\mathbbm L}=-1$. Under the assumption of an underlying ordered structure, whence a topology, we can also strengthen the definition by calling two definable subset *definably homeomorphic*, if there exists a definable (continuous) homeomorphism between them, and then build the [Grothendieck ring]{}, called the *strict [Grothendieck ring]{}* of $\mathcal M$ and denoted ${{\mathbf {Gr}^{\text{s}}(\mathcal {M})}}$, on the free Abelian group generated by homeomorphism classes of definable subsets. Note that there is a canonical surjective [homomorphism]{} ${{\mathbf {Gr}^{\text{s}}(\mathcal {M})}}\to {{\mathbf {Gr}(\mathcal {M})}}$. In the o-minimal case, the monotonicity theorem implies that both variants are equal, but this might fail in the [o-minimalistic]{} case, since cell decompositions are no longer finite (but see Corollary \[C:tamestrict\] below). In fact, in the o-minimal case, the [Grothendieck ring]{} is extremely simple, as observed by Denef and Loeser (see [@vdDomin Chap. 4, §2] for more details): \[P:omingr\] The [Grothendieck ring]{} of an o-minimal expansion of an ordered field is canonically isomorphic to the ring of integers $\mathbb Z$. By the previous discussion, the class of any open interval is equal to $-1$. The graph of a function is definably isomorphic with its domain, and so the class of any $1$-cell is equal to $-1$. Since a bounded planar $2$-cell lies in between two $1$-cells, it is definably isomorphic to an open box, and by definition of the multiplication in ${{\mathbf {Gr}(\mathcal {M})}}$, therefore its class is equal to ${\mathbbm L}^2=1$. The unbounded case is analogous, and so is the case that the $2$-cell lies in a higher Cartesian product. This argument easily extends to show that the class of a $d$-cell in ${{\mathbf {Gr}(\mathcal {M})}}$ is equal to ${\mathbbm L}^d=(-1)^d$. By Cell Decomposition, every definable subset is a finite union of cells, and hence its class in ${{\mathbf {Gr}(\mathcal {M})}}$ is an integer (multiple of $1$). We denote the canonical [homomorphism]{} ${{\mathbf {Gr}(\mathcal {M})}}\to {\mathbb Z}$ by ${\chi_{\mathcal M}(\cdot)}$ and call it the *Euler [characteristic]{}* of $\mathcal M$. Inspired by [@CluEd], we define the *Euler measure* of a definable subset $X$ in an o-minimal structure $\mathcal M$ the pair ${\mu_{\mathcal M}(X)}:=({\operatorname}{dim}(X),{\chi_{\mathcal M}(X)})\in ({\mathbb N}\cup\{-\infty\})\times{\mathbb Z}$, where we view the latter set in its lexicographical ordering. In an arbitrary first-order structure, let us say, for definable subsets $X$ and $Y$, that $ X\preceq Y$ [if and only if]{} there exists a definable injection $X\to Y$. In general, this relation, even up to definable isomorphism, will fail to be symmetric (take for instance in the reals the sets $X={{\, [0,1]\, }}$ and $Y=X\cup\{3/2\}$, where $x\mapsto x/2$ sends $Y$ inside $X$), and therefore is in general only a partial pre-order. As we will discuss below in §\[s:disc\], it does induce a partial order on isomorphism classes of discrete, definable subsets in an [o-minimalistic]{} structure, and we can even make it total by extending the class of isomorphisms (Theorem \[T:totorddisc\]). In the o-minimal case, $\preceq$ is a total pre-order by the following (folklore) result. In some sense, the rest of the paper is an attempt to extend this result to the [o-minimalistic]{} case. \[T:Euldim\] In an o-minimal expansion of an ordered field, two definable sets $X$ and $Y$ are definably isomorphic [if and only if]{} ${\mu_{\mathcal M}(X)}={\mu_{\mathcal M}(Y)}$. Moreover, $X\preceq Y$ [if and only if]{} ${\operatorname}{dim}(X)\leq {\operatorname}{dim}(Y)$ with the additional condition that ${\chi_{\mathcal M}(X)}\leq {\chi_{\mathcal M}(Y)}$ whenever both are zero-dimensional. The first statement is proven in [@vdDomin Chap. 8, 2.11]. So, suppose $X\preceq Y$. Since $X$ is definably isomorphic with a subset of $Y$, its dimension is at most that of $Y$. If both are zero-dimensional, that is to say, finite, then the pigeonhole principle gives ${\chi_{\mathcal M}(X)}={\left|X\right|}\leq{\left|Y\right|} ={\chi_{\mathcal M}(Y)}$. Conversely, assume ${\operatorname}{dim}(X)\leq {\operatorname}{dim}(Y)$. If both are finite, the assertion is clear by the same argument, so assume they are both positive dimensional. Without loss of generality, by adding a cell of the correct dimension, we may then assume that they have both the same dimension $d\geq 1$. Let $e:= {\chi_{\mathcal M}(Y)}-{\chi_{\mathcal M}(X)}$ and let $F$ consist of $e$ points disjoint from $X$ if $e$ is positive and of $-e$ open intervals disjoint from $X$ if $e$ is negative. Since ${\chi_{\mathcal {}}(F)}=e$, the Euler measure of $X\sqcup F$ and $Y$ are the same, and hence they are definably isomorphic by the first assertion, from which it follows that $X\preceq Y$. Let ${\mathcal M}$ be an ultra-o-minimal structure, say, realized as the ultraproduct of o-minimal structures $\mathcal M_i$. We define the *ultra-Euler [characteristic]{}* ${\chi_{\mathcal { M}}(\cdot)}$ as follows. Let $Y\sub M^n$ be a definable subset, say given by a formula $\varphi(x,{{\mathbf {b}}})$ with $ {{\mathbf {b}}}$ a tuple of parameters realized as the ultraproduct of tuples ${\mathbf {b}}_i$ in each $M_i$. Let $Y_i:=\varphi(\mathcal M_i,{\mathbf {b}}_i)$, so that $ Y$ is the ultraproduct of the $Y_i$, and let ${\chi_{\mathcal { M}}( Y)}$ now be the ultraproduct of the ${\chi_{\mathcal M_i}(Y_i)}$, viewed as an element of ${{\mathbb Z}_\natural}$. If $ X$ is definably isomorphic with $ Y$, via a definable bijection with graph $ G$, choose as above definable subsets $X_i$ and $G_i$ in $\mathcal M_i$ with ultraproduct equal to $ X$ and $ G$ respectively. By [Łoś’ Theorem]{}, almost each $G_i$ is the graph of a definable bijection between $X_i$ and $Y_i$, and therefore ${\chi_{\mathcal M_i}(X_i)}={\chi_{\mathcal M_i}(Y_i)}$ for almost all $i$, showing that ${\chi_{\mathcal { M}}( X)}={\chi_{\mathcal { M}}( Y)}$. Similarly, we define the *ultra-Euler measure* ${\mu_{\mathcal M}(X)}:=({\operatorname}{dim}( X),{\chi_{\mathcal M}(X)})$. Since the ultra-Euler [characteristic]{} is easily seen to be also compatible with the scissor relations , we showed (compare with Theorem \[T:Drank\] below): \[C:ulomingrot\] For an ultra-o-minimal structure $ {\mathcal M}$, we have a canonical [homomorphism]{} ${{\mathbf {Gr}(\mathcal {{ M}})}}\to {{\mathbb Z}_\natural}$. The Discrete Pigeonhole Principle --------------------------------- Before we proceed, we identify another [o-minimalistic]{} property, that is to say, a first-order property of o-minimal structures: \[P:DPP\] Given an [o-minimalistic]{} structure $\mathcal M$, if a definable map $f\colon Y\to Y$, for some $Y\sub M^n$, is injective and its image is co-discrete, meaning that $Y\setminus f(Y)$ is discrete, then it is a bijection. In particular, any definable map from a discrete subset $D$ to itself is injective [if and only if]{} it is surjective. For each formula $\varphi(x,y,{\mathbf {z}})$, we can express in a first-order way that if $\varphi(x,y,{\mathbf {c}})$, for some tuple ${\mathbf {c}}$ of parameters, defines the graph of an injective map $f\colon Y\to Y$ then $$\label{eq:DPP} \text{$Y\setminus f(Y)$ discrete implies $Y=f(Y)$.}\tag{DPP}$$ Remains to show that holds in any o-minimal structure $\mathcal M$. Indeed, if $D=Y\setminus f(Y)$, then ${\chi_{\mathcal M}(Y)}={\chi_{\mathcal M}(f(Y))}+{\chi_{\mathcal M}(D)}$. Since $f$ is injective, $Y$ and $f(Y)$ are definably isomorphic, whence have the same Euler [characteristic]{}, and so ${\chi_{\mathcal M}(D)}=0$. But a discrete subset in an o-minimal structure is finite and its Euler [characteristic]{} is then just its cardinality, showing that $D=\emptyset$. One direction in the last assertion is immediate, and for the converse, assume $f\colon D\to D$ is surjective. For each $x\in D$, define $g(x)$ as the (lexicographical) minimum of ${{f^{-1}(x)}}$, so that $g\colon D\to D$ is an injective map, whence surjective by the above, and therefore necessarily the inverse of $f$. At present, I do not know how to derive from ${{\text{DCTC}}}$. \[C:ominicgr\] An [o-minimalistic]{}  structure is o-minimal [if and only if]{} its [Grothendieck ring]{} is isomorphic to $\mathbb Z$. If $\mathcal M$ is [o-minimalistic]{}, but not o-minimal, then there exists at least one definable, discrete, infinite set $D$. By assumption, ${{[ D]}}=n$ for some integer $n$. After removing $n$ points, if $n$ is positive, or adding $-n$ points, if negative, we may suppose ${{[ D]}}=0$. By Lemma \[L:grequal\], there exists a definable subset $X$ such that $X$ and $X\sqcup D$ are definably isomorphic. By , this forces $D=\emptyset$, contradiction. \[C:discmon\] A monotone map $f\colon D\to D$ on a definable, discrete subset $D$ in an [o-minimalistic]{} structure $\mathcal M$ is either constant or an involution. Suppose $f$ is non-constant and hence $f^2$ is strictly increasing. So upon replacing $f$ by its square, we may already assume that $f$ is increasing, and we need to show that it is then the identity. Since $f$ is injective, it is bijective by Proposition \[P:DPP\]. Let $h$ be the maximum of $D$, and suppose $f(d)=h$. If $d<h$, then $h=f(d)<f(h)\in D $, contradiction, showing that $f(h)=h$. If $f$ is not the identity, then the set $Q$ of all $d\in D$ for which $f(d)\neq d$ is non-empty, whence has a maximum, say, $u<h$. In particular, if $v:=\sigma_D(u)$ is its immediate successor, then $f(u)<f(v)=v$, since $v\notin Q$, whence $f(u)<u$, since $u\in Q$. Since $u=f(a)$ for some $a\neq u$, then either $a<u$ or $v\leq a$, and hence $u=f(a)<f(u)<u$ or $v=f(v)\leq f(a)=u$, a contradiction either way. \[R:discmon\] Note that the map sending $h$ to the minimum of $D$, and equal to $\sigma_D$ otherwise is a definable permutation of $D$, but it obviously fails to be monotone. The map $x\mapsto {{\omega_\natural}}-x$ on $D=({{\mathbb N}_\natural})_{\leq {{\omega_\natural}}}$ as in Example \[E:ulomin\] is a strictly decreasing involution. It is not hard to see that if an involution exists, it must be unique: indeed, if $f$ and $g$ are both decreasing, let $a$ be the maximal element at which they disagree (it cannot be $h$ since $f(h)=l=g(h)$), and assume $f(a)<g(a)$. Since $f(\sigma(a))=g(\sigma(a))<f(a)<g(a)$, it is now easy to see that $f(a)$ does not lie in the image of $g$, contradicting that $g$ must be a bijection by . \[P:onevar\] In an [o-minimalistic]{} expansion $\mathcal M$ of an ordered field, there exists for every definable subset $Y\sub M$, two definable, discrete subsets $D, E\sub Y$ such that ${{[ Y]}}={{[ D]}}-{{[ E]}}$ in ${{\mathbf {Gr}(\mathcal {M})}}$. Since the boundary $\partial Y$ is discrete, we may remove it and assume $Y$ is open, whence a disjoint union of open intervals by Theorem \[T:onevar\]. Let us introduce some notation that will be useful later too, assuming $Y$ is open. For $y\in Y$, let $l(y)$ and $h(y)$ be respectively the maximum of $(\partial Y)_{<y}$ and the minimum of $(\partial Y)_{>y}$ (allowing $\pm\infty$). Hence ${{\, ]l(y),h(y)[\, }}$ is the maximal interval in $Y$ containing $y$, and we denote its midpoint by $m(y):={\beta({{{\, ]l(y),h(y)[\, }}})}$. Let $L(Y)$, $M(Y)$ and $R(Y)$ consist respectively of all $y$ less than, equal to, or greater than $m(y)$. Removing a maximal unbounded interval from $Y$ if necessary (whose class is equal to $-1$ as already observed above), we may assume $Y$ is bounded, so that $l(y)$ and $h(y)$ are always finite. Since the maps $f_Y\colon L(Y)\to Y\colon y\mapsto 2y-l(y)$ and $g_Y\colon L(Y)\to R(Y)\colon y\mapsto y+m(y)$ are bijections, ${{[ Y]}}={{[ L(Y)]}}={{[ R(Y)]}}$. Since the scissor relations yield ${{[ Y]}}={{[ L(Y)]}}+{{[ M(Y)]}}+{{[ R(Y)]}}$, we get ${{[ Y]}}=-{{[ M(Y)]}}$. By construction $M(Y)$ is discrete, and so we are done. The proof gives the following more general result: given any discrete subset $D_0\sub Y$, we can find disjoint discrete subsets $D,E\sub Y$ such that $D_0\sub D$ and ${{[ Y]}}={{[ D]}}-{{[ E]}}$. Indeed, let $D:=D_0\cup (Y\cap \partial Y)$ and $E:=M(Y\setminus D)$. If $\mathcal M$ merely expands an ordered group, then we have to also include the class $\mathbbm h$, that is to say, in that case we can write ${{[ Y]}}=e\mathbbm h+{{[ D]}}-{{[ E]}}$, where $e\in\{0,1,2\}$ is the number of unbounded sides of $Y$. For higher arities, we need to make a [tame]{}[ness]{} assumption: \[C:defcell\] Let $X\sub M^2$ be a definable subset in an [o-minimalistic]{} expansion $\mathcal M$ of an ordered field. If $\partial X$ is [tame]{}, then there exist definable, discrete subsets $D,E\sub X$ such that ${{[ X]}}={{[ D]}}-{{[ E]}}$ in ${{\mathbf {Gr}(\mathcal {M})}}$. In fact, the class of any [tame]{} subset in ${{\mathbf {Gr}(\mathcal {M})}}$ is of the form ${{[ D]}}-{{[ E]}}$, for some definable discrete subsets $D,E\sub M$. There is nothing to show if $X$ is discrete. Assume next that it has dimension one. Let $V:={{\operatorname}{Vert}( X)}$ be the vertical component of $X$. Since $\pi(V)$ is discrete, as we argued before, we can carry out the argument in the proof of Proposition \[P:onevar\] on each fiber separately to write ${{[ V]}}$ as the difference of two discrete classes (we leave the details to the reader, but compare with the two-dimensional case below). Removing $V$ from $X$, we may assume $X$ has no vertical components. In particular, the set $N:={{\operatorname}{Node}( X)}$ of nodes of $X$ is discrete by Proposition \[P:discnodes\]. Removing it, we may assume $X$ has no nodes, so that every point lies on a unique optimal quasi-cell by Corollary \[C:supercell\]. However, by assumption, $X$ is [tame]{}, and hence there exists a cellular map $c\colon X\to D$. Given $x\in X$, let $I_x$ be the domain $\pi({{c^{-1}(c(x))}})$ of the unique cell ${{c^{-1}(c(x))}}$ containing $x$. Let $L(X)$, $M(X)$, and $R(X)$ consist respectively of all $x\in X$ such that $\pi(x)$ lies in $L(I_x)$, $M(I_x)$, and $R(I_x)$ respectively (in the notation of the proof of Proposition \[P:onevar\]). Define $f_X\colon L(X)\to X$ and $g_X\colon L(X)\to R(X)$ by sending $x$ to the unique point on ${{c^{-1}(c(x))}}$ lying above respectively $f_{I_x}(\pi(x))$ and $g_{I_x}(\pi(x))$, showing that $X$, $L(X)$, and $R(X)$ are definably isomorphic. Since $M(X)$ is discrete and ${{[ X]}}={{[ L(X)]}}+{{[ R(X)]}}+{{[ M(X)]}}$, we are done in this case. If $X$ has dimension two, its boundary has dimension at most one, and so we have already dealt with it by the previous case. Upon removing it, we may assume $X$ is open. This time, we let $L(X)$, $M(X)$ and $R(X)$ be the union of respectively all $L({{X}_{{}}[a]})$, $M({{X}_{{}}[a]})$, and $R({{X}_{{}}[a]})$, for all $a\in\pi(X)$. The maps $(a,b)\mapsto f_{{{X}_{{}}[a]}}(b)$ and $(a,b)\mapsto g_{{{X}_{{}}[a]}}(b)$ put $L(X)$ in definable bijection with respectively $X$ and $R(X)$ (with an obvious adjustment left to the reader if the fiber ${{X}_{{}}[a]}$ is unbounded), and hence ${{[ X]}}=-{{[ M(X)]}}$. Since $M(X)$ has dimension at most one by Proposition \[P:nondisc2\], we are done by induction. Without providing the details, we can extend this argument to higher dimensions, proving the last claim, where we also must use that definable discrete subsets are univalent in an ordered field. \[R:defcell\] Inspecting the above proof, we actually proved the following: if $c\colon X\to D$ is a cellular surjective map, then $$\label{eq:eulerdisc} {{[ X]}}=\sum_{e=0}^d (-1)^e{{[ D_e]}}$$ where $D_e=c(X^{(e)})$ consist of all $a\in D$ with $e$-dimensional fiber ${{c^{-1}(a)}}$, and where $d$ is the dimension of $X$. We may reduce to the case that all fibers have the same dimension, and the assertion is then clear in the one-dimensional case, since the restriction of $c$ to $M(X)$ is a bijection. Repeating the argument therefore to $X$, we get ${{[ X]}}=-{{[ M(X)]}}={{[ M(M(X))]}}$, and now $M(M(X))$ is definably isomorphic with $D$ via $c$. Higher dimensions follow similarly by induction. In particular, if $\mathcal M$ is a [tame]{} expansion of an ordered field, then its [Grothendieck ring]{} is generated by the definable discrete subsets of $M$, and the canonical [homomorphism]{} ${{\mathbf {Gr}_{0}(\mathcal {M})}}\to {{\mathbf {Gr}(\mathcal {M})}}$ is surjective. Inspecting the above proof, we see that all isomorphisms involved are in fact homeomorphisms, and so the result also holds in the strict [Grothendieck ring]{} ${{\mathbf {Gr}^{\text{s}}(\mathcal {M})}}$. Since any function with discrete domain is continuous, we showed: \[C:tamestrict\] For a [tame]{}, [o-minimalistic]{} expansion of an ordered field, its [Grothendieck ring]{} and its strict [Grothendieck ring]{} coincide. The partial order on ${{\mathfrak D(\mathcal M)}}$ {#s:disc} -------------------------------------------------- Let ${{\mathfrak D(\mathcal M)}}$ denote the collection of isomorphism classes of definable, discrete subsets in an [o-minimalistic]{} structure $\mathcal M$. Recall that $ X\preceq Y$ if there exists a definable injection $X\to Y$. Theorem \[T:Euldim\] suggests that this relation is not very useful in higher dimensions, so we study it only on ${{\mathfrak D(\mathcal M)}}$. Assume $D$ and $E$ are discrete, definable subsets with $ D\preceq E$ and $ E\preceq D$. Hence there are definable injections $D\to E$ and $E\to D$. By Proposition \[P:DPP\], both compositions are bijections, showing that $D$ and $E$ are definably isomorphic. Since transitivity is trivial, we showed that we get a partial order on ${{\mathfrak D(\mathcal M)}}$. To obtain a partial order on the zero-dimensional [Grothendieck ring]{} ${{\mathbf {Gr}_{0}(\mathcal {M})}}$, we define ${{[ D]}}\preceq {{[ E]}}$, if there exists a definable, discrete subset $A$ such that $D\sqcup A\preceq E\sqcup A$. To show that this well-defined, assume ${{[ D]}}={{[ D']}}$ and ${{[ E]}}={{[ E']}}$. By Lemma \[L:grequal\], there exist definable, discrete subsets $F$ and $G$ such that $D\sqcup F{\cong}D'\sqcup F$ and $E\sqcup G{\cong}E'\sqcup G$. Since $D\sqcup A\preceq E\sqcup A$, we get $$D'\sqcup F\sqcup G\sqcup A{\cong}D\sqcup F\sqcup G\sqcup A\preceq E\sqcup F\sqcup G\sqcup A{\cong}E\sqcup F\sqcup G\sqcup A$$ as required. We then extend this to a partial ordering on ${{\mathbf {Gr}_{0}(\mathcal {M})}}$ by linearity. In the o-minimal case, ${{\mathbf {Gr}_{0}(\mathcal {M})}}$ is just ${\mathbb Z}$ in its natural ordering, but we will give some examples where the order is not total (though, see Theorem \[T:totorddisc\] below). Let us first prove a comparison result in a special case. In an expansion of an ordered group, we call a definable, discrete set $D$ *equidistant*, if the map $a\mapsto \sigma_D(a)-a$ is constant on all non-maximal elements of $D$, where $\sigma_D$ is the successor function. In other words, $D$ is equidistant, if for any non-maximal $a\in D$, also $a+\rho\in D$, where $\rho$ is the spread of $D$. \[P:equidist\] In an [o-minimalistic]{} expansion $\mathcal M$ of an ordered field, any two definable equidistant subset[s]{} of $M$ are comparable. Let $D,E\sub M$ be definable equidistant subsets. Since they are bounded by , we may assume after a translation that both have minimum equal to $0$, and then after taking a scaling, that both have spread $1$. Let $m$ be the maximum of (the non-empty) $D\cap E$. If $m$ is non-maximal in either set, then $m+1$ lies both in $D$ and in $E$ by assumption, contradiction. Hence $m$ is the maximum, say, of $D$, and therefore $D\sub E$, whence $ D\preceq E$. More generally, given a definable, discrete subset $D\sub M$ in an [o-minimalistic]{} expansion $\mathcal M$ of an ordered field, define the *derivative* $D'$ of $D$ as the set of all differences $\sigma_D(a)-a$, where $a$ runs over all non-maximal elements of $D$. Hence an equidistant set is one whose derivative is a singleton. Since we have a surjective map $D\setminus\{\max D\}\to D'\colon a\mapsto \sigma_D(a)-a$, it follows from the next lemma that $D'\preceq D$. \[L:imageorder\] In an [o-minimalistic]{} structure $\mathcal M$, if $g\colon X\to M^k$ is a definable map, then $ {g(D)}\preceq D$, for every discrete, definable subset $D\sub X$. This follows by considering the injective map $ g(D)\to D$ sending $a$ to the minimum of ${{g^{-1}(a)}}$. In general, this partial order will not be total. Since $D\preceq E$ implies ${{[ D]}}\preceq {{[ E]}}$, but not necessarily the converse, the former being total implies that the latter is too, but again, the converse is not clear. To construct examples, let us introduce the following notation. \[E:discoverspill\] Given a sequence ${\mathbf {a}}=(a_n)$ of real numbers, let ${{{\mathbb R}_\natural}{\langle {\mathbf {a}}\rangle}}$ be the ultraproduct of the $\mathcal R_n$, where each $\mathcal R_n$ is the expansion of the real field with a unary predicate $\tt D$ interpreting the first $n$ elements $a_1,\dots,a_n$ in the sequence. Since each $\mathcal R_n$ is o-minimal, ${{{\mathbb R}_\natural}{\langle {\mathbf {a}}\rangle}}$ is [o-minimalistic]{}. Moreover, ${{\mathbf {a}}}$ is the “finite” part of the set $D_{{\mathbf {a}}}:=\tt D({{{\mathbb R}_\natural}{\langle {\mathbf {a}}\rangle}})$ defined by $\tt D$, that is to say, $$D_{{\mathbf {a}}}\cap {\mathbb R}=\{a_1,a_2,\dots,\}.$$ so that we refer to ${{{\mathbb R}_\natural}{\langle {\mathbf {a}}\rangle}}$ as the structure obtained from ${\mathbf {a}}$ by *discrete overspill* (for a related construction, see also §\[s:Taylor\] below). In this notation, Example \[E:ulomin\] is the discrete overspill ${{{\mathbb R}_\natural}{\langle {\mathbb N}\rangle}}$ of ${\mathbb N}$ listed in its natural order. I do not know whether $\preceq$ is total on it. Any countable subset can be enumerated, including $\mathbb Q$, although this enumeration might not be order preserving. Nonetheless, we get a structure ${{{\mathbb R}_\natural}{\langle {\mathbf {q}}\rangle}}$ with $D_{{\mathbf {q}}}\cap {\mathbb R}=\mathbb Q$ (the non-standard elements of $D_{{\mathbf {q}}}$ form a proper subset of ${\mathbb Q_\natural}$ and are harder to describe as they depend on the choice of enumeration). We can repeat this construction with more than one sequence, taking one unary predicate for each. Any structure obtained by discrete overspill is [tame]{} by Remark \[R:tame\]. \[E:nontotal\] Now, if we take two unary predicates, representing, say, the sequence of prime numbers ${\mathbf {p}}$ and the sequence of powers of two ${\mathbf {t}}$, then in ${{{\mathbb R}_\natural}{\langle {\mathbf {p}},{\mathbf {t}}\rangle}}$, it seems very unlikely that the discrete sets $D_{{\mathbf {p}}}$ and $D_{{\mathbf {t}}}$ are comparable. For if they were, they would have to be definably isomorphic as they have the same ultra-Euler [characteristic]{} (equal to ${{\omega_\natural}}$, the ultraproduct of the diagonal sequence $(n)_n$). In fact, if, instead, ${\mathbf {p}}$ enumerates a computable set in ${\mathbb N}$ and ${\mathbf {t}}$ a non-computable one, then such a definable isomorphism restricted to ${\mathbb N}$ would be a computable isomorphism between these two sets, which is of course impossible. It is easy to combine these two unary sets into a single one, by letting $a_{2n}:=p_n$ and $a_{2n-1}:=-t_n$, so that then $D_{{\mathbf {a}}}\cap ({{\mathbb R}_\natural})_{\leq 0}=D_{{\mathbf {t}}}$ and $D_{{\mathbf {a}}}\cap ({{\mathbb R}_\natural})_{\geq 0}=D_{{\mathbf {p}}}$, giving an example of a single discrete overspill ${{{\mathbb R}_\natural}{\langle {\mathbf {a}}\rangle}}$ with non-total order. Nonetheless, as we will see in Theorem \[T:totorddisc\] below, the failure is due to a missing isomorphism, and so, in an appropriate expansion (no longer computable of course), both sets become definably isomorphic. \[T:ominul\] An ultra-o-minimal structure ${\mathcal M_\natural}$, given as the ultraproduct of o-minimal structures $\mathcal M_i$, is o-minimal, if for each formula $\varphi$ without parameters, there exists an $N_\varphi\in{\mathbb N}$ such that ${\left|{\chi_{\mathcal M_i}(\varphi)}\right|}\leq N_\varphi$. Let ${Y_\natural}\sub {M_\natural}$ be definable, say, given as the fiber of a $\emptyset$-definable subset ${X_\natural}\sub {M_\natural}^{1+n}$ over a tuple ${{\mathbf {b}}_\natural}$. Let $X_i\sub M_i^{1+n}$ be the corresponding $\emptyset$-definable subset, and choose ${\mathbf {b}}_i$ in $M_i$ with ultraproduct ${{\mathbf {b}}_\natural}$, so that ${Y_\natural}$ is the ultraproduct of the $Y_i:={{X_i}_{{}}[{\mathbf {b}}_i]}$. By the proof of Theorem \[T:plane\] (which in the o-minimal case does yield a finite cell decomposition), we can decompose each $X_i$ as a disjoint union of $\emptyset$-definable subsets $X_i^{(e)}$ consisting of the union of all $e$-cells in a cell decomposition of $X_i$. In fact, this proof can be carried out in the theory ${{\text{DCTC}}}$, so that it holds in any [o-minimalistic]{} structure $\mathcal M$ uniformly. For instance, assuming $n=2$ and $X=\varphi(\mathcal M)$, then $X^{(2)}$ consists exactly of all interior points that do not lie on a vertical fiber containing some node of $\partial X$, whereas $X^{(0)}$ consists of all nodes of $\partial X$ that belong to $X$, and $X^{(1)}$ of all remaining points. And so we can find formulae $\varphi^{(e)}$ that define in each model $\mathcal M\models{{\text{DCTC}}}$ the sets $X^{(e)}$, for $e\leq n+1$. Now, since each $X_i^{(e)}$ is a disjoint union of $e$-cells, its Euler [characteristic]{} is equal to $(-1)^eN_{i,e}$, where $N_{i,e}$ is the number of $e$-cells in the decomposition. By assumption (applied to the formula $\varphi^{(e)}$), this Euler [characteristic]{} is bounded in absolute value, whence so are the $N_{i,e}$, that is to say, there exist $N_e\in{\mathbb N}$ such that $N_{i,e}<N_e$ for all $i$. But then the fiber ${{X_i^{(e)}}_{{}}[{\mathbf {b}}_i]}$ admits a decomposition in at most $N_e$ cells. Since the union of the latter for all $e$ is just $Y_i$, we showed that there is a uniform bound on the number of cells, that is to say, open intervals and points, in a decomposition of $Y_i$. Since this is now first-order expressible, ${Y_\natural}$ too is a finite union of intervals, as we needed to show. Expansions of [o-minimalistic]{} structures {#s:ominexp} =========================================== Since an expansion by definable sets does not alter the collection of definable sets, we immediately have: \[L:expdef\] If $\mathcal M$ is [o-minimalistic]{} and $X\sub M^n$ is definable, then $(\mathcal M,X)$ is again [o-minimalistic]{}. So we ask in more generality, what properties does a subset of an [o-minimalistic]{} structure need to have in order for the expansion to be again [o-minimalistic]{}? Let us call such a subset *[o-minimalistic]{}* (or, more correctly, *$\mathcal M$-[o-minimalistic]{}* as this depends on the surrounding structure), where we just proved that definable subsets are. \[C:imagomin\] The image of an [o-minimalistic]{} subset under a definable map is again [o-minimalistic]{}, and so is its complement, its closure, its boundary, and its interior. More generally, any set definable from an [o-minimalistic]{} set is again [o-minimalistic]{}. It suffices to prove the last assertion. Let $X$ be an [o-minimalistic]{} subset of an [o-minimalistic]{} structure $\mathcal M$. Since $(\mathcal M,X)$ is [o-minimalistic]{}, any set definable in $(\mathcal M,X)$ is [o-minimalistic]{}  (in the expansion, whence also in the reduct) by Lemma \[L:expdef\]. To define a weaker isomorphism relation, we introduce the following notation. Let $X$ be a definable subset in a structure $\mathcal M$, say, defined by the formula (with parameters) $\varphi$, that is to say, $X=\varphi(\mathcal M)$. If $\mathcal N$ is an elementary extension of $\mathcal N$, then we set ${{X}^{{\mathcal N}}}:=\varphi(\mathcal N)$, and call it the *definitional extension* of $X$ in $\mathcal N$. Let us call two definable subsets $X$ and $Y$ of an [o-minimalistic]{} structure $\mathcal M$ *[o-minimalistic]{}[ally]{} isomorphic*, denoted $X\equiv Y$, if their definitional extensions have the same ultra-Euler measure in every ultra-o-minimal elementary extension $\mathcal M\preceq \mathcal N$, that is to say, if ${\mu_{\mathcal N}({{X}^{{\mathcal N}}})}={\mu_{\mathcal N}({{Y}^{{\mathcal N}}})}$. It is easy to see that this constitutes an equivalence relation on definable subsets. \[P:ominiso\] In an [o-minimalistic]{} expansion $\mathcal M$ of an ordered field, if two definable subsets $X$ and $Y$ are [o-minimalistic]{}[ally]{} isomorphic, then there exists an [o-minimalistic]{} expansion of $\mathcal M$ in which they become definably isomorphic. Suppose $X$ and $Y$ are [o-minimalistic]{}[ally]{} definable, and let $\mathcal N$ be some ultra-o-minimal elementary extension of $\mathcal M$, given as the ultraproduct of o-minimal structures $\mathcal N_i$. Let $X_i$ and $Y_i$ be $\mathcal N_i$-definable subsets with respective ultraproducts ${{X}^{{\mathcal N}}}$ and ${{Y}^{{\mathcal N}}}$. Since by Proposition \[P:dimulomin\] dimension is definable, ${{X}^{{\mathcal N}}}$ and ${{Y}^{{\mathcal N}}}$ have the same dimension, whence so do almost each $X_i$ and $Y_i$ by [Łoś’ Theorem]{}. By assumption, they have also the same Euler [characteristic]{} for almost all $i$, so that they are definably isomorphic by Theorem \[T:Euldim\]. Hence, there exists for almost all $i$, a definable isomorphism $f_i\colon X_i\to Y_i$. Let ${\Gamma_\natural}$ be the ultraproduct of the graphs $\Gamma(f_i)$, so that $(\mathcal N,{\Gamma_\natural})$, whence also $(\mathcal M,\Gamma)$, is [o-minimalistic]{}, where $\Gamma$ is the restriction of ${\Gamma_\natural}$ to $\mathcal M$. Moreover, by [Łoś’ Theorem]{}, ${\Gamma_\natural}$ is the graph of a bijection ${{X}^{{\mathcal N}}}\to {{Y}^{{\mathcal N}}}$, and hence its restriction $\Gamma$ is the graph of a bijection $X\to Y$, proving that $X$ and $Y$ are definably isomorphic in $(\mathcal M,\Gamma)$. I do not know whether the converse is also true: if $X$ and $Y$ are definably isomorphic in some [o-minimalistic]{} expansion $\mathcal M'$, are they [o-minimalistic]{}[ally]{} isomorphic? They will have the same Euler [characteristic]{} in any (reduct of an) ultra-o-minimal elementary extension of $\mathcal M'$ by essentially the same argument, but what about ultra-o-minimal elementary extensions of $\mathcal M$ that are not such reducts? A related question is in case $\mathcal M$ itself is already ultra-o-minimal, if two sets have the same Euler [characteristic]{}, do their definitional extensions also have the same Euler [characteristic]{} in an ultra-o-minimal elementary extension? This would follow if Euler [characteristic]{} was definable, but at the moment, we can only prove a weaker version (see Theorem \[T:defrk\]). Before we address these issues, we prove a result yielding non-trivial examples of [o-minimalistic]{}[ally]{} isomorphic sets that need not be definably isomorphic. \[C:ominiso\] In an [o-minimalistic]{} expansion $\mathcal M$ of an ordered field, if two definable subsets $X$ and $Y$ have the same dimension and the same class in ${{\mathbf {Gr}(\mathcal {M})}}$, then they are [o-minimalistic]{}[ally]{} isomorphic. By Lemma \[L:grequal\], there exists a definable subset $Z$ such that $X\sqcup Z$ and $Y\sqcup Z$ are definably isomorphic. Let $\mathcal M\preceq\mathcal N$ be an ultra-o-minimal elementary extension. Hence ${{X}^{{\mathcal N}}}\sqcup {{Z}^{{\mathcal N}}}$ and ${{Y}^{{\mathcal N}}}\sqcup {{Z}^{{\mathcal N}}}$ are definably isomorphic, and therefore $${\chi_{\mathcal N}({{X}^{{\mathcal N}}})} +{\chi_{\mathcal N}( {{Z}^{{\mathcal N}}})}={\chi_{\mathcal N}({{X}^{{\mathcal N}}}\sqcup {{Z}^{{\mathcal N}}})}={\chi_{\mathcal N}({{Y}^{{\mathcal N}}}\sqcup {{Z}^{{\mathcal N}}})}={\chi_{\mathcal N}({{Y}^{{\mathcal N}}})} +{\chi_{\mathcal N}( {{Z}^{{\mathcal N}}})}$$ showing that ${{X}^{{\mathcal N}}}$ and ${{Y}^{{\mathcal N}}}$ have the same ultra-Euler [characteristic]{}, as we needed to show. Contexts and [virtual isomorphism]{}[s]{} ----------------------------------------- To overcome the difficulties alluded to above, we must make our definitions context-dependable in the following sense. Given an [o-minimalistic]{} structure $\mathcal M$, by a *context* for $\mathcal M$, we mean an ultra-o-minimal structure $\mathcal N$ that contains $\mathcal M$ as an elementary substructure (which always exists by Corollary \[C:ulomin\]). An expansion $\mathcal M'$ of $\mathcal M$ is then called *permissible* (with respect to the context $\mathcal N$), if $\mathcal N$ can be expanded to a context $\mathcal N'$, that is to say, $\mathcal M'\preceq\mathcal N'$ and $\mathcal N'$ is again ultra-o-minimal. If $\mathcal M$ itself is ultra-o-minimal, then we may take it as its own context, but even in this case, not every expansion will be permissible, as it may fail to be an ultraproduct. From now on, we fix an [o-minimalistic]{} structure $\mathcal M$ and a context $\mathcal N$. We define a (context-dependable) *Euler [characteristic]{}* ${\chi_{\mathcal M}(\cdot)}$ (or, simply $\chi$) on $\mathcal M$ by restricting the ultra-Euler [characteristic]{} of $\mathcal N$, that is to say, by setting ${\chi_{\mathcal {}}(X)}:={\chi_{\mathcal N}({{X}^{{\mathcal N}}})}$, for any definable subset of $\mathcal M$, and we define similarly its *Euler measure* ${\mu_{\mathcal {}}(X)}:=({\operatorname}{dim}(X),{\chi_{\mathcal {}}(X)})$. We say that two definable subsets are *[virtually isomorphic]{}*, if there exists a permissible expansion of $\mathcal M$ in which they become definably isomorphic. In particular, two definable subsets that are [o-minimalistic]{}[ally]{} isomorphic are also [virtually isomorphic]{}, but the converse is unclear. We can now prove an [o-minimalistic]{} analogue of Theorem \[T:Euldim\]. \[T:viso\] In an [o-minimalistic]{} expansion $\mathcal M$ of an ordered field, two definable subsets are [virtually isomorphic]{} [if and only if]{} they have the same Euler measure. One direction is proven in the same way as Proposition \[P:ominiso\], so assume $X$ and $Y$ are [virtually isomorphic]{} definable subsets. By assumption, $\mathcal M\preceq\mathcal N$ expands into [o-minimalistic]{} structures $\mathcal M'\preceq \mathcal N'$, with $\mathcal N'$ again ultra-o-minimal, such that $X$ and $Y$ are $\mathcal M'$-definably isomorphic. Let $\mathcal N'$ be the ultraproduct of o-minimal structures $\mathcal N_i'$. Since ${{X}^{{\mathcal N'}}}$ and ${{Y}^{{\mathcal N'}}}$ are definably isomorphic, so are almost all $X_i$ and $Y_i$, where $X_i$ and $Y_i$ are $\mathcal N_i'$-definable subsets with respective ultraproducts ${{X}^{{\mathcal N'}}}$ and ${{Y}^{{\mathcal N'}}}$. In particular, $X_i$ and $Y_i$ have the same Euler measure for almost all $i$, by Theorem \[T:Euldim\]. Hence ${{X}^{{\mathcal N'}}}$ and ${{Y}^{{\mathcal N'}}}$ have the same ultra-Euler measure, by Proposition \[P:dimulomin\]. Since both invariants remain the same in the reduct $\mathcal N$, elementarity then yields ${\mu_{\mathcal {}}(X)}={\mu_{\mathcal {}}(Y)}$. \[R:totorddisc\] Any two subsets given by discrete overspill with respect to non-repeating sequences (see Example \[E:discoverspill\]) are [virtually isomorphic]{}, since they both have Euler [characteristic]{} ${{\omega_\natural}}$. O-finitism ---------- In the terminology of [@ForTame], the definable discrete sets in an [o-minimalistic]{} structure are exactly the *pseudo-finite* sets. As we already mentioned in the introduction, in the [o-minimalistic]{} context, discrete sets play the role of finite sets, and so we briefly discuss the first-order aspects of this assertion. Given a (non-empty) collection of $L$-structures $\mathfrak K$, and a subset $X\sub M^n$ in some $L$-structure $\mathcal M$, we say that $X$ is *$\mathfrak K$-finitistic*, if $(\mathcal M, X)$ satisfies every $L(\tt U)$-sentence $\sigma$ which holds in every expansion $(\mathcal K,F)$ of a structure $\mathcal K\in\mathfrak K$ by a finite set $F\sub K^n$. In case $\mathfrak K$ is the collection of o-minimal structures, we call $X\sub M^n$ *[o-finitistic]{}*. Applying the definition just to $L$-sentences $\sigma$ (not containing the predicate $\tt U$, so that $(\mathcal K,F)\models\sigma$ [if and only if]{}  $\mathcal K\models\sigma$), we see that $\mathcal M$ is then necessarily [o-minimalistic]{}. Put differently, an [o-finitistic]{} set in an [o-minimalistic]{} structure is a model of *o-finitism*, that is to say, of the theory of a finite set in an o-minimal structure. By Propositions \[P:Dedlim\], \[P:locdefdisc\], and \[P:DPP\], we have: \[C:ofindisc\] In an [o-minimalistic]{} structure, an [o-finitistic]{} set is discrete, closed, bounded, and locally definable, every non-empty intersection with an open interval has a maximum and a minimum, and every injective, definable self-map on it is an isomorphism. It seems unlikely that these properties characterize fully o-finitism. A complete axiomatization of o-finitism would be of interest in view of the following results. \[T:ofinexp\] A subset $X$ of an [o-minimalistic]{} structure $\mathcal M$ is [o-finitistic]{} [if and only if]{} it is discrete and [o-minimalistic]{}. In particular, any definable, discrete subset in an [o-minimalistic]{} structure is [o-finitistic]{}. Assume first that $X$ is [o-finitistic]{}, whence discrete by Corollary \[C:ofindisc\]. We have to show that given an $L(\tt U)$-sentence $\sigma$ holding true in every o-minimal $L(\tt U)$-structure, then $(\mathcal M,X)\models\sigma$. Let $\mathcal K$ be an o-minimal structure and let $F\sub K^n$ be a finite subset. Hence $(\mathcal K,F)$ is also o-minimal and therefore satisfies $\sigma$. Since this holds for all such expansions, $\sigma$ is true in $(\mathcal M,X)$ by o-finitism, as we needed to show. Conversely, suppose $X\sub M^n$ is discrete and [o-minimalistic]{}, that is to say, $(\mathcal M,X)$ is [o-minimalistic]{}. To show that $X$ is [o-finitistic]{}, let $\sigma$ be a sentence true in every expansion $(\mathcal K,F)$ of an o-minimal structure $\mathcal K$ by a finite subset $F\sub K^n$. Consider the disjunction $\sigma'$ of $\sigma$ with the sentence expressing that the set defined by $\tt U$ is not discrete. Hence $\sigma'$ is true in any o-minimal expansion $(\mathcal K,Y)$. Since $X$ is [o-minimalistic]{}, this means that $(\mathcal M,X)\models \sigma'$, and since $X$ is discrete, this in turn implies that $\sigma$ is true in $(\mathcal M,X)$, as we needed to show. The last assertion then follows from Lemma \[L:expdef\]. Let us call a subset of an ultra-o-minimal structure *ultra-finite*, if it is the ultraproduct of finite subsets (such a set may fail to be definable, since the definition in each component may not be uniform). An ultra-finite set is [o-finitistic]{}. As for the converse, if an [o-finitistic]{} set is an ultraproduct (a so-called *internal* set), then it must be [o-finitistic]{}, but what for external sets? In any case, we have: \[T:ulominic\] A subset $X\sub M^k$ of an [o-minimalistic]{} structure $\mathcal M$ is [o-finitistic]{} [if and only if]{} there exists an elementary extension $\mathcal M\preceq \mathcal N$ with $\mathcal N$ ultra-o-minimal and an ultra-finite subset $Y\sub N^k$, such that $X=Y\cap M^k$. Suppose $\mathcal N$ and $Y$ have the stated properties, and let $\mathcal N_i$ be o-minimal structures and $Y_i\sub N_i^k$ finite subsets, so that $\mathcal N$ and $Y$ are their respective ultraproducts. Since $(\mathcal N_i,Y_i)$ is again o-minimal, their ultraproduct $(\mathcal N,Y)$ is [o-minimalistic]{}. Since $(\mathcal M,X)$ is then an elementary substructure, the latter is also [o-minimalistic]{}. Moreover, since $Y$ is discrete, so must $X$ be, and hence $X$ is [o-finitistic]{} by Theorem \[T:ofinexp\]. Conversely, by the same theorem, if $X$ is [o-finitistic]{}, then $(\mathcal M,X)$ is [o-minimalistic]{}. Hence there exists an elementary extension $(\mathcal N,Y)$ which is ultra-o-minimal as an $L(\tt U)$-structure by Corollary \[C:ulomin\]. Write $(\mathcal N,Y)$ as an ultraproduct of o-minimal structures $(\mathcal N_i,Y_i)$. Since $X$ is discrete, so must $Y$ be by elementarity, whence so are almost all $Y_i$ by [Łoś’ Theorem]{}. The latter means that almost all are in fact finite, showing that $Y$ is ultra-finite, and the assertion follows since $X=Y\cap M^k$. Next, we give a criterion for a subset $Y\sub M$ to be [o-minimalistic]{}. By Theorem \[T:onevar\], its boundary $\partial Y$ should be discrete, and $Y^\circ=Y\setminus\partial Y$ should be a disjoint union of open intervals. Given an arbitrary set $Y\sub M$, define its *enhanced boundary* ${\DeltaY}$ as the set consisting of the following pairs: $(d,0)$ if $d\in Y$, $(d,1)$ if $d^+$ belongs to $Y$, and $(d,-1)$ if $d^-$ belongs to $Y$, where $d$ runs over all boundary points of $Y$. An enhanced boundary cannot have fibers with three points and its projection is the ordinary boundary $\partial Y$. If $Y$ is [o-minimalistic]{}, then ${\DeltaY}$ must satisfy some extra conditions: it must be bounded, discrete and closed, and, by , if $(d,1)$ belongs to it, then so must $(d',-1)$, where $d'$ is the immediate successor of $d$ in $\partial Y$. \[T:enhbd\] A one-variable subset $Y$ in an [o-minimalistic]{} structure $\mathcal M$ is [o-minimalistic]{} [if and only if]{} its enhanced boundary ${\DeltaY}$ is [o-finitistic]{} and its interior is a disjoint union of open intervals. Suppose $Y$ is [o-minimalistic]{}, so that $Y^\circ$ is a disjoint union of open intervals. Since ${\DeltaY}$ is definable from $Y$, it too is [o-minimalistic]{} by Corollary \[C:imagomin\], whence [o-finitistic]{} by Theorem \[T:ofinexp\]. To prove the converse, let $D:=\partial Y=\pi({\DeltaY})$, a bounded, closed, discrete set, and let $l$ be its minimum. Define $X\sub M$ as the set of all $x\in M$ such that one of the following three conditions holds 1. \[i:incase\] $x\in D$; 2. \[i:leftcase\] $x>l$ and $(d,1)\in{\DeltaY}$, where $d=\max D_{<x}$; 3. \[i:unbcase\] $x<l$ and $(l,-1)\in{\DeltaY}$. Since $X$ is definable from ${\DeltaY}$, it is [o-minimalistic]{} by Corollary \[C:imagomin\]. Remains to show that $X=Y$. It follows from that $X\cap \partial Y=Y\cap \partial Y$, so that it suffices to show that $X^\circ=Y^\circ$. Therefore, we may as well assume from the start that $Y$ is open. Write $Y=\sqcup_n I_n$ as a disjoint union of open intervals, and let ${{\, ]a,b[\, }}$ one of the $I_n$ (we leave the unbounded case to the reader, for which one needs ). In particular, $a\in D$ and $a^+$ belongs to $Y$, so that $(a,1)\in{\DeltaY}$. By , the entire interval ${{\, ]a,b[\, }}$ lies in $ X$, whence so does the whole of $Y$. Conversely, if $x\in X$, let $d:=\max D_{<x}$, so that $(d,1)\in{\DeltaY}$. Hence $d^+$ belongs to $Y$, and so $d$ must be an endpoint of one of the $I_n$. The other endpoint must be bigger than $d$, and hence bigger than $x$, showing that $x\in I_n\sub Y$. The virtual [Grothendieck ring]{} --------------------------------- We fix again an [o-minimalistic]{} structure $\mathcal M$ and a context $\mathcal N$. We can use [virtual isomorphism]{}[s]{} instead of definable isomorphisms in the definition of the zero-dimensional or the full [Grothendieck ring]{}, that is to say, the quotient modulo the scissor relations of the free Abelian group on [virtual isomorphism]{}  classes of respectively all discrete, definable subsets, and of all definable subsets yield the *virtual [Grothendieck ring]{}[s]{}* ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$ and ${{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}}$ respectively. We have surjective [homomorphism]{}[s]{} ${{\mathbf {Gr}_{0}(\mathcal {M})}}\to {{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$ and ${{\mathbf {Gr}(\mathcal {M})}}\to {{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}}$. \[C:omingr\] Given an [o-minimalistic]{} expansion $\mathcal M$ of an ordered field, there exist embeddings ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}{}\sub {{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}}{\hookrightarrow}{{\mathbb Z}_\natural}$, where ${{\mathbb Z}_\natural}$ is the ring of non-standard integers in the given context. Since the Euler [characteristic]{} vanishes on any scissor relation, it induces by Theorem \[T:viso\] a [homomorphism]{} $\chi\colon {{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}}\to {{\mathbb Z}_\natural}$. By the same result, its restriction to ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$ is injective. To see that $\chi$ is everywhere injective, assume ${\chi_{\mathcal {}}(X)}={\chi_{\mathcal {}}(Y)}$ for some definable subsets $X$ and $Y$. If they have the same dimension, then they are [virtually isomorphic]{}, again by Theorem \[T:viso\]. So assume $X$ has dimension $d\geq 1$ and $Y$ has lesser dimension. Let $U$ be the difference of a $d$-dimensional box minus a $(d-1)$-dimensional subbox, so that in particular ${{[ U]}}$ vanishes, whence also ${\chi_{\mathcal {}}(U)}$. As $X$ and $Y\sqcup U$ now have the same Euler measure, they are [virtually isomorphic]{} by Theorem \[T:viso\], and hence ${{[ X]}}={{[ Y]}}+{{[ U]}}={{[ Y]}}$ in ${{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}}$, as we needed to show. The injectivity of ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}\to {{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}}$ is then also clear. In particular, if $\mathcal M$ is moreover [tame]{}, then we have an equality of virtual [Grothendieck ring]{}[s]{} ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}= {{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}}$ by Corollary \[C:defcell\]. We can also extend the partial order on ${{\mathfrak D(\mathcal M)}}$ to a total order on ${{\mathfrak D^{\text{viso}}(\mathcal M)}}$, the set of [virtual isomorphism]{} classes of definable, discrete subsets. First, given definable subsets $X$ and $Y$, we say that $X\leq Y$ (or, if we want to emphasize the context, $X\leq_{\mathcal N}Y$), if $X\preceq_{\mathcal M'}Y$ in some permissible [o-minimalistic]{} expansion $\mathcal M'$ of $\mathcal M$. Clearly, if $X\preceq Y$, then $X\leq Y$. The following two results are the [o-minimalistic]{} analogues of Theorem \[T:Euldim\]. \[T:totorddisc\] In an [o-minimalistic]{} structure $\mathcal M$, two definable, discrete subsets $F$ and $G$ satisfy $F\leq G$ [if and only if]{}  ${\chi_{\mathcal {}}( F)}\leq {\chi_{\mathcal {}}( G)}$. In particular, $\leq$ is a total order on ${{\mathfrak D^{\text{viso}}(\mathcal M)}}$. Suppose first that ${\chi_{\mathcal {}}( F)}\leq {\chi_{\mathcal {}}( G)}$. Write $\mathcal N$ as the ultraproduct of o-minimal structures $\mathcal N_i$, and let $F_i$ and $G_i$ be finite sets with respective ultraproducts the definitional extensions ${{F}^{{\mathcal N}}}$ and ${{G}^{{\mathcal N}}}$ of $F$ and $G$ respectively. Since ${\chi_{\mathcal N}({{F}^{{\mathcal N}}})}\leq {\chi_{\mathcal N}({{G}^{{\mathcal N}}})}$, the cardinality of $F_i$ is at most that of $G_i$, for almost all $i$. In particular, there exists an injective map $F_i\to G_i$ for almost all $i$. Let ${\Gamma_\natural}$ be the ultraproduct of the graphs of these maps $F_i\to G_i$. Hence ${\Gamma_\natural}$ is ultra-finite and therefore its restriction $\Gamma$ to $\mathcal M$ is [o-finitistic]{} by Theorem \[T:ulominic\], whence [o-minimalistic]{} by Theorem \[T:ofinexp\]. By [Łoś’ Theorem]{} and elementarity, $\Gamma$ is the graph of an injective map $F\to G$, showing that $F\preceq_{(\mathcal M,\Gamma)} G$. Since $(\mathcal M,\Gamma)$ is permissible, $F\leq G$. The converse goes along the same lines: suppose $F\preceq_{\mathcal M'}G$, for some permissible [o-minimalistic]{} expansion $\mathcal M'$ of $\mathcal M$. By definition, there is an ultra-o-minimal expansion $\mathcal N'$ of $\mathcal N$ with $\mathcal M'\preceq \mathcal N'$. Since ${{F}^{{\mathcal N}}}\preceq_{\mathcal N'}{{G}^{{\mathcal N}}}$, we have ${\chi_{\mathcal {}}(F)}={\chi_{\mathcal N'}({{F}^{{\mathcal N}}})}\leq {\chi_{\mathcal N'}({{G}^{{\mathcal N}}})}={\chi_{\mathcal {}}(G)}$. \[P:preorddim\] In an [o-minimalistic]{} expansion $\mathcal M$ of an ordered field, we have $X\leq Y$ [if and only if]{} ${\operatorname}{dim}(X)\leq {\operatorname}{dim}(Y)$, for $X$ and $Y$ definable subsets with ${\operatorname}{dim}(Y)>0$. The direct implication is clear. For the converse, by definability of dimension, we may pass to the context of $\mathcal M$ and therefore already assume $\mathcal M$ is ultra-o-minimal, given as the ultraproduct of o-minimal structures $\mathcal M_i$. Let $X_i$ and $Y_i$ be definable subsets in $\mathcal M_i$ with respective ultraproducts $X$ and $Y$. By [Łoś’ Theorem]{}, ${\operatorname}{dim}(X_i)\leq {\operatorname}{dim}(Y_i)$, and hence $X_i\preceq Y_i$, by Theorem \[T:Euldim\], for almost all $i$. Let $f_i\colon X_i\to Y_i$ be a definable injection and let ${\Gamma_\natural}$ be the ultraproduct of the graphs $\Gamma(f_i)$. Since each $(\mathcal M_i,\Gamma(f_i))$ is again o-minimal, $(\mathcal M,{\Gamma_\natural})$ is ultra-o-minimal and hence in particular a permissible expansion. Since ${\Gamma_\natural}$ is the graph of an injective map by [Łoś’ Theorem]{}, $X\preceq_{(\mathcal M,{\Gamma_\natural})}Y$, as we needed to show. In particular, any definable, discrete subset is virtually univalent. \[C:OPP\]\[Virtual Pigeonhole Principle\] Given an [o-minimalistic]{}  structure $\mathcal M$, two definable, discrete subsets $D$ and $E$ are [virtually isomorphic]{} [if and only if]{}, for some definable subset $X$, the sets $D\sqcup X$ and $E\sqcup X$ are [virtually isomorphic]{}, [if and only if]{} ${{[ D]}}={{[ E]}}$ in ${{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}}$. One direction in the first equivalence is immediate, so assume $D\sqcup X$ and $E\sqcup X$ are [virtually isomorphic]{}. Passing to a permissible [o-minimalistic]{} expansion, we may assume that they are already definably isomorphic, say, by an isomorphism $f\colon D\sqcup X\to E\sqcup X$. By totality (Theorem \[T:totorddisc\]), we may assume that $E\leq D$, and hence after taking another permissible [o-minimalistic]{} expansion, and replacing $E$ with an isomorphic image, we may even assume that $E\sub D$. Therefore, the composition of $f$ and the inclusion $E\sqcup X\sub D\sqcup X$ is a map with co-discrete image, and hence is surjective by . However, this can only be the case if $E=D$, as we needed to show. The last equivalence is now just Lemma \[L:grequal\]. \[C:ordomingr\] For an [o-minimalistic]{} structure $\mathcal M$, its zero-dimensional, virtual [Grothendieck ring]{} ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$ is an ordered ring with respect to $\leq$. Every element in ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$ is of the form ${{[ A]}}-{{[ B]}}$, for some definable, discrete subsets $A$ and $B$ in the [o-minimalistic]{} structure $\mathcal M$. Therefore, for definable, discrete subsets $A_i$ and $B_i$, with $i=1,2$, we set ${{[ A_1]}}-{{[ B_1]}}\leq {{[ A_2]}}-{{[ B_2]}}$ [if and only if]{}  $$\label{eq:ord} A_1\sqcup B_2\leq A_2\sqcup B_1.$$ To see that this is well-defined, suppose ${{[ A_i]}}-{{[ B_i]}}={{[ A'_i]}}-{{[ B'_i]}}$, for $i=1,2$ and definable, discrete subsets $A'_i$ and $B'_i$. Therefore, ${{[ A_i\sqcup B_i']}}={{[ A_i'\sqcup B_i]}}$, whence $A_i\sqcup B_i'$ and $A_i'\sqcup B_i$ are [virtually isomorphic]{} by Corollary \[C:OPP\]. We have to show that assuming , the same inequality holds for the accented sets. Taking the disjoint union with $B_1'\sqcup B_2'$ on both sides of , yields inequalities $$\begin{aligned} (A_1\sqcup B_1')\sqcup B_2\sqcup B_2'&\leq (A_2\sqcup B_2')\sqcup B_1\sqcup B_1'\\ (A_1'\sqcup B_1)\sqcup B_2\sqcup B_2'&\leq (A_2'\sqcup B_2)\sqcup B_1\sqcup B_1'\\ (A_1'\sqcup B'_2)\sqcup (B_1\sqcup B_2)&\leq (A_2'\sqcup B_1')\sqcup (B_1\sqcup B_2)\end{aligned}$$ which by another application of Corollary \[C:OPP\] then gives $A_1'\sqcup B'_2\leq A_2'\sqcup B_1'$, as we needed to show. It is now easy to check that $\leq$ makes ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$ into a totally ordered ring. \[C:ofincut\] Every [o-finitistic]{} subset defines a cut in ${{\mathfrak D^{\text{viso}}(\mathcal M)}}$. In particular, we can put a total pre-order on the collection of [o-finitistic]{} subsets. Let $F$ be an [o-finitistic]{} subset of an [o-minimalistic]{} structure $\mathcal M$ and let $D\in{{\mathfrak D^{\text{viso}}(\mathcal M)}}$ be arbitrary. Since $(\mathcal M,F)$ is [o-minimalistic]{} by Theorem \[T:ofinexp\], we can compare $D$ and $F$ in ${{\mathfrak D(\mathcal M,F)}}$ by Theorem \[T:totorddisc\]. If $G$ is another [o-finitistic]{} subset, then we set $F\leq G$ [if and only if]{} the lower cut in ${{\mathfrak D(\mathcal M)}}$ determined by $F$ is contained in the lower cut of $G$. A note of caution: even if $F\leq G$ and $G\leq F$, for $F$ and $G$ [o-finitistic]{} subsets, they need not be [virtually isomorphic]{}. For instance, taking $D$ as in Example \[E:ulomin\], it is an [o-finitistic]{} subset of ${{\mathbb R}_\natural}$, and since ${{\mathfrak D(\mathcal {{{\mathbb R}_\natural}})}}$ is just ${\mathbb N}$ by o-minimality, its cut is $\infty$. However, $D\setminus\{{{\omega_\natural}}\}$ determines the same cut, whence $D\leq D\setminus\{{{\omega_\natural}}\}\leq D$, but we know that they cannot be definably isomorphic in any [o-minimalistic]{} expansion by . In fact, it is not clear whether two given [o-finitistic]{} subsets live in a common [o-minimalistic]{} expansion, and therefore can be compared directly. This is also why we cannot (yet?) define a [Grothendieck ring]{}  on [o-finitistic]{} subsets. Discretely valued Euler [characteristic]{}[s]{} ----------------------------------------------- In order to calculate the zero-dimensional virtual [Grothendieck ring]{}, we introduce a new type of Euler [characteristic]{}. Fix an [o-minimalistic]{} structure $\mathcal M$ and a context $\mathcal N$, and let $D$ be a definable, discrete subset. In this section, we will always view $D$ in its lexicographical order $\leq_{\text{lex}}$ (or, when there is no risk for confusion, simply denoted $\leq$). \[C:defsub\] Any definable subset of a definable, discrete subset $D$ in an [o-minimalistic]{} structure $\mathcal M$ is [virtually isomorphic]{} to an initial segment $D_{\leq a}$. The set of initial segments is a maximal chain in ${{\mathfrak D^{\text{viso}}(\mathcal M)}}$, since any two consecutive subsets in this chain differ by a single point. Hence, any definable subset $E\sub D$ must be a member of this chain up to [virtual isomorphism]{}. Clearly, such an $a$ must be unique, and so, given a non-empty definable subset $E\sub D$, we let ${\chi_{D}(E)}$ be the unique $a$ such that $E$ is [virtually isomorphic]{} with $D_{\leq a}$. We add a new symbol $\varnothing$ to $D$ and set ${\chi_{D}(\emptyset)}:=\varnothing$. For definable subsets $E_1,E_2\sub D$, we have $E_1\leq E_2$ [if and only if]{} ${\chi_{D}(E_1)}\leq {\chi_{D}(E_2)}$. Given a definable map $g$ with domain $D$, we can define by Lemma \[L:imageorder\] its *rank* as ${{\operatorname}{rk}_{}(g)}:={\chi_{D}(g(D))}$. A map is constant [if and only if]{} its rank is minimal (that is to say, equal to the minimum of its domain). By , we immediately have: \[C:defdiscmaprk\] In an [o-minimalistic]{} structure, a definable map with discrete domain is injective [if and only if]{} its rank is maximal (that is to say, equal to the maximum of its domain). Let $D\sub M$ be definable and discrete, with minimal element $l$ and maximal element $h$. For each $n$, we view the Cartesian power $D^n$ as a definable subset of $D^{n+1}$ via the map ${\mathbf {a}}\mapsto (l,{\mathbf {a}})$. We also need to take into consideration the empty set, and so we define $\varnothing$ to be lower than any element in any $D^n$, and we let $D^\infty$ be the direct limit of the ordered sets $D^n\cup \{\varnothing\}$. Under this identification, the elements of $D^n\cup \{\varnothing\}$ form an initial segment in $D^{n+1}\cup \{\varnothing\}$ with respect to the lexicographical ordering. In particular, if $E\sub D^n$ is a non-empty definable subset, then ${\chi_{D^n}(E)}={\chi_{D^{n+1}}(E')}$, where $E'$ is the image of $E$ in $D^{n+1}$. After identification therefore, we will view ${\chi_{D^n}(E)}$ simply as an element of $D^\infty$, and we just denote it ${\chi_{D}(E)}$. More generally, given an arbitrary definable subset $X\sub M^n$, we define its *$D$-valued Euler [characteristic]{}* (or, simply *Euler [characteristic]{}*) ${\chi_{D}(X)}:={\chi_{D^n}(X\cap D^n)}$. We define an addition and a multiplication on $D^\infty$ as follows. First, let us define the *disjoint union* $A\sqcup B$ of two definable subsets $A,B\sub M^n$ as the definable subset in $M^{n+1}$ consisting of all $(a,l)$ and $(b,h)$ with $a\in A$ and $b\in B$. For $a\in D^\infty$, we set $a\oplus\varnothing=\varnothing\oplus a=a$ and $a\otimes\varnothing=\varnothing\otimes a=\varnothing$. For the general case, assume $a,b\in D^n$, and let $a\oplus b$ be the Euler [characteristic]{} of the disjoint union $(D^n)_{\leq a}\sqcup(D^n)_{\leq b}\sub D^{n+1}$, and let $a\otimes b$ be the Euler [characteristic]{} of the Cartesian product $(D^n)_{\leq a}\times (D^n)_{\leq b}\sub D^{2n}$. One verifies that both operations are independent of the choice of $n$, making $D^\infty$ into a commutative semi-ring, where the zero for $\oplus$ is $\varnothing$, and where the unit for $\otimes$ is $l$, the minimum of $D$. We even can define a subtraction: if $a\leq b$ in $D^\infty$, then we define $b\ominus a$ as the Euler [characteristic]{} of $D^n\cap {{\, ]a,b]\, }}$, where $n$ is sufficiently large so that $a,b\in D^n$. This allows us to define the Grothendieck group generated by $(D^\infty, \oplus)$, defined as all pairs $(x,y)$ with $x,y\in D^\infty$ up to the equivalence $(x,y)\sim (x',y')$ [if and only if]{} $x\oplus y'=x'\oplus y$; the induced commutative ring will be denoted ${{\mathfrak Z(D)}}$, and called the ring of *$D$-integers*. To turn this into a genuine Euler [characteristic]{}, recall the construction of the *induced structure* ${\mathcal {D}_{\text{ind}}}$ on a subset $D\sub M$ of a first-order structure: for each definable subset $X\sub M^n$, we have a predicate defining in ${\mathcal {D}_{\text{ind}}}$ the subset $M\cap D^n$. If $\mathcal M$ is an ordered structure, then so is ${\mathcal {D}_{\text{ind}}}$. If $D$ is definable, then we have an induced [homomorphism]{} of [Grothendieck ring]{}[s]{} ${{\mathbf {Gr}(\mathcal {{\mathcal {D}_{\text{ind}}}})}}\to {{\mathbf {Gr}_{0}(\mathcal {M})}}$. If instead of definable isomorphism, we take [virtual isomorphism]{}, we get the virtual variant ${{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {{\mathcal {D}_{\text{ind}}}})}}\to {{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$. By the Virtual Pigeonhole Principle (Corollary \[C:OPP\]), this latter [homomorphism]{} is injective. To discuss when they are isomorphic, let us call $D$ *power dominant*, if for every definable, discrete subset $A$, there is some $n$ such that $A\leq D^n$. \[P:powdom\] In an [o-minimalistic]{} structure $\mathcal M$, a definable, discrete subset $D\sub M$ is power dominant [if and only if]{} $ {{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {{\mathcal {D}_{\text{ind}}}})}}{\cong}{{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$. Suppose first that $D$ is power dominant and let $A$ be an arbitrary definable, discrete subset. By assumption, there exists an $n$ and a definable subset $B\sub D^n$, such that $A$ is [virtually isomorphic]{} with $B$. Hence ${{[ A]}}={{[ B]}}$ in ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$, proving that it lies in the image of ${{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {{\mathcal {D}_{\text{ind}}}})}}\to {{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$. Conversely, assume that the latter map is surjective, and let $A$ be an arbitrary definable, discrete subset. Hence, there exists an $n$ and definable subsets $E,F\sub D^n$ such that ${{[ A]}}={{[ E]}}-{{[ F]}}$ in ${{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}$. By the Virtual Pigeonhole Principle (Corollary \[C:OPP\]), this means that there is a [virtual isomorphism]{} $A\sqcup F\to E$. Hence the composition $A\sub A\sqcup F\to E\sub D^n$, shows that $A\leq D^n$. To study the existence of power dominant sets, let us say, for $D$ and $E$ discrete, definable subsets, that $D\lll E$, if $D^n\leq E$ for all $n$. If neither $D\lll E$ nor $E\lll D$, then $D$ and $E$ are mutually power bounded, that is to say, there exist $m$ and $n$ such that $D\leq E^m$ and $E\leq D^m$, and we write $D\approx E$. Hence $\lll$ induces a total order relation on the set ${\text{Arch}^{\text{pow}}(\mathcal {M})}$ of $\approx$-classes of definable, discrete subsets of $\mathcal M$. The class of the empty set is the minimal element of ${\text{Arch}^{\text{pow}}(\mathcal {M})}$, the class of a singleton is the next smallest element, and the class of a two-element set is the next (and consists of all finite sets). For an o-minimal structure, these are the only three classes, whereas for a proper [o-minimalistic]{} structure, there must be at least one more class, of some infinite set. I do not know whether ${\text{Arch}^{\text{pow}}(\mathcal {M})}$ is always discretely ordered. In any case, it follows easily from the definitions that a class is maximal in ${\text{Arch}^{\text{pow}}(\mathcal {M})}$ [if and only if]{} it is the class of a power dominant set. Thus, the existence of a power dominant set corresponds to ${\text{Arch}^{\text{pow}}(\mathcal {M})}$ having a maximal element, which is especially interesting in view of Proposition \[P:powdom\] and its applications below. I conjecture that $D$ as in Example \[E:ulomin\] is power dominant (and a similar property for any set obtained by discrete overspill). This would follow from the following growth conjecture in an o-minimal $L$-expansion $\mathcal R$ of ${\mathbb R}$: does there exist, for every formula $\varphi$ in the language $L(\tt U)$, some $n\in{\mathbb N}$, such that for any finite subset $F$, the set $\varphi(\mathcal R,F)$ defined by interpreting the unary predicate $\tt U$ by $F$, if finite, has cardinality at most ${\left|F\right|}^n$. Likewise, I conjecture that the following always produces a power dominant set: let $\mathcal M$ be o-minimal and let $D$ be [o-finitistic]{}, then $D$ is power dominant in the ([o-minimalistic]{}) expansion $(\mathcal M,D)$. \[T:Drank\] In an [o-minimalistic]{} structure $\mathcal M$, every definable, discrete subset $D\sub M$ induces a ring isomorphism ${{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {{\mathcal {D}_{\text{ind}}}})}}{\cong}{{\mathfrak Z(D)}}$ by sending the class of a definable subset to its $D$-valued Euler [characteristic]{}. We already observed that the ring operations on ${{\mathfrak Z(D)}}$ are invariant under [virtual isomorphism]{}. It is now easy to see that they also respect the scissor relations  in the [Grothendieck ring]{} of ${\mathcal {D}_{\text{ind}}}$. Surjectivity follows since every element in ${{\mathfrak Z(D)}}$ is of the form $a\ominus b$ for some $n$ and some $a,b\in D^n$, and hence is the image of ${{[ (D^n)_{\leq a}]}}-{{[ (D^n)_{\leq b}]}}$. To calculate the kernel, we can write a general element as ${{[ E]}}-{{[ F]}}$, with $E,F$ definable subsets in ${\mathcal {D}_{\text{ind}}}$. Such an element lies in the kernel if ${\chi_{D}(E)}={\chi_{D}(F)}$, which means that $E$ and $F$ are [virtually isomorphic]{}, whence ${{[ E]}}={{[ F]}}$ in ${{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {{\mathcal {D}_{\text{ind}}}})}}$. Summarizing, we have the following diagram of [homomorphism]{}[s]{} among the various [Grothendieck ring]{}[s]{}, for $\mathcal M$ an [o-minimalistic]{} expansion of an ordered field: $$\label{eq:gr} \xymatrix{ {{\mathbf {Gr}(\mathcal {{\mathcal {D}_{\text{ind}}}})}}\ar@{->>}[r]\ar[d]&{{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {{\mathcal {D}_{\text{ind}}}})}}\ar^(.6){\sim}[r]\ar@{^{(}->}^i[d]&{{\mathfrak Z(D)}} \\ {{\mathbf {Gr}_{0}(\mathcal {M})}}\ar@{->>}[r]\ar[d]&{{\mathbf {Gr}^{\text{virt}}_{0}(\mathcal {M})}}\ar@{^{(}->}^j[d]\\ {{\mathbf {Gr}(\mathcal {M})}}\ar@{->>}[r]&{{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}} }$$ with $i$ an isomorphism if $D$ is power dominant by Proposition \[P:powdom\], and with $j$ an isomorphism if $\mathcal M$ is [tame]{}, by Corollary \[C:defcell\], that is to say, we proved: \[C:gromintame\] If $\mathcal M$ is a [tame]{}, [o-minimalistic]{} expansion of an ordered field admitting a definable, power dominant subset $D$, then its [o-minimalistic]{} [Grothendieck ring]{} ${{\mathbf {Gr}^{\text{virt}}_{}(\mathcal {M})}}$ is isomorphic to the ring of $D$-integers ${{\mathfrak Z(D)}}$. If we would allow classes of [o-finitistic]{}  subsets in ${\text{Arch}^{\text{pow}}(\mathcal {M})}$, then there never is a maximal element: let $D$ be any definable, discrete subset (or even any [o-finitistic]{} subset). Take an ultra-o-minimal elementary extension $\mathcal N$, and choose $D_i\sub N_i$ such that their ultraproduct is ${{D}^{{\mathcal N}}}$. Assuming univalence, let $A_i\sub N_i$ be isomorphic with $D_i^i$ and let ${A_\natural}\sub N$ be their ultraproduct. By Theorem \[T:ulominic\], the restriction ${A_\natural}\cap M$ is [o-finitistic]{} and satisfies by [Łoś’ Theorem]{} $D^n\leq A$ for all $n$, that is to say, $D\lll A$. \[T:defrk\] Let $D\sub M$ be a definable, discrete subset of an [o-minimalistic]{} structure $\mathcal M$, and let $X\sub M^{n+k}$ be any definable subset. For each $e\in D^n$, the set of parameters ${\mathbf {a}}\in M^n$ such that ${\chi_{D}({{X}_{{}}[{\mathbf {a}}]})}= e$ is [o-minimalistic]{}. If ${\mathbf {a}}$ does not belong to $D^n$, then the fiber ${{X}_{{}}[{\mathbf {a}}]}$ is empty, whence has Euler [characteristic]{} $\varnothing$. As these ${\mathbf {a}}$ form a definable subset, we may therefore replace $X$ by $X\cap D^{n+k}$ and assume already that $X$ is a definable subset of $D^n$. Let $\mathcal N$ be the context and write it as the ultraproduct of o-minimal structures $\mathcal N_i$. Choose $D_i\sub N_i$, $e_i\in D_i^n$ and $X_i\sub D_i^{n+k}$ with respective ultraproducts ${{D}^{{\mathcal N}}}$, $e$, and ${{X}^{{\mathcal N}}}$. For each $i$, let $F_i\sub N_i^n$ be the (finite) set of parameters for which the fiber has the same cardinality as $(D_i^n)_{\leq e_i}$. Hence, for each ${\mathbf {a}}\in F_i$, there exists a bijection $f_{{\mathbf {a}}}\colon {{X_i}_{{}}[{\mathbf {a}}]}\to (D_i^n)_{\leq e_i}$. Let $H_i\sub N_i^{3n}$ be the union of all $\{{\mathbf {a}}\}\times \Gamma(f_{{\mathbf {a}}})$, where ${\mathbf {a}}$ runs over all tuples in $F_i$. Let ${F_\natural}\sub N^n$ and ${H_\natural}\sub N^{3n}$ be their ultraproduct, so that both sets are ultra-finite. By [Łoś’ Theorem]{}, for each ${\mathbf {a}}\in {F_\natural}$, the fiber ${{{H_\natural}}_{{}}[{\mathbf {a}}]}$ is the graph of a bijection ${{{{X}^{{\mathcal N}}}}_{{}}[{\mathbf {a}}]}\to \left(({{D}^{{\mathcal N}}})^n\right)_{\leq e}$. Therefore, $F:={F_\natural}\cap M^n$ consists precisely of those ${\mathbf {a}}\in M^n$ for which the fiber ${{X}_{{}}[{\mathbf {a}}]}$ has $D$-valued Euler [characteristic]{} $e$ in the expansion $(\mathcal M,{H_\natural}\cap M^{3n})$ whence in $\mathcal M$, as the former is [o-minimalistic]{} by Theorem \[T:ulominic\]. For the same reason, $F$ is [o-finitistic]{}, whence [o-minimalistic]{} by Theorem \[T:ofinexp\], so that we are done. \[R:ofinarch\] In everything in this section on Euler [characteristic]{}[s]{}, we may, by passing to a suitable permissible expansion, even assume that $D$ is only [o-finitistic]{}. Archimedean reducts ------------------- As before, let $D$ be definable and discrete with respective minimum $l$ and maximum $h$. By , we have a successor function $\sigma:=\sigma_D$, defined on $D\setminus\{h\}$, with inverse ${{\sigma^{-1}}}$ defined on $D\setminus\{l\}$. Let us write $e\ll d$, if $\sigma^n(e)<d$, for all $n\in{\mathbb N}$. If neither $d\ll e$ nor $e\ll d$, then $\sigma^n(d)=e$ for some $n\in{\mathbb Z}$, and we write $d\sim_D e$. The set of $\sim_D$-equivalence classes is totally ordered by $\ll$, and is called the *Archimedean reduct* ${\operatorname}{Arch}(D)$ of $D$. \[T:archred\] The Archimedean reduct ${\operatorname}{Arch}(D)$ of a definable, discrete subset $D$ in an [o-minimalistic]{} structure $\mathcal M$ is dense. This is clear if $D$ is finite, since then there is only one Archimedean class, so assume it is infinite. If ${\operatorname}{Arch}(D)$ is not dense, there would exist $l\ll h$ in $D$ so that for no $d\in D$ we have $l\ll d\ll h$. Therefore, upon replacing $D$ with $D\cap {{\, [l,h]\, }}$, we may assume that ${\operatorname}{Arch}(D)$ consists of exactly two classes, those of $l$ and $h$ . By Corollary \[C:ulomin\] (or, Theorem \[T:ulominic\]), we can embed $\mathcal M$ elementary in an ultra-o-minimal structure $\mathcal N$ so that $D$ is the restriction of a (definable) ultra-finite set $F$ in $\mathcal N$. Let $\mathcal N_i$ and $F_i$ be respectively o-minimal structures and finite subsets in these with ultraproduct equal to $\mathcal N$ and $F$ respectively. For each $i$, let $f_i\colon F_i\to F_i$ be the map reversing the (lexicographical) order and let ${\Gamma_\natural}$ be the ultraproduct of the graphs of the $f_i$. Since this is an ultra-finite set, its restriction $\Gamma$ to $\mathcal M$ is an [o-finitistic]{} set by Theorem \[T:ulominic\]. By [Łoś’ Theorem]{}, $\Gamma$ is the graph of the order reversing permutation $f\colon D\to D$. In particular, $f$ is definable in the [o-minimalistic]{} expansion $(\mathcal M,\Gamma)$ and maps any element in the class of $l$ to an element in the class of $h$ and vice versa. By definability, there is a maximal $a\in D$ such that $f(a)\geq a$. In particular, $f(a')<a'$, where $a'$ is the successor of $a$ in $D$. A moment’s reflection then shows that then either $f(a)=a$ or $f(a)=a'$, which contradicts that no element is $\sim_D$-equivalent with its image. \[R:archred\] Similarly, given $D,E\in{{\mathfrak D^{\text{viso}}(\mathcal M)}}$, we can define $D\ll E$ if for every finite subset $F$, we have $D\cup F\leq E$. If neither $D\ll E$ nor $E\ll D$, then we say that $D$ and $E$ have the same *virtual Archimedean class*, and write $D\sim E$. This is equivalent with the existence of finite subsets $F$ and $G$ such that $D\cup F$ and $E\cup G$ are [virtually isomorphic]{}. The induced order $\ll $ on virtual Archimedean classes is dense: indeed, suppose $D\ll E$ and let $d:={\chi_{E}(D)}$ and $h:={\chi_{E}(E)}$ (i.e., the maximum of $E$). By Theorem \[T:archred\], since $d\ll h$, there is some $a\in D$ with $d\ll a\ll h$. It follows that $D\ll E_{\leq a}\ll E$. [Taylor]{} sets {#s:Taylor} =============== In this section, we work in an expansion of ${\mathbb R}$ and its ultrapower ${{\mathbb R}_\natural}$, and we introduce some notation and terminology tailored to this situation. Recall that an element in ${{\mathbb R}_\natural}$ is called *infinitesimal* if its norm is smaller than $1/n$, for all positive $n$. The *standard part* of $\alpha\in{{\mathbb R}_\natural}$, denoted ${{\alpha}\mathstrut_\sharp}$, is the supremum of all $r\in{\mathbb R}$ with $r\leq \alpha$; if ${{\alpha}\mathstrut_\sharp}$ is not infinite (that is to say, if $\alpha$ is *bounded*), then ${{\alpha}\mathstrut_\sharp}-\alpha$ is infinitesimal and ${{\alpha}\mathstrut_\sharp}$ is the unique real number with this property. If $\alpha$ is a tuple ${(\alpha_1,\dots,\alpha_{k})}$, then we define ${{\alpha}\mathstrut_\sharp}$ coordinate-wise as $({{\alpha_1}\mathstrut_\sharp},\dots,{{\alpha_k}\mathstrut_\sharp})$. For a subset $X\sub{\mathbb R}^k$, we write ${{X}\mathstrut_\sharp}$ for the set of all ${{\alpha}\mathstrut_\sharp}$ where $\alpha$ runs over all bounded elements of ${X_\natural}$ (so that $\pm\infty$ never belongs to ${{X}\mathstrut_\sharp}$), and, following the ideology from [@SchUlBook §8], we call ${{X}\mathstrut_\sharp}$ the *catapower* of $X$. We note the following simple result from non-standard analysis: \[L:cpclos\] The catapower of a subset $X\sub{\mathbb R}^k$ is equal to its closure $\bar X$. Suppose $\alpha\in {X_\natural}$ is bounded, given as the ultraproduct of elements ${\mathbf {a}}_n\in X$. Hence the ultraproduct of the sequence ${{\alpha}\mathstrut_\sharp}-{\mathbf {a}}_n$ is an infinitesimal, showing that ${\mathbf {a}}_n$ are arbitrary close to ${{\alpha}\mathstrut_\sharp}$ for almost all $n$. Put differently, there exists a subsequence of $({\mathbf {a}}_n)_n$ which converges to ${{\alpha}\mathstrut_\sharp}$, proving that ${{\alpha}\mathstrut_\sharp}$ lies in the closure of $X$. Conversely, if ${\mathbf {b}}$ lies in closure of $X$, then we can find a sequence ${\mathbf {b}}_n\in X$ converging to it, and by the same argument, ${{({{\mathbf {b}}_\natural})}\mathstrut_\sharp}={\mathbf {b}}$, where ${{\mathbf {b}}_\natural}$ is the ultraproduct of the sequence ${\mathbf {b}}_n$. In [@SchUlBook Chapter 9], we also introduce the notion of a *protopower*; since it was catered to deal with an additional ring structure, which is not needed here, we will use only the following simplified version: for $X\sub{\mathbb R}^k$, let ${ X_{\downharpoonleft_{n}}}$ be the set of points in $X$ whose coordinates have norm at most $n$, where the norm of a point is defined as the maximum of the absolute values of its coordinates. We define the *protopower* ${{{\mathbb R}}\mathstrut_\flat}$ of ${\mathbb R}$ as the ultraproduct of the ${ {\mathbb R}_{\downharpoonleft_{n}}}$. We extend this to any subset $X\sub {\mathbb R}^k$, by calling the ultraproduct of the truncations ${ X_{\downharpoonleft_{n}}}$ the *protopower* of $X$, and denote it ${{X}\mathstrut_\flat}$. In other words, ${{X}\mathstrut_\flat}={X_\natural}\cap {{{\mathbb R}}\mathstrut_\flat}^k$, where ${X_\natural}$ is the ultrapower of $X$. In particular, any protopower is bounded (in norm) by ${{\omega_\natural}}$. (To make this conform with the definitions in [@SchUlBook §9], one actually has to take the protoproduct of the structures $({\mathbb R},\frac 1n{\left|\cdot\right|})$, and the Archimedean hull of our ${{{\mathbb R}}\mathstrut_\flat}$ is then equal to this protoproduct.) By the *trace* of a subset $\Xi\sub{{\mathbb R}_\natural}^k$, denoted ${{\operatorname}{tr}(\Xi)}$, we mean the set of its real points, that is to say, ${{\operatorname}{tr}(\Xi)}=\Xi\cap{\mathbb R}^k$. If $\Xi$ is definable by a formula $\varphi$ in some expansion of ${{\mathbb R}_\natural}$, we may use the slightly ambiguous notation $\varphi({\mathbb R})$ for its trace as well. The trace of a protopower ${{X}\mathstrut_\flat}$ is equal to $X$, that is to say, $X={{\operatorname}{tr}({{X}\mathstrut_\flat})}$: indeed, ${\mathbf {a}}\in{\mathbb R}^k$ satisfies ${\mathbf {a}}\in {{X}\mathstrut_\flat}$, [if and only if]{} ${\mathbf {a}}\in{ X_{\downharpoonleft_{n}}}$ for almost all $n$ (by [Łoś’ Theorem]{}), [if and only if]{}  ${\mathbf {a}}\in X$. For given $n\in{\mathbb N}$ and a $k$-ary function $f$, let us write ${ f_{\downharpoonleft_{n}}}$ for the *truncated* function defined by sending a point ${\mathbf {a}}$ to $f({\mathbf {a}})$ if ${\left|{\mathbf {a}}\right|}\leq n$ and to zero otherwise (note that this is not the same as taking the truncation of the graph of $f$, since we allow values of arbitrary high norm). Let ${L^{\text{an}}}$ be the language of ordered fields together with a function symbol for each everywhere convergent power series (also referred to as a *globally analytic* function). Clearly, we may view ${\mathbb R}$ as an ${L^{\text{an}}}$-structure, but this is not very useful, since ${\mathbb Z}$ is definable in it (as the zero set of $\sin(\pi x)$), and therefore neither [tame]{} nor [o-minimalistic]{}. Instead, we approximate this ${L^{\text{an}}}$-structure on ${\mathbb R}$ as follows. Let ${\mathcal R^{\text{an}}}_n$ be the ${L^{\text{an}}}$-structure on ${\mathbb R}$ where each function symbol corresponding to a convergent power series $f$ is interpreted as its truncation ${ f_{\downharpoonleft_{n}}}$. By [@DvdD], each ${\mathcal R^{\text{an}}}_n$ is o-minimal (where one usually denotes ${\mathcal R^{\text{an}}}_1$ by ${\mathbb R}_{\text{an}}$), and hence their ultraproduct ${{\mathcal R^{\text{an}}}_\natural}$ is [o-minimalistic]{}. Moreover, ${{\mathcal R^{\text{an}}}_\natural}$ is [tame]{} by Corollary \[C:tame\]. Although not part of the signature, power series with a smaller radius of convergence can also be encoded in this structure, at least in one variable: using a combination of linear transformations $x\mapsto ax+b$, and the (inverse) trig functions $\tan x$ and $\arctan x$, any two open intervals (bounded or unbounded) are isomorphic via a globally analytic map. For instance, if $f$ is defined on the open interval ${{\, ]-1,1[\, }}$, then $g(x):=f(\frac 2\pi\arctan x)$ is globally analytic, and hence is definable in ${L^{\text{an}}}$. \[D:Taylor\] We call $X\sub{\mathbb R}^k$ a *[Taylor]{}* set, if there exists an ${L^{\text{an}}}$-formula $\varphi({\mathbf {x}},{\mathbf {y}})$ (without parameters), such that for each sufficiently large $n$, there exists a tuple of parameters ${\mathbf {b}}_n$ so that ${ X_{\downharpoonleft_{n}}}=\varphi({\mathcal R^{\text{an}}}_n,{\mathbf {b}}_n)$. Modifying $\varphi$ if necessary, we may even assume that this holds for all $n$, and that ${\left|{\mathbf {x}}\right|}\leq n$ is a conjunct in $\varphi$. If ${{\mathbf {b}}_\natural}$ is the ultraproduct of the ${\mathbf {b}}_n$, then the protopower ${{X}\mathstrut_\flat}$ is equal to $\varphi({{\mathcal R^{\text{an}}}_\natural},{{\mathbf {b}}_\natural})$ by [Łoś’ Theorem]{}, and hence $X={{\operatorname}{tr}({{X}\mathstrut_\flat})}$. Any set realized as a *protopower* of a [Taylor]{} set will be called an *analytic protopower*, giving a one-one correspondence between [Taylor]{} sets and analytic protopowers. We refer to the defining formula $\varphi({\mathbf {x}},{{\mathbf {b}}_\natural})$ of ${{X}\mathstrut_\flat}$ as the *analytic* formula for $X$, and we express this by writing $X=\varphi({\mathbb R})$ (this does not mean that $X$ is definable, since the parameters might be non-standard; in the terminology of §\[s:locdef\], a [Taylor]{} set is in general only locally ${L^{\text{an}}}$-definable). Not every definable subset is an analytic protopower (equivalently, not every ${L^{\text{an}}}({{\mathbb R}_\natural})$-formula is analytic): let $\Theta$ be defined by $(\exists y)\, xy=1{\wedge}\sin(\pi y)=0$. Its trace ${{\operatorname}{tr}(\Theta)}$ is equal to the set of reciprocals of positive natural numbers and cannot be a [Taylor]{} set by Lemma \[L:discTaylor\] below. Any quantifier free ${L^{\text{an}}}({{\mathbb R}_\natural})$-formula is analytic, so that in particular, any globally real analytic variety is [Taylor]{}. [Taylor]{} sets are closed under (finite) Boolean combinations, but not under definable (analytic) images, nor under projections. In particular, the [Taylor]{} sets do not form a first-order structure. \[L:discTaylor\] A real discrete subset is [Taylor]{} [if and only if]{} it is closed. Moreover, a discrete [Taylor]{} set intersects any bounded set in finitely many points. If $X$ is discrete, then ${ X_{\downharpoonleft_{n}}}$ must be finite by o-minimality, and hence $X$ cannot have an accumulation point whence is closed. Conversely, if $X$ is discrete and closed, then it is the zero set of some analytic function $f$ (taking sums of squares allows us to reduce to a single equation), and hence ${ X_{\downharpoonleft_{n}}}$ is defined in ${\mathcal R^{\text{an}}}_n$ by ${ f_{\downharpoonleft_{n}}}({\mathbf {x}})=0$, and ${\left|{\mathbf {x}}\right|}\leq n$. \[L:Taylorul\] A subset $X\sub {\mathbb R}^k$ is [Taylor]{} [if and only if]{} its protopower ${{X}\mathstrut_\flat}$ is ${{\mathcal R^{\text{an}}}_\natural}$-definable. Recall that ${{X}\mathstrut_\flat}$ is the ultraproduct of the truncations ${ X_{\downharpoonleft_{n}}}$. One direction has already been observed. Assume ${{X}\mathstrut_\flat}$ is ${{\mathcal R^{\text{an}}}_\natural}$-definable, say ${{X}\mathstrut_\flat}=\varphi({{\mathcal R^{\text{an}}}_\natural},{\mathbf {b}})$, for some ${L^{\text{an}}}$-formula $\varphi$ and some tuple of parameters ${\mathbf {b}}$. Writing ${\mathbf {b}}$ as the ultraproduct of tuples ${\mathbf {b}}_n$, it follows from [Łoś’ Theorem]{} that ${ X_{\downharpoonleft_{n}}}=\varphi({\mathcal R^{\text{an}}}_n,{\mathbf {b}}_n)$ for almost all $n$. Enlarging the tuple of parameters if necessary, we may assume that $n$ is one of the entries of ${\mathbf {b}}_n$. Choosing for each $n$ some $m>n$ such that ${ X_{\downharpoonleft_{m}}}=\varphi({\mathcal R^{\text{an}}}_m,{\mathbf {b}}_m)$, we get ${ X_{\downharpoonleft_{n}}}=\varphi({\mathcal R^{\text{an}}}_n,{\mathbf {b}}_m){\wedge}{\left|{\mathbf {x}}\right|}\leq n$, showing that $X$ is [Taylor]{}. We can rephrase this as a criterion for analytic protopowers: \[C:Taylorul\] A protopower ${{X}\mathstrut_\flat}\sub {{\mathbb R}_\natural}^k$ is analytic [if and only if]{} it is ${{\mathcal R^{\text{an}}}_\natural}$-definable [if and only if]{} its trace $X$ is [Taylor]{}. In terms of formulae, we might paraphrase this as: an ${L^{\text{an}}}({{\mathbb R}_\natural})$-formula $\varphi$ is analytic [if and only if]{} $\varphi({{\mathbb R}_\natural})$ is the ultraproduct of the $\varphi({ {\mathbb R}_{\downharpoonleft_{n}}})$. Thus, an open interval in ${{\mathbb R}_\natural}$ is an (analytic) protopower [if and only if]{} its endpoints are either real or equal to $\pm{{\omega_\natural}}$: indeed, suppose ${{\, ]\alpha,\beta[\, }}$ is a protopower, and let ${{\alpha}\mathstrut_\sharp}$ and ${{\beta}\mathstrut_\sharp}$ be the respective standard parts of $\alpha$ and $\beta$. Hence $I:={{\, ]\alpha,\beta[\, }}\cap {\mathbb R}$ is a (not necessarily open) interval with endpoints ${{\alpha}\mathstrut_\sharp}$ and ${{\beta}\mathstrut_\sharp}$. If ${{\alpha}\mathstrut_\sharp}$ is finite, then ${ I_{\downharpoonleft_{n}}}$ is an interval with left endpoint ${{\alpha}\mathstrut_\sharp}$ for $n$ sufficiently large, and hence the same is true for the ultraproduct of these truncations. By Corollary \[C:Taylorul\], this forces ${{\alpha}\mathstrut_\sharp}=\alpha$. In the other case, the left endpoint of ${ I_{\downharpoonleft_{n}}}$ is $-n$, and hence their ultraproduct has left endpoint $-{{\omega_\natural}}$, showing that $\alpha=-{{\omega_\natural}}$. The same argument applies to $\beta$, proving the claim. \[E:panal\] By Lemma \[L:discTaylor\], every closed, discrete subset, whence in particular any subset of ${\mathbb Z}$, is [Taylor]{}. To give a non-discrete example, consider the spiral $C\sub {\mathbb R}^2$ with parametric equations $x=\exp \tau\sin \tau$ and $y=\exp \tau\cos \tau$, for $\tau\in{\mathbb R}$. If $(x,y)\in{ C_{\downharpoonleft_{n}}}$, then $\exp\tau=\sqrt{x^2+y^2}\leq n\sqrt 2$ and hence $\tau\leq \log(n\sqrt 2)\leq n$. In particular, the negative values of $\tau$ can be larger in absolute value than $n$. Hence $C$ is not [Taylor]{}. However, if $C^+$ is the ‘positive’ part, given by the same equations but only for $\tau\geq 0$, then ${ C^+_{\downharpoonleft_{n}}}$ is defined in ${\mathcal R^{\text{an}}}_n$ by $x={ \exp_{\downharpoonleft_{n}}} (\tau){ \sin_{\downharpoonleft_{n}}}(\tau)$, $y={ \exp_{\downharpoonleft_{n}}}(\tau){ \cos_{\downharpoonleft_{n}}}(\tau)$, and $\tau\leq { \log_{\downharpoonleft_{n}}}(n\sqrt 2)$, showing that $C^+$ is [Taylor]{} (see Corollary \[C:polarTaylor\] below). \[P:closTaylor\] The closure, interior, frontier, and boundary of a [Taylor]{} set is again [Taylor]{}. Since all concepts are obtained by either taking closures or Boolean combinations, it suffices to show that the closure $\bar X$ of a [Taylor]{} set $X$ is again [Taylor]{}. Let $\varphi(x,z)$ be an analytic formula for $X$, so that ${ X_{\downharpoonleft_{n}}}=\varphi({\mathcal R^{\text{an}}}_n,{\mathbf {b}}_n)$, for some parameters ${\mathbf {b}}_n$ and all $n$. If $\psi(x,z)$ is the formula $(\forall a>0)(\exists y){\left|x-y\right|}<a{\wedge}\varphi(y,z)$, then $\psi({\mathcal R^{\text{an}}}_n,{\mathbf {b}}_n)$ defines the closure of ${ X_{\downharpoonleft_{n}}}$. It is now easy to check that the latter is equal to ${ \bar X_{\downharpoonleft_{n}}}$, showing that $\bar X$ is [Taylor]{}. \[R:closTaylor\] From the proof it is also clear that if ${{X}\mathstrut_\flat}$ is the protopower of $X$, then the closure $\overline{{{X}\mathstrut_\flat}}$ of ${{X}\mathstrut_\flat}$ is the protopower of $\bar X$, and the analogous properties for the other topological operations. Inspecting the above proofs and examples, we can single out the following geometric feature of [Taylor]{} sets.[^3] \[P:quantanalform\] Let $X\sub {\mathbb R}^{k+1}$ be a [Taylor]{} set and let $Y\sub {\mathbb R}^k$ be its projection onto the first $k$ coordinates. If there exists $K\in{\mathbb N}$ such that ${ Y_{\downharpoonleft_{n}}}$ is contained in the projection of ${ X_{\downharpoonleft_{Kn}}}$, for all sufficiently large $n$, then $Y$ is again [Taylor]{}. Let $\varphi({\mathbf {x}},y,{{\mathbf {c}}_\natural})$ be the analytic formula defining $X$, and choose tuples ${\mathbf {c}}_n$ with ultraproduct equal to ${{\mathbf {c}}_\natural}$, so that ${ X_{\downharpoonleft_{n}}}$ is defined in ${\mathcal R^{\text{an}}}_n$ by $\varphi({\mathbf {x}},y,{\mathbf {c}}_n)$. Let $\tilde \varphi({\mathbf {x}},y,{{\mathbf {c}}_\natural})$ be the formula obtained from $\varphi$ by replacing every power series $f({\mathbf {x}},y)$ occurring in it by the power series$f({\mathbf {x}}, Ky)$, and put $\psi({\mathbf {x}}, {{\mathbf {c}}_\natural}):=(\exists y)\tilde \varphi({\mathbf {x}},y,{{\mathbf {c}}_\natural})$. I claim that $\psi$ is an analytic formula with $\psi({\mathbb R})=Y$. To this end, we have to show that ${ Y_{\downharpoonleft_{n}}}=\psi({\mathcal R^{\text{an}}}_n,{{\mathbf {c}}}_n)$, for almost all $n$. One inclusion is clear, so assume ${\mathbf {a}}\in{ Y_{\downharpoonleft_{n}}}$, for some $n$. Hence ${\left|{\mathbf {a}}\right|}\leq n$ and there exists $b\in{\mathbb R}$ such that $({\mathbf {a}},b)\in X$. By assumption, we can choose ${\left|b\right|}\leq Kn$. Let $b':=b/K$, so that ${\left|b'\right|}\leq n$. Since then ${\mathcal R^{\text{an}}}_n\models\tilde\varphi({\mathbf {a}},b')$, as the point $({\mathbf {a}},b')$ has norm at most $n$, whence agrees on any power series with its $n$-th truncation, we get ${\mathcal R^{\text{an}}}_n\models\psi({\mathbf {a}})$, as required. Given a $C_1$-function $f\colon{\mathbb R}\to {\mathbb R}$ on an open interval ${{\, ]a,b[\, }}$, we say that $f$ is *increasing* at $b$ if $$\lim_{x\to b^-}f'(x)>0$$ with a similar definition for decreasing or at the left endpoint. \[C:polarTaylor\] Let $f$ be a power series converging on a half-open interval ${{\, [a,b[\, }}$. If $f$ is increasing at $b$, then the curve $C\sub {\mathbb R}^2$ with polar equation $R=f(\theta)$, for $a\leq\theta<b$, is [Taylor]{}. By the discussion at the beginning of this section, we may make an order-preserving, analytic change of variables so that $f$ becomes convergent on ${\mathbb R}_{\geq0}$. By L’H\^ opital’s rule, the limit of $f(x)/x$ for $x\to\infty$ exists and is positive. Hence, we may choose $K\in{\mathbb N}$ large enough so that $1/K<f(x)/x$ for all $x\geq K$. Let $X\sub{\mathbb R}^3$ be the semi-analytic set given by $x=f(z)\sin(z)$, $y=f(z)\cos(z)$, and $a\leq z<b$, so that $C$ is just the projection of $X$ onto the first two coordinates. By Proposition \[P:quantanalform\], it suffices to show that ${ C_{\downharpoonleft_{n}}}$ is contained in the projection of ${ X_{\downharpoonleft_{(K\sqrt 2)n}}}$, for all $n$. To this end, let $(a,b)\in{ C_{\downharpoonleft_{n}}}$, so that $a=f(\theta)\sin\theta$ and $b=f(\theta)\cos\theta$, for some $\theta\geq0$. In particular, $f(\theta)= \sqrt {a^2+b^2}\leq n\sqrt 2$. There is nothing to prove if $\theta\leq K$, so let $\theta>K$ and hence $1/K<f(\theta)/\theta$. The result now follows since $\theta< Kf(\theta)\leq (K\sqrt 2)n$. Of course, a similar criterion exists if the domain is open at the left endpoint, where the function now has to be decreasing. Any [Taylor]{} set is of the form $\varphi({\mathbb R})$ for some ${L^{\text{an}}}({{\mathbb R}_\natural})$-formula $\varphi$, that is to say, is a trace of an ${{\mathcal R^{\text{an}}}_\natural}$-definable subset. For each such trace $X:=\varphi({\mathbb R})$, we can define its *dimension* ${\operatorname}{dim}(X)$ to be the dimension of $\varphi({{\mathcal R^{\text{an}}}_\natural})$. In general, this notion is not well behaved: the trace of the discrete, zero-dimensional set given by the formula $(\exists y>0) \sin(\pi y)=0{\wedge}\sin(\pi xy)=0$ is equal to $\mathbb Q$, a non-discrete set. Fortunately, [Taylor]{} sets behave tamely, as witnessed, for instance, by the following planar trichotomy (compare with Theorem \[T:triplanar\]): \[T:planardimTaylor\] A non-empty [Taylor]{} subset $X\sub {\mathbb R}^2$ is either 1. zero-dimensional, discrete, and closed; 2. one-dimensional, nowhere dense, but at least one projection has non-empty interior; 3. two-dimensional with non-empty interior. Let ${{X}\mathstrut_\flat}$ be the protopower of $X$ and $d$ its dimension. By Proposition \[P:dimulomin\], almost all truncations ${ X_{\downharpoonleft_{n}}}$ have dimension $d$ . Hence, if $d=0$, then almost all (whence all) ${ X_{\downharpoonleft_{n}}}$ are finite and $X$ is closed and discrete. If $d=2$, then almost all (whence all) ${ X_{\downharpoonleft_{n}}}$ have non-empty interior, whence so does $X$. Finally, if $d=1$, (almost) all ${ X_{\downharpoonleft_{n}}}$ are nowhere dense, and some projection has interior. Therefore, $X$ itself has the same properties. In view of Remark \[R:closTaylor\], the dimension of the frontier ${{\operatorname}{fr}( X)}$ of a [Taylor]{} set $X$ is strictly less than its dimension ${\operatorname}{dim}(X)$. Hence, by the same argument as for Corollary \[C:const\], we immediately get: \[C:constTaylor\] Any [Taylor]{} set is constructible. Next, we study maps in this context. For $X\sub {\mathbb R}^k$ and $Y\sub {\mathbb R}^l$, let us call a map $f\colon X\to Y$ *[Taylor]{}*, if its graph is a [Taylor]{} set. \[C:defmapTaylor\] The domain and image of a [Taylor]{} map are [Taylor]{}, and so is any fiber. Likewise, if the graph of an ${{\mathcal R^{\text{an}}}_\natural}$-definable map $\gamma\colon\Xi\to \Theta$ is a protopower, then so are $\Xi$ and $\gamma(\Xi)$, as well as every fiber ${{\gamma^{-1}({\mathbf {b}})}}$ with ${\mathbf {b}}\in Y$. Moreover, the trace of $\gamma$ induces a [Taylor]{} map $g\colon {{\operatorname}{tr}(\Xi)}\to{{\operatorname}{tr}(\gamma(\Xi))}$, and any [Taylor]{} map is obtained in this way. The first assertion follows from the last assertion by Corollary \[C:Taylorul\]. So assume the graph $\Gamma(\gamma)$ is a protopower. Without loss of generality, we may assume $\Theta=\gamma(\Xi)$, that is to say, that $\gamma$ is surjective. Let $G:={{\operatorname}{tr}(\Gamma(\gamma))}$ be the trace, so that the ultraproduct of the ${ G_{\downharpoonleft_{n}}}$ is equal to $\Gamma(\gamma)$, and let $X:={{\operatorname}{tr}(\Xi)}$ and $Y:={{\operatorname}{tr}(\Theta)}$ be the respective traces of domain and image. It follows that ${ X_{\downharpoonleft_{n}}}$ is defined by the formula $(\exists {\mathbf {y}}) ({\mathbf {x}},{\mathbf {y}})\in { G_{\downharpoonleft_{n}}}$, and hence $X$ is [Taylor]{} (alternatively, use Proposition \[P:quantanalform\]). Moreover, $\gamma$ restricted to ${ X_{\downharpoonleft_{n}}}$ takes values inside ${ Y_{\downharpoonleft_{n}}}$, that is to say, induces a map $g_n\colon { X_{\downharpoonleft_{n}}}\to { Y_{\downharpoonleft_{n}}}$. It follows that the ultraproduct of the $g_n$ is equal to $\gamma$. By [Łoś’ Theorem]{}, almost all $g_n$ are surjective. Therefore, the respective ultraproducts of ${ X_{\downharpoonleft_{n}}}$ and ${ Y_{\downharpoonleft_{n}}}$ are $\Xi$ and $\Theta$, proving that both sets are analytic protopowers by Corollary \[C:Taylorul\]. Moreover, the union $g$ of the $g_n$ is the restriction of $\gamma$ to $X$. Fix ${\mathbf {b}}\in Y$, let $\Phi:={{\gamma^{-1}({\mathbf {b}})}}$ its fiber and $F:={{\operatorname}{tr}(\Phi)}$ the latter’s trace. One checks that ${ F_{\downharpoonleft_{n}}}={{g_n^{-1}({\mathbf {b}})}}$, and hence the ultraproduct of the ${ F_{\downharpoonleft_{n}}}$ is equal to $\Phi$, proving that $\Phi$ is an analytic protopower. \[R:Taylorcell\] In particular, a horizontal [Taylor]{} $1$-cell in ${\mathbb R}^2$ must be the graph of a continuous, [Taylor]{} map, and similarly, a [Taylor]{} $2$-cell in ${\mathbb R}^2$ is the region between two [Taylor]{} graphs. Let $X$ be a [Taylor]{} set with protopower ${{X}\mathstrut_\flat}$. Since ${{\mathcal R^{\text{an}}}_\natural}$ is [tame]{}, we can find a surjective, cellular map $\delta\colon {{X}\mathstrut_\flat}\to \Delta$ with $\Delta$ a discrete, closed set. I conjecture that we may take $\delta$ to be a protopower too. Assuming this, taking traces yields a [Taylor]{} map $d\colon X\to {{\operatorname}{tr}(\Delta)}$, whose fibers are all [Taylor]{} cells, and hence defined by means of continuous [Taylor]{} maps. This yields a [Taylor]{} cell decomposition of $X$ which is finite on each compact subset by Lemma \[L:discTaylor\]. \[C:DPPTaylor\] Any discrete [Taylor]{} set $D$ satisfies DPP in the sense that a [Taylor]{} map $D\to D$ is injective [if and only if]{} it is surjective. Let $g\colon D\to D$ be [Taylor]{} and let ${{g}\mathstrut_\flat}\colon {{D}\mathstrut_\flat}\to {{D}\mathstrut_\flat}$ be its protopower, that is to say, given as the ultraproduct of the restrictions $g_n:={\left.g\right|_{{{ D_{\downharpoonleft_{n}}}}}}$. By [Łoś’ Theorem]{}, ${{g}\mathstrut_\flat}$ is injective (surjective) [if and only if]{} almost all $g_n$ are, whence [if and only if]{} $g$ is, and the result now follows easily from the [o-minimalistic]{} DPP (Proposition \[P:DPP\]). \[C:discTaylor\] A [Taylor]{} map $g\colon X\to Y$ is continuous outside a set of dimension strictly less than the dimension of $X$. In particular, one-variable [Taylor]{} maps are monotone outside a discrete, closed ([Taylor]{}) subset. We may assume, for the purposes of this proof that $g$ is surjective, so that, in particular, both $X$ and $Y$ are [Taylor]{}, by Corollary \[C:defmapTaylor\]. By the same result, taking protopowers yields a definable map ${{g}\mathstrut_\flat}\colon {{X}\mathstrut_\flat}\to {{Y}\mathstrut_\flat}$ whose restriction to $X$ is equal to $g$. By Theorem \[T:disctu\], the set of discontinuities $\Delta$ of ${{g}\mathstrut_\flat}$ has dimension strictly less than ${\operatorname}{dim}({{X}\mathstrut_\flat})={\operatorname}{dim}(X)$. Replacing $\Delta$ by its closure,[^4] which does not change the dimension, we may assume $\Delta$ is closed. I claim that $g$ is continuous outside the trace $D:={{\operatorname}{tr}(\Delta)}$. Indeed, if $a\in X\setminus D$, then by the non-standard criterion for continuity, we have to show that for every $\alpha$ infinitesimally close to $a$, their images under ${g_\natural}$ remain infinitesimally close, where ${g_\natural}$ is the ultrapower of $g$. However, since ${{g}\mathstrut_\flat}$ is the ultrapower of the restrictions ${\left.g\right|_{{{ X_{\downharpoonleft_{n}}}}}}$, both maps agree on bounded elements, and so we have to show that ${{g}\mathstrut_\flat}(a)$ and ${{g}\mathstrut_\flat}(\alpha)$ are infinitesimally close. This does hold indeed for $\alpha$ sufficiently close to $a$ since $a\notin \Delta$ and $\Delta$ is closed. In the one-variable case, we may choose $\Delta$ so that ${{g}\mathstrut_\flat}$ is monotone on any interval with endpoints in $\Delta$, and clearly, $g$ is then monotone on $D$. It follows from Lemma \[L:discTaylor\] and Theorem \[T:planardimTaylor\] that $D$ is [Taylor]{}. \[R:genctuTaylor\] Using the discussion in Remark \[R:genctu\], we can choose $\Delta$ in the above statement also to be [Taylor]{} in higher dimensions. If $f\colon X\to Y$ is [Taylor]{} and bijective, then its inverse is also [Taylor]{}, and we will say that $X$ and $Y$ are *analytically isomorphic*. In the definition of a [Grothendieck ring]{}, it was not necessary that the collection of subsets formed a first-order structure, only that they were preserved under Boolean combinations and products. Since this is true also of [Taylor]{} sets, we can define the *analytic* [Grothendieck ring]{} ${{\mathbf {Gr}^{\text{an}}}}$ as the free Abelian group of analytic isomorphism classes of [Taylor]{} sets modulo the scissor relations. \[P:angr\] Sending the class of a [Taylor]{} set $X$ to the class of its protopower ${{X}\mathstrut_\flat}$ induces a natural ring [homomorphism]{} of [Grothendieck ring]{}[s]{} ${{\mathbf {Gr}^{\text{an}}}}\to {{\mathbf {Gr}(\mathcal {{{\mathcal R^{\text{an}}}_\natural}})}}$. To show that the map ${{[ X]}}\to {{[ {{X}\mathstrut_\flat}]}}$ is well-defined, suppose $f\colon X\to Y$ is an analytic isomorphism. The ultraproduct of the truncations $f_n\colon { X_{\downharpoonleft_{n}}}\to { Y_{\downharpoonleft_{n}}}$ induces then a definable map ${{f}\mathstrut_\flat}\colon {{X}\mathstrut_\flat}\to {{Y}\mathstrut_\flat}$, and by [Łoś’ Theorem]{}, this is again a bijection. [10]{} Raf Cluckers and M[á]{}rio Edmundo, *Integration of positive constructible functions against [E]{}uler characteristic and dimension*, J. Pure Appl. Algebra **208** (2007), no. 2, 691–698. Jan Denef and Lou van den Dries, *$p$-adic and real subanalytic sets*, Ann. of Math. **128** (1988), 79–138. Alf Dolich, Chris Miller, and Charlie Steinhorn, *Structures having o-minimal open core*, Trans. Amer. Math. Soc. **362** (2010), no. 3, 1371–1411. Antongiulio Fornasiero, *Tame structures and open cores*, ArXiv 1003.3557, 2010. Jan Kraj[í]{}[č]{}ek, *Uniform families of polynomial equations over a finite field and structures admitting an [E]{}uler characteristic of definable sets*, Proc. London Math. Soc. (3) **81** (2000), no. 2, 257–284. Jan Kraj[í]{}[č]{}ek and Thomas Scanlon, *Combinatorics with definable sets: [E]{}uler characteristics and [G]{}rothendieck rings*, Bull. Symbolic Logic **6** (2000), no. 3, 311–330. Chris Miller, *Expansions of dense linear orders with the intermediate value property*, J. Symbolic Logic **66** (2001), no. 4, 1783–1790. [to3em]{}, *Tameness in expansions of the real field*, Logic [C]{}olloquium ’01, Lect. Notes Log., vol. 20, Assoc. Symbol. Logic, Urbana, IL, 2005, pp. 281–316. Chris Miller and Patrick Speissegger, *Expansions of the real line by open sets: o-minimality and open cores*, Fund. Math. **162** (1999), no. 3, 193–208. Anand Pillay, *An introduction to stability theory*, Oxford Logic Guides, vol. 8, The Clarendon Press Oxford University Press, New York, 1983. Hans Schoutens, *The use of ultraproducts in commutative algebra*, Lecture Notes in Mathematics, vol. 1999, Springer-Verlag, 2010. Carlo Toffalori and Kathryn Vozoris, *Notes on local o-minimality*, MLQ Math. Log. Q.Math. Log. Q. **55** (2009), no. 6, 617–632. Lou van den Dries, *Tame topology and o-minimal structures*, LMS Lect. Note Series, Cambridge University Press, Berlin, 1998. Kathryn Vozoris, *The complex field with a predicate for the integers*, Ph.D. thesis, University of Illinois at Chicago, 2007. [^1]: The concept of ‘tameness’ is quite vague and often depends on the particular author’s taste, the present author not excluded; see §\[s:tame\]. [^2]: This is a slightly stronger version of what is called in the literature *local o-minimality*, but which agrees with it in the case of an expansion of an ordered field. [^3]: The corresponding syntactic characterization of analytic formulae is not yet clear to me. [^4]: In fact, this is not needed since one can show that $\Delta$ is already closed.
{ "pile_set_name": "ArXiv" }
--- abstract: 'The possibility of detecting mutations in a DNA from force measurements (as a first step towards sequence analysis) is discussed theoretically based on exact calculations. The force signal is associated with the domain wall separating the zipped from the unzipped regions. We propose a comparison method (“differential force microscope”) to detect mutations. Two lattice models are treated as specific examples.' address: - 'Institute of Physics, Bhubaneswar 751 005, India' - 'International School for Advanced Studies (SISSA) and INFM, Via Beirut 2-4, 34014, Trieste, Italy' author: - 'Somendra M. Bhattacharjee[@eml1]' - 'D. Marenduzzo[@eml1]' title: 'DNA sequence from the unzipping force? : one mutation problem' --- ł [87.14Gg, 87.80Fe, 05.20Gg]{} The possibility of a force induced unzipping transition [@Boc1; @Boc2; @smb; @Nel; @maren; @gerland] has opened up new ways of exploring the properties of biomolecules. Since the critical force for a double stranded DNA depends on its sequence, an inverse problem can be posed: “Can the DNA sequence be detected from the force required to unzip it ?” A still simpler question would be: “Can the mutations in a DNA be detected by force measurements?”. A positive answer to either of these questions would lead to the possibility of testing (and detecting) mutations and of sequence determination in a non-destructive way. Based on a few simple models used earlier for the DNA denaturation and unzipping transitions[@smb; @Nel; @maren], we show here the signature of mutations on the $f$-vs-$r$ curve (force versus relative distance of the end points of the two strands). The inverse problem is then to get the position of the mutations from such an experimentally realizable curve. Our emphasis in this paper is on the one base pair mutation problem also known as point mutation. This is not just of academic interest. The replication of DNA is a high-fidelity process, thanks to the inbuilt proof-reading and repair mechanisms, so that mistakes are very rare though even one could play havoc. Our proposal (solution of the inverse problem), based on exact calculations, is to obtain the force difference (“differential force”) between identically stretched native and mutant DNAs. This could be done by using [*e.g.*]{} two atomic force microscope tips: we name this apparatus a “differential force microscope”. The position of the mutation can be obtained from a calibration curve involving the extremum position (or its value) of the differential force curve. In our models, we can find the nature of the mutation from the sign of the differential force. Let us model a double stranded (ds) DNA by two N-monomer polymers interacting at the same contour length (or monomer index) $j$ of the strands through a contact attractive potential $-\epsilon_j(\epsilon_j>0)$, which might depend on the contour length. The interaction energy is $H=-\sum_{j=1}^N \delta_{{\bf r}_j,{\bf 0}}\epsilon_j$, where ${\bf r}_j$ denotes the relative distance of the $j-$th base pair and $\delta$ denotes the Kronecker delta. Other features like the self- and mutual-avoidance, base stacking energy, helicity, etc are ignored in this study (but see below) in order to focus on the base pairing energy. Such a model exhibits a denaturation transition at a model-dependent temperature $T=T_m$[@den] from a low $T$ double stranded configuration to a high $T$ phase of two unpaired single strands. A few definitions: Starting with a sequence $\{\epsilon_j|j=1,\ldots,N\}$ or bases of a DNA (to be called the native DNA), we define a mutant DNA as one with almost the same base sequence as the native molecule [*except* ]{} a few base pairs. A homogeneous DNA or homo-DNA is a DNA with identical base pairs (i.e. $\epsilon_j =\epsilon$ for all $j$) while a heterogeneous DNA is one with heterogeneity in the sequence ($j$-dependent $\epsilon_j$). Notice that the model we have defined does not consider base pair stacking interaction and therefore does not distinguish [*e.g.*]{} between an AT and a TA (or CG and GC) base pair. Consequently, the mutations we are referring to are only those involving AT (or TA) $\leftrightarrow $ CG (or GC). Two specific examples are considered here because of their exact tractability (analytical and numerical): (Mi) a two dimensional $d=2$ directed DNA model (with the strands directed along $(1,1)$) with base-pair interaction and mutual avoidance (forbidden crossing of the strands), and (Mii) Two Gaussian polymers with the base-pair interaction in $d$-dimensions. In both cases, only the relative chain, involving the separation [**r**]{} of the bases at the same contour length, need be considered. These models in addition to the denaturation transition also show, for $T<T_m$, an unzipping transition in presence of a force at the free end ($j=N$)[@smb; @Nel; @maren]. In this paper, we consider the conjugate fixed distance ensemble and henceforth restrict ourselves to $T<T_m$. Consequently, we put $\beta=(k_B T)^{-1}=1$ unless explicitly shown, where $k_B$ is the Boltzmann constant. For any ds DNA having their first monomers ($j=1$) joined and their last monomers ($j=N$) at a relative lattice distance ${\bf r}$, the force ${\bf f}_N({\bf r})$ required to maintain this relative distance ${\bf r}$ is ${\bf f}_N({\bf r})\equiv \nabla {\cal F}({\bf r})$ where ${\cal F} ({\bf r})$ represents the free energy of the system in the fixed-${\bf r}$ ensemble$^{\footnotemark[4]}$. By definition, $\int_0^\infty{\bf f}_N({\bf r})\cdot d {\bf r}$ = ${\cal F}(\infty)-{\cal F}(0)$ which gives the free energy of binding or the work required to unzip completely DNA. The two ensembles, fixed-force and fixed-stretch, are expected to give identical results in the $N\rightarrow \infty$ limit, though for finite $N$ inequivalence might be expected [@Neu] (this is indeed the picture valid for the homo-DNA), but a more serious situation arises if, [*e.g.*]{} for the scalar case, $\partial f/\partial x <0$ in the fixed-stretch ensemble because, in the fixed-force ensemble, $\frac{\partial\langle x\rangle}{\partial \beta f}$ $\equiv \langle x^2\rangle-\langle x\rangle^2\ge 0$, where $\langle\cdot\rangle$ denotes thermal averaging. This is the case for a heterogeneous DNA as shown in Fig. 1. The regions of “wrong” sign (reminiscent of preshocks in Burgers turbulence[@note1] or of “slip” in Ref.[@Boc2]) go away only after quenched averaging but do survive in the thermodynamic limit for each individual realization. [*The absence of self-averaging in the system is encouraging, because it implies that individual features, typical of a single realization of the sequence, are maintained in the characteristic curves.*]{} We draw attention to the overlap of the $f$ vs $x$ curves over a range of $x$ in Fig. 1 for different lengths with identical sequence near the open end. This overlap is a sign that the modulation is a characteristic of the sequence, and we have found this to be true quite generally. Let us now consider the one mutation problem where the $k$-th pairing energy of the native DNA has been changed from $\epsilon_k$ to $\epsilon'_k$. For this one site change in $H$, the partition function $Z_{N|k}({\bf r},N)$ is given by \[mutation\_one\] Z\_[N|k]{}([**r**]{},N)=Z\_N([**r**]{},N)+c\_k Z\_N([**r**]{},N,k). where $c_k\equiv\l(\exp{\l(\beta(\epsilon'_k-\epsilon_k)\r)}-1\r)$ and $Z_N({\bf r},N)$ is the native DNA partition function with last monomers at a relative distance ${\bf r}$, while $Z_{N}({\bf r},N\mid {\bf 0},k)$ is with the additional constraint of the $k-$th pair being zipped. Notice that Eq. \[mutation\_one\] applies for both heterogeneous and homogeneous cases (and also for self avoiding strands). We define a zipping probability $P({\bf r},N|{\bf 0},k)\equiv$ $\frac{Z_{N}({\bf r},N|{\bf 0},k)} {Z_N({\bf r},N)}$, which is the conditional probability that site $k$ is zipped when the free ends are at a distance ${\bf r}$. The force difference (to be called the differential force) to keep the free ends of both the mutant and the native DNA at the same distance ${\bf r}$ can be written as $$\begin{aligned} \delta {\bf f}_{N|k}(\bf r)&=&-\ \frac{c_k}{1+c_kP({\bf r},N|{\bf 0},k)} \nabla P({\bf r},N|{\bf 0},k)\\ & =& \frac{c_kP({\bf r},N|{\bf 0},k)}{1+c_kP({\bf r},N|{\bf 0},k)} ({\bf f}_{N}({\bf r}|{\bf 0},k)-{\bf f}_{N}({\bf r})). \end{aligned}$$ \[mut3\] Here ${\bf f}_{N} ({\bf r}|{\bf 0},k)$ is a generalized force, the one necessary to keep the free ends of the native DNA at a relative distance ${\bf r}$ when the $k-$th monomers are zipped. Except $c_k$, all other quantities in Eq. \[mut3\] refer to the native DNA. This fact allows the inverse problem to be tackled as we show explicitly for a few cases. Since for DNA we have generally two possible choices for $\epsilon$, ($\epsilon_1$ and $\epsilon_2$), the sign of $c_k$ determines the sign of $\delta f$ in Eq. \[mut3\]. Thus, in our simple model, the nature of the mutation can be identified from the sign of the differential force curve. Eqs. \[mutation\_one\],\[mut3\] can be generalized to more than one mutations and to models with other local energy parameters, though they become algebraically more involved. These more complicated cases will be discussed elsewhere. We consider the simplest case here. If ${\bf r}\equiv (x,0,\ldots,0)$ is the direction of stretching, the quantity $P({\bf r},N|{\bf 0},k)$ has a kink-like behavior (as in Fig. \[fig:2\]). In the fixed stretch ensemble, the chain separates into an unzipped and a zipped region separated by a domain wall, which to a very good approximation can be fit by a $\tanh$ function: $$\begin{aligned} \label{eq:3} P({\bf r},N&|&{\bf 0},k)\equiv P({\bf 0},N|{\bf 0},k)e^{-\beta \int d{\bf r}\cdot ({\bf f}_{N}({\bf r}|{\bf 0},k)- {\bf f}_N({\bf r}))} \nonumber\\ &\approx&P_0\l(1+\tanh{\chi}\r)/2, \, \chi\equiv{\l[x-x_{\rm d}(k)\r]} /{w_{\rm d}(k)}, \end{aligned}$$ where $x_{\rm d}(k)$ and $w_{\rm d}(k)$ are the position and the width of the wall (kink), and $P_0$ is a constant. Eq. \[eq:3\] suggests that $\delta f(x)\equiv w_{\rm d}(k)^{-1} {\tilde {\sf f}}(\chi)$, where ${\tilde {\sf f}}(\chi)$ is a scaling function. It is possible to extend the above analysis to the general case where more than one mutation is present. For example, the partition function for the case with two mutations at positions $k_1$ and $k_2$, in an obvious notation, is $$\begin{aligned} \label{eq:4} Z_{N|k_1,k_2}({\bf r})=Z_N({\bf r})&+&\sum_{i=1,2}c_{k_i}Z_{N}({\bf r}N|{\bf 0}k_i)+\nonumber\\ &&c_{k_1}c_{k_2}Z_{N}({\bf r}N|{\bf 0}k_1|{\bf 0}k_2).\end{aligned}$$ The last term of the above equation, representing correlation of the mutations, gives an additional contribution to the differential force over and above the individual contributions of the mutations. This additional contribution is negligible if the two mutations are far away or, more quantitatively, not in the same domain wall. We now use these general results for cases (Mi),(Mii). In the two dimensional model (Mi), the partition function for two directed chains having their last monomers at a relative lattice distance $x$ (along $(1,-1)$ and in unit of the elementary square diagonal), and their first monomers joined, can be written in terms of N monomer-to-monomer transfer matrices $W_j$ ($j=1,\ldots,N$) with matrix elements $\langle x'|W_j|x\rangle\equiv$ $\l(\l(\exp{\l(\beta\epsilon_j\r)}-1\r)\delta_{x',0}+1\r)$ $(2\delta_{x',x}+\delta_{x',x+1}+\delta_{x',x-1})$ $|x\rangle,|x'\rangle$ being the position vectors (with the constraint $x,x'\ge 0$). For a homo-DNA with contact energy $\epsilon$, the largest eigenvalue of $W$ determines the free energy and the thermodynamic properties in the limit $N\rightarrow\infty$. For $T<T_m\equiv\frac{\epsilon}{k_B\log{\frac{4}{3}}}$, the melting temperature, $\beta f(x)=\cosh^{-1} (\frac{1}{2z_0}-1)$ with $z_0=\sqrt{X} -X,X\equiv(1-e^{-\beta\epsilon})$. Indeed, in the fixed-stretch ensemble, any finite ($x\ll N$) stretch puts the chain on the phase coexistence curve. Exploiting the equivalence of the ensembles valid for for homo-DNA , we find $\beta f_N(x)\equiv \hat{F}(x/N)$, where \[characteristic\] (y)&=&2 \^[-1]{}({y,y\_0}),( y\_0\^), is a piecewise continuous non-analytic function. For finite $N$, there is no singularity but the approximation of Eq. \[characteristic\] still works quite well. At $x\sim y_0 N$ the force curve increases sharply (see also Fig.1, curves (a) and (c)). We now come to the explicit results for the one-mutation case where one $\epsilon$ is replaced by $\epsilon' <\epsilon$. For a homo-DNA, by starting from Eq. \[characteristic\], using Eqs. \[mut3\],\[eq:3\], one can find analytical approximations to the shapes of the previously introduced $P(x,N|0,k)$ and $\delta f$ in terms of piecewise continuous functions. We find [*e.g.*]{} $P(x,N|0,k)$ = ${P(0,N|0,k)} \exp{\l(g(x))\r)}$, where (if $Ny_0<\bar{k} \equiv N-k$): $$\begin{aligned} g(x)&=&0 \quad {\rm when}\ x<\bar{k}y_0,\\ &=&g_{\bar{k}}(x) \ {\rm if} \ \bar{k}y_0<x<Ny_0 \\ &=&g_{\bar{k}}(x)-g_N(x) \ {\rm if} \ Ny_0<x<\bar{k} \end{aligned}$$ \[eq:2\] and $g_{{k}}(x)=\log{\l[\l(\frac{1+\frac{x}{{k}}}{1+y_0}\r)^ {x+{k}} \l(\frac{1-y_0}{1-\frac{x} {{k}}}\r)^{x-{k}}\r]}$. Eq. \[mut3\] now simplifies because $f_{N}(x|0,k)=f_{\bar{k}}(x)$. The characteristics of the differential force curve $\delta f$ vs $x$, such as the extremum value, its position and the width, $\delta f_{\rm max}(k)$, $x_f(k)$ and $w_f(k)$, can be determined from Eqs. \[eq:2\] as \[results\] f\_[max]{}(k)[\~]{} \[w\_f(k)\]\^[-1]{} \~[|[k]{}]{}\^[-1/2]{}, x\_f(k) =[|[k]{}]{}[x]{}\_f(|[k]{}) with ${\tilde x}_f(0)=1$ and $\lim_{\bar{k}\to\infty}{\tilde x}_f (\bar{k})=y_0$, where $y_0$ is defined in Eq. \[characteristic\]. The area of the peak, which yields the difference (with respect to the native case) in the work necessary to unzip completely the molecule, is constant as expected. The scaling form introduced after Eq. \[eq:3\] suggests that the differential force is significant only in the domain wall region and the width of the domain wall $w_{\rm d}(k) \sim w_f(k)$ as we do see in the numerical results. For model (Mii), with one-dimensional Gaussian polymers ($T_m=\infty$), $P(x,N|0,k)$ has been calculated exactly by a transfer matrix method and is shown in Fig.2a. The validity of Eq. \[results\] ($\delta f_{\rm max}(k)\sim w_f(k)^{-1}$) is apparent from the data-collapse of the various $\delta f$ curves in Fig.2b where $\frac{\delta f}{\delta f_{\rm max}(k)}$ is plotted against $(x-x_f(k)) \delta f_{\rm max}(k)$. The peak force difference, $\delta f_{\rm max}(k)$ as a function of the mutant position $k$ is in accord with the $\bar{k}^{-1/2}$ law of Eq. \[results\]. The results for model (Mi) are shown in Fig. \[fig:3\]. For $d>1$ (${\bf r}=(x,0,\ldots,0)$ as above) the situation is similar as, [ *e.g.*]{}, the data-collapse of Fig. \[fig:2\] and Eq. \[results\] remain valid. Coming to the case of one mutation on a heterogeneous DNA consisting of two energies $\epsilon_1$ and $\epsilon_2 >\epsilon_1$ chosen with equal probability, the shapes of the zipping probability curves are found to be similar to the homo-DNA case, in fact indistinguishable on the plot of Fig. \[fig:2\], with $x_{\rm d}(k)$ and $w_{\rm d}(k)$ sequence dependent. This indicates the validity of the domain wall interpretation even for a heterogeneous DNA. The mutation involves a change of the energy at site $k$ (i.e. $\epsilon_1 \leftrightarrow \epsilon_2$). The signals $\delta f$ for various mutations are shown in Fig. \[fig:4\]. As already mentioned, in our models the sign of $\delta f$ tells us the nature of the mutation. These individual curves can again be collapsed on to a single one as for homo-DNA, though the nature of collapse is not as good, mainly because the area under the curve is no longer strictly a constant. This reflects the importance of local sequences around the mutation point. Figure 3 (curves b, c) shows the $k$-dependence of $x_f(k)$. Unlike for homo-DNA, $\delta f_{\rm max}(k)$ does not seem to have the simple form of curve (a) in Fig. 3. Although the linearity is maintained for $x_f(k)$ as for the homogeneous case, there are regions of nonmonotonicity at small scales which hamper the inversion. Fig. 3 gives a basis for a calibration curve. This could be $x_f(k)$ or $\delta f_{\rm max}$ vs k (or both) for a homogeneous DNA, though for heterogeneous DNA, we find the $x_f$-vs-$k$ curve to be more reliable. Given the value of $x_f(k)$, one can look up in Fig.3 for the corresponding $k$. The accuracy of the method relies on the ability to resolve close-by mutation points, [*i.e.*]{} mutation points in the same domain wall. There are differences in the full profiles of the $\delta f$ curves for mutations at, say, $k$, $k+1$, $k+2$, though translating that information back to the identification of the position is yet to be achieved. A better resolution is in any case obtained by changing the point at which the strands are pulled. We now argue on how our calculation can compare with a potential experiment. In [@Boc1; @Boc2], the typical force arising in the unzipping is between $10$ and $15$ pN, while the resolution is set below $0.2$ pN, so in percentage it is $<1.3-2\%$. Dynamical effects (important in [@peyrard]) are almost negligible already at the lowest stretching velocity used in [@Boc1; @Boc2] ($20$ nm/s) and are less important as the velocity is lowered. Our values for the typical force at $T=1$ are at the border of this present day resolution (see Fig. 3, where it appears that the resolution is around $1\%$ for $k\sim 500$, [*i.e.*]{} on the middle of the chains). In principle they can be improved above it by lowering $T$, though it is not clear to what extent this will work because the experimental temperature range is rather limited. In Ref.[@thompson-siggia], the authors suggest that there will be experimental difficulties which would hamper the acquisition of the base by base sequence of DNA by means of force measurements (but would however allow to get information over groups of $\sim 10$ bases). This difficulty, though absent in our exact analysis of the models, might also set a lower limit on the error on the position of the mutation. Summarizing, we prove that mutations are detectable in the theoretical models. Numbers coming from our models suggest that this measurement might be a benchmark for present day real technology. We finally propose an algorithm for sequencing DNA from the unzipping force in our models. This is defined so that the energy of the ${j}-$th base pair is $\epsilon_2$ if the average force at stretch $x=N-j$ is above the force signal of a homo-DNA with an attractive energy $\epsilon\equiv\frac{\epsilon_1+\epsilon_2}{2}$ ($T$ is low enough that $y_0\sim 1$). Our algorithm differs from the discussion in [@Boc2] because $T$ and extra constraints (see below) play crucial roles. If we define the “score” of the algorithm as the fraction of base pairs correctly predicted, from our data we observe that for any finite-size sample the score is 100% for $T<T_0(N){\sim}{N}^{-\psi}$, with $0\le\psi\le 1$ for $N\to\infty$. However, $N_0$ monomers near the open end can be sequenced at $T\simeq T_0(N_0)$, no matter how big the total $N$ is. Once this is done, we restart this time keeping the corresponding bases at position $N_0$ from the open end at a distance $x$ with the constraint that the monomers at $N$ are at a distance $x'>N_0-x$, to prevent rejoining in the already unzipped $N_0$ monomers. In this way we would sequence another $N_0$ monomers and so on. We have verified this for models (Mi) and (Mii). In conclusion, we studied the ${\bf f}$ vs. ${\bf r}$ characteristic curve in the fixed stretch ensemble for simple models of DNA focusing on the base pairing energy only. We have seen that for a homo-DNA, the force difference between a native and the corresponding one mutation case when pulled to the same distance contains enough signature to locate the position of the mutation. This could be the basis of a differential force microscope to detect mutations. For a heterogeneous DNA, the mutation point cannot be localized always as accurately as for homo-DNA. Accuracy could be achieved by taking cognizance of the full features of the $\delta f$ curve. We have shown that the differential force curve can be understood as due to the domain wall separating the zipped and the unzipped phases as the strands are pulled apart. Moreover, we found (Fig. 1) that the modulations in the force curve are connected to the local sequence. This holds promise of extension of our proposal to cases beyond point mutation. [*Acknowledgements:*]{} SMB thanks ICTP for hospitality. We thank A. Maritan and F. Seno for useful discussions. References {#references .unnumbered} ========== [99]{} email: (SMB) [email protected]  (DM) [email protected] , [*Proc. Natl. Acad. Sci. USA*]{}, [**94**]{}, [11935]{} ([1997]{}); [T. Strunz [*et al.*]{}]{}, [*Proc. Natl. Acad. Sci. USA*]{}; M. Rief, H. Clausen-Schaumann, and H.E. Gaub, [*Nature Struct. Biol.*]{} [**6**]{}, 346 (1999). , [*Phys. Rev. Lett.*]{}, [**79**]{}, 4489 (1997), [*Phys. Rev. E*]{}, [**58**]{}, 2386 (1998). S. M. Bhattacharjee cond-mat/9912297, [*J. Phys. A*]{} [**33**]{}, L423 (2000); 9003(E) (2000); cond-mat/0010132. D. K. Lubensky, D. R. Nelson, [*Phys. Rev. Lett.*]{} [**85**]{}, 1572 (2000), cond-mat/0107423; K. L. Sebastian, [*Phys. Rev. E*]{}, [**62**]{}, [1128]{},[(2000)]{}, H. Zhou, cond-mat/0007015. D. Marenduzzo [*et. al.*]{}, [*Phys. Rev. Lett.*]{} [**88**]{} 028102 (2002), D. Marenduzzo, A. Trovato, A. Maritan, [*Phys. Rev. E*]{} [**64**]{}, 031901 (2001). U. Gerland, R. Bundschuh, and T. Hwa, [*Biophys. J.*]{} [**81**]{}, 1324 (2001). D. Poland, H. A. Scheraga, [*J. Chem. Phys.*]{} [**45**]{}, 1464 (1966); [M. Peyrard, A. R. Bishop]{}, [*Phys. Rev. Lett.*]{}, [**62**]{}, [2755]{}, ([1989]{}). S. Mukherji, Phys. Rev. E [**50**]{}, R2407(1994). R. M. Neumann, [*Phys. Rev. A*]{}, [**31**]{}, R3516 (1985), R. A. Guyer, J. A. Y. Johnson, [*Phys. Rev. A*]{}, [**32**]{}, 3661 (1985), J. T. Titantah [*et al.*]{} [*Phys. Rev. E*]{}, [**60**]{}, 7010 (1999). M. Peyrard, [*Europhys. Lett.*]{}, [**44**]{}, 271 (1998). R. E. Thompson, E. D. Siggia, [*Europhys. Lett.*]{}, [**31**]{}, 335 (1995).
{ "pile_set_name": "ArXiv" }
--- author: - 'Xiaoqiang Wang$^1$, Dabin Zheng\*$^1$ and Yan Zhang$^2$' date: | 1. Hubei Province Key Laboratory of Applied Mathematics,\ Faculty of Mathematics and Statistics, Hubei University, Wuhan 430062, China\ 2. School of Computer Science and Information Engineering, Hubei University, Wuhan 430062, China title: 'Two classes of $p$-ary linear codes and their duals[^1]' --- 1.0in 1.0in [**Abstract.**]{} Let $\mathbb{F}_{p^m}$ be the finite field of order $p^m$, where $p$ is an odd prime and $m$ is a positive integer. In this paper, we investigate a class of subfield codes of linear codes and obtain the weight distribution of $$\begin{split} \mathcal{C}_k=\left\{\left(\left( {\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)+c\right)_{x \in \mathbb{F}_{p^m}}, {\rm Tr}_1^m(a)\right) : \, a,b \in \mathbb{F}_{p^m}, c \in \mathbb{F}_p\right\}, \end{split}$$ where $k$ is a nonnegative integer. Our results generalize the results of the subfield codes of the conic codes in [@Hengar]. Among other results, we study the punctured code of $\mathcal{C}_k$, which is defined as $$\mathcal{\bar{C}}_k=\left\{\left( {\rm Tr}_1^m\left(a x^{{p^k}+1}+bx\right)+c\right)_{x \in \mathbb{F}_{p^m}} : \, a,b \in \mathbb{F}_{p^m}, \,\,c \in \mathbb{F}_p\right\}.$$ The parameters of these linear codes are new in some cases. Some of the presented codes are optimal or almost optimal. Moreover, let $v_2(\cdot)$ denote the 2-adic order function and $v_2(0)=\infty$, the duals of $\mathcal{C}_k$ and $\mathcal{\bar{C}}_k$ are optimal with respect to the Sphere Packing bound if $p>3$, and the dual of $\mathcal{\bar{C}}_k$ is an optimal ternary linear code for the case $v_2(m)\leq v_2(k)$ if $p=3$ and $m>1$. [*Keywords.*]{} Linear code, subfield code, weight distribution, exponential sum, Sphere Packing bound. [*2010 Mathematics Subject Classification.*]{} 94B05, 94B15 0.0in 0.0in Introduction ============ Let $p$ be an odd prime and $\mathbb{F}_{p^m}$ be a finite field of size $p^m$. An $[n, k,d]$ code $\mathcal{C}$ over the finite field $\mathbb{F}_{p^m}$ is a $k$-dimensional linear subspace of $\mathbb{F}_{p^m}^n$ with the minimum Hamming distance $d$. An $[n,k,d]$ code is called [*distance-optimal*]{} or [*dimension-optimal*]{} if there does not exist $[n,k,d+1]$ code or $[n,k+1,d]$ code, respectively. The Hamming weight of a codeword $\mathbf{c}=(c_0, c_1, \cdots, c_{n-1}) \in \mathcal{C}$ is the number of nonzero $c_i$ for $0\leq i\leq n-1$. Let $A_i$ denote the number of nonzero codewords with Hamming weight $i$ in $\mathcal{C}$. The [*weight enumerator*]{} of $\mathcal{C}$ is defined by $1+A_1x+A_2x^2+\cdots+A_nx^n$. The sequence $(1, A_1, \cdots, A_n)$ is called the [*weight distribution*]{} of $\mathcal{C}$. The weight distribution of a code not only gives the error correcting ability of the code, but also allows the computation of the error probability of error detection and correction [@Klove2007]. Hence, the study of the weight distribution of a linear code is important in both theory and applications. The reader can refer to [@Ding2015]-[@DDing2015], [@Dinh2015] and the references therein. It is well known that the minimum Hamming distances of linear codes play an important role in measuring error-correcting performance. Thus, how to find optimal linear codes with new lengths and minimum distances is one of the central topics in coding theory. In recent years, a series of work have been done. Ding and Helleseth [@Ding2013] presented some optimal ternary cyclic codes of parameters $[3^m-1, 3m-2m-1, 4]$ according to the Sphere Packing bound by utilizing almost perfect nonlinear monomials and a number of other monomials over $\mathbb{F}_{3^m}$. Moreover, they proposed nine open problems about optimal ternary cyclic codes. Recently, three of those were solved [@Han2019; @Li2014; @Li2015] by solving the certain equations over finite fields and some new optimal ternary cyclic codes with parameters $[3^m-1, 3m-2m-1, 4]$ and $[3^m-1, 3m-2m-2,5]$ were obtained. Along this line, some other $p$-ary optimal linear codes have been studied [@Fanarxiv; @Fan2016; @Xu2016; @Zhou2019]. From these results, it is easy to see that most of the work in fining optimal linear codes focuses on constructing optimal ternary linear codes since analyzing the solutions of certain equations over $\mathbb{F}_{3}$ is easier than analyzing the solutions of certain equations over $\mathbb{F}_{p}$, where $p>3$. In this paper, we obtain a lot of $p$-ary optimal linear codes for any odd prime $p$ by using the Pless power moment identities and the weight distributions of the duals of the linear codes. To the best of our knowledge, subfield codes were first considered in [@Canteaut2000] and [@Carlet1998], but the authors do not use the name “subfield codes". The definition of subfield codes were first given by [@Cannon2013 p.5117] and a Magma function for subfield codes is actual operated in the Magma package. Recently, following the Magma definition of subfield codes, Ding and Heng in [@Dingar] proved some basic results about the subfield codes of linear codes and determined the weight distributions of the subfield codes of ovoid codes. Let ${\rm PG}(2,\mathbb{F}_{p^m})$ denote the projective plane over the finite field $\mathbb{F}_{p^m}$. An [*oval*]{} in ${\rm PG}(2,\mathbb{F}_{p^m})$ is a set of ${p^m}+1$ points such that no three of them are collinear. A [*conic*]{} in ${\rm PG}(2,\mathbb{F}_{p^m})$ is a set of ${p^m}+1$ points that are zeros of a nondegenerate quadratic form in three variables. Obviously, a conic is an oval. Segre [@Segre1955] states that in ${\rm PG}(2,\mathbb{F}_{p^m})$ every oval is also a conic if $p$ is odd. Define a conic as follows: $$\begin{split} \mathcal{O}=\{(x^2,x,1): x \in \mathbb{F}_{p^m}\}\cup\{(1,0,0)\}. \end{split}$$ Assume that $\mathbb{F}_{p^m}=\{x_1,x_2,\cdots,x_{p^m}\}$ and $f(x)$ is a polynomial over $\mathbb{F}_{p^m}$. Let $\mathcal{ C}$ be a $[p^m+1,3]$ linear code with the generator matrix $$\label{eq:matrixA'} \begin{split} G_f = \left( \begin{array}{cccccc} f(x_1) & f(x_2) & \cdots & f(x_{p^m}) & 1 \\ x_1 & x_2 & \cdots & x_{p^m} & 0 \\ 1 & 1 & \cdots & 1 & 0 \end{array} \right). \end{split}$$ In [@Hengar], Heng and Ding constructed a class of conic codes of length $p^m+1$ over $\mathbb{F}_{p^m}$ with generator matrix in (\[eq:matrixA’\]) for $f(x)=x^2$, and determined the weight distributions of the subfield codes of the conic codes. Along the line of the work in [@Hengar], this paper constructs a class of linear codes of length $p^m+1$ over $\mathbb{F}_{p^m}$ with generator matrix in (\[eq:matrixA’\]) for $f(x)=x^{p^k+1}$, and determines the weight distributions of the subfield codes of these linear codes, where the subfield code is defined as $$\label{code0} \begin{split} \mathcal{C}_k=\left\{\left(\left( {\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)+c\right)_{x \in \mathbb{F}_{p^m}}, {\rm Tr}_1^m(a)\right) : \, a,b \in \mathbb{F}_{p^m}, c \in \mathbb{F}_p\right\}. \end{split}$$ Moreover, from the weight distribution of $\mathcal{C}_k$, we obtain the weight distribution of linear code $$\mathcal{\bar{C}}_k=\left\{\left( {\rm Tr}_1^m\left(a x^{{p^k}+1}+bx\right)+c\right)_{x \in \mathbb{F}_{p^m}} : \, a,b \in \mathbb{F}_{p^m}, \,\,c \in \mathbb{F}_p\right\}.$$ Hence, we generalize the results about the weight distributions of the subfield codes of conic codes in [@Hengar]. The parameters of these $p$-ary linear codes are new in some cases. Some of the codes presented are optimal or almost optimal. Moreover, from the weight distributions of $\mathcal{C}_k$ and $\mathcal{\bar{C}}_k$ and the Pless power moment identities, we obtian that the duals of $\mathcal{C}_k$ and $\mathcal{\bar{C}}_k$ are optimal with respect to the Sphere Packing bound for many cases. The remainder of this paper is organized as follows. Section $2$ recalls some preliminary results. In Section $3$ $($Section $4$$)$, the weight distribution of $\mathcal{C}_k$ and the dual of the parameters of $\mathcal{C}_k$ for the case $v_2(m)\leq v_2(k)$ $(v_2(m)> v_2(k))$ are determined, where $v_2(\cdot)$ denotes the 2-adic order function and $v_2(0)=\infty$. In Section $5$, from the weight distribution of $\mathcal{C}_k$ and the Sphere Packing bound, we obtain the weight distribution of $\mathcal{\bar{C}}_k$ and the parameters of the dual of $\mathcal{\bar{C}}_k$. Section 6 concludes the paper. Preliminaries ============= Let $\mathbb{F}_{p^m}$ be a finite field and $\mathbb{F}_{p^m}^* = \mathbb{F}_{p^m} \setminus \{0\}$. Let $\psi$ be a multiplicative character of $\mathbb{F}_{p^m}^*$ and $\chi$ be a canonical additive character of $\mathbb{F}_{p^m}$. The [*Gaussian sum*]{} $G(\psi, \chi)$ is defined by $$G(\psi, \chi) = \sum_{x\in \mathbb{F}_{p^m}^*} \psi(x)\chi(x) .$$ It is very difficult to determine explicit values of Gaussian sums in general. Until now, Gaussian sums are known only for a few special cases. The quadratic Gaussian sums are known and given in the following lemma. [@Lidl1983 Theorem 5.15]\[lemGxn\] Let $\mathbb{F}_{p^m}$ be a finite field. Then $$G(\eta,\chi)=\left\{ \begin{array}{lcl} (-1)^{m-1}p^{\frac{m}{2}}, & {\rm if}\, \,\, p\equiv 1 \pmod 4,\\ (-1)^{m-1}i^mp^{\frac{m}{2}}, & {\rm if}\, \,\, p\equiv 3 \pmod 4, \end{array} \right.$$ where $\eta$ is the quadratic multiplicative character of $\mathbb{F}_{p^m}$. Let $\mathbb{F}_q$ denote the finite field with $q$ elements, where $q$ is an odd prime power. By identifying the finite field $\mathbb{F}_{q^s}$ with an $s$-dimensional vector space $\mathbb{F}_{q}^s$ over $\mathbb{F}_{q}$, a function $f$ from $\mathbb{F}_{q^s}$ to $\mathbb{F}_q$ can be viewed as an $s$-variable polynomial over $\mathbb{F}_{q}$. The function $f(x)$ is called a quadratic form if it is a homogenous polynomial of degree two as follows: $$f(x_1, x_2, \cdots, x_s) = \sum_{1\leq i\leq j\leq s} a_{ij} x_ix_j , \,\, \, a_{ij}\in \mathbb{F}_{q},$$ where we fix a basis of $\mathbb{F}_q^s$ over $\mathbb{F}_q$ and identify $x\in \mathbb{F}_{q^s}$ with a vector $(x_1, x_2, \cdots, x_s)\in \mathbb{F}_q^s$. The rank of the quadratic form $f(x)$ is defined as the codimension of $\mathbb{F}_q$-vector space $$V = \{ x\in \mathbb{F}_q^s\, \, |\,\, f(x+z)-f(x)-f(z) = 0, \,\, \mbox{for all}\,\, z\in \mathbb{F}_{q}^s \} ,$$ which is denoted by rank$(f)$. Then $|V| = q^{s-{\rm rank}(f)}$. The following lemma gives a general result on an exponential sum of a quadratic function from $\mathbb{F}_{p^m}$ to $\mathbb{F}_p$. \[lem:quadraticsum\] ([@Liu2018 Lemma 2.1]) Let $Q(x)$ be a quadratic function from $\mathbb{F}_{p^m}$ to $\mathbb{F}_p$ with rank $r(r\neq0)$, $\zeta_p$ be a $p$-th primitive root of unity and $\eta$ be the quadratic multiplicative character of $\mathbb{F}_{p^m}$. Then for any $z\in \mathbb{F}_{p}^*$, $$\sum_{x\in \mathbb{F}_{p^m}} \zeta_p^{z Q(x)}=\eta^r(z)\sum_{x\in \mathbb{F}_{p^m}} \zeta_p^{ Q(x)}.$$ The following is a well known result about quadratic forms. \[cor:quadraticsum\][@Draper2007 Corollary 7.6] Let $v_2(\cdot)$ denote the 2-adic order function and $v_2(0)=\infty$. Let $a \in \mathbb{F}_{p^m}^*$ and $k$ be a nonnegative integer. Assume that $Q(x)={\rm Tr}_1^m\left(ax^{p^k+1}\right)$. [(i)]{} If $v_2(m)\leq v_2(k)$, $$\begin{split} \sum_{x\in \mathbb{F}_{p^m}} \zeta_p^{Q(x)}=\eta(a)(-1)^{m-1}i^{\frac{(p-1)^2m}{2}}p^{\frac{m}{2}}. \end{split}$$ [(ii)]{} If $v_2(m)=v_2(k)+1$, $$\begin{split} \sum_{x\in \mathbb{F}_{p^m}} \zeta_p^{Q(x)}= \begin{cases} p^{\frac{m+\gcd(2k,m)}{2}}, & \text{if\,\,\,\, $a^{\frac{(p^k-1)(p^m-1)}{p^{\gcd(2k,m)}-1}}=-1$,} \\ -p^{\frac{m}{2}}, & \text{otherwise}. \end{cases} \end{split}$$ [(iii)]{} If $v_2(m)>v_2(k)+1$, $$\begin{split} \sum_{x\in \mathbb{F}_{p^m}} \zeta_p^{Q(x)}= \begin{cases} -p^{\frac{m+\gcd(2k,m)}{2}}, & \text{if\,\,\,\, $a^{\frac{(p^k-1)(p^m-1)}{p^{\gcd(2k,m)}-1}}=1$,} \\ p^{\frac{m}{2}}, & \text{otherwise}. \end{cases} \end{split}$$ Next, we recall some known results about the subfield codes, which were introduced in [@Dingar]. Let $\mathcal{C}$ be an $[n,k]$ code over $\mathbb{F}_{q^m}$ with the generator matrix $$\begin{split} G= \left( \begin{array}{cccc} g_{11} & g_{12} & \cdots & g_{1n} \\ g_{21} & g_{22} & \cdots & g_{2n} \\ \vdots & \vdots & \ddots & \vdots\\ g_{k1} & g_{k2} & \cdots & g_{kn} \end{array} \right). \end{split}$$ Assume that $\{\alpha_1, \alpha_2, \cdots, \alpha_m\}$ is a basis of $\mathbb{F}_{q^m}$ over $\mathbb{F}_q$. Ding and Heng in [@Dingar] constructed a new $[n,k']$ code $\mathcal{C}^{(q)}$ over $\mathbb{F}_q$ with the generator matrix $$\begin{split} G^{(q)}= \left( \begin{array}{cccc} G_1^{(q)} \\ G_2^{(q)} \\ \vdots \\ G_k^{(q)} \end{array} \right), \end{split}$$ where $$\begin{split} \ G^{(q)}_i = \left( \begin{array}{cccc} {\rm Tr}_1^m(g_{i1}\alpha_1) & {\rm Tr}_1^m(g_{i2}\alpha_1) & \cdots & {\rm Tr}_1^m(g_{in}\alpha_1) \\ {\rm Tr}_1^m(g_{i1}\alpha_2) & {\rm Tr}_1^m(g_{i2}\alpha_2) & \cdots & {\rm Tr}_1^m(g_{in}\alpha_2) \\ \vdots & \vdots & \ddots & \vdots\\ {\rm Tr}_1^m(g_{i1}\alpha_m) & {\rm Tr}_1^m(g_{i2}\alpha_m) & \cdots & {\rm Tr}_1^m(g_{in}\alpha_m) \\ \end{array} \right). \end{split}$$ They called $\mathcal{C}^{(q)}$ the [*subfield code*]{} of $\mathcal{C}$ and gave the trace representation of $\mathcal{C}^{(q)}$ as follows. \[subfieldcode\] Let $\mathcal{C}$ be an $[n,k]$ code over $\mathbb{F}_{q^m}$. Let $G=[g_{i,j}]_{1\leq i \leq k, 1\leq j \leq n}$ be a generator matrix of $\mathcal{C}$. Then the trace representation of the subfield code $\mathcal{C}^{(q)}$ is given by $$\mathcal{C}^{(q)}=\left\{\left( {\rm Tr}_1^m\left(\sum_{i=1}^ka_ig_{i1}\right),\cdots, {\rm Tr}_1^m\left(\sum_{i=1}^ka_ig_{i1}\right)\right)\,:\, a_1,\cdots, a_k \in \mathbb{F}_{q^m}\right\}.$$ Let $q=p$ be an odd prime and the generator matrix of $\mathcal{C}$ be defined in (\[eq:matrixA’\]). From Lemma \[subfieldcode\] we have the trace representation of the subfield code of $\mathcal{C}$ as follows: $$\begin{split} \mathcal{C}_f=\left\{\left(\left( {\rm Tr}_1^m(af(x)+bx)+c\right)_{x \in \mathbb{F}_{p^m}}, {\rm Tr}_1^m(a)\right)\, : \, a,b \in \mathbb{F}_{p^m}, c \in \mathbb{F}_p\right\}. \end{split}$$ When $a=0$, the Hamming weights and their corresponding frequencies of the codewords of the form $\mathbf{c}(0,b,c)$ can be easily determined as follows. If $a=b=0$, then $$\begin{split} \mathbf{c}(0,0,c)=(\underbrace{c,c, \cdots, c}_{\small{p^m}},0), \end{split}$$ i.e., $$\label{111} \begin{split} \operatorname{wt_H}(\mathbf{c}(0,0,c))= \begin{cases} 0, & \text{occur\,\,\,\, $1$\,\,\,\,\,\,\, time,} \\ p^m, & \text{occur $p-1$ times}. \end{cases} \end{split}$$ If $a=0$, $b\neq 0$, then $$\begin{split} \mathbf{c}(0,b,c)=\left(\left({\rm Tr}_1^m(bx)+c\right)_{x \in \mathbb{F}_{p^m}}, 0\right). \end{split}$$ It is easy for us to get that $$\label{112} \begin{split} \operatorname{wt_H}(\mathbf{c}(0,b,c))= p^{m-1}(p-1)\,\,\, & \text{occur} \,\, p(p^m-1) \,\, \text{times}. \end{split}$$ When $a\neq0$, to determine the Hamming weights and their corresponding frequencies of $\mathbf{c}(a,b,c)$ in $\mathcal{C}_f$ for $(a,b,c)$ running through $(\mathbb{F}_{p^m}^*, \mathbb{F}_{p^m}, \mathbb{F}_p)$ is a very hard problem for the general polynomial $f(x)$. In the following sections, we always assume that $f(x)=x^{p^k+1}$, where $k$ is a nonnegative integer. In order to obtain the parameters of the dual codes of the discussed subfield codes, we need the Pless power moment identities on linear codes. Let $\mathcal{C}$ be a $p$-ary $[n, k]$ code over $\mathbb{F}_p$, and denote its dual by $\mathcal{C}^{\perp}$. Let $A_i$ and $A^{\perp}_i$ be the number of codewords of weight $i$ in $\mathcal{C}$ and $\mathcal{C}^{\perp}$, respectively. The first five Pless power moment identities are as follows ([@MacWilliam1997], p. 131): $$\begin{split} &\sum_{i=0}^nA_i=p^k;\\ &\sum_{i=0}^niA_i=p^{k-1}(pn-n-A_1^{\perp});\\ &\sum_{i=0}^ni^2A_i=p^{k-2}[(p-1)n(pn-n+1)-(2pn-p-2n+2)A_1^{\perp}+2A_2^{\perp}];\\ &\sum_{i=0}^ni^3A_i=p^{k-3}[(p-1)n(p^2n^2-2pn^2+3pn-p+n^2-3n+2)-(3p^2n^2-3p^2n-6pn^2+12pn\\ &\hskip 48pt+p^2-6p+3n^2-9n+6)A_1^{\perp}+6(pn-p-n+2)A_2^{\perp}-6A_3^{\perp}];\\ &\sum_{i=0}^ni^4A_i=p^{k-4}[(p-1)n(p^3n^3-3p^2n^3+6p^2n^2-4p^2n+p^2+3pn^3-12pn^2+15pn-6p-n^3+6n^2-11n+6)\\ &\hskip 48pt-(4p^3n^3-6p^3n^2+4p^3n-p^3-12p^2n^3+36p^2n^2-38p^2n+14p^2+12pn^3-54pn^2+78pn-36p\\ &\hskip 48pt-4n^3+24n^2-44n+24)A_1^{\perp}+(12p^2n^2-24p^2n+14p^2-24pn^2+84pn-72p+12n^2-60n+72)A_2^{\perp}\\ &\hskip 48pt-(24pn-36p-24n+72)A_3^{\perp}+24A_4^{\perp}]. \end{split}$$ The following two lemmas on the bound of linear codes are well-known. (Sphere Packing bound) Let $\mathcal{C}$ be a $p$-ary $[n, k,d]$ code. Then $$p^n\geq p^k\sum_{i=0}^{\lfloor \frac{d-1}{2} \rfloor}\left( \begin{array}{cccc} n \\ i \\ \end{array} \right)(p-1)^i.$$ [@Rouayheb2007]\[bound2\] Let $q$ be an odd prime power and $A_q(n,d)$ be the maximum number of codewords of a $q$-ary code with length $n$ and Hamming distance at least $d$. If $q\geq 3$, $t=n-d+1$ and $r=\lfloor min\{\frac{n-t}{2}, \frac{t-1}{q-2}\}\rfloor$, then $$\begin{split} A_q(n,d)\leq \frac{q^{t+2r}}{\sum_{i=0}^r\left(\begin{array}{cccc} t+2r \\ i \\ \end{array} \right)(q-1)^i}. \end{split}$$ The weight distribution of $\mathcal{C}_k$ for $v_2(m)\leq v_2(k)$ ================================================================== Through out this section, let $\eta$ and $\eta_0$ denote the quadratic multiplicative character of $\mathbb{F}_{p^m}$ and $\mathbb{F}_p$, respectively. Our main task is to determine the weight distribution of $\mathcal{C}_k$ for the case $v_2(m)\leq v_2(k)$ and give the parameters of the dual of $\mathcal{C}_k$, where $\mathcal{C}_k$ is defined in (\[code0\]). We start with the following lemma, which is given by [@Coulter2002] and [@Lidl1983]. [@Coulter2002; @Lidl1983]\[eqSab\] Let $v_2(\cdot)$ denote the 2-adic order function and $v_2(0)=\infty$. Let $a\in \mathbb{F}_{p^m}^*$, $b \in \mathbb{F}_{p^m}$, and $k,m$ be two integers such that $v_2(m)\leq v_2(k)$. Let $$\begin{split} S(a,b)=\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{{\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)}, \end{split}$$ and $x_b$ be the unique solution of the equation $a^{p^k}x^{p^{2k}}+ax+b^{p^k}=0$. Then $$S(a,b)=\left\{ \begin{array}{lcl} (-1)^{m-1}p^{\frac{m}{2}}\eta(-a)\zeta_p^{{\rm Tr}_1^m\left(-ax_b^{p^k+1}\right)}, & {\rm if}\, \,\, p\equiv 1 \pmod 4,\\ (-1)^{m-1}p^{\frac{m}{2}}i^{3m}\eta(-a)\zeta_p^{{\rm Tr}_1^m\left(-ax_b^{p^k+1}\right)}, & {\rm if}\, \,\, p\equiv 3 \pmod 4, \end{array} \right.$$ where $\zeta_p$ is a $p$-th primitive root of unity. For any $a \in \mathbb{F}_{p^m}^*$, it is clear that $a^{p^k}x^{p^{2k}}+ax=0$ is a linearized permutation polynomial over $\mathbb{F}_{p^m}$ for the case $v_2(m)\leq v_2(k)$. Hence, the equation $a^{p^k}x^{p^{2k}}+ax+b^{p^k}=0$ has a unique solution. Now, we determine the possible values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ for $(a,b,c)$ running through $(\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m}, \mathbb{F}_p)$. \[pro:ww\] Let $a\in \mathbb{F}_{p^m}^*$, $b \in \mathbb{F}_{p^m}$ and $c \in \mathbb{F}_p$. Let $x_b$ be the unique solution of the equation $a^{p^k}x^{p^{2k}}+ax+b^{p^k}=0$. Assume that $v_2(m)\leq v_2(k)$ and $c_0=c-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)$. If $v_2(m)=0$, then the possible distinct values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ are $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c))=\begin{cases} (p-1)p^{m-1}, & \text{if \,\,${\rm Tr}_1^m(a)=0$ and $c_0=0$}, \\ (p-1)p^{m-1}+1, & \text{if \,\,${\rm Tr}_1^m(a)\neq 0$ and $c_0= 0$}, \\ (p-1)p^{m-1}-p^{\frac{m-1}{2}}, & \text{if \,\,${\rm Tr}_1^m(a)=0$ and $\eta(ac_0)=1$,} \\ (p-1)p^{m-1}-p^{\frac{m-1}{2}}+1, & \text{if \,\,${\rm Tr}_1^m(a)\neq 0$ and $\eta(ac_0)=1$,} \\ (p-1)p^{m-1}+p^{\frac{m-1}{2}}, & \text{if \,\,${\rm Tr}_1^m(a)=0$ and $\eta(ac_0)=-1$,} \\ (p-1)p^{m-1}+p^{\frac{m-1}{2}}+1, & \text{if \,\,${\rm Tr}_1^m(a)\neq 0$ and $\eta(ac_0)=-1$.} \\ \end{cases} \end{split}$$ If $0<v_2(m)\leq v_2(k)$, then the possible distinct values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ are $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c))=\begin{cases} (p-1)p^{m-1}-\varepsilon_0(p-1)p^{\frac{m-2}{2}}, & \text{if \,\,${\rm Tr}_1^m(a)=0$, $\eta(a)=-1$ and $c_0=0$,} \\ (p-1)p^{m-1}-\varepsilon_0(p-1)p^{\frac{m-2}{2}}+1, & \text{if \,\,${\rm Tr}_1^m(a)\neq0$, $\eta(a)=-1$ and $c_0=0$,} \\ (p-1)p^{m-1}-\varepsilon_0p^{\frac{m-2}{2}}, & \text{if \,\,${\rm Tr}_1^m(a)=0$, $\eta(a)=1$ and $c_0\neq0$,} \\ (p-1)p^{m-1}-\varepsilon_0p^{\frac{m-2}{2}}+1, & \text{if \,\,${\rm Tr}_1^m(a)\neq0$, $\eta(a)=1$ and $c_0\neq0$,} \\ (p-1)p^{m-1}+\varepsilon_0(p-1)p^{\frac{m-2}{2}}, &\text{if \,\,${\rm Tr}_1^m(a)=0$, $\eta(a)=1$ and $c_0=0$,} \\ (p-1)p^{m-1}+\varepsilon_0(p-1)p^{\frac{m-2}{2}}+1, & \text{if \,\,${\rm Tr}_1^m(a)\neq0$, $\eta(a)=1$ and $c_0=0$,} \\ (p-1)p^{m-1}+\varepsilon_0p^{\frac{m-2}{2}}, & \text{if \,\,${\rm Tr}_1^m(a)=0$, $\eta(a)=-1$ and $c_0\neq0$,} \\ (p-1)p^{m-1}+\varepsilon_0p^{\frac{m-2}{2}}+1, & \text{if \,\,${\rm Tr}_1^m(a)\neq0$, $\eta(a)=-1$ and $c_0\neq0$,} \\ \end{cases} \end{split}$$ where $\varepsilon_0=(-1)^{\frac{(p-1)m}{4}}$. [*Proof*]{}. The weight of the codeword $\mathbf{c}(a,b,c)$ is related to the exponential sum $S(a,b)$. According to Lemma \[eqSab\], we only discuss the case $p \equiv 1\pmod 4$, and the case $p \equiv 3\pmod 4$ can be shown by the similar way. By the definition of $\mathcal{C}_k$, the last entry of codeword $\mathbf{c}(a,b,c)$ is $0$ (or not $0$) if ${\rm Tr}_1^m(a)=0$ ( or ${\rm Tr}_1^m(a)\neq 0$). So, we prove the theorem according to the following four cases. [**Case 1**]{}: $v_2(m)=0$ and ${\rm Tr}_1^m(a)=0$. By Lemma \[eqSab\] we have $$\label{eq:cabc1} \begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c))&=p^m-\frac{1}{p}\sum_{z \in \mathbb{F}_p}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{z\left({\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)+c\right)}\\ &=(p-1)p^{m-1}-\frac{1}{p}\sum_{z \in \mathbb{F}_p^*}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{z\left({\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)+c\right)}\\ &=(p-1)p^{m-1}-\frac{1}{p}\sum_{z \in \mathbb{F}_p^*}\zeta_p^{zc}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{{\rm Tr}_1^m\left(zax^{p^k+1}+zbx\right)}\\ &=(p-1)p^{m-1}-p^{\frac{m-2}{2}}\sum_{z \in \mathbb{F}_p^*}\eta(-za)\zeta_p^{z\left(c-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)\right)}. \end{split}$$ If $c={\rm Tr}_1^m\left(ax_b^{p^k+1}\right)$, then $\operatorname{wt_H}(\mathbf{c}(a,b,c))=(p-1)p^{m-1}$. If $c\neq{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)$, recalling that $c_0=c-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)$, by Lemma \[lemGxn\] and (\[eq:cabc1\]) we have $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c))&=(p-1)p^{m-1}-p^{\frac{m-2}{2}}\sum_{z \in \mathbb{F}_p^*}\eta(-za)\zeta_p^{zc_0}\\ &=(p-1)p^{m-1}-p^{\frac{m-2}{2}}\sum_{z \in \mathbb{F}_p^*}\eta(-za)\eta^2 (c_0)\zeta_p^{zc_0}\\ &=(p-1)p^{m-1}-p^{\frac{m-2}{2}}\eta(-ac_0)\sum_{z \in \mathbb{F}_p^*}\eta(zc_0)\zeta_p^{zc_0}\\ &=(p-1)p^{m-1}-p^{\frac{m-2}{2}}\eta(ac_0)\sum_{z \in \mathbb{F}_p^*}\eta_0(zc_0)\zeta_p^{zc_0}\\ &=(p-1)p^{m-1}-p^{\frac{m-2}{2}}\eta(ac_0)G(\eta_0,\chi_0)\\ &=\begin{cases} (p-1)p^{m-1}-p^{\frac{m-1}{2}}, & \text{if $\eta(ac_0)=1$,} \\ (p-1)p^{m-1}+p^{\frac{m-1}{2}}, & \text{if $\eta(ac_0)=-1$,} \\ \end{cases} \end{split}$$ where $\chi_0$ is a canonical additive character of $\mathbb{F}_{p}$. In the fourth equality, we used the fact that $\eta_0(zc_0)=\eta(zc_0)$ and $\eta(-1)=1$ since $m$ is odd and $p\equiv1 \pmod 4$. [**Case 2**]{}: $v_2(m)=0$ and ${\rm Tr}_1^m(a)\neq0$. Then $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ can be expressed as $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c))=p^m+1-\frac{1}{p}\sum_{z \in \mathbb{F}_p}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{z\left({\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)+c\right)}. \end{split}$$ Using the same way as in Case $1$, we have $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c)) &=\begin{cases} (p-1)p^{m-1}+1, & \text{if $c_0=0$,} \\ (p-1)p^{m-1}-p^{\frac{m-1}{2}}+1, & \text{if $\eta(ac_0)=1$,} \\ (p-1)p^{m-1}+p^{\frac{m-1}{2}}+1, & \text{if $\eta(ac_0)=-1$.} \\ \end{cases} \end{split}$$ [**Case 3**]{}: $0<v_2(m)\leq v_2(k)$ and ${\rm Tr}_1^m(a)=0$. By Lemma \[eqSab\] and (\[eq:cabc1\]) we have $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c))&=(p-1)p^{m-1}+p^{\frac{m-2}{2}}\sum_{z \in \mathbb{F}_p^*}\eta(-za)\zeta_p^{z\left(c-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)\right)}\\ &=(p-1)p^{m-1}+p^{\frac{m-2}{2}}\eta(a)\sum_{z \in \mathbb{F}_p^*}\zeta_p^{z\left(c-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)\right)}\\ &=\begin{cases} (p-1)p^{m-1}+(p-1)p^{\frac{m-2}{2}}, & \text{if $\eta(a)=1$ and $c_0 =0$,} \\ (p-1)p^{m-1}-(p-1)p^{\frac{m-2}{2}}, & \text{if $\eta(a)=-1$ and $c_0 =0$,} \\ (p-1)p^{m-1}-p^{\frac{m-2}{2}}, & \text{if $\eta(a)=1$ and $c_0 \neq0$,} \\ (p-1)p^{m-1}+p^{\frac{m-2}{2}}, & \text{if $\eta(a)=-1$ and $c_0 \neq0$.} \\ \end{cases} \end{split}$$ In the second equality, we used the fact that $\eta(-z)=1$ for $z \in \mathbb{F}_p^*$ since $m$ is even. [**Case 4**]{}: $0<v_2(m)\leq v_2(k)$ and ${\rm Tr}_1^m(a)\neq0$. Using the same way as in Case $3$, we have $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c)) =\begin{cases} (p-1)p^{m-1}+(p-1)p^{\frac{m-2}{2}}+1, & \text{if $\eta(a)=1$ and $c_0 =0$,} \\ (p-1)p^{m-1}-(p-1)p^{\frac{m-2}{2}}+1, & \text{if $\eta(a)=-1$ and $c_0 =0$,} \\ (p-1)p^{m-1}-p^{\frac{m-1}{2}}+1, & \text{if $\eta(a)=1$ and $c_0 \neq0$,} \\ (p-1)p^{m-1}+p^{\frac{m-1}{2}}+1, & \text{if $\eta(a)=-1$ and $c_0 \neq0$.} \\ \end{cases} \end{split}$$ From Case $1$, Case $2$, Case $3$ and Case $4$, the result follows. $\square$ In the following, we determine the weight distribution of $\mathcal{C}_k$. To this end, we first determine the lower bound of the minimum Hamming distance of $\mathcal{C}_k^{\perp}$. Let $\mathcal{C}_k$ be the linear code defined in (\[code0\]), then the minimum Hamming distance of $\mathcal{C}_k^{\perp}$ is greater than or equal to $3$. [*Proof.*]{} We only need to prove that there does not exist a codeword $\mathbf{c}^{\perp}\in\mathcal{C}_k^{\perp}$ with the Hamming weight $1$ and $2$. There are following cases. [**Case 1**]{}: Assume that $a=b=0$ and $c =1$ in (\[code0\]), then $\mathbf{c}=(\mathbf{1},0) \in \mathcal{C}_k,$ where $\mathbf{1}$ is a vector of length $p^m$ with all entries being $1$. If there exists a codeword $\mathbf{c}^{\perp}\in \mathcal{C}_k^{\perp}$ with Hamming weight $1$, then $\mathbf{c}^{\perp}$ must have the form $(\mathbf{0},\alpha) \in \mathcal{C}_k^{\perp},$ where $\mathbf{0}$ is a vector of length $p^m$ with all entries being $0$ and $\alpha \in \mathbb{F}_p^*$. It is obvious that for any ${\rm Tr}_1^m(a)\neq 0$, there is a codeword $(\mathbf{u},{\rm Tr}_1^m(a)) \in \mathcal{C}_k$ such that $(\mathbf{u},{\rm Tr}_1^m(a))(\mathbf{0},\alpha)\neq 0$, where $\mathbf{u}$ is a vector of length $p^m$. Hence, there is not a codeword in $\mathcal{C}_k^{\perp}$ with Hamming weight $1$. [**Case 2**]{}: Assume that there is a codeword $\mathbf{c}^{\perp} \in \mathcal{C}^{\perp}$ with ${\rm wt_H}(\mathbf{c}^{\perp})=2$. Obviously, $\mathbf{c}^{\perp}$ must have the form $(\mathbf{0},\alpha,\mathbf{0},-\alpha,\mathbf{0})$ since $(\mathbf{1},0) \in \mathcal{C}$, where $\alpha \in \mathbb{F}_p^*$. So, let $c=0$, for any $a,b \in \mathbb{F}_{p^m}$, there exist two different elements $\beta,\gamma \in \mathbb{F}_{p^m}$ such that $${\rm Tr}_1^m\left(a\beta^{p^k+1}+b\beta\right)=-{\rm Tr}_1^m\left(a\gamma^{p^k+1}+b\gamma\right).$$ This means that $$\label{gamma} \begin{split} {\rm Tr}_1^m\left(a\left(\beta^{p^k+1}+\gamma^{p^k+1}\right)+b(\beta+\gamma)\right)=0 \end{split}$$ for any $a,b \in \mathbb{F}_{p^m}.$ It is easy to see that (\[gamma\]) holds if and only if $$\begin{split} \begin{cases} \beta^{p^k+1}+\gamma^{p^k+1}=0, \\ \beta+\gamma=0, \end{cases} \end{split}$$ which is impossible. Combining with Case $1$ and Case $2$, the result follows. $\square$ The following result [@Hengar] is useful for us to determine the weight distribution of $\mathcal{C}_k$. [@Hengar Lemma 14]\[lemeven\] Let $\mathbb{F}_{p^m}$ be the finite field of $p^m$ elements, where $p$ is an odd prime. Let $\eta$ be the quadratic multiplicative character of $\mathbb{F}_{p^m}^*$. If $m$ is even, then $$\begin{split} \left|\left\{a \in \mathbb{F}_{p^m}^*\,|\, \eta(a)=\pm 1\,\, and\,\, {\rm Tr}_1^m(a)=0\right\}\right|= \frac{p^{m-1}-1\mp (p-1)p^{\frac{m-2}{2}}(-1)^{\frac{(p-1)m}{4}}}{2}, \end{split}$$ and $$\begin{split} \left|\{a \in \mathbb{F}_{p^m}^*\,|\, \eta(a)=\pm1\,\, and\,\, {\rm Tr}_1^m(a)\neq0\}\right|= \frac{(p-1)\left(p^{m-1}\pm p^{\frac{m-2}{2}}(-1)^{{\frac{(p-1)m}{4}}}\right)}{2}. \end{split}$$ With above preparations, we now give the weight distribution of $\mathcal{C}_k$ for the case $v_2(m)\leq v_2(k)$. \[theorem1\] Let $k,m$ be two integers such that $v_2(m)\leq v_2(k)$. Let $\mathcal{C}_k$ be the linear code defined in (\[code0\]). [(i)]{} If $v_2(m)=0$, then $\mathcal{C}_k$ is a $\left[p^m+1,2m+1,(p-1)p^{m-1}-p^{\frac{m-1}{2}}\right]$ code with the weight distribution in Table \[Table1\]. Its dual has parameters $[p^m+1,p^m-2m,4]$, which is optimal with respect to the Sphere Packing bound for $p\geq 5$. Weight Multiplicity --------------------------------------- --------------------------- -- -- $0$ $1$ $p^m$ $p-1$ $(p-1)p^{m-1}$ $p(p^m-1)+p^m(p^{m-1}-1)$ $(p-1)p^{m-1}+1$ $p^m(p^m-p^{m-1})$ $(p-1)p^{m-1}\pm p^{\frac{m-1}{2}}$ $p^m(p^{m-1}-1)(p-1)/2$ $(p-1)p^{m-1}\pm p^{\frac{m-1}{2}}+1$ $p^{2m-1}(p-1)^2/2$ : The weight distribution of $\mathcal{C}_k$ for $v_2(m)=0$ []{data-label="Table1"} [(ii)]{} If $0<v_2(m)\leq v_2(k)$, then $\mathcal{C}_k$ is a $\left[p^m+1,2m+1,(p-1)\left(p^{m-1}-p^{\frac{m-2}{2}}\right)\right]$ code with the weight distribution in Table \[Table2\]. Its dual has parameters $[p^m+1,p^m-2m,4]$, which is optimal with respect to the Sphere Packing bound for $p\geq 5$. Weight Multiplicity -------------------------------------------------------------- ------------------------------------------------------------------------ -- -- $0$ $1$ $p^m$ $p-1$ $(p-1)p^{m-1}$ $p(p^m-1)$ $(p-1)\left(p^{m-1}+\varepsilon_0p^{\frac{m-2}{2}}\right)$ $p^m\left(p^{m-1}-1-\varepsilon_0(p-1)p^{\frac{m-2}{2}}\right)/2$ $(p-1)\left(p^{m-1}+\varepsilon_0p^{\frac{m-2}{2}}\right)+1$ $p^m(p-1)\left(p^{m-1}+\varepsilon_0p^{\frac{m-2}{2}}\right)/{2}$ $(p-1)\left(p^{m-1}-\varepsilon_0p^{\frac{m-2}{2}}\right)$ $p^m\left(p^{m-1}-1+\varepsilon_0(p-1)p^{\frac{m-2}{2}}\right)/2$ $(p-1)\left(p^{m-1}-\varepsilon_0p^{\frac{m-2}{2}}\right)+1$ $p^m(p-1)\left(p^{m-1}-\varepsilon_0p^{\frac{m-2}{2}}\right)/2$ $(p-1)p^{m-1}-\varepsilon_0p^{\frac{m-2}{2}}$ $p^m(p-1)\left(p^{m-1}-1-\varepsilon_0(p-1)p^{\frac{m-2}{2}}\right)/2$ $(p-1)p^{m-1}-\varepsilon_0p^{\frac{m-2}{2}}+1$ $p^m(p-1)^2\left(p^{m-1}+\varepsilon_0p^{\frac{m-2}{2}}\right)/2$ $(p-1)p^{m-1}+\varepsilon_0p^{\frac{m-2}{2}}$ $p^m(p-1)\left(p^{m-1}-1+\varepsilon_0(p-1)p^{\frac{m-2}{2}}\right)/2$ $(p-1)p^{m-1}+\varepsilon_0p^{\frac{m-2}{2}}+1$ $p^m(p-1)^2\left(p^{m-1}-\varepsilon_0p^{\frac{m-2}{2}}\right)/2$ : The weight distribution of $\mathcal{C}_k$ for $0<v_2(m)\leq v_2(k)$ []{data-label="Table2"} where $\varepsilon_0=(-1)^{\frac{(p-1)m}{4}}$. [*Proof.*]{} We only prove the case $p \equiv 1\pmod 4$, and the case $p \equiv 3\pmod 4$ can be shown by the similar way. There are two cases to be discussed. [**Case 1**]{}: $v_2(m)=0$. From (\[111\]), (\[112\]) and Proposition \[pro:ww\], when $(a,b,c)$ runs over $(\mathbb{F}_{p^m}, \mathbb{F}_{p^m}, \mathbb{F}_p)$, the number for $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ being $p^m$ is $p-1$. We now determine the multiplicities of the other possible values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ for $(a,b,c)$ running through $(\mathbb{F}_{p^m},\mathbb{F}_{p^m}, \mathbb{F}_p)$. Define $$\label{eq:quadraticformT} N_{\epsilon_0,\epsilon_1}=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m},\mathbb{F}_{p^m},\mathbb{F}_p)\, |\, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}+\epsilon_0p^{\frac{m-1}{2}}+\epsilon_1\right\}\right|,$$ where $\epsilon_0=\pm 1$, and $\epsilon_1\in\{0,1\}$. For any $a \in \mathbb{F}_{p^m}^*$, $a^{p^k} x^{p^{2k}} + ax $ is a permutation polynomial over $\mathbb{F}_{p^m}$ since $v_2(m)=0$. When $b$ runs through $\mathbb{F}_{p^m}$, the corresponding solution $x_b$ of $a^{p^k} x^{p^{2k}} + ax + b^{p^k}=0$ runs through $\mathbb{F}_{p^m}$. So, the number of the pairs $(c, x_b)$ such that $c={\rm Tr}_1^m\left(ax_b^{p^k+1}\right)$ is $p^m$ when $(b,c)$ runs over $(\mathbb{F}_{p^m}, \mathbb{F}_p)$. Hence, from (\[111\]), (\[112\]) and Proposition \[pro:ww\], we have $$\begin{split} N_{0,0}&=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m},\mathbb{F}_{p^m},\mathbb{F}_p)\,\,|\,\operatorname{wt_H}(\mathbf{c}(a,b,c))=(p-1)p^{m-1} \right\}\right|\\ &=\left|\left\{(a, b,c) \in (0, \mathbb{F}_{p^m},\mathbb{F}_p)\,\,|\,\operatorname{wt_H}(\mathbf{c}(a,b,c))=(p-1)p^{m-1} \right\}\right|\\ &+\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\,\,|\,\operatorname{wt_H}(\mathbf{c}(a,b,c))=(p-1)p^{m-1} \right\}\right|\\ &=p(p^m-1)+\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\,\,|\,{\rm Tr}_1^m(a)=0,\,\, c_0=0\right\}\right|\\ &=p(p^m-1)+p^m(p^{m-1}-1), \end{split}$$ and $$\begin{split} N_{0,1}&=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m},\mathbb{F}_{p^m},\mathbb{F}_p)\,\,|\,\operatorname{wt_H}(\mathbf{c}(a,b,c))=(p-1)p^{m-1}+1 \right\}\right|\\ &=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m},\mathbb{F}_{p^m},\mathbb{F}_p)\,\,|\,{\rm Tr}_1^m(a)\neq0,\,\, c_0=0\right\}\right|\\ &=p^m(p^m-p^{m-1}). \end{split}$$ Recall that $c_0=c-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)$. It is easy to see that $c_0$ runs through $\mathbb{F}_p$ if $c$ runs through $\mathbb{F}_p$ for any $(a,b) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m})$. So, when $(a,b,c)$ runs through $(\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m}, \mathbb{F}_p)$, by Proposition \[pro:ww\], we have $N_{1,0}=N_{-1,0}$ and $N_{1,1}=N_{-1,1}$ since $\eta(ac_0)=\eta(a)\eta_0(c_0)$, where $\eta$ and $\eta_0$ are quadratic multiplicative characters of $\mathbb{F}_{p^m}^*$ and $\mathbb{F}_{p}^*$, respectively. Hence, from the first three Pless power moment identities, we have $$\begin{split} \begin{cases} N_{1,0}=N_{-1,0}=\frac{p^m(p^{m-1}-1)(p-1)}{2}, \\ N_{1,1}=N_{-1,1}=\frac{p^{2m-1}(p-1)^2}{2}. \end{cases} \end{split}$$ So, we obtain the weight distribution of $\mathcal{C}_k$ in Table \[Table1\] for the case $v_2(m)=0$. [**Case 2**]{}: $0<v_2(m)\leq v_2(k)$. From (\[111\]), (\[112\]) and Proposition \[pro:ww\], when $a=0$ and $(b,c)$ runs over $(\mathbb{F}_{p^m}, \mathbb{F}_p)$, the number of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ being $p^m$ or $(p-1)p^{m-1}$ is $p-1$ or $p(p^m-1)$, respectively. We now determine the multiplicities of the other possible values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ for $(a,b,c)$ running through $(\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m}, \mathbb{F}_p)$. Define $$\label{eq:quadraticformT} N_{\epsilon_0,\epsilon_1,\epsilon_2}=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\, |\, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}+\epsilon_0p^{\frac{m}{2}}+\epsilon_1p^{\frac{m-2}{2}}+\epsilon_2\right\}\right|,$$ where $\epsilon_0\in\{0,1,-1\}$, $\epsilon_1\in \{1,-1\}$ and $\epsilon_2 \in \{0,1\}$. Then, by Lemma \[lemeven\] we have $$\begin{split} N_{1,-1,0}&=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\, |\, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}+p^{\frac{m}{2}}-p^{\frac{m-2}{2}}\right\}\right|,\\ &=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\,\,|\,{\rm Tr}_1^m(a)=0,\,\,\eta(a)=1,\,\,c_0=0\right\}\right|\\ &=p^m\left(p^{m-1}-1-(p-1)p^{\frac{m-2}{2}}\right)/2, \end{split}$$ and $$\begin{split} N_{1,-1,1}&=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\, |\, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}+p^{\frac{m}{2}}-p^{\frac{m-2}{2}}+1\right\}\right|\\ &=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\,\,|\,{\rm Tr}_1^m(a)\neq0,\,\,\eta(a)=1,\,\,c_0=0\right\}\right|\\ &=p^m(p-1)\left(p^{m-1}+p^{\frac{m-2}{2}}\right)/2. \end{split}$$ By the similar calculations, we can get the values of $N_{-1,1,0}$, $N_{-1,1,1}$, $N_{0,-1,0}$, $N_{0,-1,1}$, $N_{0,1,0}$ and $N_{0,1,1}$. Then the weight distribution of $\mathcal{C}_k$ is obtained in Table \[Table2\] for the case $0<v_2(m)\leq v_2(k)$. From Table \[Table1\], Table \[Table2\] and the first five Pless power moments identities, we have that the dual of $\mathcal{C}_k$ has parameters $[p^m+1,p^m-2m,4]$ for the case $v_2(m)\leq v_2(k)$. By the Sphere Packing bound, we have $$\begin{split} p^{p^m+1}\geq p^{p^m+1-(2m+1)}\left(\sum_{i=0}^{\lfloor\frac{{\rm d_H}(\mathcal{C}_k^{\perp})-1}{2}\rfloor}\left( \begin{array}{cccc} p^m+1 \\ i \\ \end{array} \right)(p-1)^i\right). \end{split}$$ Hence, ${\rm d_H}(\mathcal{C}_k^{\perp})\leq 6$ if $p=3$ and ${\rm d_H}(\mathcal{C}_k^{\perp})\leq 4$ if $p\geq 5$. Therefore, when $p\geq 5$, $\mathcal{C}_k^{\perp}$ is optimal with respect to the Sphere Packing bound. $\square$ Let $k=0$ in Theorem \[theorem1\], then we obtain [@Hengar Theorem 16]. This means that [@Hengar Theorem 16] can be seen as a special case of Theorem \[theorem1\]. \[example1\] Let $\mathcal{C}_k$ be the linear code in Theorem \[theorem1\]. [(1)]{} Let $m=3$, $p=5$ and $k=1$. Then $\mathcal{C}_k$ has parameters $[126,7,95]$ and its dual has parameters $[126,119,4]$. [(2)]{} Let $m=4$, $p=3$ and $k=0$. Then $\mathcal{C}_k$ has parameters $[82,9,48]$ and its dual has parameters $[82,73,4]$. All of these codes and the duals are optimal or almost optimal with respect to the tables of best codes known maintained at http://www.codetables.de. The weight distribution of $\mathcal{C}_k$ for $v_2(m)> v_2(k)$ =============================================================== In this section, we always assume that $v_2(m)> v_2(k)$, $d=\gcd(m,k)$, $\xi$ is a primitive element of $\mathbb{F}_{p^m}$ and $a_0=\xi^{\frac{p^m-1}{2(p^d-1)}}$, where $m$ and $k$ are positive integers. In the following, we determine the weight distribution of the code $\mathcal{C}_k$ for the case $v_2(m)> v_2(k)$, where $\mathcal{C}_k$ is defined in (\[code0\]). Recall that $$\begin{split} S(a,b)=\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{{\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)}. \end{split}$$ The possible values of exponential sum $S(a,b)$ for the case $v_2(m)> v_2(k)$ have been given in [@Coulter1998] and [@Coulter2002]. [@Coulter1998; @Coulter2002]\[lemeven1\] Let $a,b \in \mathbb{F}_{p^m}$ with $a\neq 0$. Then $S(a,b)=0$ unless the equation $a^{p^k}x^{p^{2k}}+ax+b^{p^k}=0$ is solvable. Let $\xi$ be a primitive element of $\mathbb{F}_{p^m}$ and $a_0=\xi^{\frac{p^m-1}{2(p^d-1)}}$. Assume $a^{p^k}x^{p^{2k}}+ax+b^{p^k}=0$ is solvable, then there are two possibilities. [(i)]{} If $a\neq a_0\xi^{s\left(p^d+1\right)}$ for any integer $s$, then the equation has a unique solution $x_b$ for any $b \in \mathbb{F}_{p^m}$, and $$S(a,b)=(-1)^{\frac{m}{2d}}p^{\frac{m}{2}}\zeta_p^{-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)}.$$ [(ii)]{} If $a= a_0\xi^{s\left(p^d+1\right)}$ for some integer $s$, then the equation is solvable if and only if ${\rm Tr}_{2d}^m\left(b\gamma^{-s}\right)=0$, where $\gamma \in \mathbb{F}_{p^m}^*$ is the unique element satisfying $\gamma^{\frac{p^k+1}{p^d+1}}=\xi$. In such cases, $$S(a,b)=-(-1)^{\frac{m}{2d}}p^{\frac{m}{2}+d}\zeta_p^{-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)},$$ where $x_b$ is a solution of $a^{p^k}x^{p^{2k}}+ax+b^{p^k}=0$ and $d=\gcd(k,m)$. Now, we determine the possible values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ for $(a,b,c)$ running through $(\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m}, \mathbb{F}_p)$. \[pro:ww1\] Let $c \in \mathbb{F}_p$ and the other notations be as in Lemma \[lemeven1\]. Let $c_0=c-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)$. If $a\neq a_0\xi^{\left(p^d+1\right)s}$ for any integer $s$, then the possible distinct values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ are $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c))=\begin{cases} (p-1)p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}, & \text{if ${\rm Tr}_1^m(a)=0$ and $c_0\neq 0$,} \\ (p-1)p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}+1, & \text{if ${\rm Tr}_1^m(a)\neq 0$ and $c_0\neq 0$,} \\ (p-1)p^{m-1}-(-1)^{\frac{m}{2d}}(p-1)p^{\frac{m-2}{2}}, & \text{if ${\rm Tr}_1^m(a)=0$ and $c_0=0$,} \\ (p-1)p^{m-1}-(-1)^{\frac{m}{2d}}(p-1)p^{\frac{m-2}{2}}+1, & \text{if ${\rm Tr}_1^m(a)\neq 0$ and $c_0=0$.} \\ \end{cases} \end{split}$$ If $a=a_0\xi^{\left(p^d+1\right)s}$ for some integer $s$, then the possible distinct values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ are $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c))=\begin{cases} (p-1)p^{m-1}, & \text{if ${\rm Tr}_1^m(a)=0$ and ${\rm Tr}_{2d}^m(b\gamma^{-s})\neq0$}, \\ (p-1)p^{m-1}+1, & \text{if ${\rm Tr}_1^m(a)\neq 0$ and ${\rm Tr}_{2d}^m(b\gamma^{-s})\neq0$}, \\ (p-1)p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}, & \text{if ${\rm Tr}_1^m(a)=0$, ${\rm Tr}_{2d}^m(b\gamma^{-s})=0$ and $c_0\neq 0$,} \\ (p-1)p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}+1, & \text{if ${\rm Tr}_1^m(a)\neq 0$, ${\rm Tr}_{2d}^m(b\gamma^{-s})=0$ and $c_0\neq 0$,} \\ (p-1)p^{m-1}+(-1)^{\frac{m}{2d}}(p-1)p^{\frac{m+2d-2}{2}}, & \text{if ${\rm Tr}_1^m(a)=0$, ${\rm Tr}_{2d}^m(b\gamma^{-s})=0$ and $c_0=0$,} \\ (p-1)p^{m-1}+(-1)^{\frac{m}{2d}}(p-1)p^{\frac{m+2d-2}{2}}+1, & \text{if ${\rm Tr}_1^m(a)\neq 0$, ${\rm Tr}_{2d}^m(b\gamma^{-s})=0$ and $c_0=0$.} \\ \end{cases} \end{split}$$ [*Proof*]{}. We only prove the case ${\rm Tr}_1^m(a)=0$. The case ${\rm Tr}_1^m(a)\neq 0$ can be proved similarly and omit the details here. It is easy to see that $p^d+1\,\big|\, \frac{p^m-1}{p-1}$ since $v_2(m)> v_2(k)$ and $d=\gcd(m,k)$. Then, $z \in \mathbb{F}_p^*$ can be expressed as $z=\xi^{(p^d+1)i}$, where $i$ is a positive integer. This means that $az= a_0\xi^{(p^d+1)s_1}$ if and only if $a= a_0\xi^{(p^d+1)s}$ for any $z \in \mathbb{F}_p^*$, where $s$, $s_1$ are positive integers. There are three cases. [**Case 1**]{}: $a\neq a_0\xi^{(p^d+1)s}$ for any integer $s$. By Lemma \[lemeven1\] we have $$\label{even1} \begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c))&=p^m-\frac{1}{p}\sum_{z \in \mathbb{F}_p}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{z\left({\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)+c\right)}\\ &=(p-1)p^{m-1}-\frac{1}{p}\sum_{z \in \mathbb{F}_p^*}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{z\left({\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)+c\right)}\\ &=(p-1)p^{m-1}-\frac{1}{p}\sum_{z \in \mathbb{F}_p^*}\zeta_p^{zc}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{{\rm Tr}_1^m\left(zax^{p^k+1}+zbx\right)}\\ &=(p-1)p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}\sum_{z \in \mathbb{F}_p^*}\zeta_p^{z\left(c-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)\right)}\\ &=\begin{cases} (p-1)p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}, & if\,\, c\neq{\rm Tr}_1^m\left(ax_b^{p^k+1}\right),\\ (p-1)p^{m-1}-(-1)^{\frac{m}{2d}}(p-1)p^{\frac{m-2}{2}}, & if\,\, c={\rm Tr}_1^m\left(ax_b^{p^k+1}\right). \end{cases} \end{split}$$ [**Case 2**]{}: $a=a_0\xi^{(p^d+1)s}$ for some integer $s$ and ${\rm Tr}_{2d}^m(b\gamma^{-s})\neq0$. From the third equality of (\[even1\]) we have $$\operatorname{wt_H}(\mathbf{c}(a,b,c))=(p-1)p^{m-1}-\frac{1}{p}\sum_{z \in \mathbb{F}_p^*}\zeta_p^{zc}S(za,zb),$$ where $$\begin{split} S(za,zb)=\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{{\rm Tr}_1^m\left(zax^{p^k+1}+zbx\right)}. \end{split}$$ In this case, we have $S(za,zb)=S(a,b)=0$. Hence, $\operatorname{wt_H}(\mathbf{c}(a,b,c))=(p-1)p^{m-1}.$ [**Case 3**]{}: $a=a_0\xi^{(p^d+1)s}$ for some integer $s$ and ${\rm Tr}_{2d}^m(b\gamma^{-s})=0$. By Lemma \[lemeven1\] and the third equality (\[even1\]) we obtain $$\begin{split} \operatorname{wt_H}(\mathbf{c}(a,b,c)) & =(p-1)p^{m-1}-\frac{1}{p}\sum_{z \in \mathbb{F}_p^*}\zeta_p^{zc}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{{\rm Tr}_1^m\left(zax^{p^k+1}+zbx\right)}\\ &=(p-1)p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}\sum_{z \in \mathbb{F}_p^*}\zeta_p^{z\left(c-{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)\right)}\\ &=\begin{cases} (p-1)p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}, & \text{if $c\neq{\rm Tr}_1^m\left(ax_b^{p^k+1}\right)$,} \\ (p-1)p^{m-1}+(-1)^{\frac{m}{2d}}(p-1)p^{\frac{m+2d-2}{2}}, & \text{if $c={\rm Tr}_1^m\left(ax_b^{p^k+1}\right)$.} \\ \end{cases} \end{split}$$ Combining with Case $1$, Case $2$ and Case $3$, the result follows. $\square$ In order to determine the weight distribution of $\mathcal{C}_k$ for the case $v_2(m)>v_2(k)$, we need the following lemma. Let $m$ and $d$ be two positive integers such that $d\, | \, m$ and $v_2(m)>v_2(d)$. Let $\xi$ be a primitive element of $\mathbb{F}_{p^m}$, $a_0=\xi^{\frac{p^m-1}{2(p^d-1)}}$ and $a=a_0\xi^{(p^d+1)s}$ for some integer $s$. Let $N(a)$ be the number of $a \in \mathbb{F}_{p^m}$ such that ${\rm Tr}_1^m(a)=0$, then $$\begin{split} &N(a)=\begin{cases} \frac{p^{m-1}+p^{\frac{m+2d}{2}}-p^{\frac{m+2d-2}{2}}-1}{p^d+1}, & \text{if $v_2(m)=v_2(d)+1$,}\\ \frac{p^{m-1}-p^{\frac{m+2d}{2}}+p^{\frac{m+2d-2}{2}}-1}{p^d+1}, & \text{if $v_2(m)>v_2(d)+1$.} \end{cases} \end{split}$$ [*Proof*]{}. Let $G$ be the cyclic group generated by $\xi^{p^d+1}$. Obviously, $p^d+1 \, | \, p^m-1$ since $d\, | \, m$ and $v_2(m)>v_2(d)$, then by Lemma \[lem:quadraticsum\] we have $$\begin{split} N(a)&= \frac{1}{p}\sum_{z \in \mathbb{F}_p}\sum_{s=0}^{\frac{p^m-1}{p^d+1}-1}\zeta_p^{z{\rm Tr}_1^m\left(a_0\xi^{\left(p^d+1\right)s}\right)} \\ &=\frac{1}{p}\sum_{z \in \mathbb{F}_p}\left(\frac{1}{p^d+1}\sum_{x \in \mathbb{F}_{p^m}^*}\zeta_p^{z{\rm Tr}_1^m\left(a_0x^{p^d+1}\right)}\right)\\ &=\frac{p^m-p}{p(p^d+1)}+\frac{1}{p(p^d+1)}\sum_{z \in \mathbb{F}_p^*}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{z{\rm Tr}_1^m\left(a_0x^{p^d+1}\right)}\\ &=\frac{p^m-p}{p(p^d+1)}+\frac{1}{p(p^d+1)}\sum_{z \in \mathbb{F}_p^*}\eta^r(z)\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{{\rm Tr}_1^m\left(a_0x^{p^d+1}\right)}\\ &=\frac{p^m-p}{p(p^d+1)}+\frac{p-1}{p(p^d+1)}\sum_{x \in \mathbb{F}_{p^m}}\zeta_p^{{\rm Tr}_1^m\left(a_0x^{p^d+1}\right)},\\ \end{split}$$ where $r$ is the rank of the quadratic form ${\rm Tr}_1^m\left(a_0x^{p^d+1}\right)$. In the last equality, we used the fact that the rank of the quadratic form ${\rm Tr}_1^m\left(a_0x^{p^d+1}\right)$ is even, which has been given in [@Coulter19980; @Draper2007]. Clearly, $$\begin{split} a_0^{\frac{(p^d-1)(p^m-1)}{p^{\gcd(2d,m)}-1}}=\xi^{{\frac{p^m-1}{2(p^d-1)}}\cdot{\frac{(p^d-1)(p^m-1)}{p^{\gcd(2d,m)}-1}}}=\begin{cases} -1, & \text{if $v_2(m)=v_2(d)+1$,}\\ 1, & \text{if $v_2(m)>v_2(d)+1$.} \end{cases} \end{split}$$ Hence, by Lemma \[cor:quadraticsum\] we have $$\begin{split} N(a)&=\begin{cases} \frac{p^{m-1}+p^{\frac{m+2d}{2}}-p^{\frac{m+2d-2}{2}}-1}{p^d+1}, & \text{if $v_2(m)=v_2(d)+1$,}\\ \frac{p^{m-1}-p^{\frac{m+2d}{2}}+p^{\frac{m+2d-2}{2}}-1}{p^d+1}, & \text{if $v_2(m)>v_2(d)+1$.} \end{cases} \end{split}$$ The result follows. $\square$ With above preparations, we now give the weight distribution of $\mathcal{C}_k$ for the case $v_2(m)>v_2(k)$. \[thm2\] Let $m, k$ be positive integers with $v_2(m)>v_2(k)$, $d=\gcd(k,m)$ and $\mathcal{C}_k$ be the linear code defined in (\[code0\]). The following statements hold. [(1)]{} If $v_2(m)=v_2(k)+1$ $(m\neq 2k)$, then $\mathcal{C}_k$ is a $\left[p^m+1, 2m+1,(p-1)\left(p^{m-1}-p^{\frac{m+2d-2}{2}}\right)\right]$ code with the weight distribution in Table \[Table3\]. Its dual has parameters $[p^m+1,p^m-2m,4]$, which is optimal with respect to the Sphere Packing bound for $p>3$. Weight Multiplicity ----------------------------------------------------------------------- ----------------------------------------------------------------------------------------------------------------- -- -- $0$ $1$ $p^m$ $p-1$ $(p-1)p^{m-1}$ $\left(p^{m-d}-p^{m-2d}\right)\left(p^{m}+p^{\frac{m+2d+2}{2}}-p^{\frac{m+2d}{2}}-p\right)+p\left(p^m-1\right)$ $(p-1)p^{m-1}+1$ $(p-1)\left(p^{m-d}-p^{m-2d}\right)\left(p^m-p^{\frac{m+2d}{2}}\right)$ $(p-1)p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}$ $(p-1)(p^{2m+d-1}-p^{\frac{3m+2d}{2}}+p^{\frac{3m+2d-2}{2}}-p^{m+d})/(p^d+1)$ $(p-1)p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}+1$ $(p-1)^2\left(p^{2m+d-1}+p^{\frac{3m+2d-2}{2}}\right)/(p^d+1)$ $(p-1)\left(p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}\right)$ $\left(p^{2m+d-1}-p^{\frac{3m+2d}{2}}+p^{\frac{3m+2d-2}{2}}-p^{m+d}\right)/(p^d+1)$ $(p-1)\left(p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}\right)+1$ $(p-1)\left(p^{2m+d-1}+p^{\frac{3m+2d-2}{2}}\right)/(p^d+1)$ $(p-1)p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}$ $\left(p^{m-1}+p^{\frac{m+2d}{2}}-p^{\frac{m+2d-2}{2}}-1\right)(p-1)p^{m-2d}/(p^d+1)$ $(p-1)p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}+1$ $(p-1)^2\left(p^{2m-2d-1}-p^{\frac{3m-2d-2}{2}}\right)/(p^d+1)$ $(p-1)\left(p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}\right)$ $\left(p^{m-1}+p^{\frac{m}{2}+d}-p^{\frac{m}{2}+d-1}-1\right)p^{m-2d}/(p^d+1)$ $(p-1)\left(p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}\right)+1$ $(p-1)\left(p^{2m-2d-1}-p^{\frac{3m-2d-2}{2}}\right)/(p^d+1)$ : [ The weight distribution of $\mathcal{C}_k$ for $v_2(m)>v_2(k)$ with $m\neq 2k$ ]{} []{data-label="Table3"} [(2)]{} If $v_2(m)>v_2(k)+1$, then $\mathcal{C}_k$ is a $\left[p^m+1, 2m+1,(p-1)p^{m-1}-p^{\frac{m+2d-2}{2}}\right]$ code with the weight distribution in Table \[Table3\]. Its dual has parameters $[p^m+1,p^m-2m,4]$, which is optimal with respect to the Sphere Packing bound for $p>3$. [(3)]{} If $m=2k$, then $\mathcal{C}_k$ is a $\left[p^m+1, \frac{3m}{2}+1,(p-1)p^{m-1}-p^{\frac{m-2}{2}}\right]$ code with the weight distribution in Table \[Table4\]. Its dual has parameters $[p^m+1,p^m-\frac{3m}{2},4]$, which is optimal with respect to the Sphere Packing bound for $p>3$. Weight Multiplicity ------------------------------------------------- ----------------------------------- -- -- $0$ $1$ $p^m$ $p-1$ $(p-1)p^{m-1}$ $p^{m+1}-p$ $(p-1)p^{m-1}-p^{\frac{m-2}{2}}$ $(p-1)(p^{\frac{3m}{2}-1}-p^{m})$ $(p-1)p^{m-1}-p^{\frac{m-2}{2}}+1$ $(p-1)^2p^{\frac{3m}{2}-1}$ $(p-1)\left(p^{m-1}+p^{\frac{m-2}{2}}\right)$ $p^{\frac{3m}{2}-1}-p^{m}$ $(p-1)\left(p^{m-1}+p^{\frac{m-2}{2}}\right)+1$ $(p-1)p^{\frac{3m}{2}-1}$ : The weight distribution of $\mathcal{C}_k$ for $m=2k$ []{data-label="Table4"} [*Proof.*]{} We only prove the case $v_2(m)=v_2(k)+1$. The case $v_2(m)>v_2(k)+1$ can be shown by the similar way. The multiplicities of all possible values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ for $(a,b,c)$ running through $(0,\mathbb{F}_{p^m}, \mathbb{F}_p)$ have been given in (\[111\]), (\[112\]). In the following, we only consider the multiplicities of all possible values of $\operatorname{wt_H}(\mathbf{c}(a,b,c))$ for $(a,b,c)$ running through $(\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m}, \mathbb{F}_p)$. There are two cases to be discussed. [**Case 1**]{}: $a\neq a_0\xi^{(p^d+1)s}$ for any integer $s$. Define $$\begin{split} H_{\epsilon_0,\epsilon_1}=&\big|\big\{ (a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\,\,|\,\, a\neq a_0\xi^{(p^d+1)s} \,\, \text{for any $s$}, \\ &\,\,\,\,\,\,\,\,\,\, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}+\epsilon_0p^{\frac{m}{2}}-p^{\frac{m-2}{2}}+\epsilon_1\big\}\big|, \end{split}$$ where $\epsilon_0,\epsilon_1\in\{0,1\}$. By Proposition \[pro:ww1\] we have $$\begin{split} H_{0,0}&=\left|\left\{(a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\, |\, a\neq a_0\xi^{(p^d+1)s} \,\, \text{for any $s$}, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}-p^{\frac{m-2}{2}}\right\}\right|\\ &=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\, |\, {\rm Tr}_1^m(a)=0,\,\, c_0\neq 0 \right\}\right|\\ &-\left|\left\{(a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\, |\, a=a_0\xi^{(p^d+1)s} \,\,\text{for some $s$}, {\rm Tr}_1^m(a)=0,\,\, c_0\neq 0 \right\}\right|\\ &=(p^{m-1}-1)(p-1)p^m-\frac{\left(p^{m-1}+p^{\frac{m+2d}{2}}-p^{\frac{m+2d-2}{2}}-1\right)(p-1)p^m}{p^d+1}\\ &=\frac{(p-1)\left(p^{2m+d-1}-p^{\frac{3m+2d}{2}}+p^{\frac{3m+2d-2}{2}}-p^{m+d}\right)}{p^d+1}, \end{split}$$ and $$\begin{split} H_{0,1}&=\left|\left\{(a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\,\, |\, a\neq a_0\xi^{(p^d+1)s} \,\, \text{for any $s$}, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}-p^{\frac{m-2}{2}}+1\right\}\right|\\ &=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\, |\, {\rm Tr}_1^m(a)\neq0,\,\, c_0\neq 0 \right\}\right|\\ &-\left|\left\{ (a, b,c) \in ({ {\mathbb F}}_{p^m}^*\mathbb{F}_{p^m},\mathbb{F}_p)\, |\,a=a_0\xi^{(p^d+1)s} \,\,\text{for some $s$},\,\, {\rm Tr}_1^m(a)\neq 0,\,\, c_0\neq 0 \right\}\right|\\ &=\left|\left\{(a,b,c) \in (\mathbb{F}_{p^m}^*,\mathbb{F}_{p^m},\mathbb{F}_p)\, |\, {\rm Tr}_1^m(a)\neq0,\,\, c_0\neq 0 \right\}\right|\\ &-\left|\left\{ (a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\, |\, a=a_0\xi^{(p^d+1)s} \,\,\text{for some $s$},\,\, c_0\neq 0 \right\}\right|\\ &+\left|\left\{ (a, b,c) \in (\mathbb{F}_{p^m},\mathbb{F}_p)\, |\,a=a_0\xi^{(p^d+1)s} \,\,\text{for some $s$},\,\, {\rm Tr}_1^m(a)=0,\,\, c_0\neq 0 \right\}\right|\\ &=(p-1)^2p^{2m-1}-\frac{(p^m-1)(p-1)p^m}{p^d+1}+\frac{\left(p^{m-1}+p^{\frac{m+2d}{2}}-p^{\frac{m+2d-2}{2}}-1\right)(p-1)p^m}{p^d+1}\\ &= \frac{(p-1)^2\left(p^{2m+d-1}+p^{\frac{3m+2d-2}{2}}\right)}{p^d+1}. \end{split}$$ By the similar calculations, we can get the values of $H_{1,0}$ and $H_{1,1}$. [**Case 2**]{}: $a=a_0\xi^{(p^d+1)s}$ for some integer $s$. Define $$\begin{split} T_{\epsilon_0,\epsilon_1,\epsilon_2}=&\big|\big\{ (a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\, |\, a=a_0\xi^{(p^d+1)s} \text{ for some $s$}, \\ &\,\,\,\,\,\,\,\,\,\, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}-\epsilon_0p^{\frac{m+2d}{2}}+\epsilon_1p^{\frac{m+2d-2}{2}}+\epsilon_2\big\}\big|, \end{split}$$ where $\epsilon_0,\epsilon_1,\epsilon_2\in\{0,1\}$. By Proposition \[pro:ww1\] we have $$\begin{split} T_{0,0,0}&=\left|\left\{(a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\, |\, a=a_0\xi^{(p^d+1)s} \text{ for some $s$}, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}\right\}\right|\\ &=p\left|\left\{ (a, b) \in ( { {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m}) \, |\, a=a_0\xi^{(p^d+1)s} \,\,\text{for some $s$},\, {\rm Tr}_1^m(a)=0,\, {\rm Tr}_{2d}^m(b\gamma^{-s})\neq0 \right\}\right|\\ &=\frac{\left(p^m-p^{m-2d}\right)\left(p^{m}+p^{\frac{m+2d+2}{2}}-p^{\frac{m+2d}{2}}-p\right)}{p^d+1}\\ &=\left(p^{m-d}-p^{m-2d}\right)\left(p^{m}+p^{\frac{m+2d+2}{2}}-p^{\frac{m+2d}{2}}-p\right),\\ \end{split}$$ $$\begin{split} T_{0,0,1}&=\left|\left\{ (a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\, |\,a=a_0\xi^{(p^d+1)s} \text{ for some $s$}, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}+1\right\}\right|\\ &=p\left|\left\{(a, b)\in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m})\, |\, a=a_0\xi^{(p^d+1)s} \,\,\text{for some $s$},\, {\rm Tr}_1^m(a)\neq 0,\, {\rm Tr}_{2d}^m(b\gamma^{-s})\neq0 \right\}\right|\\ &=p\left|\left\{ (a, b) \in ( { {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m})\, |\, a=a_0\xi^{(p^d+1)s} \,\,\text{for some $s$},\, {\rm Tr}_{2d}^m(b\gamma^{-s})\neq0 \right\}\right|-T_{0,0,0}\\ &=\frac{\left(p^{m}-p^{m-2d}\right)\left(p^{m+1}-p\right)}{p^d+1}-\frac{\left(p^m-p^{m-2d}\right)\left(p^{m}+p^{\frac{m+2d+2}{2}}-p^{\frac{m+2d}{2}}-p\right)}{p^d+1}\\ &=(p-1)\left(p^{m-d}-p^{m-2d}\right)\left(p^m-p^{\frac{m}{2}+d}\right), \end{split}$$ and $$\begin{split} T_{0,1,1}&=\left|\left\{ (a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\, |\,a=a_0\xi^{(p^d+1)s} \,\text{for some $s$}, \, \operatorname{wt_H}(\mathbf{c}(a,b,c)) = (p-1)p^{m-1}+p^{\frac{m+2d-2}{2}}+1\right\}\right|\\ &=\left|\left\{ (a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\, |\, a=a_0\xi^{(p^d+1)s} \,\text{for some $s$},\, {\rm Tr}_1^m(a)\neq0,\,\,{\rm Tr}_{2d}^m(b\gamma^{-s})=0,\,\, c_0\neq 0 \right\}\right|\\ &=\left|\left\{(a, b,c) \in ({ {\mathbb F}}_{p^m}^*, \mathbb{F}_{p^m},\mathbb{F}_p)\, |\,a=a_0\xi^{(p^d+1)s} \text{ for some $s$},\, {\rm Tr}_{2d}^m(b\gamma^{-s})=0,\,\, c_0\neq 0 \right\}\right|\\ &-\left|\left\{ (a,b,c) \in (\mathbb{F}_{p^m},\mathbb{F}_p)\, |\,a=a_0\xi^{(p^d+1)s} \,\,\text{for some $s$},\, {\rm Tr}_1^m(a)=0,\,\,{\rm Tr}_{2d}^m(b\gamma^{-s})=0,\,\, c_0\neq 0 \right\}\right|\\ &=\frac{(p^{m}-1)(p-1)p^{m-2d}}{p^d+1}-\frac{\left(p^{m-1}+p^{\frac{m+2d}{2}}-p^{\frac{m+2d-2}{2}}-1\right)(p-1)p^{m-2d}}{p^d+1}\\ &=\frac{(p-1)^2\left(p^{2m-2d-1}-p^{\frac{3m-2d-2}{2}}\right)}{p^d+1}. \end{split}$$ By the similar calculations, we get the values of $T_{1,1,1}$, $T_{1,1,0}$ and $T_{0,1,0}$. Combining with (\[111\]), (\[112\]), Case $1$ and Case $2$, we get Table \[Table3\]. When $v_2(m)\geq v_2(k)$, from the first five Pless power moment identities and the weight distribution of $\mathcal{C}_k$, we obtain that the dual of $\mathcal{C}_k$ is a $[p^m+1,p^m-2m,4]$ code, which is optimal with respect to the Sphere Packing bound for the case $p\geq 5$. When $m=2k$, i.e., $m=2d$, from Table \[Table3\], we know that the Hamming weight $0$ occurs $p^{\frac{m}{2}}$ times if $(a,b,c)$ runs through $(\mathbb{F}_{p^m},\mathbb{F}_{p^m}, \mathbb{F}_p)$. This means that every codeword in $\mathcal{C}_k$ repeats $p^{\frac{m}{2}}$ times. Hence, in this case, $\mathcal{C}_k$ is degenerate and its dimension is $\frac{3m}{2}$. Substituting $d=\frac{m}{2}$ to Table \[Table3\] and dividing each frequency by $p^\frac{m}{2}$, we get Table \[Table4\]. From the first five Pless power moments identities and the weight distribution of $\mathcal{C}_k$, we have the dual of $\mathcal{C}_k$ is a $[p^m+1,p^m-\frac{3m}{2},4]$ code, which is optimal with respect to the Sphere Packing bound for the case $p\geq 5$. $\square$ \[example2\] Let $\mathcal{C}_k$ be the linear code in Theorem \[thm2\]. [(1)]{} Let $m=2$, $p=5$ and $k=1$. Then $\mathcal{C}_k$ has parameters $[26,4,19]$ and its dual has parameters $[26,22,4]$. [(2)]{} Let $m=4$, $p=3$ and $k=1$. Then $\mathcal{C}_k$ has parameters $[82,9,45]$ and its dual has parameters $[82,73,4]$. The duals of these codes are optimal or almost optimal with respect to the tables of best codes known maintained at http://www.codetables.de. The punctured code of $\mathcal{C}_k$ ===================================== In this section, we study the linear code $$\label{eqbarC} \begin{split} \mathcal{\bar{C}}_k=\left\{\left( {\rm Tr}_1^m\left(ax^{p^k+1}+bx\right)+c\right)_{x \in \mathbb{F}_{p^m}}: \, a,b \in \mathbb{F}_{p^m}, c \in \mathbb{F}_p\right\}. \end{split}$$ Note that $\mathcal{\bar{C}}_k$ can be viewed as a punctured code of $\mathcal{C}_k$ in Section $3$ and Section $4$. Hence, the weight distribution of $\mathcal{\bar{C}}_k$ can be easily derived from that of $\mathcal{C}_k$. We here give the weight distribution of $\mathcal{\bar{C}}_k$ and the parameters of the dual of $\mathcal{\bar{C}}_k$. \[thm3\] Let $d=\gcd(k,m)$ and $\mathcal{\bar{C}}_k$ be the linear code defined in (\[eqbarC\]). The following statements hold. [(1)]{} If $v_2(m)=0$, then $\mathcal{\bar{C}}_k$ is a $\left[p^m, 2m+1, (p-1)p^{m-1}- p^{\frac{m-1}{2}}\right]$ code with the weight distribution in Table \[Table5\]. Its dual has parameters $[p^m,p^m-2m-1,4]$ if $p>3$, and its dual has parameters $[p^m,p^m-2m-1,5]$ if $p=3$ and $m>1$. Weight Multiplicity -------------------------------------- --------------------- -- -- $0$ $1$ $p^m$ $p-1$ $(p-1)p^{m-1}$ $(p+p^m)(p^m-1)$ $(p-1)p^{m-1}\pm p^{\frac{m-1}{2}}$ $p^m(p^m-1)(p-1)/2$ : The weight distribution of $\mathcal{\bar{C}}_k$ for $v_2(m)=0$ []{data-label="Table5"} [(2)]{} If $0<v_2(m)\leq v_2(k)$, then $\mathcal{\bar{C}}_k$ is a $\left[p^m, 2m+1,(p-1)\left(p^{m-1}-p^{\frac{m-2}{2}}\right)\right]$ code with the weight distribution in Table \[Table6\]. Its dual has parameters $[p^m,p^m-2m-1,4]$ if $p>3$, and its dual has parameters $[p^m,p^m-2m-1,5]$ if $p=3$. Weight Multiplicity -------------------------------------------------- ----------------------- -- -- $0$ $1$ $p^m$ $p-1$ $(p-1)p^{m-1}$ $p(p^m-1)$ $(p-1)\left(p^{m-1}\pm p^{\frac{m-2}{2}}\right)$ $(p^{2m}-p^m)/2$ $(p-1)p^{m-1}\pm p^{\frac{m-2}{2}}$ $(p-1)(p^{2m}-p^m)/2$ : The weight distribution of $\mathcal{\bar{C}}_k$ for $0<v_2(m)\leq v_2(k)$ []{data-label="Table6"} [(3)]{} If $v_2(m)=v_2(k)+1$ $(m\neq 2k)$, then $\mathcal{\bar{C}}_k$ is a $\left[p^m, 2m+1,(p-1)\left(p^{m-1}-p^{\frac{m+2d-2}{2}}\right)\right]$ code with the weight distribution in Table \[Table7\]. Its dual has parameters $[p^m,p^m-2m-1,4]$. [(4)]{} If $v_2(m)> v_2(k)+1$, then $\mathcal{\bar{C}}_k$ is a $\left[p^m, 2m+1,(p-1)p^{m-1}-p^{\frac{m+2d-2}{2}}\right]$ code with the weight distribution in Table \[Table7\]. Its dual has parameters $[p^m,p^m-2m-1,4]$. Weight Multiplicity --------------------------------------------------------------------- --------------------------------------------------------- -- -- $0$ $1$ $p^m$ $p-1$ $(p-1)p^{m-1}$ $p^{2m-d+1}-p^{2m-2d+1}+p^{m+1}-p^{m-d+1}+p^{m-2d+1}-p$ $(p-1)p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}$ $(p-1)(p^{2m+d}-p^{m+d})/(p^d+1)$ $(p-1)\left(p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m-2}{2}}\right)$ $(p^{2m+d}-p^{m+d})/(p^d+1)$ $(p-1)p^{m-1}-(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}$ $(p-1)(p^{2m-2d}-p^{m-2d})/(p^d+1)$ $(p-1)\left(p^{m-1}+(-1)^{\frac{m}{2d}}p^{\frac{m+2d-2}{2}}\right)$ $(p^{2m-2d}-p^{m-2d})/(p^d+1)$ : The weight distribution of $\mathcal{\bar{C}}_k$ for $v_2(m)>v_2(k)$ []{data-label="Table7"} [(5)]{} If $m=2k$, then $\mathcal{\bar{C}}_k$ is a $\left[p^m, \frac{3m}{2}+1,(p-1)p^{m-1}-p^{\frac{m-2}{2}}\right]$ code with the weight distribution in Table \[Table8\]. Its dual has parameters $[p^m,p^m-\frac{3m}{2}-1,4]$. Weight Multiplicity ----------------------------------------------- ------------------------------- -- -- $0$ $1$ $p^m$ $p-1$ $(p-1)p^{m-1}$ $p^{m+1}-p$ $(p-1)p^{m-1}-p^{\frac{m-2}{2}}$ $(p-1)(p^{\frac{3m}{2}}-p^m)$ $(p-1)\left(p^{m-1}+p^{\frac{m-2}{2}}\right)$ $p^{\frac{3m}{2}}-p^m$ : The weight distribution of $\mathcal{\bar{C}}_k$ for $m=2k$ []{data-label="Table8"} [(6)]{} The dual of $\mathcal{\bar{C}}_k$ is an optimal code with respect to the Sphere Packing bound if $p>3$, and the dual of $\mathcal{\bar{C}}_k$ is an optimal ternary code for the case $v_2(m)\leq v_2(k)$ if $p=3$ and $m>1$. [*Proof.*]{} The weight distribution of $\mathcal{\bar{C}}_k$ can be derived directly from Theorem \[theorem1\] and Theorem \[thm2\]. In the following, we will show the minimum Hamming distance of the dual of $\mathcal{\bar{C}}_k$. There are two cases. [[**Case 1**]{}: $v_2(m)> v_2(k)$]{}. By the first five Pless power moment identities, we get that ${\rm d_H}(\mathcal{\bar{C}}^{\perp}_k)=4$, which is optimal with respect to the Sphere Packing bound for the case $p>3$. [[**Case 2**]{}: $v_2(m)\leq v_2(k)$]{}. By the first five Pless power moment identities, we obtain that ${\rm d_H}(\mathcal{\bar{C}}^{\perp}_k)=4$ if $p> 3$ and ${\rm d_H}(\mathcal{\bar{C}}^{\perp}_k)>4$ if $p=3$. Clearly, the linear code $\mathcal{\bar{C}}^{\perp}_k$ is optimal with respect to the Sphere Packing bound if $p> 3$. We now show ${\rm d_H}(\mathcal{\bar{C}}^{\perp}_k)=5$ if $p=3$ and $m>1$ (If $m=1$ and $p=3$, then $\mathcal{\bar{C}}_k$ is a $[3,3,1]$ code and $\mathcal{\bar{C}}^{\perp}_k=\mathbf{0}$). On on hand, if $p=3$, we have ${\rm d_H}(\mathcal{\bar{C}}^{\perp}_k)\leq 6$ by the Sphere Packing bound. On the other hand, assume that there exists a ternary linear code with parameters $[3^m, 3^m-2m-1,6]$. Applying Lemma \[bound2\], we have $q=3$, $n=3^m$, $t=3^m-5$, $r=2$, and $$3^{3^m-2m-1}\leq \frac{3^{3^m-1}}{1+2(3^m-1)^2},$$ which is impossible if $m>1$. Hence, ${\rm d_H}(\mathcal{\bar{C}}^{\perp}_k)=5$ if $p=3$ and $m>1$, which is optimal with respect to Lemma \[bound2\]. $\square$ It is clear that $\mathcal{\bar{C}}_k$ has fewer distinct weights than $\mathcal{C}_k$ and the dual of $\mathcal{\bar{C}}_k$ is optimal for many cases. This means that it is worthwhile to study the linear code $\mathcal{\bar{C}}_k$. \[example3\] Let $\mathcal{\bar{C}}_k$ be the linear code in Theorem \[thm3\]. [(1)]{} Let $m=3$, $p=5$ and $k=1$. Then $\mathcal{C}_k$ has parameters $[125,7,95]$ and its dual has parameters $[125,118,4]$. [(2)]{} Let $m=3$, $p=3$ and $k=1$. Then $\mathcal{C}_k$ has parameters $[27,7,15]$ and its dual has parameters $[27,20,5]$. [(3)]{} Let $m=4$, $p=3$ and $k=2$. Then $\mathcal{C}_k$ has parameters $[81,7,51]$ and its dual has parameters $[81,74,4]$. All of these codes and the duals are optimal with respect to the tables of best codes known maintained at http://www.codetables.de. Conclusion =========== This paper continued the work of Heng and Ding in [@Hengar], and investigated more subfield codes of linear codes. Some linear codes presented in this paper are optimal or almost optimal. To our knowledge, many presented codes have new parameters. Specifically, the main work is summarized as follows: [$\bullet$]{} In Section $3$ and Section $4$, we obtained the weight distribution of the subfield code $\mathcal{C}_k$, which is defined in (\[code0\]). [$\bullet$]{} In section $5$, the weight distribution of $\mathcal{\bar{C}}_k$ is determined, where $\mathcal{\bar{C}}_k$ defined in (\[eqbarC\]) is a punctured code of $\mathcal{C}_k$. [$\bullet$]{} The parameters of the duals of $\mathcal{C}_k$ and $\mathcal{\bar{C}}_k$ are determined. The duals of $\mathcal{C}_k$ and $\mathcal{\bar{C}}_k$ are optimal codes with respect to the Sphere Packing bound if $p>3$, and the dual of $\mathcal{\bar{C}}_k$ is optimal for the case $v_2(m)\leq v_2(k)$ if $p=3$ and $m>1$. [$\bullet$]{} In Theorem \[theorem1\], the dual of $\mathcal{C}_k$ is a $p$-ary MDS code with parameters $[p+1,p-2,4]$ if $m=1$. [$\bullet$]{} In Theorem \[thm3\], the dual of $\mathcal{\bar{C}}_k$ is a $p$-ary MDS code with parameters $[p,p-3,4]$ if $m=1$ and $p>3$. [$\bullet$]{} Table \[Table5\] demonstrated a class of $p$-ary MDS codes with parameters $[p, 3, p-2]$ if $m=1$. [$\bullet$]{} Example \[example1\], Example \[example2\] and Example \[example3\] showed some optimal or almost optimal codes with respect to the tables of best codes known maintained at http://www.codetables.de. [100]{} J. Cannon, W. Bosma, C.Fieker, E. Stell, Handbook of Magma Functions, Version 2.19, Sydney, 2013. A. Canteaut, P. Charpin, H. Dobbertin, Weight divisibility of cyclic codes, highly nonlinear functions on $\mathbb{F}_{2^n}$, and crosscorrelation of maximum-length sequences, SIAM Disc. Math., 13 (1) (2000) 105-138. C. Carlet, P. Charpin, V. Zinoviev, Codes, bent functions and permutations suitable For DES-like cryptosystems, Des. Codes Cryptogr., 15 (2) (1998) 125-156. R. S. Coulter, Explicit evaluations of some Weil sums. Acta Arith. 83 (1998) 241-251. R. S. Coulter, Further evaluations of Weil sums, Acta Arith., 86 (1998) 217-226. R. S. Coulter, The number of rational points of a class of Artin-Schreier curves, Finite Fields Appl., 8 (2002) 397-413. C. Ding, Z. Heng, The subfield codes of ovoid codes, DOI 10.1109/TIT.2019.2907276. C. Ding, T. Helleseth, Optimal ternary cyclic codes from monomials, IEEE Trans. Inf. Theory, 59(9) (2013) 5898-5904. C. Ding, Linear codes from some 2-designs, IEEE Trans. Inf. Theory, 61 (6) (2015) 3265-3275. C. Ding, A construction of binary linear codes from Boolean functions, Discrete Math., 339 (9) (2016) 2288- 2303. K. Ding, C. Ding, Binary linear codes with three weights. IEEE Commun. Lett., 18 (2014) 1879-1882. K. Ding, C. Ding, A class of two-weight and three-weight codes and their applications in secret sharing. IEEE Trans. Inf. Theory, 61 (11) (2015) 5835-5842. H. Q. Dinh, C. Li, Q. Yue, Recent progress on weight distributions of cyclic codes over finite fields, J. Algebra Comb. Disc. Struc. & Appl., 2 (2015) 39-63. S. Draper, X. Hou, Explicit evaluation of certain exponential sums of quadratic functions over $\mathbb{F}_{p^n}$, $p$ odd, arxiv:0708.3619v1. (2007). S. Y. EI Rouayheb, C. N. Georghiades, E. Soljanin, A. Sprintson, Bounds on codes based on graph theory, IEEE Int. Symp. on Information Theory. Nice, France, June, (2007), 1876-1879. J. Fan, Optimal $p$-ary cyclic codes with minimum distance four, arXiv: 1706.09188v2. C. Fan, N. Li, Z. Zhou, A class of optimal ternary cyclic codes and their duals, Finite Fields Appl., 37 (2016) 193-202. D. Han, H. Yan, On an open problem about a class of optimal ternary cyclic codes, Finite Fields Appl., 59 (2019) 335-343. Z. Heng, C. Ding, The subfield codes of Hyperoval and Conic codes, DOI: 10.1016/j.ffa.2018.12.006. H. Liu, X. Wang, D. Zheng, On the weight distributions of a class of cyclic codes, Discrete Math., 341 (2018) 759-771. T. Kl[ø]{}ve, Codes for Error Detection, Hackensack, NJ: world Scientific, 2007. N. Li, C. Li, T. Helleseth, C. Ding, X. Tang, Optimal ternary cyclic codes with minimum distance four and five, Finite Fields Appl., 30 (2014) 100-120. N. Li, Z. Zhou, T. Helleseth, On a conjecture about a class of optimal ternary cyclic codes, Seventh International Workshop on Signal Design and its Applications in Communications (IWSDA), DOI: 10.1109/IWSDA.2015.7458415 (2015). R. Lidl, H. Niederreiter, Finite Fields, Encyclopedia of Mathematics, Vol. 20, Cambridge University Press, Cambridge, 1983. F.J. MacWilliams, N.J.A. Sloane, The Theory of Error-Correcting Codes, North-Holland Publishing Company, 1997. B. Segre, Ovals in a finite projective plane Canad. J. Math., 7 (1955) 414-416. G. Xu, X. Cao, S. Xu, Optimal $p$-ary cyclic codes with minimum distance four from monomials, Cryptogr. Commun., 8(4) (2016) 541-554. Y. Zhou, X. Kai, S. Zhu, J. Li, On the minimum distance of negacyclic codes with two zeros, Finite Fields Appl., 55 (2019) 143-150. [^1]: \*Corresponding author. E-Mail addresses: [email protected] (X. Wang), [email protected](D. Zheng), [email protected](Y. Zhang)
{ "pile_set_name": "ArXiv" }
--- abstract: 'In [@P] R. Pellikaan introduced a two variable zeta-function $Z(t,u)$ for a curve over a finite field ${{\mathbb{F}}}_q$ which, for $u=q$, specializes to the usual zeta-function and he proved, among other things, rationality: $Z(t,u)=(1-t)^{-1}(1-ut)^{-1}P(t,u)$ with $P(t,u)\in{{\mathbb{Z}}}[t,u]$. We prove that $P(t,u)$ is absolutely irreducible. This is motivated by a question of J. Lagarias and E. Rains about an analogous two variable zeta-function for number fields.' author: - Niko Naumann title: 'On the irreducibility of the two variable zeta-function for curves over finite fields' --- [MSC2000: 11G20 (primary), 14G10 (secondary)]{} Introduction {#sec:1} ============ Let $X$ be a proper, smooth, geometrically connected curve of genus $g$ over the finite field ${{\mathbb{F}}}_q$. The zeta-function of $X/{{\mathbb{F}}}_q$ can be written as a power series $$Z(t)=\sum_D\frac{q^{h^0(D)}-1}{q-1}t^{deg(D)}.$$ Here the sum is over ${{\mathbb{F}}}_q$-rational divisor classes of $X$ and\ $h^0$($D$):= dim$_{{{\mathbb{F}}}_q}$ $H^0(X,{\cal O}(D))$. Writing $b_{nk}$ for the number of divisor classes of degree $n$ and with $h^0(D)$=$k$ this becomes $$Z(t)=\sum_{n\ge 0}\sum_{k\ge 1} b_{nk}\frac{q^k-1}{q-1}t^n.$$ In [@P] R. Pellikaan observed that the classical proof of rationality and the functional equation for $Z(t)$ go through when $q$ is treated as a variable in this expression. He thus introduced the following power series in [@P], Def. 3.1: $$Z(t,u):=\sum_{n\ge 0}\sum_{k\ge 1} b_{nk}\frac{u^k-1}{u-1}t^n.$$ This is called the two variable zeta-function of the curve. We will denote by $h$ the class-number of $X/{{\mathbb{F}}}_q$, i.e. $h=$$|$Pic$^0$($X$)$|$. Then Pellikaan proved: \[rationality\] We have $$Z(t,u)=(1-t)^{-1}(1-ut)^{-1}P(t,u) \mbox{ with } P\in{{\mathbb{Z}}}[t,u].$$ Furthermore:\ 1) deg$_t$ $P$=$2g$, deg$_u$ $P$=$g$.\ 2) In the expansion $P(t,u)=\sum_{i=0}^{2g} P_i(u)t^i$ one has $P_0(u)=1$,\ deg$_u$ $P_i(u)\leq i/2+1$ and $P_{2g-i}(u)=u^{g-i}P_i(u)$ for $0\leq i\leq 2g$.\ 3) $P(1,u)=h$. Here deg$_u$ and deg$_t$ denote the degree of a polynomial in the indicated variable. The above results are all taken from [@P], Prop. 3.5 and we copied only those needed later on. Note that the statement deg$_u$ $P_i(u)\leq i/2$ in \[loc. cit.\] is a misprint. Indeed, we will see below that one always has deg$_u$ $P_1(u)=1$ (unless $g=0$). As expected we have $P(t,u)=1$ in case $g=0$.\ In [@GS] G. van der Geer and R. Schoof used analogies from Arakelov-theory to define a two variable zeta-function for number fields along the above lines. As the number field case will serve only as a motivation in this note we refer to the original sources [@GS] and [@LR] for definitions and to [@D] for a comparison between them. Suffice it to say that in [@LR], section 8 we find an entire function $\xi_{{{\mathbb{Q}}}}(w,s)$ of two complex variables which for $w=1$ equals Riemann’s $\xi$-function. In particular, the zeroes of $\xi_{{{\mathbb{Q}}}}(1,s)$ are precisely the non-trivial zeros of the Riemann zeta-function. One is thus led to study the zero-locus of $\xi_{{{\mathbb{Q}}}}(w,s)$. J. Lagarias and E. Rains ask whether it might be the closure of a single irreducible complex-analytic variety of multiplicity one. The corresponding question in the geometric case seems to be whether the zero-locus of $P(t,u)$ is irreducible.\ This is indeed the case: \[irr\] In the above situation, $P(t,u)$ is irreducible in ${{\mathbb{C}}}(u)[t]$. As an illustration we discuss the cases $g=1$ and $g=2$:\ For $g=1$ setting $N:=|X({{\mathbb{F}}}_q)|$ we have $$P(t,u)=1+(N-1-u)t+ut^2,$$ c.f. [@P], example 3.4. This polynomial is reducible in ${{\mathbb{C}}}(u)[t]$ if and only if $N=0$ in which case we have $P(t,u)=(1-t)(1-ut)$. But it is well-known that a curve of genus one over a finite field always has a rational point, i.e. $N\neq 0$.\ In case $g=2$ let the usual zeta-function of $X$ be $$Z(t)=(1-t)^{-1}(1-qt)^{-1}L(t)$$ with $$L(t)=1+at+bt^2+qat^3+q^2t^4$$ for certain $a,b\in{{\mathbb{Z}}}$. As $X$ is hyperelliptic, Prop. 4.3. of [@P] can be used to compute $$P(t,u)=1+((a+q)-u)t+((q(q-1)+aq+b)-(a+q-1)u)t^2+((a+q)-u)ut^3+u^2t^4.$$ This will not be used in the sequel and we omit the proof.\ In order not to lead intuition astray we point out that in general $Z(t,u)$ is not determined by $Z(t)$, see [@P], example 4.4.\ After a lengthy computation with discriminants one sees that a neccessary condition for this $P(t,u)$ to be reducible is $$b+a(q+1)+(q^2+1)=0.$$ But this expression equals $L(1)=h\neq 0$ !\ The fact that $h\neq 0$ enters in the general proof of theorem \[irr\] precisely through the condition $\beta\neq 0$ of lemma \[irrcrit\] below. Note, however, that condition 2) of this lemma cannot be dropped. So one needs one more result on $P(t,u)$, contained in Prop. \[leadingcoeff\], which follows from Clifford’s theorem.\ I would like to thank C. Deninger for posing the above problem and for useful discussions on the topic. Proof ===== We will use the following criterion for irreducibility: \[irrcrit\] Let $k$ be a field, $F\in k[u,t]$ and assume:\ 1) $F$ is monic in $t$.\ 2) the leading coefficient of $F$ as a polynomial in $u$ is irreducible in $k[t]$.\ 3) there are $\alpha,\beta\in k$, $\beta\neq 0$ with $F(u,\alpha)=\beta$. Then $F$ is irreducible in $k(u)[t]$. This lemma will be applied to $$F=\tilde{P}(t,u):= t^{2g}P(t^{-1},u)\in{{\mathbb{C}}}[u,t].$$ Note that the irreducibility of $\tilde{P}$ in ${{\mathbb{C}}}(u)[t]$ will imply the irreducibility of $P$ because $P(0,u)=P_0(u)=1\neq 0$ by theorem \[rationality\]. The advantage of $\tilde{P}$ is that it is monic in $t$ and so satisfies condition 1) of lemma \[irrcrit\]. Also 3) is satisfied (with $\alpha=1$, $\beta=h$) according to theorem \[rationality\], 3). (of lemma \[irrcrit\]) Assume to the contrary that $F=fg$ in $k(u)[t]$ with $f$ and $g$ of positive degree and monic. One knows, c.f. for example [@E], Prop 4.11, that the coefficient of $f$ and $g$ are integral over $k[u]$ and as $k[u]$ is integrally closed we have $f,g\in k[u,t]$. So we can consider the decomposition $F=fg$ as polynomials in $u$ and infer from 2) that the leading coefficient of $f$ as a polynomial in $u$ lies in $k[t]^*=k^*$ (upon exchanging $f$ and $g$ if neccessary). In particular $n:=$deg$_u$ $f(u,t)=$deg$_u$ $f(u,\alpha)$. Substituting 3) gives $\beta=f(u,\alpha)g(u,\alpha)$ in $k[u]$. As $\beta \neq 0$ we get $n=0$, i.e. $f$ is constant in $u$ hence $f\in k^*$, contradiction. We are left with verifying condition 2) of lemma \[irrcrit\] for the given $\tilde{P}$, i.e. the leading coefficient of $\tilde{P}$ as a polynomial in $u$ is irreducible in $k[t]$. We will in fact determine this coefficient: \[leadingcoeff\] For $g\geq 1$: $\tilde{P}(t,u)=(1-t)u^g+O(u^{g-1})$. We already know deg$_u$ $\tilde{P}=g$. Also the assertion is clear for $g=1$ from the formula for $P(t,u)$ recalled in the introduction. We assume $g\ge 2$. Looking at $$\tilde{P}(t,u)= t^{2g}+P_1(u)t^{2g-1}+\cdots +P_g(u)t^g+uP_{g-1}(u)t^{g-1}+\cdots +u^gt^0$$ and using the bound deg$_u$ $P_i(u)\leq i/2+1$ we see that $u^g$ can only occur in the last three terms: $u^{g-2}P_2(u)t^2+u^{g-1}P_1(u)t+u^g$. So the proof is completed by the following result on $P_1$ and $P_2$. For $g\geq 1$:\ 1) deg$_u$ $P_1(u)=1$ and the leading coefficient is $-1$.\ 2) deg$_u$ $P_2(u)\leq 1$. This is again clear for $g=1$. We assume $g\ge 2$ and write $P_i(u)=\sum_k\alpha_{ik}u^k$, $\alpha_{ik}\in{{\mathbb{Z}}}$. As we already know deg$_u$ $P_1(u)\leq 1$ and deg$_u$ $P_2(u)\leq 2$ we need to show $\alpha_{11}=-1$ and $\alpha_{22}=0$. Recalling the notation $b_{nk}$ from the introduction we have the following [**Claim**]{}: $b_{12}=\alpha_{00}+\alpha_{11}$ and $b_{23}=\alpha_{00}+\alpha_{11}+\alpha_{22}$. Granting this we observe that Clifford’s thorem, c.f. [@H], IV, thm 5.4, gives $b_{12}=b_{23}=0$. Recalling also $\alpha_{00}=1$, because $P_0(u)=1$, and substituting gives indeed $\alpha_{11}=-1$ and $\alpha_{22}=0$. To prove the above claim we write the rational expression for $Z(t,u)$ in theorem \[rationality\] in terms of coefficients: $$Z(t,u)=\sum_{n\geq 0}\sum_{k\geq 1} b_{nk}\frac{u^k-1}{u-1}t^n=(\sum_{i\geq 0}t^i)(\sum_{j\geq 0} (ut)^j)(\sum_{l\geq 0}\sum_{k\geq 0} \alpha_{lk} u^kt^l).$$ This gives \ 1) for $\nu$,$\alpha\geq 0$ : $\sum_{k\geq\alpha +1} b_{\nu k}=\sum_{\mu,i\geq 0,\mu+i\leq\nu}\alpha_{i,\alpha-\mu}$\ 2) for $\nu\geq 0$, $\mu\geq 1$: $b_{\nu\mu}=\sum_{i=0}^{\nu}(\alpha_{i,\mu-\nu-1+i}-\alpha_{i\mu})$ We omit the details of this straightforward computation except to say that for 1) one uses $\frac{u^k-1}{u-1}=1+\cdots+u^{k-1}$ and 2) follows from 1) by a telescope-summation. In the formulation of the lemma it is understood that $\alpha_{nk}=0$ whenever $k<0$. We get from 2):\ $$b_{23}=\alpha_{00}-\alpha_{03}+\alpha_{11}-\alpha_{13}+\alpha_{22}-\alpha_{23}$$ and $$b_{12}=\alpha_{00}-\alpha_{02}+\alpha_{11}-\alpha_{12}.$$ But we know $\alpha_{02}=\alpha_{03}=\alpha_{12}=\alpha_{13}=\alpha_{23}=0$ because deg$_u$ P$_i$(u)$\leq i/2+1$.\ This concludes the proof of the claim, hence of theorem \[irr\]. [99999]{} C. Deninger, A regularized product expansion for the two-variable zeta-function of van der Geer and Schoof, preprint, 2002. D. Eisenbud, Commutative Algebra with a View Towards Algebraic Geometry, GTM 150, Springer, 1995. G. van der Geer, R. Schoof, Effectivity of Arakelov Divisors and the Theta Divisor of a Number Field, Selecta Math. (N.S.), [**6**]{}, 2000, no. 4, pp 377-398. R. Hartshorne, Algebraic Geometry, GTM 52, Springer, 1977. J. Lagarias, E. Rains, On a Two-Variable Zeta Function for Number Fields, arXiv:math.NT/0104176 v5 7 Jul 2002. R. Pellikaan, On special divisors and the two variable zeta function of algebraic curves over finite fields, in: Arithmetic, Geometry and Coding Theory 4, Luminy, 1993, Walter de Gruyter, Berlin, 1996, pp 175-184. Mathematisches Institut der WWU Münster\ Einsteinstr. 62\ 48149 Münster\ Germany\ e-mail: [email protected]\ \
{ "pile_set_name": "ArXiv" }
--- abstract: 'We study the superfluid properties of a system of fully polarized dipolar bosons moving in the $XY$ plane. We focus on the general case where the polarization field forms an arbitrary angle $\alpha$ with respect to the $Z$ axis, while the system is still stable. We use the diffusion Monte Carlo and the path integral ground state methods to evaluate the one-body density matrix and the superfluid fractions in the region of the phase diagram where the system forms stripes. Despite its oscillatory behavior, the presence of a finite large-distance asymptotic value in the $s$-wave component of the one-body density matrix indicates the existence of a Bose condensate. The superfluid fraction along the stripes direction is always close to 1, while in the $Y$ direction decreases to a small value that is nevertheless different from zero. These two facts confirms that the stripe phase of the dipolar Bose system is a clear candidate for an intrinsic supersolid without the presence of defects as described by the Andreev-Lifshitz mechanism.' author: - 'R. Bombin, J. Boronat and F. Mazzanti' title: Dipolar Bose Supersolid Stripes --- Supersolid many-body systems appear in nature when two continuous U(1) symmetries are broken. The first one is associated to the translational invariance of the crystalline structure, while the second one corresponds to the appearance of a non-trivial global phase of the superfluid state [@Boninsegni_2012]. Supersolid phases were predicted to exist in Helium already in the late 60’s [@Andreev_69], though their experimental observation has been ellusive. In fact, the claims for detection made at the beginning of this century have been refuted, as the observed behavior is not caused by finite non-conventional rotational inertia but rather to elastic effects [@Kim_2012]. In this way, a neat observation of supersolidity in $^4$He is still lacking. In fact, it is not clear yet whether a pure, defect-free supersolid structure like the one that would be expected in $^4$He really exists. Recently, the issue of supersolidity has emerged again, but now in the field of ultracold atoms. Two different experimental teams have claimed that spatial local order and superfluidity have been simultaneously observed in lattice setups [@Leonard_17] and in stripe phases [@Li_17]. In this way, the definition of what a supersolid really is seems to still be under discussion [@Anderson_2017]. Superfluid properties of solid-like phases are also of fundamental interest in quantum condensed matter. One of these is the stripe phase, where the system presents spatial order in one direction but not in the others. For instance, stripe phases have been of major interest since 1990, when non-homogeneous metallic structures with broken spatial symmetry were found to favor superconductivity [@Bianconi_00; @Bianconi_13]. More recently, stripe phases have been observed in Bose-Einstein condensates with synthetically created spin-orbit coupling [@Li_17], where the momentum dependence of the interaction induces spatial ordering along a single direction in some regions of the phase diagram [@Li_13]. Stripe phases have also been discussed in the context of quantum dipolar physics, including very recent theoretical and experimental analysis of metastable striped gases of $^{164}$Dy [@Wenzel_2017]. Due to the anisotropic character of the dipolar interaction, in some regions of the phase diagram dipoles arrange in stripes, both in Fermi [@Yamaguchi_10; @Sun_10] and Bose [@Macia_2012; @Macia_2014] systems. In some cases the presence of this phase has been reported to exist even in the isotropic limit [@Parish_12]. Though the presence of stripe phases in dipolar systems is well established and has been recently observed [@Kadau_2016], it is not yet clear whether the system exhibits superfluid properties (thus forming supersolid stripes) or not. In a previous work we determined the phase diagram [@Macia_2014] of the two-dimensional system of Bose dipoles at zero temperature, tracing the transition lines between the solid, gas and stripe phases. The formation and excitation spectrum of the stripe phase, where the system acquires crystal order in one direction while being fluid on the other, was previously analyzed in Ref. [[@Macia_2012]]{}. In this Letter we investigate the superfluid properties of the stripe phase as a function of the density and polarization angle. Our results show that dipolar stripes are a special form of supersolid, and we quantify the superfluid density and condensate fraction all along the superstripe phase. In the following we consider a system of $N$ fully polarized dipolar bosons of mass $m$ moving on the XY plane. All dipoles are considered to be aligned along a fixed direction in space given by a polarization (electric or magnetic) field, which is contained on the XZ plane and forming and angle $\alpha$ with respect to the Z axis. The model Hamiltonian describing the system becomes then $$H = -{\hbar^2 \over 2m}\sum_{j=1}^N \nabla_j^2 + {C_{dd} \over 4\pi} \sum_{i<j}^N \left[ {1 - 3\lambda^2 \cos^2\theta_{ij} \over r_{ij}^3} \right] \ , \label{Hamiltonian}$$ with $\lambda=\sin\alpha$, and $(r_{ij},\theta_{ij})$ the polar coordinates associated to the position vector of particle $j$ with respect to particle $i$. The constant $C_{dd}$ is proportional to the square of the (electric or magnetic) dipole moment of the components, assumed all of them to be identical. In the following we use dimensionless units obtained from the characteristic dipolar length $r_0=m C_{dd}/(4\pi\hbar^2)$. We quantify the superfluid properties of the system evaluating both the one-body density matrix and its asymptotic value (the condensate fraction), and the superfluid density. In order to do that we employ stochastic methods. We use two different quantum Monte Carlo techniques that are known to provide exact values for the energy of the system within residual statistical noise: the diffusion Monte Carlo (DMC) [@Hammond_94; @Kosztin_96] and the path integral Monte Carlo (PIGS) [@Sarsa_2000; @Rota_2010] methods. The DMC simulations have been performed using a second order propagator [@Chin_90], while a fourth order propagator has been employed in the PIGS calculations [@Chin_02]. In all cases, a variational model of the ground state wave function $\Psi_T$ is used. In the DMC method, the guiding wave function is used for importance sampling but the ground state estimation of any observable commuting with the Hamiltonian is exact. In PIGS simulations, $\Psi_T$ acts as a boundary condition at the end points of the open chains representing the set of particles. It is then propagated in imaginary time to the center of the chains, where expectation values are evaluated. In this way, any contribution orthogonal to the exact ground state is wiped out. Two different models have been used in this work. In the DMC simulations, $\Psi_T$ has been taken to be of the Jastrow form, with a two-body correlation factor that results from the zero-energy solution of the two-body problem associated to Eq. (\[Hamiltonian\]) as derived in Ref. [@Macia_11], matched with a long-range phononic extension as discussed in the same reference. This model must be modified when describing the stripe phase, including a one-body term $f_1({\bf r})$ that allows for the formation of the stripes along the $Y$ direction $$f_1(\textbf{r})=\exp\left[ \eta_s\cos\left({2\pi n_s y\over L_y}\right)\right] \ , \label{onebodystripes}$$ with $L_y$ the box side length along the $Y$ direction, and $n_s$ the number of stripes in the simulation box. Notice that these two parameters are not independent, as one must guarantee the simulation box is commensurated for a fixed number of particles. In Eq. (\[onebodystripes\]), $\eta_s$ is a variational parameter that is consistently found to be zero in the gas phase, and non-zero in the stripe phase. For the PIGS simulations, we have adopted a much simpler model based on the zero-energy solution of the isotropic ($\alpha=0$) problem, matched with a phononic tail as in Ref. [@Macia_2014]. Despite its simplicity, we have found no differences with the results obtained when using the same model as in the DMC case. ![(color online). Phase diagram of the 2D dipolar Bose gas at zero temperature. Letters indicate the set of points corresponding to fixed density and polarization angles explored in this work. []{data-label="fig_PhaseDiag"}](Diag_Phase_FINAL_2.png){width="48.00000%"} Since we are analyzing superfluid properties, we have performed several calculations spanning a wide range of densities and polarization angles in the regions of the phase diagram where the system is in stripe form. Notice that, in the solid phase, the system arranges in a triangular lattice that completely breaks the continuous translational symmetry [@Macia_2014], while in the stripe phase this symmetry is broken only in one direction (the Y axis in our setup). For the sake of comparison, we have also explored two additional points where the system remains either as a gas or as a solid. The set of points explored in this work is shown in the phase diagram, Fig. \[fig\_PhaseDiag\], and a summary of the results obtained for these points is reported in Table \[table\_1\]. $n r_0^2$ $\alpha$ $n_0$ $\rho_s$ $\rho_s^x$ $\rho_s^y$ --- ----------- ---------- ------------ ---------- ------------ ------------ A 512 0,50 0.00030(4) 0.86(8) 1.06(8) 0.61(8) B 512 0.53 0.00055(6) 0.62(6) 0.99(8) 0.26(3) C 512 0.55 0.0029(3) 0.53(5) 0.92(8) 0.14(2) D 512 0.57 0.0031(3) 0.49(5) 0.95(8) 0.043(4) E 512 0.60 0.0047(5) 0.49(5) 0.95(8) 0.027(3) F 400 0.50 0.0038(3) 1.05(8) 1.07(8) 1.04(8) G 400 0.55 0.0042(4) 0.63(6) 1.001(7) 0.26(3) H 400 0.60 0.0052(4) 0.55(5) 1.07(8) 0.028(3) I 256 0.55 0.015(1) 1.05(8) 1.03(8) 1.08(8) J 256 0.60 0.011(1) 0.54(5) 1.00(8) 0.080(6) K 128 0.60 0.071(4) 0.95(7) 0.97(7) 0.93(7) L 512 0.20 0 0 0 0 M 256 0.40 0.019(2) 1 1 1 : Superfluid densities and condensate fraction for the points shown in Fig. \[fig\_PhaseDiag\]. Figures in parenthesis are the errorbars.[]{data-label="table_1"} A direct measure of the off-diagonal long-range order present in the system is provided by the one-body density matrix (OBDM) $$\begin{aligned} n_1({\bf r}_{11'}) \!\! & = & \!\! \Omega \int d{\bf r}_2\cdots d{\bf r}_N \\ & & \Psi_0( {\bf r}_1, {\bf r}_2,\ldots,{\bf r}_N) \Psi_0( {\bf r}'_1, {\bf r}_2,\ldots,{\bf r}_N) \ , \nonumber \label{obdm-b1}\end{aligned}$$ with $\Psi_0$ the ground state wave function and $\Omega$ the volume of the container. In this way, $n_1({\bf r})$ is normalized such that $n_1(0)=1$, while $n_1(|{\bf r}_{11'}|\to\infty)\to n_0$ if there is off-diagonal long-range order, with $n_0$ the condensate fraction. Notice that, in 2D, $n_0$ can be non-zero only at $T=0$. ![(color online). One Body Density Matrix of the 2D dipolar Bose system at the density $n r_0^2=512$ in the stripe phase for $\alpha=0.55$ (filled circles), and in the solid phase for $\alpha=0.20$ (empty squares). Purple circles and green squares: cuts along the X direction; blue circles and orange squares: cuts along the Y direction. Distance $r$ is measured in units of r$_0$. Errorbars are smaller than $10\%$ of each measure and have not been included for the sake of clarity.[]{data-label="fig_OBDM"}](nr_av_Np252n51200a055.pdf){width="49.00000%"} Figure \[fig\_OBDM\] shows a comparison of the one-body density matrix of the system at points C and L of Fig. \[fig\_PhaseDiag\], corresponding to the same density $nr_0^2=512$ but different polarization angles. In all cases $n_1({\bf r})$ depends on the direction due to the anisotropy of the interaction. The lower curves show two cuts of $n_1({\bf r})$ along the $X$ and $Y$ directions, when the system is in the solid phase (point L), while the upper curves show the same quantities for the system in the stripe phase (point C). As it can be seen, all curves show an oscillatory behavior that is partially a consequence of the anisotropy of the interaction [@Macia_11]. Most remarkably, the curves corresponding to the solid phase decay exponentially to zero, while the ones for the stripe phase saturate to a common value that corresponds to $n_0$. The condensate fraction, which appears only in the $s$-wave term of the partial wave expansion of $n_1({\bf r})$, has been obtained by fitting a constant to the intermediate-distance tail in regions near (but not at) half the box side where the results are stable. All values in the third column of Table \[table\_1\] have been obtained in this way. At large densities, where increasing the polarization angle makes the system change from the solid to the stripe phase, the condensate fraction increases with increasing $\alpha$. This is not surprising since the dipolar interaction is overall less repulsive when approaching the line of collapse, at the critical angle $\alpha_c\approx 0.615$. The situation is reversed at lower densities, when the system changes from the gas to the stripe form (point I and J for instance). In this case and close to the transition line, the condensate fraction is expected to approach higher values, as the gas is less interacting. Perpendicular cuts at fixed polarization angle and increasing density leads always to a reduction in $n_0$, consistent with the fact that particles have less effective space. In any case the largest values of $n_0$ are achieved near the gas-stripe transition line at the lowest possible densities. In this way, the large-distance limit of the OBDM of the stripe phase is always non-zero, as happens with other supersolid systems. Even though the presence of non-zero condensate fraction value already points towards a superfluid behavior, it is possible to evaluate directly the superfluid response of the system in DMC. At finite temperature, the superfluid fraction $\rho_s$ is estimated from the winding number [@Pollock_87], which takes into accounts the diffusion of world lines at large imaginary times. At $T=0$, this is equivalent [@Zhang_95] to measuing the diffusion of the center of mass of the system in the infinite imaginary time limit, according to the expression $$\rho_s = \lim_{\tau\to\infty} {1\over 4 N \tau} \left( {D_s(\tau) \over D_0} \right) \ , \label{superflu_b1}$$ where $D_s(\tau)=\left\langle ({\bf R}_{CM}(\tau) - {\bf R}_{CM}(0) )^2 \right\rangle$ and $D_0=\hbar^2/(2m)$. For the 2D system analyzed, we identify the $X$ and $Y$ components of this expression with the superfluid fractions along the $X$ and $Y$ directions, according to $\rho_s=(\rho_s^x+\rho_s^y)/2$. ![(color online). Superfluid fractions along the $X$ direction $\rho_s^x$ (red crosses), along the $Y$ direction $\rho_s^y$ (green squares) and total $\rho_s$ (blue stars). The upper panel shows the dependence of these quantities on the polarization angle at the fixed density $n r_0^2=512$. The lower panel corresponds to $\alpha=0.6$ and different densities. In all cases the system remains in the stripe phase. []{data-label="fig_rhos"}](rho_s.pdf){width="49.00000%"} Figure \[fig\_rhos\] shows our results for $\rho_s^x$, $\rho_s^y$ and the total $\rho_s$ for two perpendicular cuts on the phase diagram. The upper panel corresponds to a fixed density $nr_0^2=512$ and different angles in the region where the system remains in the stripe phase. The lower panel corresponds to a fixed angle $\alpha=0.6$ but different densities, also in the stripe phase. The cut at $nr_0^2=512$ and increasing $\alpha$ shows that the $X$ component of the superfluid fraction is always close to $1$, while the $Y$ component decreases to 0, leading to the overall value $\rho_s\approx 1/2$ near $\alpha=\alpha_c$. Remarkably, the total superfluid fraction $\rho_s$ is larger close to the transition line to the solid phase, decreasing as $\alpha$ increases. In this way, the superfluid response is discontinuous across the solid-stripe transition. The fact that $\rho_s^y$ (and thus $\rho_s$) decrease when $\alpha$ increases is once again a consequence of the anisotropic character of the dipolar interaction, which becomes less repulsive along the $X$ direction with increasing $\alpha$. Close to $\alpha_c$ the interaction along the $X$ direction is weak and particles can easily flow in each stripe, but the confinement of the stripes is stronger and the system becomes more localized along the $Y$ direction. This is confirmed by the fact that the optimal values of $\eta_s$ in Eq. (\[onebodystripes\]) is larger when $\alpha$ approaches $\alpha_c$ at fixed density. A similar situation is found when the density is increased at constant $\alpha$. The lower panel of Fig. \[fig\_rhos\] show the different components of the superfluid fraction at $\alpha=0.6$ and increasing density. Once again we observe that $\rho_s^y$ decays to values close to zero already at $nr_0^2=256$, thus confirming that at high densities the confinement of the different stripes is very strong. Only point K in that line presents a large $\rho_s^y$ value, but that point is essentially in the gas-stripe transition line, and we know the total superfluid fraction $\rho_s=1$ in the gas phase. Contrarily to what happens when moving from the stripe to the solid phase, in the gas-stripe transition the change in $\rho_s$, $\rho_s^x$ and $\rho_s^y$ appears to be continuous. ![(color online). One-body density matrix along the $X$ (blue open squares) and $Y$ (green stars) at $\alpha=0.6$ and $nr_0^2=512$. The solid lines are fits of the form $A x^{-1/\eta}$ with for fixed $\eta$ obtained from the slope of the static structure factor near the origin. The inset shows a snapshot of the PIGS simulation, where some of the particle exchanges are highlighted in black. []{data-label="fig_n1_Lutt"}](nr_av_Np252n512a06_NEW.png){width="49.00000%"} At this point, and according to the previous results, one could wonder whether stripes are so tightly confined that no particle exchanges between different stripes is possible. If that was the case, one could also think that each stripe may behave as an isolated, (quasi) 1D system. In fact and according to the results in the last column of Table \[table\_1\], in some regions the $Y$ component of the superfluid fraction acquires very low values. However, it never vanishes. This indicates that, in fact, particle exchange between different stripes is always possible, though it becomes unlikely in the limits commented above. Taking that into account, one can look for traces of a (quasi) 1D behavior in the regions where $\rho_a^y\sim 0$. One way to do that is to analyze the system as a Luttinger liquid, and to check for consistency in the values of the corresponding Luttinger parameters. In order to do that we have extracted the sound velocity $c$ from a fit of the form $|k|/2c$ to the low-$k$ behavior of the static structure factor $S({\bf k})$ evaluated both in DMC and PIGS. Once with it, we have performed a fit of the form $n_1(u)=A u^{-1/\eta}$ with $\eta=2\pi n/c$  [@Luttinger_81] to the $X$ and $Y$ components of $n_1({\bf r})$, with the results shown in Fig, \[fig\_n1\_Lutt\]. As it can be seen, the fit reproduces better the tail of $n_1({\bf r})$ along the $X$ direction, while strong oscillations in the $Y$ component are clearly visible and $n_1({\bf r})$ for ${\bf r}=(0,y)$ differs significantly from the fit. It must be kept in mind, though, that the large distance behavior of $n_1({\bf r})$ in Luttinger liquid theory is a decaying power law not compatible with a finite condensate fraction value, while we have seen before that the stripe phase OBDM presents a large-distance asymptotic value $n_0\neq 0$. In this way, the curve fits well the calculated $X$-component of the OBDM at intermediate distances only. The inset in Fig. \[fig\_n1\_Lutt\] shows a snapshot of the system after thermalization in PIGS, for the same conditions $nr_0^2=512$ and $\alpha=0.6$, where a pair of examples where particle exchange between different stripes is visible, have been highlighted. It is worth recalling that since simulations in PIGS are done with open chains (with variational wave functions at the end points), it is hardly possible to see long exchange lines crossing the whole simulation box. In summary, we have performed DMC and PIGS simulations to analyze the supersolid properties of dipolar Bose stripes in two dimensions for polarization angles before collapse. We have evaluated the one-body density matrix to find that it always presents a finite (though in some regions, quite small) condensate fraction value, in contrast to the continuously decaying tail it presents in the solid phase. We have also evaluated the superfluid fraction along the $X$ and $Y$ directions to find that, at large densities and/or polarization angles, the $Y$-component becomes very small, though it never vanishes. At high densities and polarization angles the stripes are tightly confined and the intermediate distance behavior of the OBDM along the stripe direction has a dependence on the distance that is somehow compatible with a Luttinger liquid model. However, particle exchanges, always visible in configuration snapshots, lead to a finite condensate fraction value and an overall superfluid behavior that, together with the existence of Bragg peaks [@Macia_2014], confirm the supersolid character of that phase. This work has been supported by the Ministerio de Economia, Industria y Competitividad (MINECO, Spain) under grant No. FIS2014-56257-C2-1-P. [99]{} M. Boninsegni, and N. V. Prokof’ev, Rev. Mod. Phys. [**84**]{}, 759 (2012). A. F. Andreev, and I. M. Lifshitz, JETP [**29**]{}, 1107 (1969). E. Kim, and M. H. W. Chan, Nature [**427**]{}, 225 (2004). D. Y. Kim, and M. H. W. Chan, Phys. Rev. Lett. [**109**]{}, 155301 (2012). J. Léonard, A. Morales, P. Zupancic, T. Esslinger, and T. Donner, Nature [**543**]{}, 87 (2017). J. R. Li, J. Lee, W. Huang, S. Burchesky, B. Shteynas, F.Ç. Top, A. O. Jamison, and W. Ketterle, Nature [**543**]{}, 91 (2017). P. W. Anderson, Physics World [**30**]{}, 21 (2017). A. Bianconi, Int. J. Mod. Phys. B, [**14**]{}, 3289 (2000). A. Bianconi, D. Innocenti, and G. Campi, Journal of Superconductivity and Novel Magnetism, [**26**]{}, 2585 (2013). Y. Li, G. I. Martone, L. P. Pitaevskii, and S. Stringari, Phys. Rev. Lett. [**110**]{}, 235302 (2013). M. Wenzel, F. Böttcher, T. Langen, I. Ferrier-Barbut, and T. Pfau, arXiv:1706.09388. Y. Yamaguchi, T. Sogo, T. Ito, and T. Miyakawa, Phys. Rev. A [**82**]{}, 013643 (2010). K. Sun, C. Wu, and S. Das Sarma, Phys. Rev. B [**82**]{}, 075105 (2010). A. Macia, D. Hufnagl, F. Mazzanti, J. Boronat, and R. E. Zillich, Phys. Rev. Lett. [**109**]{}, 235307 (2012) A. Macia, J. Boronat, and F. Mazzanti, Phys. Rev. A [**90**]{}, 061601(R) (2014). M. M. Parish, and F. M. Marchetti, Phys. Rev. Lett. [**108**]{}, 145304 (2012). H. Kadau, M. Schmitt, M. Wenzel, C. Wink, T. Maier, I. Ferrier-Barbut, T. Pfau, Nature 530, 194-197 (2016). B. L. Hammond, W. A. Lester Jr., and P. J. Reynolds, [*Monte Carlo Methods in Ab InitioQuantum Chemistry*]{}, (World Scientific, Singapore, 1994). I. Kosztin, B. Faber, and K. Schulten, Am. J. Phys. [**64**]{}(5), 633 (1996). S. A. Chin, Phys. Rev. A [**42**]{}, 6991 (1990). A. Sarsa, K. E. Schmidt, and W. R. Magro, J. Chem. Phys. [**113**]{}, 1366 (2000). R. Rota, J. Casulleras, F. Mazzanti, and J. Boronat, Phys. Rev. E [**81**]{}, 016707 (2010). S. A. Chin, and C. R. Chen, J. Chem. Phys. [**117**]{}, 1409 (2002). A. Macia, F. Mazzanti, J. Boronat, and R. E. Zillich, Phys. Rev. A [**84**]{}, 033625 (2011). E. L. Pollock and D. M. Ceperley, Phys. Rev. B [**36**]{}, 8343 (1987). S. Zhang, N. Kawashima, J. Carlson, and J. E. Gubernatis, Phys. Rev. Lett. [**74**]{}, 1500 (1995). F. D. M. Haldane, Phys. Rev. Lett. [**47**]{}, 1840 (1981).
{ "pile_set_name": "ArXiv" }
--- author: - 'Yuna G. Kwon ' -  Masateru Ishiguro - Jungmi Kwon - Daisuke Kuroda - Myungshin Im - Changsu Choi - Motohide Tamura - Takahiro Nagayama - Nobuyuki Kawai - 'Jun-Ichi Watanabe' date: 'Received March 26, 2019 / Accepted July 8, 2019' title: 'Near-Infrared Polarimetric Study of Near-Earth Object 252P/LINEAR: An Implication of Scattered Light from the Evolved Dust Particles' --- Introduction \[sec:intro\] ========================== Comets, which are among the least-altered leftovers from the early solar system, have probably preserved their primitive structures inside, whereas their surfaces become different from initial states after repetitive orbital revolutions around the Sun. Solar radiation depletes near-surface volatiles and grains with high surface-to-volume ratios [@Prialnik2004], while concurrent sintering effects would lead to the growth of grain-to-grain contact areas and condensation of them [@Kossacki1994; @Thomas1994; @Gundlach2018]. As a result, resurfacing by the inert refractory layer (the so-called ‘dust mantle’) makes the surface drier and more consolidated than the bulk nuclei [@Biele2015; @Spohn2015; @Kossacki2015]. The cometary dust that we observe generally originates from such a near-surface layer or bounded boulders ejected from the last apparition [@Rotundi2015; @Fulle2018]. Conversely, fresh particles tend to be ejected from the interiors of the nuclei in rather erratic ways, such as during a sudden brightness enhancement near perihelion or outbursts (e.g., @Ishiguro2010 [@Ishiguro2016]). The polarimetry of comets is a useful diagnostic for investigating the physical properties of dust particles, such as their sizes and porosities [@Kiselev2015]. In particular, polarimetric observations at near-infrared wavelengths (NIR; 1.2 $\mu$m–2.3 $\mu$m) have been predicted to set constraints on the porosity of dust grains by covering more dust monomers (basic constitutional units of dust aggregates with $\sim$0.1 $\mu$m radii; e.g., @Kimura2006) within a single wavelength than in the optical [@Kolokolova2010]. Electromagnetic interaction between the monomers in aggregates depolarizes the light, randomizing directional information of the scattered light, so that the lower the porosity of a dust particle is, the greater the depolarization as the wavelength increases. It is thus theoretically expected that NIR polarimetry would maximize the difference between the porosities of cometary dust, showing a more distinct contrast between the evolved, hardened particles and fresh, likely fluffier dust particles compared to the optical polarimetry [@Kolokolova2010]. As of now, a dozen comets have been polarimetrically observed at NIR, with a bias toward dynamically new or long-periodic comets. Three of them were observed with multiple NIR filters (quasi-)simultaneously (i.e., comets C/1995 O1 (Hale-Bopp), 1P/Halley, and C/1975 V1 (West)), with the last one being the only comet observed at large phase angles of $\alpha$ &gt; 70 (@Kiselev2015 and references therein). Here, we present the results of tracking the evolution of the activity of comet 252P/LINEAR (hereafter “252P”) over four months in 2016, with multiband NIR imaging polarimetric and optical imaging observations. 252P attracted attention due to its close approach to the Earth ($\sim$0.036 au) and its possible dynamical pairing with comet P/2016 BA14 (PanSTARRS). Additionally, 252P had been one of the weakly active comets [@Ye2016a], despite its recent ejection into the near-Earth orbit ($q$ &lt; 1.3 au, where $q$ denotes the perihelion distance), which led @Ye2016a [@Ye2016b] to suggest that 252P probably has a volatile-poor origin. However, we questioned such a scenario for the comet based on (i) the strong C$_{\rm 2}$-rich coma of 252P observed in 2016 [@McKay2017], (ii) observational evidence suggesting the existence of icy chunks near the nucleus of the comet [@Li2017; @Coulson2017], and (iii) an estimated elapsed time of 252P in the near-Earth orbit ($\sim$3 $\times$ 10$^{\rm 2}$ yr; @Tancredi2014), which could suffice for the growth of the dust mantle ($\sim$1–10 yr near the Sun; @Jewitt2002 [@Kwon2016]). What if the long-lasting dormancy of 252P might originate in insulation by the sturdy mantle, not in a lack of volatiles? How does the light scattered by such evolved dust particles behave in comparison with the light scattered by freshly ejected particles from the interior? Our science objective is to characterize the activity of this near-Earth comet in its 2016 apparition. The favorable observational geometry of the comet in 2016 enabled us to sample the bright short-periodic comet with a high spatial resolution using ground-based telescopes. This paper is divided as follows. Section \[sec:obsdata\] describes the observational methods and data analyses, and Section \[sec:res1\] presents the results. Compared with the optical imaging observations conducted over $\sim$four months, the NIR polarimetric observations only overlap for a short period of $\sim$two weeks before the perihelion passage. Hence in Section \[sec:res1\], we cope with the observational results of the data both taken at pre-perihelion, which are discussed in Section \[sec:discuss\]. Finally, Section \[sec:sum\] presents a summary. The collateral post-perihelion photometric results are present in Appendix \[sec:postphot\], showing the optical colorimetric results favorable for the presence of icy particles, and the results of backward dynamical simulation and discussion on the size of the comet are described in Appendices \[sec:orbit\] and \[sec:size\], respectively. -1ex -- -------------- --------------- -- ----- ------- --------- --------- -------- ------- Median UT 2016+ 02/13 11:00 90 30.0 1.085 0.215 57.7 322.7 02/22 10:53 100 30.0 1.042 0.168 67.1 332.8 02/24 10:48 50 60.0 1.034 1.057 69.2 335.2 02/26 10:38 50 60.0 1.027 0.146 71.3 337.6 02/28 11:12 25 60.0 1.021 0.135 73.4 340.0 03/01 10:44 50 60.0 1.015 0.125 75.4 342.5 03/05 10:39 100 30.0 1.006 0.103 79.4 347.5 03/07 10:44 65 30.0 1.002 0.092 81.3 350.0 03/09 11:04 100 5.0 0.999 0.081 83.1 352.6 03/04 20:16 60 20.0 1.007 0.106 78.8 346.7 03/09 20:27 60 10.0 0.999 0.079 83.5 353.1 03/11 18:12 20 5.0 0.997 0.069 85.0 355.5 03/12 21:19 100 5.0 0.997 0.063 85.8 356.9 03/13 19:43 100 5.0 0.996 0.059 86.3 358.1 03/14 18:49 100 5.0 0.996 0.054 86.7 359.4 (Perihelion) (03/15 06:38) (…) (…) (0.996) (0.052) (86.8) (0.0) 03/15 19:56 100 5.0 0.996 0.050 87.0 0.7 03/24 19:47 12 30.0 1.005 0.041 77.6 12.2 03/25 19:31 40 60.0 1.007 0.044 76.1 13.4 03/26 19:20 50 60.0 1.009 0.048 74.8 14.7 03/27 18:53 100 60.0 1.012 0.052 73.5 15.9 03/28 18:55 95 60.0 1.014 0.056 72.2 17.1 04/05 19:30 35 60.0 1.041 0.098 63.3 26.8 04/11 18:09 145 60.0 1.068 0.132 57.1 33.6 04/15 17:24 150 60.0 1.089 0.154 53.0 38.0 04/17 17:15 65 60.0 1.100 0.166 51.0 40.1 04/18 17:44 180 60.0 1.106 0.171 50.0 41.2 04/19 16:49 80 60.0 1.112 0.177 49.0 42.2 04/22 17:02 30 120.0 1.130 0.194 46.1 45.2 04/24 17:38 65 120.0 1.143 0.206 44.2 47.2 04/25 16:59 80 120.0 1.150 0.212 43.3 48.1 04/28 16:31 50 120.0 1.170 0.229 40.5 51.0 04/29 17:18 75 120.0 1.178 0.235 39.6 51.9 04/30 17:14 85 120.0 1.185 0.241 38.7 52.8 05/02 16:56 30 120.0 1.199 0.253 37.0 54.6 05/04 17:15 60 120.0 1.214 0.266 35.2 56.4 05/11 17:15 70 120.0 1.268 0.310 29.8 62.1 05/12 17:13 70 120.0 1.276 0.317 29.1 62.9 05/14 17:17 45 120.0 1.292 0.331 27.7 64.5 05/17 17:19 20 120.0 1.316 0.352 25.8 66.7 05/18 16:34 80 120.0 1.324 0.359 25.2 67.4 05/20 16:37 75 120.0 1.341 0.373 24.2 68.9 06/02 18:00 9 120.0 1.453 0.480 19.6 77.4 06/10 17:58 15 120.0 1.524 0.555 19.1 82.0 -- -------------- --------------- -- ----- ------- --------- --------- -------- ------- \[t1\] -1ex -1ex UT 2016+ Filter $P$ $^{\rm a}$ $\sigma_{\rm P}$ $^{\rm b}$ $\theta_{\rm P}$ $^{\rm c}$ $\sigma_{\theta_{\rm P}}$ $^{\rm d}$ $P_{\rm r}$ $^{\rm e}$ $\sigma_{P_{\rm r}}$ $^{\rm f}$ $\theta_{\rm r}$ $^{\rm g}$ $m_{\rm IRSF}$ $^{\rm h}$ --------------------------- ------------- ---------------- ----------------------------- ----------------------------- -------------------------------------- ------------------------ --------------------------------- ----------------------------- --------------------------- -- MAR 04 $J$ 19.46 0.85 $-$0.68 1.01 18.12 1.41 10.69 13.28 $\pm$ 0.19 ($\phi$$^{\rm i}$ = 78.7) $H$ 19.57 0.79 $-$0.46 0.47 18.17 0.92 10.91 12.60 $\pm$ 0.18 $J$ 26.04 0.99 $-$0.34 0.29 25.19 0.99 7.33 12.49 $\pm$ 0.17 $H$ 24.00 0.91 $-$0.38 0.18 23.23 0.90 7.28 11.80 $\pm$ 0.11 ($\phi$ = 82.4) $K_{\rm S}$ 24.05 0.96 $-$0.24 0.54 23.25 1.04 7.43 11.60 $\pm$ 0.15 $J$ 28.19 1.30 $-$0.43 0.37 27.61 1.31 5.79 12.74 $\pm$ 0.12 $H$ 27.88 1.13 $-$0.37 0.10 27.31 1.11 5.85 12.24 $\pm$ 0.10 ($\phi$ = 83.9) $K_{\rm S}$ 27.74 1.56 $-$0.15 1.85 27.12 2.11 6.10 12.08 $\pm$ 0.20 $J$ 32.57 1.22 0.27 0.39 31.92 1.23 5.73 11.13 $\pm$ 0.11 $H$ 31.69 1.14 0.23 0.36 31.07 1.16 5.69 11.20 $\pm$ 0.11 ($\phi$ = 84.5) $K_{\rm S}$ 27.35 1.06 0.24 0.51 26.81 1.10 5.70 11.58 $\pm$ 0.12 $J$ 29.60 1.11 0.21 0.44 29.13 1.13 5.13 11.12 $\pm$ 0.13 $H$ 28.73 1.03 0.21 0.36 28.27 1.04 5.13 11.22 $\pm$ 0.10 ($\phi$ = 85.2) $K_{\rm S}$ 28.36 1.06 0.18 0.46 27.92 1.09 5.10 11.29 $\pm$ 0.11 $J$ 27.37 1.02 0.21 0.46 27.02 1.05 4.64 11.86 $\pm$ 0.17 $H$ 26.03 0.94 0.23 0.36 25.69 0.95 4.66 11.41 $\pm$ 0.10 ($\phi$ = 85.6) $K_{\rm S}$ 26.19 0.98 0.19 0.56 25.85 1.02 4.62 11.27 $\pm$ 0.20 $J$ 28.10 1.11 0.19 0.59 27.73 1.15 4.68 11.84 $\pm$ 0.12 $H$ 27.28 1.01 0.26 0.38 26.91 1.02 4.75 11.37 $\pm$ 0.10 ($\phi$ = 85.9) $K_{\rm S}$ 26.87 1.05 0.29 0.46 26.49 1.08 4.77 11.25 $\pm$ 0.19 \[t2\] -1ex Observations and data analyses \[sec:obsdata\] ============================================== Observations \[sec:obs\] ------------------------ The journal of our observational geometry and instrument settings is shown in Table \[t1\]. NIR imaging polarimetric observations were conducted from UT 2016 March 02 to March 15 in daily cadence using a polarimeter SIRPOL of the NIR camera SIRIUS attached to the InfraRed Survey Facility (IRSF) 1.4 m diameter telescope at the South African Astronomical Observatory (3222$\arcmin$46$\arcsec$E, $-$2048$\arcmin$39$\arcsec$S, 1798 m), Sutherland, South Africa. SIRPOL, a single-beam polarimeter of a half-wave plate (HWP) rotator with a wire grid polarizer located upstream of the SIRIUS camera at the Cassegrain focus, simultaneously provides multiband NIR wide-field polarimetry (7.7$\times$ 7.7, with the image scale of 0.45pixel$^{\rm -1}$) at the $J$ (center wavelength $\lambda_{\rm C}$ = 1.25 $\mu$m), $H$ ($\lambda_{\rm C}$ = 1.65 $\mu$m), and $K_{\rm S}$ ($\lambda_{\rm C}$ = 2.25 $\mu$m) bands [@Nagayama2003; @Kandori2006]. A microcontroller controls the HWP rotator angle by driving a stepping motor and takes images in the sequence of 0, 45, 22.5, and 67.5. Because of the unavailability of the nonsidereal tracking mode, we set the exposure time of 5–20 sec so that the elongation of the comet was smaller than $\sim$2. Target observations were conducted in units of 10 sets, each of which consists of exposures at four HWP angles (0, 45, 22.5, and 67.5) dithered by $\sim$30 pixels (13.5, which was large enough to avoid the coma signals). We interleaved sky observations $+$400 pixels in x and $-$400 pixels in y directions ($\pm$3$\arcmin$) off the detector center with every 10 sets of target observations. Seeing ranged from 0.9–1.8and was primarily 1.2. Preperihelion Johnson $R$ band ($\lambda_{\rm C}$ = 0.62 $\mu$m) imaging observations were obtained from UT 2016 February 13 to March 09 using the 0.4 m diameter Lee Sang Gak telescope (LSGT) at the Siding Spring Observatory (14903$\arcmin$52$\arcsec$E, $-$3116$\arcmin$24$\arcsec$S, 1165m), New South Wales, Australia. We used the SBIG ST-10 camera, the CCD chip of which has 2184 $\times$ 1472 pixels with a 6.8 $\mu$m pixel pitch. It provides an image scale of 0.48with a field of view of 17.5$\times$ 11.8[@Im2015]. Because of the unavailability of the nonsidereal tracking mode, we set the exposure time of 5–60 sec so that the elongation of the comet was smaller than $\sim$1. Seeing was approximately 1.2. Postperihelion simultaneous multiband (Sloan Digital Sky Survey (SDSS) g$'$ and Johnson-Cousins $R_{\rm C}$ and $I_{\rm C}$ bands, the $\lambda_{\rm C}$ of which are 0.48, 0.66, and 0.80 $\mu$m, respectively) imaging observations were performed from UT 2016 March 24 to June 10 using the 0.5 m diameter telescope at the Okayama Astrophysical Observatory (OAO-50; 13335$\arcmin$36$\arcsec$E, $+$3434$\arcmin$33$\arcsec$N, 360 m), Okayama, Japan. We employed the Multicolor Imaging Telescopes for Survey and Monstrous Explosions (MITSuME) system, which consists of three 1024 $\times$ 1024 CCD chips with a 24.0 $\mu$m pixel pitch [@Kotani2005]. It covers the field of view of 26$\arcmin$ $\times$ 26$\arcmin$ with a pixel resolution of 1.53. Nonsidereal guiding was conducted so that we obtained integrations of 30–120 sec, each depending on the signal-to-noise ratio (SNR) of the comet. Airmass was variable in the range 1.1–5.1and seeing was primarily 2.5. Data analyses \[sec:data\] -------------------------- The raw observational (polarimetric and photometric) data were preprocessed with standard techniques for imaging data: bias and dark subtractions, flat-fielding, frame registration, and sky subtraction in IRAF. The pixel coordinates of the images were converted into celestial coordinates using WCSTools [@Mink1997], with field stars listed in the 2MASS catalog [@Skrutskie2006] for the IRSF data and with stars listed in the UCAC-3 catalog [@Zacharias2010] for the LSGT and OAO data. Since incomplete sky subtraction can produce a spurious false degree of linear polarization ($P$) signals at NIR, we took special care for the background subtraction at the level of each HWP angle image. Each set of sky frames was median-combined by matching the stars with identical wcs information and was subtracted from the target images taken at the same HWP angle. If there are still remaining counts after reduction by the standard sky subtraction process, we subtracted an arbitrary constant value measured at areas of blank sky well outside the coma to make the sky count nearly zero. The resulting object images were median-combined nightly by matching the instantaneous location of the comet as the center to eliminate contaminations (e.g., noise from background stars and cosmic rays) and to improve the SNR. In this process, we used data only if the SNR of the comet in a single image exceeds three. The $P$ of the comet was derived by performing photometry of focused images using APPHOT in IRAF. Since the SNR of the resulting combined images ($\lesssim$ 80) is not high enough to perform imaging polarimetry, we decided to perform aperture polarimetry. Therefore, we integrated all signals within an aperture, losing the spatial information inside. We implemented a photopolarimetric aperture of 5 pixels (2.25, corresponding to the projected physical distance of 82–173 km during the polarimetric observation) in radius in all ($J$, $H$, and $K_{\rm S}$) bands. The Stokes $I$ and, subsequently, Stokes $Q$ and $U$ were calculated using $$I = \frac{(I_{\rm 0} + I_{\rm 45} + I_{\rm 22.5} + I_{\rm 67.5})}{2}~, \label{eq:eq1}$$ and $$Q = I_{\rm 0} - I_{\rm 45}, ~~~~~~~U = I_{\rm 22.5} - I_{\rm 67.5} ~, \label{eq:eq2}$$ where $I_{\rm X}$ denotes the intensity taken at a HWP angle of X in degrees. The $P$ and the polarization position angle of the strongest electric vector $\theta_{\rm P}$ can be calculated using $$P = \frac{\sqrt{Q^{\rm 2} + U^{\rm 2}}}{I}~, \label{eq:eq3}$$ and $$\theta_{\rm P} = \frac{1}{2}~{\rm arctan}~\biggl(\frac{U}{Q}\biggl)~. \label{eq:eq4}$$ The above quantities ought to be corrected for instrumental effects, that is, for internal polarization, polarization efficiencies of each filter and position angle offset of SIRPOL. The polarization efficiencies of the $J$, $H$, and $K_{\rm S}$ filters are known to be 95.5 %, 96.3 %, and 98.5 %, respectively, and the polarization position angle offset of the instrument is 105 in the bands [@Kandori2006]. For instrumental polarization ($P_{\rm inst}$), as mentioned in detail in previous studies using SIRPOL (e.g., @Kandori2006 [@Kwon2015]) as well as in the recently updated results [@Kusune2015], the $P_{\rm inst}$ of SIRPOL has been considered to be negligible. Since 252P was observed at high phase angles ($\alpha$ = 78.8–87.0), where the majority of comets exhibit $P$ $\gtrsim$ 20 % [@Kiselev2015], an influence of $P_{\rm inst}$ on our results should be inconsequential. Finally, the corrected $P$ and $\theta_{\rm P}$ were converted into quantities with respect to the scattering plane (the plane on which the Sun–comet–Earth are located) as follows: $$P_{\rm r} = P~{\rm cos}~(2 \theta_{\rm r})~, \label{eq:eq5}$$ and $$\theta_{\rm r} = \theta_{\rm P} - \biggl(\phi \pm \frac{\pi}{2}\biggl)~, \label{eq:eq6}$$ where $\phi$ represents the position angle of the scattering plane, the sign of which ($\pm$ in Eq. \[eq:eq6\]) was chosen to satisfy 0 $\le$ ($\phi$ $\pm$ $\pi$/2) $\le$ $\pi$ [@Chernova1993]. Uncertainties were estimated in a standard error propagation law. We tabulated the resulting polarimetric parameters and their errors in Table \[t2\]. For polarimetric analyses, we excluded all data taken on UT 2016 March 02–03 and the data in the $K_{\rm S}$ band on March 04 because of the faint comet brightness (SNR $\lesssim$ 3) and temporal malfunction of the $K_{\rm S}$ band detector, respectively, and the data on March 05–08 because of high and fluctuating humidity conditions. Photometric aperture sizes of the imaging (LSGT and OAO) and Stokes $I$ (IRSF) data were all restricted to the projected physical distance of 1000 km from the center of the nucleus (corresponding to 13–35 pixels for the LSGT data, 2–22 pixels for the OAO data, and 29–61 pixels for the IRSF data during the observing epochs) in radius. Sky brightness was subtracted by the circular annuli with widths of 5 pixels just outside the employed aperture. Flux calibration of the g$'$$R_{\rm C}I_{\rm C}$ filters was obtained from images of nearby comparison stars listed in the AAVSO Photometric All Sky Survey DR9 catalog (APASS; @Henden2016). We assumed a systematic error of the catalog of $\sim$0.1 mag to take into account the calibration uncertainties [@Tonry2018]. SDSS magnitudes of the APASS catalog were transformed to the Johnson-Cousins magnitudes using the prescriptions for the main-sequence stars by @Jester2005 and @Jordi2005. In addition, we converted the IRSF magnitude system, which follows the Mauna Kea Observatories Near-Infrared filter system [@Tokunaga2002], into the Johnson-Cousins $R_{\rm C}$ magnitudes to know the approximate trend of the cometary activity during the IRSF observations compared to the optical data. We analyzed the Stokes $I$ of the $H$-band data as a representative, simply due to the highest SNR of the comet therein. Flux calibration of the IRSF magnitudes was obtained from the field stars of the 2MASS Point Source Catalog [@Curti2003] using the conversion equations of @Kato2007. The $H$ magnitudes were then converted into the Johnson $R$ magnitudes, assuming a solar-like spectrum of the comet for simplicity (i.e., $R-H$ = 1.055; @Binney1998), and finally converted into the $R_{\rm C}$ magnitudes using the prescriptions of $R$ $-$ $R_{\rm C}$ = $-$0.17 [@Fernie1983]. We tabulated the apparent magnitudes of the IRSF data in Table \[t2\]. ![image](F1.eps){width="13cm"} -1ex Results \[sec:res1\] ==================== -- ------------- ------- -- ------------------ -- -------------------- -- ------------------ Median UT 2016+ 02/13 11:00 6.41 18.67 $\pm$ 0.30 6.78 $\pm$ 2.50 0.38 $\pm$ 0.14 02/22 10:53 8.21 18.23 $\pm$ 0.30 7.51 $\pm$ 2.50 0.43 $\pm$ 0.14 02/24 10:48 8.78 18.53 $\pm$ 0.30 5.33 $\pm$ 2.45 0.30 $\pm$ 0.23 02/26 10:38 9.43 18.13 $\pm$ 0.24 7.23 $\pm$ 2.20 0.41 $\pm$ 0.14 02/28 11:12 10.19 17.76 $\pm$ 0.25 9.48 $\pm$ 2.25 0.55 $\pm$ 0.16 03/01 10:44 11.07 17.78 $\pm$ 0.25 8.70 $\pm$ 2.20 0.50 $\pm$ 0.15 03/05 10:39 13.44 15.88 $\pm$ 0.18 44.14 $\pm$ 1.48 2.85 $\pm$ 1.02 03/07 10:44 15.03 15.81 $\pm$ 0.14 44.10 $\pm$ 1.25 2.82 $\pm$ 1.00 03/09 11:04 17.02 15.76 $\pm$ 0.14 43.56 $\pm$ 1.24 2.62 $\pm$ 0.95 03/24 19:47 34.30 15.42 $\pm$ 0.12 71.05 $\pm$ 3.24 8.13 $\pm$ 1.61 03/25 19:31 31.96 16.64 $\pm$ 0.13 24.27 $\pm$ 3.36 3.38 $\pm$ 1.65 03/26 19:20 29.30 17.08 $\pm$ 0.22 16.94 $\pm$ 4.25 1.62 $\pm$ 1.12 03/27 18:53 27.05 17.33 $\pm$ 0.19 14.07 $\pm$ 3.79 1.08 $\pm$ 0.97 03/28 18:55 25.11 16.90 $\pm$ 0.25 21.77 $\pm$ 4.55 1.43 $\pm$ 1.27 04/05 19:30 14.35 15.92 $\pm$ 0.14 71.04 $\pm$ 3.40 2.23 $\pm$ 1.69 04/11 18:09 10.65 14.89 $\pm$ 0.13 225.84 $\pm$ 3.36 3.94 $\pm$ 1.65 04/15 17:24 9.13 14.75 $\pm$ 0.16 291.78 $\pm$ 5.15 4.64 $\pm$ 1.79 04/17 17:15 8.47 15.27 $\pm$ 0.26 192.36 $\pm$ 4.60 3.08 $\pm$ 2.30 04/18 17:44 8.22 14.44 $\pm$ 0.17 427.23 $\pm$ 8.00 7.17 $\pm$ 1.83 04/19 16:49 7.95 14.55 $\pm$ 0.17 399.17 $\pm$ 7.48 6.95 $\pm$ 1.86 04/22 17:02 7.25 14.81 $\pm$ 0.26 346.12 $\pm$ 8.48 9.55 $\pm$ 2.30 04/24 17:38 6.83 14.27 $\pm$ 0.21 604.79 $\pm$ 13.15 21.04 $\pm$ 2.07 04/25 16:59 6.63 14.57 $\pm$ 0.23 473.73 $\pm$ 10.75 18.63 $\pm$ 2.14 04/28 16:31 6.14 14.99 $\pm$ 0.14 351.14 $\pm$ 5.62 13.66 $\pm$ 1.68 04/29 17:18 5.98 14.79 $\pm$ 0.12 435.16 $\pm$ 6.45 16.94 $\pm$ 1.60 04/30 17:14 5.84 14.66 $\pm$ 0.12 504.63 $\pm$ 7.54 19.77 $\pm$ 1.59 05/02 16:56 5.56 14.26 $\pm$ 0.22 768.66 $\pm$ 17.25 31.25 $\pm$ 2.09 05/04 17:15 5.29 14.56 $\pm$ 0.20 614.42 $\pm$ 12.53 25.07 $\pm$ 1.98 05/11 17:15 4.54 14.54 $\pm$ 0.19 747.90 $\pm$ 14.98 26.37 $\pm$ 1.93 05/12 17:13 4.44 14.48 $\pm$ 0.22 808.69 $\pm$ 17.57 30.03 $\pm$ 2.09 05/14 17:17 4.25 14.86 $\pm$ 0.33 596.15 $\pm$ 17.25 21.62 $\pm$ 2.66 05/17 17:19 4.00 14.87 $\pm$ 0.29 628.16 $\pm$ 16.44 21.77 $\pm$ 2.44 05/18 16:34 3.92 14.99 $\pm$ 0.30 571.81 $\pm$ 15.25 20.33 $\pm$ 2.54 05/20 16:37 3.77 15.04 $\pm$ 0.29 566.12 $\pm$ 14.69 20.24 $\pm$ 2.44 06/02 18:00 2.93 16.36 $\pm$ 0.14 194.27 $\pm$ 3.37 6.04 $\pm$ 1.68 06/10 17:58 2.53 16.55 $\pm$ 0.27 165.99 $\pm$ 4.74 4.46 $\pm$ 2.35 -- ------------- ------- -- ------------------ -- -------------------- -- ------------------ \[t3\] -1ex Photometric results I: Detection of two discontinuous activations \[sec:phot1\] ------------------------------------------------------------------------------- Figure \[Fig1\] shows the temporal evolutions of (a) the absolute magnitude $m_{\rm R}$(1, 1, 0), (b) $Af\rho$ parameter, and (c) dust mass-loss rate $\dot{M}_{\rm d}$ of 252P as a function of the Julian Date (JD) during the period from UT 2016 February 13 to June 10. Arrows in panel (a) mark two presumable points at which the comet presented discontinuous brightness enhancements. At a glance, the first activation seemed to occur between March 01 and 05 (i.e., between JD 2457448.5 and 2457452.5), which is roughly consistent with the amateur reports of the webpage by S. Yoshida[^1]. To investigate a change in the intrinsic brightness of the comet, it is necessary to eliminate the effect of the varying viewing geometry. Thus, we converted the apparent magnitudes into the reduced absolute magnitudes, which correspond to the magnitude at a hypothetical point in space ($r_{\rm H}$ = $\Delta$ = 1 au and $\alpha$ = 0), using $$m_{\rm R}(1,~1,~0)=m_{\rm R}(r_{\rm H},~\Delta,~\alpha)-5\log_{\rm 10}(r_{\rm H}\Delta)-2.5\log_{\rm 10} (\Phi(\alpha)), \label{eq:eq7}$$ where $m_{\rm R}$($r_{\rm H}$, $\Delta$, $\alpha$) is the apparent magnitude of the comet, and $\Phi(\alpha)$ is the phase function of the coma dust. We adopted a commonly used empirical scattering phase function, i.e., 2.5 $\log_{\rm 10}$$(\Phi(\alpha))$ = $b$$\alpha$, where the phase coefficient of $b$ = 0.035 mag deg$^{\rm -1}$ was assumed (see, e.g., @Lamy2004, p. 223). Although the errors associated with the Poisson noise of the OAO data were on the order of 0.001–0.01 mag, we added the standard deviations of the field stars for the differential photometry (always $\lesssim$0.2 mag) and the employed photometric error of the APASS catalog ($\sim$0.1 mag) to derive the resulting uncertainties. We then estimated the $Af\rho$ parameters, a proxy of dust production rate of the comet ($A$ is the albedo of dust particles, and $f$ is their packing density within the aperture radius of $\rho$; @A'Hearn1984) to compare the activity level in 2016 with that of the 2000 apparition [@Ye2016b] from $$Af\rho = {\rm Y}~\biggl[\frac{\Delta}{{\rm au}}\biggr]^2~\biggl[\frac{r_{\rm H}}{{\rm au}}\biggr]^2~\biggl[\frac{\rho}{{\rm cm}}\biggr]^{-1} \times 2.512^{(m_\odot -~m_{\rm R})}, \label{eq:eq8}$$ in which $m_\odot$ is the $R_{\rm C}$ band magnitude of the Sun at $r_{\rm H}$ = 1 au ($m_\odot$ = $-$27.11; @Drilling2000), Y is the unitary transformation factor of 8.95 $\times$ 10$^{\rm 26}$ for distances [@Kwon2016], and $\rho$ is the considered aperture size (10$^{\rm 3}$ km = 10$^{\rm 8}$ cm in this study). The resulting $Af\rho$ values range from 5.3–9.5 cm, with an average of 7.5 cm before the first activation, already showing an $\sim$13 times higher level a month prior to its perihelion passage compared to the 2000 apparition ($\sim$0.6 cm; @Ye2016b). Both $m_{\rm R}$(1, 1, 0) and $Af\rho$ values escalated sharply between March 01 and 05 by $\sim$2 mag and by $\sim$35 cm, respectively, implying a sudden increase in the number density of dust particles within the coma encircled by the aperture radius of 1000 km. A lower limit of the dust mass-loss rate can be estimated from the optical photometry, using an equation in @Luu1992: $$\dot{M}_{\rm d} = \frac{1.1 \times 10^{-3} \pi \rho_{\rm d} \bar{a} \eta r_{\rm obj}^{2}}{\varrho r_{\rm h}^{1/2} \Delta}, \label{eq:eq9}$$ where $\rho_{\rm d}$ is the mass density of dust particles (nominal 1000 kg m$^{\rm -3}$ was assumed), $\bar{a}$ is the average of small particle sizes (1 $\mu$m was assumed; the radius of the most effective scatterers in the optical), $r_{\rm obj}$ is the radius of the comet (300 $\pm$ 30 m; @Li2017), and $\varrho$ is the photometric aperture size in arcseconds (Table \[t3\]). $\eta$ is the ratio of the mean optical scattering cross-section of the coma dust ($C_{\rm c}$) to the nucleus cross-section ($C_{\rm n}$). We assumed a spherical nucleus with radius of $r_{\rm obj}$, i.e., $C_{\rm n}$ = $\pi r_{\rm obj}^{\rm 2}$. The $C_{\rm c}$ can be computed in the same manner as in @Luu1992: $$p_{\rm R}C_{\rm c} = 2.24\times10^{22}\pi~(r_{\rm H}\Delta)^2~10^{0.4(m_\odot -~m_{\rm R}(1, 1, 0))}, \label{eq:eq10}$$ in which $p_{\rm R}$ is the geometric albedo (0.04 was assumed). As a result, the average $\dot{M}_{\rm d}$ prior to the first activation is 0.4 $\pm$ 0.2 kg s$^{\rm -1}$, increasing by $\sim$5.5 times during the first activation. Note that the above estimates of $\dot{M}_{\rm d}$, assuming the size of the most efficient scatterers of $\sim$1 $\mu$m in the optical wavelength, could become significantly higher if we consider the large-sized dust particles (e.g., $\bar{a}$ = 100 $\mu$m–1 mm). Despite the simplified assumptions we employed, it would be informative to monitor a long-term variation of the activity level of the comet. We summarized our photometry in Table \[t3\]. ![image](F2.eps){width="13.5cm"} The above loose estimate of the activating point from the optical photometry (between March 01–05) would be narrowed using the Stokes $I$ maps of the IRSF data taken UT 2016 March 04 and 09 (open circles in Figure \[Fig1\]a). Figure \[Fig2\]a shows azimuthally averaged surface brightness profiles of 252P from the LSGT ($R$ band) and IRSF ($H$ band) data on a logarithmic scale. All brightness points were normalized to the radial distance of 1.25. Radial gradients of the points were then compared with the slopes of $m$ = $-$1 and $m$ = $-$1.5, which are typical of cometary dust expanding with initial ejection velocity under the solar radiation field [@Jewitt1987]. Compared to the triangles, both were decreasing steadily along the $m$ = $-$1 slope to $\sim$4 and along the $m$ = $-$1.5 slope outwards; squares therein show flatter distributions, and in particular, red squares on March 09 clearly show a shallower slope in the inner coma. 2 $\times$ 2 binned $H$-band images of the IRSF data are shown in panels (b) and (c), with the surface brightness being normalized between 0 and 1. A development of the central whitish-yellow part before versus after the activation is evident. On March 04, a feeble coma encircled the nucleus. The dust coma on March 09, however, was enlarged, being broadly elongated in the direction of the negative velocity vector ([**$-$v**]{}) with respect to the nucleus position (star symbol). Based on (i) the sudden increases of the photometric parameters of the LSGT data between March 01 and 05 and of the IRSF data between March 04 and 09 but (ii) nearly identical and steady radial profiles of the comet on March 01 and 04, we concluded that the first brightness enhancement most likely occurred on UT 2016 March 04–05. Similarly, we presented the multiband photometric results of the postperihelion reactivation of 252P that occurred on UT 2016 March 27–28 in the Appendix \[sec:postphot\]. Photometric results II: NIR dust color \[sec:irsfphot\] ------------------------------------------------------- From the $I$ of the IRSF data (Eq. \[eq:eq1\]), we obtained the colorimetric results of 252P dust. Figure \[Fig3\] shows temporal evolution of the $m_{\rm IRSF}$ difference of $J-H$ and $H-K_{\rm S}$ as black and purple circles, respectively. Values of the magnitude differences calculated from $m_{\rm IRSF}$ in Table \[t2\] are listed in Table \[t4\]. Overall, the temporal evolution in the NIR dust color of 252P appears to be in opposition to that of the comet’s brightness. Before the activation (March 04.84), the $J-H$ dust color was 0.68 $\pm$ 0.26 mag, which is consistent with the color range of cometary dust measured by @Jewitt1986. As time passed soon after the ignition, both $J-H$ and $H-K_{\rm S}$ values first decreased to the minima: from 0.69 $\pm$ 0.20 mag (March 09.85) and 0.50 $\pm$ 0.16 mag (March 11.76) to $-$0.07 $\pm$ 0.16 mag (March 12.89) and $-$0.10 $\pm$ 0.16 mag (March 13.82) for the $J-H$, and from 0.20 $\pm$ 0.19 mag (March 09.85) and 0.16 $\pm$ 0.22 mag (March 11.76) to $-$0.38 $\pm$ 0.16 mag (March 12.89) and $-$0.07 $\pm$ 0.15 mag (March 13.82) for the $H-K_{\rm S}$. The bluest color (i.e., minimal $m_{\rm IRSF}$ difference) occurred on March 12–13 both for the $J-H$ and $H-K_{\rm S}$, when the comet showed the maximum brightness (Figure \[Fig1\]a). Subsequently, the colors reddened back to the pre-activation values: 0.45 $\pm$ 0.20 mag (March 14.78) and 0.47 $\pm$ 0.19 mag (March 15.83) for the $J-H$, and 0.14 $\pm$0.22 mag (March 14.78) and 0.12 $\pm$ 0.21 mag (March 15.83) for the $H-K_{\rm S}$. ![Temporal evolution of the NIR color indices ($m_{\rm diff}$) of 252P dust measured from the IRSF data (Table \[t2\]). Black and purple circles denote the colors of $J-H$ and $H-K_{\rm S}$, respectively.[]{data-label="Fig3"}](F3.eps){width="9cm"} ------------- ----------------- -------------------- -------------------- -------------------- -------------------- Median UT Epoch$^{\rm 1}$ 2016+ number 03/04 20:16 1 0.68 $\pm$ 0.26 $-$ 0.13 $\pm$ 1.68 $-$ 03/09 20:27 2 0.69 $\pm$ 0.20 0.20 $\pm$ 0.19 $-$4.90 $\pm$ 1.34 0.03 $\pm$ 1.38 03/11 18:12 3 0.56 $\pm$ 0.16 0.16 $\pm$ 0.22 $-$0.75 $\pm$ 0.72 $-$0.32 $\pm$ 2.38 03/12 21:19 4 $-$0.07 $\pm$ 0.16 $-$0.38 $\pm$ 0.16 $-$2.13 $\pm$ 1.69 $-$7.10 $\pm$ 1.60 03/13 19:43 5 $-$0.10 $\pm$ 0.16 $-$0.07 $\pm$ 0.15 $-$2.15 $\pm$ 1.54 $-$0.58 $\pm$ 1.51 03/14 18:49 6 0.45 $\pm$ 0.20 0.14 $\pm$ 0.22 $-$3.33 $\pm$ 1.42 0.27 $\pm$ 1.39 03/15 19:56 7 0.47 $\pm$ 0.19 0.12 $\pm$ 0.21 $-$2.05 $\pm$ 1.54 $-$0.70 $\pm$ 1.49 ------------- ----------------- -------------------- -------------------- -------------------- -------------------- \[t4\] -1ex Polarimetric results I: phase angle dependence \[sec:pol1\] ----------------------------------------------------------- Low albedos and porous aggregate structures of cometary dust particles have led to a general dependence of $P$ with respect to the scattering plane ($P_{\rm r}$) on the phase angle $\alpha$, parameterized by a shallow branch of negative polarization with an average minimum polarization $P_{\rm min}$ $\approx$ $-$1.5 % at $\alpha_{\rm min}$ $\approx$ 10 , an inversion angle $\alpha_{\rm 0}$ at $\alpha$ $\approx$ 22, and a maximum polarization $P_{\rm max}$ $\approx$ 25–30 % at $\alpha_{\rm max}$ $\approx$ 95 in the optical and NIR [@Kiselev2015]. To investigate the polarimetric behavior of 252P, we first gleaned the archival NIR polarimetric data of cometary dust from the database of comet polarimetry (DOCP; @Kiselev2010, as shown in Figure 22.3 of @Kiselev2015) and later literature [@Kuroda2015; @Kwon2017]. Fitting the average phase curve in each NIR band was obtained by employing the empirical trigonometric function of @Penttila2005: $$P_{\rm r}(\alpha) = b~(\sin \alpha)^{c_{\rm 1}} \times \cos \left(\frac{\alpha}{2}\right)^{c_{\rm 2}} \times \sin (\alpha - \alpha_{\rm 0})~, \label{eq:eq11}$$ where $b$, $c_{\rm 1}$, $c_{\rm 2}$, and $\alpha_{\rm 0}$ are the wavelength-dependent parameters for characterizing the $P_{\rm r}$–$\alpha$ dependence. The best fit (minimum $\chi$$^{\rm 2}$) parameters weighted by the square of the errors in the $J$, $H$, and $K$ ($K_{\rm S}$) bands are described in the captions of Figures \[Fig4\], \[Fig5\], and \[Fig6\], respectively. We consider that the fitting results of the inversion angle and maximum polarization degrees are less reliable because of the few or lack of data observed at small ($\alpha$ $\lesssim$ 25) and large ($\alpha$ $\gtrsim$ 110) phase angle regions, despite the small fitting errors. Figures \[Fig4\]–\[Fig6\] present $P_{\rm r}$ of comets in the $J$, $H$, and $K$ (and $K_{\rm S}$) bands, respectively, as a function of phase angle $\alpha$. The average $\alpha$ dependencies are shown as the solid curves with colored 3$\sigma$ areas. The curves are basically the results of ‘interpolation’; thus, the error ranges of the four fitting parameters are not as large, except for the case of the $J$ band, in which there are no available data points at $\alpha$ &lt; 30. In this case, we forced the curve to fit the point (0, 0) just for visualization. ![$P_{\rm r}$ of comets in the $J$ band ($\lambda_{\rm C}$ = 1.25 $\mu$m) as a function of phase angle. The data for comets C/1995 O1 (Hale-Bopp), C/1975 V1 (West), and 1P/Halley are quoted from the database of comet polarimetry (DOCP; @Kiselev2010), and the data for comets 209P/LINEAR and C/2013 US10 (Catalina) are from @Kuroda2015 and @Kwon2017, respectively. The red symbols show the results for 252P. The solid curve denotes the interpolated average $\alpha$ dependence of the comets, as a result of Eq. \[eq:eq11\]. The rose-colored area covers the 3$\sigma$ region of the average trend. The best fit (minimum $\chi$$^{\rm 2}$) parameters weighted by the square of the errors in $J$ band are $b$ = 38.64 $\pm$ 1.74 %, $c_{\rm 1}$ = 0.77 $\pm$ 0.31, $c_{\rm 2}$ = 0.70 $\pm$ 0.11, and $\alpha_{\rm 0}$ = 21.85$\pm$ 2.02.[]{data-label="Fig4"}](F4.eps){width="9cm"} -1ex ![Same as Figure \[Fig4\] but in the $H$ band ($\lambda_{\rm C}$ = 1.65 $\mu$m). All the comet data except those for 252P are quoted from the DOCP. The best fit (minimum $\chi$$^{\rm 2}$) parameters weighted by the square of the errors in $H$ band are $b$ = 26.57 $\pm$ 0.16 %, $c_{\rm 1}$ = 0.63 $\pm$ 0.04, $c_{\rm 2}$ = (3.41 $\pm$ 0.14) $\times$ 10$^{\rm -9}$, and $\alpha_{\rm 0}$ = $\alpha_{\rm 0}$ = 18.76$\pm$ 0.41.[]{data-label="Fig5"}](F5.eps){width="9cm"} -1ex ![Same as Figure \[Fig4\] but in the $K$ and $K_{\rm S}$ bands ($\lambda_{\rm C}$ = 2.20 and 2.25 $\mu$m, respectively). All comet data with the exception of those for C/2013 US10 (Catalina) [@Kwon2017] and 252P are quoted from the DOCP. The best fit (minimum $\chi$$^{\rm 2}$) parameters weighted by the square of the errors in $K$ and $K_{\rm S}$ bands are $b$ = 31.47 $\pm$ 0.18 %, $c_{\rm 1}$ = 0.65 $\pm$ 0.06, $c_{\rm 2}$ = (6.55 $\pm$ 0.18) $\times$ 10$^{\rm -10}$, and $\alpha_{\rm 0}$ = 16.97$\pm$ 0.44.[]{data-label="Fig6"}](F6.eps){width="9cm"} -1ex At first glance, the $P_{\rm r}$ values of cometary dust distribute quite homogeneously along the average phase curves, especially in the $K$ and $K_{\rm S}$ bands, regardless of the dynamical class of comets or the observational conditions (e.g., the observing geometry). The similarity in $P_{\rm r}$ may imply that dust grains of comets have similarities to a large extent in the physical (e.g., structure and size distribution) and/or compositional properties (e.g., complex refractive index). After a careful examination, however, we found that some comets still show deviations from the average trends. In particular, long period comet C/1995 O1 (Hale-Bopp) consistently shows higher $P_{\rm r}$($\alpha$) values than the majority of comets in all NIR bands [@Jones2000], let alone in the optical [@Kwon2017], whereas short period comets 10P/Tempel 2 and 55P/Tempel-Tuttle [@Kelley2004] apparently have slightly lower $P_{\rm r}$($\alpha$) values than the average in the $K$ band. It is noteworthy that 252P showed an abrupt change in $P_{\rm r}$($\alpha$) during our observations, especially before and soon after the sudden activation on March 04–05. In Figures \[Fig4\] and \[Fig5\], $P_{\rm r}$ values on March 04 (i.e., before the first activation) were significantly lower (by $\sim$7 % in the $J$ and by $\sim$5 % in the $H$ bands, but no available $K_{\rm S}$ band data due to the low SNR of &lt; 3) than the expected $P_{\rm r}$ at given $\alpha$ (i.e., the fitted curves). Throughout the activation, however, $P_{\rm r}$ increased by $\sim$13 % to the maxima on March 12–13. The $P_{\rm r}$ at perihelion ($\alpha$ $\approx$ 87 on March 15.38) seems to return to normal. Such a temporal $P_{\rm r}$($\alpha$) change of the comet apparently looks similar to the patterns of photometrically driven parameters, i.e., $m_{\rm R}$, Af$\rho$, and $\dot{M}_{\rm d}$(Section \[sec:phot1\]). The polarization vectors of the comet ($\theta_{\rm r}$) were roughly aligned perpendicular to the scattering plane (within 4.6$-$10.9from the normal direction to the scattering plane) in the $J$, $H$, and $K_{\rm S}$ bands but presented a decreasing trend as $\alpha$ increased ($\theta_{\rm r}$ in Table \[t2\]). It should be expected that a randomly distributed scattering medium tends to signal its strongest intensity of the polarized light to the normal direction of the scattering plane (i.e., $\theta_{\rm r}$ $\sim$ 0; @Bohren1983). Breaking such a condition of coma dust might lead to a nonzero value and some systematic trends of $\theta_{\rm r}$; however, we could not provide a conclusive explanation for this phenomenon from our results. It might be due to an increase in the accuracy of measurements as the values of the Stokes parameters increase. Alternatively, one of the causes might come from a substructure in the coma of 252P, indecomposable by the aperture polarimetry in this study. A series of high-resolution images taken by HST on March 14 and April 04 2016 suggested that 252P had a strong sunward jet, the influence of which vanished at an aperture size of &gt;50–60 pixels (corresponding to 80–100 km in radial distance). As mentioned in Section \[sec:data\], our aperture size integrated all signals within the projected radial distances of 82 km (on March 04) to 173 km (on March 15) during the observations, which is similar to or slightly larger than the jet scales. ![$P_{\rm r}$ of 252P from the IRSF data (Table \[t2\]) as a function of the wavelength. Numbers on the left side of the data indicate the order of the observation epochs (1: March 04, 2: March 09, 3: March 11, 4: March 12, 5: March 13, 6: March 14, and 7: March 15), and numbers on the right side are the phase angles of the corresponding epochs. Red on the fourth epoch (March 12) denotes the day when the comet shows the maximum NIR brightness (open circles in Figure \[Fig1\]a). []{data-label="Fig7"}](F7.eps){width="7cm"} -1ex Polarimetric result II: spectral dependence \[sec:pol2\] -------------------------------------------------------- In addition to the $\alpha$-dependence, $P_{\rm r}$ exhibits spectral dependence (so-called ‘polarimetric color’, [*PC*]{}), which is defined as $$PC = \frac{\Delta P_{\rm r}}{\Delta \lambda} = \frac{P_{\rm r}(\lambda_{\rm 2}) - P_{\rm r}(\lambda_{\rm 1})}{\lambda_{\rm 2} - \lambda_{\rm 1}}, \label{eq:eq12}$$ where $P_{\rm r}$($\lambda_{\rm 1}$) and $P_{\rm r}$($\lambda_{\rm 2}$) are the $P_{\rm r}$ values in percent measured at the center wavelengths of $\lambda_{\rm 1}$ $\mu$m and $\lambda_{\rm 2}$ $\mu$m, respectively, and $\Delta \lambda$ is the difference between the two wavelengths ($\lambda_{\rm 2}$$-$$\lambda_{\rm 1}$, when $\lambda_{\rm 2}$ &gt; $\lambda_{\rm 1}$) in $\mu$m. Positive [*PC*]{} is conventionally labeled as ‘red’, and negative [*PC*]{} is labeled as ‘blue’. [*PC*]{} also depends on the phase angle, but, in general, [*PC*]{} of cometary dust is red over $\lambda$ = 0.5–1.6 $\mu$m and seems to turn blue at longer wavelengths at $\alpha$ &gt; 25 [@Kolokolova2004]. Figure \[Fig7\] shows the polarimetric color of 252P from the IRSF data. A sequence of the observation epochs and corresponding phase angles are shown on the left and right sides of the data points, respectively. Red on the fourth epoch (March 12) indicates the day when 252P showed the maximum NIR brightness during the IRSF observation (Figure \[Fig1\]a). The derived values of the [*PC*]{} and their errors are listed in Table \[t4\]. Before the activation, [*PC*]{} was nearly neutral (1; March 04). After the activation (2–7; March 09–15), however, blue [*PC*]{} became dominant over the $J$ and $H$ bands (i.e., $-$2.55 % $\mu$m$^{\rm -1}$ on average), while neutral-to-blue [*PC*]{} was shown over the $H$ and $K_{\rm S}$ bands (i.e., $-$0.26 % $\mu$m$^{\rm -1}$ on average), except for the steep negative slope of [*PC*]{} at the fourth epoch (4; red line on March 12). Blue-dominating [*PC*]{} of 252P over the $J$ and $H$ bands is distinctively different from other comets observed at high phase angles, which showed moderate $P_{\rm r}$ increase (e.g., $\sim$2.9 % $\mu$m$^{\rm -1}$ for comet 1P/Halley at $\alpha$ = 65 and $\sim$3.1 % $\mu$m$^{\rm -1}$ for comet C/1975 V1 (West) at $\alpha$ = 66; @Kiselev2015) at this domain. Overall, 252P exhibits similar blue [*PC*]{} trends at NIR during the activation, except for the steepest negative [*PC*]{} over the $H$ and $K_{\rm S}$ bands at the fourth epoch (red points) when the comet exhibited the maximum brightness with the highest $P_{\rm r}$ in shorter wavelengths (open circles in Figure \[Fig1\]a). Plausible scenarios to produce the observed polarimetric properties of the comet, together with the colorimetric behaviors, will be discussed in Section \[sec:dis1\]. Discussion \[sec:discuss\] ========================== Abrupt change in polarization degrees near the first activation \[sec:dis1\] ---------------------------------------------------------------------------- As shown in Sections \[sec:pol1\] and \[sec:pol2\], both NIR polarization degree $P_{\rm r}$ and polarimetric color [*PC*]{} of 252P precipitously changed upon the first activation point only within the change in $\delta$$r_{\rm H}$ = 0.008 au, which is one of the biggest and rapid changes ever observed. It is reminiscent of comet D/1999 S4 (LINEAR), a totally disintegrated comet whose $P_{\rm r}$ in the optical increased sharply around its perihelion [@Kiselev2002]. By definition, $P_{\rm r}$ is the ratio of the difference of light intensities measured between perpendicular and parallel directions against the scattering plane to the total intensity. A change in $P_{\rm r}$ thus indicates a change of physical and/or compositional properties of dust particles responsible for light scattering as the comet approached the Sun. Since we observed the comet with the identical instrument in simultaneous multiband imaging polarimetric mode, therewith correcting the systematic effects, it is likely that we have detected the real phenomena. We considered three cases that could be affecting the polarimetric variations, together with the NIR colorimetric results (Section \[sec:irsfphot\]): changes in the (i) composition, (ii) effective size and (iii) porosity of dust particles. ### Change in composition of dust particles \[sec:dis1-1\] The primary constituents of cometary dust are silicates (mainly in the form of olivine and pyroxene), carbonaceous materials (amorphous carbon and organics), Fe-bearing sulfides, and ice [@Levasseur-Regourd2018]. Unlike asteroids, the optical properties of which are largely differentiated with respect to heliocentric distance [@DeMeo2014; @Belskaya2017], comets seem to possess fairly homogeneous bulk optical properties with a low geometric albedo of $\sim$0.04 [@Lamy2004], showing a nearly uniform distribution on the polarimetric phase curves, regardless of dynamical classes [@Kwon2017]. However, any difference in a comet might be expected if we consider the grain properties with depth in the nucleus, perhaps as a result of mainly solar irradiation on the outer layers. Notably, solar heating tends to deplete near-surface volatiles (e.g., @Prialnik2004). As such, the observed discontinuous brightening of 252P (Section \[sec:phot1\]), as well as its jet structure detected in a high-resolution images [@Li2017] in 2016 apparition, might cast a possibility for fresh particles to be ejected from the interior of the nucleus. Laboratory experiments and numerical modeling of comet analogs showed that changes in dust compositions could make a difference to total absorptivity, which leads to a change in polarimetric properties: more transparent particles are subject to multiple scattering, so that the scattered light would be more depolarized. Absorptive particles, however, suppress the multiple scattering, which results in higher $P_{\rm r}$. Hence, increasing absorption with wavelength may lead to red [*PC*]{}, whereas decreasing absorption with wavelength may cause blue [*PC*]{} while all other properties (i.e., size and porosity) remained unchanged (@Gustafson1999 [@Kolokolova2001; @Kolokolova2004; @Kimura2006] and references therein). Therefore, to satisfy the observed increase in $P_{\rm r}$ of 252P, first, the ejection of more absorptive particles is required. For carbonaceous materials, their absorption coefficients on wavelength increases in the NIR [@Rouleau1991; @Greenberg1996; @Kolokolova1997], such that they produce a red [*PC*]{}, not our observed blue [*PC*]{}. On the contrary, the ejection of the compact icy or solid absorbing particles [@Warren1984; @Warren2019], whose scattering largely depends on the single scattering of the whole particles but not as much on the absorption coefficients as fluffy ones [@Kolokolova1997; @Gustafson1999], makes them contribute to the increase in $P_{\rm r}$ and probably to the blue [*PC*]{} if the size of the particles are as large as $>$a few tens of $\mu$m (see Section \[sec:dis1-2\] for detail discussions on the size and porosity). The idea of the ejection of compact icy particles during the discontinuous brightening around the perihelion might be reconciled with the existence of sintered subsurface icy layer [@Kossacki1994; @Kossacki2015]. Meanwhile, the bluing of the NIR color data (Section \[sec:irsfphot\]) could put an additional constraint, if this is a purely compositional effect, that more transparent materials were ejected during the activation. However, the dominance of such material is incompatible with the sharp increase of $P_{\rm r}$, which indicates that factors other than the compositional effect should be the primary cause of the NIR color agent. In summary, variations in the composition of dust particles may not be a primary factor provoking the sudden change in polarimetric parameters of 252P preperihelion, but instead other properties, such as porosity and size of dust particles, should be considered as the responsible factors for the observed phenomena of the comet. ### Change in particle size and porosity \[sec:dis1-2\] Scattering regimes can be broadly classified into three types depending on the size parameter $X$ of dust ($X$ := 2$\pi$$a$ / $\lambda$, where $a$ is the particle radius and $\lambda$ is the observation wavelength): the Rayleigh, Mie, and geometrical optics regimes in order of increasing $a$. Relatively high $P_{\rm r}(\alpha)$ can be achieved in both the Rayleigh ($X$ $\ll$ 1; $a$ $\ll$ 0.2 $\mu$m in the $J$, $a$ $\ll$ 0.3 $\mu$m in the $H$, and $a$ $\ll$ 0.4 $\mu$m in the $K_{\rm S}$ bands) and geometrical optics ($X$ $\gg$ 10; $a$ $\gg$ 2.0 $\mu$m in the $J$, $a$ $\gg$ 2.6 $\mu$m in the $H$, and $a$ $\gg$ 3.6 $\mu$m in the $K_{\rm S}$ bands) regimes. For Rayleigh-like tiny particles in the Mie scattering regime ($X$ $\sim$ 1–10; an order of 0.1–1 $\mu$m dust in the NIR), the greater their contribution to the scattered light is, the higher the observed $P_{\rm r}(\alpha)$ and the redder [*PC*]{} [@Kolokolova2004]. Unfortunately, the dominance of such particles would not explain the observed results for 252P, particularly the observed bluing of the NIR dust color (Figure \[Fig3\]) and prevailing blue [*PC*]{} over the $J$ and $H$ bands upon the activation (Figure \[Fig7\]), and seems to conflict with the results of previous studies showing a minor role of very small-sized particles in mass density, i.e., dynamics [@Fulle2015; @Rotundi2015; @Fulle2016], and in light scattering of cometary dust [@Jewitt1986; @Kolokolova2007]. Therefore, we dismiss the possibility that the observed polarimetric changes were [*largely*]{} stimulated by the increase of Rayleigh(-like) particles, although not ruling out their possible existence evidenced by the change in the optical brightness (Section \[sec:phot1\]) and observed jet-like structure in the optical [@Li2017]. Meanwhile, such small dust particles may thermally depolarize the $K_{\rm S}$ band data (e.g., blue [*PC*]{} over the $H$ and $K_{\rm S}$ bands in Figure \[Fig7\]), but we precluded the predominance of such a possibility based on (i) the $P_{\rm r}(\alpha)$ of 252P in the $K_{\rm S}$ band near the perihelion having similar values to the points of comet C/1975 V1 (West) (gray squares; @Oishi1978) from which thermal flux was subtracted and on (ii) the case of C/1975 V1 (West), the thermal depolarization effect of which was negligible at $r_{\rm H}$ $\gtrsim$ 0.9 au [@Oishi1978]. Large dust particles in the geometrical optics regime ($a$ &gt; a few tens of $\mu$m) can be further considered with two different porosities: fluffy (porous) and compact particles, two of which are in fact two main populations of dust particles of comet 67P/Churyumov-Gerasimenko (67P) collected by Rosetta/GIADA [@Fulle2015]. In case of large fluffy particles (tensile strength of &lt;10$^{\rm 5}$ N m$^{\rm -2}$ (@Mendis1991) and high charge-to-mass ratio (@Fulle2015)), they behave similarly to that of the individual constituent particles in dynamics [@Mukai1992], in light scattering [@Kolokolova2011], and in thermal emissions [@Wooden2002; @Kolokolova2007]. Accordingly, the polarimetric properties of scattered light by such large fluffy dust particles would be akin to those by particles approaching the Rayleigh regime (but are larger than $X$ = 1) (i.e., an increase in $P_{\rm r}$ with the red [*PC*]{}), which again contradicts the observed results of 252P. Finally, we consider to what extent large, compact dust particles contributed to the observed polarimetric properties of 252P. A dominance of such particles was corroborated by the results of Rosetta/GIADA, showing that mm-to-cm-scaled compact chunks ejected near perihelion, along with the release of the upper desiccated layer with a higher refractory-to-ice mass ratio than that inside the nucleus, accounted for $>$85 % of the dust mass loss and luminosity function of the comet (e.g., @Blum2017 [@Fulle2018]). As the interaction energy between ambient dipoles (i.e., monomers) are inversely proportional to the cube of the distance between the two [@Jackson1965], less porous materials are more subject to the scattered light from neighboring dust constituents. Accordingly, as the wavelength of the incident light increases (i.e., more monomers are covered in a single wavelength), $P_{\rm r}$ of such compact particles decreases more rapidly than that of the porous particles due to the enhanced electromagnetic interaction [@Gustafson1999; @Kolokolova2010]. Compared to the non/long-periodic comets observed at similar phase angles, it is more likely that the near-surface of 252P has a paucity of small, fluffy dust particles owing to its more frequent perihelion passages in the near-Earth orbit (e.g., @Li1998 [@Kolokolova2007]). The circumstance would emerge as the unusual blue [*PC*]{} of the comet. For this reason, the dominance of signals by the ejection of large, compact particles into the coma would be attributable to both a sudden increase in $P_{\rm r}$ and the enhanced blue [*PC*]{} upon the first activation of 252P. ### Tentative conclusions on particle properties ejected during the activation The ejection of large (i.e., $a$ is at least &gt;a few tens of $\mu$m in the geometrical optics regime) and compact (likely as low as the porosity of $\sim$30–65 % of the near-surface Rosetta/Philae landing site of 67P; @Spohn2015) particles can also be deduced from the morphology of the dust coma and the color change of the apparent magnitudes of the IRSF data. Figure \[Fig2\]c shows a broad extension of the dust cloud to the trailing direction ([**-v**]{}) with respect to the photocenter, alluding to the ejection of dust particles that are insensitive to the solar radiation pressure at $r_{\rm H}$ $\lesssim$ 1 au. As such, a primary factor to distribute the ejected particles in the [**-v**]{} direction should rather be an explosive mechanism, such as a rocket force from sublimating gas with nonzero ejection velocity [@Kelley2013], which is most likely responsible for the discontinuous brightening of the comet. Concurrently, the $J-H$ and $H-K_{\rm S}$ colors of the comet decreased with time (Figure \[Fig3\]): from the typical color of cometary dust (e.g., @Jewitt1986) before the activation to the blue–neutral color. It may also support our scenario that the dust particles, which are brighter at the longer wavelengths, contributed more to the intensity of the comet than the smaller or large porous dust particles. Taken as a whole, the above conjectures may answer our first question: the evolved near-surface dust layer would primarily define the activity of the comet as long as the incoming solar heat flux is sufficiently large. Implication for evolved surface \[sec:dis2\] -------------------------------------------- Our second question is how useful NIR polarimetry would be to discriminate the behaviors of fresh and evolved dust particles. The discussion on the observed polarimetric properties largely induced by the evolved dust particles of 252P in Section \[sec:dis1\] could advocate for the potential usefulness of this approach. Returning to Figure \[Fig6\], the $P_{\rm r}$($\alpha$) of 11 comets in the K band, where the number of observed comets to date is the largest among NIR bands, follows the average trend quite well, while a careful check reveals that non/long-periodic comets tend to show, on average, higher $P_{\rm r}$($\alpha$) values, particularly than for two Jupiter-Family comets (JFCs), 10P/Tempel 2 and 55P/Tempel-Tuttle [@Kelley2004]. Interestingly, the two are characterized by weak-to-absent 10 $\mu$m silicate emission features and subtle temperature excess for the blackbody temperature [@Lynch1995; @Lynch2000], both suggesting a lack of small $\mu$m-sized and/or fluffy dust particles [@Hanner1999; @Wooden2002; @Lisse2002]. Given that JFCs orbit closer to the Sun than non/long-periodic comets, which offers a more favorable environment to develop the consolidated dust mantles on their surfaces, it would be understandable for JFCs to show lower $P_{\rm r}$($\alpha$) at NIR than those of less-heated comets. The orbital evolution of 252P in the near-Earth orbit over 300 yr or even more would be in favor of this scenario (Appendix \[sec:orbit\]). However, for the moment, we should be cautious in drawing any firm conclusions. A single $P_{\rm r}$($\alpha$) point of a comet does not provide much information on its physical and compositional properties, but rather, we need (i) (quasi-)simultaneous multiband polarimetric data to measure the [*PC*]{} of dust particles for estimation of the porosity of dust and/or (ii) comets observed at multiple observation epochs, including the high $\alpha$ region, to trace the variation of the $P_{\rm r}$($\alpha$) and [*PC*]{}. Further NIR polarimetric observations undertaken in a well-organized manner are highly desirable to investigate secular evolutions of polarimetric parameters of cometary dust on the orbital motion as well as to search for any systematic differences in them between different dynamical groups of comets. Summary \[sec:sum\] =================== We present multiband NIR imaging polarimetric observations around the perihelion passage and optical imaging observations of comet 252P/LINEAR taken over four months in its 2016 apparition. We summarized the main results as follows. 1. We detected two discontinuous brightness enhancements of 252P: one in the inbound (on UT 2016 March 04–05) orbit and the other in the outbound (March 27–28) orbit. A month prior to the perihelion passage, 252P already showed $\sim$13 times higher $Af\rho$ values than that of the 2000 apparition. 2. Upon the first activation, $m_{\rm R}(1, 1, 0)$ and $Af\rho$ of 252P derived from the optical $R$ and $R_{\rm C}$ bands data increased by $\sim$2 mag and $\sim$35 cm, respectively. The attendant $\dot{M}_{\rm d}$ for assumed 1 $\mu$m particles increased to $\sim$5.5 times the initial state. In the meantime, both the $J-H$ and $H-K_{\rm S}$ dust color of 252P had decreased, showing the bluest color in the middle of the activation. 3. Before the first activation, the $P_{\rm r}$ values of 252P were far lower than the average trend of comets at given $\alpha$, by $\sim$7 % and $\sim$5 % in the $J$ and $H$ bands, respectively. Upon and soon after the activation, however, the $P_{\rm r}$ increased in all NIR bands by $\sim$13 % at most, showing the distinctive development of the blue [*PC*]{} over 1.25–2.25 $\mu$m similar to, but fiercer than the behaviors of a fragmenting comet D/1999 S4 (LINEAR) [@Kiselev2002]. In particular, the blue-dominating [*PC*]{} at the $J$–$H$ bands ($-$2.55 % $\mu$m$^{\rm -1}$ on average) is different from other comets observed, which show moderately red [*PC*]{} at similar phase angles. 4. The most likely implication of the sudden change in the observed polarimetric properties of the comet during the activation (i.e., increase $P_{\rm r}$ with the blue [*PC*]{}) as well as the bluing of the NIR dust color would be the dominance of large (i.e., well located in the geometrical optics regime at NIR), compact particles predominantly ejected from the desiccated dust layer. The paucity of small, fluffy dust particles around the nucleus of 252P would be ascribed to the more intense solar heating effects the comet has received in the near-Earth orbit (Appendix \[sec:orbit\]) than for the less-heated non/long-periodic comets observed to date. We thank the referee, L., Kolokolova, whose careful reading and constructive comments improved our manuscript. Y.G.K. was supported by the Global Ph.D. Fellowship Program through a National Research Foundation of Korea (NRF) grant funded by the Ministry of Education (NRF-2015H1A2A1034260). This work at Seoul National University was also supported by the NRF funded by the Korean Government (MEST), No. 2015R1D1A1A01060025. J.K. was supported by Astrobiology Center of NINS. M.T. was supported by MEXT/JSPS KAKENHI grant Nos. 18H05442, 15H02063, and 22000005. The IRSF project is a collaboration between Nagoya University and the South African Astronomical Observatory (SAAO) supported by Grants-in-Aid for Scientific Research on Priority Areas (A; No. 10147207 and No. 10147214) and Optical & Near-Infrared Astronomy Inter-University Cooperation Program, from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan and the National Research Foundation (NRF) of South Africa. A’Hearn, M. F., Schleicher, D. G., Millis, R. L., Feldman, P. D., & Thompson, D. T. 1984, , 89, 579 Barnes, T. F., A’Hearn, M. F., & Kolokolova, L., (Eds.), Properties of Comet Nuclei (PDS4 Format), urn:nasa:pds:compil-comet:nuc properties::1.0, NASA Planetary Data System, 2019 Belskaya, I. N., Fornaxier, S., Tozzi, G. P., et al. 2017, , 284, 30 Biele, J., Ulamec, S., Maibaum, M., et al. 2015, Science, 349, aaa9816 Binney, J., & Merrifield, M., 1998, Galactic Astronomy, Princeton Univ. Press, Princeton, NJ Blum, J., Gundlach, B., Krause, M., et al. 2017, , 469, 755 Bohren, C., & Huffman, D., 1983, Absorption and Scattering of Light by Small Particles. Wiley & Sons, New York Bonev, T., Boehnhardt, H., & Borisov, G., 2008, A&A, 480, 277 Chambers, J. E., 1999, in Impact of Modern Dynamics in Astronomy, ed. Henrard, J., & Ferraz-Mello, S., 449 Chernova, G. P., Kiselev, N. N., & Jockers, K. 1993, , 103, 144 Coulson I., M., Cordiner, M. A., Kuan, Y.-J., et al. 2017, , 153, 169 Curti, R. M., Skrutskie, M. F., van Dyk, S., et al. 2003, VizieR Online Data Catalog. p. 2246, Available at: http://adsabs.harvard.edu/abs/2003yCat.2246....0C DeMeo, F. E., & Carry, B. 2014, Nature, 505, 629 Drilling, J. S., & Landolt, A. U. 2000, in Astrophysical Quantities, ed. A. N. Cox (4th edn.; New York: Springer), 381 Fern[á]{}ndez, J. A., Tancredi, G., Rickman, H., & Licandro, J., 1999, A&A, 352, 327 Fernie, J. D., 1983. PASP, 95, 782 Fulle, M., Corte, D., Rotundi, V., et al. 2015, ApJL, 802, 12 Fulle, M., Altobelli, N., Buratti, B., et al. 2016, , 462, 2 Fulle, M., Blum, J., Green, S. F., et al. 2018, , 482, 3326 Greenberg, J. M., & Li, A., 1996, A&A, 309, 258 Greenstreet, S., Ngo, H., & Gladman, B., 2012, , 217, 355 Gundlach, B., Ratte, J., Blum, J., Oesert, J., & Gorb, S. N., 2018, , 479, 5272 Gustafson, B.[Å]{}.S., & Kolokolova, L., 1999, J. Geophys. Res., 104, 31711 Hanner, M. S., 1999, Space Sci. Rev., 90, 99 Henden, A. A., Templeton, M., Terrell, D., et al. 2016, VizieR Online Data Catalog: II/336 Im, M., Choi, C., & Kim, K., 2015, JKAS, 48, 207 Ishiguro, M., Watanabe, J.-I., Sarugaku, Y., et al. 2010, , 714, 1324 Ishiguro, M., Kuroda, D., Hanayama, H., et al. 2016, , 152, 169 Jackson, J. D., 1965, Joh Willey & Sons, 641 Jester, S., Schneider, D. P., Richards, G. T., et al. 2005, , 130, 873 Jewitt, D., & Meech, K., 1986, , 310, 937 Jewitt, D., & Meech, K., 1987, , 317, 992 Jewitt, D. 2002, , 123, 1039 Jewitt, D. 2004, Comets II, 659 Jones, T. J., & Gehrz, R. D., 2000, , 143, 338 Jordi, K., Grebel, E. K., & Ammon, K., 2005, A&A, 460, 339 Kandori, R., Kusakabe, N., Tamura, M., et al. 2006, Proc. SPIE, 6269, 626951 Kato, D., Nagashima, C., Nagayama, T., et al. 2007, PASJ, 59, 615 Keller, H. U., Mottola, S., Hviid, S. F., et al. 2017, , 469, 357 Kelley, M. S., Woodward, C. E., Jones, T. J., Reach, W. T., & Johnson, J., 2004, , 127, 2398 Kelley, M. S., Lindler, D. J., Bodewits, D., et al. 2013, , 222, 634 Kimura, H., Kolokolova, L., & Mann, I., 2006, A&A, 449, 1243 Kiselev, N., Jockers, K., & Rosenbush, V., 2002, EM&P, 90, 167 Kiselev, N., Velichko, S., Jockers, K., Rosenbush, V., & Kikuchi, S. 2010, NASA Planetary Data System, Database of Comet Polarimetry (DOCP), EAR-C-COMPIL-5-COMET-POLARIMETRY-V1.0, 8124 Kiselev, N., Rosenbush, V., Kolokolova, L., & Levasseur-Regourd, A.-Ch. 2015, in Polarimetry of Stars and Planetary Systems, ed. L. Kolokolova, J. Hough, & A. Levasseur-Regourd (Cambridge: Cambridge Univ. Press), 379 Kolokolova, L., & Jockers, K., 1997, P&SS, 45, 1543 Kolokolova, L., Jockers, K., Gustafson, B.[Å]{}.S., & Lichtenberg, G., 2001, J. Geophys. Res., 106, 10113 Kolokolova, L., Hanner, M. S., Levasseur-Regourd, A.-C., & Gustafson, B.[Å]{}.S. 2004, in Comets II, ed. M. Festou, H. U. Keller, & H. A. Weaver (Tucson, AZ: Univ. Arizona Press), 577 Kolokolova, L., Kimura, H., Kiselev, N., & Rosenbush, V., 2007, A&A, 463, 1189 Kolokolova, L., & Kimura, H., 2010, A&A, 513, A40 Kolokolova, L., 2011, ASPC, 449, 345 Kossacki, K. J., Kömle, N. I., Kargl, G., & Steiner, G. 1994, Planet. Space Sci., 42, 393 Kossacki, K. J., Spohn, T., Hagermann, A., Kaufmann, E., K[ü]{}hrt, E., 2015, , 260, 464 Kotani, T., Kawai, N., Yanagisawa, K., et al. 2005, NCimC, 28, 755 Kuroda, D., Ishiguro, M., Watanabe, M., et al. 2015, , 814, 156 Kusune, T., Sugitani, K., Miao, J., et al. 2015, , 798, 60 Kwon, J., Tamura, M., Hough, J., et al. 2015, , 824, 85 Kwon, Y. G., Ishiguro, M., Hanayama, H., et al. 2016, , 818, 67 Kwon, Y. G., Ishiguro, M., Kuroda, D., et al. 2017, , 154, 173 Lamy, P. L., Toth, I., Fernandez, Y. R., & Weaver, H. A. 2004, in Comets II, ed. M. C. Festou et al. (Tuczon, AZ: Univ. Arizona Press), 223 Levasseur-Regourd A. C., Jessica, A., Herv[é]{}, C., et al., 2018, Space Sci. Rev., 214, 64 Li, A., & Greenberg, J. M., 1998, A&A, 338, 364 Li, J. Y., Kelley, M. S., Samarasinha, N. H., et al. 2017, , 154, 136 Lisse, C. M., A’Hearn, M. F., Fern[á]{}ndez, Y. R., & Peschke, S. B., 2002, in IAU Coll. 181: Dust in the Solar System and Other Planetary Systems, A Search for Trends in Cometary Dust Emission, ed. S. F. Green et al. (Canterbury, UK: ASP), 259 Luu, J. X., & Jewitt, D. C. 1992, , 104, 2243 Lynch, D. K., Hackwell, J. A., Edelsohn, D., et al. 1995, , 114, 197 Lynch, D. K., Russell, R. W., & Sitko, M. L., 2000, , 144, 187 McKay, A. J., Kelley, M. S. P., Bodewits, D., et al. 2017, ACM, 3b4, <http://acm2017.uy/abstracts/Parallel3.b.4.pdf> Meech, K. J., & Svore[ň]{}, J. 2004, in Comets II, ed. M. Festou, H. U. Keller, & H. A. Weaver (Tucson, AZ: Univ. Arizona Press), 317 Mendis, D. A., 1991, Astrophys. Space Sci., 176, 163 Mink, D. J. 1997, in ASP Conf. Ser. 125, Astronomical data Analysis Software and Systems VI, ed. G. Hunt & H. E. Payne (San Francisco, CA: ASP), 249 Morbidelli, A., Bottke, Jr., W. F., Froeschl[é]{}, C., & Michel, P., 2002, in Asteroids III, ed. W. F. Bottke, Jr. et al. (Tucson, AZ: Univ. Arizona Press), 409 Mukai, T., Ishimoto, H., Kozasa, T., Blum, J., & Greenberg, J. M., 1992, A&A, 262, 315 Nagayama, T., Nagashima, C., Nakajima, Y., et al. 2003, Proc. SPIE, 4841, 459 Naidu, S. P., Benner, L. A. M., Brozovic, M., et al. 2016, in AAS/DPS Meeting 48 Abstract, 219.05 Ohtsuka, K., Sekiguchi, T., Kinoshita, D., et al. 2006, A&A, 450, L25 Oishi, M., Kawara, K., Kobayashi, Y., et al. 1978, PASJ, 30, 149 Penttil[ä]{}, A., Lumme, K., Hadamcik, E., & Levasseur-Regourd, A.-C. 2005, A&A, 432, 1081 Perovich, D. K., 1998, , 103, 5563 Prialnik, D., Benkhoff, J., & Podolak, M. 2004, Comets II, 359 Rickman, H., Fernandez, J. A., & Gustafson, B. A. S. 1990, A&A, 237, 524 Reddy, V., 2016, NASA/JPL News, Comet Flying by Earth Observed with Radar and Infrared, 2016 March 24 Rotundi, A., Sierks, H., Della, C., et al., 2015, Science, 347, aaa3905 Rouleau, F., & Martin, P. G., 1991, , 377, 526 Rudawska, R., Vaubaillon, J., T[ó]{}th, J., & Raetz, S., 2016, in AAS/DPS Meeting 48 Abstract, 218.01 Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, , 131, 1163 Solontoi, M., Ivezi[ć]{}, $\check{Z}$., Juri[ć]{}, M., et al. 2012, , 218, 571 Southworth, R. B., & Hawkins, G. S. 1963, Smithsonian Contributions to Astrophysics (Washington D.C.: Smithsonian Institution), Vol. 7, 261 Spohn, T., Knollenber, J., Ball, A. J., et al. 2015, Science, 349, aab0464 Tancredi, G., 2014, , 234, 66 Tokunaga, A. T., Simons, D. A. & Vacca, W. D., 2002, PASP, 114, 180 Tonry, J. L., Denneau, L., Flewelling, H., et al. 2018, , 867, 105 Thomas, H., Ratke, L., & Kochan, H., 1994, Adv. Space Res. 12, 207 Warren, S. G., 1984, Appl. Opt. 23, 1206. Warren, S. G., 2019, Phil. Trans. R. Soc. A, 377:20180161 Wooden, D., 2002, EM&P, 89, 247 Ye, Q.-Z., Hui, M.-T., Brown, P. G., et al. 2016, , 264, 48 Ye, Q.-Z., Brown, P. G., & Wiegert, P. A., 2016, , 818, L29 Zacharias, N., Finch, C., Girard, T., et al. 2010, , 139, 2184 Postperihelion activation of 252P \[sec:postphot\] ================================================== photometric results ------------------- Changes in the postperihelion activity level of 252P were quantified by the photometry, as in Section \[sec:phot1\]. The intrinsic brightness of 252P (Figure \[Fig1\]a; derived from Eq. \[eq:eq7\]) soon after perihelion passage seemed to decline steeply, but it began to rebound from the minimum of $\sim$17.3 mag with a nearly 16 times flux increase on UT 2016 March 27–28 (right arrow in Figure \[Fig1\]a). Such re-ignition again became tranquil after peaking in approximately late April to early May at $\sim$14.3 mag. The changes in $Af\rho$ and $\dot{M}_{\rm d}$ patterns also showed similar trends with that of $m_{\rm R}$(1, 1, 0), despite the former two having their maxima $\sim$ a week after the peak of $m_{\rm R}$(1, 1, 0). ![image](F9.eps){width="12cm"} -1ex Over 1.5 months after the second activation, $Af\rho$ of 252P (Figure \[Fig1\]b; derived from Eq. \[eq:eq8\]) increased by $\sim$57 times from 14.1 $\pm$ 3.8 cm to 808.7 $\pm$ 17.6 cm. We compared the $Af\rho$ values with the [*Hubble Space Telescope*]{} (HST) data in the broadband F625W filter ($\lambda_{\rm C}$ = 0.625 $\mu$m) taken on UT 2016 March 14 and April 04 (16.8 $\pm$ 0.3 cm and 57 $\pm$ 1 cm, respectively, from @Li2017; cross points in Figure \[Fig1\]b). The consistent differences by $\sim$20 cm between the HST and OAO data observed at similar epochs exist, but we suspect that the different photometric aperture sizes employed (by more than two orders of magnitude, i.e., $\rho$ $\lesssim$ 10 km for the HST data and $\rho$ = 1000 km for the OAO data) might cause the difference in $Af\rho$. Although, by definition, $Af\rho$ should be a constant value regardless of the aperture size under the steady expansion in the coma, such a condition would not be met in the vicinity of the nucleus. Considering that $Af\rho$ increases outward to reach an equilibrium (to $\sim$1000 km at $r_{\rm H}$ $\sim$ 1 au; see, e.g., @Bonev2008), it might be explainable that the OAO data in this study consistently showed higher $Af\rho$ than those of the HST data. Nonetheless, the HST data can be interpreted in line with the OAO data, in that $Af\rho$ of 252P in postperihelion was higher than that in preperihelion. Similarly, variations in $\dot{M}_{\rm d}$ (Figure \[Fig1\]c; derived from Eq. \[eq:eq9\] and Eq. \[eq:eq10\]) strongly support the postperihelion reactivation of 252P, although the overall morphology of the evolutionary trend follows a concave shape, increasing and decreasing both in proportion to the third power of $r_{\rm H}$, unlike the convex shapes of the $m_{\rm R}$(1, 1, 0) and $Af\rho$ trends. From the minimum of 1.1 $\pm$ 1.0 kg s$^{\rm -1}$ on March 27 through the maximum of 31.3 $\pm$ 2.1 kg s$^{\rm -1}$ on May 02 to another minimum of 4.5 $\pm$ 2.4 kg s$^{\rm -1}$ on June 10, the total mass-loss during the second activation event was $\sim$7.8 $\times$ 10$^{\rm 7}$ kg, when we assumed nominal $\bar{a}$ = 1 $\mu$m dust grains. Taken as a whole, the derived parameters indicate that the second activation of 252P was likely on UT 2016 March 27–28. Variation in the coma color --------------------------- ![Temporal evolution in the postperihelion absolute magnitudes of 252P from the multiband OAO data. $m_{\rm X}$ denotes the absolute magnitude in X band, with an aperture size of 1000 km (corresponding to 33.6–2.5) in radius. Differences in $m_{\rm X}$ between two bands are shown as blue ($m_{\rm g'}$ $-$ $m_{\rm R_{\rm C}}$) and orange ($m_{\rm R_{\rm C}}$ $-$ $m_{\rm I_{\rm C}}$) circles. All the error bars include the 1$\sigma$ errors on the measurements, calibration of the field stars with the assumed internal catalog error of $\sim$0.1, and propagated errors in magnitude subtraction of two bands.[]{data-label="Fig10"}](F10.eps){width="9cm"} -1ex As a byproduct of our multiband imaging observation, we produced color images to clarify to what extent the fluxes from each band contributed to the postperihelion reactivation of the comet. The g$'$, $R_{\rm C}$, and $I_{\rm C}$ band images were allocated as green, blue, and red hues, respectively. Since we employed the broadband filters, gas contaminations in flux estimation would be inevitable, particularly by the molecular emissions of C$_{\rm 2}$ Swan bands in the g$'$, NH$_{\rm 2}$ $\alpha$ bands in the $R_{\rm C}$, and CN red bands in the $I_{\rm C}$ filters [@Meech2004]. Figure \[Fig9\] presents wide-field composite images, together with the dates of observation and 5$\arcmin$ scale bars in each panel, the fields of view of which are much larger than the applied photometric aperture size. In these images, greenish and bluish components spherically expanded from the nucleus once the comet became activated, whereas a whitish component elongated along the antisolar direction ([**$-$r$_\odot$**]{}). As the activity of the comet peaked, a greenish coma covered the entire field of view, developing a significant bluish green color in the innermost coma. The gas coma of 252P at this term spherically extended to &gt;10$^{\rm 4}$ km from the nucleus in the sky plane (panels (d) to (f)). Meanwhile, a reddish component was observably dominant, only after the reactivating event was over (panel (h)). Figure \[Fig10\] shows the temporal evolution of the 252P coma color in postperihelion more quantitatively. All reduced magnitudes were derived with the 1000 km aperture size (corresponding to 33.6–2.5) in radius, well inside the central-most regions of the images in Figure \[Fig9\]. The magnitude differences between the g$'-R_{\rm C}$ and $R_{\rm C}-I_{\rm C}$ bands are given as blue ($m_{\rm g'}$ $-$ $m_{\rm R_{\rm C}}$) and orange ($m_{\rm R_{\rm C}}$ $-$ $m_{\rm I_{\rm C}}$) circles, respectively. The putative activation point (on March 27–28) is marked as an arrow. Overall, a blueward color change (i.e., predominant brightening in the $R_{\rm C}$ band flux) is apparent in the inner coma. Before the activation, the comet was slightly red, showing on average $m_{\rm g'}$ $-$ $m_{\rm R_{\rm C}}$ $\sim$ 0.5 and $m_{\rm R_{\rm C}}$ $-$ $m_{\rm I_{\rm C}}$ $\sim$ 0.4 (almost identical within the error bars, though), which are physically indistinguishable from those of short-period comets in the inner solar system [@Solontoi2012]. During the activation, however, fluxes in the $R_{\rm C}$ band became significantly enhanced, resulting in the redder color of $m_{\rm g'}$ $-$ $m_{\rm R_{\rm C}}$ and more neutral color of $m_{\rm R_{\rm C}}$ $-$ $m_{\rm I_{\rm C}}$. Such magnitude differences had remained nearly constant at the value of $m_{\rm g'}$ $-$ $m_{\rm R_{\rm C}}$ = 0.9 $\pm$ 0.1 (the standard deviation of the nominal values) and $m_{\rm R_{\rm C}}$ $-$ $m_{\rm I_{\rm C}}$ = 0.2 $\pm$ 0.1 over the following 1.5 months. After peaking the maximal $Af\rho$ and $\dot{M}_{\rm d}$, the coma color again turned to reddish-neutral, albeit with large uncertainties, consistent with the status before the activation. Orbital evolution of 252P/LINEAR and its dynamical association with P/2016 BA14 (PanSTARRS) \[sec:orbit\] ========================================================================================================= In this section, we investigated orbital evolution of 252P and its dynamical association with P/2016 BA14 (PanSTARRS) to search an existence of any possible dynamical characteristics of the objects attributable to the observed photo-polarimetric properties of 252P. Comet P/2016 BA14 (PanSTARRS; hereafter “BA14”), a possible dynamical pair of 252P with remarkably similar orbital elements [@Rudawska2016], possesses a geometric albedo of &lt;3 % [@Reddy2016] and diametric size of &gt;1 km [@Naidu2016] or 1–2 km [@Li2017]. Unlike the capricious behaviors of 252P, BA14 showed a &gt;6 mag larger r$'$-band magnitude with a fraction of active area on the surface of $\sim$0.01 %, i.e., being rather asteroidal in the 2016 apparition [@Li2017]. To trace the dynamic history of the two back to the time when they were ejected into the near-Earth orbit, we conducted a backward dynamical simulation with the Mercury 6 integrator [@Chambers1999] under the gravity of the Sun and eight planets. We generated 1000 clones randomly distributed within the 1-$\sigma$ Gaussian distribution, with the orbital elements quoted on the JPL Small-Body Database Browser site. They were integrated 1000 yr backwards with 8 day time step in the general Bulirsch-Stoer mode without considering the Yarkovsky force (i.e., consistent with the negligible activity levels before the 2016 apparition). The applied orbital elements and their errors are summarized in Table \[t4\].\ Object $a$ $^{\rm a}$ $e$ $^{\rm b}$ $i$ $^{\rm c}$ $\omega$ $^{\rm d}$ $\Omega$ $^{\rm e}$ $M$ $^{\rm f}$ -------- --------------------------------- --------------------------------- --------------------------------- --------------------------------- --------------------------------- --------------------------------- 3.0470 0.6731 10.4223 343.3110 190.9499 10.5137 (6.3018 $\times$ 10$^{\rm -5}$) (6.8756 $\times$ 10$^{\rm -6}$) (6.3172 $\times$ 10$^{\rm -5}$) (5.4848 $\times$ 10$^{\rm -5}$) (6.2681 $\times$ 10$^{\rm -5}$) (3.000 $\times$ 10$^{\rm -4}$) 3.0216 0.6662 18.9181 351.9022 180.5312 40.8120 (8.4014 $\times$ 10$^{\rm -7}$) (9.2653 $\times$ 10$^{\rm -8}$) (1.7085 $\times$ 10$^{\rm -6}$) (4.5491 $\times$ 10$^{\rm -6}$) (2.209 $\times$ 10$^{\rm -6}$) (1.7161 $\times$ 10$^{\rm -5}$) \[t6\] -1ex In addition, to assess the orbital similarity of two bodies in the near-Earth orbit, we estimated a traditional [*D*]{}$_{\rm SH}$ parameter [@Southworth1963], which is a distance between the orbits of two bodies in five-dimensional orbital element space ($q$, $e$, $i$, $\omega$, and $\Omega$) calculated by $$\begin{aligned} D_{\rm SH} = {} & \Bigg[\big(e_{\rm B} - e_{\rm A}\big)^{\rm 2} + \big(q_{\rm B} - q_{\rm A}\big)^{\rm 2} \\ & + \bigg(2 \sin \frac{I_{\rm AB}}{2} \bigg)^{\rm 2} + \bigg[\big(e_{\rm A} + e_{\rm B}\big) \sin \frac{\Pi_{\rm AB}}{2}\bigg]^{\rm 2}\Bigg]^{\rm 1/2}, \label{eq:eq15} \end{aligned}$$ where $$I_{\rm AB} = \arccos \Big[ \cos i_{\rm A} \cos i_{\rm B} + \sin i_{\rm A} \sin i_{\rm B} \cos \big( \Omega_{\rm A} - \Omega_{\rm B} \big) \Big] \label{eq:eq16}$$ and $$\Pi_{\rm AB} = \big(\omega_{\rm A} - \omega_{\rm B} \big) + 2 \arcsin \Bigg[\cos \frac{(i_{\rm A} + i_{\rm B})}{2} \sin \frac{(\Omega_{\rm A} - \Omega_{\rm B})}{2} \sec \frac{I_{\rm AB}}{2} \Bigg]. \label{eq:eq17}$$ The subscripts A and B denote the two bodies to be compared (here, A for 252P and B for BA14), $q$ is the perihelion distance in au, $e$ is the eccentricity, $i$ is the inclination, $\omega$ is the argument of perihelion, and $\Omega$ is the longitude of the ascending node. In case of $D_{\rm SH}$ $\leq$ 0.20, we tend to consider that the two bodies are probably within the association range [@Southworth1963]. For instance, a minimum $D_{\rm SH}$ value of the possible pair, (3200) Phaethon and 2005 UD, shows $\sim$0.04 [@Ohtsuka2006]. ![Time evolution of (a) the perihelion distance $q$ and (b) the eccentricity $e$ of 252P. Red lines denote the evolution of a particle with the average values at the epoch (May 11.0 2016, the Julian date of 2457519.5), while each gray line represents that of an individual clone, the orbital elements of which follow 1-$\sigma$ standard deviation of the Gaussian distribution. []{data-label="Fig8"}](F8.eps){width="9cm"} -1ex Figure \[Fig8\] shows the time evolution of (a) the perihelion distance [*q*]{} and (b) the eccentricity [*e*]{} of 252P. Red lines denote the evolution of a particle with the average values at the epoch (May 11.0 2016, the Julian date of 2457519.5), while each gray line represents that of an individual clone, the orbital elements of which have 1-$\sigma$ standard deviation on the Gaussian distribution. Our simulation shows that 252P would have closely encountered Jupiter at the minimal distance of [*d*]{}$_{\rm min}$ $\sim$ 0.02 au, 314 yr ago. Due to the high uncertainty, we could only trace back the orbital evolution of the comet from the current epoch to $\sim$310 yr in the past, during which the [*q*]{} and [*e*]{} of the comet have been changed by $\delta$[*q*]{} $\sim$ $-$1 au (decreasing) and by $\delta$[*e*]{} $\sim$ +0.2 (increasing). Interestingly, based on the current orbital elements of 252P, the Near-Earth Object (NEO) model of @Greenstreet2012 suggests two most probable source regions of the comet: the JFC source ($\sim$81.8 %) and the outer main asteroidal belt ($\sim$16.2 %). In either case, 252P should have eventually been transported to the current orbit by means of the resonance with Jupiter and the perturbations of terrestrial planets (e.g. @Morbidelli2002). Based on the analysis, we conjectured that the decreasing $q$, together with its repetitive close encounters with planets (e.g., with Jupiter within $\sim$0.05 au at $\sim$320 yr ago and with the Earth within $\sim$0.02 au since $\sim$170 yr to the present), might introduce internal stresses to 252P, which would render the nucleus overall more fragile and would in turn affect the rapid polarimetric variation in its perihelion passage. The average $D_{\rm SH}$ value derived from the osculating orbital elements of 252P and BA14 is $\sim$0.16, which implies a positive dynamical correlation of two bodies, although not as tight as that of (3200) Phaethon and 2005 UD [@Ohtsuka2006]. Among all the possible combinations of the 1-$\sigma$ clones, we obtained the minimal $D_{\rm SH}$ value of $\sim$0.15. Owing to the close encounters of the objects to the planets and their non-gravitational force, it is hard to know exactly when they were fragmented each other from this study. Contrast in behaviors of large and small comet nuclei in the near-Earth orbit \[sec:size\] ========================================================================================== This section describes our conjecture of the reason why 252P and BA14 exhibited such different activity patterns during the 2016 apparition, despite their possible dynamical association resulting in similar orbital evolution in the near-Earth orbit. The remarkably higher activity of 252P in 2016 compared to the previous apparitions, alongside its discontinuous brightness enhancements near the perihelion, is far different from the behavior of the comet over the past apparitions. We conjecture that the effects of such orbital evolution on comets in near-Earth orbits might vary greatly depending on comet nuclei size. This conjecture would be exemplified in different observational characteristics of 252P and BA14 in the 2016 apparition, despite their similar orbital evolution over a few centuries (i.e., $D_{\rm SH}$ $\leq$ 0.20 in Appendix \[sec:orbit\]). 252P is on the small end of sizes (300$\pm$30 m; @Li2017) of JFCs, for which a typical comet nucleus would range from one to a few km in size [@Fernandez1999; @Barnes2019]. The surface gravity of 252P would then be less than a factor of at least one-fifth the gravity at the surface of ordinary JFCs. The concomitant smaller escape velocity permits large particles, mm–cm in size, to readily escape the gravitational influence of 252P, while a larger nucleus would be more favorable to accumulate dust particles on its surface. The small size can also contribute to the spin evolution of the nuclei such that the rotational stability tends to be proportional to the fourth power of the radius of the nucleus ($\tau_{\rm ex}$ $\propto$ $r_{\rm n}^{\rm 4}$; @Jewitt2004). 252P thus has a $\gtrsim$10$^{\rm 3}$ times smaller excitation timescale than that of a typical JFC. Indeed, a sudden increase of nongravitational force of 252P by a factor of $\sim$10 but negligible change of BA14 during the 2016 apparition [@Li2017] may reflect this higher rotational instability of the smaller comet. Taken as a whole, such consequences of the small nucleus size together with the orbital evolution in the near-Earth orbit (Appendix \[sec:orbit\]) would enhance erratic sublimation patterns of the comet, probably resulting from highly changeable mass transfer on the surface (e.g., landslides; @Keller2017) due to the enhanced solar radiation, and/or in unstable development of the dust mantle on the surface (e.g., @Rickman1990). They may in turn affect the abnormal variations of the photo-polarimetric parameters of 252P near its perihelion. [^1]: http://www.aerith.net/comet/catalog/0252P/2016.html
{ "pile_set_name": "ArXiv" }
--- abstract: 'In this paper, we find all integers $x$ such that $x^{2}-1$ has only prime factors smaller than $100$. This gives some interesting numerical corollaries. For example, for any positive integer $n$ we can find the largest positive integer $x$ such that all prime factors of each of $x, x+1,\ldots, x+n$ are less than 100.' --- [**On the largest prime factor of $x^2-1$**]{}\ Florian Luca and Filip Najman **Keywords** Pell equation, Compact representation, Lucas sequence. **Mathematics subject classification (2000)** 11D09, 11Y50.\ Introduction {#sec:1} ============ For any integer $n$ we let $P(n)$ be the largest prime factor of $n$ with the convention $P(0)=P(\pm 1)=1$. Our main result in this paper is the determination of all integers $x$ satisfying the inequality $$\label{eq:1} P(x^2-1)<100.$$ We also give some interesting corollaries to our main result. For simplicity, we will consider only solutions with positive values of $x$. Before stating the results, some history. In 1964, Lehmer [@leh] found all positive integer solutions $x$ to the inequality $P(x(x+1))\le 41$. Notice that this amounts to finding all odd positive integers $y=2x+1$ such that $P(y^2-1)\le 41$. There are $869$ such solutions. In [@Lu], the first author found all positive integer solutions of the similar looking equation $P(x^2+1)<100$. There are $156$ of them. In both [@leh] and [@Lu], the method of attack on this question was the following. Assume that $x$ is a positive integers such that $P(x^2\pm 1)\le K$, where $K=41$ if the sign is $-$ and $K=100$ if the sign is $+$. Then we can write $$\label{eq:2} x^2\pm 1=dy^2,$$ where $d$ is squarefree, and $P(dy)\le K$. When the sign is $-$, then there are $13$ possible primes $p\le 41$ which can participate in the factorization of $d$. When the sign is $+$, then $p\le 100$ but either $p=2$ or $p$ is a prime which is congruent to $1$ modulo $4$. There is a total of $12$ such primes. In both cases, we can write equation in the form $$x^2-dy^2=\mp 1.$$ Thus, our possible values for $x$ appear as the first coordinate of one of the solutions of at most $2^{13}-1=8191$ Pell equations. For a given Pell equation, the sequence $(y_n/y_1)_{n\ge 1}$ forms a Lucas sequence with real roots. The [*Primitive Divisor Theorem*]{} for Lucas sequences with real roots (see, for example, [@car], or the more general result from [@BHV] which applies to all Lucas sequences) says that if $n>6$, then $y_n$ has a prime factor which is at least as large as $n-1$. In particular, for the cases treated in [@leh] (and [@Lu]), it suffices to check the first $42$ (respectively $100$) values of the component $x$ of the Pell equations involved there and among these one finds all possible solutions of the equations considered there. We follow the same approach in the present paper. The first step is, as previously, to determine the form of all the possible solutions. This is done as explained above via the Primitive Divisor Theorem for the second component of the solutions to Pell equations. Since there are $25$ primes $p<100$, this leads to first solving $2^{25}-1$ Pell equations $x^2-dy^2=1$, the largest one of them having $d$ with $37$ decimal digits. This is clearly impossible using standard algorithms like the continued fractions, as the computations would be too slow and the fundamental solutions of some of the involved Pell equations would have hundreds of millions of decimal digits. Instead, the Pell equations are solved by first computing the regulator of the ring of integers of the corresponding quadratic field, and then from the regulator obtaining a compact representation of the fundamental solution of the Pell equation. The only algorithm known fast enough to compute the huge amount of regulators needed is Buchmann’s subexponential algorithm. The output of this algorithm gives exactly the regulator under the Generalized Riemann Hypothesis and unconditionally is only a multiple of the regulator. Next, we check which of the solutions to the Pell equations will lead to solutions to equation . Finally, a check is performed that proves our search misses no solutions, thus eliminating the apparent dependence of our results on the Generalized Riemann Hypothesis. [**Acknowledgements.**]{} Part of this work was done while F. L. visited the Mathematics Department of the University of Zagreb in February, 2009. He thanks the people of this Department for their hospitality. Both authors thank Professor Andrej Dujella and the anonymous referee for useful suggestions. F. L. was also supported in part by Grants SEP-CONACyT 79685 and PAPIIT 100508. F. N. was supported by the Ministry of Science, Education and Sports, Republic of Croatia, Grant 037-0372781-2821. Application of the Primitive Divisor Theorem {#sec:2} ============================================ Here, we explain in more detail the applicability of the Primitive Divisor Theorem alluded to in Section \[sec:1\] to our problem. Let $x$ be an integer such that $x^2-1$ is a product of only the primes up to $97$. We can then write $x^2-1=dy^2$, or, equivalently, $x^2-dy^2=1$, where $$d=2^{a_1}\cdot 3^{a_2}\cdots 97^{a_{25}},\ a_i\in \{0,1\}\text{ for }i=1, \ldots, 25.$$ The only restriction for $d$ above is that not all the $a_i$s can be zero. This is a Pell equation, so $(x,y)=(x_n,y_n)$, where $x_n+y_n\sqrt d= (x_1+y_1\sqrt d)^n$ for some positive integer $n$ and $x_1+y_1\sqrt d$ is the fundamental solution. As we have $2^{25}-1$ possibilities for $d$, to get a finite number of possible solutions for our equation, we need an upper bound for $n$. Let $d$ be fixed, $\eta=x_1+y_1\sqrt d$ and $\zeta=x_1-y_1\sqrt d$. It can be easily seen that $$\left\{ \begin{matrix} x_n+y_n\sqrt d & = & (x_1+y_1\sqrt d)^n=\eta^n,\\ x_n-y_n\sqrt d & = & (x_1-y_1\sqrt d)^n=\zeta^n. \end{matrix} \right.$$ From here, we deduce that $$x_n=\frac{\eta^n+\zeta^n}{2}\qquad \text{ and } \qquad y_n=\frac{\eta^n-\zeta^n}{2\sqrt d}.$$ It is easy to see that $y_1$ divides $y_n$. We define $u_n=y_n/y_1$. As $u_n=(\eta^n-\zeta^n)/(\eta-\zeta)$, it follows that $(u_n)_{n\geq 0}$ is a Lucas sequence of the first kind with real roots $\eta$ and $\zeta$. By a result of Carmichael (see [@car]) known as the Primitive Divisor Theorem, it follows that if $n>12$, then $u_n$ has a *primitive divisor* $p$. This primitive divisor has several particular properties, the most important one for us being that it satisfies the congruence $p\equiv\pm 1 \pmod n$. This implies that for $n>98$, there exists a prime $p\geq 101$ dividing $u_n$. Thus, $n\leq 98$. The algorithm {#sec:3} ============= We now give our algorithm. Let $$S=\{2,3,5,7,11,13,17,19,23,29,31,37,\ldots,97\},$$ be the set of all primes $p\le 100$. The set $S$ has $25$ elements. Let $R_d$ be the regulator of the ring of integers of the quadratic field $\mathbb Q(\sqrt d)$ and let $x_1+y_1 \sqrt d$ be the fundamental solution of the Pell equation $x^2-dy^2=1$. [For all $D\in \mathcal P(S)$\ $\{$\ $1.\ d=\prod_{p\in D} p$\ $2.$ Compute $mR_d$\ $3.$ Compute a compact representation of $x_m+y_m\sqrt d$\ $4.$ For $i=1$ to $25$ compute ${\rm ord}_{p_i}y_m$\ $5.\ z=2^{{\rm ord}_{p_1}y_m}\cdot \ldots 97^{{\rm ord}_{p_{25}}y_m}$\ $6.$ Compute all the convergents ${p_n}/{q_n}$ of the continued fraction expansion of $\sqrt d$ having $q_n<z$, and check whether $p_n^2-dq_n^2=1$\ $7.$ If $R_d-\log 2-\log \sqrt d \approx \log z$\ $\{$\ $m=1$ and $x_1$ is a solution\ $8.$ For $i=2$ to $98$\ $\{$\ For $j=1$ to $25$ compute ${\rm ord}_{p_j}y_i$\ $z=2^{{\rm ord}_{p_1}y_i}\cdot \ldots 97^{{\rm ord}_{p_{25}}y_i}$\ If $i\cdot R_d-\log 2-\log \sqrt d\approx \log z$, $x_i$ is a solution\ $\}$\ $\}$\ $\}$\ ]{}\ The algorithm searches through all $2^{25}-1$ possible $d$. In step $1$, a value for $d$ is chosen. In step 2, Buchmann’s subexponential algorithm is used to compute $R_d$. This algorithm returns a multiple $mR_d$ of the regulator $R_d$, unconditionally . If the Generalized Riemann Hypothesis is true, then $m=1$. We will remove the dependence of our algorithm on the Generalized Riemann Hypothesis in step 6. A *compact representation* of an algebraic number $\beta \in \mathbb Q (\sqrt d)$ is a representation of $\beta$ of the form $$\beta=\prod_{j=1}^k \left(\frac{\alpha_j}{d_j}\right)^{2^{k-j}}, \label{cr}$$ where $d_j\in \mathbb Z,\ \alpha_j={(a_j+b_j\sqrt d)}/2 \in \mathbb Q(\sqrt d),\ a_j,b_j\in \mathbb Z,\ j=1,\ldots, k$, and $k$, $\alpha$ and $d_j$ have $O(\log \sqrt d)$ digits. In step 3, a compact representation of $x_m+y_m\sqrt d$ is constructed using the algorithms described in [@mm]. The reason for using compact representations is that the standard representation of the fundamental solution of the Pell equation has $O(\sqrt d)$ digits. Using the standard representation would make these computations impossible. More details about compact representations can be found in [@jw]. Since we only have the compact representation of the $x_m+y_m\sqrt d$, in step 4, we use algorithms from [@fn] to perform modular arithmetic on the compact representation. The $p$-adic valuations of $y_m$ when $p$ is one of the first 25 primes are computed in the following way. We define $v=\prod_{i=1}^{25}p_i$. We first compute $y_m \pmod v$. If $y_m\equiv 0 \pmod v$, we conclude that $y_m$ is divisible by all 25 primes. Next, for all primes that we now know divide $y_m$, we compute $y_m\pmod{p_i^{c}}$, where $c$ is a sufficiently large constant. We first take $c=15$ and if $y_m\equiv 0\pmod{p_i^{c}}$ still, we then replace $c$ by $2c$ and repeat the computation. In step 5, $z$, which is defined to be the part of $y_m$ divisible by the first 25 primes, is computed. The purpose of step 6 is to remove the dependence of this algorithm on the Generalized Riemann Hypothesis. If the Generalized Riemann Hypothesis is false, it is possible that, without this check, we could miss some solutions in our search. Suppose therefore that $m>1$ and that $y_k$ is such a solution that we are missing, meaning that $y_k$ is divisible only by some of the first 25 primes for some $k$ that is not a multiple of $m$. But $y_1\mid y_k$, so $y_1$ is divisible only by some of the first 25 primes. Also $y_1\mid y_m$, meaning that $y_1$ divides the part of $y_m$ that is divisible by the first 25 primes, which in our notation is $z$. In other words, $y_1 \mid z$, implying $y_1<z$. As $y_1$ has to be the denominator of a convergent of the continued fraction expansion of $\sqrt d$, it follows that by checking that the relation $p_n^2-dq_n^2\neq 1$ holds for all $n$ satisfying $q_n<z$, we arrive at a contradiction. This proves that either $m=1$, or that $y_k$ has a prime factor larger than 100 for all positive integers $k$. This implies that our algorithm finds all solutions to equation unconditionally, and if $y_m=z$, then $m=1$. In practice, $z$ will be a relatively small number, so usually only about $10$ convergents will need to be computed. Since $x_1\approx y_1 \sqrt d$, it follows that $$R_d=\log(x_1+y_1\sqrt d)\approx \log(2y_1\sqrt d)=\log 2+\log \sqrt d +\log y_1,$$ so we can determine whether $y_1=z$ by the test in step 7. In this test, we took that $a\approx b$ if $|a-b|<0.5$. If $y_1\neq z$, then $y_1\geq 101\cdot z$, so $\log y_1>4.61 +\log z$. This shows that great numerical precision is not needed here. Just in case, we used 10 digits of precision. Step 8 checks whether any of the $x_n$ are solutions for $n=2$ to $98$. With the purpose of speeding up the algorithm, the fundamental solution was not powered when computing ${\rm ord}_{p_j}y_i$. Instead, we computed the fundamental solution modulo the appropriate integer and then powering modulo that integer. Also, usually not all $n$ have to be checked. This is a consequence of the fact that if a prime $p$ divides $y_k$, then it divides $y_{lk}$ for every positive integer $l$. For example, if we get that $y_2$ is divisible by a prime larger than 100, then $y_4$, $y_6, \ldots$ do not need to be checked. The running time of the algorithm is dominated by the computation of the regulator in step 2. Step 2 makes one the algorithm run in subexponential time. It is the only part of the algorithm that is not polynomial. The computations were performed on a Intel Xeon E5430. The computation of the regulators took around 12 days of CPU time, while the rest of the computations took around 3 days of CPU time. Suppose that one wants to find the solutions of $P(x^2-1)<K$ using this algorithm. By the Prime Number Theorem, there are approximately $K/ \log K$ primes up to $K$. This means that the algorithm will loop $2^{K \log K}$ times. The product of all these primes will be of size $O(e^K)$. This means that step 2 will run in $O(e^{\sqrt{K\log K}})$. Thus, the time complexity of the algorithm is $O({\rm exp}(K/\log K +\sqrt{K\log K}))$. Results {#sec:4} ======= - The largest three solutions $x$ of the equation $P(x^2-1)<100$ are $$x=\left\{ \begin{matrix} 19182937474703818751,\\ 332110803172167361,\\ 99913980938200001. \end{matrix} \right.$$ - The largest solution $x$ of $P(x^4-1)<100$ is $x=4217$. - The largest solution $x$ of $P(x^6-1)<100$ is $x=68$. - The largest $n$ such that $P(x^{2n}-1)<100$ has an integer solution $x>1$ is $n=10$, the solution being $x=2$. - The largest $n$ such that $P(x^{2n}-1)<100$ has an integer solution $x>2$ is $n=6$, the solution being $x=6$. - The equation $P(x^2-1)<100$ has $16167$ solutions. - The greatest power $n$ of the fundamental solution of the Pell equation $(x_1+y_1\sqrt d)^n$ which leads to a solution of our problem is $(2+\sqrt 3)^{18}$; i.e., $n=18$ for $d=3$. The case $d=3$ also gives the most solutions, namely $10$ of them. *Proof:*\ The proof is achieved via a computer search using the algorithm from Section \[sec:3\]. Part b) is proved by finding the largest square of all the $x$, c) by finding the largest a cube, etc.\ The largest solution $x$ has $20$ decimal digits, the second largest has $18$ digits, followed by $5$ solutions with $17$ digits and $10$ solutions with $16$ digits. All of the mentioned large solutions are odd. The largest even solution $x$ has $15$ digits. \[tm\][Theorem]{} Write the equation as $$x^2-1=2^{a_1}\cdots97^{a_{25}}.$$ Then the following hold: - The solution with the largest number of $a_i\neq 0$ is $x=9747977591754401$. For this solution, $17$ of the $a_i$s are non-zero. - The solution with the largest $\sum_{i=1}^{25}{a_i}$ is $x= 19182937474703818751$. For this solution, $\sum_{i=1}^{25}{a_i}=47$. - The single largest $a_i$ appearing among all solutions is $a_1=27$ and corresponds to the solution $x=4198129205249$. *Proof:*\ Again, this is done via a computer search using the algorithm from Section \[sec:3\].\ Dabrowski [@dab] considered a similar problem, where the prime factors of $x^2-1$ consist of the first $k$ primes $p_1,\ldots, p_k$. He formulated the following conjecture. \[tm\][Conjecture]{} The Diophantine equation $$x^2-1=p_1^{\alpha_1}\cdots p_k^{\alpha_k}$$ has exactly $28$ solutions $(x;\alpha_1,\ldots,\alpha_k)$ in positive integers, as follows: - $(3;3)$, - $(5;3,1),(7;4,1),(17;5,2)$, - $(11; 3, 1, 1), (19; 3, 2, 1), (31; 6, 1, 1), (49; 5, 1, 2), (161; 6, 4, 1),$ - $(29; 3, 1, 1, 1), (41; 4, 1, 1, 1), (71; 4, 2, 1, 1), (251; 3, 2, 3, 1), (449; 7, 2, 2, 1),$ $(4801; 7, 1, 2, 4),(8749; 3, 7, 4, 1),$ - $(769; 9, 1, 1, 1, 1), (881; 5, 2, 1, 2, 1), (1079; 4, 3, 1, 2, 1), (6049; 6, 3, 2, 1, 2),$ $(19601; 5, 4, 2, 2, 2),$ - $(3431; 4, 1, 1, 3, 1, 1), (4159; 7, 3, 1, 1, 1, 1), (246401; 8, 6, 2, 1, 1, 2),$ - $(1429; 3, 1, 1, 1, 1, 1, 1), (24751; 5, 2, 3, 1, 1, 1, 1), (388961; 6, 4, 1, 4, 1, 1, 1),$ - $(1267111; 4, 3, 1, 1, 3, 1, 1, 2).$ \[dabslut\] The main result of [@dab] is that Conjecture \[dabslut\] is true for $k\leq 5$. From our data, we confirm Dabrowski’s conjecture in a much wider range. \[tm\][Theorem]{} Conjecture \[dabslut\] is true for $k\leq 25$. *Proof:*\ This is done by simply factoring all $x^2-1$, where $x$ is a solution of .\ The next theorem follows also trivially from our results. Recall that a positive integer $n$ is [*$K$-smooth*]{} if $P(n)\le K$. In particular, the main result of our paper is the determination of all the $100$-smooth positive integers of the form $x^2-1$. \[tm\][Corollary]{} Let $t$ be the largest odd solution, $t=19182937474703818751$ and $s$ be the largest even solution, $s=473599589105798$ of equation . - The largest consecutive $100$-smooth integers are $x$ and $x+1$ where $x=(t-1)/2$ is the largest solution of $P(x(x+1))<100.$ - The largest consecutive even $100$-smooth integers are $t-1$ and $t+1$. - The largest consecutive odd $100$-smooth integers are $s-1$ and $s+1$. - The largest triangular $100$-smooth integer is $(t^2-1)/8$. As we mentioned in the Introduction, the problem of finding two consecutive $K$-smooth integers was examined by Lehmer in [@leh] in the sixties. At that time, he was able to solve the above problem for the values $K\le 41$. The advance of both computing power and theoretical arguments (namely, the compact represenations) allow us to solve the much harder problem of finding consecutive $K$-smooth integers for any $K\leq 100$. Note that, as was already mentioned at the end of Section \[sec:3\], the difficulty of this problem grows exponentially with $K$. Our results can also be applied to finding $k$ consecutive $K$-smooth integers, for any integer $k$. We obtain the following results. \[tm\][Corollary]{} The largest integer $x$ satisfying $$P(x(x+1)\ldots(x+n))<100,$$ for a given $n$, are given in the following table: $n$ $x$ ----- ----------------------- $1$ $9591468737351909375$ $2$ $407498958$ $3$ $97524$ $4$ $7565$ $5$ $7564$ $6$ $4896$ $7$ $4895$ $8$ $284$ *Proof:*\ To find $k$ consecutive $100$ smooth integers, we first create a list of all pairs of consecutive $100$-smooth integers. Every odd solution of our starting problem will give us one such pair. This is because if $x$ is an odd solution, then $(x-1)/2$ and $(x+1)/2$ are consecutive $100$-smooth integers. Once this list is created, we search for overlaps in these pairs and obtain our results.\ In the recent paper [@ST], Shorey and Tijdeman proved several extensions of some irreducibility theorems due to Schur. The main results of [@ST] rely heavily on Lehmer’s results from [@leh]. Thus, replacing, for example Lemma 2.1 in [@ST] by our results, it is likely that the main results from [@ST] can be extended in a wider range. **Remark.** The tables produced by our computations can be found on the webpage [http://web.math.hr/$\sim$fnajman]{}. [1]{} Yu. Bilu, G. Hanrot, P. M.  Voutier, ‘Existence of primitive divisors of Lucas and Lehmer numbers. With an appendix by M. Mignotte’, [*J. Reine Angew. Math.*]{} [**539**]{} (2001), 75–122. J.  Buchmann, *A subexponential algorithm for the determination of class groups and regulators of algebraic number fields*, Seminaire de Theorie des Nombres (1990), 27-41. R.  D.  Carmichael, *On the numerical factors of arithmetic forms $\alpha^n \pm \beta^n$*, Ann. of. Math. **15** (1913), 30–70. A.  Dabrowski, *On the Brocard-Ramanujan problem and generalizations*, Preprint, 2009. M.  J.  Jacobson Jr., H.  C . Williams, *Solving the Pell Equation*, Springer, 2009. D.  H.  Lehmer, *On a problem of Störmer*, Illinois J. Math **8** (1964), 57–79. F. Luca, ‘Primitive divisors of Lucas sequences and prime factors of $x\sp 2+1$ and $x\sp 4+1$’, [*Acta Acad. Paedagog. Agriensis Sect. Mat. (N.S.)* ]{} [**31**]{} (2004), 19–24. M.  Maurer, *Regulator approximation and fundamental unit computation for real quadratic orders*, PhD thesis, Technische Universit¨at Darmstadt, Fachbereich Informatik, Darmstadt, Germany, 2000. F.  Najman, *Compact representation of quadratic integers and integer points on some elliptic curves*, Rocky Mountain J. Math., to appear. T. N.  Shorey and R. Tijdeman, *Generalizations of some irreducibility results by Schur*, Preprint, 2009. \ \ *E-mail address:* [email protected]\ \ \ *E-mail address:* [email protected]
{ "pile_set_name": "ArXiv" }
--- abstract: 'The goal of this paper is twofold; first, show the equivalence between certain problems in geometry, such as view-obstruction and billiard ball motions, with the estimation of covering radii of lattice zonotopes. Second, we will estimate upper bounds of said radii by virtue of the Flatness Theorem. These problems are similar in nature with the famous lonely runner conjecture.' address: - 'Institut für Mathematik, Freie Universität Berlin, Arnimallee 2, 14195 Berlin, Germany' - 'Technische Universität Berlin, Institut für Mathematik, Sekretariat MA 4-1, Stra[ß]{}e des 17. Juni 136, D-10623 Berlin, Germany' author: - Matthias Henze - 'Romanos-Diogenes Malikiosis' bibliography: - 'jointbib.bib' title: 'On the covering radius of lattice zonotopes and its relation to view-obstructions and the lonely runner conjecture' --- [^1] [^2] Introduction ============ The purpose of this article is to exhibit and utilize the equivalence of certain geometric problems in different settings: [**(a)**]{} billiard ball motions inside a cube avoiding an inner cube, [**(b)**]{} lines in a multidimensional torus avoiding a smaller “copy” of the torus, [**(c)**]{} views unobstructed by a lattice arrangement of cubes, and [**(d)**]{} covering radii of lattice zonotopes. The equivalence of the first three problems has been shown in the works of Wills [@willslrc], Cusick [@cusickviewob], and Schoenberg [@Sch76], among others; the equivalence to estimating covering radii of zonotopes is the novelty here. The latter interpretation gives the possibility to use techniques from discrete geometry and the geometry of numbers in order to tackle these problems. In a nutshell the four settings are related as follows: A billiard ball motion inside the cube $[0,1]^m$ can be unfolded into a line in the torus ${\ensuremath{\mathbb{T}}}^m=[0,1)^m$, by reflecting appropriately its pieces between the boundary of the cube. Then, through periodization of this configuration we obtain a lattice arrangement of lines in the space $\R^m$. So, if the billiard ball motion intersects an inner cube, say $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m$, then the corresponding line in ${\ensuremath{\mathbb{T}}}^m$ will intersect a smaller “copy” of the torus, and the line in $\R^m$ will intersect a lattice arrangement of cubes. An equivalent condition to the latter case is having a cube intersecting a lattice arrangement of lines in $\R^m$. Then, under an appropriate projection, we get a zonotope intersecting a lattice. In order to get bounds on the size of the cubes under question, we will chiefly work in the zonotope setting. The other three settings and their equivalences have been investigated in some detail before. The keen reader might recognize that when the line in question passes through the origin we essentially deal with the *lonely runner problem*. In order to make all this precise, some notation is in order. For standard notions in convex geometry and the theory of lattices, we refer the reader to the textbooks of Gruber [@gruber2007convex] and Martinet [@martinet2003perfect], respectively. A billiard ball motion inside the unit cube is denoted by $\operatorname{bbm}({\ensuremath{\boldsymbol}}u_0,{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})$, where ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ shall denote its initial direction and ${\ensuremath{\boldsymbol}}u_0$ its starting point. There is a linear subspace $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ of $\R^m$ that uniquely corresponds to every such ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$ (see Subsection \[subsect\_zonotopes\_vo\]). The orthogonal projection $C_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ of the unit cube $C_m=[0,1]^m$ onto $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ is a zonotope with vertices in ${\ensuremath{\mathbb{Z}}}^m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. Next, take an invertible linear map $T:V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\to\R^n$, for which $T({\ensuremath{\mathbb{Z}}}^m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})={\ensuremath{\mathbb{Z}}}^n$, and denote the zonotope $T(C_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})$ by $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. We further let ${\ensuremath{\mathbf{1}}}_m=(1,\ldots,1)^\intercal$ be the all-one-vector in $\R^m$. The discussed equivalences can now be summarized as follows. \[equivform\] Let ${\ensuremath{\boldsymbol}}{u}_0\in\R^m$, ${\ensuremath{\boldsymbol}}{\alpha}\in\R^m$, $0\leq{\ensuremath{\varepsilon}}\leq1/2$, and $n=\dim(V_{{\ensuremath{\boldsymbol}}{\alpha}})$. The following statements are equivalent. 1. The $\operatorname{bbm}({\ensuremath{\boldsymbol}}{u}_0,{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}})$ in $[0,1]^m$ intersects $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m$. 2. The line $\left\{{\ensuremath{\boldsymbol}}{u}_0+t{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}} \mid t\in{\ensuremath{\mathbb{R}}}\right\}/\,\Z^m$ in ${\ensuremath{\mathbb{T}}}^m$ intersects $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m$. 3. The view from ${\ensuremath{\boldsymbol}}{u}_0$ with direction ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ is obstructed by $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m+{\ensuremath{\mathbb{Z}}}^m$. 4. $\left((1-2{\ensuremath{\varepsilon}})Z_{{\ensuremath{\boldsymbol}}{\alpha}}-\bar{{\ensuremath{\boldsymbol}}{u}}_0\right)\cap\Z^n\neq{\ensuremath{\varnothing}}$, where $\bar{{\ensuremath{\boldsymbol}}{u}}_0=T({\ensuremath{\boldsymbol}}{u}_0-{\ensuremath{\varepsilon}}{\ensuremath{\mathbf{1}}}_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})$. It should be noted that property [**(V4)**]{} does not depend on the choice of the map $T$, as long as it satisfies $T({\ensuremath{\mathbb{Z}}}^m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})={\ensuremath{\mathbb{Z}}}^n$. Furthermore, when [**(V4)**]{} fails, we have a zonotope avoiding a lattice, and through Khinchin’s *Flatness Theorem* [@khinchin1948a], we obtain an estimate on the largest possible ${\ensuremath{\varepsilon}}$ that one can choose in Theorem \[equivform\] under natural constraints on ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$. Since our methods are efficient in the last setting, we supply the text with the relevant tools and definitions. In Section \[sect\_rationally\_uniform\] it will be apparent why we further restrict the direction vectors ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$, requiring that they may be *rationally uniform*: Let ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in{\ensuremath{\mathbb{R}}}^m$ and write $\dim_\Q({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})$ for the dimension of the $\Q$-vector space generated by the coordinates of ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$. Then, ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ is called [*rationally uniform*]{} if every $\dim_\Q({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})$ coordinates of ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ are linearly independent over $\Q$. If ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ is rationally uniform, then $\operatorname{bbm}({\ensuremath{\boldsymbol}}{u}_0,{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})$ is also called such. For ${\ensuremath{\boldsymbol}}u,{\ensuremath{\boldsymbol}}v\in\R^n$, let $[{\ensuremath{\boldsymbol}}{u},{\ensuremath{\boldsymbol}}{v}]$ denote the line segment with endpoints ${\ensuremath{\boldsymbol}}{u}$ and ${\ensuremath{\boldsymbol}}{v}$. A *zonotope* $Z{\ensuremath{\subseteq}}\R^n$ is the Minkowski sum of line segments $[{\ensuremath{\boldsymbol}}{a}_i,{\ensuremath{\boldsymbol}}{b}_i]$, $1\leq i\leq m$, where the vectors ${\ensuremath{\boldsymbol}}{b}_i-{\ensuremath{\boldsymbol}}{a}_i$ generate the whole space $\R^n$. Here, we shall only consider lattice zonotopes, that is, zonotopes whose vertices lie on a lattice ${\ensuremath{\Lambda}}$. So, we consider $Z=\sum_{i=1}^{m}[{\ensuremath{\boldsymbol}}{0},{\ensuremath{\boldsymbol}}{z}_i]$, where ${\ensuremath{\boldsymbol}}{z}_1,\dotsc,{\ensuremath{\boldsymbol}}{z}_m\in{\ensuremath{\Lambda}}$ generate $\R^n$. Also, we may require that every $n$ vectors from the generators of the zonotope form a basis of $\R^n$; we prove that these are precisely the zonotopes $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ that are obtained from rationally uniform ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$. Let $S\subset\R^n$ be a finite subset. We say that $S$ is in *linear general position* ([LGP]{}), if any $n$ points in $S$ are linearly independent. Estimating the *covering radius* of a zonotope will provide us with bounds for the largest possible ${\ensuremath{\varepsilon}}$ described above. This quantity is defined generally for convex bodies as follows. Let $K{\ensuremath{\subseteq}}\R^n$ be a convex body, and let ${\ensuremath{\Lambda}}$ be a lattice. Then, the *covering radius* of $K$ with respect to ${\ensuremath{\Lambda}}$, denoted by $\mu(K,{\ensuremath{\Lambda}})$, is the smallest positive real number $\rho$ for which the translates of $\rho K$ by ${\ensuremath{\Lambda}}$ cover the entire space $\R^n$, that is, $$\mu(K,{\ensuremath{\Lambda}})=\inf\{\rho>0\mid\rho K+{\ensuremath{\Lambda}}=\R^n\}.$$ For $\Lambda=\Z^n$, we abbreviate $\mu(Z)=\mu(Z,\Z^n)$. ![Illustrating [**(V3)**]{}, [**(V1)**]{}, and [**(V4)**]{} for the parameters ${\boldsymbol u}_0=(1/6,1/2,5/6)^\intercal$, ${\boldsymbol{\ensuremath{\alpha}}}=(1,2,1)^\intercal$, and ${\ensuremath{\varepsilon}}=1/3$.](vo121.png "fig:"){height="4cm"} ![Illustrating [**(V3)**]{}, [**(V1)**]{}, and [**(V4)**]{} for the parameters ${\boldsymbol u}_0=(1/6,1/2,5/6)^\intercal$, ${\boldsymbol{\ensuremath{\alpha}}}=(1,2,1)^\intercal$, and ${\ensuremath{\varepsilon}}=1/3$.](bbm121.png "fig:"){height="4cm"} ![Illustrating [**(V3)**]{}, [**(V1)**]{}, and [**(V4)**]{} for the parameters ${\boldsymbol u}_0=(1/6,1/2,5/6)^\intercal$, ${\boldsymbol{\ensuremath{\alpha}}}=(1,2,1)^\intercal$, and ${\ensuremath{\varepsilon}}=1/3$.](zonotope121.pdf "fig:"){height="3.5cm"} We can now state our main result with respect to the covering radius of lattice zonotopes. \[thmCoveringRadiusAsymptotics\] Let $S=\{{\ensuremath{\boldsymbol}}{z}_1,\ldots,{\ensuremath{\boldsymbol}}{z}_m\}{\ensuremath{\subseteq}}\Z^n$ and let $Z=\sum_{i=1}^m[{\ensuremath{\boldsymbol}}{0},{\ensuremath{\boldsymbol}}{z}_i]$ be the lattice zonotope generated by $S$. If $S$ is in [LGP]{}, we have $$\mu(Z) \leq \frac{c\,n\log{n}}{m-n+1},$$ for some absolute constant $c>0$. Moreover, there is an $S\subset\Z^n$, $|S|=m$, in [LGP]{} such that $\mu(Z) \geq \frac1{m-n+1}$. The logarithmic factor in the above estimate is most likely not needed and it has its roots in the application of the Flatness Theorem to our problem. It is commonly believed that the *flatness constant* is of order $O(n)$ rather than the currently known bound $O(n\log{n})$ (see the discussion in Section \[sect\_flatness\]). A beautiful result of Schoenberg [@Sch76] regarding billiard ball motions inside the unit cube is the following. \[Sch\] Every nontrivial billiard ball motion inside the unit cube $[0,1]^m$ intersects the cube $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m$ if and only if ${\ensuremath{\varepsilon}}\leq 1/(2m)$. The number of such motions touching the cube for ${\ensuremath{\varepsilon}}=1/(2m)$ is essentially finite. A nontrivial billiard ball motion is one that is not contained in a translate of a coordinate hyperplane. As we shall see below, these motions correspond exactly to initial directions ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ every coordinate of which is nonzero. An example of a billiard ball motion that intersects the cube $[\frac1{2m},1-\frac1{2m}]^m$ only in the boundary is given by the parameters $${\ensuremath{\boldsymbol}}u_0={{\left({0,\frac{1}{m},\ldots,\frac{m-1}{m}}\right)}}^\intercal\quad\text{and}\quad{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}={\ensuremath{\mathbf{1}}}_m=(1,\ldots,1)^\intercal.$$ We later reformulate Schoenberg’s result to the statement that the inequality $\mu(Z)\leq n/(n+1)$ holds for every lattice zonotope $Z\subseteq\R^n$ that is generated by $n+1$ vectors in [LGP]{}. This interpretation motivates the investigation of billiard ball motions whose initial directions have restricted rational dependencies. Based on the equivalence of [**(V1)**]{}-[**(V4)**]{}, we define, for any $n\leq m$, $$\begin{aligned} {\ensuremath{\varepsilon}}(n,m)=\sup\big\{{\ensuremath{\varepsilon}}\geq0 \mid &\,{\bf (Vi)}\text{ hold for any }{\ensuremath{\boldsymbol}}u_0\in\R^m\text{ and any rationally}\\ &\text{ uniform }{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in(\R{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m\text{ with}\,\dim_{\Q}({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})\geq m-n\big\}.\end{aligned}$$ The trivial cases are ${\ensuremath{\varepsilon}}(0,m)=1/2$ and ${\ensuremath{\varepsilon}}(m,m)=0$, whereas Theorem \[Sch\] translates into ${\ensuremath{\varepsilon}}(m-1,m)=1/(2m)$. We show in Section \[subsect\_rationally\_uniform\] that ${\ensuremath{\varepsilon}}(n,m)=\frac12\left(1-\sup \mu(Z)\right)$, where the supremum is taken over all lattice zonotopes $Z\subseteq\R^k$ generated by $m$ vectors of $\Z^k$ in [LGP]{}, and where $k\leq n$. Therefore, Theorem \[thmCoveringRadiusAsymptotics\] translates into the bounds $$\begin{aligned} \frac{m-O(n\log{n})}{2(m-n+1)}\leq{\ensuremath{\varepsilon}}(n,m)\leq\frac{m-n}{2(m-n+1)}.\label{eqnEpsAsympBounds}\end{aligned}$$ The fact that the logarithmic factor in the lower estimate of ${\ensuremath{\varepsilon}}(n,m)$ might not be needed, as well as the implied constant in the case covered by Theorem \[Sch\] leads us to formulate: \[mainconj\] Let $S={{\left\{{{\ensuremath{\boldsymbol}}{z}_1,\dotsc,{\ensuremath{\boldsymbol}}{z}_m}\right\}}}{\ensuremath{\subseteq}}\Z^n$ and let $Z$ be the lattice zonotope generated by $S$. If $S$ is in [LGP]{}, then $\mu(Z)\leq \frac{n}{m}$. This is equivalent to ${\ensuremath{\varepsilon}}(n,m)\geq\frac{m-n}{2m}$. By virtue of Theorem \[equivform\], for ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$ with $\dim_{\Q}({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})\geq m-n$, Conjecture \[mainconj\] has the following equivalent forms: - Every rationally uniform $\operatorname{bbm}({\ensuremath{\boldsymbol}}{u}_0,{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})$ intersects ${{\left[{\frac{m-n}{2m},\frac{m+n}{2m}}\right]}}^m$. - For every ${\ensuremath{\boldsymbol}}{u}_0\in\R^m$ and every rational uniform ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$, the line ${{\left\{{{\ensuremath{\boldsymbol}}{u}_0+t{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\mid t\in \R}\right\}}}/\,\Z^m$ in ${\ensuremath{\mathbb{T}}}^m$ intersects ${{\left[{\frac{m-n}{2m},\frac{m+n}{2m}}\right]}}^m$. - From every point ${\ensuremath{\boldsymbol}}u_0\in\R^m$, the view with rationally uniform direction ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ is obstructed by ${{\left[{\frac{m-n}{2m},\frac{m+n}{2m}}\right]}}^m+\Z^m$. The paper is organized as follows. In the next section, we familiarize the reader with Schoenberg’s terminology concerning billiard ball motions, albeit in a more general setting, and we rephrase Theorem \[Sch\] in the context of lines in a multidimensional torus. Finally we pass to lattice zonotopes through orthogonal projections, thus establishing the equivalence of the properties [**(V1)**]{}–[**(V4)**]{} in Theorem \[equivform\]. In Section \[sect\_rationally\_uniform\], we restrict our attention to rationally uniform vectors, justifying the definition of ${\ensuremath{\varepsilon}}(n,m)$ above. We show that these vectors are associated with lattice zonotopes generated by vectors in LGP and vice versa. We briefly discuss the *associated zonotope* of $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, originally studied by Shephard [@shephard1974combinatorial], and establish the monotonicity properties of ${\ensuremath{\varepsilon}}(n,m)$. As the ${\ensuremath{\varepsilon}}(n,m)$ can be expressed in terms of covering radii of lattice zonotopes generated by vectors in [LGP]{}, we apply Khinchin’s Flatness Theorem in order to produce nontrivial bounds; Theorem \[thmCoveringRadiusAsymptotics\] is thus proved in Section \[sect\_flatness\]. In the last section, we draw some connections to the Lonely Runner Problem, where we state with Corollary \[corEquivLRP\] the analogous statement to Theorem \[equivform\] in this setting. We also rephrase this problem in terms of zonotopes centered at a point of order $2$ modulo ${\ensuremath{\mathbb{Z}}}^n$ having a lattice point, with the hope that it could be useful towards establishing better bounds than the existing ones (Conjecture \[conjZonotopalLRC\]). Aside from that, we prove that it suffices to consider integer velocities in the Lonely Runner Conjecture. This reduction was originally stated by Wills [@willslrc], thereafter taken for granted, until a proof appeared in [@sixrunners], which however depends on solving the Lonely Runner Conjecture in lower dimensions. In Lemma \[LRCintegerReduction\] we prove the reduction to integer velocities unconditionally. Finally, we show that the Lonely Runner Conjecture implies a more refined statement, namely Conjecture \[strongLRC\], where we take the dimension of the $\Q$-span of the velocities into account. Billiard ball motions, multidimensional tori, and zonotopes {#sect_theorem_equivs} =========================================================== Billiard ball motions {#subsectBBM} --------------------- In [@Sch76], Schoenberg defines a billiard ball motion inside a cube as rectilinear and uniform and it is reflected in the usual way when striking any of the cube’s faces. Since the boundary of a cube is not smooth, a little care should be taken when this motion hits the boundary of the cube in lower-dimensional faces. Let ${\ensuremath{\boldsymbol}}{u}_0\in[0,1]^m$ and let ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}=({\ensuremath{\alpha}}_1,\dotsc,{\ensuremath{\alpha}}_m)\in\R^m$. Initially, the billiard ball motion starting at ${\ensuremath{\boldsymbol}}{u}_0$ with inital direction ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ has the form ${\ensuremath{\boldsymbol}}{u}_0+t{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$, $t\geq0$. Assume that $t_0$ is the first instance when this motion hits the boundary of the standard cube $[0,1]^m$, say, at the relative interior of the facet $x_j={\ensuremath{\varepsilon}}_j$ for $j\in J{\ensuremath{\subseteq}}{{\left\{{1,2,\dotsc,m}\right\}}}$, where ${\ensuremath{\varepsilon}}_j=0$ or $1$. Then, the motion is reflected and follows the path ${\ensuremath{\boldsymbol}}{u}_0+t_0{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}+(t-t_0){\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'$ for $t>t_0$ (until it hits the boundary again), where ${\ensuremath{\alpha}}'_j=-{\ensuremath{\alpha}}_j$ for $j\in J$ and ${\ensuremath{\alpha}}'_j={\ensuremath{\alpha}}_j$ otherwise. It is clear that such a motion is completely determined by the starting point ${\ensuremath{\boldsymbol}}{u}_0$ and the initial direction ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$. A *billiard ball motion* starting at ${\ensuremath{\boldsymbol}}{u}_0$ and with initial direction ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$, that is reflected naturally as described above, is denoted by $\operatorname{bbm}({\ensuremath{\boldsymbol}}{u}_0,{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}})$. We call such a motion *nontrivial* when all the coordinates of ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ are nonzero, and *rationally uniform*, when ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ is. If ${\ensuremath{\boldsymbol}}{v}_0$ is the symmetric point of ${\ensuremath{\boldsymbol}}{u}_0+t_0{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ with respect to ${\ensuremath{\frac{1}{2}{\ensuremath{\mathbf{1}}}_{m}}}$, then it is clear from the above that the path ${\ensuremath{\boldsymbol}}{v}_0-(t-t_0){\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'$ is symmetric to the path ${\ensuremath{\boldsymbol}}{u}_0+t_0{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}+(t-t_0){\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'$ with respect to ${\ensuremath{\frac{1}{2}{\ensuremath{\mathbf{1}}}_{m}}}$. Hence, for every ${\ensuremath{\varepsilon}}\geq0$, the $\operatorname{bbm}({\ensuremath{\boldsymbol}}{u}_0,{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}})$ avoids the cube $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m$, if and only if the line ${\ensuremath{\boldsymbol}}{u}_0+t{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$, $t\in\R$, in the torus ${\ensuremath{\mathbb{T}}}^m=\R^m/\Z^m$ avoids the same cube. In the latter case, the coordinates are taken $\bmod\,1$. Furthermore, the latter case happens if and only if the line ${{\left\{{{\ensuremath{\boldsymbol}}{u}_0+t{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}} \mid t\in\R}\right\}}}$ avoids the set $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m+{\ensuremath{\mathbb{Z}}}^m$. This however can be stated as a view-obstruction property, namely that the view from ${\ensuremath{\boldsymbol}}{u}_0$ with direction ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ is not obstructed by $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m+{\ensuremath{\mathbb{Z}}}^m$. In summary, these considerations prove the equivalence of the statements [**(V1)**]{}, [**(V2)**]{}, and [**(V3)**]{} in Theorem \[equivform\]. By Theorem \[Sch\], when we restrict ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ to have no coordinate equal to zero, the infimum of such ${\ensuremath{\varepsilon}}$ is equal to $1/(2m)$. An equivalent statement is that the $l^{\infty}$-distance from ${\ensuremath{\frac{1}{2}{\ensuremath{\mathbf{1}}}_{m}}}$ to any line in ${\ensuremath{\mathbb{T}}}^m$ not parallel to a coordinate hyperplane is at most $\frac{m-1}{2m}$. Note that ${\ensuremath{\mathbb{T}}}^m$ inherits the $l^{\infty}$-distance from $\R^m$ as a quotient space. Lines in a multidimensional torus {#multitorus} --------------------------------- Now that the connection between billiard ball motions and lines in a torus has been established, we want to determine the shape of a line in a multidimensional torus, or equivalently, the shape of its periodization into the whole space $\R^m$ by the standard lattice ${\ensuremath{\mathbb{Z}}}^m$. So, for ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}\in\R^m$, we want to describe ${\ensuremath{\mathbb{Z}}}^m+\R{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$. Ideas similar to those discussed below have been briefly elaborated on in [@sixrunners Lem. 8], and to a greater extent by the second author in [@CyclotomicPyjama] with respect to a different problem. We define the lattice $${\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}:=\left\{{\ensuremath{\boldsymbol}}{l}\in{\ensuremath{\mathbb{Z}}}^m\mid{\ensuremath{\boldsymbol}}{l}\cdot{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}=0\right\}$$ and the set $$\mathcal{E}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}:=\left\{{\ensuremath{\boldsymbol}}{\xi}\in\R^m\mid{\ensuremath{\boldsymbol}}{l}\cdot{\ensuremath{\boldsymbol}}{\xi}\in{\ensuremath{\mathbb{Z}}},\forall{\ensuremath{\boldsymbol}}{l}\in{\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}\right\}.$$ For a first description of $\Z^m+\R{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$, we need a special case of Kronecker’s approximation theorem. \[thmKronecker\] Let ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}},{\ensuremath{\boldsymbol}}{\ensuremath{\beta}}\in{\ensuremath{\mathbb{R}}}^k$. For every ${\ensuremath{\varepsilon}}>0$ there is an integer $q\in\Z$ and a vector ${\ensuremath{\boldsymbol}}p\in{\ensuremath{\mathbb{Z}}}^k$ such that $${{\left\|{q{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}-{\ensuremath{\boldsymbol}}p-{\ensuremath{\boldsymbol}}{\ensuremath{\beta}}}\right\|}}_{\infty}<{\ensuremath{\varepsilon}},$$ if and only if for every ${\ensuremath{\boldsymbol}}r\in{\ensuremath{\mathbb{Z}}}^k$ with ${\ensuremath{\boldsymbol}}r\cdot{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in{\ensuremath{\mathbb{Z}}}$ we also have ${\ensuremath{\boldsymbol}}r\cdot{\ensuremath{\boldsymbol}}{\ensuremath{\beta}}\in{\ensuremath{\mathbb{Z}}}$. \[mainlemma\] Let ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}\in\R^m$. Then, $\mathcal{E}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}$ is the closure of ${\ensuremath{\mathbb{Z}}}^m+\R{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$. Inclusion is obvious: let ${\ensuremath{\boldsymbol}}{\xi}\in{\ensuremath{\mathbb{Z}}}^m+\R{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ be arbitrary. So, ${\ensuremath{\boldsymbol}}{\xi}={\ensuremath{\boldsymbol}}{\nu}+x{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ for some ${\ensuremath{\boldsymbol}}{\nu}\in{\ensuremath{\mathbb{Z}}}^m$ and $x\in\R$. Now, for arbitrary ${\ensuremath{\boldsymbol}}{l}\in{\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}$, we have $${\ensuremath{\boldsymbol}}{l}\cdot{\ensuremath{\boldsymbol}}{\xi}={\ensuremath{\boldsymbol}}{l}\cdot{\ensuremath{\boldsymbol}}{\nu}+x({\ensuremath{\boldsymbol}}{l}\cdot{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}})={\ensuremath{\boldsymbol}}{l}\cdot{\ensuremath{\boldsymbol}}{\nu}\in{\ensuremath{\mathbb{Z}}},$$ since, by definition, ${\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}{\ensuremath{\subseteq}}{\ensuremath{\mathbb{Z}}}^m$. Hence, ${\ensuremath{\boldsymbol}}{\xi}\in\mathcal{E}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}$. Now let ${\ensuremath{\boldsymbol}}{x}=(x_1,\dotsc,x_m)\in\mathcal{E}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}$ be arbitrary. If ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}={\ensuremath{\boldsymbol}}{0}$, then obviously $\mathcal{E}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}={\ensuremath{\mathbb{Z}}}^m$ and there is nothing more to prove. Thus, we assume that ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}\neq{\ensuremath{\boldsymbol}}{0}$, and without loss of generality we may assume that ${\ensuremath{\alpha}}_m\neq0$. Moreover, we can assume that ${\ensuremath{\alpha}}_m=1$, since both $\mathcal{E}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}$ and ${\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}$ are invariant under multiplication of ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ by a nonzero number. We show that we can approximate ${\ensuremath{\boldsymbol}}{x}$ by elements of ${\ensuremath{\mathbb{Z}}}^m+\R{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ as close as we want. It suffices to prove that the sequence $$(x_m+s){\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}-{\ensuremath{\boldsymbol}}x=((x_m+s){\ensuremath{\alpha}}_1-x_1,\dotsc,(x_m+s){\ensuremath{\alpha}}_{m-1}-x_{m-1},s)^\intercal,\, s\in{\ensuremath{\mathbb{Z}}}$$ has terms arbitrarily close to ${\ensuremath{\mathbb{Z}}}^m$, or equivalently, the sequence $$(s{\ensuremath{\alpha}}_1,\dotsc,s{\ensuremath{\alpha}}_{m-1})^\intercal,\, s\in{\ensuremath{\mathbb{Z}}}$$ has terms arbitrarily close to $(x_1-x_m{\ensuremath{\alpha}}_1,\dotsc,x_{m-1}-x_m{\ensuremath{\alpha}}_{m-1})^\intercal+{\ensuremath{\mathbb{Z}}}^{m-1}$. Consider the projection $\pi:\R^m\rightarrow\R^{m-1}$ that “forgets” the last coordinate. In view of Kronecker’s Theorem \[thmKronecker\], the closure of the subgroup of ${\ensuremath{\mathbb{T}}}^{m-1}$ generated by $$\pi({\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}})+{\ensuremath{\mathbb{Z}}}^{m-1}=({\ensuremath{\alpha}}_1,\dotsc,{\ensuremath{\alpha}}_{m-1})^\intercal+{\ensuremath{\mathbb{Z}}}^{m-1}$$ is the set of all $(a_1,\dotsc,a_{m-1})^\intercal+{\ensuremath{\mathbb{Z}}}^{m-1}$ for which $l_1a_1+\dotsb+l_{m-1}a_{m-1}\in{\ensuremath{\mathbb{Z}}}$ whenever $l_1,\dotsc,l_{m-1}\in{\ensuremath{\mathbb{Z}}}$ and $l_1{\ensuremath{\alpha}}_1+ \dotsb+l_{m-1}{\ensuremath{\alpha}}_{m-1}\in{\ensuremath{\mathbb{Z}}}$. In other words, it suffices to prove that $$\pi({\ensuremath{\boldsymbol}}{l})\cdot(x_1-x_m{\ensuremath{\alpha}}_1,\dotsc,x_{m-1}-x_m{\ensuremath{\alpha}}_{m-1})^\intercal\in{\ensuremath{\mathbb{Z}}},$$ whenever ${\ensuremath{\boldsymbol}}{l}\in{\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}$. So, when ${\ensuremath{\boldsymbol}}{l}\in{\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}$, by definition of ${\ensuremath{\boldsymbol}}x\in\mathcal{E}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}$ we have ${\ensuremath{\boldsymbol}}{l}\cdot{\ensuremath{\boldsymbol}}{x}\in{\ensuremath{\mathbb{Z}}}$. Hence, ${\ensuremath{\boldsymbol}}{l}\cdot({\ensuremath{\boldsymbol}}{x}-x_m{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}})\in{\ensuremath{\mathbb{Z}}}$. The latter is obviously equal to the desired inner product, completing the proof. When the coordinates of ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ are linearly independent over ${\ensuremath{\mathbb{Q}}}$, then ${\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}=\{{\ensuremath{\boldsymbol}}{0}\}$ and $\mathcal{E}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}=\R^m$. Therefore, by Lemma \[mainlemma\], ${\ensuremath{\mathbb{Z}}}^m+\R{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}$ is dense in $\R^m$, or equivalently, the set $$\left\{\left(\{x{\ensuremath{\alpha}}_1\},\{x{\ensuremath{\alpha}}_2\},\dotsc,\{x{\ensuremath{\alpha}}_m\}\right)\mid x\in\R\right\}$$ is dense in ${\ensuremath{\mathbb{T}}}^m$, where $\{y\}=y-\lfloor y \rfloor$ denotes the fractional part of $y\in\R$. In particular, this means that ${\ensuremath{\varepsilon}}(0,m)=1/2$ as claimed in the introduction. Zonotopes and view-obstructions {#subsect_zonotopes_vo} ------------------------------- We now finish the proof of Theorem \[equivform\] by providing the details of the zonotopal description of the view-obstruction problem under consideration. Our arguments are based on a more illuminating description of ${\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. Define $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ to be the linear span of ${\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, and let ${\ensuremath{\Lambda}}^\star_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}=\{{\ensuremath{\boldsymbol}}{x}\in V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}} \mid {\ensuremath{\boldsymbol}}{l}\cdot{\ensuremath{\boldsymbol}}{x}\in\Z, \forall{\ensuremath{\boldsymbol}}{l}\in\Lambda_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\}$ be the dual lattice of ${\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ inside $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, where $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ is now considered as an inner product subspace of $\R^m$, with respect to the standard inner product. \[lattarr\] For every ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$ holds ${\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}={\ensuremath{\Lambda}}^\star_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\oplus V^{\perp}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. Let ${\ensuremath{\boldsymbol}}z\in {\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ be arbitrary, and let ${\ensuremath{\boldsymbol}}z={\ensuremath{\boldsymbol}}x+{\ensuremath{\boldsymbol}}y$, where ${\ensuremath{\boldsymbol}}x\in V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ and ${\ensuremath{\boldsymbol}}y\in V^{\perp}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. It suffices to prove that ${\ensuremath{\boldsymbol}}l\cdot{\ensuremath{\boldsymbol}}x\in{\ensuremath{\mathbb{Z}}}$, for all ${\ensuremath{\boldsymbol}}l\in{\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\subseteq V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. For any such ${\ensuremath{\boldsymbol}}l$, we have ${\ensuremath{\boldsymbol}}l\cdot{\ensuremath{\boldsymbol}}y=0$, hence ${\ensuremath{\boldsymbol}}l\cdot{\ensuremath{\boldsymbol}}x={\ensuremath{\boldsymbol}}l\cdot{\ensuremath{\boldsymbol}}z$, and by definition ${\ensuremath{\boldsymbol}}l\cdot{\ensuremath{\boldsymbol}}z\in{\ensuremath{\mathbb{Z}}}$, thus proving ${\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}{\ensuremath{\subseteq}}{\ensuremath{\Lambda}}^\star_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\oplus V^{\perp}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. For the reverse inclusion, let ${\ensuremath{\boldsymbol}}x+{\ensuremath{\boldsymbol}}y={\ensuremath{\boldsymbol}}z\in{\ensuremath{\Lambda}}^\star_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\oplus V^{\perp}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ be arbitrary with ${\ensuremath{\boldsymbol}}x\in {\ensuremath{\Lambda}}^\star_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ and ${\ensuremath{\boldsymbol}}y\in V^{\perp}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. Also, let ${\ensuremath{\boldsymbol}}l\in{\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\subseteq V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ be arbitrary. By definition, ${\ensuremath{\boldsymbol}}l\cdot {\ensuremath{\boldsymbol}}y=0$ and ${\ensuremath{\boldsymbol}}l\cdot {\ensuremath{\boldsymbol}}x\in{\ensuremath{\mathbb{Z}}}$, therefore ${\ensuremath{\boldsymbol}}l\cdot {\ensuremath{\boldsymbol}}z\in{\ensuremath{\mathbb{Z}}}$, thus proving that $z\in {\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. Let ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$. By the previous result, ${\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ consists of infinitely many parallel affine subspaces, in a lattice arrangement. By Lemma \[mainlemma\], the condition [**(V3)**]{} is equivalent to the following: &E\_[[$\boldsymbol$]{}[$\alpha$]{}]{}\[[$\varepsilon$]{},1-[$\varepsilon$]{}\]\^m-[$\boldsymbol$]{}u\_0. & Next, denote the unit cube $[0,1]^m$ by $C_m$ and assume that [**(V3)**]{} holds for all ${\ensuremath{\boldsymbol}}u_0\in\R^m$, which is certainly true when ${\ensuremath{\varepsilon}}\leq1/(2m)$ according to Theorem \[Sch\] and the remarks in Section \[subsectBBM\]. In other words, every translate of $(1-2{\ensuremath{\varepsilon}})C_m$ intersects ${\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. Having this in mind, we project everything orthogonally onto $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. First, we note that ${\ensuremath{\boldsymbol}}e_i\in {\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ for all $1\leq i\leq m$, as ${\ensuremath{\mathbb{Z}}}^m{\ensuremath{\subseteq}}{\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ by definition, where the ${\ensuremath{\boldsymbol}}e_i$ are the standard basis vectors. Then, ${\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}={\ensuremath{\Lambda}}^\star_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ by Proposition \[lattarr\], and $C_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ is a zonotope generated by ${\ensuremath{\boldsymbol}}e_i|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}={\ensuremath{\boldsymbol}}w_i\in {\ensuremath{\Lambda}}^\star_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, $1\leq i\leq m$. In order to describe $C_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, let $\dim(V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})=n$ and let ${{\left\{{{\ensuremath{\boldsymbol}}{a}_1,\dotsc,{\ensuremath{\boldsymbol}}{a}_n}\right\}}}$ be a basis of ${\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ with ${\ensuremath{\boldsymbol}}{a}_i=(a_{i1},\dotsc,a_{im})$, $1\leq i\leq n$. Furthermore, let ${{\left\{{{\ensuremath{\boldsymbol}}{a}_1^\star,\dotsc,{\ensuremath{\boldsymbol}}{a}_n^\star}\right\}}}$ be the dual basis of ${\ensuremath{\Lambda}}^\star_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, so that ${\ensuremath{\boldsymbol}}{a}_i\cdot{\ensuremath{\boldsymbol}}{a}_j^\star={\ensuremath{\delta}}_{ij}$, where $\delta_{ij}$ is the Kronecker delta. Since ${\ensuremath{\boldsymbol}}{a}_i\cdot{\ensuremath{\boldsymbol}}{w}_j={\ensuremath{\boldsymbol}}{a}_i\cdot{\ensuremath{\boldsymbol}}e_j=a_{ij}$, we have $${\ensuremath{\boldsymbol}}{w}_j=\sum_{i=1}^n a_{ij}{\ensuremath{\boldsymbol}}{a}_i^\star.$$ Now, $C_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}=\sum_{j=1}^{m}[{\ensuremath{\mathbf{0}}},{\ensuremath{\boldsymbol}}{w}_j]$, so if we apply the linear transformation $T:V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\to\R^n$ that sends the basis ${{\left\{{{\ensuremath{\boldsymbol}}{a}_1^\star,\dotsc,{\ensuremath{\boldsymbol}}{a}_n^\star}\right\}}}$ to the standard basis of $\R^n$, we find that the zonotope $Z=T(C_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})$ is generated by the columns of the matrix ${\ensuremath{\boldsymbol}}A=(a_{ij})$, whose rows are given by ${\ensuremath{\boldsymbol}}{a}_1^\intercal,\dotsc,{\ensuremath{\boldsymbol}}{a}_n^\intercal$. We also denote this zonotope by $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ to stress its dependence on ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$. Note that $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ also depends on the choice of the basis of $\Lambda_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, however, different such choices lead to *unimodularly equivalent* zonotopes, that is, they are the same up to a linear transformation ${\ensuremath{\boldsymbol}}U\in\Z^{n\times n}$ with $\det({\ensuremath{\boldsymbol}}U)=\pm 1$. As the value of ${\ensuremath{\varepsilon}}$ in Theorem \[equivform\] is invariant under unimodular transformations of the problem, we can safely ignore this dependence in the sequel. By Proposition \[lattarr\], property [**(V3)**]{} is equivalent to the fact that ${\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}={\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\star$ intersects $(1-2{\ensuremath{\varepsilon}})C_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}-({\ensuremath{\boldsymbol}}{u}_0-{\ensuremath{\varepsilon}}{\ensuremath{\mathbf{1}}}_m)|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, which in turn is equivalent to the fact that ${\ensuremath{\mathbb{Z}}}^n$ has nontrivial intersection with $(1-2{\ensuremath{\varepsilon}})Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}-\bar{{\ensuremath{\boldsymbol}}{u}}_0$, where $\bar{{\ensuremath{\boldsymbol}}{u}}_0=T(({\ensuremath{\boldsymbol}}{u}_0-{\ensuremath{\varepsilon}}{\ensuremath{\mathbf{1}}}_m)|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})$, completing the proof of Theorem \[equivform\]. A case for rationally uniform directions {#sect_rationally_uniform} ======================================== This section begins with an extended investigation of the zonotopes that arise in Theorem \[equivform\], in fact, we see that for every direction vector ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ there are two lattice zonotopes that are associated to each other. Once this is achieved, we find that there is a correspondence between rationally uniform directions and lattice zonotopes that are generated by vectors in [LGP]{}. Finally, we first solve the view-obstruction problem in the most general situation, before concentrating on the more illuminating case of rationally uniform directions. The associated zonotope of Z\_ ------------------------------- We need to prepare ourselves with an auxiliary statement from linear algebra. For abbreviation, we write $[m]=\{1,\ldots,m\}$, and $\binom{[m]}{k}$ for the family of $k$-element subsets of $[m]$. Moreover, for any $I,J\in\binom{[m]}{k}$ and any matrix ${\ensuremath{\boldsymbol}}V\in\R^{m\times m}$, we write ${\ensuremath{\boldsymbol}}V_{I,J}\in\R^{k\times k}$ for the matrix that remains when striking out all rows of ${\ensuremath{\boldsymbol}}V$ not indexed by $I$ and all columns of ${\ensuremath{\boldsymbol}}V$ not indexed by $J$. We shall denote the complement $[m]{\ensuremath{\setminus}}I$ by $I^c$. \[lemMinorRelations\] Let ${\ensuremath{\boldsymbol}}V=({\ensuremath{\boldsymbol}}v_1,\ldots,{\ensuremath{\boldsymbol}}v_m)\in\R^{m\times m}$ be a matrix whose columns correspond to a basis of $\R^m$ and let ${\ensuremath{\boldsymbol}}V^\star=({\ensuremath{\boldsymbol}}v^\star_1,\ldots,{\ensuremath{\boldsymbol}}v^\star_m)$ be the matrix corresponding to the dual basis, that is, ${\ensuremath{\boldsymbol}}v_i\cdot{\ensuremath{\boldsymbol}}v^\star_j=\delta_{ij}$, for every $i,j\in[m]$. For any $k\in[m]$ and any $I,J\in\binom{[m]}{k}$, we have $$\det({\ensuremath{\boldsymbol}}V_{I,J})=\pm\det({\ensuremath{\boldsymbol}}V)\cdot\det({\ensuremath{\boldsymbol}}V^\star_{I^c,J^c}).$$ Our arguments are based on the theory of the exterior algebra of $\R^m$, for which we refer the reader to [@birkhoffmaclane1979algebra Ch. XVI]. The main ingredient is the fact that for any matrix ${\ensuremath{\boldsymbol}}M\in\R^{m\times m}$, we have $$\det({\ensuremath{\boldsymbol}}M)\cdot {\ensuremath{\boldsymbol}}e_1\wedge\ldots\wedge{\ensuremath{\boldsymbol}}e_m=({\ensuremath{\boldsymbol}}M{\ensuremath{\boldsymbol}}e_1)\wedge\ldots\wedge({\ensuremath{\boldsymbol}}M{\ensuremath{\boldsymbol}}e_m).$$ Note also, that $\det({\ensuremath{\boldsymbol}}V_{I,J})=\pm\det(\tilde{{\ensuremath{\boldsymbol}}V}_{I,J})$, where $\tilde{{\ensuremath{\boldsymbol}}V}_{I,J}=\left({\ensuremath{\boldsymbol}}e_i \mid i\notin I, {\ensuremath{\boldsymbol}}v_j \mid j\in J\right)$. Applying these two identities repeatedly yields $$\begin{aligned} \det({\ensuremath{\boldsymbol}}V_{I,J})\cdot {\ensuremath{\boldsymbol}}e_1\wedge\ldots\wedge{\ensuremath{\boldsymbol}}e_m&=\pm\bigwedge_{i\notin I}{\ensuremath{\boldsymbol}}e_i\wedge\bigwedge_{j\in J}{\ensuremath{\boldsymbol}}v_j\\ &=\pm\det({\ensuremath{\boldsymbol}}V)\cdot \bigwedge_{i\notin I}({\ensuremath{\boldsymbol}}V^{-1})_i\wedge\bigwedge_{j\in J}{\ensuremath{\boldsymbol}}e_j\\ &=\pm\det({\ensuremath{\boldsymbol}}V)\cdot\det({\ensuremath{\boldsymbol}}V^{-1}_{J^c,I^c})\cdot {\ensuremath{\boldsymbol}}e_1\wedge\ldots\wedge{\ensuremath{\boldsymbol}}e_m.\end{aligned}$$ Therefore, $\det({\ensuremath{\boldsymbol}}V_{I,J})=\pm\det({\ensuremath{\boldsymbol}}V)\cdot\det({\ensuremath{\boldsymbol}}V^{-1}_{J^c,I^c})$. Since ${\ensuremath{\boldsymbol}}V^\star={\ensuremath{\boldsymbol}}V^{-\intercal}$, we have $\det({\ensuremath{\boldsymbol}}V^{-1}_{J^c,I^c})=\det({\ensuremath{\boldsymbol}}V^{-\intercal}_{I^c,J^c})=\det({\ensuremath{\boldsymbol}}V^\star_{I^c,J^c})$, which finishes the proof. One could write down the sign that appears in Lemma \[lemMinorRelations\] explicitely, depending on the chosen sets $I$ and $J$. However, we refrain here from doing so, since it is not important for our purposes. Extending the investigation from the previous section, we find that a given ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$ actually gives rise to two lattice zonotopes that are associated to each other in the sense of Shephard [@shephard1974combinatorial]. In order to make this precise, let as before ${\ensuremath{\boldsymbol}}A=({\ensuremath{\boldsymbol}}a_1,\ldots,{\ensuremath{\boldsymbol}}a_n)^\intercal$, where ${{\left\{{{\ensuremath{\boldsymbol}}a_1,\ldots,{\ensuremath{\boldsymbol}}a_n}\right\}}}$ is a basis of $\Lambda_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. Now, we can extend this basis to a basis ${{\left\{{{\ensuremath{\boldsymbol}}a_1,\ldots,{\ensuremath{\boldsymbol}}a_n,\ldots,{\ensuremath{\boldsymbol}}a_m}\right\}}}$ of $\Z^m$ since $\Lambda_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}=V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\cap\Z^m$. Taking the dual basis ${{\left\{{{\ensuremath{\boldsymbol}}a_1^\star,\ldots,{\ensuremath{\boldsymbol}}a_m^\star}\right\}}}$ thereof, we find that ${{\left\{{{\ensuremath{\boldsymbol}}a_{n+1}^\star,\ldots,{\ensuremath{\boldsymbol}}a_m^\star}\right\}}}$ is a basis of $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp\cap\Z^m$, and we write ${\ensuremath{\boldsymbol}}A^\perp=({\ensuremath{\boldsymbol}}a_{n+1}^\star,\ldots,{\ensuremath{\boldsymbol}}a_m^\star)^\intercal$ (see, e.g. [@martinet2003perfect Ch. 1]). Now, using the map $T:\R^m\to\R^n\times\R^{m-n}$ that sends ${\ensuremath{\boldsymbol}}a_i^\star$ to the coordinate unit vector ${\ensuremath{\boldsymbol}}e_i$, for all $i\in[m]$, we obtain the lattice zonotopes $$Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}=T\left(C_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\right)\subseteq\R^n\quad\text{and}\quad Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp=T\left(C_m|V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp\right)\subseteq\R^{m-n},$$ which are generated by the columns of the matrices ${\ensuremath{\boldsymbol}}A$ and ${\ensuremath{\boldsymbol}}A^\perp$, respectively. It is well-known that a zonotope $Z=\sum_{i=1}^m[{\ensuremath{\boldsymbol}}0,{\ensuremath{\boldsymbol}}z_i]\subseteq\R^n$ can be dissected into translates of the parallelepipeds $Z_I=\sum_{i\in I}[{\ensuremath{\boldsymbol}}0,{\ensuremath{\boldsymbol}}z_i]$, $I\in\binom{[m]}{n}$ (see, e.g. [@shephard1974combinatorial]). The volume of the parallelepipeds that $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ and $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp$ are composed of are related as follows. \[propVolParalleps\] Let ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$. Then, for every $I\in\binom{[m]}{n}$, we have $$\operatorname{vol}_n((Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})_I)=\operatorname{vol}_{m-n}((Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp)_{I^c}).$$ In particular, if ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\Z_{>0}^m$ with $\gcd(\alpha_1,\ldots,\alpha_m)=1$, then for every $i\in[m]$, $$\alpha_i=\operatorname{vol}_{m-1}((Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})_{[m]{\ensuremath{\setminus}}{{\left\{{i}\right\}}}}).$$ The volume of a parallelepiped is given by the absolute value of the determinant of its generators. So, we can just apply Lemma \[lemMinorRelations\] to the matrices ${\ensuremath{\boldsymbol}}V=({\ensuremath{\boldsymbol}}a_1,\ldots,{\ensuremath{\boldsymbol}}a_m)$ and ${\ensuremath{\boldsymbol}}V^\star=({\ensuremath{\boldsymbol}}a_1^\star,\ldots,{\ensuremath{\boldsymbol}}a_m^\star)$, and obtain $$\operatorname{vol}_n((Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})_I)=|\det({\ensuremath{\boldsymbol}}V_{I,[n]})|=|\det({\ensuremath{\boldsymbol}}V)|\cdot|\det({\ensuremath{\boldsymbol}}V^\star_{I^c,[m]{\ensuremath{\setminus}}[n]})|=\operatorname{vol}_{m-n}((Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp)_{I^c}),$$ where we used that $\det({\ensuremath{\boldsymbol}}V)=\pm1$. In the special case that ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\Z^m_{>0}$ with $\gcd(\alpha_1,\ldots,\alpha_m)=1$, we have $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp=\operatorname{lin}\{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\}$ and hence $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp\cap\Z^m=\Z{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$. Therefore, we have ${\ensuremath{\boldsymbol}}A^\perp={\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}^\intercal$ and thus $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp=\sum_{i=1}^m[0,\alpha_i]$, implying the claim. Rationally uniform vectors {#subsect_rationally_uniform} -------------------------- Next, we study the situation when ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ is *rationally uniform*. We show that ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ possesses this property, if and only if the vectors generating $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ are in [LGP]{}. Two auxiliary statements are needed. Let ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}\in\R^m$ and let $\Lambda_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}},V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ be defined as above, in particular, let $n=\dim(V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})$. Then, $\dim_\Q({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})=m-n$. Let $f:\Q^m\to\R$ be the $\Q$-linear map defined by $f({\ensuremath{\boldsymbol}}\ell)={\ensuremath{\boldsymbol}}\ell\cdot{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$. Clearly, $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}=\operatorname{lin}(\ker(f))$ and $\dim_\Q({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})=\dim_\Q(\operatorname{\ensuremath{\operatorname{im}}}(f))$. Hence, by the rank-nullity theorem we get $m=\dim_\Q(\ker(f))+\dim_\Q(\operatorname{\ensuremath{\operatorname{im}}}(f))=n+\dim_\Q({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})$. \[propMinorsAndSubalphas\] Let ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$ and let $\{{\ensuremath{\boldsymbol}}a_1,\ldots,{\ensuremath{\boldsymbol}}a_n\}$ be a basis of $\Lambda_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. Define ${\ensuremath{\boldsymbol}}A\in\Z^{n\times m}$, as before, as the matrix whose $i$th row is given by ${\ensuremath{\boldsymbol}}a_i$. For a subset $I\subseteq[m]$, let ${\ensuremath{\boldsymbol}}A_I$ be the submatrix of ${\ensuremath{\boldsymbol}}A$ consisting of those columns of ${\ensuremath{\boldsymbol}}A$ that are indexed by $I$, and define ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}_I$ analogously. Then, if $|I|=n$, $$\det({\ensuremath{\boldsymbol}}A_I)\neq 0\quad\Longleftrightarrow\quad\dim_\Q({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}_{I^c})=m-n.$$ Let $L_I=\operatorname{lin}\{{\ensuremath{\boldsymbol}}e_i \mid i\in I\}$. Since by definition $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp=\ker({\ensuremath{\boldsymbol}}A)$ and $\dim(V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp)+\dim(L_I)=m$, the following equivalences hold: $$\begin{aligned} \det({\ensuremath{\boldsymbol}}A_I)\neq 0 &\Longleftrightarrow \forall{\ensuremath{\boldsymbol}}\ell\in\R^m,{\ensuremath{\operatorname{supp}}}({\ensuremath{\boldsymbol}}\ell)\subseteq I,{\ensuremath{\boldsymbol}}A\cdot{\ensuremath{\boldsymbol}}\ell=0 \Longrightarrow {\ensuremath{\boldsymbol}}\ell={\ensuremath{\boldsymbol}}0\\ &\Longleftrightarrow V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp\cap L_I=\{{\ensuremath{\boldsymbol}}0\}\\ &\Longleftrightarrow V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\cap L_{I^c}=V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\cap L_I^\perp=\{{\ensuremath{\boldsymbol}}0\}.\end{aligned}$$ As $\dim(\Lambda_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})=\dim(V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})$ and $\operatorname{lin}_\Q(\Lambda_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})$ is dense in $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, the latter is equivalent to the fact that every ${\ensuremath{\boldsymbol}}\ell\in\Z^m$ with ${\ensuremath{\boldsymbol}}\ell\cdot{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}=0$ and ${\ensuremath{\operatorname{supp}}}({\ensuremath{\boldsymbol}}\ell)\subseteq I^c$ is already the zero-vector, that is, ${\ensuremath{\boldsymbol}}\ell={\ensuremath{\boldsymbol}}0$. In other words, the coordinates of ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ that are indexed by $I^c$ are linearly independent over $\Q$, which means that $\dim_\Q({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}_{I^c})=m-n$ as claimed. \[cor\_ratunif\_lgp\]  \ 1. The vector ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$ is rationally uniform if and only if the zonotope $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ is generated by vectors in [LGP]{}. 2. The zonotope $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ is generated by vectors in [LGP]{} if and only if the associated zonotope $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp$ is. 1): By the analysis in Subsection \[subsect\_zonotopes\_vo\], we can take a map $T$ such that $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ is generated by the columns of the matrix ${\ensuremath{\boldsymbol}}A$, as described in Proposition \[propMinorsAndSubalphas\]. The conclusion of the latter proves the desired fact. Part 2) is a direct consequence of Lemma \[lemMinorRelations\] and Proposition \[propMinorsAndSubalphas\]. Shephard [@shephard1974combinatorial] proves Corollary \[cor\_ratunif\_lgp\] 2) combinatorially in the language of *cubical zonotopes*. So far, we have only seen zonotopes attached to a vector ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$, and Corollary \[cor\_ratunif\_lgp\] gives us a characterization of such zonotopes that are generated by vectors in [LGP]{}. This construction can be reversed: Suppose we have a lattice zonotope $Z{\ensuremath{\subseteq}}\R^n$, generated by $m$ vectors in [LGP]{}. Let ${\ensuremath{\boldsymbol}}A$ be the $n\times m$ matrix whose columns correspond to these vectors. Taking the rows of ${\ensuremath{\boldsymbol}}A$, we obtain a basis of an $n$-dimensional lattice ${\ensuremath{\Lambda}}$ in $\R^m$. Let $V$ be the space spanned by ${\ensuremath{\Lambda}}$, and let $V^{\perp}$ be its orthogonal complement. Now, one can choose a vector ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in V^{\perp}$, that is not orthogonal to any lattice vector, except from those of ${\ensuremath{\Lambda}}$. Indeed, if a vector ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}$ does not belong to $V$, then ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}^{\perp}\cap V^{\perp}$ is a codimension-one subspace of $V^{\perp}$. So, ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ needs to avoid countably many hyperplanes of $V^{\perp}$, and as is well-known, this is indeed possible, because countably many hyperplanes cannot cover the entire space. We conclude by noting that in this case, ${\ensuremath{\Lambda}}={\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, and the zonotopes $Z$ and $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ are unimodularly equivalent. If, for ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$, the conditions [**(V1)**]{}-[**(V4)**]{} in Theorem \[equivform\] hold for all ${\ensuremath{\boldsymbol}}u_0\in\R^m$ and some ${\ensuremath{\varepsilon}}>0$, then we can deduce that any translate of $(1-2{\ensuremath{\varepsilon}})Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ contains a lattice point, or in other words, $$\begin{aligned} \mu(Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})&\leq 1-2{\ensuremath{\varepsilon}}.\end{aligned}$$ This interpretation of the covering radius is classical; we refer, for instance, to [@gruberlekker1987geometry §13]. Therefore, Theorem \[equivform\] implies that $$\begin{aligned} \sup\mu(Z) &= 1-2{\ensuremath{\varepsilon}}(n,m),\label{eqnEpsSupMu}\end{aligned}$$ where the supremum is taken over all lattice zonotopes $Z{\ensuremath{\subseteq}}\R^k$ with $m$ generators in [LGP]{}, and where $k\leq n$. In the case $n=m-1$, Theorem \[Sch\] gives us ${\ensuremath{\varepsilon}}(m-1,m)=1/(2m)$, and thus, a reinterpretation of Schoenberg’s result is as follows. \[zonocodim1\] Let $Z=\sum_{i=1}^{n+1}[{\ensuremath{\mathbf{0}}},{\ensuremath{\boldsymbol}}z_i]$, where ${{\left\{{{\ensuremath{\boldsymbol}}z_1,\dotsc,{\ensuremath{\boldsymbol}}z_{n+1}}\right\}}}{\ensuremath{\subseteq}}{\ensuremath{\mathbb{Z}}}^n$ is in [LGP]{}. Then $\mu(Z)\leq \frac{n}{n+1}$. Equality is attained for ${\ensuremath{\boldsymbol}}z_i={\ensuremath{\boldsymbol}}e_i$, $1\leq i\leq n$, ${\ensuremath{\boldsymbol}}z_{n+1}={\ensuremath{\mathbf{1}}}_n$. The case $n=1$ can also be treated easily in terms of this zonotopal approach. Indeed, a one-dimensional lattice zonotope $Z$ that is generated by $m$ vectors in [LGP]{} is just an interval with integer endpoints and length at least $m$. Its covering radius is just the inverse of its length, thus $\mu(Z)\geq1/m$, which in view of  translates into ${\ensuremath{\varepsilon}}(1,m)\leq(m-1)/(2m)$. An example showing that this is actually an identity is given by any ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in\R^m$ such that ${\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}=\Z\cdot{\ensuremath{\boldsymbol}}1_m$, because this yields an interval $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ of length exactly $m$. It is easy to see that any ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}=({\ensuremath{\alpha}}_1,\ldots,{\ensuremath{\alpha}}_{m-1},-\alpha_1-\ldots-\alpha_{m-1})^\intercal$ with ${{\left\{{\alpha_1,\ldots,\alpha_{m-1}}\right\}}}$ linearly independent over $\Q$ will do the job. Hence, ${\ensuremath{\varepsilon}}(1,m)=(m-1)/(2m)$. Together with the trivial cases $n=0$ and $n=m$ the above considerations show that Conjecture \[mainconj\] holds for all $n\in{{\left\{{0,1,m-1,m}\right\}}}$. The question remains in the intermediate cases, where we obtain a weaker bound, namely Theorem \[thmCoveringRadiusAsymptotics\] (see Section \[sect\_flatness\] for details). The monotonicity of the function (n,m) -------------------------------------- Remember that one of our main interests is to determine, or at least estimate, the supremum among all ${\ensuremath{\varepsilon}}\geq0$ such that the equivalent statements in Theorem \[equivform\] hold for every starting point ${\ensuremath{\boldsymbol}}u_0\in\R^m$ and every nontrivial direction ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in(\R{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m$ with $\dim_{\Q}({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})=m-n$. We first see that in this generality the problem can be solved exactly but depends only on the parameter $n$. To this end, let us consider the numbers $$\begin{aligned} {\overline}{\ensuremath{\varepsilon}}(n,m)=\sup\big\{{\ensuremath{\varepsilon}}\geq0 \mid &\,{\bf (Vi)}\text{ holds for any }{\ensuremath{\boldsymbol}}u_0\in\R^m\text{ and }{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in(\R{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m\\ &\text{ such that }\dim_{\Q}({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})\geq m-n\big\}.\end{aligned}$$ \[propPropertiesBarEpsFunction\] Let $n,m\in\N$ be such that $n\leq m$. Then, 1. ${\overline}{\ensuremath{\varepsilon}}(n,m+1)\leq{\overline}{\ensuremath{\varepsilon}}(n,m)$, and 2. if $n<m$, then ${\overline}{\ensuremath{\varepsilon}}(n,m)=\frac1{2(n+1)}$. 1): For the sake of brevity let ${\overline}{\ensuremath{\varepsilon}}={\overline}{\ensuremath{\varepsilon}}(n,m+1)$. Let ${\ensuremath{\boldsymbol}}{u}'_0\in\R^m$ and ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'\in(\R\setminus\{0\})^m$ with $\dim_\Q({\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}')\geq m-n$ be arbitrary. Now, let ${\ensuremath{\boldsymbol}}{u}_0=({\ensuremath{\boldsymbol}}{u}'_0,t)\in\R^{m+1}$, for some $t\in\R$, and let ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}=({\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}',a)$, for some $a\in\R\setminus\{0\}$ that is rationally independent from the entries of ${\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'$. Note that $\dim_\Q({\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}})=\dim_\Q({\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}')+1\geq m+1-n$ and ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in(\R{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^{m+1}$. By definition of ${\overline}{\ensuremath{\varepsilon}}$, we have $({\ensuremath{\boldsymbol}}{u}_0+{\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})\cap[{\overline}{\ensuremath{\varepsilon}},1-{\overline}{\ensuremath{\varepsilon}}]^{m+1}\neq{\ensuremath{\varnothing}}$. The choice of $a$ implies that no multiple $qa$, $q\in\Z\setminus\{0\}$, can be written as ${\ensuremath{\boldsymbol}}{\ell}'\cdot{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'$, for some ${\ensuremath{\boldsymbol}}{\ell}'\in\Z^m$. Therefore, $\Lambda_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}=\Lambda_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'}\times\{0\}$, and hence $$\begin{aligned} \pi({\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}})&=\left\{\pi({\ensuremath{\boldsymbol}}{\xi}) \mid {\ensuremath{\boldsymbol}}{\xi}\in\R^{m+1},{\ensuremath{\boldsymbol}}{\ell}\cdot{\ensuremath{\boldsymbol}}{\xi}\in\Z,\,\forall {\ensuremath{\boldsymbol}}{\ell}\in\Lambda_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}}\right\}\\ &=\left\{{\ensuremath{\boldsymbol}}{\xi}'\in\R^m \mid {\ensuremath{\boldsymbol}}{\ell}'\cdot{\ensuremath{\boldsymbol}}{\xi}'\in\Z,\,\forall {\ensuremath{\boldsymbol}}{\ell}'\in\Lambda_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'}\right\}={\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'},\end{aligned}$$ where $\pi:\R^{m+1}\to\R^m$ is the projection that forgets the last coordinate. As a consequence we obtain that $$({\ensuremath{\boldsymbol}}{u}'_0+{\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}'})\cap[{\overline}{\ensuremath{\varepsilon}},1-{\overline}{\ensuremath{\varepsilon}}]^m=\pi({\ensuremath{\boldsymbol}}{u}_0+{\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{{\ensuremath{\alpha}}}})\cap\pi([{\overline}{\ensuremath{\varepsilon}},1-{\overline}{\ensuremath{\varepsilon}}]^{m+1})\neq{\ensuremath{\varnothing}},$$ and hence ${\overline}{\ensuremath{\varepsilon}}(n,m)\geq{\overline}{\ensuremath{\varepsilon}}={\overline}{\ensuremath{\varepsilon}}(n,m+1)$ as desired. 2): In view of part 1), we have $${\overline}{\ensuremath{\varepsilon}}(n,m)\leq{\overline}{\ensuremath{\varepsilon}}(n,m-1)\leq\ldots\leq{\overline}{\ensuremath{\varepsilon}}(n,n+1)\leq{\ensuremath{\varepsilon}}(n,n+1)=\frac1{2(n+1)},$$ where the last equation is due to Theorem \[Sch\]. For the lower bound, we consider the vector $${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}=({\ensuremath{\alpha}}_1,\ldots,{\ensuremath{\alpha}}_{m-n},{\ensuremath{\alpha}}_1,\ldots,{\ensuremath{\alpha}}_1)\in(\R{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m,$$ where ${{\left\{{{\ensuremath{\alpha}}_1,\ldots,{\ensuremath{\alpha}}_{m-n}}\right\}}}$ is linearly independent over $\Q$. A basis of the lattice $\Lambda_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$ is given by $\{{\ensuremath{\boldsymbol}}e_1-{\ensuremath{\boldsymbol}}e_{m-n+1},{\ensuremath{\boldsymbol}}e_1-{\ensuremath{\boldsymbol}}e_{m-n+2},\ldots,{\ensuremath{\boldsymbol}}e_1-{\ensuremath{\boldsymbol}}e_m\}$, so we find that $Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}=[0,-1]^n+[{\ensuremath{\boldsymbol}}0,{\ensuremath{\boldsymbol}}1_n]$. This is exactly the same zonotope $Z_{{\overline}{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}}$ that is induced by ${\overline}{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}={\ensuremath{\boldsymbol}}1_{n+1}\in\R^{n+1}$. By the example following Theorem \[Sch\] we know that $\mu(Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})=\mu(Z_{{\overline}{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}})=\frac{n}{n+1}$, and hence ${\overline}{\ensuremath{\varepsilon}}(n,m)\geq\frac12(1-\mu(Z_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}))=\frac1{2(n+1)}$. Our intuition says that the more we restrict rational dependencies in the direction vector ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ the larger we can choose ${\ensuremath{\varepsilon}}$ in Theorem \[equivform\]. Proposition \[propPropertiesBarEpsFunction\] above shows that in order for this to be true, we need to impose stronger conditions than only $\dim_{\Q}({\ensuremath{\boldsymbol}}{\ensuremath{\alpha}})\geq m-n$. In fact, these remarks explain and justify the restriction to rationally uniform vectors in the definition of ${\ensuremath{\varepsilon}}(n,m)$, and hence in the formulation of Conjecture \[mainconj\]. In order to study this situation in more detail, we use the zonotopal definition of ${\ensuremath{\varepsilon}}(n,m)$ provided by [**(V4)**]{} together with Corollary \[cor\_ratunif\_lgp\], that is, $$\begin{aligned} {\ensuremath{\varepsilon}}(n,m)=\sup\big\{{\ensuremath{\varepsilon}}\geq0 \mid &\,\mu(Z,\Z^k)\leq1-2{\ensuremath{\varepsilon}}\text{ for all lattice zonotopes }Z\subseteq\R^k\\ &\,\text{generated by }m\text{ vectors in {LGP}\ and where }k\leq n\big\}.\end{aligned}$$ The following monotonicity properties of the function ${\ensuremath{\varepsilon}}(n,m)$ are compatible with our conjecture that ${\ensuremath{\varepsilon}}(n,m)=(m-n)/(2m)$ (see the remark after Conjecture \[mainconj\]). \[propPropertiesEpsFunction\] Let $n,m\in\N$ be such that $n\leq m$. Then, we have 1. ${\ensuremath{\varepsilon}}(n,m)\leq{\ensuremath{\varepsilon}}(n-1,m)$ and ${\ensuremath{\varepsilon}}(n,m)\leq{\ensuremath{\varepsilon}}(n,m+1)$, 2. ${\ensuremath{\varepsilon}}(n,m)\leq{\ensuremath{\varepsilon}}(n-1,m-1)\leq\ldots\leq{\ensuremath{\varepsilon}}(1,m-n+1)=\frac{m-n}{2(m-n+1)}$. 1): The inequality ${\ensuremath{\varepsilon}}(n,m)\leq{\ensuremath{\varepsilon}}(n-1,m)$ follows directly from the definition of ${\ensuremath{\varepsilon}}(n,m)$. For the second inequality, let $Z'=\sum_{i=1}^{m+1}[{\ensuremath{\boldsymbol}}0,{\ensuremath{\boldsymbol}}z_i]\subseteq\R^k$, for some $k\leq n$, be generated by lattice vectors in [LGP]{}. Dropping the last generator gives us a zonotope $Z=\sum_{i=1}^m[{\ensuremath{\boldsymbol}}0,{\ensuremath{\boldsymbol}}z_i]$, which in view of $k\leq n<m+1$ is of the same dimension as $Z'$ and generated by vectors in [LGP]{}. Since, clearly $Z\subseteq Z'$, this implies $\mu(Z',\Z^k)\leq\mu(Z,\Z^k)\leq1-2{\ensuremath{\varepsilon}}(n,m)$. 2): It suffices to show that ${\ensuremath{\varepsilon}}(n,m)\leq{\ensuremath{\varepsilon}}(n-1,m-1)$. Let $Z'=\sum_{i=1}^{m-1}[{\ensuremath{\boldsymbol}}0,{\ensuremath{\boldsymbol}}z'_i]$ be generated by $\{{\ensuremath{\boldsymbol}}z'_1,\ldots,{\ensuremath{\boldsymbol}}z'_{m-1}\}\subseteq\Z^k$ in [LGP]{}, for some $k\leq n-1$. We define ${\ensuremath{\boldsymbol}}z_m={\ensuremath{\boldsymbol}}e_{k+1}$ and ${\ensuremath{\boldsymbol}}z_i=({\ensuremath{\boldsymbol}}z'_i,h_i)$, for $i\in[m-1]$, where $h_i\in\Z$ are chosen such that $\{{\ensuremath{\boldsymbol}}z_1,\ldots,{\ensuremath{\boldsymbol}}z_m\}$ is in [LGP]{}. Now, the zonotope $Z=\sum_{i=1}^m[{\ensuremath{\boldsymbol}}0,{\ensuremath{\boldsymbol}}z_i]\subseteq\R^{k+1}$ projects onto $Z'$, that is, $Z'=\pi(Z)$, and clearly $\Z^k=\pi(\Z^{k+1})$, where $\pi:\R^{k+1}\to\R^k$ forgets the last coordinate. Therefore, we have $\mu(Z',\Z^k)\leq\mu(Z,\Z^{k+1})\leq1-2{\ensuremath{\varepsilon}}(n,m)$, which implies the desired inequality. Bounds on (n,m) via the Flatness Theorem {#sect_flatness} ======================================== In this section, we prove Theorem \[thmCoveringRadiusAsymptotics\]. Our arguments are based on the so-called *Flatness Theorem*, which states that every convex body that does not contain lattice points in its interior is necessarily flat in a lattice direction. \[thmFlatness\] Let $ n \in \N $ and let $ K \subseteq \R^n $ be a convex body with $ \operatorname{int}(K+{\ensuremath{\boldsymbol}}t) \cap \Z^n = {\ensuremath{\varnothing}}$, for some ${\ensuremath{\boldsymbol}}t\in\R^n$. There exists a vector $ {\ensuremath{\boldsymbol}}{v} \in \Z^n\setminus\{{\ensuremath{\boldsymbol}}0\}$ and a constant $ \operatorname{Flt}(n) $ only depending on $n$ such that $$w(K, {\ensuremath{\boldsymbol}}{v}) := \max_{{\ensuremath{\boldsymbol}}{x} \in K} {\ensuremath{\boldsymbol}}{x}\cdot {\ensuremath{\boldsymbol}}{v} - \min_{{\ensuremath{\boldsymbol}}{x} \in K} {\ensuremath{\boldsymbol}}{x}\cdot {\ensuremath{\boldsymbol}}{v}\le \operatorname{Flt}(n).$$ There has been quite a lot of research on estimating the dimensional constant $\operatorname{Flt}(n)$. We refer to the textbook by Barvinok [@barvinok2002acourse Ch. VII.8] for a proof of Khinchin’s result, a discussion of the history of the problem, as well as for the current state of the art. The best result to date for the class of $o$-symmetric convex bodies, that is, convex bodies $K\subseteq\R^n$ such that $K=-K$, is due to Banaszczyk [@banaszczyk1996inequalities], who proved that $$\begin{aligned} \text{If }K\subseteq\R^n\text{ is }o\text{-symmetric, then }\operatorname{Flt}(n)\leq c\,n\log{n},\text{ for some }c>0.\label{eqnFlatnessBound}\end{aligned}$$ It is generally believed that the optimal such bound is $\operatorname{Flt}(n)\in O(n)$. Recall that the covering radius $\mu(K)$ of $K$ is the minimal dilation $\mu>0$ such that every translate of $\mu K$ contains a point of $\Z^n$. Therefore, the Flatness Theorem reformulates as $$\begin{aligned} \mu(K)w(K)\leq\operatorname{Flt}(n),\label{eqnFlatnessReformulation}\end{aligned}$$ where $w(K)=\min_{{\ensuremath{\boldsymbol}}{v}\in\Z^n\setminus\{{\ensuremath{\boldsymbol}}{0}\}}w(K,{\ensuremath{\boldsymbol}}{v})$ denotes the *lattice width* of $K$. As a consequence, in order to obtain upper bounds on the covering radius $\mu(Z)$ of a lattice zonotope $Z$, it suffices to study its lattice width. \[lemLatticeWidth\] Let $S=\{{\ensuremath{\boldsymbol}}{z}_1,\ldots,{\ensuremath{\boldsymbol}}{z}_m\}\subset\Z^n$, $m\geq n$, and let $Z=\sum_{i=1}^m[{\ensuremath{\boldsymbol}}{0},{\ensuremath{\boldsymbol}}{z}_i]$ be the lattice zonotope generated by $S$. If $S$ is in [LGP]{}, we have $$w(Z)\geq m-n+1.$$ For every $m\geq n$ there exists such a set $S$ in [LGP]{} attaining the bound. Let ${\ensuremath{\boldsymbol}}{v}\in\Z^n\setminus\{{\ensuremath{\boldsymbol}}{0}\}$. Since all vertices of $Z$ are lattice points, the width of $Z$ in direction ${\ensuremath{\boldsymbol}}{v}$ is one less than the number of lattice planes parallel to ${\ensuremath{\boldsymbol}}{v}^\perp=\{{\ensuremath{\boldsymbol}}{x}\in\R^n \mid {\ensuremath{\boldsymbol}}{v}\cdot{\ensuremath{\boldsymbol}}{x}=0\}$ that intersect $Z$. More precisely, writing $L_{{\ensuremath{\boldsymbol}}{v}}=\operatorname{lin}\{{\ensuremath{\boldsymbol}}{v}\}$, we have $$w(Z,{\ensuremath{\boldsymbol}}{v})={\lvert\left\{{\ensuremath{\boldsymbol}}{y}\in\Z^n|L_{{\ensuremath{\boldsymbol}}{v}} \ \big|\ ({\ensuremath{\boldsymbol}}{v}^\perp+{\ensuremath{\boldsymbol}}{y})\cap Z\neq{\ensuremath{\varnothing}}\right\}\rvert}-1,$$ where $\Z^n|L_{{\ensuremath{\boldsymbol}}{v}}$ is the lattice that arises as the projection of $\Z^n$ onto $L_{{\ensuremath{\boldsymbol}}{v}}$. Hence, we need to estimate the number of points of $\Z^n|L_{{\ensuremath{\boldsymbol}}{v}}$ in the (one-dimensional) projected zonotope $Z|L_{{\ensuremath{\boldsymbol}}{v}}=\sum_{i=1}^m[0,{\ensuremath{\boldsymbol}}{z}_i|L_{{\ensuremath{\boldsymbol}}{v}}]$. Since $S$ is in [LGP]{}, the hyperplane ${\ensuremath{\boldsymbol}}{v}^\perp$ contains at most $n-1$ generators of $Z$. Hence, there are at least $m-(n-1)$ nonzero generators of $Z|L_{{\ensuremath{\boldsymbol}}{v}}$, which implies that the segment $Z|L_{{\ensuremath{\boldsymbol}}{v}}$ contains at least $m-(n-1)+1$ points of $\Z^n|L_{{\ensuremath{\boldsymbol}}{v}}$. From the above considerations this yields $w(Z,{\ensuremath{\boldsymbol}}{v})\geq m-n+1$, and as ${\ensuremath{\boldsymbol}}v\in\Z^n{\ensuremath{\setminus}}{{\left\{{{\ensuremath{\boldsymbol}}0}\right\}}}$ was arbitrary, we get $w(Z)\geq m-n+1$ as desired. In order to show that the bound cannot be improved in general, we construct a set of generators in [LGP]{}, that satisfies the above projection properties extremally for a particular lattice direction ${\ensuremath{\boldsymbol}}{v}$. For ${\ensuremath{\boldsymbol}}{v}={\ensuremath{\boldsymbol}}e_1$ such a set is given by $$\begin{aligned} S=\left\{{\ensuremath{\boldsymbol}}e_1,\ldots,{\ensuremath{\boldsymbol}}e_n\right\} \cup \left\{(1,\ell,\ell^2,\ldots,\ell^{n-1})^\intercal \mid \ell=1,\ldots,m-n\right\}.\label{eqnExtremalGeneratingSet}\end{aligned}$$ The determinant formula for the Vandermonde matrix readily implies that $S$ is indeed in [LGP]{}. The upper bound follows by combining the Flatness Theorem in its formulation , with Lemma \[lemLatticeWidth\] above and Banaszczyk’s bound  on $\operatorname{Flt}(n)$. Note, that we can apply the latter since zonotopes are $o$-symmetric up to translation. We have seen in Lemma \[lemLatticeWidth\] that the lattice zonotope $Z$ generated by the set $S$ in  has lattice width $w(Z)=m-n+1$. Now, for any convex body $K\subseteq\R^n$, one has $\mu(K)w(K)\geq1$. In fact, assuming that $K$ is scaled such that $w(K)=1$, we find a vector ${\ensuremath{\boldsymbol}}v\in\Z^n{\ensuremath{\setminus}}{{\left\{{{\ensuremath{\boldsymbol}}0}\right\}}}$ such that $w(K,{\ensuremath{\boldsymbol}}v)=1$, which means that up to a translation $K$ is sandwiched between two parallel consecutive lattice planes orthogonal to ${\ensuremath{\boldsymbol}}v$. Hence, there exists a translate of $K$ whose interior does not contain lattice points, and thus $\mu(K)\geq1$. Together with the previous observations, the covering radius of the lattice zonotope $Z$ can now be bounded by $\mu(Z)\geq1/(m-n+1)$, as desired. The construction of lattice zonotopes $Z$ with $\mu(Z)\geq1/(m-n+1)$ in the above proof together with  shows the upper bound $${\ensuremath{\varepsilon}}(n,m)=\frac12\left(1-\sup \mu(Z)\right)\leq\frac12\left(1-\frac1{m-n+1}\right)=\frac{m-n}{2(m-n+1)}.$$ This coincides with the bound derived by the monotonicity of the function ${\ensuremath{\varepsilon}}(n,m)$ in Proposition \[propPropertiesEpsFunction\]. A Reformulation of the Lonely Runner Conjecture {#sect_LRC} =============================================== In this section, we consider the special case of billiard ball motions, or equivalently view-obstruction problems, where the starting point ${\ensuremath{\boldsymbol}}u_0={\ensuremath{\boldsymbol}}0$. This restricted variant was independently introduced by Wills [@willslrc] as a problem in Diophantine approximation (see the description  below) and by Cusick [@cusickviewob] in the form of the view-obstruction formulation. Goddyn came up with the nowadays very popular interpretation which he coined the *Lonely Runner Conjecture*. It states that if $m$ runners with nonzero constant velocities run on a circular track of length $1$, with common starting point, then in a certain moment in time all runners are away from the starting point, having distance at least $1/(m+1)$. This conjecture has been proven for all $m\leq6$ but is open for all other cases; see [@barajasserra2008the] for the proof for $m=6$ and more background information on the problem[^3]. As a corollary to Theorem \[equivform\], we may summarize the equivalent interpretations as follows. We use the symbol ${\ensuremath{\boldsymbol}}v$ instead of ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ in the sequel to stress that the problem only depends on the velocities of the runners. \[corEquivLRP\] Let ${\ensuremath{\boldsymbol}}v\in\R^m$, $0\leq{\ensuremath{\varepsilon}}\leq 1/2$, and $n=\dim(V_{{\ensuremath{\boldsymbol}}v})$. The following statements are equivalent. 1. The $\operatorname{bbm}({\ensuremath{\boldsymbol}}0,{\ensuremath{\boldsymbol}}v)$ in $[0,1]^m$ intersects $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m$. 2. The line $\left\{t{\ensuremath{\boldsymbol}}v \mid t\in{\ensuremath{\mathbb{R}}}\right\}/\,\Z^m$ in ${\ensuremath{\mathbb{T}}}^m$ intersects $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m$. 3. The view from ${\ensuremath{\boldsymbol}}0$ with direction ${\ensuremath{\boldsymbol}}v$ is obstructed by $[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m+{\ensuremath{\mathbb{Z}}}^m$. 4. $\left((1-2{\ensuremath{\varepsilon}})Z_{{\ensuremath{\boldsymbol}}v}+{\ensuremath{\boldsymbol}}c\right)\cap\Z^n\neq{\ensuremath{\varnothing}}$, where ${\ensuremath{\boldsymbol}}c=T({\ensuremath{\varepsilon}}{\ensuremath{\mathbf{1}}}_m | V_{{\ensuremath{\boldsymbol}}v})$. As before we want to determine the maximum ${\ensuremath{\varepsilon}}\in[0,1/2]$ such that any of these equivalent statements hold for all velocity vectors ${\ensuremath{\boldsymbol}}v$, under certain natural constraints. Analogously to ${\ensuremath{\varepsilon}}(n,m)$ and ${\overline}{\ensuremath{\varepsilon}}(n,m)$, we therefore define $$\begin{aligned} {\ensuremath{\varepsilon}}_0(n,m)=\sup\{{\ensuremath{\varepsilon}}\geq0 \mid &\,{\bf (Li)}\text{ holds for all rationally uniform }{\ensuremath{\boldsymbol}}v\in(\R\setminus\{0\})^m\\ &\text{ with}\,\dim_\Q({\ensuremath{\boldsymbol}}v)\geq m-n\},\end{aligned}$$ and $$\begin{aligned} {\overline}{\ensuremath{\varepsilon}}_0(n,m)=\sup\{{\ensuremath{\varepsilon}}\geq0 \mid &\,{\bf (Li)}\text{ holds }\forall{\ensuremath{\boldsymbol}}v\in(\R\setminus\{0\})^m\text{ with}\,\dim_\Q({\ensuremath{\boldsymbol}}v)\geq m-n\}.\end{aligned}$$ The Lonely Runner Conjecture now claims that $${\ensuremath{\varepsilon}}_0(m-1,m)={\overline}{\ensuremath{\varepsilon}}_0(m-1,m)=1/(m+1),$$ and an extremal velocity vector would be given by ${\ensuremath{\boldsymbol}}v=(1,2,\ldots,m)^\intercal$. As we shall see in more detail below, in fact integral velocities alone determine ${\ensuremath{\varepsilon}}_0(m-1,m)$. On the other hand, due to a result by Czerwi[ń]{}ski [@czerwinski2012random], we know that random velocity vectors allow for ${\ensuremath{\varepsilon}}$ arbitrarily close to $1/2$ in Corollary \[corEquivLRP\]. Since random vectors have linearly independent entries over $\Q$, this is in the spirit of the trivial identity ${\ensuremath{\varepsilon}}_0(0,m)={\overline}{\ensuremath{\varepsilon}}_0(0,m)=1/2$. Using Theorem \[thmCoveringRadiusAsymptotics\] in the formulation  and similar ideas as in Propositions \[propPropertiesBarEpsFunction\] 1) and \[propPropertiesEpsFunction\] 1), we can interpolate between these two extremal situations. \[corLRPBounds\] For any $n,m\in\N$ with $n\leq m$, we have 1. ${\ensuremath{\varepsilon}}_0(m-1,m)\leq\ldots\leq{\ensuremath{\varepsilon}}_0(1,m)\leq{\ensuremath{\varepsilon}}_0(0,m)=\frac12$, 2. ${\ensuremath{\varepsilon}}_0(n,m)\geq{\ensuremath{\varepsilon}}(n,m)\geq\frac{m-O(n\log{n})}{2(m-n+1)}$, and 3. ${\overline}{\ensuremath{\varepsilon}}_0(n,m)\leq{\overline}{\ensuremath{\varepsilon}}_0(n,m+1)$. These inequalities say that the more rational dependencies we allow in the velocity vector ${\ensuremath{\boldsymbol}}v$ the smaller we have to choose ${\ensuremath{\varepsilon}}$ in Corollary \[corEquivLRP\]. In spite of these partial results, the main interest is of course to determine ${\ensuremath{\varepsilon}}_0(m-1,m)$. Wills [@willslrc] first stated that the problem reduces to the case of nonzero velocities $v_1,v_2,\ldots,v_m\in\Z{\ensuremath{\setminus}}{{\left\{{0}\right\}}}$ with $\gcd(v_1,\ldots,v_m)=1$, or equivalently, to vectors ${\ensuremath{\boldsymbol}}v\in(\R{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m$ with $\dim_\Q({\ensuremath{\boldsymbol}}v)=1$. A proof of such a statement though, appeared only in Bohman, Holzman & Kleitman [@sixrunners Lem. 8]. However, their lemma is not without restrictions; it states that if the conjecture is true for nonzero rational velocities in $m-1$ dimensions, then it is also true for irrational velocities in $m$ dimensions, thus leaving only the rational case to prove in $m$ dimensions. We also remark that this statement is included in Corollary \[corLRPBounds\] 3) as the special case $n=m-1$. Using the tools that we developed so far, we drop the dependence on lower dimensions, and show that the Lonely Runner Conjecture reduces to nonzero integer velocities in any dimension unconditionally. \[LRCintegerReduction\] For any $m\in\N$, we have $${\ensuremath{\varepsilon}}_0(m-1,m)=\sup{{\left\{{{\ensuremath{\varepsilon}}>0 \mid {\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\cap[{\ensuremath{\varepsilon}},1-{\ensuremath{\varepsilon}}]^m\neq{\ensuremath{\varnothing}}\text{ for all }{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in({\ensuremath{\mathbb{Z}}}{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m}\right\}}}.$$ Restricting the attention to vectors with integral coordinates clearly cannot decrease the considered supremum, showing that ${\ensuremath{\varepsilon}}_0(m-1,m)$ is less than or equal to the right hand side of the claimed identity. In order to show the reverse inequality, it suffices to show that for every ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}\in(\R{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m$ there is a ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}\in({\ensuremath{\mathbb{Z}}}{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m$, such that ${\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\beta}}}{\ensuremath{\subseteq}}{\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$. In order to see this, observe that by definition $\dim(V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})=\dim(V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\cap{\ensuremath{\mathbb{Z}}}^m)=\dim({\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}})$ and analogously $\dim(V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^{\perp}\cap{\ensuremath{\mathbb{Z}}}^m)=\dim(V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^{\perp})>0$. So there is a nonzero ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}\in V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^{\perp}\cap{\ensuremath{\mathbb{Z}}}^m$. We claim that in fact there is such a ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}$ in $({\ensuremath{\mathbb{Z}}}{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m$. Let us assume the contrary, that is, every ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}\in V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^{\perp}\cap{\ensuremath{\mathbb{Z}}}^m$ has at least one coordinate equal to zero. Suppose for the moment that there exist nonzero ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}=({\ensuremath{\beta}}_1,\dotsc,{\ensuremath{\beta}}_m)$ and ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}'=({\ensuremath{\beta}}'_1,\dotsc,{\ensuremath{\beta}}'_m)$ in $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^\perp\cap\Z^m$ that have no common zero coordinate, that is, for every $j\in[m]$, either ${\ensuremath{\beta}}_j\neq0$ or ${\ensuremath{\beta}}'_j\neq0$. It is easy to see that this means that the vector $${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}+(2{\lvert{\ensuremath{\beta}}_1\rvert}+\dotsb+2{\lvert{\ensuremath{\beta}}_m\rvert}){\ensuremath{\boldsymbol}}{\ensuremath{\beta}}'\in V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^{\perp}\cap{\ensuremath{\mathbb{Z}}}^m$$ has no coordinate equal to zero, contradicting the assumption. Hence, all ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}\in V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^{\perp}\cap{\ensuremath{\mathbb{Z}}}^m$ have a common zero coordinate, implying that a coordinate vector belongs to $V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, and thus ${\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}$ has a zero coordinate, a contradiction. Thus, there is a ${\ensuremath{\boldsymbol}}{\ensuremath{\beta}}\in V_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}^{\perp}\cap({\ensuremath{\mathbb{Z}}}{\ensuremath{\setminus}}{{\left\{{0}\right\}}})^m$, and by definition we have ${\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}\subseteq{\ensuremath{\Lambda}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\beta}}}$, implying that ${\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\beta}}}{\ensuremath{\subseteq}}{\ensuremath{\mathcal{E}}}_{{\ensuremath{\boldsymbol}}{\ensuremath{\alpha}}}$, as desired. So, without loss of generality let ${\ensuremath{\boldsymbol}}v=(v_1,\ldots,v_m)\in{\ensuremath{\mathbb{Z}}}^m_{>0}$ (replacing $v_i$ by $-v_i$ does not change the conclusion of the conjecture). In view of [**(L2)**]{}, we have ${\ensuremath{\varepsilon}}_0(m-1,m)=1/(m+1)$ if and only if the inequalities $$\begin{aligned} \frac{1}{m+1}&\leq {{\left\{{tv_i}\right\}}}\leq \frac{m}{m+1},\quad 1\leq i\leq m\label{eqnDiophApprDescr}\end{aligned}$$ hold for some $t\in\R$, which is Wills’ original formulation given in [@willslrc]. Note that these inequalities together with Corollary \[corLRPBounds\] 3) would imply the following more general statement. \[strongLRC\] For any $v_1,\dotsc,v_m\in\R_{>0}$ with $\dim_{\Q}(v_1,\dotsc,v_m)\geq d$, there is some $t\in \R$, such that $$\frac{1}{m+2-d}\leq {{\left\{{tv_i}\right\}}}\leq \frac{m+1-d}{m+2-d},\quad 1\leq i\leq m.$$ However, we want to take a more detailed look at condition [**(L4)**]{}, that is, $\left((1-2{\ensuremath{\varepsilon}})Z_{{\ensuremath{\boldsymbol}}v}+{\ensuremath{\boldsymbol}}c\right)\cap\Z^n\neq{\ensuremath{\varnothing}}$, where ${\ensuremath{\boldsymbol}}c=T({\ensuremath{\varepsilon}}{\ensuremath{\mathbf{1}}}_m | V_{{\ensuremath{\boldsymbol}}v})$. To this end, let $Z_{{\ensuremath{\boldsymbol}}v}=\sum_{i=1}^m[{\ensuremath{\boldsymbol}}0,{\ensuremath{\boldsymbol}}z_i]$, where ${\ensuremath{\boldsymbol}}z_i=T({\ensuremath{\boldsymbol}}e_i|V_{{\ensuremath{\boldsymbol}}v})\in\Z^{m-1}$ and where $T$ is the linear map from Subsection \[subsect\_zonotopes\_vo\]. Writing ${\ensuremath{\boldsymbol}}x=\frac12\sum_{i=1}^m{\ensuremath{\boldsymbol}}z_i$ for the center of $Z_{{\ensuremath{\boldsymbol}}v}$ and putting ${\ensuremath{\varepsilon}}=1/(m+1)$, we get $$\begin{aligned} (1-2{\ensuremath{\varepsilon}})Z_{{\ensuremath{\boldsymbol}}v}+{\ensuremath{\boldsymbol}}c&=\frac{m-1}{m+1}(Z_{{\ensuremath{\boldsymbol}}v}-{\ensuremath{\boldsymbol}}x)+\frac{m-1}{m+1}{\ensuremath{\boldsymbol}}x+\frac1{m+1}\sum_{i=1}^mT({\ensuremath{\boldsymbol}}e_i|V_{{\ensuremath{\boldsymbol}}v})\\ &=\frac{m-1}{m+1}(Z_{{\ensuremath{\boldsymbol}}v}-{\ensuremath{\boldsymbol}}x)+{\ensuremath{\boldsymbol}}x.\end{aligned}$$ With the analysis above, putting $n=m-1$, we obtain the following reformulation of the lonely runner conjecture. \[conjZonotopalLRC\] Let $Z$ be a zonotope generated by $n+1$ vectors of ${\ensuremath{\mathbb{Z}}}^n$ in LGP, and let ${\ensuremath{\boldsymbol}}{x}$ be the center of $Z$. Then, $$\left({\ensuremath{\boldsymbol}}{x}+\frac{n}{n+2}(Z-{\ensuremath{\boldsymbol}}{x})\right)\cap{\ensuremath{\mathbb{Z}}}^n\neq {\ensuremath{\varnothing}}.$$ In the case that ${\ensuremath{\boldsymbol}}x\in{\ensuremath{\mathbb{Z}}}^n$, there would be nothing to prove, so we can assume otherwise. Then, ${\ensuremath{\mathbb{Z}}}^n$ and ${\ensuremath{\boldsymbol}}x$ generate a lattice $\Lambda=\Z^n\cup({\ensuremath{\boldsymbol}}x+\Z^n)$. Shifting everything by ${\ensuremath{\boldsymbol}}x$, so that $Z$ is $o$-symmetric with respect to the origin, our desired (nonempty) intersection becomes $$\frac{n}{n+2}Z\cap(\Lambda{\ensuremath{\setminus}}{\ensuremath{\mathbb{Z}}}^n).$$ In other words, we wish to prove that if we dilate $Z$ by a factor of at most $n/(n+2)$, we get a nontrivial point of $\Lambda$ not contained in ${\ensuremath{\mathbb{Z}}}^n$. This alludes to the notion of the *first restricted successive minimum*, as defined in [@henkthiel]: $${\ensuremath{\lambda}}_1(Z,\Lambda{\ensuremath{\setminus}}{\ensuremath{\mathbb{Z}}}^n)=\min{{\left\{{{\ensuremath{\lambda}}\geq0 : {\ensuremath{\lambda}}Z\cap (\Lambda{\ensuremath{\setminus}}{\ensuremath{\mathbb{Z}}}^n)\neq{\ensuremath{\varnothing}}}\right\}}}.$$ Hence, yet another reformulation of Conjecture \[conjZonotopalLRC\] with respect to this definition is the following: [*Let $Z{\ensuremath{\subseteq}}\R^n$ be a zonotope generated by $n+1$ vectors of ${\ensuremath{\mathbb{Z}}}^n$ in LGP, and which is translated as to be $o$-symmetric. Let $\Lambda$ be a lattice such that $\Lambda{\ensuremath{\setminus}}{\ensuremath{\mathbb{Z}}}^n$ is a translate of $\Z^n$. Then $${\ensuremath{\lambda}}_1(Z,\Lambda{\ensuremath{\setminus}}{\ensuremath{\mathbb{Z}}}^n)\leq\frac{n}{n+2}.$$*]{} In view of Schoenberg’s result on the general view-obstruction problem (see Theorem \[Sch\]), we know that an upper bound of $n/(n+1)$ holds true. It is quite interesting that this has not been significantly improved; the interested reader may consult [@perarnauserra2016correlation] for an informative discussion of this matter. So, even if it is difficult to prove the Lonely Runner Conjecture in the geometric setting, it is reasonable to ask whether the intersection $$\frac{n+2-c}{n+2}Z\cap (\Lambda{\ensuremath{\setminus}}{\ensuremath{\mathbb{Z}}}^n)$$ is nonempty for some absolute constant $c>1$. Any such result would be a significant advance to the problem. [^1]: The first author is supported by the Freie Universität Berlin within the Excellence Initiative of the German Research Foundation. [^2]: The second author is supported by a Postdoctoral Fellowship from the Alexander von Humboldt Foundation. [^3]: This case corresponds to seven runners, as the seventh runner is considered to have velocity equal to zero, being fixed at the starting point.
{ "pile_set_name": "ArXiv" }
--- author: - | D. Cremades, L. E. Ibáñez and F. Marchesano\ Departamento de Física Teórica C-XI and Instituto de Física Teórica C-XVI,\ Universidad Autónoma de Madrid, Cantoblanco, 28049 Madrid, Spain. title: Intersecting Brane Models of Particle Physics and the Higgs Mechanism --- Introduction ============ Chirality is probably the deepest property of the standard model (SM) of particle physics. It thus seems that, in trying to build a string theory description of the SM, the first thing we have to obtain is massless chiral fermions. There is a number of known ways in order to achieve chirality in string theory models. From the point of view of explicit D-brane model building there are however only a few known ways. One of the simplest is to locate D-branes (e.g., D3-branes) at some (e.g., orbifold or conifold) singularity in transverse space [@dm]. Explicit models with branes at singularities have been built [@aiqu; @evenmore] with three generations and massless spectrum close to that of the SM or some simple left-right symmetric extension (see also [@bjl; @bailin; @gerardo]). Another alternative way in order to get massless chiral fermions in D-brane models is intersecting branes [@bdl]. Under certain conditions, D-branes intersecting at angles give rise to massless chiral fermions localized at the intersections. Explicit D-brane models with realistic three-generation particle spectrum lying at intersecting branes have been constructed in the last couple of years [@bgkl; @afiru; @afiru2; @bkl; @imr; @bklo; @csu; @cim1; @pheno; @blumy] (see also [@cosmo]). One of the nice features of this type of constructions is that family replication appears naturally as a consequence of the fact that in a compact space branes may intersect a multiple number of times. Another attractive property of this scenario is that quarks and leptons corresponding to different generations are localized at different points in the transverse compact space. It has been suggested [@afiru2; @imr] that this may provide a geometrical understanding of the observed hierarchy of quark and lepton masses. In ref.[@imr] a particularly interesting class of models was found in which the massless chiral fermion spectrum is identical to that of the SM of particle physics. They were obtained from sets of intersecting D6-brane wrapping an (orientifolded) six-torus [@bgkl]. These models are non-supersymmetric but it has recently been found [@cim1] that analogous models with one $\cn = 1$ supersymmetry at each brane intersection may be obtained by appropriately varying the geometry (complex structure) of the torus. In these models each intersection respects in general a [*different*]{} $\cn = 1$ supersymmetry so that the model is globally non-supersymmetric but in some sense locally supersymmetric. These type of models are called in [@cim1] [*quasi-supersymmetric*]{} (Q-SUSY) and have the property that quantum corrections to scalar masses appear only at two-loops. The different $\cn = 1$ structure of each intersection in this class of models may be represented in terms of [*SUSY quivers*]{}. This Q-SUSY property may be of phenomenological interest in order to stabilize the hierarchy between the weak scale and a fundamental scale of order 10-100 TeV present in low string scale models. The present article has two main purposes: [*i)*]{} We construct explicit D6-brane Q-SUSY models with three quark-lepton generations and study some of their properties. All of these models have four stacks of D6-branes : the [*baryonic*]{}, [*leptonic*]{}, [*left*]{} and [*right*]{} stacks and either the SM group or a slight generalization with an extra $U(1)$. We show that there are just four classes of quivers (see fig. \[quivers\]) with four stacks of branes and no $SU(3)$ anomalies yielding realistic models. We call them the triangle, linear, square and rombic quivers. Each of these classes of quivers differ on the $\cn = 1$ SUSY preserved by the quark, lepton and Higgs sector. In some models ([*triangle quiver*]{}) all of them preserve the same $\cn = 1$ SUSY whereas in others e.g., left-handed fermions preserve one supersymmetry and right-handed fermions a different one ([*linear quiver*]{}). The ([*square quiver*]{}) class corresponds to Q-SUSY versions of the models of [@imr] and have the SM fermion spectrum. They have the property that left-handed quarks, right-handed quarks, left-handed leptons and right-handed leptons have each a different $\cn = 1$ SUSY inside a $\cn = 4$ living in the bulk. Finally, models constructed from the ([*rombic quiver*]{}) have the property that all quarks and the Higgs set respect the same $\cn = 1$ supersymmetry whereas the leptonic sector does not. The example provided has a left-right symmetric gauge group. Whereas in the linear and square quiver models the Higgs sector is non-SUSY, it is so in the other two, providing explicit examples of models with weak scale stabilization. There are a couple of results which look fairly general: - Imposing supersymmetry at all the intersections implies necessarily the presence of a very definite extra $U(1)$, the $U(1)_{B-L}$ familiar from left-right symmetric models. - If in addition we insist on some left- and right-handed fermions to respect the [*same*]{} $\cn = 1$ SUSY (as happens e.g. in the triangular and rombic quivers), the brane configurations are forced to have intersections giving rise to SM Higgs multiplets [^1], providing for a rationale for the very existence of the SM Higgs sector. We also show how slight variations of torus moduli give rise to Fayet-Iliopoulos (FI) terms for the $U(1)$’s in the theory. These FI-terms are only present for the subset of $U(1)$’s which become massive by combining with antisymmetric $B_2$ fields of the closed string sector of the theory. Since those $U(1)$’s are massive, the scalar D-term masses look like explicit soft SUSY-breaking masses for the sparticles charged under those $U(1)$’s. We finally analyze the structure of gauge coupling constants in this class of models. For these intersecting brane models gauge couplings do not unify at the string scale. Rather they are inversely proportional to the volume wrapped by each brane. In the case of Q-SUSY models those volumes get a particularly simple expression in terms of the torus moduli, and we give explicit formulae for them. [*ii)* ]{} We describe how the SM Higgs mechanism in intersecting brane world models has a nice geometrical interpretation as a brane recombination process, in which the branes giving rise to $U(2)_L\times U(1)$ recombine into a single brane related to $U(1)_{em}$. In the SM Higgs mechanism the rank of the gauge group is reduced. The stringy counterpart of this rank-reduction is brane recombination [^2]. The chiral fermion spectrum in these models is determined by the intersection numbers $I_{ab}$ which are topological in character. As branes recombine, $|I_{ab}|$ decreases, signaling that some chiral fermions become massive. This is the stringy version of the fermions getting masses from Yukawa couplings after the Higgs mechanism takes place. We exemplify this brane recombination interpretation of the Higgs mechanism in some of the specific examples introduced in the paper, although the general physics applies to any intersecting brane model. As we said, having some $\cn = 1$ SUSY at all brane intersections implies generically the presence of a $B-L$ massless generator in the spectrum. We discuss how this symmetry may be Higgssed away in terms of a brane recombination process in which the [*leptonic*]{} and [*right*]{} branes recombine into a single brane. That recombination process gives masses to the right-handed neutrinos and we argue that, after the SM Higgs/recombination mechanism takes place, the left-handed neutrinos get at some level Majorana masses. The structure of this paper is as follows. In the next two sections we give short introductions to both the D6-brane toroidal models introduced in [@imr] and to the concept of quasi-supersymmetric models of ref.[@cim1]. In Section 4 we construct several three generation models with SM gauge group (or some simple extension) and with some quasi-supersymmetry. We analyze SUSY-breaking effects from FI-terms for some of those models in Section 5 whereas we provide formulae for the gauge coupling constants in Section 6. In Section 7 we interpret the SM Higgs mechanism in terms of recombination of intersecting branes. Some final comments and general conclusions are presented in Section 8. A couple of appendices concerning the presence of extra $U(1)$’s in Q-SUSY models and models with D6-branes wrapping non-factorizable cycles on the tori are provided. The Standard Model at intersecting branes revisited =================================================== In this section we summarize the construction of the Standard Model fermionic spectrum arising from intersecting brane worlds, as presented in [@imr]. We refer the reader to this paper for details regarding this construction. We will consider Type IIA string theory compactified on a factorized 6-torus $T^2 \times T^2 \times T^2$. In this setup, we introduce sets of D6-branes with their 7-dimensional volume containing four-dimensional Minkowski space and wrapping 3-cycles $[\Pi]$ of $T^6$ [@bgkl; @afiru; @bkl] (for related constructions see also [@inter1]). We will further assume that these 3-cycles can be factorized as three 1-cycles, each of them wrapping on a different $T^2$. We denote by $(n_a^i,m_a^i)$, $i = 1,2,3$ the wrapping numbers of each $D6_a$-brane, on the $i^{th}$ torus, $n_a^i$ ($m_a^i$) being the number of times the brane is wrapping around the basis vector $e_1^i$ ($e_2^i$) defining the lattice of the $i^{th}$ torus, as depicted in figure \[compact2\]. Actually, we will study an orientifolded version of the theory obtained by modding out by $\Omega \car$ [@bgkl], where $\Omega$ is the worldsheet parity operator and $\car = \car_{(4)} \car_{(6)} \car_{(8)}$ is a reflection operator with respect to the real axis of each internal complex dimension $Z_j = X_{2j+2} + iX_{2j+3}$, $j = 1,2,3$ (that is, $\car_{(4)}: Z_1 \mapsto \bar Z_1$, etc.). Since we must have $\Om \car$ symmetry in our models, for each $D6_a$ brane we introduce we must also add its mirror image $\Om \car D6_a$ or $D6_{a^*}$, whose geometrical locus will be determined by a reflection on the real axes $X_4, X_6, X_8$. Being identified, both branes will give rise to the same unitary gauge group. This symmetry under $\Om \car$ does also imply that we are allowed to consider only some specific choices of $T^2$. We may either have rectangular tori (as the first two tori in fig. \[compact2\]) or tilted tori (as the third torus in fig. \[compact2\]). When we have a tilted torus, we can easily describe our configurations in terms of fractional wrapping numbers, where the $m$’s may take $\IZ/2$ values (see [@bkl; @imr] and Appendix I for more detailed discussions). The number of times two branes $D6_a$ and $D6_b$ intersect in $T^6$ is a topological quantity known as the intersection number. When dealing with factorizable branes it can be easily expressed as I\_[ab]{} =  (n\_a\^1m\_b\^1-m\_a\^1n\_b\^1)(n\_a\^2m\_b\^2-m\_a\^2n\_b\^2)(n\_a\^3m\_b\^3-m\_a\^3n\_b\^3). \[internumber\] Open strings stretching around the intersections give rise to chiral fermions in bifundamental representations $(N_a,{\bar N}_b)$ or $(N_a,{N}_b)$ under the gauge group of the two branes $ U(N_a)\times U(N_b)$. Thus, these configurations yield $I_{ab}$ copies of the same bifundamental representation. In [@imr] it was shown how one can build a configuration giving just three generations of a $SU(3)\times SU(2)\times U(1)_Y$ gauge group. To achieve this construction one must consider four stacks of D6-branes whose multiplicities and associated gauge group is shown in table \[SMbranes\]. Now, given this brane content, it is relatively easy to achieve the desired SM spectrum as arising from massless fermions living at brane intersections. It is enough to select branes’ wrapping numbers $(n_a^i,m_a^i)$ in such a way that the intersection numbers $I_{ij}$, $i,j = a,b,c,d$ are given by [@imr] [lcl]{} I\_[ab]{} =  1, & & I\_[ab\*]{} = 2,\ I\_[ac]{} =  -3, & & I\_[ac\*]{} = -3,\ I\_[bd]{} =  -3, & & I\_[bd\*]{} = 0,\ I\_[cd]{} =  3, & & I\_[cd\*]{} = -3, \[intersec2\] all other intersections vanishing. Here a negative number denotes that the corresponding fermions should have opposite chirality to those with positive intersection number. The massless fermionic spectrum arising from (\[intersec2\]) is shown in table \[tabpssm\], as well as the charges with respect to the four $U(1)$’s. Note in this respect that eventually only one $U(1)$ (the hypercharge) remains massless, the other three become massive by swallowing certain closed RR fields, as discussed below. In [@imr] a general class of solutions was given for the wrapping numbers $(n_a^i,m_a^i)$ giving rise to a SM spectrum. These are shown in table \[solution\]. In this table we have several discrete parameters. First we consider $\beta^i =1 - b^i$, with $b^i = 0, 1/2$ being the T-dual NS B-background field discussed in [@bkl] (see also [@Bachas; @bfield; @fluxes] for previous related discussions). As shown there, the addition of this background is required in order to get and odd number of quark-lepton generations. From the point of view of branes at angles $\beta^i = 1$ stands for a rectangular lattice for the $i^{th}$ torus, whereas $\beta^i = 1/2$ describes a tilted lattice allowed by the $\Om\car$ symmetry. Notice that in the third torus one always has $\beta^3 = 1/2$, as in figure \[compact2\]. We also have two phases $\epsilon, \tilde \eps = \pm 1$ and the parameter $\rho$ which can only take the values $\rho = 1, 1/3$. Furthermore, each of these families of D6-brane configurations depend on four integers ($n_a^2,n_b^1,n_c^1$ and $n_d^2$). Any of these choices lead exactly to the same massless fermion spectrum of table \[tabpssm\]. The wrapping numbers of the complete set of D6-branes have to verify the RR tadpole cancellation conditions [@bgkl] [c]{} \_a N\_a n\_a\^1n\_a\^2n\_a\^3 = 16\ \_a N\_a n\_a\^1m\_a\^2m\_a\^3 = 0\ \_a N\_a m\_a\^1n\_a\^2m\_a\^3 = 0\ \_a N\_a m\_a\^1m\_a\^2n\_a\^3 = 0 \[tadpoleO6\] which just state that the total RR-charge of the configuration has to vanish. The orientifold modding leads to the presence of 8$\b^1\b^2\b^3$ orientifold planes wrapping the cycle $(1/\b^1,0)(1/\b^2,0)(1/\b^3,0)$, each with net RR charge -2 (compared to the charge of a pair of mirror D6-branes), as the first condition shows. These conditions automatically guarantee cancellation of chiral anomalies. In this class of models the last three conditions are automatically verified whereas the first requires:  +   +   =   16  . \[tadsm\] Note however that one can always relax this constraint by adding extra D6-branes with no intersection with the previous ones and not contributing to the rest of the tadpole conditions. In particular, any set of branes with wrapping numbers of the form N\_h (1/\_1,0)( 1/\_2,0)(n\_h\^3,m\_h\^3) \[hidden\] does not intersect the branes from table \[SMbranes\] whereas contributes with $N_hn_h^3/(\beta_1\beta_2)$ to the first condition (without affecting the other three). We will call such D6-brane a [*hidden brane*]{}. As advanced in [@cim1], we can implement this same idea in a more general situation. Let us consider an arbitrary configuration of D6-branes whose low energy chiral spectrum has neither chiral nor mixed anomalies. Since tadpoles imply anomaly cancellation but not the other way round, it could happen that the brane content giving rise to such anomaly-free spectrum did not satisfy tadpoles. We must then ‘complete’ our model by simply adding the necessary brane content to cancel RR charges. In particular, we can complete our brane configuration by a single brane, let us call it $H$-brane. Unlike in (\[hidden\]), in general this $H$-brane will not have a simple expression, but will be wrapping a non-factorizable cycle [^3] $[\Pi_H] $ of $T^6$. The important point to notice is that, as long as the low-energy fermion spectrum is anomaly-free, this $H$-brane will have no net intersection with any brane belonging to our initial configuration, just as in our previous example. Thus, it will constitute a hidden sector, and its presence will not imply the existence of new chiral massless fermions in our low energy spectrum. Alternatively, and following the recent proposal in [@uranga] (see also [@flujos]), we could equally well turn on an explicit NS-NS background flux $H_{NS}$ in our configuration. This flux will have an associated homology class $[H_{NS}]$, behaving as an RR source just as a D6-brane in this same homology class would do. Thus, we can consider configurations where tadpoles cancel by a combination of D6-brane content and NS-NS flux, the presence of the latter not implying new chiral spectrum. Furthermore, the presence of these fluxes may relax even more the model-building constraints, since the addition of $H_{NS}$ can compensate the anomaly generated by an anomalous $D = 4$ chiral spectrum. The reason for this is that the presence of $H_{NS}$ will induce a Wess-Zumino term in the low energy Lagrangian, which will cancel the $SU(N_a)^3$ and $U(1)_a-SU(N_b)^2$ anomaly developed from the naive ‘fermion content’ point of view (see [@uranga]). Unlike the addition of some $H$-brane to complete an anomalous configuration, such background flux will not imply new chiral content to our spectrum[^4]. We will use the addition of both RR sources in some of the explicit models built in Section 4. However, in order to limit the arbitrariness, we will only consider the possible presence of H-flux induced Wess-Zumino terms for anomalous massive $U(1)$’s so that none of the anomalies of the SM gauge groups have to be canceled via Wess-Zumino terms. One of the most interesting aspects of this class of theories is the structure of Abelian gauge symmetries. The four $U(1)$ symmetries $Q_a$, $Q_b$, $Q_c$ and $Q_d$ have clear interpretation in terms of global symmetries of the standard model. Indeed, $Q_a$ is $3B$, $B$ being the baryon number, and $Q_d$ is nothing but lepton number [^5] . Concerning $Q_c$, it is twice $I_R$, the third component of right-handed weak isospin familiar from left-right symmetric models. Finally $Q_b$ has the properties of a Peccei-Quinn symmetry, having mixed $SU(3)$ anomalies. Two out of the four $U(1)$’s have triangle anomalies which are canceled by a generalized Green-Schwarz mechanism [@imr]. The anomalous symmetries correspond to generators $Q_b$ and $3Q_a+Q_d$, the latter being identified with $(9B+L)$. For a general brane configuration the anomaly cancellation mechanism goes as follows. There are four closed string antisymmetric fields $B_2^{I}$, $I = 0,1,2,3$ which couple to the four $U(1)_a$ fields associated to each set of $N_a$ D6-branes in the form: \_[M\_4]{} \_a N\_a( m\^1\_a m\^2\_a m\^3\_a  B\_2\^0   +  \_I n\^J\_a n\^K\_a m\^I\_a   B\_2\^I  ) F\_a , I=J=K \[bf\] On the other hand, the Poincare duals of those antisymmetric fields, denoted $C^I$, $I=0,1,2,3$ couple to both Abelian and non-Abelian fields $F_b$ as follows: \_[M\_4]{} \_b ( n\^1\_b n\^2\_b n\^3\_b C\^0  +  \_I n\^I\_b m\^J\_b m\^K\_b C\^I ) F\_bF\_b  . \[cff\] The combined effect of these two couplings cancel the residual $U(1)_a\times U(N_b)^2$ triangle anomalies by the tree level exchange of the four RR fields. At the same time the $B_2^I\wedge F$ couplings in (\[bf\]) give masses to the following four linear combinations of Abelian fields [c]{} B\_2\^0  :\_a N\_a m\^1\_a m\^2\_a m\^3\_a F\_a,\ B\_2\^1  :\_a N\_a m\^1\_a n\^2\_a n\^3\_a F\_a,\ B\_2\^2  :\_a N\_a n\^1\_a m\^2\_a n\^3\_a F\_a,\ B\_2\^3  :\_a N\_a n\^1\_a n\^2\_a m\^3\_a F\_a. \[caplillos\] Since in the specific models we are constructing at least one of the $m_a$’s of each brane vanishes, the first linear combination is trivially vanishing and there is always a massless $U(1)$. The three linear combinations which are massive are : [c]{} Q\_[b]{}\ 3Q\_a + Q\_d\ Q\_m = ( [3\^2 n\_a\^2]{} Q\_a + 6\^1 n\_b\^1 Q\_[b]{} - 2\^1n\_c\^1 Q\_c - 3\^2 n\_d\^2 Q\_d - 2N\_h m\_h\^3 Q\_h ) \[massive\] where for the sake of generality we have included the effect of a set of $N_h$ parallel “hidden” branes as discussed above. The first two massive generators are model independent and correspond to the anomalous $U(1)$’s, whereas $Q_m$ is model-dependent and anomaly-free. The massless $U(1)$’s are: [*i) $m_h^3\ =\ 0$*]{} In this case the presence of a hidden brane does not affect the form of the massless linear combinations which are given by: Q\_0 = [16]{} Q\_a + [r2]{} Q\_c - [12]{} Q\_d , r = [\^2 2n\_c\^1\^1]{}(n\_a\^2+3n\_d\^2) \[hyper\] and the hidden brane charge $Q_h$. Note that for $r=-1$ the above generator is the standard hypercharge generator, as discussed in [@imr]. In this case only this standard hypercharge generator remains massless [^6]. [*ii)  $m_h^3\ \neq \ 0$*]{} In this case the hidden brane affects the form of the visible linear combination. The massless linear combinations are: & & Q\_1 = [16]{} Q\_a - [12]{} Q\_c - [12]{} Q\_d + (1+r)[[n\_c\^1\^1]{} ]{}Q\_h\ & & Q\_2 =  [13]{} Q\_a - Q\_d + [[rn\_c\^1\^1]{}]{}Q\_h \[uunose\] Note that the first of these charges couples like standard hypercharge to the SM particles, whereas the second one couples like $B-L$. Thus, in this more general case two massles $U(1)$’s coupling to quarks and leptons remain in the massless spectrum. The general solutions yielding the SM spectrum have the following geometrical properties. The Baryonic ($a$) and Leptonic ($d$) stacks are parallel in the first complex dimension and that is why they do not intersect. Thus, no lepto-quarks fields appear in our massless spectrum. On the other hand, Left ($b$) and Right ($c$) stacks are parallel in the second complex dimension, again not intersecting. Something similar happens for each brane $i$ and its mirror $i^*$, $i = a,b,c,d$. Strings exchanged between stacks $b$ and $c$ have the quantum numbers of SM Higgs (and Higgsino) fields and eventually we would be interested in some of these scalar states to be relatively light, so as to play the role of Higgs fields. So we will asume that the distance between $a$ and $d$ branes is bigger than that between branes $b$ and $c$, so that the latter can provide us with a Higgs system, as we will describe in more detail in section 7. Note that we will consider the different distances between branes to be fixed quantities. Hence, they will play the role of external parameters of our models, very much like the geometrical moduli. Notice that the whole of the previous SM construction has been achieved by using factorizable D6-branes, that is, branes whose geometrical locus can be described as product of 1-cycles, each wrapping a different $T^2$. This kind of construction is preferred from a model-building point of view, since it has a particularly simple associated geometry. This allows us to easily compute some phenomenologically important quantities as, for instance, scalar masses at intersections. However, we could have also constructed a model in a more general set-up, where the building blocks of our configuration were D6-branes wrapping [*general*]{} 3-cycles of our six-torus. In this way, we could have equally well realised a D6-brane tadpole-free configuration giving rise to a SM spectrum, the factorizable models just presented being a particular subfamily of the latter. For simplicity and better visualization of our constructions, we will give as examples models where the relevant physics arises at factorizable D6-branes configurations, so as to extract the phenomenologically interesting quantities more straightfordwardly. However, it is important to notice that reached this point it is not possible to ignore non-factorizable D6-branes anymore. In particular, when studying the recombination process of two branes into a third one (corresponding to some Higgs mechanism) such non-factorizable branes will naturally arise. As we will see, this is due to the fact that conservation of RR charge imposes that this third brane should wrap a 3-cycle which is the sum of the two former 3-cycles. In general, this will yield a non-factorizable brane. In order to appropiately describe these less intuitive objects, we have made use of a “$q$-basis” formalism, which is described in Appendix I. For some results regarding the geometry of branes wrapping general cycles see also [@bbh]. Let us also emphasize that in the present paper we will be dealing mostly with the open string sector of the theory, which is the one which may give rise to the SM physics. We will not discuss here the closed string NS potential and its associated NS tadpoles. We will rather take the value of NS geometric moduli as well as brane positions as frozen external parameters defining the geometry of the configuration. Some work on the NS-tadpole structure of this class of theories may be found in, e.g., ref.[@bklo; @cim1; @blumy]. Intersecting D6-branes and Supersymmetry: SUSY & Q-SUSY models ============================================================== An interesting question is whether one can construct intersecting brane models analogous to those discussed in the previous section and with $\cn = 1$ supersymmetry. As discussed in refs.[@bgkl; @cim1], the answer to this question is no, if we restrict ourself to purely toroidal models (no orbifolds) with factorized D6-branes and impose RR tadpole cancellation conditions. Models with $\cn = 1$ SUSY may be constructed if an additional $\IZ_2\times \IZ_2$ orbifolding is perfomed [@csu], but at the cost of losing the simplicity of the spectrum, as exotic chiral particles beyond the SM appear. On the other hand it was shown in [@cim1] the possibility of having all brane intersections preserving [*some*]{} unbroken $\cn = 1$ SUSY, although not necessarily the same one. These configurations were called quasi-supersymmetric, (Q-SUSY). Roughly speaking, a Q-SUSY field theory consists of different subsectors each preserving at least $\cn = 1$ supersymmetry but, not being the same supersymmetry for every sector, the system as a whole has $\cn = 0$. Thus, we expect corrections that spoil the boson-fermion mass degeneracy appearing in a truly supersymmetric theory, this effects appearing only at two-loops in perturbation theory. There might also be some [*massive*]{} sectors not respecting any supersymmetry at all. In any case, the radiative corrections will be fairly supressed with respect to an ordinary $\cn = 0$ field theory. In particular, scalars will have a two-loop protection against quadratic divergences. This may be of interest when dealing with scenarios where the fundamental scale of physics is two or three orders of magnitude above the electroweak scale [@cim1]. A two-loop protection would be enough to understand a 2-3 orders of magnitude hierarchy between the weak scale and a fundamental scale of order 10-100 TeV. Let us describe the D-brane setup that allows for an explicit realization of this Q-SUSY structure. Just as in the construction of the Standard Model presented in last section, we will be dealing with Type IIA D6-branes wrapping factorized 3-cycles in $T^2 \t T^2 \t T^2$. Any factorized D6-brane constitutes a $\oh$BPS state that preserves half of the supersymmetries coming from a plain toroidal compactification. This will be reflected in the $D6_aD6_a$ sector of the theory (that is, strings beginning and ending on the same $D6_a$-brane), yielding $D = 4$ $\cn = 4$ super Yang-Mills as its low energy spectrum. When considering a pair of such factorized branes, say $a$ and $b$, each of them will preserve some $\cn = 4$ subalgebra of the whole $\cn = 8$ bulk superalgebra. Whether these two different subalgebras do have some overlap or not will determine if the $D6_aD6_b$ sector is also supersymmetric (see figure \[venn\]). In general, the answer will be positive whenever these two branes are related by a $SU(3)$ rotation in compact dimensions [@bdl]. When dealing with D6-branes at angles, this can be easily computed in terms of the spectrum living at the $D6_aD6_b$ sector. Whereas the Ramond sector always yields a massless chiral fermion, the lightest states coming from the Neveu-Schwarz sector will be four scalars with masses [@afiru] [c]{}\ 12(-|\_[ab]{}\^1|+|\_[ab]{}\^2|+|\_[ab]{}\^3|)\ 12(|\_[ab]{}\^1|-|\_[ab]{}\^2|+|\_[ab]{}\^3|)\ 12(|\_[ab]{}\^1|+|\_[ab]{}\^2|-|\_[ab]{}\^3|)\ 1-12(|\_[ab]{}\^1|+|\_[ab]{}\^2|+|\_[ab]{}\^3|), \[scalars\] where $\vt_{ab}^i$ represent the angles between branes $a$ and $b$ in the $i^{th}$ torus (see fig. \[compact2\]). The number of SUSY’s shared by sectors $D6_aD6_a$ and $D6_bD6_b$ will correspond to the number of massless scalars in (\[scalars\]). Indeed, if for instance $\vt_{ab}^1 = \vt_{ab}^2 + \vt_{ab}^3$, then the first of such scalars will be massless, and the $D6_aD6_b$ sector will consist of $|I_{ab}|$ $\cn = 1$ chiral multiplets. In an orientifold compactification, however, a generic D6-brane $a$ cannot exists on its own, but will be accompanied by its mirror image $a^*$. In order to achieve our Q-SUSY structure, we will require any brane and its mirror to preserve at least one common SUSY. In fact, we will require them to share a $\cn = 2$ subalgebra [^7]. As discussed in [@cim1], there exist six types of such branes. In order to describe them, let us define a [*twist*]{} vector $v_\a = (\th_\a^1,\th_\a^2,\th_\a^3)$ whose three entries contain the relative angles $\th_\a^i$ between a D6-brane $\a$ and the horizontal axis of the $i^{th}$ torus (see figure \[twist\]). In terms of this twist vector, the six different types of branes can be listed as [rcc]{} [type of brane]{} & [twist vector]{} & [SUSY]{}\ a\_1 [branes]{} : & v\_[a1]{} = (0,\_[a1]{},\_[a1]{}) & r\_2, r\_3\ a\_2 [branes]{} : & v\_[a2]{} = (0,\_[a2]{},-\_[a2]{}) & r\_1, r\_4\ b\_1 [branes]{} : & v\_[b1]{} = (\_[b1]{},0,\_[b1]{}) & r\_1, r\_3\ b\_2 [branes]{} : & v\_[b2]{} = (\_[b2]{},0,-\_[b2]{}) & r\_2, r\_4\ c\_1 [branes]{} : & v\_[c1]{} = (\_[c1]{},\_[c1]{},0) & r\_1, r\_2\ c\_2 [branes]{} : & v\_[c2]{} = (\_[c2]{},-\_[c2]{},0) & r\_3, r\_4 \[vectors\] where we have suppressed the upper index $i$. We have named the generators of each $\cn = 1$ algebra by $r_j$, $j = 1,2,3,4$, just in order to keep track of the supersymmetries present on each sector (for a more detailed description see [@cim1]). Each type of brane in (\[vectors\]) shares a different $\cn = 2$ subalgebra with their mirror brane, which can be represented by a pair of generators [^8]. Notice that the twist vector of a mirror brane $\a^*$ will be given by $v_{\a^*} = - v_\a$, whereas the angles between two branes $\a$ and $\b$ are the components of $v_{\a\b} := v_\b - v_\a$. When considering a configuration where several of such $\cn = 2$ branes appear, we must also consider sectors corresponding to D6-brane intersections. The full Q-SUSY structure can be encoded in a hexagonal quiver-like diagram, shown in figure \[hexagon\]. Each node of this diagram represents one of the branes in (\[vectors\]), and at the same time its mirror image. Notice that gauge groups at intersecting D6-branes models always arise from $D6_\a D6_\a$ sectors, hence they will be localized at these nodes. Just as in the toroidal case, each gauge group will correspond to a $\cn = 4$ subsector of the theory, while the matter content living on the $D6_\a D6_{\a^*}$ sector presents a non-chiral, generically massive $\cn = 2$ spectrum. Each of the branes in (\[vectors\]) will generically intersect with four other types of brane. Such intersection is represented in our hexagonal quiver by a link joining two nodes. In general, a non-vanishing intersection number $I_{\a\b}$ does signal that some chiral spectrum lives at the $D6_{\a}D6_{\b}$ sector. It can easily be seen that any of these sectors will also preserve some $\cn = 1$ SUSY, yielding chiral multiplets transforming in bifundamentals. Let us take, for instance, the link joining $a_2$ with $b_2$. Each node, as a combined system of a brane and its mirror, preserves two supersymmetries, having $r_4$ in common. Thus, the link between these two nodes will contain $I_{a_2b_2} \ (N_{a_2},\bar{N}_{b_2})\ +\ I_{a_2b_2^*} \ (N_{a_2},N_{b_2})$ chiral multiplets under the supersymmetry generator $r_4$. Finally, branes corresponding to opposite nodes (as for instance $a_1$ and $a_2$) do never intersect, since they are parallel in one of the tori. This $D_{\a_1}D_{\a_2}$ sector will contain a non-chiral, generically massive, $\cn = 0 $ spectrum. We thus see that any low energy effective field theory coming from a D-brane configuration whose building blocks belong to (\[vectors\]) will yield a Q-SUSY field theory. In particular, the general hexagonal construction depicted in fig. \[hexagon\] will contain as massless sectors $\cn = 4$ vector multiplets and $\cn = 1$ chiral multiplets, whereas some non-chiral $\cn = 2$ and $\cn = 0$ massive sectors may also arise. As we mentioned at the beginning of this section, this will protect scalar masses from one-loop corrections, the first non-vanishing contributions coming from two-loops in perturbation theory [@cim1]. There are contributions coming both from light particle exchange (fig.(\[loop5\]-a)) and heavy non-SUSY particle exchange (fig.(\[loop5\]-b)). All will give corrections to scalar masses of order $\alpha_i/(4\pi)M_s$. For the $\cn = 4$ gauginos the only source of masses will be the one-loop exchange of massive $\cn = 0$ sectors of the theory (see figure (\[loop5\]-c)). These general considerations will equally well apply to any subquiver of the hexagonal construction just presented. In particular, we will be interested in models containing four stacks of branes. This is essentially because, just as in the explicit example presented in last section, we need at least four branes in order to arrange the SM chiral spectrum as coming from bifundamental representations. Let us then consider some four-stack subquivers arising from the hexagonal construction. Having four stacks implies that at least there is a pair of them with vanishing intersection, thus yielding either a $\cn = 2$ or a $\cn = 0 $ sector. Without loss of generality, we will take two $a$-type branes to be such pair, and one of them to be of one definite kind, say $a_2$. In order to have a non-trivial chiral spectrum, we need the other two branes to be of type $\a_i$, with $\a \neq a$. Our general four-stack configuration will be of the form $\left(a_2, a_i^\prime, \b_j, \g_k \right)$. There are some inequivalent choices: - Either $i = 2$, having a massive $\cn = 2$ $D6_aD6_{a^\prime}$ sector,\ or $i = 1$, having a massive $\cn = 0$ $D6_aD6_{a^\prime}$ sector. - Either $\b = \g$, having a non-chiral spectrum on the $D6_\b D6_\g$ sector,\ or $\b \neq \g$, having a $\cn = 1$ chiral spectrum. - Either $j = k$, so that at least one of the $a$ branes preserves the same supersymmetry with both $\b$ and $\g$ branes,\ or $j \neq k$, so that both supersymmetries are different. We thus see that we have eight inequivalent possibilities when considering four-stack subquivers, as the rest of them just amount of some relabeling of the nodes. However, performing a simple anomaly cancellation analysis, we can reject half of them as phenomenologically uninteresting. This comes from the fact that, when performing our model-building, we will usually choose $(a,a^\prime)$ to be the Baryonic and Leptonic branes, respectively. Being a non-intersecting pair of branes, this will avoid massless lepto-quarks in our spectrum. Now, in order to cancel $SU(3)$ anomalies, we need fermions transforming both in [**3**]{} and in ${\bf \bar 3}$ representations, same number of them on each. It can be easily shown that, in order for this to be true in our models, we must have either $\b = \g$ and $j \neq k$, or $\b \neq \g$ and $j = k$. So we are finally led to consider only four subquivers, which are depicted in figure \[quivers\]. In the next section we will try to combine these four Q-SUSY structures with phenomenologically appealing brane configurations, following the philosophy described in Section 2 for getting the SM from general intersecting branes. Model-building: three generation models with SUSY intersections ================================================================ In this section we are going to consider the construction of explicit intersecting D6-brane models of phenomenological interest. In particular we will focus in the construction of models with quasi-SUSY in the sense described in previous section and ref.[@cim1] with three quark-lepton generations and SM group (or some simple left-right extension). We will concentrate in the class of branes discussed above in which each individual brane preserves an unbroken $\cn = 2$ supersymmetry with its mirror brane, having null intersection with it. As was discussed above, in this class of brane configurations only up to three $U(1)$’s may become massive by combining with three of the four antisymmetric fields $B_{\mu \nu }^i$ present in the massless spectrum [@imr]. We will be considering four stacks of branes $a,b,c$ and $d$ each of them containing $N_a=3$, $N_b=2$, $N_c= 1,2$ and $N_d=1$ parallel branes. Thus to start with our gauge group will be either $U(3)_a\times U(2)_b\times U(1)_c\times U(1)_d$ or $U(3)_a\times U(2)_b\times U(2)_c\times U(1)_d$. The most general phenomenologically interesting Q-SUSY models one can consider involving four stacks of branes may be depicted by the four SUSY-quivers shown in fig. \[quivers\]. Locating the four stacks of branes of the SM at the four corners of those quivers we will get Q-SUSY models with the SM gauge group (and possibly some extra $U(1)$ symmetry as discussed below). Without loss of generality, we will take the Baryonic stack to to be of type $a_2$, whereas the Left stack can also be chosen to be of some definite type different from $a$, say $b_2$. Those two stacks will yield a gauge group $U(3)\times U(2)$. Clearly, the leptonic brane must be non-intersecting with the baryonic brane, since we want to avoid lepto-quark particles. It must then be of the same type $a$ as the latter. Depending on what kind of brane one chooses for the right and leptonic stacks, different model building possibilities will be obtained: [*i) Square quiver (fig.(\[qsimr\])).*]{} In this case there are four types of intersections respecting four different types of SUSY. They will correspond to Q-SUSY versions of the three generation configurations of ref.[@imr]. [*ii) Linear quiver (fig.(\[linear\])).*]{} This quiver may be put into a line by putting the leptonic $a_2$’ brane on top of the baryonic stack $a_2$. It may be considered as a variant of the square quiver in which the left-handed fermions have one supersymmetry and right-handed leptons a different one. [*iii) Rombic quiver (fig.(\[rombic\])).*]{} In this third quiver three different SUSY’s are present. One main difference with the previous cases is the presence of Higgs chiral multiplets at the intersections of right and left stacks. Note that the baryonic and Higgs sub-sector will respect the same $\cn = 1$ supersymmetry, whereas the leptonic sector respect other ones. [*iv) SUSY-triangle (fig.(\[MSSM\])).* ]{} In this case baryon and lepton branes are of the same ($a_2$) type so that the quiver may be deformed into a triangle by locating one on top of each other. Note that now all intersections preserve the same $\cn = 1$ SUSY and, as in the rombic quiver, there are massless Higgs multiplets at the intersections. It turns out that is is easy to find families of examples corresponding to the [*square* ]{} and [*linear*]{} quivers. On the contrary, the [*triangle*]{} and [ *rombic*]{} quivers are more restricted configurations and examples are more scarce. Let us emphasize that all of these brane configurations lead in general to RR-tadpoles. Those may be canceled in a model by model basis by adding extra D-branes (with no intersection with the ones in the quivers, so that the massless chiral spectrum is not modified) and/or the addition of explicit $H_{NS}$-flux, as described in Section 2. There are some general aspects of these Q-SUSY configurations which can be given without the need of going into the details of the models. One can see that the following properties hold: [*1)*]{} “ In any square, rombic and linear quiver there is, at least, one $U(1)$ generator which remains massless (does not combine with an antisymmetric B-field). In the triangle quiver there are two such massless $U(1)$’s.” [*2)* ]{} “In any SUSY or Q-SUSY (factorizable) D6-brane configuration yielding SM fermion spectrum the $U(1)_c$ (right) gauge generator is massless.” [*3)*]{} ‘If we insist that the left- and right-handed counterparts of the same field respect the [*same*]{} $\cn = 1$ SUSY, then necessarily the brane configuration must have intersections at which multiplets with the quantum numbers of SM Higgs arise’. The first two general properties are proven in Appendix II. The second one follows because it turns out that supersymmetry at all intersections forces the right-stack to have wrapping numbers $m_c^2=0,$ $n_c^1 = n_c^3 = 0$. Notice that this massless generator $U(1)_c$ may or may not coincide with the one(s) mentioned in point [*1)*]{}. Concerning the third one, it is a direct consequence of the quiver classification above: in order for the [*Baryonic*]{} stack (of type $a_i$) to have intersections giving rise to e.g., left- and right-handed quarks respecting the same $\cn = 1$ SUSY, the [*Left*]{} and [*Right*]{} branes must be of type $b_i$ and $c_i$ respectively (or viceversa, see fig.(\[quivers\])). But branes of type $b_i$ and $c_i$ necessarily intersect, giving rise to massless states with the quantum numbers of SM Higgs multiplets. This is the link drawn between $b_2$ and $c_2$ in the lowest quivers in fig.(\[quivers\]). Due to these properties, there is a number of limitations on the possible model building, which we will discuss in a model by model basis below. As a general consequence we will see that in general Q-SUSY forces the hypercharge to come along with an extra $U(1)_{B-L}$ gauge boson. This general property can only be evaded if Q-SUSY is only an approximate symmetry, as will be discussed in the case of the square SUSY-quiver. Otherwise a gauged $B-L$ must be present, although of course it may be eventually Higgssed. This is an interesting result by itself since it shows an intriguing connection between the presence of supersymmetry and the gauging of $(B-L)$. It is also intriguing the third general property above, since it gives an interesting connection between the presence of SUSY and the existence of Higgs multiplets. We will present specific examples corresponding to the four types of quivers discussed above. Before presenting the models let us emphasize that these are not the unique models that one can construct starting with the four quiver configurations. Other variations may be constructed depending on how we cancel RR-tadpoles (by explicit factorizable branes and/or non-factorizable branes and/or explicit RR H-flux). One can also easily build models with extended gauge symmetries like models with left-right and/or Pati-Salam symmetries. But we think the examples provide the reader with enough information to look for other interesting variations. A general summary of the models we construct here is given at the end of the paper in table \[propmodels\]. The square quiver ----------------- In order to describe explicit examples of square quivers let us start from the generic family of models already presented in table \[solution\]. As we previously said, this family is parametrized by two phases $\epsilon, \tilde \eps = \pm1$, the NS background on the first two tori $\beta^i = 1-b^i = 1, 1/2$, four integers $n_a^2,n_b^1,n_c^1,n_d^2$ and a parameter $\rho = 1,1/3$. As we also remarked, without loss of generality one can choose the baryonic(left)-stacks to be of type $a_2$ ($b_2$). This implies setting $\epsilon ={\tilde \epsilon} = 1$. In order to find out whether some SUSY is unbroken at the intersections, let us display the angles that compose the twist vector $v_\a$ for each brane $\a = a, b, c, d$. Those angles are shown in table \[gral\_angles\] in terms of the ratios $U^i=R_2^i/R_1^i$, where $R_1^i$,$R_2^i$, $i=1,2,3$ are the radii of the tori. In order for some SUSY to be preserved at each intersection, one needs to have angles $\vt_{\a\b}^1 \pm \vt_{\a\b}^2 \pm \vt_{\a\b}^3 \in \IZ$ for some choice of signs, where $\vt_{\a\b}^i$, $i=1,2,3$ are the angles formed by any pair of branes $\a$ and $\b$ in the three 2-tori [@cim1]. Looking at table \[gral\_angles\] one can easily see that for certain choices of the $U^i$ and particular choices of some of the integers involved one can manage so that all intersection preserve one SUSY. Specifically, if the conditions n\_c\^1 = 0 & & =,    \^1=1 ,\ n\_b\^1 &gt; 0 , & &  n\_a\^2 = 3n\_d\^2= n\_d\^2  &gt; 0 ,\ U\^1 =   U\^3 , & & U\^2 = [[n\_a\^2 ]{}]{}   U\^3 , \[tune1\] are met, the angles of the branes have the general form of table \[QSUSY\_angles\], where $\alpha_1=tg^{-1}(U^3/6)$ and $\alpha_2=tg^{-1}(U^3/2)$. Note that with these choices the branes $a,b,c$ and $d$ become $\cn = 2$ branes of types $a_2$, $b_2$, $b_1$ and $a_1$ respectively. Thus we recover the square Q-SUSY structure of fig.\[qsimr\]. Altogether the wrapping numbers of the resulting models with Q-SUSY properties are displayed in table \[QSUSYsquare\]. One can check that there are simple choices of the integer parameters and extra branes with no intersection with the SM branes such that all RR tadpoles are canceled. The massless fermionic spectrum is the same as in the models in [@imr], but now all of them come with a scalar superpartner. We thus have obtained a family of square Q-SUSY models parametrized by the two integers $n_a^2,n_b^1$ and the NS parameter $\beta^2$, yielding just the spectrum of the standard model. In addition now each chiral fermion will have one massless scalar partner, so there will be the coresponding squarks and sleptons. Thus the initial spectrum (before taking into account the SUSY-breaking effects) will be rather similar to that of the MSSM. It turns out however that the choice of parameters in order to get Q-SUSY (in particular setting $n_c^1=0$) changes the structure of the $U(1)$’s in the model. Indeed, as we already mentioned above, this choice implies that the $Q_c$ generator is massless. In the absence of any extra brane (like that given by the last line in table \[QSUSYsquare\]) we would have just $Q_c$ as our only massless generator, rather than hypercharge. Let us analyze this point in more detail. If we substitute in equation (\[caplillos\]) the wrapping numbers of the Q-SUSY SM model above, we obtain for the couplings of the antisymmetric fields to the $U(1)$’s the result B\_2\^1 &&  [[2]{} ]{}F\^[b]{}\ B\_2\^2 &&  3 \^2 (3F\^a +  F\^[d]{})\ B\_2\^3 &&   ( F\^a - F\^[b]{}  +  F\^d  + N\_hF\^[h]{}) \[bfs\] Note that the coupling to $B_2^0$ is zero, because $\prod_i m_a^i=0 \ \forall a$. Thus we see that, in the absence of additional branes ($N_h=0$) all $U(1)$’s but $U(1)_c$ gain masses by combining with antisymmetric tensors. On the other hand, if there are additional $h$ branes present the situation changes. Indeed, recalling eq.(\[uunose\]), we can see that we would have two massless U(1)’s: $U(1)_c$ and the linear combination $U(1)_a-3 U(1)_d +\frac{3n_a^2\beta^2}{N_hm_h^3}U(1)_h$. In fact, a linear combination of these two gives us a $U(1)$ charge whose coupling to SM fermions is precisely that of the standard hypercharge: Q\_[Y]{}=Q\_a-Q\_c-Q\_d-  Q\_[h]{}   . \[y\] In addition to hypercharge, an extra massless $U(1)$ with analogous couplings to that of a $(B-L)$ generator will remain in the massless spectrum. This is a first realization of our comment above that Q-SUSY requires the gauging not only of hypercharge but of the $B-L$ generator. This latter symmetry may eventually be spontaneously broken, in a way analogous to that discussed in subsection 7.4. An alternative way to guarantee the presence of just standard hypercharge at the massless level will be explained in section (5.1). An important point is in order concerning the Higgs sector in square quivers of the SM. In these models no Higgs multiplets appear at any of the intersections of the square. As discussed in Section 2 and in [@imr], scalar fields with the quantum numbers of SM Higgs appear if the branes $b$ and $c$ (which are parallel along the second torus) approach to each other. However, since these two branes preserve opposite types of supersymmetries, the combined $bc$ system preserves no SUSY at all. This is reflected in the fact that the scalars in that combined system may become tachyonic when the branes are sufficiently close. Those tachyons signal the recombination of the branes and spontaneous symmetry breaking, as described in Section 7. However, since this Higgs system is not supersymmetric, in this case the Q-SUSY property of the rest of the spectrum will not be sufficient to stabilize a hierarchy of masses between the weak scale and the string scale, a (modest) fine-tuning being required to keep apart those two scales. This is one of the main motivations to consider the rombic and triangular quivers which we discuss below. We will describe more aspects of electroweak symmetry breaking for the square quiver in section 7. The linear quiver ----------------- In some way this may be understood as a variation of the square quiver in which we flip the type of leptonic brane from type $a_1$ to $a_2$. Due to this change the left-handed quarks and leptons share now the same SUSY, whereas the right-handed ones respect a different one (see fig.(\[linear\])). One can check that the conditions to get Q-SUSY at the intersections are analogous to those in the square quiver eq.(\[tune1\]) and the wrapping numbers giving rise to the SM fermion content are shown in table \[QSUSYlinear\]. One can see that the wrapping numbers are identical to those of the square quiver except for the leptonic brane. This slight change has however an impact in the massless $U(1)$ spectrum. The couplings of the RR fields $B_2^i$ to the $U(1)$’s is identical to that in the square quiver eq.(\[bfs\]) except for a flip in sign of the coefficient of the $B_2^3\wedge F^d$ term. As a consequence one observes that two $U(1)$’s will remain necessarily massless: $U(1)_c$ and $U(1)_a-3 U(1)_d$. Thus in this model again we will have not only a massless hypercharge $Y=Q_a/6+Q_c/2-Q_d/2$ but also an extra $B-L$ generator. One can check that this conclusion does not change even if we add an extra brane as we did in the square quiver case. Note also that, for $\rho =1$, the baryonic and leptonic stacks would have the same wrapping numbers and hence one can unify them into a Pati-Salam stack with gauge group $U(4)_{PS}$ if one locates both stacks on top of each other. The simple set of branes discussed above are not enough to cancel RR tadpoles in this model. In fact one can check that for the massive anomalous $U(1)_b$ generator the Green-Schwarz cancellation mechanism will only cancel partially its anomalies and one would in general need to add explicit $H_{NS}$-flux to cancel tadpoles, as discussed in Section 2. This $U(1)$ being anyway massive, decouples from the observable massless spectrum. Before getting into the rombic class of models let us make a comment concerning the possibility of building left-right symmetric models with gauge group $SU(3)\times SU(2)_L\times SU(2)_R \times U(1)_{B-L}$, i.e., the simplest non-Abelian extension of the SM. It is easy to construct such a models starting from the square and linear quivers by just putting a couple of $(c)$-branes instead of one. We will not present here an analysis of these models, which is straightforward. Let us just make a comment concerning the connection between the property of Q-SUSY and the existence of a gauged $B-L$. In these models, in order to get the desired quantum numbers, we have to ensure that there is a massless $U(1)_{B-L}$ remaining at the massless level (i.e., no couplings to $B_2^i$ fields) and no other linear combination of $U(1)$’s. It turns out that in the linear quiver models imposing the Q-SUSY conditions analogous to eq.(\[tune1\]) automatically guarantees that there is a $U(1)_{B-L}$ gauge boson in the massless spectrum. This connection between Q-SUSY and the presence of $B-L$ is another example of the intriguing connection that we find upon model building in between the presence of Q-SUSY (or SUSY) and the existence of a massless $B-L$ generator. The rombic quiver ----------------- The general structure of the rombic quiver is presented in figure \[rombic\]. In order to give an example of such Q-SUSY configuration we are going to present a model with a left-right symmetric $SU(2)_L\times SU(2)_R\times U(1)_{B-L}$ gauge group. Getting such gauge group will again imply considering four stacks of branes, following the same general philosophy described in section 2 in order to get the Standard Model. In particular, we will be considering four sets of branes with the wrapping numbers of table \[trombic\]. Note that in this brane configuration all the three tori have NS background (i.e., $\beta^i=1/2$, for all $i$). From the above configuration one finds for the intersection numbers: [lcl]{} I\_[ab]{} =  3, & & I\_[ab\*]{} = 0,\ I\_[ac]{} =  -3, & & I\_[ac\*]{} = 0,\ I\_[bc]{} =  3, & & I\_[bc\*]{} = 0,\ I\_[cd]{} =  -3, & & I\_[cd\*]{} = 0,\ I\_[bd]{} =  0, & & I\_[bd\*]{} = 3, whose associated massless chiral spectrum is presented in table \[trombic2\]. Unlike the square quiver case, this model has electroweak Higgs fields in the massless chiral spectrum. The property of Q-SUSY is obtained for U\^1 = U\^2 = U\^3 \[tune2\] in which case each brane respects some $\cn = 2$ supersymmetry and the branes $a,b,c$ and $d$ are of type $a_2,b_2,c_2$ and $a_1$ respectively. Let us analyze the structure of the $U(1)$’s. In this case one has that the RR fields $B_2^{i}$, $i=1,2,3$ couple to the $U(1)$’s as follows: [rcl]{} B\_2\^1 &&   6(- F\^b + F\^c)\ B\_2\^2 &&  3 ( 3F\^a - 2F\^c + F\^d )\ B\_2\^3 &&  3( -3F\^a + 2F\^b + F\^d ) \[brombic\] From here one concludes that there is a unique massless $U(1)$ given by: Q\_[B-L]{} =  -[[2]{}3]{}Q\_a -  Q\_b - Q\_c which couples to the fermionic spectrum as the standard $(B-L)$ generator of left-right symmetric models. An advantage of this model with respect to the square and linear quivers is that the subsector formed by $Q_R,Q_L$ and the electroweak Higgs fields $H$ all respect the same $\cn = 1$ supersymmetry. Thus the quark subsector of the theory is supersymmetric. Since we know experimentally that the Yukawa couplings of the leptons are in general very small, it is enough to have an approximate SUSY in the quarks sector in order to stabilize the hierarchy between the weak scale and a possible cut-off scale of order 10-100 TeV. As in the linear model, the set of branes discussed above are not enough to cancel RR tadpoles in this model. In fact one can check that for the massive anomalous $U(1)_d$ generator the Green-Schwarz cancellation mechanism will only cancel partially its anomalies. As in the linear models, one would in general need to add explicit $H_{NS}$-flux to cancel tadpoles. Again $U(1)_d$ is massive and decouples anyway at low energies. Let us finally mention that the left-right symmetry may eventually break down to the Standard Model by brane recombination, in a way analogous to that explained for the other models in Section 7. The reader can verify, following the methods developed in that section, that after the recombination of one of the $(c)$ branes with the $(d)$ brane into a single one, the resulting configuration yields the fermion content and the gauge group of the SM. This would correspond, from the effective action point of view, to giving a vev to the right-handed sneutrinos, what eventually would break $B-L$ into the hypercharge. The $\cn = 1$ triangle quiver ----------------------------- Let us now show an example based in the triangle quiver in fig. \[MSSM\]. Note that, as shown in fig. \[quivers\], the baryonic ($a$) and leptonic ($d$) branes are of the same ($a_2$) type and could be put on top of each other in the quiver, leading to a triangular shape. In this example all of the intersections will preserve the same $\cn = 1$ SUSY so that, at least locally, the models will look fully supersymmetric. As we remarked at the beginning of this section, this quiver predicts the necessary presence of Higgs multiplets appearing at the intersections of [*Left* ]{} and [*Right* ]{} brane stacks. Imposing the that left- and right-handed fermion multiplets respect the same $\cn = 1$ SUSY forces the existence of these Higgs multiplets. Consider the wrapping numbers in table \[SUSYmodel\]. It is easy to check that if the conditions 2U\^1 = 2\^2 U\^2 = U\^3 \[tune3\] are met, the brane configuration indeed respects the same supersymmetry at all intersections. Here again $U^i=R_2^i/R_1^i$, where $R_1^i$,$R_2^i$, $i=1,2,3$ are the radii of the tori. From the brane wrapping numbers above one can obtain the intersection numbers [lcl]{} I\_[ab]{} =  1, & & I\_[ab\*]{} = 2,\ I\_[ac]{} =  -3, & & I\_[ac\*]{} = -3,\ I\_[bd]{} =  -1, & & I\_[bd\*]{} = 2,\ I\_[cd]{} =  3, & & I\_[cd\*]{} = -3,\ I\_[bc]{} =  -1/\^2 , & & I\_[bc\*]{} = -1/\^2 , \[intersicsm\] which gives rise to the massless spectrum in table \[espectrosm\]. Let us now study the structure of the couplings of the $U(1)$ fields to the antisymmetric B-fields, which determine which of them become massive. One finds the couplings: [rcl]{} B\_2\^1 &&   [2]{} F\^b\ B\_2\^2 &&  3\^2 ( 3F\^a  + F\^d )\ B\_2\^3 &&   - [12]{}( 9F\^a + [2\^2]{} F\^b + 3F\^d )  . \[bmssm\] One concludes from here that both $U(1)_b$ and $3U(1)_a+U(1)_d$ generators become massive. The massless anomaly-free generators are thus $U(1)_c$ and $U(1)_a-3U(1)_d$, which correspond with the third component of right-handed isospin and B-L, respectively. The hypercharge is given by the linear combination of generators: U(1)\_Y = [ [U(1)\_a]{}6 ]{}   - [ [U(1)\_c]{}2]{}  - [ [ U(1)\_d]{} 2]{}  . Thus, as expected, the massless hypercharge comes along with a massless B-L generator, if supersymmetric intersections are imposed. Note also that, as in the linear quiver case, if one locates the leptonic stack of branes $d$ on top of the baryonic stack one would get an enlarged $U(4)_{PS}$ Pati-Salam gauge symmetry. Note that there are two versions of the model with $\beta^2 = 1,1/2$. Interestingly enough, for the $\beta^2=1$ case [*the massless chiral spectrum is exactly the same of the MSSM*]{}. For $\beta^2=1/2$ the Higgs sector is doubled, although the rest of the spectrum remains the same. As in the rombic example, the simple brane configuration by itself would give rise to RR tadpoles. Those may be canceled by the addition of either some (non-factorizable) brane system with no intersection with the SM branes (for the $\beta^2=1/2$ case) or some explicit $H_{NS}$-flux (for the $\beta^2=1$ case). Before ending this model-building section let us make a final comment. The square quiver models have an attractive point which the others do not have. Square quiver models are just a subclass of those presented in [@imr], and the latter have the attractive feature that the number of generations is related to the number of colours by cancellation of $U(1)_b$ anomalies. In order for this to work there cannot be extra fermion doublets, like the Higgsinos appearing in the triangle and rombic models, which contribute to $U(1)_b$ anomalies. In the case of the linear quiver models one can check that the $U(1)_b$ anomalies do not cancel and Wess-Zumino terms (induced by the presence of $H_{NS}$-flux) must appear (in addition to the Green-Schwarz mechanism terms) to complete $U(1)_b$ anomaly cancellation. Thus also in this case the number of generations/colours argument is not present. On the other hand, as we mentioned above, the triangle and rombic quivers predict the generic presence of massless Higgs multiplets to give rise to electroweak symmetry breaking. SUSY-breaking and Fayet-Iliopoulos terms ======================================== The Q-SUSY examples discussed in the previous section may present three different sources of supersymmetry breaking: [*1)*]{} Soft SUSY-breaking masses from loop graphs, as discussed in Section 3. These will in general be present in the four types of models considered, since all of them will have non-SUSY massive sectors with masses generically of order $M_s$. They will induce generically one-loop gaugino masses as well as two-loop scalar masses$^2$. These effects will be of order $\alpha/(4\pi)M_s$, $M_s$ being the string scale. [*2)*]{} Fayet-Iliopoulos terms. These may be present in any of the models and have a nice geometrical interpretation, as discussed in [@csu; @cim1]. Indeed, in order to obtain SUSY at the brane intersections, we had to appropriately tune the $U^i$ complex structure parameters (see eqs. (\[tune1\]),(\[tune2\]), (\[tune3\])). A slight departure from these adjustments is seen in the effective Lagrangian as the turning on of FI-terms for the $U(1)$ fields of the models. This source of SUSY-breaking may be large or small, depending on the values given to the $U^i$. [*3)*]{} Some models have explicit SUSY-breaking in some sector. In particular we already mentioned that in the square and linear quivers the Higgs sector is non-SUSY and so will be the Yukawa couplings. In the case of rombic models the Higgs, and quark sectors respect SUSY at tree level but there will be explicit SUSY-breaking in the leptonic sector (non-SUSY Yukawa couplings). This is expected because, although all intersections preserve some SUSY, Yukawa couplings involve in general different intersections respecting different supersymmetries. As we mentioned, since the leptonic Yukawas are known to be small, the Higgs mass will be sufficiently protected from loop corrections so as to avoid fine-tuning. Finally, in the triangle quiver case all intersections respect the same SUSY and there is a supersymmetric superpotential. The first source of SUSY-breaking will depend on the particular massive non-SUSY sector of the given model. On the other hand following ref.[@cim1] we can give a relatively model-independent discussion of the FI-terms in this class of theories. Consider a D6-brane which is forming angles with the orientifold plane given by $(\vt^1, \vt^2, \vt^1+\vt^2 + \d, 0)$, $\vt^i > 0$. Then, for $\d =0$ one would recover one unbroken supersymmetry. For small non-vanishing $\d $ one can approximate the effect of this by turning on a Fayet-Iliopoulos term for the $U(1)$ field associated to the brane and given by \_a - [2]{}M\_s\^2 . \[Fayet4\] In order to compute the mass of a scalar living at the intersection of two branes, let us take two D6-branes $D6_a$ and $D6_b$ whose separation from the same SUSY limit is given by $\d_a$ and $\d_b$, respectively. Then the mass of this scalar is given by m\_[ab]{}\^2 = (- q\_a \_a - q\_b \_b) (\_a - \_b) M\_s\^2 . \[FImass\] and in fact one can check that this result is in agreement with the masses for the scalars lying at a intersection obtained from the string mass formulae (see ref.[@cim1] for details). Thus for small deviations from a SUSY configuration, the masses of the scalars at an intersection may be understood as coming from Fayet-Iliopoulos terms. Let us emphasize some points to try to avoid confusion. Normally, when one talks about a FI-term in a SUSY field theory, one is taking about a massless $U(1)$. On the other hand in our models with exact Q-SUSY it is the massive $U(1)$’s (masses of order $M_s$) that get FI-terms. This may be understood because a FI-term is the SUSY partner of a $B\wedge F$ coupling [@Dine], and those $U(1)$’s with a non-vanishing coupling to some $B$-field are the ones which become massive. On the other hand these massive $U(1)$’s decouple well below the string scale. This means that, e.g., there are no quartic scalar couplings proportional to the gauge coupling constants of massive $U(1)$’s. Thus from the effective low-energy point of view the masses of scalars coming from FI-terms of these massive $U(1)$’s will look rather like standard soft SUSY-breaking masses. It is about these scalar masses that we will be talking in this section. FI-terms and sfermion masses in the square and linear quivers ------------------------------------------------------------- Let us now consider the turning of FI-terms in the square quiver. In fact we are going to consider a slight generalization of the square quiver which has the relative advantage that the only unbroken $U(1)$ is hypercharge. Indeed, in order to get exact SUSY at all intersections we imposed $n_c^1=0$. This was required so that the brane $c$ makes a ${\pi\over2}$ angle with the O-plane in tori 1 and 3, preserving in this way two supersymmetries with the orientifold and one supersymmetry in each intersection with branes $a$ and $d$. This had as a consequence that the single $U(1)$ massless (in the absence of extra branes) was $Q_c$, rather than hypercharge. On the other hand, if we are going to break SUSY anyway by adding FI-terms, we can also relax the Q-SUSY condition and allow for approximate SUSY at the intersections involving the $c$-brane. Let us start from the model of table \[solution\] and let us impose the conditions n\_b\^1 &gt; 0 , & & n\_a\^2 = 3n\_d\^2 &gt; 0 ,\ U\^1 =U\^3 , & & U\^2 = [[n\_a\^2]{}]{} U\^3. \[ello2\] Note that we have not imposed the condition $n_c^1=0$, and thus we do not need to impose any condition on $\beta^1$, $\rho$. The conditions (\[ello2\]) will ensure SUSY for all intersections not involving brane $c$. Now, the condition to get a massless hypercharge generator is: n\_c\^1=- [\^22 \^1]{}(n\_a\^2+3n\_d\^2) = - [\^2\^1]{}3n\_d\^2 , since we have chosen $\eps = \tilde \eps = 1$. This implies that in order to avoid multiple wrappings [@imr] we also need to set $\rho=1/3$ (and hence $n_a^2 = n_d^2$). We present in table \[solution2\] the wrapping numbers of these models. Consider now a slight departure of the moduli $U^{1,2}$ from their assigned values in (\[ello2\]): [c]{} U\^1 = U\^3 +\^1\ U\^2 = [[n\_a\^2]{}]{}U\^3+\^2 \[tune4\] and let us also define    [2]{}-tan\^[-1]{}(U\^1)  = -tan\^[-1]{}  . \[deltona\] This parameter $\Delta $ will measure the departure from supersymmetry of the brane $c$. Under these conditions the angles formed by the branes with the orientifold plane are the ones shown in table \[apLSUSY\_angles\], where $\alpha_1=tan^{-1}(\frac{ U^3}{6})$ and $\alpha_2=tan^{-1}(\frac{ U^3}{2})$. There one has (for small $\d^i$) \_a & = &  ;\ \_b & = &  ;\ \_d & = & \[deltas\] Note also that (again for $\delta^1\ll1$): tan (1-) so that one can check that $\Delta $ can be made quite small for not too large $U_3$. From table \[apLSUSY\_angles\] one sees that for all practical purposes the effect of not setting $n_c^1=0$ induces SUSY-breaking which (for small $\Delta$) can be parametrized as a FI-term for the $U(1)_c$ generator. We can now compute the masses of the sparticles in terms of the FI-terms following the discussion given in ref.[@cim1]. The results are shown in table \[sparticles2\]. Notice that we could formally put the masses depending on $\Delta$ as coming from a Fayet-Iliopoulos term. However, and unlike the other cases, supersymmetry in these intersections would only be recovered in the $U^3\to\infty$ limit. An important point to note is that there is a wide range of parameters for which all squark and slepton (mass)$^2$ are positive and hence there are no unwanted charge and colour-breaking minima. It is enough to satisfy $\delta_a,\delta_d> \Delta$ and $\d_a > |\d_b|$ in order to have all (mass)$^2$ positive. Note also that (for a given choice of discrete parameters) the squark and slepton masses come determined from just three independent parameters $\d_1, \d_2$ and $U^3$. In some limits the expressions for these masses becomes particularly simple. For example, if $\d_1=0$ one gets m\^2\_[[Q]{}\_L]{} = [ [\_a]{}2]{}M\_s\^2  &;&  m\^2\_[[L]{}]{} = [ [\_d]{}2]{}M\_s\^2\ m\^2\_[[U]{}\_R,[D]{}\_R]{} = ( [ [\_a]{}2]{}   [ 2]{} ) M\_s\^2 & ;&  m\^2\_[[E]{}\_R,[N]{}\_R]{} = ( [ [\_d]{}2]{}   [ 2]{}) M\_s\^2 \[msql\] Note that in this case the ratio of [*average* ]{} squark masses versus slepton masses is controlled only by the value of $U^3$:  =  [ [\_a]{} ]{} = [13]{}  [ [1+(U\^3/2)\^2]{} ]{}  = [[g\_a\^2]{} ]{} \[massessql\] The last equality follows from eqs.(\[coup\])and (\[longis\]) in the next subsection. In this connection note that using those equations we can rewrite eqs.(\[deltas\]) as \_a =  g\_a\^2(\^2\^2)   ;  \_b = g\_b\^2([[\^1]{}]{})  ;  \_d = g\_d\^2(\^2\^2) \[deltas2\] Thus the (masses)$^2$ of sfermions from FI-terms are proportional to the square of the $U(1)$ coupling constants, very much like if they were coming from a one-loop effect [^9]. In any event, it is interesting how one can obtain the masses of squarks and sleptons in this scheme in terms of a few geometrical parameters. Notice however that one should add to these masses the contributions coming from the loops involving massive non-SUSY particles, as described above. One interesting point concerns the right-handed sneutrino. From the above mass structure one can check that there is a region of parameter space for which one might have a tachyonic mass for the ${\tilde {\nu }}_R$ but a positive one for the rest of the scalars. It is enough to have $\d_a> \Delta $ but $\d_d < \Delta$. One can check that this is possible provided $U^3 > \sqrt{12}$. If this happens lepton number is broken and the $B-L$ gauge boson becomes massive swallowing a linear combination of ${\tilde {\nu }}_R$’s and antisymmetric $B$-fields. From the brane point of view the $c$ and $d^*$ branes will recombine into a single one and the obtained structure is similar to the models discussed in subsection 7.4 . Another point is in order concerning FI-terms. We are used to the fact that in [*supersymmetric*]{} theories the masses that scalars get from a FI-term are proportional to the $U(1)$ charges of their fermionic partners. Note that in the case of the Q-SUSY theories that is in general not the case, as observed in ref.[@cim1]. Indeed, consider for example the FI-term associated to the baryonic symmetry $U(1)_a$. One can see from table \[sparticles2\] that both left- and right-handed squarks get positive masses, although their fermionic partners have opposite baryon number. This is a reflection of the fact that different $\cn = 1$ SUSY’s are preserved at the $ab$ and at the $ac$ intersections. One can repeat an analogous discussion for the linear quiver case. It is easy to check that one obtains the same results for the masses coming from FI-terms than the ones in the square quiver, table \[sparticles2\]. The only difference is that one necessarily has $\Delta =0$ in the linear case. Indeed, if $\Delta \not=0$ is taken, then the $Q_c$ generator is massive and the unbroken generator would be $B-L$, rather than hypercharge. This is different to what happens in the square quiver case in which one can chose parameters so that it is hypercharge which remains massless. FI-terms and sfermion masses in the triangle and rombic quivers ---------------------------------------------------------------- One can also easily compute the masses of the scalars coming from turning on FI-terms in the other triangle and rombic type of models. Note that in these cases there are several (or all) intersections respecting the [*same*]{} $\cn = 1$ supersymmetry. In these $\cn = 1$ subsectors the usual fact will hold and the masses of scalars from FI-terms will be proportional to the $U(1)$ fermionic charges. Consider for example the triangular case. By going slightly away from the supersymmetric configuration U\^1 =  [12 ]{}U\^3  + \_1  ; U\^2 =  [1 ]{}U\^3  + \_2   \[tune5\] it is easy to check that one induces FI terms for the $U(1)$’s as follows: \_a  = \_d  = - [ [\_2 \^2/6]{} ]{} M\_s\^2 ;  \_b  = - [ [\_1/2]{} ]{} M\_s\^2 \[fitriangle\] and $\xi_c = 0$ (recall that $U(1)_c$ is massless). Note that there are only two independent parameters for the FI-terms, which corresponds to the existence of two massive (anomalous) $U(1)$’s, $U(1)_b$ and $3U(1)_a+U(1)_d$ (more physically $(9B+L)$). Now, unlike the case of the square quiver, the masses for the scalars at the intersections will be just proportional to these charges. Looking at the particle spectrum in the triangle models (table \[espectrosm\]) we see that left- and right-handed squarks will have [*opposite*]{} masses from the $U(1)_a$ baryonic FI-term. Thus, FI-terms are unable to give positive masses for all squarks and sleptons at the same time. There is nothing wrong with that, only that the FI-terms contribution cannot be the dominant source of scalar masses if we want to avoid unwanted charge/colour breaking minima, the one-loop contributions should also be important. A similar general conclusion holds for the rombic model, since the quiver contains a triangle with three intersections respecting the same SUSY. Gauge coupling constants ======================== Unlike the heterotic case, the gauge coupling constants of brane models in which each gauge factor lives in a different stack of branes have no unification. Rather, the size of each coupling constant square is inversely proportional to the volume wrapped by the corresponding brane. The physical gauge couplings will depend on [*i)*]{} The size of the gauge couplings at the string scale. [*ii)*]{} The running between the string scale $M_s$ and the weak scale. For a string scale of order 10-100 TeV there is running for two to three orders of magnitude. The effect of this running may be enhanced due to the fact that in between the string scale, and the weak scale there is not only the spectrum of the SUSY SM but also of the adjoint scalars and fermions corresponding to the $\cn = 4$ structure of the gauge sectors. Those fields are expected to get masses of order $\alpha/(4\pi) M_s$, which may be well below the string scale. In addition close to the string scale there can be important threshold corrections due to the excitation of KK , winding and “gonion” states [@afiru2] some of which may be below the string scale. All these running effects are difficult to evaluate in a model independent way. However one can obtain easy and closed formulae for the size of the gauge couplings at the string scale. The formulae for the size of gauge couplings get simplified in the case of Q-SUSY and SUSY models compared to the general non-SUSY intersection models. As noted in e.g. [@afiru2], the tree-level value of the different gauge coupling constants at the string scale are controlled by the length of the wrapping cycles, i.e.,  =  [[M\_s\^3]{}(2)\^2 [\_[II]{}]{}]{}  ||l\_i|| \[coup\] where $M_s$ is the string scale, $\lambda_{II}$ is the Type II string coupling, and $||l_i||$ is the length of the cycle of the i-th set of branes  =  ((n\_i\^1R\_1\^[(1)]{})\^2+(m\_i\^1R\_2\^[(1)]{})\^2) ((n\_i\^2R\_1\^[(2)]{})\^2+(m\_i\^2R\_2\^[(2)]{})\^2) ((n\_i\^3R\_1\^[(3)]{})\^2+(m\_i\^3R\_2\^[(3)]{})\^2)  . \[length\] The ratio of the $SU(3)$ and $SU(2)_L$ couplings is thus given by : = \[relacion1\] For models without a non-Abelian extended gauge structure (like the square and triangle quiver examples above) the hypercharge is a linear combination of $U(1)_a$, $U(1)_c$ and $U(1)_d$. One has for the hypercharge coupling: \_Y\^[-1]{}=\_[a]{}\^[-1]{}+(\_[c]{}\^[-1]{}+ \_[d]{}\^[-1]{}) where $\alpha_a$ is the coupling of the $U(1)_a$, which in our normalization verifies at the string scale $\alpha_a=\alpha_{QCD}/6$. We thus have: & = & + (+ )\ & = & +(+ ) \[hiperra\] We can now evaluate these ratios for the different models. It turns out that for Q-SUSY models the complicated non-linear expression for the volume of the cycles (\[length\]) is substantially simplified and the dependence on moduli becomes linear [@cim1]. [*i) Square and linear quiver*]{} In this case we have for the wrapped volume of each brane [rcl]{} ||l\_a|| & = & [3n\_a\^2]{}R\_1\^[(1)]{}R\_1\^[(2)]{}R\_1\^[(3)]{} (1+(U\^3/6)\^2)\ ||l\_b||& = & R\_1\^[(1)]{}R\_1\^[(2)]{}R\_1\^[(3)]{} (1+(U\^3/2)\^2)\ ||l\_c|| & = & R\_1\^[(1)]{}R\_1\^[(2)]{}R\_1\^[(3)]{}(U\^3)\^[2]{}\ ||l\_d|| & = & [3n\_a\^2]{}R\_1\^[(1)]{}R\_1\^[(2)]{}R\_1\^[(3)]{} (1+(U\^3/2)\^2) \[longis\] Here we have $\rho =1/3$ in the square quiver and both choices $\rho=1,1/3$ in the linear quiver case. With these values for the volumes of the cycles one has for the ratio of $SU(3)$ and $SU(2)_L$ couplings: == [[n\_b\^1]{}]{} [ [1+(U\^3/2)\^[2]{}]{} ]{} \[relacion2\] and for the hypercharge coupling = + ( [ [ 3+(3/4+[[n\_b\^1]{} ]{}) (U\^3)\^2 ]{} ]{} ) [*ii) Triangle quiver* ]{} In this case we have for the wrapped volume of each brane [rcl]{} ||l\_a|| & = & 9 R\_1\^[(1)]{}R\_1\^[(2)]{}R\_1\^[(3)]{} (1+(U\^3/6)\^2)\ ||l\_b|| & = & R\_1\^[(1)]{}R\_1\^[(2)]{}R\_1\^[(3)]{} (1+(U\^3/2)\^2)\ ||l\_c|| & = & R\_1\^[(1)]{}R\_1\^[(2)]{}R\_1\^[(3)]{}(U\^3)\^[2]{}\ ||l\_d|| & = & ||l\_a|| \[longissm\] so that we find for the ratio of strong and weak couplings = =  [ [1]{} ]{} [ [1+(U\^3/2)\^2]{} ]{} \[relacion3\] and for the hypercharge coupling =  +    ( 1+ [1]{} [ [(U\^3)\^2]{} ]{} ) \[relacion4\] As we said, this gives us the ratios of the coupling constants at the string scale, and one should compute the running of the couplings in the region in between the weak scale and the string scale. There can be important effects if there are extra particles in that energy region from e.g., KK or winding states, gonions etc. One can in fact check that if one neglects the effect of the running, it is not possible to find values for $U^3$ such that one reproduces at the string scale the experimental weak scale ratios $\alpha_{QCD}^{-1} \ : \ \alpha_L^{-1} \ : \ \alpha_Y^{-1} \ = \ 8.3 \ : \ 29.6 \ : \ 98.4$. In the case of the square quiver it is easy to obtain the weak scale result for $ \frac{\alpha_{QCD}}{\alpha_L}$ but the ratio $\frac{\alpha_{QCD}}{\alpha_Y}$ turns out to be too small. In the case of the triangle model presented above it is also difficult to obtain $ \frac{\alpha_{QCD}}{\alpha_L}$ large enough. This is due to the fact that this particular example has very large winding numbers $n_a^i$ for the baryonic brane, which then tends to give too small $\alpha _{QCD}$. It should be interesting to do a systematic search for other triangle models giving rise to larger values for $\alpha _{QCD}$. Note that these statements concern these particular models with just 4 stacks of branes discussed above. If for example there is an extra brane $h$ contributing to the hypercharge group as in eq.(\[y\]) there is an extra piece ${1\over 4}\frac{||l_h||}{||l_a||}$ to be added to eq.(\[hiperra\]) and in the square quiver case one can then easily adjust $\frac{\alpha_{QCD}}{\alpha_Y}$ by varying the extra brane parameters. The SM Higgs mechanism as a brane recombination process ======================================================= All the above constructions involve the unbroken electroweak gauge symmetry $SU(2)_L\times U(1)_Y$ which we know is spontaneously broken by the Higgs mechanism. The different models have scalars in their spectrum with the quantum numbers of SM Higgs fields to do the job. Whereas the mechanism is understood at the field theory level, there should exist a string version of the SM Higgs mechanism in terms of the underlying branes. Specifically, from the string theory point of view we know that the gauge group of the SM originates from open strings starting and ending on D-branes. The gauge group $SU(2)_L$ originates on two parallel branes whereas hypercharge comes from a linear combination of $U(1)$’s attached to different branes. In the SM Higgs mechanism the rank of the gauge group is reduced. As we will now discuss, the stringy counterpart of this rank-reduction is brane recombination. Brane recombination is a process in which two intersecting branes fuse into a single one (see fig.\[fussion\]). In particular, it is known that tachyons appearing in string theory signal an instability with respect to the decay of the system into another one with lower energy. Tachyons appearing at a pair of interesting branes show the instability with respect to the recombination of both branes into a single one with less energy, i.e., less wrapped volume and lying in the same homology class than the initial pair. In a brane recombination process the number of massless chiral fermions decreases. Consider a couple of intersecting branes $\alpha $ and $\beta $ which recombine into a single final brane $\gamma$. Consider an spectator brane $\rho $ which intersects both branes $\alpha $ and $\beta $ ( i.e., $I_{\rho \alpha }\not= 0$, $I_{\rho \beta }\not= 0$) but not participating in the recombination. The net number of chiral fermions before recombination $n_i$ n\_i =  |I\_|+|I\_|   |I\_ + I\_|   = |I\_|  = n\_f \[ctheorem\] is bigger than the net number of chiral fermions after recombination $n_f$. The effective field theory interpretation is that after a Higgs mechanism some of the chiral fermions may acquire masses from (not-necessarily renormalizable) Yukawa couplings. In what follows we will show how general arguments about brane recombination as above yield results which are in agreement with the low-energy field theory expectations. However we will also see that brane recombination has some extra stringy ingredients: it cannot just be described by the lightest tachyon getting a vev, higher excitations of the tachyon must be involved in the full process of brane recombination. This seems to indicate that a full description of the process would require string field theory, rather than an effective field theory Lagrangian. Our argumentation on recombination will only make use of the computation of intersection numbers before and after brane recombination, imposing conservation of RR charge. Thus it will be mostly topological rather than geometrical in nature, i.e., we will be able to say what fermions and gauge bosons become massive in a brane recombination process, but not the size of the masses acquired. The understanding of the Higgs mechanism as a brane recombination process has nothing to do with the presence or not of some SUSY at the brane intersections. So our discussion below also applies to general non-SUSY intersecting models. However for concreteness we will first discuss how the Higgs/recombination process occurs in the square quiver models and will later describe the case of the triangle quiver in which an interesting departure between brane recombination and effective field theory Higgsing occurs. The electroweak Higgs system in the square quiver -------------------------------------------------- We will consider for simplicity the version of the square quiver discussed in section (5.1) in which only the hypercharge remains massless, although the discussion in this subsection applies in fact to the larger class of models discussed in [@imr]. In the square quiver there are not generic intersections giving rise to massless scalars with the quantum numbers of the SM Higgs fields. On the other hand open strings stretched between the ‘$b$’ and ‘$c$’ stacks of branes do have the quantum numbers of Higgs fields appropriate to yield electroweak symmetry-breaking. These two stacks of branes are parallel on the second torus, and that is why they generically do not intersect. However if one approaches those stacks to each other scalar Higgsses with the quantum numbers of table \[higgsses\] appear in the light spectrum [@imr]. These states corresponding to open strings stretching between branes $b$ and $c$ (denoted by $h^\pm$) and between branes $b$ and $c^*$ (denoted by $H^\pm$) have masses (see [@imr]) \^2  &=& [X\^2\_[bc\^\*]{} 4]{} M\_s\^2   ||\^1\_[bc\^\*]{}|-|\^3\_[bc\^\*]{}||  =  M\_s\^2  ([X\^2\_[bc\^\*]{}4]{}    (2\_2-+\_b) ),\ [m\_[h\^]{}]{}\^2  &=& [X\^2\_[bc]{} 4]{} M\_s\^2   ||\^1\_[bc]{}|-|\^3\_[bc]{}||  =  M\_s\^2  ([X\^2\_[bc]{}4]{}    (2\_2++\_b) ), \[mashigs\] where $X_{bc^*}$ ($X_{bc}$) is the distance (in $\alpha^{\prime \frac 12}$ units) in transverse space along the second torus. One also has $\alpha_2=tg^{-1}(U^3/2)$ and $\Delta$, $\d_b$ were defined in (\[deltona\],\[deltas\]). There are also fermionic partners (“Higgsinos”) of these two types of complex scalar fields. The above scalar mass spectrum can be interpreted as arising from a field theory mass matrix (H\_1\^\*  H\_2) ( **[M\^2]{} ) (** [c]{} H\_1\ H\_2\^\* ) +(h\_1\^\*  h\_2) ( **[m\^2]{} ) (** [c]{} h\_1\ h\_2\^\* ) + h.c. where = M\_s\^2 ( [cc]{} X\_2\^[(bc\^\*)]{}& (2\_2-+\_b)\ (2\_2-+\_b) & X\_2\^[(bc\^\*)]{}\ ),  [**m\^2**]{}=(bc\^\*,-bc,) The fields $H_i$ and $h_i$ are thus defined as H\^=[12]{}(H\_1\^\*H\_2);  h\^=[12]{}(h\_1\^\*h\_2)  . It is clear from the above formulae that when the distance between each pair of stacks is small enough, some of the scalars $H^{\pm}$ and $h^{\pm}$ become tachyonic, which will be the signal of spontaneous symmetry breaking in the SM. In this process the weak vector bosons $Z_0$ and $W^{\pm}$ get a mass but also the fermions get masses proportional to their Yukawa couplings to the different Higgs fields. The form of the Yukawa couplings among the SM fields in table \[tabpssm\] and the different Higgs fields are essentially fixed by conservation of gauge symmetries and have the general form [@imr]: y\^U\_jQ\_LU\_R\^j h\_1  + y\^D\_jQ\_LD\_R\^jH\_2  +\ y\^u\_[ij]{}q\_L\^iU\_R\^j H\_1  + y\^d\_[ij]{}q\_L\^iD\_R\^jh\_2  +\ y\^L\_[ij]{}L\^iE\_R\^jH\_2  +  y\^N\_[ij]{}L\^iN\_R\^jh\_1  +  h.c. \[yuki\] where $i=1,2$ and $j=1,2,3$. With this Yukawa structure one observes that, for example, if only the Higgs fields of type $H_i$ get vevs, two down-like and one up-like quarks still remain massless, as well as all neutrinos. If in addition the Higgs of type $h_i$ get a vev, all fermions get a mass. As discussed in previous sections, it is clear that this Higgs sector of the square quiver is not supersymmetric. However one can obtain a electroweak scale well below the string scale $M_s$ by “modest tuning” the distances $X^2_{bc^*}/2\pi = 2\alpha_2 - \Delta + \d_b$, $X^2_{bc}/2\pi = 2\alpha_2 + \Delta + \d_b$. Let us turn now to a description of the electroweak Higgs mechanism in the brane recombination language. The mechanism of brane recombination ------------------------------------ We are now going to show how the process of brane recombination yield results consistent with the above field theory description of the Higgs mechanism. For this to happen we have to show that [*i)*]{} The final gauge group after brane recombination is indeed just $SU(3)\times U(1)_{em}$. [*ii)*]{} The quarks and leptons become massive in the expected way. Before proceeding let us first make a few comments about the recombination process. In general, even if the two initial intersecting branes $\alpha$ and $\beta$ wrap factorizable cycles, the resulting recombined brane $\gamma$ will not be factorizable. In this case the form of the recombined cycle will not have an easy geometrical description. Still, many properties of the recombination process will only depend on the general homology class of the cycles. In the case of non-factorizable cycles one has to work with a $2\times 2\times 2=8$-dimensional basis for the RR-charges of the cycles. Instead of wrapping numbers $n^i,m^i$, one has to work with 8-dimensional vectors $ \vec{q}_k$, $k=1-8$ (see Appendix I). For a given system of branes to be recombinable into another the initial and final configurations must lie in the same homology class, the recombination must preserve RR charge: \_[s=initial branes]{} N\_s (1+) \_s  = \_[f=final branes]{} N\_f (1+) \_f where $\Omega \vec{q}_s$ gives the charge vector of the orientifold mirror of each brane. Consider for example the case of the Higgs fields $H_i$ in our class of models. These states come from the open string exchange between branes of type $b$ and $c^*$. Thus this Higgs field taking a vev should be related to the recombination of one of the two branes of type $b$ (let us call it $b_1$) with the brane $c^*$ into a single non-factorizable ‘$e$’ brane with charge: 2 \_b+\_c=(\_[b\_1]{}+ \_c)+\_[b\_2]{} \_e +\_[b\_2]{} Instead of two $b$-branes (which contained the $SU(2)_L$ charged current electroweak interactions) we are now left with only one. Thus the electroweak symmetry has been broken. We can now check which chiral fermions, if any, remain in the massless spectrum by computing the new intersection numbers. The latter are easy to compute knowing that $I_{\alpha \beta}$ is a bilinear quantity in terms of $\vec{q}_\alpha $ and $\vec{q}_\beta$ (see Appendix I). In our case we simply have $I_{ae}=I_{ab_1}+I_{ac^*}$ and $I_{ae^*}=I_{ab_1^*}+I_{ac}$. Thus the new intersection numbers $after$ recombination are: [lcl]{} I\_[ab\_2]{} =  1, & & I\_[ab\_2\^\*]{} = 2,\ I\_[ae]{} =  -2, & & I\_[ae\^\*]{} = -1,\ I\_[b\_2d]{} =  -3, & & I\_[b\_2d\^\*]{} = 0,\ I\_[de]{} =  0, & & I\_[de\^\*]{} = -3, \[intersec3\] Note that the number of chiral fermions has been reduced, only three quark flavours and some three leptons remain massless. This result is as expected from field theory arguments. Indeed, looking at (\[yuki\]) we see that if Higgs fields of type $H_i$ get vevs, two up-like and one down-like quarks become massive and also charged leptons do, due to Yukawa couplings. In figure \[recombination\] we can obtain an intuitive image of what may be happening. The Yukawa coupling amplitudes are obtained from correlators involving vertices of quarks and Higgs fields on the vertices of a triangle in the extra compact dimensions. The different Yukawa couplings are expected to be exponentially suppressed [@afiru2; @imr] by the area of the corresponding triangle i.e., $Y_{abc^*}\ =\ \exp{(- A_{abc^*})}$. After the brane $c^*$ has recombined with one of the branes $b$, the triangle is smoothed out at the $bc^*$ vertex. This corresponds to the Higgs fields $H_i$ getting a vev. Note that the net contribution to $I_{ae}$ due to the quarks at the other two vertices is zero, these quarks are no-longer chiral, have acquired masses. At the end of the above recombination process we have four stack of branes $a$, $b_2$, $e$ and $d$, suggesting a gauge group $SU(3)\times U(1)_a\times U(1)_b\times U(1)_e \times U(1)_d$. Now, if our interpretation of the Higgs mechanism as a recombination process is correct, we should just obtain electromagnetic charge as our only surviving $U(1)$ boson in the recombined system. Thus as a test we should check that indeed it is the electromagnetic charge with survives the recombination process. The $U(1)$ combination with remains light $U(1)_{massless}$ is the one with no coupling to the RR closed string fields $B_2^I$, $I=0,1,2,3$. This is the combination U(1)\_[massless]{} = \_[k=a,b\_2,e,d]{} c\^kU(1)\_k, with coefficients $c^k$ such that \_kc\^k[d]{}\_k\^I = 0 ,  I=0,1,2,3 \[cons\] where ${\tilde d}_k^I$ are the coefficients of the couplings of RR fields to the four recombined $U(1)$’s, i.e., the coefficients in the couplings ${\tilde d}_k^I\ B_2^I\wedge F_k$. The corresponding coefficients for the starting cycles $d_k^I$ are given in eq.(\[caplillos\]). Using linearity of the RR charge in the recombination one can compute the new 4 ${\tilde d}_k^I$ coefficients in terms of the ones before recombination $d_k^I$ as follows: - [$\tilde d_k^I= d_k^I$]{} for those brane stacks $k$ not taking part in the recombination. - [$\tilde d_k^I=\frac{1}{N_\alpha}d_\alpha^I+\frac{1}{N_\beta}d_\beta^I$, for the new brane resulting from the recombination.]{} Using this and imposing eq.(\[cons\]) before and after recombination one can compute explicitly the new ${\tilde d}_k^I$’s and obtain which $U(1)$’s remain massless. In the present case, if before recombination we had as the only surviving $U(1)$ the hypercharge of eq.(\[hyper\]) (with $r=-1$), after recombination there is only one remaining massless generator given by Q\_[em]{}=Q\_a+Q\_e-Q\_d- Q\_[b\_2]{}   . \[em\] One can compute the charges of the chiral fermions that remain massless and one finds the residual spectrum (obtained from the intersection numbers (\[intersec3\]) ) after recombination to be: [**Gauge Group:**]{}  SU(3)$\times$ $U(1)_{em}$ [**Matter Content:**]{}  $2(3)_{-\frac{1}{3}}+2(\bar3)_{\frac{1}{3}}+ 1(3)_{\frac{2}{3}}+ 1(\bar3)_{-\frac{2}{3}}+3(1)_{0}+3(1)_{0} $ Thus we see we have one massless u-quark, two massless d-like quarks (i.e. down and strange) and massless (left and right) neutrinos, as expected from the field theory arguments from the standard Yukawa coupling of the Higgs fields $H_i$ to fermions. This is interesting and in fact not far from the experimental situation in which three quark flavours are relatively light (leading to Gellman’s flavour $SU(3)$ ) and neutrinos are almost massless. From the effective Lagrangian point of view, we could have in addition a vev for the other Higgs fields of type $h_i$. If this is the case the rest of the fermions would now become massive, but still a charge generator should remain unbroken. From the brane recombination point of view this corresponds to the recombination of the new brane ‘$e$’ with the ‘$b_2^*$’ brane which was an spectator in the first recombination. The new recombined brane ‘$f$’ will have a RR charge vector $\vec{q}_f \equiv(\vec{q}_e+\Omega \vec{q}_{b_2})$. We are thus only left with three stacks of branes, $a$,$d$ and $f$. One can easily check that the intersection numbers between these branes is zero. For example, $I_{af}=I_{ae }+I_{ab_2^*}=-2+2=0$. Thus there are no massless fermions left after this second recombination, again as expected from field theory arguments. One might worry that, since we have now only three stacks of branes but we still have four RR fields $B_2^I$, all $U(1)$’s could become massive and we would be left with no photon in the low energy spectrum. This turns out not to be the case. One can easily check that a $U(1)$ generator Q\_[massless]{} =  Q\_a+Q\_f-Q\_d   \[em2\] remains unbroken. This massless generator can be identified with electromagnetism by noting that e.g., the massive fields stretched between branes $a$ and $d$ have charges which correspond to the electric charge they had before this last recombination. Note that from the effective field theory point of view there are other (less interesting) field directions of the scalar potential, particularly if we include in the complete potential additional massive scalars which appear at the intersections of branes $b$ and $c$ with their mirrors $b^*$ and $c^*$. In particular, open strings stretched between the branes $b$ and $b^*$ contain (massive) scalars transforming like symmetric (triplet) and antisymmetric (singlet) $SU(2)_L$ representations which couple to the SM Higgs doublets $H_i$ and $h_i$ and modify the scalar potential. These other field directions could in principle lead to unwanted vacua with breaking of electromagnetic $U(1)$. From the field theory point of view this can be controlled by making the unwanted fields heavy, by appropriately separating each brane from its mirror. These different Higgsing possibilities also exist in the brane recombination language as coming from different choices for brane recombination. The brane recombination discussed above describing the SM Higgs mechanism correspond to the recombination of branes: b\_1 + b\_2\^\*  + c\^\*     e  + b\_2\^\*   f \[cadena\] Recombining e.g., $b_1 + b_2 + c^*$ or $b_1 + b_2 + c$ would have lead to other less interesting final configurations with e.g., broken electromagnetic charge. Thus from the phenomenological point of view these other possibilities should be somehow energetically disfavoured. We have seen how the Higgs mechanism in the SM may be understood as a brane recombination process in which the three branes giving rise to the $U(2)_b\times U(1)_c$ recombine into a single brane giving rise to a final (in general non-factorizable) brane $f$. There is only one massless Abelian generator which can be identified with the standard electromagnetic charge. Thus the actual string vacuum after recombination involves only three stacks of branes $a$,$d$ and $f$ and the gauge group is just $SU(3)_{QCD}\times U(1)_{em}$. There is a couple of questions which appear in such an interpretation: [*i)*]{} The recombination language shows us how the quarks and leptons become massive at each step, but how can one explain the observed existence of large hierarchies of fermion masses? We already mentioned that in the initial picture in which the $a,b,c,d$ branes wrap factorizable cycles, there are Yukawa couplings between Higgs fields and chiral fermions (see fig.(\[recombination\])). The worldsheet with one right-handed fermion, one left-handed fermion and one Higgs field have a triangular shape with those particles at the corners [@afiru2]. As we mentioned above, the corresponding coupling may be exponentially suppressed. After the branes recombine (i.e., after the Higgs get vevs) the vertex where the Higgs field lies is smoothed out (see fig.(\[recombination\])). Still, if the vev of the Higgs is small compared to the string scale, this will amount to a small perturbation, fermion masses would still be exponentially suppressed by the area left between the recombined brane $e$ and the baryonic $a$-stack ($d$-stack in the case of leptons), which is the smoothed out triangle. Thus the fermion masses may still have hierarchical ratios. [*ii)*]{} The actual final vacuum contains only the stacks $a$, $d$ and $f$ and only $SU(3)_{QCD}\times U(1)_{em}$ as gauge group. If we put energy in the system, we should be able to see the electroweak gauge bosons $Z_0$ and $W^{\pm }$ produced, with masses of order the vev of the Higgs fields. Which are those states in the final recombined system? The $Z_0$ is neutral and should correspond to open strings beginning and ending on the same recombined brane ‘$f$’. In particular, it should correspond to an open string stretching between the opposite smoothed portions of that brane in fig.(\[recombination\]). On the other hand open strings stretching between the recombined brane $f$ and its mirror $f^*$ have charge $= \pm 1$ (see eq.(\[em2\])) and should give rise to the charged $W^{\pm}$ bosons. Higgs mechanism and brane recombination in the triangle quiver --------------------------------------------------------------- Let us now discuss for comparison the Higgs mechanism and brane recombination in the triangle quiver. Those models have an extra massless $U(1)_{B-L}$ but we will concentrate first on electroweak symmetry breaking. One important difference is that in these models there are massless chiral multiplets corresponding to Higgs fields at the intersections of the stacks $(b)$ and $(c)$ (see table \[espectrosm\]). In the case $\beta^2=1$ there is in fact a single Higgs set and the charged chiral massless spectrum is just that of the MSSM (plus right-handed neutrinos). Looking at the charges in table \[espectrosm\] one sees that the only allowed Yukawa superpotential couplings are: y\^u\_[ij]{}q\_L\^iU\_R\^j [|H]{}  + y\^d\_[ij]{}q\_L\^iD\_R\^j H  +\ y\^L\_[ij]{}l\^iE\_R\^jH  +  y\^N\_[ij]{}l\^iN\_R\^j[|H]{}  +  h.c. \[yuki2\] with $i=1,2$ and $j=1,2,3$. Thus one would say that one of the generations of quarks and leptons does not get masses at this level. Let us now see how would be the Higgs electroweak symmetry breaking in the brane recombination language. As in the square quiver case, let us assume that one of the two $(b)$ branes, e.g., $b_1$ recombines with $c^*$. Looking at table \[espectrosm\] we see that this should correspond to ${\bar H}$ getting a vev. After the recombination $b_1+c^*\rightarrow e$ into a single brane $e$ the electroweak symmetry is broken. We are left with four stacks of branes $a$, $b_2$, $e$ and $d$ and the intersection numbers are given by [lcl]{} I\_[ab\_2]{} =  1, & & I\_[ab\_2\^\*]{} = 2,\ I\_[ae]{} =  -2, & & I\_[ae\^\*]{} = -1,\ I\_[db\_2]{} =  1, & & I\_[db\_2\^\*]{} = 2,\ I\_[de]{} =  -2, & & I\_[de\^\*]{} = -1,\ I\_[b\_2e]{} =  -[1\^2]{}, & & I\_[b\_2e\^\*]{} = -[1\^2]{},\ I\_[ee\^\*]{} = [2\^2]{} \[intersec5\] with $\beta^2=1,1/2$. It is easy to check now that (apart from the $B-L$ symmetry which remains unaffected by this process and will be discussed later) there is an unbroken massless gauge generator which can be identified with electromagnetism U(1)\_[em]{} = [ [U(1)\_a]{}6 ]{}  - [ [U(1)\_d]{}2]{}  - [ [ U(1)\_[b\_2]{}]{} 2]{}  + [ [ U(1)\_[e]{}]{} 2]{}  . and is not rendered massive by couplings to RR B-fields. The massless chiral spectrum after recombination obtained from the above intersection numbers [^10] is (for $\beta^2=1$) $\begin{array}{ll} {\bf Gauge Group:} & SU(3) \times U(1)_{em} \times [U(1)_{B-L}] \\ {\bf Matter Content:} & 2(3)_{-\frac{1}{3}}\ +\ 2(\bar3)_{\frac{1}{3}} +\ 1(3)_{\frac{2}{3}}\ +\ 1(\bar3)_{-\frac{2}{3}}\ + 2(1)_{-1}\ +\ 2(1)_{1} \\ & +\ (1)_{0}+(1)_{0}\ +\ [(1)_{1}+(1)_{0}+(1)_{-1}] \end{array}$ Thus at this level, with only $b_1+c^*$ recombination, we are left with two D-quarks, one U-quark, two charged leptons and one (Dirac) neutrino. In addition there are charged and neutral Higgsinos (last three states in brackets). Comparing this massless spectrum with the one expected from the field theory Yukawa couplings with a vev for the Higgs ${\bar H}$, we see that one extra D-quark and one charged lepton became massive after recombination, which was not expected from the effective field theory Lagrangian. It seems that this puzzling result can be understood as follows. Consider the brane intersection $bc^*$ giving rise the the Higgs field ${\bar H}$ (see table \[espectrosm\]). In addition to the ${\bar H}$ massless chiral multiplet, there are at this intersection other [*massive*]{} vector-like pairs with the opposite $Q_b$ and $Q_c$ charges to ${\bar H}$. These are in some way $\cn = 4$ partners of the massless Higgs multiplet (see section (4.1) of ref.[@cim1]). Thus at the intersection nothere are massive $Y=-1/2$ fields $H'$ with $Q_{b,c}$ charges qual to $(1,1)$. Note that such $H'$ fields have precisely the gauge quantum numbers required to couple to $Q_L$ and $D_R$ (see table \[espectrosm\]), so that if a vev is induced for $H'$ a D-quark (and a charged lepton) will become massive. Thus what it seems is happening here is that when the branes $b_1$ and $c^*$ recombine at this intersection it is not only the lightest (tachyonic) scalar which is involved but also massive excitations. Recall however that our recombination arguments are purely topological in character, they tell us who becomes massive and who remains massless, but it does not tells us how big are the masses. It is reasonable to expect that since massive modes are involved, the masses of these D-quark and charged lepton are small, so that they can perhaps be identified with the d-quark and the electron. As in the square quiver example, further recombination of the spectator brane $b_2^*$ with the brane $e$, into a final brane $f$, $b_2^* + e\rightarrow f$ renders massive all of the fermions. Indeed, using bilinearity of the intersection numbers it is easy to check that all intersection numbers vanish. Still an electromagnetic generator U(1)\_[em]{} = [ [U(1)\_a]{}6 ]{}  - [ [U(1)\_d]{}2]{}  + [ [ U(1)\_[f]{}]{} 2]{}  . can be checked to remain in the massless spectrum, i.e., it does not receive any mass from couplings to B-fields. Thus the electroweak symmetry breaking process is completed. We have not addressed here the question of what triggers electroweak symmetry breaking in the triangle quiver. The Higgs multiplets are massless to start with but the scalars get in general masses from loops and FI-terms, as explained in the previous section. In addition there is the one-loop contribution from top-quark loops which will tend to induce a negative mass$^2$ to the ${\bar H}$ higgs in the usual way [@ir]. We are assuming here that those effects combined yield a negative $mass^2$ to the Higgs fields. Breaking of B-L and neutrino masses from brane recombination ------------------------------------------------------------ We have seen that in all of the four classes of models constructed (except for the square quiver with $\Delta \not=0$) there is an extra gauge boson corresponding to $B-L$ in the massless spectrum. This extra $U(1)$ may be Higgssed in a variety of ways but perhaps the simplest would be to give a vev to some right-handed sneutrino, which are always present in these models. We would like to discuss in this section how this process would be interpreted in the brane recombination language and how it implies the appearance of Majorana neutrino masses [^11]. Let us consider for definiteness the triangle quiver model. Its massless chiral spectrum is the one of the MSSM with right-handed neutrinos. In addition to hypercharge we have the $B-L$ generator at the massless level. The breaking of this extra gauge symmetry could be obtained if some right-handed sneutrino ${\tilde {\nu }}_R$ gets a vev. The stringy counterpart of this process would be the recombination of branes $c$ and $d$ in the model. So let us assume there is a recombination $c + d \rightarrow j$ into a final brane $j$. We will be left with three stacks of branes now $a$,$b$ and $j$. Using the results in (\[intersicsm\]) and the linearity of intersection numbers one gets [lcl]{} I\_[ab]{} =  1, & & I\_[ab\_2\^\*]{} = 2,\ I\_[aj]{} =  -3, & & I\_[aj\^\*]{} = -3,\ I\_[bj]{} =  -1 - , & & I\_[bj\^\*]{} = 2 - ,\ I\_[jj\^\*]{} =  -6. \[intersec8\] One can then easily check that there is only one $U(1)$ with no couplings to RR-fields and hence remains massless. It is given by U(1)\_Y = [ [U(1)\_a]{}6 ]{}  -  [ [U(1)\_j]{}2]{}  . and corresponds to standard hypercharge. The resulting fermionic spectrum is shown in table \[espectrosm2\], for the case $\b^2 = 1$. Note that this fermion spectrum is the one of the SM [*without right-handed neutrinos*]{}. The breaking of electro-weak symmetry may then proceed as in the previous subsection. Doublet Higgs scalars may be provided both by the scalar partners of the fermionic doublets in the table but also they could be provided by the $H,{\bar H}$ fields of the original model in table \[espectrosm\] if they remained relatively light upon recombination, which is something which will depend on the detailed geometry of the recombined branes. Note also that in the recombination process not only $U(1)_{B-L}$ has acquired a mass but also the right-handed neutrinos have dissapeared from the massless spectrum. In addition, compared to the spectrum in table \[espectrosm\], a pair of $SU(2)_L$ doublets have gained masses. The latter was expected from the point of view of the effective field-theory Higgs mechanism. Looking at the superpotential in (\[yuki2\]) one observes that a vev for ${\tilde {\nu }}_R$’s mixes leptons with Higgsinos. On the other hand the fact that the ${ {\nu }}_R$’s get massive is less obvious from the effective field theory point of view, since there are apparently no renormalizable Yukawa couplings giving (Majorana) masses to them. As in the case of the unexpected massive fermions upon EW symmetry breaking discussed in the previous section, an understanding of this fact seems to imply that brane recombination involves in the process also the effects of massive fields. There are massive chiral fields with opposite charges to those of the ${\tilde {\nu }}_R$’s at the intersections. In general dim=5 couplings of type $(\nu_R\nu_R{\tilde {\nu }}_R^*{\tilde {\nu }}_R^*)$ will give masses to $\nu_R$’s once the sneutrinos get vevs [^12]. Whatever the low-energy field theory interpretation, it is a fact that the right-handed neutrinos get massive upon $c+d$ brane recombination. This is interesting in itself since it has always been difficult to find mechanisms in string-theory giving rise to Majorana masses for neutrinos. It seems that brane recombination gives one possible answer. One interesting question is what happens now with the left-handed neutrinos. The gauge group of the above model after this recombination is just $SU(3)\times SU(2)\times U(1)_Y$. Once the ${\tilde {\nu }}_R$’s get vevs, lepton number is violated and there is no longer distinction between sleptons and Higgs fields [^13]. Since there are no $\nu_R$’s left in the massless spectrum one may argue that there are no possible Dirac fermions for neutrinos and an effective field theory analysis would suggest that the left-handed neutrinos should remain massless after electroweak symmetry breaking. The brane recombination analysis tells us that this will not be the case and after full brane recombination of the branes involved in lepton number violation [*and*]{} electroweak symmetry breaking: b\_2 + b\_1\^\* + c + d f there are no massless fermions left. Indeed, using bilinearity of intersection numbers plus eq.(\[intersicsm\]) it is easy to check that $I_{af}=I_{af^*}=I_{ff^*}=0$. So somehow left-handed neutrinos have managed to become massive also after full brane recombination. As in previous cases a possible field theory interpretation of the neutrino masses is the importance of the massive states at the intersections. Dimension 5 operators of the form $(LL{\bar H}{\bar H})$ can give Majorana masses to left-handed neutrinos if the scalars in ${\bar H}$ get a vev. An important point would be to know the size of neutrino masses after the full recombination/Higgsing process. Unfortunately the intersection numbers are topological numbers which count the net number of massless fermions but do not give as any information on the size of the masses of the non-chiral fermions. The masses (Yukawa couplings) are geometrical quantities which depend on the precise locations of the wrapping branes. Note in particular that although the initial $a,b,c,d$ stacks of branes of this model are factorizable branes with intersections respecting the same supersymmetry, after any recombination takes place the resulting recombined brane is in general non-factorizable and we do not know the precise shape of the cycle that the brane is wrapping. Thus we cannot compute in detail aspects like Yukawa couplings unless we get that geometrical information. On the other hand it is reasonable to expect that, if the ${\tilde {\nu }}_R$-vevs are of order the string scale, since the left-handed neutrinos get their masses involving massive modes of order $M_s$, see-saw-like Majorana masses of order $m_{\nu}\propto |H|^2/M_s$, with a model dependent coefficient which will depend on the geometry and can perfectly be rather small. So neutrino masses within the experimental indications could be obtained. In the scheme we are discussing, after full recombination of the [*left* ]{}, [*right*]{} and [*lepton*]{} branes into a single brane, the observed SM particles and interactions would come from open string exchange between a couple of brane stacks (see fig.\[artistic\] for an artist’s view): [*Baryonic stack*]{}. It contains three parallel branes (and mirrors) and the gauge group $SU(3)\times U(1)_a$, with $U(1)_a$ gauging baryon number. QCD gauge bosons originate in this stack. [*Electro-Weak stack (f)* ]{}. It contains only one brane (plus mirror) with a $U(1)_{EW}$ gauge boson in its worldvolume. It results from the recombination of [*left* ]{}, [*right*]{} and [*lepton*]{} branes. A linear combination of $U(1)_a$ and $U(1)_{EW}$ may be identified with the photon and it is massless. The orthogonal combination gets massive by combining with a RR B-field, but its symmetry remains as a global symmetry in perturbation theory, guaranteeing proton stability, as in the examples in previous sections. The $W^{\pm}$,$Z^0$ bosons correspond to massive string states stretching between different sectors of the electroweak brane ($f$) or its mirror ($f^*$). The masses of fermions of the SM will depend on the detailed geography of the configuration. All quarks and leptons are massive but some of their masses may be very small due to the fact that they are located at different distances from the location of the Higgs fields, and some Yukawa couplings are naturally exponentially suppressed, as discussed in ref.[@afiru2]. Thus every SM parameter would have a reflection in terms of the detailed geometry of the underlying [*baryonic*]{} and [*electro-weak*]{} stacks. Final comments and conclusions ============================== Let us make a number of general comments about the explicit intersecting D6-brane models constructed in the previous sections. They consist of Type IIA string compactifications on $T^6$ along with an orientifold operation. We have concentrated on configurations with four stacks of D6-branes : $a$ ([*baryonic*]{}), $b$ ([*left*]{}), $c$ ([*right*]{}), and $d$ ([*leptonic*]{}). This is the simplest structure capable of giving rise to all fermions of the SM in bifundamental representations of the underlying group. By varying the geometry of the tori, one can obtain models in which all brane intersections respect [*some* ]{} $\cn = 1$ supersymmetry. There are four classes of models with four stacks of branes with SUSY-quiver structures shown in fig.(\[quivers\]). They are called respectively square, linear, rombic and triangle quivers depending on the SUSY-quiver structure. We have built explicit models corresponding to these four classes. Some general characteristics of these models are displayed in table \[propmodels\]. The four classes of models have quarks, leptons and Higgs multiplets respecting in general different SUSY’s, as recorded in the table. In general, in addition to the SM gauge group, the property of Q-SUSY (or SUSY) seems to force the presence of an additional $U(1)$ generator corresponding to $U(1)_{B-L}$, an abelian symmetry well known from left-right symmetric models. This property is quite intriguing since there is in principle no obvious connection between the SUSY properties of a theory and the gauge groups one is gauging. In the case of the square quiver one can give mass to that additional $U(1)$ by departing from the SUSY limit (i.e. $\Delta \not=0$). The models from the square and linear quiver have non-SUSY Higgs sectors. If, as suggested in ref.[@cim1], one insists in getting one-loop protection of the Higgs particles to stabilize the weak scale, models with triangle or rombic quiver structure would be required. In particular, in the case of the triangle quiver, specific configurations with the minimal SUSY Higgs sector can be obtained. The latter triangle models have (for $\beta^2 = 1$) the chiral content of the MSSM (with right-handed neutrinos). From this point of view the triangle class of models are particularly attractive. They are also attractive because they predict the presence of a SM Higgs sector. Indeed we showed how, in order for left- and right-handed multiplets to respect the same $\cn = 1$ SUSY, brane configurations are forced to have intersections at which Higgs sectors arise. This is an interesting property which is not present in schemes which follow the unification route like SUSY-GUT, $CY_3$ and/or Horava-Witten heterotic compactifications etc. [^14]. On the other hand, we have seen that going slightly away from the Q-SUSY limit, one can parametrize the corresponding SUSY-breaking in terms of FI-terms. As remarked in the table, one can check that in the square and linear quivers all squarks and sleptons can get positive $(mass)^2$ from FI-terms whereas that is not the case for the triangle and rombic quivers. In the latter cases the leading source of SUSY-breaking scalar masses should come from loop effects, as in fig.(\[loop5\]). The general consistency of any of these D6-brane configurations require global cancellation of the total RR charge. In the case of the square quiver it is easy to find simple choices of D6-branes wrapping factorized cycles on the torus and with no intersection with the SM branes, so that all RR-tadpoles cancel. The other three types of quivers are equally consistent from this point of view, although in general non-factorizable extra branes and additional RR H-flux may be required to cancel tadpoles. In D6-brane models like this, the standard way [@aadd] to lower the string scale $M_s$ compared with the Planck scale $M_{Planck}$ by making large some of the torus radii cannot be performed [^15]. Note also that the intersecting brane structure is not necessarily linked to a low string scale (i.e., $M_s\approx $ 10-100 TeV) hypothesis. Consider in particular a triangle quiver configuration, where all intersections preserve the [*same* ]{} $\cn = 1$ SUSY. In principle one can consider this triangle as part of some bigger $\cn = 1$ brane configuration somewhat analogous to the class of models in ref.[@csu]. One could then translate most of our discussion (FI-terms, brane recombination, gauge coupling constants) on triangle quivers to that configuration. This is what we mean by indicating in the table that the string scale for the triangle quiver case is $\sim 10$ TeV-$M_{Planck}$. Independently of the particular models discussed, we have presented a brane interpretation of the SM Higgs mechanism. It is important to realize that the familiar brane interpretation of the Higgs mechanism in terms of the separation of parallel branes is not appropriate for the Higgs mechanism of the SM. Brane separation does not lower the rank of the gauge group and corresponds to adjoint Higgsing. We claim that the appropriate brane interpretation of the SM Higgs mechanism (and analogous Higgsings lowering the rank) is brane recombination [^16] . In our approach the non-Abelian weak interactions $SU(2)_L$ live on the worldvolume of two parallel branes i.e., $b_1$,$b_2$. If the Higgs field comes from open strings exchanged between branes $b$ and $c^*$ (as in the models constructed), what will happen is that one of the two parallel $b$-branes (say, $b_1$) will fuse with $c^*$ giving rise to a single brane $b_1+c^*\rightarrow e$. This is the brane recombination mechanism. Since each brane comes along with its own $U(1)$, it is obvious that the rank has been reduced in the process. At the same time the number of chiral fermions (computed from the intersection numbers of the residual branes left) may be shown to decrease (or vanish), corresponding to the Higgs field giving masses to chiral fermions through Yukawa couplings. An analogous brane recombination interpretation exists for a process in which a right-handed sneutrino gets a vev. In this case lepton number is broken and at the same time it is shown that both right-handed and left-handed neutrinos get independent (i.e., Majorana masses). If one consider this process in the specific models constructed (like the square or the triangle models) one finds that the distinction between Higgs multiplets and sleptons dissappear, and the properties of the models are somewhat similar to R-parity violating SUSY models with lepton number violation. In this respect we note that it seems quite difficult within the brane intersection scheme to have lepton number violating neutrino masses without having at the same time L-violating dimension-four couplings, both things come along once brane recombination takes place. At the end of the day, in these schemes the observed SM would have a description in terms of two final recombined brane-stacks : a [*baryonic*]{} stack and an [*electroweak stack*]{} supporting a $SU(3)\times U(1)_{em}$ gauge group. Every SM parameter would have a reflection in terms of the detailed geometry of the underlying [*baryonic*]{} and [*electro-weak* ]{} stacks. **Acknowledgements** We are grateful to G. Aldazábal, C. Kokorelis, R. Rabadán and A. Uranga for useful discussions. The research of D.C. and F.M. was supported by the Ministerio de Educación, Cultura y Deporte (Spain) through FPU grants. This work is partially supported by CICYT (Spain) and the European Commission (RTN contract HPRN-CT-2000-00148). Appendix I: Branes wrapping general cycles. =========================================== As we have mentioned in the text, when dealing with more general intersecting D-brane constructions or considering a generic configuration after some brane recombination has taken place, we are naturally led to consider branes wrapping general cycles. In our particular setup, a D$6_a$-brane wrapping a general cycle will be located in a 3-submanifold of $T^6 = T^2 \times T^2 \times T^2$, thus corresponding to an element $[\Pi_a]$ of $H_3(T^6, \IZ)$, which is the group of homology classes of 3-cycles [@nakahara]. It turns out that $H_3(T^6, \IZ)$ is a discrete vector space, so any of its elements can be represented by a vector with integer entries. This vector space has dimension 20. One particular subset of $H_3(T^6, \IZ)$ is given by what we have called [*factorizable cycles*]{}. These are 3-cycles that can be expressed as products of 1-cycles on each $T^2$ (see figure \[compact2\] for an example)[^17]. Any of those cycles can be expressed by 6 integers as $[\Pi_a] = \prod_{i = 1}^3 [(n_a^i, m_a^i)]$, where $n_a^i, m_a^i \in \IZ$ describe the 1-cycle the D$6_a$-brane is wrapping on the $i^{th}$ torus. Factorizable cycles can be easily described geometrically, which allows us to compute many phenomenologically interesting quantities. For instance, we can compute the lightest bosonic spectrum living at the intersection of two factorizable D6-branes $a$ and $b$ by simply computing the angles they form on each torus and using (\[scalars\]). Thanks to this, we were able to check whether certain $\cn = 1$ supersymmetries were preserved at each intersection and define Q-SUSY or SUSY models in this way. Unfortunately, we will be unable to study which supersymmetries are left, if any, after a brane recombination process has taken place. Due to this fact, we have constructed our particle physics models mostly using branes wrapping factorizable cycles, although non-factorized branes will be generically unavoidable after brane recombination. It turns out that factorizable 3-cycles are not a vector subspace of the homology group $H_3(T^6, \IZ)$. Indeed, the sum of two factorizable cycles is not, in general, a factorizable cycle. This is an important point, since the homology class $[\Pi_a]$ where a D$6_a$-brane lives determines its RR charges. When two branes $a$ and $b$ fuse into a third one $c$ in a brane recombination process the total RR charge should be conserved, which implies that the final brane will lie in a 3-cycle such that $[\Pi_c] = [\Pi_a] + [\Pi_b]$. Even if we start with a configuration where every brane is factorizable, we are finally led to consider non-factorizable branes as well. To this regard, we will consider the smallest vector subspace of $H_3(T^6, \IZ)$ that contains factorizable 3-cycles. This is $[H_1(T^2, \IZ)]^3$, and its dimension is $2^3 = 8$. Following [@torons], we define a basis on this subspace by $q$ comp. 3-cycle factor. comp. ----------- ----------------------------------- --------------- $q_1$ $[a_1] \times [a_2] \times [a_3]$ $n^1 n^2 n^3$ $q_2$ $[b_1] \times [b_2] \times [b_3]$ $m^1 m^2 m^3$ $q_3$ $[a_1] \times [b_2] \times [b_3]$ $n^1 m^2 m^3$ $q_4$ $[b_1] \times [a_2] \times [a_3]$ $m^1 n^2 n^3$ $q_5$ $[b_1] \times [a_2] \times [b_3]$ $m^1 n^2 m^3$ $q_6$ $[a_1] \times [b_2] \times [a_3]$ $n^1 m^2 n^3$ $q_7$ $[b_1] \times [b_2] \times [a_3]$ $m^1 m^2 n^3$ $q_8$ $[a_1] \times [a_2] \times [b_3]$ $n^1 n^2 m^3$ where each element of this basis can be expressed as a product of 1-cycles $[a_i]$, $[b_i]$ of the $i^{th}$ $T^2$. Each general cycle $[\Pi_a]$ under consideration can then be expressed by a vector $\q_a$, whose 8 integer components are defined above. In addition, a factorizable 3-cycle will correspond to vector $\q$ whose components are given in the third column above [^18]. See [@torons] for more details on this construction and some other features involving non-factorizable cycles. The usefulness of this $\q$-basis formalism comes from the fact that quantities as the intersection number of two branes $a$ and $b$ can be easily expressed as bilinear products involving an intersection matrix. That is, it can be expressed as I\_[ab]{} = \_a\^[ t ]{}I \_b, \[numinter\] where the intersection matrix is given by I = ( [cccccccc]{} 0 & 1 & & & & & &\ -1 & 0 & & & & & &\ & & 0 & 1 & & & &\ & & -1 & 0 & & & &\ & & & & 0 & 1 & &\ & & & & -1 & 0 & &\ & & & & & & 0 & 1\ & & & & & & -1 & 0\ ) \[matrizD6\] When dealing with orientifold compactifications each generic brane $a$ must be accompanied by its mirror image $a^*$. In this $q$-basis formalism, the corresponding vectors can be related under the action of a linear operator $\Om$, such that \_a = \_[a\^\*]{}, \^2 = [Id]{}. \[accionom\] As we already mentioned, the geometrical action associated to $\Om$ amounts to a reflection on each complex internal dimension. In terms of a 1-cycle $(n^i, m^i)$ wrapping on the $i^{th}$ $T^2$, this translates into : (n\^i, m\^i) (n\^i, - m\^i - 2b\^i n\^i), \[action\] where $b^i = 0, \oh$ is the T-dual discrete NS background defined in Section 2, and related to the complex structure of the $i^{th}$ torus. From this we can deduce the action of the operator $\Om$ on a general vector $\q$ describing a D6-brane: = ( [cccccccc]{} 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0\ -8b\^[1]{}b\^[2]{}b\^[3]{} & -1 & -2b\^[1]{} & -4b\^[2]{}b\^[3]{} & -2b\^[2]{} & -4b\^[1]{}b\^[3]{} & -2b\^[3]{} & -4b\^[2]{}b\^[1]{}\ 4b\^[2]{}b\^[3]{} & 0 & 1 & 0 & 0 & 2b\^[3]{} & 0 & 2b\^[2]{}\ -2b\^[1]{} &0 & 0 & -1 &0 &0 &0 &0\ 4b\^[1]{}b\^[3]{} &0 &0 & 2b\^[3]{} & 1 & 0 &0 &2b\^[1]{}\ -2b\^[2]{} &0 &0 &0 & 0 & -1 &0 &0\ 4b\^[1]{}b\^[2]{} &0 &0 &2b\^[2]{} &0 &2b\^[1]{} & 1 & 0\ -2b\^[3]{} &0 &0 &0 &0 &0 & 0 & -1\ ) Notice that this linear operator should leave fixed the cycle where the O6-plane lies, which is \[\_[ori]{}\] = \_[i=1]{}\^3 ([1 1 - b\^[i]{}]{}\[a\_i\] - 2 b\^[i]{} \[b\_i\]), \[cicloori\] and that can be easily translated into a vector $\q_{ori}$. The chiral massless spectrum arising at general cycles intersections is given by [c]{} \_[a&lt;b]{}\ \_a \[specori2\] where ${\bf S}_a$ (${\bf A}_a$) stands for the (anti)symmetric representation of the gauge group $U(N_a)$. Here $\b^i = 1 - b^i$, as defined in the text. All intersections involved in (\[specori2\]) can be computed in this $q$-basis formalism, as for instance I\_[ab]{} = \_a\^[ t ]{}I \_b, I\_[a,ori]{} = \_a\^[ t ]{}I \_[ori]{}. Remarkably enough, RR tadpoles cancellation have an extremely simple expression in this formalism \_a N\_a (\_a + \_a ) = Q\_[ori]{} \_[ori]{} \[tadpolesq\] where $Q_{ori}$ represents the relative charge between the O6-plane and the D6-branes and is given by $Q_{ori} = 32 \b^1\b^2\b^3$. Notice that we can define the operators $P_\pm = \oh (1 \pm \Om)$, which satisfy P\_\^2 = P\_, & & P\_P\_= 0, \[proyectores\] thus being projector operators on the $q$-basis space. Notice that the projector $P_+$ is involved in condition (\[tadpolesq\]), which means that only some of the components of $\q_a$ are relevant for tadpoles. On the other hand, the projector $P_-$ is involved on the coupling to branes of the RR $B_2$ fields that mediate the generalized GS mechanism (see Section 2 and [@afiru; @imr] for a proper definition of these fields). Indeed, a general D6$_{a}$-brane whose vector is $\q_a$ will couple to these antisymmetric four-dimensional fields as $P_- \q_a$. Notice that since $P_-$ is a projector, it will only couple to four fields. In the same way that was done for factorizable branes in (\[caplillos\]), we can give an explicit expression for these couplings in terms of the components of the vector $\q$ of this brane. [l]{} B\_2\^0  : N\_a ( b\^1b\^2b\^3 q\_1 + q\_2 + b\^1 q\_3 + b\^2b\^3 q\_4 + b\^2 q\_5 + b\^1b\^3 q\_6 + b\^3 q\_7 + b\^1b\^2 q\_8 ) F\_a,\ B\_2\^1  : N\_a ( b\^1 q\_1 + q\_4 ) F\_a,\ B\_2\^2  : N\_a ( b\^2 q\_1 + q\_6 ) F\_a,\ B\_2\^3  : N\_a ( b\^3 q\_1 + q\_8 ) F\_a. \[caplillos5\] In order to relate these expressions with the ones presented in the text, let us recall that in (\[caplillos\]), we were expressing our D6-brane configurations in terms of [*fractional*]{} 1-cycles. These fractional 1-cycles are defined as [@bkl; @cim1] (n\^i,m\^i)\_[[frac]{}]{} (n\^i,m\^i) + b\^[i]{} (0,n\^i), \[fwrapping\] so when computing the coupling of a fractional brane to, say, the field $B_2^1$, we have the coefficient N\_a ( b\^1 q\_1 + q\_4 ) = N\_a ( b\^1 n\^1n\^2n\^3 + m\^1n\^2n\^3 ) = N\_a m\^1\_[[frac]{}]{} n\^2\_[[frac]{}]{} n\^3\_[[frac]{}]{}, \[demofrac\] so it reduces to the previous expression. Notice that, apart from this appendix, we have used the fractional notation on the whole text, without any subindex. Appendix II: Extra $U(1)$’s and Q-SUSY structure ================================================ One of the most interesting aspects regarding intersecting brane world models involves the massive $U(1)$ structure, arising from couplings with antisymmetric $B_2$ fields, as shown in Section 2 and more extensively in [@imr]. In particular, we are interested in the abelian gauge symmetries that remain after those couplings have been taken into account. When dealing with Q-SUSY models of factorizable branes there are some general results that can be stated regarding such massless $U(1)$’s. Indeed, in this second appendix we will try to elucidate the number of massless $U(1)$’s in terms of the different Q-SUSY structures presented in Section 3. Let us start from a generic brane content consisting of four stacks of factorizable D6-branes, which contains each of the SUSY-quivers considered in Section 3. This brane content is presented in table \[generic\]. Notice that we must impose $m_\g^2\cdot m_\g^3 = 0$ for the brane $\g$ to belong to the hexagonal structure depicted in figure \[hexagon\]. We should also impose this condition in order to avoid chiral exotic matter appearing in the $cc^*$ sector. This brane content will yield the most general Q-SUSY quiver of four stacks of factorizable branes, modulo renumbering of the tori. However, as we already mentioned in the text, without loss of generality we can take the branes $\a$, $\b$ to be of type $a_2$, $b_2$, respectively. This amounts to take $\eps_\a = \eps_\b = -1$ in table \[generic\]. Given a specific brane content, we can easily compute which of the $U(1)$’s will remain massless in our configuration by looking at the couplings (\[caplillos\]). For our purposes, it will be useful to encode such information in matrix notation. In general, for a configuration of $K$ stacks of branes we can define the $\B$ as a $4 \times K$ matrix, containing on each column the coupling of the $i^{th}$ brane to the four $B_2$ fields. When dealing with factorizable branes, such matrix has the form = ( [c]{} …\ [c]{} N\_i m\_i\^1 m\_i\^2 m\_i \^3\ N\_i m\_i\^1 n\_i\^2 n\_i \^3\ N\_i n\_i\^1 m\_i\^2 n\_i \^3\ N\_i n\_i\^1 n\_i\^2 m\_i \^3\ [c]{} …\ ). \[B\] Given a general linear combination of $U(1)$ fields, whose generator is Q\_X = \_[j = 1]{}\^K c\_X\^j Q\_j, \[combi\] it will remain as a massless linear combination of the low energy spectrum whenever it does not couple to any of the $B_2$ fields, that is, if \_X = 0, \_X\^[ t]{} = ( c\_X\^j  ). \[kernel\] So we can see that each massless combination of $U(1)$’s corresponds to a vector $\q_X$ belonging to the kernel of $\B$, now seen as a linear operator. In particular the number of massive $U(1)$’s on each configuration equals [*Rank*]{}($\B$). This simple observation will help us to elucidate how many $U(1)$’s remain massless on each Q-SUSY configuration arising from table \[generic\]. Let us distinguish two different cases - [* $m_\g^3\ =\ 0$*]{} This choice contains both rombic and triangular quiver structures. Our $\B$ matrix will have the form = ( [cccc]{} 0 & 0 & 0 & 0\ 0 & N\_ m\_\^1 n\_\^2 n\_\^3 & N\_ m\_\^1 n\_\^2 n\_\^3 & 0\ N\_ n\_\^1 m\_\^2 n\_\^3 & 0 & \_\^2 N\_ n\_\^1 m\_\^2 n\_\^3 & N\_ n\_\^1 m\_\^2 n\_\^3\ - N\_ n\_\^1 n\_\^2 m\_\^3 & - N\_ n\_\^1 n\_\^2 m\_\^3 & 0 & \_N\_ n\_\^1 n\_\^2 m\_\^3\ ), \[B2\] from which it can easily be seen that the rank of $\B$ will be always lower than 4, so at least one $U(1)$ remains massless. Now, having a Q-SUSY structure of any sort will imply some topological restrictions[^19], such as = [m\_\^3/n\_\^3 m\_\^2/n\_\^2]{}. Indeed, if we define \_1 = [n\_\^2 m\_\^3 m\_\^2 n\_\^3]{}, & & \_2 = [n\_\^1 m\_\^3 m\_\^1 n\_\^3]{}, \[defs\] then it can easily be seen that having a Q-SUSY structure implies $\B$ taking the form = ( [cccc]{} 0 & 0 & 0 & 0\ 0 & N\_ m\_\^1 n\_\^2 n\_\^3 & N\_ m\_\^1 n\_\^2 n\_\^3 & 0\ N\_ n\_\^1 m\_\^2 n\_\^3 & 0 & \_\^2 (\_2/\_1) N\_ m\_\^1 n\_\^2 n\_\^3 & N\_ n\_\^1 m\_\^2 n\_\^3\ - \_1 N\_ n\_\^1 m\_\^2 n\_\^3 & - \_2 N\_ m\_\^1 n\_\^2 n\_\^3 & 0 & \_\_1 N\_ n\_\^1 m\_\^2 n\_\^3 ). The minor determinants of this matrix will be proportional to [l]{} [det]{}\_[(1,2,3)]{} 1 + \_\^2\ [det]{}\_[(1,2,4)]{}, [det]{}\_[(1,3,4)]{} 1 + \_\ [det]{}\_[(2,3,4)]{} \_\^2\_-1 \[minors\] For a triangular quiver we should impose $\eps_\g^2 = \eps_\d = -1$, so every minor determinant will vanish and the rank of $\B$ will be two. Thus, we will have precisely 2 massless surviving $U(1)$’s, at least if none of the entries in (\[B2\]) vanishes. In any case, when trying to build either Standard or Left-Right symmetric models, one extra abelian group will arise. When dealing with a rombic quiver, however, we should impose instead $\eps_\g^2 = - \eps_\d = -1$, and as a result there will be two nonvanishing minor determinants. Thus, we will have just one massless $U(1)$. The other choices of phases correspond to some other SUSY-quivers not considered in this paper. - [* $m_\g^2\ =\ 0$*]{} This second choice will contain square and linear Q-SUSY structures. Proceeding in the same manner as done above, we find that in order to have a Q-SUSY structure our $\B$ matrix should take the form = ( [cccc]{} 0 & 0 & 0 & 0\ 0 & N\_ m\_\^1 n\_\^2 n\_\^3 & N\_ m\_\^1 n\_\^2 n\_\^3 & 0\ N\_ n\_\^1 m\_\^2 n\_\^3 & 0 & 0 & N\_ n\_\^1 m\_\^2 n\_\^3\ - \_1 N\_ n\_\^1 m\_\^2 n\_\^3 & - \_2 N\_ m\_\^1 n\_\^2 n\_\^3 & \_\^3 \_2  N\_ m\_\^1 n\_\^2 n\_\^3 & \_\_1 N\_ n\_\^1 m\_\^2 n\_\^3 ), the minor determinants now being proportional to [l]{} [det]{}\_[(1,2,3)]{}, [det]{}\_[(2,3,4)]{} 1 + \_\^3\ [det]{}\_[(1,2,4)]{}, [det]{}\_[(1,3,4)]{} 1 + \_ \[minors2\] The square quiver case amounts to take $\eps_\g^3 = \eps_\d = 1$, which implies [*Rank*]{}($\B$) = 3 and just one massless $U(1)$. For the Linear quiver, in turn, we must take $\eps_\g^3 = - \eps_\d = 1$, again with the same result. Apart from these general considerations, let us notice that, when trying to get the Standard Model from a general orientifold configuration whith the brane content of table \[generic\], the hypercharge generator will have an associated vector of the form \_Y = ( [c]{} 1/6\ 0\ / 2\ /2 ), where $\eps$, $\tilde\eps$ are model-dependent phases. For this vector to belong to the kernel of the general matrix in (\[B2\]), we must impose, among others, the condition $m_\g^1 n_\g^2 n_\g^3 = 0$. Now, notice that $m_\g^1 \neq 0$, or else there will be no intersection with branes $\a$ and $\d$, so we are finally led to consider $n_\g^2 = 0$ or $n_\g^3 = 0$. Since this will imply that the brane $\g$ has a twist vector v\_= (\_, 2, 0) v\_= (\_, 0, 2), \[rtwist\] then in order to belong to one of the types of branes in (\[vectors\]), we must impose $\th_\g = \frac \pi2$, which in turn implies that $n_c^1 = 0$. As a result, the brane $\g$ will not couple to any of the $B_2$ fields in (\[caplillos\]), as can be seen by direct substitution in (\[B\]). The $\B$ matrix will be effectively reduced to the $3 \times 3$ minor $(1,2,4)$, and its rank will depend exclusively on $\eps_\d$. Indeed, when $\eps_\d = -1$ we find that there are two massless $U(1)$’s whose generators are [c]{} Q\_,\ (N\_ n\_\^1 m\_\^2 n\_\^3) Q\_- (N\_ n\_\^1 m\_\^2 n\_\^3) Q\_, \[massless2\] whereas in the case $\eps_\d = 1$ only $Q_\g$ remains massless. [99]{} M. R. Douglas and G. W. Moore, [*“D-branes, Quivers, and ALE Instantons,”*]{} hep-th/9603167.\ M. R. Douglas, B. R. Greene and D. R. Morrison, [*“Orbifold resolution by D-branes,”*]{} Nucl. Phys. B [**506**]{}, 84 (1997), hep-th/9704151. G. Aldazábal, L. E. Ibáñez, F. Quevedo and A. M. Uranga, [*“D-branes at singularities: A bottom-up approach to the string embedding of the standard model,”*]{} JHEP [**0008**]{}, 002 (2000), hep-th/0005067. G. Aldazábal, L. E. Ibáñez, F. Quevedo, [*“Standard-like models with broken supersymmetry from type I string vacua,”*]{} JHEP [**0001**]{}, 031 (2000), hep-th/9909172.   [*“A D-brane alternative to the MSSM,”*]{} JHEP [**0002**]{}, 015 (2000), hep-ph/0001083.\ M. Cvetic, A. M. Uranga and J. Wang, [*“Discrete Wilson lines in N = 1 D = 4 type IIB orientifolds: A systematic exploration for Z(6) orientifold,”*]{} Nucl. Phys. B [**595**]{}, 63 (2001), hep-th/0010091. D. Berenstein, V. Jejjala and R. G. Leigh, [*“The standard model on a D-brane,”*]{} Phys. Rev. Lett.  [**88**]{}, 071602 (2002), hep-ph/0105042. D. Bailin, G. V. Kraniotis and A. Love, [*“Searching for string theories of the standard model,”*]{} hep-th/0108127.\ L. L. Everett, G. L. Kane, S. F. King, S. Rigolin and L. T. Wang, [*“Supersymmetric Pati-Salam models from intersecting D-branes,”*]{} hep-ph/0202100. L. F. Alday and G. Aldazábal, [*“In quest of ’just’ the standard model on D-branes at a singularity,”*]{} JHEP [**0205**]{}, 022 (2002), hep-th/0203129. M. Berkooz, M. R. Douglas and R. G. Leigh, [*“Branes intersecting at angles,”*]{} Nucl. Phys. B [**480**]{}, 265 (1996), hep-th/9606139;\ H. Arfaei and M. M. Sheikh Jabbari, [*“Different D-brane interactions,”*]{} Phys. Lett. B [**394**]{}, 288 (1997) hep-th/9608167.\ V. Balasubramanian and R. G. Leigh, [*“D-branes, moduli and supersymmetry,”*]{} Phys. Rev. D [**55**]{}, 6415 (1997), hep-th/9611165.\ M. M. Sheikh Jabbari, [*“Classification of different branes at angles,”*]{} Phys. Lett. B [**420**]{}, 279 (1998), hep-th/9710121. R. Blumenhagen, L. Görlich, B. Körs and D. Lüst, [*“Noncommutative compactifications of type I strings on tori with magnetic flux”,*]{} JHEP [**0010**]{}, 006 (2000), hep-th/0007024.   [*“Magnetic flux in toroidal type I compactification,”*]{} Fortsch. Phys.  [**49**]{}, 591 (2001), hep-th/0010198. G. Aldazábal, S. Franco, L. E. Ibáñez, R. Rabadán and A. M. Uranga, [*“D = 4 chiral string compactifications from intersecting branes,”*]{} J. Math. Phys.  [**42**]{}, 3103 (2001), hep-th/0011073. G. Aldazábal, S. Franco, L. E. Ibáñez, R. Rabadán and A. M. Uranga, [*“Intersecting brane worlds,”*]{} JHEP [**0102**]{}, 047 (2001), hep-ph/0011132. R. Blumenhagen, B. Körs and D. Lüst, [*“Type I strings with F- and B-flux,”*]{} JHEP [**0102**]{}, 030 (2001), hep-th/0012156. L. E. Ibáñez, F. Marchesano and R. Rabadán, [*“Getting just the standard model at intersecting branes,”*]{} JHEP [**0111**]{}, 002 (2001), hep-th/0105155.\ L. E. Ibáñez, [*“Standard Model Engineering with Intersecting Branes,”*]{} hep-ph/0109082. R. Blumenhagen, B. Körs, D. Lüst and T. Ott, [*“The standard model from stable intersecting brane world orbifolds,”*]{} Nucl. Phys. B [**616**]{}, 3 (2001), hep-th/0107138.   [*“Intersecting brane worlds on tori and orbifolds,”*]{} hep-th/0112015. M. Cvetic, G. Shiu and A. M. Uranga, [*“Three-family supersymmetric standard like models from intersecting branes ”*]{}, Phys. Rev. Lett.  [**87**]{}, 201801 (2001), hep-th/0107143.   [*“Chiral four-dimensional N = 1 supersymmetric type IIA orientifolds from intersecting branes”*]{} Nucl. Phys. B [**615**]{}, 3 (2001), hep-th/0107166.   [*“Chiral type II orientifold constructions as M theory on G(2) holonomy spaces ”*]{}, hep-th/0111179. D. Cremades, L. E. Ibáñez and F. Marchesano, [*“SUSY quivers, intersecting branes and the modest hierarchy problem,”*]{} hep-th/0201205. D. Bailin, G. V. Kraniotis and A. Love, [*“Standard-like models from intersecting D4-branes,”*]{} Phys. Lett. B [**530**]{}, 202 (2002), hep-th/0108131.\ H. Kataoka and M. Shimojo, [*“SU(3) x SU(2) x U(1) chiral models from intersecting D4/D5 branes,”*]{} hep-th/0112247.\ G. Honecker, [*“Intersecting brane world models from D8-branes on $(T^2 \t T^4/Z_3)/\Om\car_1$ type IIA orientifolds,”*]{} JHEP [**0201**]{}, 025 (2002), hep-th/0201037. R. Blumenhagen, B. Körs and D. Lüst, [*“Moduli stabilization for intersecting brane worlds in type 0’ string theory,”*]{} Phys. Lett. B [**532**]{}, 141 (2002), hep-th/0202024. J. García-Bellido, R. Rabadán and F. Zamora, [*“Inflationary scenarios from branes at angles,”*]{} JHEP [**0201**]{}, 036 (2002), hep-th/0112147.\ R. Blumenhagen, B. Körs, D. Lüst and T. Ott, [*“Hybrid inflation in intersecting brane worlds,”*]{} hep-th/0202124. R. Blumenhagen, L. Görlich and B. Körs, [*“Supersymmetric 4D orientifolds of type IIA with D6-branes at angles,”*]{} JHEP [**0001**]{}, 040 (2000), hep-th/9912204.   [*“A new class of supersymmetric orientifolds with D-branes at angles,”*]{} hep-th/0002146.\ S. Förste, G. Honecker and R. Schreyer, [*“Supersymmetric $Z_N \times Z_M$ Orientifolds in 4D with D-Branes at Angles,”*]{} Nucl. Phys. B [**593**]{}, 127 (2001), hep-th/0008250.   [*“Orientifolds with branes at angles,”*]{} JHEP [**0106**]{}, 004 (2001), hep-th/0105208.\ G. Honecker, [*“Non-supersymmetric orientifolds with D-branes at angles,”*]{} hep-th/0112174. C. Bachas, [*“A Way to Break Supersymmetry,”*]{} hep-th/9503030. M. Bianchi, G. Pradisi and A. Sagnotti, [*“Toroidal compactification and symmetry breaking in open string theories,”*]{} Nucl. Phys. B [**376**]{}, 365 (1992).\ M. Bianchi, [*“A note on toroidal compactifications of the type I superstring and other superstring vacuum configurations with 16 supercharges,”*]{} Nucl. Phys. B [**528**]{}, 73 (1998) hep-th/9711201.\ E. Witten, [*“Toroidal compactification without vector structure,”*]{} JHEP [**9802**]{}, 006 (1998), hep-th/9712028.\ C. Angelantonj, [*“Comments on open-string orbifolds with a non-vanishing B(ab),”*]{} Nucl. Phys. B [**566**]{}, 126 (2000), hep-th/9908064.\ . Angelantonj and R. Blumenhagen, [*“Discrete deformations in type I vacua,”*]{} Phys. Lett. B [**473**]{}, 86 (2000) hep-th/9911190.\ Z. Kakushadze, [*“Geometry of orientifolds with NS-NS B-flux,”*]{} Int. J. Mod. Phys. A [**15**]{}, 3113 (2000), hep-th/0001212. R. Blumenhagen, L. Görlich, B. Körs and D. Lüst, [*“Asymmetric orbifolds, noncommutative geometry and type I string vacua,”*]{} Nucl. Phys. B [**582**]{}, 44 (2000), hep-th/0003024.\ C. Angelantonj, I. Antoniadis, E. Dudas and A. Sagnotti, [*“Type-I strings on magnetised orbifolds and brane transmutation,”*]{} Phys. Lett. B [**489**]{}, 223 (2000), hep-th/0007090.\ C. Angelantonj and A. Sagnotti, [*“Type-I vacua and brane transmutation,”*]{} hep-th/0010279. A. M. Uranga, [*“D-brane, fluxes and chirality,”*]{} JHEP [**0204**]{}, 016 (2002), hep-th/0201221. K. Dasgupta, G. Rajesh and S. Sethi, [*“M theory, orientifolds and G-flux,”*]{} JHEP [**9908**]{}, 023 (1999), hep-th/9908088.\ S. B. Giddings, S. Kachru and J. Polchinski, [*“Hierarchies from fluxes in string compactifications,”*]{} hep-th/0105097.\ S. Kachru, M. Schulz and S. Trivedi, [*“Moduli stabilization from fluxes in a simple IIB orientifold,”*]{} hep-th/0201028. S. Kachru and J. McGreevy, [*“Supersymmetric three-cycles and (super)symmetry breaking,”*]{} Phys. Rev. D [**61**]{}, 026001 (2000), hep-th/9908135.\ M. Mihailescu, I. Y. Park and T. A. Tran, [*“D-branes as solitons of an N = 1, D = 10 non-commutative gauge theory,”*]{} Phys. Rev. D [**64**]{}, 046006 (2001), hep-th/0011079.\ E. Witten, [*“BPS bound states of D0-D6 and D0-D8 systems in a B-field,”*]{} JHEP [**0204**]{}, 012 (2002), hep-th/0012054.\ R. Blumenhagen, V. Braun and R. Helling, [*“Bound states of D(2p)-D0 systems and supersymmetric p-cycles,”*]{} Phys. Lett. B [**510**]{}, 311 (2001), hep-th/0012157. M. Dine, N. Seiberg and E. Witten, [*“Fayet-Iliopoulos Terms In String Theory,”*]{} Nucl. Phys. B [**289**]{}, 589 (1987). J. J. Atick, L. J. Dixon and A. Sen, [*“String Calculation Of Fayet-Iliopoulos D Terms In Arbitrary Supersymmetric Compactifications,”*]{} Nucl. Phys. B [**292**]{}, 109 (1987).\ M. Dine, I. Ichinose and N. Seiberg, [*“F Terms And D Terms In String Theory,”*]{} Nucl. Phys. B [**293**]{}, 253 (1987). L. E. Ibáñez and G.G. Ross, [*“$SU(2)\times U(1)$ symmetry breaking as a radiative effect of supersymmetry breaking in GUT’s,”*]{} Phys. Lett. B [**110**]{}, 215 (1982). J. D. Lykken, [*“Weak Scale Superstrings,”*]{} Phys. Rev. D [**54**]{}, 3693 (1996), hep-th/9603133.\ N. Arkani-Hamed, S. Dimopoulos and G. R. Dvali, [*“The hierarchy problem and new dimensions at a millimeter,”*]{} Phys. Lett. B [**429**]{}, 263 (1998), hep-ph/9803315.\ I. Antoniadis, N. Arkani-Hamed, S. Dimopoulos and G. R. Dvali, [*“New dimensions at a millimeter to a Fermi and superstrings at a TeV,”*]{} Phys. Lett. B [**436**]{}, 257 (1998), hep-ph/9804398.\ G. Shiu and S. H. Tye, [*“TeV scale superstring and extra dimensions,”*]{} Phys. Rev. D [**58**]{}, 106007 (1998), hep-th/9805157. M. Nakahara, [*“Geometry, Topology And Physics,”*]{}, [*Bristol, UK: Hilger (1990) 505 p. (Graduate student series in physics)*]{}. R. Rabadán, [*“Branes at angles, torons, stability and supersymmetry,”*]{} Nucl. Phys. B [**620**]{}, 152 (2002), hep-th/0107036. [^1]: This is in contrast with other popular embeddings of the SM like SUSY-GUT, $CY_3$ or Horava-Witten heterotic compactifications etc., in which the presence of light SM Higgs multiplets is somewhat ad-hoc and misterious. [^2]: Note that the SM Higgs mechanism cannot be described by the familiar process in which two parallel branes separate. Brane separation does not lower the rank and corresponds to adjoint Higgsing, which is not what the SM Higgs mechanism requires. [^3]: See Appendix I for a discussion of non-factorizable cycles. [^4]: If we added such a D6-brane $H$ in order to complete an [*anomalous* ]{} configuration, its net intersection number would not vanish for some brane $a$ contained in it. Clearly, in this case we would not be allowed to call this extra brane a [*hidden brane*]{}. [^5]: In all the brane intersection models discussed in this article the baryon number symmetry $U(1)_a$ is automatically gauged. Although the corresponding generator becomes eventually massive, baryon number remains as an accidental symmetry in perturbation theory [@imr]. Thus the proton is perturbatively stable. [^6]: As noted in [@imr], if in addition one has $n_c^1 = 0$, then an extra generator $(1/3)Q_a-Q_d$ (which corresponds to B-L) remains also massless. [^7]: From the model-building point of view, this is required in order to avoid massless chiral multiplets in the $D6_aD6_{a^*}$ sector, which does only provide us with matter transforming as symmetric and antisymmetric representations. We will forbid such kind of exotic spectrum right from the start. [^8]: In a D-brane language, the twist vector $v_a$ contains the relative angles between the D6-brane $a$ and the O6-plane, who lies on the cycle $(1/\b^1,0)(1/\b^2,0)(1/\b^3,0)$. Hence, it encodes the supersymmetries shared by both. If we name the generators of the $\cn =4$ algebra of the O6-plane by $r_1, r_2, r_3, r_4$, then the six types of branes described above will correspond to different choices of a $\cn = 2$ subalgebra. [^9]: In fact in some sense they do, since tree-level closed string couplings are one-loop from the open string channel point of view. [^10]: In order to compute the chiral fermion spectrum from an intersection of a brane $j$ with its mirror $j^*$ one should use the general formula (\[specori2\]) of Appendix I. Since in the present case $I_{j,ori}=0$, the number of chiral fermions is just half the intersection number. This is also true for the spectrum in table (\[espectrosm2\]). [^11]: Note that in the models of ref.[@imr] and in the subset of them formed by the models of section (5.1) in the present paper, there is no extra $U(1)_{B-L}$ which should be spontaneously broken. Thus it is not required to spontaneously break lepton number by giving a vev to a ${\tilde {\nu }}_R$. In such a case there can only be Dirac neutrino masses, as argued in ref.[@imr]. In models with an extra $U(1)_{B-L}$ we will be forced to break lepton number conservation anyhow and at some level Majorana masses will appear, as discussed below. [^12]: In addition there will be mixing of the $\nu_R$’s with the gauginos of the $U(1)_{B-L}$ generator. [^13]: This is rather similar to R-parity violating models with lepton number violation. [^14]: Also note that in those unification schemes the stability of the proton relies on the asumed presence of symmetries like R-parity or generalizations as well as some doublet-triplet splitting mechanism. In the present case proton stability is a consequence of the gauge character of baryon symmetry. [^15]: For a discussion of this point see [@bgkl; @afiru; @cim1] and references therein. [^16]: An analogous proposal in the context of D4-brane models was put forward in ref.[@afiru2]. [^17]: Notice that for this definition to make sense we also need our $T^6$ to be factorized as $T^2 \times T^2 \times T^2$. [^18]: In the construction of the $q$-basis we are not considering fractional wrapping numbers, so both $n$ and $m$ are integer, see below. [^19]: Notice that, although supersymmetry between two stacks of branes always implies a geometrical condition given by the angles they form, the [*ability*]{} for a full configuration of branes to be Q-supersymmetric does imply some topological restrictions.
{ "pile_set_name": "ArXiv" }
--- address: - 'Petersburg Nuclear Physics Institute of NRC “Kurchatov Institute”, Gatchina 188300, Russia' - 'St. Petersburg Electrotechnical University LETI, Prof. Popov Str. 5, 197376 St. Petersburg, Russia' - 'Helmholtz Institute Mainz, Johannes Gutenberg University, 55099 Mainz, Germany' - 'Department of Physics, University of California at Berkeley, Berkeley, California 94720-7300, USA' - 'Nuclear Science Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA' author: - 'Mikhail G. Kozlov [^1]' - Dmitry Budker title: Sensitivity coefficients to variation of fundamental constants --- Transition frequencies in atoms and molecules depend on the values of the fundamental constants (FC). The energies of all stationary states of these systems are proportional to the atomic energy unit $E_h=4\pi\hbar c R_\infty=m_e e^4/\hbar^2$ and additionally depend on the fine structure constant $\alpha=e^2/(\hbar c)$ and the electron-to-proton mass ratio $\mu=m_e/m_p$. Here $m_e$ and $m_p$ are the electron and proton masses, $e$ is the elementary-charge magnitude, $\hbar$ is Plank’s constant, $R_\infty$ is Rydberg constant, and $c$ is the speed of light. In principle, all energy levels depend on the nuclear radii (isotope field shifts), but this dependence is usually weak. The hyperfine energy additionally depends on the nuclear moments, most importantly, on the $g$ factors and quadrupole moments of the nuclei. All nuclear parameters, in turn, depend on the QCD energy scale $\Lambda_\mathrm{QCD}$ and on quark masses. When we search for a possible temporal variation of FC we are studying time-variation of atomic (molecular) frequencies. It is difficult to imagine how to compare directly the frequency today with itself yesterday. Using additional devices, like resonators, or delay lines, introduces time-evolution of the frequency caused, for example, by a possible change of the resonator size, or shape. Instead, we need to measure the ratio of two frequencies and look how this ratio varies in time. This ratio changes only when two frequencies have different dependences on FC. Clearly, the dependence on the atomic energy unit cancels out. Because of that, it is convenient to use atomic units when discussing the sensitivity to a variation of FC. The dependence of the atomic optical transitions between different electronic configurations on the mass ratio $\mu$ and on the nuclear radii and moments is weak. We conclude that frequency ratios for atomic optical transitions depend mainly on $\alpha$. This dependence appears due to the relativistic corrections to the energy. Such corrections scale as $\alpha^2 Z^2\times\mathrm{IP}$, where $Ze$ is nuclear charge and IP is the ionization potential. For light atoms $Z\approx 1$ and the corrections are small. For heavy atoms, where $Z\gg 1$, and/or for highly charged ions, where $\mathrm{IP}\gg E_h$, the relativistic corrections become large. Consequently, $\alpha$-dependence of the optical transitions is weak for light atoms and strong for heavy atoms and highly charged ions. For molecules, due to the vibrational and rotational energy-level structure, the transition frequencies generally depend on both $\alpha$ and $\mu$. Because of the interplay of different contributions to the molecular energy, the spectra of molecules are much richer and occasionally their dependence on FC may be significantly enhanced. Dimensional & dimensionless sensitivity coefficients {#dimensional-dimensionless-sensitivity-coefficients .unnumbered} ---------------------------------------------------- In 1999 Dzuba, Flambaum, and Webb [@DFW99b; @DFW99a] introduced $q$ factors, which define how atomic energies depend on $\alpha$: $$\delta E = 2q_\alpha\frac{\delta \alpha}{\alpha_0}, \label{eq:q}$$ where $\alpha_0$ is the present value of the fine structure constant. The $q$ factor has dimension of energy. A transition frequency $\hbar\omega_{i,f}=E_f-E_i$ depends on $\alpha$ as: $${\hbar}\delta \omega_{i,f} % = 2\left(q_{\alpha\,f}-q_{\alpha\,i}\right)\frac{\delta \alpha}{\alpha_0} % \equiv 2q_{\alpha\,i,f}\frac{\delta \alpha}{\alpha_0} = 2\left(q_{\alpha}^f-q_{\alpha}^i\right)\frac{\delta \alpha}{\alpha_0} \equiv 2q_{\alpha}^{i,f}\frac{\delta \alpha}{\alpha_0} \,. \label{eq:q_if}$$ It is often convenient to introduce a dimensionless sensitivity coefficient $K_\alpha$ for an $i \to f$ transition: $$\frac{\delta \omega}{\omega_0} = K_\alpha \frac{\delta \alpha}{\alpha_0}\,, \qquad K_\alpha=\frac{2q_\alpha}{\omega_0}\,, \label{eq:K}$$ where we skip the indexes $i,f$ for brevity. In order to find $K_\alpha$, we first calculate the parameters $q_\alpha$ for both levels and then insert $q_\alpha=q_\alpha^f-q_\alpha^i$ in the second equation of . Similarly to Eqs. (\[eq:q\],\[eq:K\]) we can define the sensitivity coefficients to $\mu$-variation: $$\begin{aligned} \delta E = q_\mu\frac{\delta \mu}{\mu_0}, \label{eq:q_mu} \qquad \frac{\delta \omega}{\omega_0} = K_\mu \frac{\delta \mu}{\mu_0}\,, \qquad K_\mu=\frac{q_\mu}{\omega_0}\,. %\label{eq:K_mu}\end{aligned}$$ Note, that there is an extra factor of 2 in Eq.  compared to the first equation in , which was introduced in [@DFW99b; @DFW99a] because atomic energies depend on $\alpha^2$. Let us consider the relevance of the coefficients $q$ and $K$ for various experiments. According to Eqs. (\[eq:q\],\[eq:q\_mu\]) the factors $q$ give the absolute energy (frequency) shifts for a given change of the FC. On the other hand, the factors $K$ determine the relative frequency change. Therefore, the latter are important when the relative accuracy of the frequency measurement is fixed. For example, this may be the case, when the lines are Doppler broadened. Then the linewidth is proportional to the transition frequency, $\Gamma_D/\omega = \Delta v/c$, and shifts down to a certain fraction of the linewidth can be experimentally resolved ($\Delta v$ is the width of velocity distribution). This situation is typical for high-redshift astrophysical observations. In high-precision laboratory measurements, on the contrary, Doppler-free spectroscopy and optical combs are often used. Then the experiments are characterized by the absolute accuracy of the frequency measurements. In this case, the $q$ factors are more relevant. If there is an accidental degeneracy of two energy levels $i$ and $f$ with different $q$ factors, then according to Eq.  the transition sensitivity coefficient $K^{i,f}$ is inversely proportional to the frequency $\omega_{i,f}$ and is strongly enhanced.[^2] However, an actual experimental sensitivity to the FC variation is enhanced *only if* the transition frequency $\omega_{i,f}$ is directly measured with high relative accuracy. The sensitivity coefficients $K$ are *irrelevant* for *indirect* measurements, when small frequency $\omega$ appears as a difference of two measured frequencies, $\omega=\omega_a- \omega_b$. Here the absolute experimental error does not depend on $\omega$ and the relative experimental error is inversely proportional to $\omega$. This counterbalances growth of the sensitivity coefficient $K$ [@NBL04]. This is one of the misunderstandings, which may be found in the literature. Dependence on the units and conventions {#dependence-on-the-units-and-conventions .unnumbered} --------------------------------------- The factors $q$ have dimension of energy and clearly depend on the choice of units. In atomic physics, some calculations are done using atomic units while others use relativistic units, where the energy unit is $E_r=m_e c^2$ and $E_h=\alpha^2 E_r$. Consequently, the calculated $q$ factors are different depending on the units used. Moreover, even if we use the same units, the result depends on the choice of the reference energy. For example, for many-electron atoms the energy of a level $i$ can be defined as the core and valence parts, $E_i= E_\mathrm{core}+\tilde{E}_i$, and the core energy $E_\mathrm{core}$ is often dropped. The core energy is typically much larger than the valence energy $\tilde{E}_i$ and, therefore, $|q_i|\gg |\tilde{q}_i|$. Different definitions of the core may lead to very different values of $\tilde{q}_i$. Of course, when we calculate $q$ factors for the transitions, the core contribution cancels out and transition coefficients $q^{i,f}=\tilde{q}^{i,f}$ depend only on the choice of units. In practice, the $q$ factors are usually calculated with respect to the ground state, thus $q^{i}=q^{0,i}$. The dependence of the $q$ factors on units is usually recognized. However, it is often assumed that the dimensionless coefficients $K$ are independent of units. This, in fact, is not the case. According to Eq. , $$K_\alpha = \frac{\partial \omega}{\partial \alpha} \frac{\alpha}{\omega}\,. \label{eq:K_der}$$ If $\omega$ is in atomic units, then in relativistic units the frequency $\tilde{\omega}$ is $$\tilde{\omega}=\alpha^2 \omega\,,$$ and respective sensitivity coefficient is equal to $$\tilde{K}_\alpha =K_\alpha+2\,.$$ As pointed out above, in experiments, we always measure the ratio of two frequencies. For example, in an absolute measurement, we determine the ratio of an atomic frequency to the frequency of the cesium frequency standard. The variation of the frequency ratio due to a variation of $\alpha$ is: $$\frac{\delta\left(\omega_m/\omega_n\right)}{\omega_m/\omega_n} =\left(K_\alpha^m-K_\alpha^n\right) \frac{\delta\alpha}{\alpha}\,. \label{eq:K_ratio}$$ We see that the variation of the ratio of frequencies depends on the difference of the sensitivity coefficients and does not depend on the choice of the units. We conclude, that the values of the sensitivity coefficients $q$ and $K$ depend on the conventions used. However, the differences in the sensitivity coefficients $K$ are independent of the choice of the system of units and the reference frequency. Therefore, it is always necessary to look at the differences in the sensitivity coefficients of the levels and transitions. Lattice constant {#lattice-constant .unnumbered} ---------------- Up to now we focused on the energies and frequencies. This is because frequency measurements are by far the most accurate measurements in physics. However, in the literature there are discussions of a possible variation of the lattice constant of crystals [@StaFla15; @StaFla16a]. The dependence of the lattice constant on the FC can be tested in frequency measurements if one uses a crystal to make an optical cavity.[^3] The frequency of the $N^\mathrm{-th}$ longitudinal mode of the cavity of the length $L_c$ is: $\omega^N_c= \pi N c/L_c$. The sensitivity coefficient $K_\alpha^c$ of this frequency to $\alpha$-variation is linked to the sensitivity coefficient $K_\alpha^a$ of the lattice constant $a$: $$\frac{\delta L_c}{L_c} =\frac{\delta a}{a} =K_\alpha^a \frac{\delta\alpha}{\alpha}\,. \label{eq:K_a}$$ In atomic units $c=\alpha^{-1}$ and therefore, $$\begin{aligned} &\frac{d\omega^N_c}{d\alpha} =\frac{d}{d\alpha}\left(\frac{\pi N}{\alpha L_c}\right) =-\omega^N_c \left(\frac{1}{\alpha}+\frac{1}{L_c}\frac{dL_c}{d\alpha}\right)\,, \quad\Rightarrow \nonumber\\ &K_\alpha^c=-\left(1 + K_\alpha^a\right)\,. \label{eq:K_c}\end{aligned}$$ For crystals made of light elements $|K_\alpha^a|\ll 1$ and $K_\alpha^c\approx -1$. The ratio of some atomic frequency $\omega_\mathrm{at}$ to the cavity frequency has sensitivity to $\alpha$-variation, which is given by Eq. ; $$\frac{\delta\left(\omega_\mathrm{at}/\omega^N_c\right)} {\omega_\mathrm{at}/\omega^N_c} \approx\left(K_\alpha^\mathrm{at}+1\right) \frac{\delta\alpha}{\alpha}\,. \label{eq:K_ratio1}$$ This is particularly relevant in view of the recent high-precision searches for alpha variation involving just such a comparison [@WMBC16; @KOBK18]. Acknowledgement {#acknowledgement .unnumbered} --------------- This work was inspired by discussions with A. Borschevsky, J. Ye, Y. Stadnik, and V. Flambaum. <span style="font-variant:small-caps;"></span> \[1\][\#1]{} [\[1\]]{} <span style="font-variant:small-caps;">V.A. Dzuba</span>, <span style="font-variant:small-caps;">V.V. Flambaum</span>, and <span style="font-variant:small-caps;">J.K. Webb</span> **82**, 888 (1999). <span style="font-variant:small-caps;">V.A. Dzuba</span>, <span style="font-variant:small-caps;">V.V. Flambaum</span>, and <span style="font-variant:small-caps;">J.K. Webb</span> **59**, 230 (1999). <span style="font-variant:small-caps;">A.T. [Nguyen]{}</span>, <span style="font-variant:small-caps;">D. [Budker]{}</span>, <span style="font-variant:small-caps;">S.K. [Lamoreaux]{}</span>, and <span style="font-variant:small-caps;">J.R. [Torgerson]{}</span> **69**, 022105 (2004). <span style="font-variant:small-caps;">Y.V. [Stadnik]{}</span> and <span style="font-variant:small-caps;">V.V. [Flambaum]{}</span> **114**, 161301 (2015). <span style="font-variant:small-caps;">Y.V. [Stadnik]{}</span> and <span style="font-variant:small-caps;">V.V. [Flambaum]{}</span> **93**, 063630 (2016). <span style="font-variant:small-caps;">P. [Wcis[ł]{}o]{}</span>, <span style="font-variant:small-caps;">P. [Morzy[ń]{}ski]{}</span>, <span style="font-variant:small-caps;">M. [Bober]{}</span>, <span style="font-variant:small-caps;">A. [Cygan]{}</span>, <span style="font-variant:small-caps;">D. [Lisak]{}</span>, <span style="font-variant:small-caps;">R. [Ciury[ł]{}o]{}</span>, and <span style="font-variant:small-caps;">M. [Zawada]{}</span> **1**, 0009 (2016). <span style="font-variant:small-caps;">C. Kennedy</span>, <span style="font-variant:small-caps;">E. Oelker</span>, <span style="font-variant:small-caps;">T. Bothwell</span>, <span style="font-variant:small-caps;">D. Kedar</span>, <span style="font-variant:small-caps;">L. Sonderhouse</span>, <span style="font-variant:small-caps;">E. Marti</span>, <span style="font-variant:small-caps;">S. Bromley</span>, <span style="font-variant:small-caps;">J. Robinson</span>, and <span style="font-variant:small-caps;">J. Ye</span> p.H06.00005 (2018). [^1]: Corresponding authorE-mail:  [^2]: This is not generally true for non-accidental degeneracy. Consider, for example, the fine structure splitting. In atomic units it scales as $\alpha^2 Z^2$ and the sensitivity coefficients are $K_\mathrm{fs}=2$ independently of the size of the splitting. [^3]: Amorphous materials are expected to depend on $\alpha$ in the same way as crystals.
{ "pile_set_name": "ArXiv" }
--- author: - The ATLAS Collaboration bibliography: - 'bjetpaper.bib' date: 'December 19, 2011' title: '\' --- Introduction ============ The production of $b$-quarks in proton–proton collisions at the Large Hadron Collider (LHC) provides an important test of perturbative QCD (pQCD). Calculations of the $b$-quark production cross-section have been performed at next-to-leading order (NLO) in pQCD [@Frixione:1997ma]. These calculations can be matched to different partonshower and hadronisation models to produce final states that can be compared to those measured in collision data. Cross-sections for $b$-jet production in high energy $p\overline p$ collisions have been measured at the ${\rm Sp\overline pS}$ [@Albajar:1986iu; @Albajar:1988th] and Tevatron [@Abe:1993hr; @Abe:1993sj; @Abbott:1999se; @Abbott:2000iv] colliders. The experiments measured cross-sections different from those predicted by QCD at the time. This led to substantial improvements in the experimental methods and theoretical calculations. It is therefore of great interest to test the theoretical predictions at the higher centre-of-mass energy provided by the LHC. Moreover, the measurement of the $b$-jet cross-sections is an important ingredient in understanding other processes involving the production of $b$-quarks, which represent substantial backgrounds in many searches for new physics. Measurements of $b$-hadron production at $\sqrt{s} = 7$TeV in the forward region have been reported by LHCb [@Aaij:2010gn] and in the central region by CMS [@Khachatryan:2011hf; @Khachatryan:2011wq]. This paper describes measurements of the inclusive $b$-jet and $b\overline b$-dijet production cross-sections performed with the ATLAS detector at the LHC. Jets are reconstructed from energy clusters in the calorimeter using the [anti-$k_t$]{} algorithm [@Cacciari:2008gp], with jet radius parameter $R=0.4$. The relatively long lifetime of hadrons containing $b$-quarks is exploited to obtain a jet sample enriched in $b$-jets by selecting jets with a reconstructed secondary vertex significantly displaced from the primary vertex. The number of $b$-jets in this enriched sample is derived from a fit to the invariant mass distribution of the charged particle tracks in the secondary vertex, assuming the pion mass for the individual particles. This is referred to as secondary vertex mass hereafter. The inclusive cross-section is measured for jets containing $b$- or ${\ensuremath{\overline{b}}\xspace}$-quarks as a function of the transverse momentum, [$p_{\rm T}$]{}, and rapidity, [$y$]{}, for jets with $20 < {\ensuremath{p_{\rm T}}\xspace}< 400$GeV and $|y| < 2.1$. The requirement $|y| < 2.1$ ensures that jets are contained within the acceptance of the inner tracking detectors. In the kinematic region $30 < {\ensuremath{p_{\rm T}}\xspace}< 140$GeV, muonbased $b$-tagging is used to provide a complementary, and largely independent, cross-section measurement as a function of jet [$p_{\rm T}$]{}. The [$b\overline{b}$]{}-dijet cross-section is measured for the leading and sub-leading jet in the event as a function of the dijet invariant mass, [$m_{\rm jj}$]{}, the azimuthal angle difference between the two jets, $\Delta\phi$, and the angular variable $\chi = \exp{\left|y_1 - y_2\right|}$ for jets with ${\ensuremath{p_{\rm T}}\xspace}> 40$GeV and $|y| < 2.1$. The variable $\chi$ is defined such that the cross-section of $2\to 2$ elastic scattering of point-like massless particles is approximately constant as a function of $\chi \simeq \frac{1 + \cos\theta^*}{1 - \cos\theta^*}$, where $\theta^*$ is the centre-of-mass scattering angle. To measure the cross-sections as a function of $\chi$, an additional acceptance requirement is used that restricts the boost of the dijet system to $|y_{\rm boost}| = \frac{1}{2}|y_1 + y_2| < 1.1$. This reduces the sensitivity to parton distribution function (PDF) uncertainties at small values of $x$, where $x$ is the fraction of the proton’s momentum carried by the parton participating in the hard scattering. The resulting angular distributions provide a test of pQCD that is relatively insensitive to PDF uncertainties. The measured cross-sections are corrected for all experimental effects using simulated events, to allow comparison with theoretical predictions. The data used for these measurements were collected by the ATLAS detector in 2010 and correspond to an integrated luminosity of $34.0 \pm 1.2$pb$^{-1}$. A detailed description of the luminosity determination can be found in Refs. [@Aad:2011dr; @ATLAS-CONF-2011-011]. The ATLAS detector ================== The ATLAS detector [@Aad:2008zzm] consists of an inner tracking system, immersed in a 2T axial magnetic field, surrounded by electromagnetic calorimeters, hadronic calorimeters and a muon spectrometer. The ATLAS reference system has the origin at the nominal interaction point. The $x$- and $y$-axes define the transverse plane, the azimuthal angle $\phi$ is measured around the beam axis, $z$, and the polar angle $\theta$ with respect to the $z$-axis. The pseudorapidity is defined as $\eta = -\ln\left(\tan\left(\theta/2\right)\right)$. The inner detector (ID) has full coverage in $\phi$ and covers the pseudorapidity range $|\eta| < 2.5$. The ID consists of silicon pixel and microstrip detectors, surrounded by a transition radiation tracker (up to $|\eta| = 2.0$). The electromagnetic calorimeter is a lead-liquid argon sampling calorimeter covering $|\eta| < 3.2$. Hadronic calorimetry in the barrel ($|\eta| < 1.7$) is provided by a scintillator tile calorimeter using steel as the absorber material. The end-cap hadronic calorimeter uses liquid argon with copper absorber plates and extends up to $|\eta| = 3.2$. Additional forward calorimeters extend the calorimetric coverage to $|\eta| < 4.9$, outside the acceptance of this measurement. The outer region of the detector is formed by a muon spectrometer that uses a toroidal magnetic field with a bending power of 1.5–5.5Tm in the barrel and 1.0–7.5Tm in the end-caps. Three layers of muon chambers provide precision tracking in the bending plane up to $|\eta| = 2.7$ and the trigger for muons up to $|\eta| = 2.4$. The trigger system uses three consecutive trigger levels to record a selection of interesting events. The first level trigger (L1) is based on custombuilt hardware that processes the data with a fixed latency of $2.5$s. The second level and the event filter, collectively referred to as the high level trigger (HLT), are software-based triggers running on computing farms. Their average execution times are $40$ms and $4$s respectively, with a design output rate of $3$kHz and $200$Hz respectively. Most of the events used in the measurements presented here are selected by the calorimeterbased triggers. At L1, the electromagnetic and hadronic calorimeters are read out using trigger towers with a granularity of $\Delta\phi \times \Delta\eta = 0.1 \times 0.1$, with jet identification based on transverse energy in a sliding window of $4\times 4$ or $8\times 8$ trigger towers. At the beginning of data-taking in 2010 only the L1 triggers were active, while in the later runs the HLT was used to refine the jet selection further. Events containing jets with $20 < {\ensuremath{p_{\rm T}}\xspace}< 40$GeV were triggered using the minimum bias trigger scintillators (MBTS) [@Aad:2010ir]. The MBTS consist of 32 scintillator counters arranged in two discs located at $\pm 3.56$m from the interaction point, covering $2.09 < |\eta| < 3.84$. The hit multiplicity in the MBTS provides a highefficiency trigger for jet events, independent of the jet [$p_{\rm T}$]{}, with negligible bias. Monte Carlo samples and theoretical predictions =============================================== Simulated events produced by the [Pythia]{} 6.423 [@Sjostrand:2006za] event generator are used for the baseline comparisons and to evaluate corrections. [Pythia]{}implements leading-order (LO) pQCD matrix elements for $2 \to 2$ processes, [$p_{\rm T}$]{}-ordered partonshowers calculated in a leading-logarithmic approximation and an underlying event simulation using multi-parton interactions. It uses the Lund string model [@Andersson:1983ia] for hadronisation. All events were generated using a specially tuned set of parameters denoted as AMBT1 [@Aad:2010ir] with MRST LO$^{*}$ [@Sherstnev:2007nd] parton-density functions. The generated particles are passed through a full simulation [@Aad:2010wqa] of the ATLAS detector and trigger based on GEANT4 [@Agostinelli:2002hh]. Finally, the simulated events are reconstructed and selected using the same analysis chain as is used for the collision data, with the same trigger and event selection criteria. The flavour of jets is defined by matching jets to hadrons with ${\ensuremath{p_{\rm T}}\xspace}> 5$GeV. The jet is considered a $b$-jet if a $b$-hadron is found within $\Delta R = \sqrt{\Delta\phi^2 + \Delta\eta^2} = 0.3$ of the jet axis; otherwise, if a $c$-hadron is found within the same distance the jet is labeled as a $c$-jet. All other jets are considered lightflavour jets. The measured cross-sections are compared to NLO predictions derived using [POWHEG]{} [@Nason:2004rx; @Frixione:2007vw; @Alioli:2010xd; @Frixione:2007nw] and [MC@NLO]{} [@Frixione:2002ik; @Frixione:2003ei], both using the MSTW 2008 NLO PDFs [@Martin:2009iq] and a $b$-quark mass of 4.95GeV. To perform the partonshowering, [POWHEG]{}is interfaced to [Pythia]{} 6 and [MC@NLO]{}to [Herwig]{} 6 [@Corcella:2000bw]. For Herwig, the AUET1 [@ATL-PHYS-PUB-2010-014] tune is used. In contrast to [Pythia]{}, [Herwig]{}uses an angular-ordered partonshower model and a cluster hadronisation model. Event and jet selection ======================= The events used in the lifetimebased analysis are triggered by the L1 or HLT jet triggers, with the exception of the $20 < {\ensuremath{p_{\rm T}}\xspace}< 40$GeV bin in the inclusive cross-section measurement where the MBTS trigger is used. The trigger efficiency for $b$-jets using these trigger selections is estimated to be above 97% in all cases and typically close to 100%. For the muonbased cross-section measurement the combination of a jet and a muon trigger is required, which results in an efficiency ranging from about 35% for jets with ${\ensuremath{p_{\rm T}}\xspace}< 50$GeV to 65% for jets with ${\ensuremath{p_{\rm T}}\xspace}> 105$GeV. While this efficiency is lower, the different trigger prescale factors allocated a much higher rate to the jet-muon trigger than to the inclusive jet triggers for a similar jet [$p_{\rm T}$]{}threshold. Quality selections are applied to the reconstructed jets to ensure that they are not produced by poorly calibrated detector regions or noisy calorimeter cells [@ATLAS-CONF-2010-038]. Additionally, the charged particle tracks contained in the jets are required to be of adequate quality for $b$-tagging [@ATLAS-CONF-2010-042] and a good reconstructed primary vertex is required that contains at least 10 tracks with ${\ensuremath{p_{\rm T}}\xspace}> 150$MeV. The combined efficiency of the reconstruction and the quality requirements is determined to be above 96% for $b$-jets. The secondary vertex $b$-tagging algorithm used, SV0 [@ATLAS-CONF-2010-042], aims at reconstructing the position of the displaced vertex from the charged decay products of long-lived particles in a jet. The SV0 algorithm reconstructs two-track vertices from tracks inside a cone of $\Delta R = 0.4$ around the jet axis that are significantly displaced from the primary vertex, based on the threedimensional impact parameter significance. Quality requirements are applied to the two-track vertices to reject vertices that are compatible with the primary vertex, are located at a radius consistent with one of the pixel detector layers or contain tracks that have an invariant mass consistent with a $K_S^0$ meson, a $\Lambda^0$ baryon or a photon conversion. A single secondary vertex is then fitted to all the tracks which contribute to any of the remaining two-track vertices in the jet. The signed decay length significance of the secondary vertex, $L/\sigma_L$, is used to select a jet sample enriched in $b$-jets. The sign of the decay length is given by the sign of the projection of the decay length vector onto the jet axis. Jets with $L/\sigma_L > 5.85$ are referred to as $b$-tagged jets. The selection at 5.85 is chosen such that it produces a 50% $b$-tagging efficiency for $b$-jets in simulated [$t\overline{t}$]{}events. $b$-tagging efficiency ---------------------- The efficiency of the chosen selection on $L/\sigma_L$ is estimated with a data-driven method that uses jets containing a muon. The number of $b$-jets before and after $b$-tagging can be obtained using the variable [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}, which is defined as the momentum of the muon transverse to the combined muon plus jet axis. Muons originating from $b$-hadron decays have a harder [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}spectrum than muons in $c$- and lightflavour jets. Templates of the [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}shape are constructed for each jet flavour separately. The templates for $b$- and $c$-jets are extracted from Monte Carlo simulation, while the lightflavour template is obtained from a light-jet enriched data sample. These are then fitted [@Barlow:1993dm] to the measured [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}spectrum of muons in jets to obtain the fraction of $b$-jets before and after requiring a $b$-tag. The fit determines the relative contributions of the $b$-, $c$- and lightflavour templates such that their sum best describes the shape of the [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}distribution in data. Having obtained the flavour composition of jets containing muons from the [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}fits, the $b$-tagging efficiency is defined as $$e_b^{\rm data} =\frac{f_b^{\rm tag}\cdot N^{\rm tag}}{f_b\cdot N}\cdot C \,,$$ where $f_b$ and $f_b^{\rm tag}$ are the fractions of $b$-jets before and after $b$-tagging is applied, and $N$ and $N^{\rm tag}$ are the total number of jets in those two samples. The factor $C$ corrects the efficiency for biases introduced by differences between data and simulation in the modelling of the $b$-hadron direction and by heavyflavour contamination of the [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}template for lightflavour jets. The magnitude of these corrections is typically a few percent. Examples of fits to the [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}distribution before and after the $L/\sigma_L$ requirement are shown in Fig. \[fig:ptrel\_fits\]. The [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}method can be used to determine the $b$-tagging efficiency for $b$-jets containing $b$-hadrons that decay semileptonically. Studies have been performed to show that this determination can be extended to all $b$-jets and a systematic uncertainty due to this generalization is assigned to the $b$-tagging efficiency for all $b$-jets. A detailed account of the systematic uncertainties in the $b$-tagging efficiency calibration is given in Ref. [@ATLAS-CONF-2011-089]. The discriminating power of the [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}method decreases with increasing jet [$p_{\rm T}$]{}, hence this method can only provide a data-driven determination of the $b$-jet tagging efficiency for jet [$p_{\rm T}$]{}values up to about 140GeV. For jets with ${\ensuremath{p_{\rm T}}\xspace}> 140$GeV, the $b$-tagging efficiency is derived from simulation and multiplied by a correction factor of $0.88 \pm 0.18$ that accounts for the difference between data and simulation observed in the [$p_{\rm T}$]{}range 90–140GeV. Comparisons between data and simulation as a function of jet [$p_{\rm T}$]{}show that the simulation models the data equally well in all regions of jet [$p_{\rm T}$]{}in which data measurements are available, so the above extrapolation is well motivated. Moreover, detailed comparisons between data and simulation as a function of jet [$p_{\rm T}$]{}, in terms of the quantities that affect $b$-tagging, show that the effect of any mismodelling of the $b$-tagging performance at higher jet [$p_{\rm T}$]{}values is within the systematic uncertainties assigned to the $b$-tagging efficiency. The efficiency after applying the requirement of $L/\sigma_L > 5.85$ ranges from 20% for $b$-jets of ${\ensuremath{p_{\rm T}}\xspace}< 40$GeV and $|y| > 1.2$ to 55% for central $b$-jets with [$p_{\rm T}$]{}of about 100GeV. The [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}distribution can also be used as a discriminant variable to measure the inclusive $b$-jet cross-section directly. While this method is statistically limited and cannot be used beyond 140GeV, as mentioned above, it does provide a useful cross-check for the lifetimebased measurement. Many of the systematic uncertainties are different and the sample of jets used is statistically largely independent from that used in the lifetimebased measurement. The muon-based cross-section measurement is described in Section \[sec:results\]. $b$-jet purity -------------- In the lifetimebased measurement, the fraction of $b$-jets in the $b$-tagged sample of jets, referred to as the purity of the sample, is determined by performing a template fit to the secondary vertex mass distribution. The templates for $b$-, $c$- and lightflavour jets are extracted from Monte Carlo simulation. The average invariant mass of a secondary vertex increases when going from lightflavour jets via $c$-jets to $b$-jets, making it possible to separate the flavours by determining the relative fractions of the templates that best describe the vertex mass distribution in data. For the inclusive cross-section measurement, the number of $b$-, $c$- and lightflavour jets is fitted by maximizing a binned likelihood function that takes into account the statistical uncertainties in both the data and the templates. The fit is performed for each [$p_{\rm T}$]{}and [$y$]{}region separately, in vertex mass bins of 200MeV. In the dijet cross-section measurement, the fraction of $b$-jet pairs is determined from a template fit to the sum of the vertex masses of the two $b$-tagged jets. This fit uses two templates: the $b$-template, where both jets are matched to a $b$-hadron in simulation; and a non-$b$ template, where at least one of the two jets is a $c$- or lightflavour jet. In order to reduce the effect of the limited statistics in simulation, a parameterization is used to smooth the templates. The fit is performed for each kinematic region separately. Typical fit results in the inclusive and dijet measurements are shown in Fig. \[fig:fits\]. Results {#sec:results} ======= All the measured cross-sections are corrected for experimental effects using a bin-by-bin correction, so as to represent particlelevel cross-sections of jets containing $b$-hadrons. The correction is obtained from [Pythia]{}simulated dijet events by calculating the cross-sections for both particlelevel $b$-jets (including muons and neutrinos) and reconstructed $b$-jets. The correction factors are derived bin-by-bin in each distribution by taking the ratio of the two cross-sections. Systematic uncertainties ------------------------ The dominant systematic uncertainties in both the inclusive and the dijet cross-section measurements come from the $b$-jet energy scale calibration, and the determination of the $b$-tagging efficiency and purity. The systematic uncertainties, including those on the muonbased measurement which will be discussed in Section \[sec:ptrel\], are summarized in Table \[tab:systematics\]. [llll]{} Syst. uncertainty & Inclusive $b$-jet & [$b\overline{b}$]{}-dijet & Muon-based\ Jet energy scale & 10–20% & 10–20% & 15–20%\ $b$-tagging efficiency & 5–20% & 30–50% & –\ $b$-jet purity fit & 3–8% & 20–30% & 8–18%\ Luminosity & 3.4% & 3.4% & 3.4%\ Other sources & 2% & 2% & 3%\ Jets are calibrated to the hadronic scale using the inclusive jet energy scale calibration [@Aad:2010wv; @ATLAS-CONF-2011-032], which is based on [$p_{\rm T}$]{}- and $\eta$-dependent correction factors derived from Monte Carlo simulation and validated with test beam measurements. The uncertainty on this jet energy scale varies between 2% and 6% depending on the jet [$p_{\rm T}$]{}and rapidity region. For heavy-flavour jets, two studies were performed to estimate additional contributions to the jet energy scale uncertainty that account for flavour-dependent systematic uncertainties. Firstly, the uncertainty on the calorimeter response for $b$-jets due to their different particle composition has been evaluated using single hadron response studies [@ATLAS-CONF-2011-028]. This method compares the relative response of $b$-tagged jets in [$t\overline{t}$]{}events with that of inclusive jets in QCD dijet events. For jets within $|\eta| < 0.8$ and $20 < {\ensuremath{p_{\rm T}}\xspace}< 250$GeV, this difference is found to be negligible ($<$0.5%). Secondly, systematic uncertainties for $b$-jets were studied in Monte Carlo simulation by comparing particlelevel jets to reconstructed jets. The variations that were studied include the modelling of fragmentation, hadronisation, partonshowers and the underlying event, but also variations in soft-physics tunes and the effects of the uncertainty on the material description. The $b$-jet energy scale uncertainty obtained using these two methods is validated in data by comparing the total transverse momentum of the calorimeter jet to that of the charged particle tracks associated to it [@ATLAS-CONF-2011-067]. It is found that there is an additional 2.5% uncertainty on the $b$-jet energy scale with respect to the uncertainty on the energy scale of inclusive jets. This extra uncertainty is added in quadrature. When propagated to the cross-section measurements, this leads to an uncertainty of 10% to 20%, depending on the kinematic region. The most important contributions to the systematic uncertainty on the $b$-tagging efficiency originate from the modelling of muons in jets in the simulation, the generalization of the efficiency from $b$-jets with muons to inclusive $b$-jets, and the limited statistics of the templates used for the [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}fits. More details about the $b$-tagging efficiency uncertainty can be found in Ref. [@ATLAS-CONF-2011-089]. The resulting uncertainty on the cross-sections amounts to between 5% and 20% for the inclusive $b$-jet cross-section, and between 30% and 50% for the dijet cross-section. The systematic uncertainties from the purity fits account for the observed differences between jets in collision data and those in the Monte Carlo simulation used to derive the templates. The uncertainty is derived from studies of the secondary vertex mass distribution in light-jet enriched samples and $b$-jet enriched samples. The light-jet enriched sample is obtained by selecting jets with a negative decay length. For the $b$-jet enriched samples, two methods are used: the first requires another $b$-tagged jet to be present in the event, while the second selects secondary vertices with high track multiplicities. The observed differences in the secondary vertex mass distribution are then used to correct the template shapes and re-evaluate the fits. The difference in the cross-section is found to be between 3% and 8% and this is assigned as the purity fit systematic uncertainty. For the $b\overline{b}$-dijet cross-section, the most important contribution to the systematic uncertainty on the $b\overline b$-fraction is due to the limited template statistics. The effect of the statistical uncertainty of the templates is estimated by varying the shape parameters of the parameterized templates within their uncertainties and re-evaluating the cross-section. The resulting uncertainty is between 20% and 30%. The systematic uncertainty on the luminosity determination is 3.4% [@ATLAS-CONF-2011-011]. The remaining sources of systematic uncertainty, such as the effect of possible differences in the cross-section shapes between data and simulation on the bin-by-bin corrections, differences in the jet energy resolution between data and simulation, the trigger efficiency and the jet selection efficiency, lead to a combined systematic uncertainty of about 2%. The effect of different shower and hadronisation models is included in the jet energy scale uncertainty. The impact of changing the shape of the [$p_{\rm T}$]{}distribution on the bin-by-bin corrections was found to be much less than 1%. Using [Herwig]{}instead of [Pythia]{}to derive the correction factors gives statistically consistent results. Muonbased $b$-jet cross-section {#sec:ptrel} ------------------------------- The [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}method, used for calibrating the $b$-tagging efficiency, is also used to obtain an independent measurement of the inclusive $b$-jet cross-section in the range $30 < {\ensuremath{p_{\rm T}}\xspace}< 140$GeV. This measurement uses jets containing a muon of ${\ensuremath{p_{\rm T}}\xspace}> 4$GeV within a cone of $\Delta R = 0.4$ from the jet axis. The flavour composition of this sample is extracted from a template fit to the muon [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}distribution. The templates for $b$- and $c$-jets are obtained from Monte Carlo simulation. Two data-driven techniques are employed to extract the shape of the muon [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}in lightflavour jets. The first takes the shape from jets with negative decay length in data, which is then corrected for $b$-jet contamination using simulation. The second method uses inclusive jets without a muon; the template is then obtained by converting each track inside the jet into a muon and weighting the resulting [${\ensuremath{p_{\rm T}}\xspace}^{\rm rel}$]{}by a probability to simulate hadron decays in flight. The $b$-jet fraction is evaluated using both methods, taking the average as the central value and assigning the difference between them as a systematic uncertainty. The dominant sources of systematic uncertainties in this measurement are the $b$-jet energy scale (15–20%) and the purity fits (8–18%). Contributions to the purity fit systematics include limited template statistics and uncertainties in the modelling of semileptonic $b$-hadron decays and $b$-fragmentation. The first modelling error is estimated by varying the muon momentum distribution in the rest frame of the $b$-hadron between that measured by DELPHI [@Abdallah:2005cx] and that measured by BABAR [@Aubert:2004td]. The second is measured by varying the fraction of the $b$-jet energy carried by the $b$-hadron by $\pm 5$% and rederiving the $b$-jet templates in the simulation. Apart from the $b$-jet energy scale, the systematic uncertainties are to a large extent specific to the muonbased measurement. This makes the comparison with the lifetimebased cross-section measurement a useful cross-check. Cross-section results and discussion ------------------------------------ The double-differential inclusive $b$-jet cross-section is shown in Fig. \[fig:xsec\] as a function of jet [$p_{\rm T}$]{}in four different rapidity regions. Figure \[fig:intxsec\] shows the single differential cross-section as a function of [$p_{\rm T}$]{}, integrated over the entire rapidity range of $|y| < 2.1$. In the [$p_{\rm T}$]{}range where the lifetimebased and the muonbased measurements overlap, both results are shown. The data are compared to NLO predictions derived with [POWHEG]{}and [MC@NLO]{}. In addition, the data are compared to the [Pythia]{}prediction. [Pythia]{}, as a leading-logarithmic partonshower generator, is not expected to predict the correct normalization. The [Pythia]{}prediction is scaled by a factor $\times 0.67$ in order to match the measured integrated cross-section, allowing a comparison of the cross-section shapes. All three calculations describe the general features of the cross-section reasonably well. ![Inclusive double-differential $b$-jet cross-section as a function of [$p_{\rm T}$]{}for the different rapidity ranges. The data are compared to the predictions of [Pythia]{}, [POWHEG]{}and [MC@NLO]{}. The leading-order [Pythia]{}prediction is scaled $(\times 0.67)$ to the measured integrated cross-section.[]{data-label="fig:xsec"}](inclusive_xsec_pty.pdf){width="48.50000%"} ![Differential $b$-jet cross-section as a function of [$p_{\rm T}$]{}for $b$-jets with $|y| < 2.1$. The data are compared to the predictions of [Pythia]{}, [POWHEG]{}and [MC@NLO]{}. In the region $30 < {\ensuremath{p_{\rm T}}\xspace}< 140$GeV the muon-based cross-section measurement is also shown. For the muon-based measurement only the [POWHEG]{}prediction is shown.[]{data-label="fig:intxsec"}](inclusive_xsec_pt.pdf){width="48.50000%"} To allow for a better comparison between the data and the NLO predictions, Fig. \[fig:xsecratio\] shows the ratio of the measured cross-section to the NLO theory predictions for $|y| < 2.1$ (top) and for each rapidity region separately. The plot for the full rapidity acceptance also allows a direct comparison between the lifetimebased and the muonbased cross-section measurements in the overlapping [$p_{\rm T}$]{}range, indicating a good agreement between the two measurements. Good agreement is also observed between the measured cross-section and the NLO predictions obtained using [[POWHEG]{}+ [Pythia]{}]{}in all rapidity regions. [[MC@NLO]{}+ [Herwig]{}]{}, however, predicts a significantly different behaviour of the double-differential cross section, as shown in Fig. \[fig:xsecratio\_mcatnlo\]. When the cross-section is integrated over the full rapidity acceptance this effect averages out somewhat and [[MC@NLO]{}+ [Herwig]{}]{}shows better agreement with data. It has been checked that the qualitative behaviour remains the same when [POWHEG]{}is interfaced to [Herwig]{}instead of [Pythia]{}, implying that the observed rapidity dependence in [[MC@NLO]{}+ [Herwig]{}]{}is not resulting from the partonshower Monte Carlo program. On the other hand, [[POWHEG]{}+ [Herwig]{}]{}appears to predict a cross-section that is consistently lower than the [[POWHEG]{}+ [Pythia]{}]{}prediction. This would suggest that the deficit of [[MC@NLO]{}+ [Herwig]{}]{}compared to the data in Fig. \[fig:xsecratio\_mcatnlo\], may be partly due to the [Herwig]{}partonshowering. Comparison to the inclusive (all-flavour) jet cross-section measurement [@Aad:2010wv], shows that the fraction of jets containing a $b$-hadron is approximately 5% in the kinematic region where the two measurements overlap, $60 < {\ensuremath{p_{\rm T}}\xspace}< 400$GeV and $|y| < 2.1$. The [$b\overline{b}$]{}-dijet cross-section is shown as a function of dijet mass in Fig. \[fig:dijet-mass\]. It should be noted that nearby [$b\overline{b}$]{}-pairs, as expected for example from gluon splitting, are generally not resolved as separate jets. Also, since the measurement refers to the leading and sub-leading jet in the event, the contribution from gluon splitting is expected to be small. The [$b\overline{b}$]{}-dijet cross-section is compared to [Pythia]{}and the NLO predictions obtained using [POWHEG]{}and [MC@NLO]{}. The [Pythia]{}prediction is again normalized to the measured integrated cross-section, here using a factor of $\times0.85$. The [Pythia]{}normalization is not expected to be the same as that used in the inclusive cross-section, given the different event selection used. All theory predictions show good agreement with the measured cross-section. Figure \[fig:dijet-phi\] shows the fractional [$b\overline{b}$]{}-dijet cross-section as a function of the azimuthal angle between the two jets, $\Delta\phi$. The dijets selected in this measurement show a pronounced back-to-back configuration in the transverse plane that is generally well reproduced by QCD generators. The $b\overline b$-dijet cross-section as a function of the angular variable $\chi$ is shown in Fig. \[fig:dijet-chi\] for dijets with $|y_{\rm boost}| < 1.1$. The $\chi$ distribution is well reproduced by the theoretical calculations. The distribution flattens for large invariant mass values. ![The [$b\overline{b}$]{}-dijet cross-section as a function of dijet invariant mass for $b$-jets with ${\ensuremath{p_{\rm T}}\xspace}>40$GeV and $|y|<2.1$. The data are compared to the MC predictions of [Pythia]{}, [POWHEG]{}and [MC@NLO]{}. The leading-order [Pythia]{}prediction is scaled to the measured integrated cross-section. The shaded regions around the MC predictions reflect the statistical uncertainty only.[]{data-label="fig:dijet-mass"}](dijet_xsec_mass.pdf){width="48.50000%"} ![The [$b\overline{b}$]{}-dijet cross-section as a function of the azimuthal angle difference between the two jets for $b$-jets with ${\ensuremath{p_{\rm T}}\xspace}>40$GeV, $|y|<2.1$ and a dijet invariant mass of ${\ensuremath{m_{\rm jj}}\xspace}> 110$GeV. The data are compared to the theory predictions of [Pythia]{}, [POWHEG]{}and [MC@NLO]{}. The shaded regions around the MC predictions reflect the statistical uncertainty only.[]{data-label="fig:dijet-phi"}](dijet_xsec_phi.pdf){width="48.50000%"} \ In the NLO calculations, the renormalization and factorization scales are set equal to the transverse energy of the hardest parton: $Q^2 = E_{\rm T}^2 = m_b^2 + {\ensuremath{p_{\rm T}}\xspace}^2$. To estimate the potential impact of higher order terms not included in the NLO calculation on the theory predictions, the renormalization scale is varied from half to twice its default value. Similarly, to estimate the impact of the choice of the scale where the PDF evolution is separated from the matrix element, the factorization scale is varied up and down by a factor of two. The effect of each of these variations on the NLO cross-section prediction is estimated using [POWHEG]{}and found to be approximately 20% for all kinematic regions. Finally, the uncertainty on the PDFs is estimated by deriving the NLO predictions using the NNPDF [@Ball:2010de] and CTEQ 6.6 [@Nadolsky:2008zw] PDFs, resulting in a difference of approximately 10% for all kinematic regions. Conclusions =========== The inclusive $b$-jet and [$b\overline{b}$]{}-dijet production cross-sections have been measured in proton–proton collisions at a centre-of-mass energy of 7TeV, using data with an integrated luminosity of 34pb$^{-1}$ recorded by the ATLAS detector. The inclusive $b$-jet cross-section was measured as a function of jet [$p_{\rm T}$]{}in the range $20 < {\ensuremath{p_{\rm T}}\xspace}< 400$GeV and rapidity in the range $|y| < 2.1$. The [$b\overline{b}$]{}-dijet cross-section was measured as a function of dijet invariant mass in the range $110 < {\ensuremath{m_{\rm jj}}\xspace}< 760$GeV, as a function of the azimuthal angle difference and of the angular variable $\chi$. The measurements are dominated by systematic uncertainties, mainly coming from the $b$-jet energy scale and the determination of the $b$-jet tagging efficiency and purity. The measured cross-sections have been compared to next-to-leading order QCD predictions derived using [POWHEG]{}interfaced to [Pythia]{}and [MC@NLO]{}interfaced to [Herwig]{}. The inclusive cross-section measured over $|y| < 2.1$ for $b$-jets identified by the presence of a secondary vertex is compared to a largely independent cross-section measurement that uses muon-based $b$-tagging in the range $30 < {\ensuremath{p_{\rm T}}\xspace}< 140$GeV. The two measurements show good agreement. The inclusive $b$-jet cross-section is found to be in good agreement with the [[POWHEG]{}+ [Pythia]{}]{}prediction over the full kinematic range. [[MC@NLO]{}+ [Herwig]{}]{}, however, predicts a significantly different behaviour of the double-differential cross section that is not observed in the data. The normalized leading-order [Pythia]{}prediction shows broad agreement with the measured cross-section. [[POWHEG]{}+ [Pythia]{}]{}and [[MC@NLO]{}+ [Herwig]{}]{}show good agreement with the measured [$b\overline{b}$]{}-dijet cross-sections, as does the normalized leading-order [Pythia]{}generator. Acknowledgements {#acknowledgements .unnumbered} ================ 1500 We thank CERN for the very successful operation of the LHC, as well as the support staff from our institutions without whom ATLAS could not be operated efficiently. We acknowledge the support of ANPCyT, Argentina; YerPhI, Armenia; ARC, Australia; BMWF, Austria; ANAS, Azerbaijan; SSTC, Belarus; CNPq and FAPESP, Brazil; NSERC, NRC and CFI, Canada; CERN; CONICYT, Chile; CAS, MOST and NSFC, China; COLCIENCIAS, Colombia; MSMT CR, MPO CR and VSC CR, Czech Republic; DNRF, DNSRC and Lundbeck Foundation, Denmark; ARTEMIS, European Union; IN2P3-CNRS, CEA-DSM/IRFU, France; GNAS, Georgia; BMBF, DFG, HGF, MPG and AvH Foundation, Germany; GSRT, Greece; ISF, MINERVA, GIF, DIP and Benoziyo Center, Israel; INFN, Italy; MEXT and JSPS, Japan; CNRST, Morocco; FOM and NWO, Netherlands; RCN, Norway; MNiSW, Poland; GRICES and FCT, Portugal; MERYS (MECTS), Romania; MES of Russia and ROSATOM, Russian Federation; JINR; MSTD, Serbia; MSSR, Slovakia; ARRS and MVZT, Slovenia; DST/NRF, South Africa; MICINN, Spain; SRC and Wallenberg Foundation, Sweden; SER, SNSF and Cantons of Bern and Geneva, Switzerland; NSC, Taiwan; TAEK, Turkey; STFC, the Royal Society and Leverhulme Trust, United Kingdom; DOE and NSF, United States of America. The crucial computing support from all WLCG partners is acknowledged gratefully, in particular from CERN and the ATLAS Tier-1 facilities at TRIUMF (Canada), NDGF (Denmark, Norway, Sweden), CC-IN2P3 (France), KIT/GridKA (Germany), INFN-CNAF (Italy), NL-T1 (Netherlands), PIC (Spain), ASGC (Taiwan), RAL (UK) and BNL (USA) and in the Tier-2 facilities worldwide. The ATLAS Collaboration {#the-atlas-collaboration .unnumbered} =======================
{ "pile_set_name": "ArXiv" }
--- abstract: 'The effects of introducing a harmonic spatial inhomogeneity into the Kalb-Ramond field, interacting with the Maxwell field according to a ‘string-inspired’ proposal made in earlier work are investigated. We examine in particular the effects on the polarization of synchrotron radiation from cosmologically distant (i.e. of redshift greater than 2) galaxies, as well as the relation between the electric and magnetic components of the radiation field. The rotation of the polarization plane of linearly polarized radiation is seen to acquire an additional contribution proportional to the square of the frequency of the dual Kalb-Ramond axion wave, assuming that it is far smaller compared to the frequency of the radiation field.' address: - '[$^1$]{}Centre for Theoretical Studies and Department of Physics, Indian Institute of Technology, Kharagpur 721302, India' - '[$^2$]{}Institute of Mathematical Sciences, Chennai 600113, India' - '[$^{3,4}$]{}Department of Physics, Jadavpur University, Calcutta 700 032, India' author: - 'Sayan Kar[^1]$^1$, Parthasarathi Majumdar[^2]$^2$, Soumitra SenGupta[^3]$^3$, and Saurabh Sur[^4]$^4$' title: 'Cosmic optical activity from an inhomogeneous Kalb-Ramond field' --- Introduction ============ The behaviour of electromagnetic waves in a curved background spacetime with torsion and its cosmological consequences, has been an area of some interest in recent years. This is in view of its prospective implications for the low energy approximation to string theory. One way to investigate these implications is by identifying spacetime torsion [@pmss], with the massless antisymmetric second rank tensor field, i.e., the Kalb-Ramond (KR) field, existing in most supergravity theories and as such in the massless sector of the most viable string theories [@gsw]. Certain physically observable phenomena result from the above analysis: a cosmic optical activity involving the rotation of the plane of polaization of linearly polarized synchrotron rotation from high redshift galaxies investigated recently [@skpm] being an example. The angle through which the polarization plane rotates is shown to be proportional, to leading order in inverse conformal time (which decreases with redshift), to the rate of change of the KR dual axion field. It is also independent of the frequency of radiation. This latter aspect is in contrast to the well-known Faraday rotation, and hence a new phenomenon. Another point to note is that the KR field, argued to be responsible for the effect, has been treated in [@skpm] as a perturbation on the Maxwell equations in a standard Friedmann-Robertson-Walker (FRW) cosmological background with both matter and radiation domination. Thus, it is assumed to have a negligible effect on shaping cosmological background spacetime. One way of thinking about this is to imagine that the KR axion decouples from the radiation or matter (dust) fluid shaping cosmic geometry far prior to dust-photon decoupling, leaving behind a ‘Cosmic KR Background’ which affects incoming radiation from distant galaxies, albeit rather softly. As the universe expands further, this effect will no doubt gradually subside. The point made in [@skpm] is that the effect may yet be observable in this epoch. This viewpoint has received support from [@pdpj] where the the effects of a time-dependent dilaton field are additionally incorporated, while demonstrating that the earlier findings have a degree of robustness. However, in these earlier assays, it is assumed that the axion and dilaton fields are spatially homogeneous, depending only on the conformal time coordinate $\eta$. This is possibly quite justified considering the overall homogeneity of the universe over cosmological distance scales. In this paper, we generalize the scenario in [@skpm] by introducing a spatial inhomogeneity of a particular type into the KR axion field $H$: the massless Klein-Gordon equation obeyed by the axion quite naturally leads us to take the spatial dependence in the form of a plane wave propagating in space. However, taking into consideration the fact that the effects caused by the KR field should be confined to a very feeble disturbance on the overall homogeneity of space, we assume the the frequency of the axionic wave to be far smaller compared to that of the electromagnetic wave. This modification produces some interesting features, such as, it alters the mutual orthogonality of the electric and magnetic field vectors while inflicting a change on the Poynting equation as well. Furthermore, the inhomogeneity produces an additional rotation of the plane of polarization of radiation over that found in [@skpm]. This additional contribution turns out to be porportional to the square of the frequency of the axion field while being independent of the wavelength of the radiation. The paper is organized as follows. For completeness and as a background, we briefly recapitulate in Section 2 the basic tenets of [@skpm], which lead to a modified set of Maxwell equations. The harmonic spatial dependence assumed for the axion field further modifies these equations as well as the one for the H field itself. Demanding that the waves retain their forms obtained in [@skpm] in the limit $H$ becomes space-independent, we obtain the equations representing circularly-polarized states. Carrying out a standard WKB type procedure, we solve these equations and calculate the rotation angle of the plane of polarization of the fields for a flat background spacetime in Section 3 and for a spatially flat spacetime in Section 4, considering separately the radiation and matter dominated cases therein. We make a few concluding remarks on our results in Section 5. Gauge invariant Einstein-Cartan-Maxwell-Kalb-Ramond coupling ============================================================ Modified field equations ------------------------ The action for gauge-invariant Einstein-Cartan-Maxwell-Kalb-Ramond coupling is taken to be of the form [@pmss]: $$S = \int~ d^{4}x \sqrt{-g} ~\left[~\frac{\tilde{R} (g,T)}{\kappa} - \frac{1}{4} F_{\mu \nu} F^{\mu \nu} - \frac{1}{2} \tilde{H}_{\mu \nu \lambda} \tilde{H}^{\mu \nu \lambda} + \frac{1}{\sqrt{\kappa}} T^{\mu \nu \lambda} \tilde{H}_{\mu \nu \lambda}~\right]$$ $\tilde{R}(g,T)$ being the scalar curvature for the Einstein-Cartan spacetime where the connection contains the torsion tensor $T_{\alpha \mu \nu}$ (supposed to be antisymmetric in all its indices) in addition to the Christoffel term; $\kappa = 16 \pi G$ is the coupling constant; and $\tilde{H}_{\mu \nu \lambda}$ is the KR field strength three-tensor modified by U(1) Chern-Simons term arising form the quantum consistency of an underlying string theory: $$\tilde{H}_{\mu \nu \lambda} = \partial_{[\mu} B_{\nu \lambda]} + \frac{1}{3} \sqrt{\kappa} A_{[\mu} F_{\nu \lambda]},$$ with, $B_{\nu \lambda}$ being the antisymmetric KR potential which is considered to be the possible source of torsion. $\tilde{R}$ is related to the scalar curvature $R$ of purely Riemannian (torsion-free) space-time by $$\tilde{R}(g,T) = R(g) + T_{\mu \nu \lambda} T^{\mu \nu \lambda},$$ The fact that the augmented KR field strength three tensor plays the role of spin angular momentum density (which is the source of torsion [@hehl]) can be evidenced directly from Eq.(1) where the torsion tensor $T_{\mu \nu \lambda}$, being an auxiliary field, obeys the constraint equation $$T_{\mu \nu \lambda} = \sqrt{\kappa} \tilde{H}_{\mu \nu \lambda}.$$ Substituting Eq.(4) in the action (1) and varying the latter with respect to $B_{\mu \nu}$ and $A_{\mu}$ respectively, two sets of field equations are obtained $$D_{\mu} \tilde{H}^{\mu \nu \lambda} \equiv \frac{1}{\sqrt{- g}} \partial _{\mu} (\sqrt{- g} \tilde{H}^{\mu \nu \lambda}) = 0$$ and $$D_{\mu} F^{\mu \nu} = \sqrt{\kappa} \tilde{H}^{\mu \nu \lambda}F_{\lambda \mu}.$$ In addition, there is also the Maxwell - Bianchi identity $$D_{\mu}~ ^{*}F^{\mu \nu} \equiv \frac{1}{\sqrt{- g}} \partial_{\mu} (\sqrt{- g}~ ^{*}F^{\mu \nu}) = 0.$$ Note here that all covariant derivatives are defined with the Christoffel connection and the Maxwell field strength is the standard 2-form $F=dA$. Now, expressing the KR field strength three tensor $H_{\mu \nu \lambda} \equiv \partial_{[\mu} B_{\nu \lambda]}$ as the Hodge-dual to the derivative of the spinless pseudoscalar field $H$ (the axion): $$H_{\mu \nu \lambda} = \epsilon_{\mu \nu \lambda}^{\rho} D_{\rho} H.$$ and substituting in Eqs.(6) and (7), the modified generally covariant Maxwell’s equations are obtained in three-vectorial form $$\begin{aligned} {\bf D\/}\cdot{\bf E\/} &=& 2 \sqrt{\kappa}~{\bf D\/} H\cdot{\bf B\/} \\ D_{0} {\bf E\/} ~-~ {\bf D\/} \times {\bf B\/} &=& -~2 \sqrt{\kappa}~[~D_{0}H~{\bf B\/} ~-~ {\bf D\/}H \times {\bf E\/}~] ~+~ O(\kappa)\\ {\bf D\/}\cdot{\bf B\/} &=& 0 \\ D_{0} {\bf B\/} ~-~ {\bf D\/} \times {\bf E\/} &=& 0 \end{aligned}$$ where $D_{\mu}$ stands for the covariant derivative. On dropping all the higher order terms, as a first approximation, and retaining terms only of the order of $\sqrt{\kappa}$, in a spatially flat isotropic FRW background with metric $$ds^{2} = R^{2}(\eta) (d\eta^{2} - d{\bf x\/}^{2}),$$ the set of equations (9)-(12) take the form $$\begin{aligned} \nabla \cdot {{\bf \tilde{E}\/}}&=& 2 \nabla H \cdot {{\bf \tilde{B}\/}}\\ \partial_{\eta} {{\bf \tilde{E}\/}}~-~ \nabla \times {{\bf \tilde{B}\/}}&=& -2[ ~\partial_{\eta} H {{\bf \tilde{B}\/}}- \nabla H \times {{\bf \tilde{E}\/}}~] \\ \nabla \cdot {{\bf \tilde{B}\/}}&=& 0 \\ \partial_{\eta} {{\bf \tilde{B}\/}}~+~ \nabla \times {{\bf \tilde{E}\/}}&=& 0 \end{aligned}$$ where $\eta$ the conformal time coordinate, defined by $d\eta = dt/R(t)$, $R$ is the cosmological scale factor; and ${{\bf \tilde{E}\/}}= R^{2}{\bf E\/}$ and ${{\bf \tilde{B}\/}}= R^{2}{\bf B\/}$. $H$ is redefined by absorbing the $\sqrt{\kappa}$ in it. It is easy to show from the very form of the KR field strength, viz., $H_{\mu \nu \lambda} = \partial_{[\mu}B_{\nu \lambda]}$, that it satisfies the Bianchi identity $$\epsilon^{\mu \nu \lambda \rho} \partial_{\rho} H_{\mu \nu \lambda} = 0$$ which immediately implies that $H$ satisfies the wave equation $$D_{\rho} D^{\rho} H = 0$$ In isotropic spatially flat universe this equation reduces to $$(\partial_{\eta}^{2} ~-~ \nabla^{2}) H ~= ~-~2 \frac{\dot{R}}{R} \dot{H}$$ where $H$ is taken to be a general function of both space and time coordinates, and the over-dot implies partial differentiation with respect to $\eta$. Assuming a general wave solution for $H$, viz., $$H(\eta,{\bf x\/}) ~=~ H_{0}(\eta)~ \cos ~{{{\bf p \/}}\cdot {\bf x\/}}$$ Eq.(20) enables us to get an equation for $H$ $$\ddot{H} ~+~ 2 \frac{\dot{R}}{R} \dot{H} ~+~ p^{2} H ~=~ 0$$ We should point out here that as a consequence of our prior assumption that the overall homogeneity of the universe over long distance scales is not much disturbed by the inclusion of the spatial part in $H$, we are taking $p$ to be much less compared to the wave number for the electromagnetic radiation. Modifications in electro-magnetic orthogonality and the Poynting equation -------------------------------------------------------------------------- From the field equations (14) - (17) we derive the wave equations for the electric and magnetic fields $$\begin{aligned} \Box {{\bf \tilde{B}\/}}~\equiv~ (~\partial_{\eta}^{2} ~-~ \nabla^{2}~)~ {{\bf \tilde{B}\/}}&=& 2~ \nabla \times (~\dot{H} {{\bf \tilde{B}\/}}) ~-~ 2~ \nabla \times (~\nabla H \times {{\bf \tilde{E}\/}}~) \\ \Box {{\bf \tilde{E}\/}}~\equiv~ (~\partial_{\eta}^{2} ~-~ \nabla^{2}~)~ {{\bf \tilde{E}\/}}&=& 2~ \nabla \times (~\dot{H} {{\bf \tilde{E}\/}}) ~+~ 2~ \nabla H \times \dot{{{\bf \tilde{E}\/}}} ~-~ \nabla (~\nabla \cdot {{\bf \tilde{E}\/}}~) ~-~ 2~ \ddot{H} {{\bf \tilde{B}\/}}\end{aligned}$$ These equations indeed reduce to the pure Maxwell equations in the limit $H ~\rightarrow ~0$, or a constant. Treating the axion $H$ as a [*tiny*]{} purturbation over the Maxwell equations we argue that the solutions should have a form not much departing from the usual plane wave structure with the wave vector ${{\bf k \/}}$ perpendicular to both the electric and magnetic vectors. Moreover, the field equations (14) and (15) enable us to derive $$\nabla \dot{H} \cdot {{\bf \tilde{B}\/}}= 0$$ which, in view of the specific form of $H$, viz., $H_{0}(\eta)~ \cos ({{{\bf p \/}}\cdot {\bf x\/}})$, implies $${{\bf p \/}}\cdot {{\bf \tilde{B}\/}}= 0$$ provided $\dot{H} \ne 0$, which is the general case we are handling. This orthogonality of ${{\bf p \/}}$ and ${{\bf \tilde{B}\/}}$ \[Eq.(26)\] makes it easier to assume, for simplicity, that ${{\bf p \/}}$ can be taken to be orthogonal to ${{\bf \tilde{E}\/}}$ as well, i.e., ${{\bf p \/}}$ is either parallel or antiparallel to ${{\bf k \/}}$. This is fairly justified, as it looks, from the similar wave nature of the electric and magnetic fields, at least in the limiting pure Maxwellian case as we are treating the modification caused by the KR field as a small purturbation over the Maxwellian behaviour. In fact, we may choose to observe the electromagnetic radiation which is travelling in the direction of propagation of the KR field. Considering the z-direction to be the propagation direction of the electromagnetic waves and as such for the axion (following the abovementioned assumption) we reduce the four-dimesional problem to a two-dimensional one with $\eta$ and $z$ being the only variables. The field equations now reduce to simpler forms $$\begin{aligned} \Box {{\bf \tilde{B}\/}}&=& 2~ \dot{H} \nabla \times {{\bf \tilde{B}\/}}~+~ 2 \partial_{\eta} (H' {\bf \hat{e}_{z}} \times {{\bf \tilde{B}\/}}) ~+~ 2~ H'' {{\bf \tilde{E}\/}}\\ \Box {{\bf \tilde{E}\/}}&=& 2~ \dot{H} \nabla \times {{\bf \tilde{E}\/}}~+~ 2 \partial_{\eta} (H' {\bf \hat{e}_{z}} \times {{\bf \tilde{E}\/}}) ~-~ 2~ \ddot{H} {{\bf \tilde{B}\/}}\end{aligned}$$ where the over-dot and prime denote respectively the partial differentiations with respect to $\eta$ and $z$; and ${\bf \hat{e}_{z}}$ is the unit vector along the z-direction. Now, it is easy to show from the field equations (15) and (17) that $$\partial_{\eta}~({{\bf \tilde{E}\/}}\cdot {{\bf \tilde{B}\/}})~ = ~{{\bf \tilde{B}\/}}\cdot \nabla \times {{\bf \tilde{B}\/}}~- {{\bf \tilde{E}\/}}\cdot \nabla \times {{\bf \tilde{E}\/}}~-~ 2~\dot{H}~{{\bf \tilde{B}\/}}^2 ~+~ 2~\nabla H \cdot {{\bf \tilde{E}\/}}\times {{\bf \tilde{B}\/}}$$ Considering the limiting plane wave behaviour of the solutions of the electromagnetic wave equations as $~H ~\rightarrow ~0~$, the magnetic and the electric fields can be expressed $$\begin{aligned} {{\bf \tilde{B}\/}}(\eta,z) &=& {{\bf \tilde{B}_{0}\/}}(\eta,z) ~ e^{- i k z} \\ {{\bf \tilde{E}\/}}(\eta,z) &=& {{\bf \tilde{E}_{0}\/}}(\eta,z) ~ e^{- i k z}\end{aligned}$$ When the KR field actually vanishes, ${{\bf \tilde{E}_{0}\/}}, {{\bf \tilde{B}_{0}\/}}~\equiv~$ constant vectors $\times~ e^{i k \eta}$  in the plane wave solutions of the pure Maxwell equations. Eq.(29) then gives $~{{\bf \tilde{E}\/}}\cdot \nabla \times {{\bf \tilde{E}\/}}~=~ {{\bf \tilde{B}\/}}\cdot \nabla \times {{\bf \tilde{B}\/}}~=~ 0$, i.e., $\partial_{\eta} ({{\bf \tilde{E}\/}}\cdot {{\bf \tilde{B}\/}}) ~=~ 0$. Moreover, since ${{\bf \tilde{B}\/}}$ is in the direction of $\nabla \times {{\bf \tilde{E}\/}}$, as is evident from Eq.(17), it follows that ${{\bf \tilde{E}\/}}\cdot {{\bf \tilde{B}\/}}~=~ 0$. When the KR field is present, but is only time-dependent (the case in [@skpm]), the vectors ${{\bf \tilde{E}_{0}\/}}$ and ${{\bf \tilde{B}_{0}\/}}$ in the solutions (30) and (31) again depend on time only and Eq.(29) reduces to $$\partial_{\eta}~({{\bf \tilde{E}_{0}\/}}\cdot {{\bf \tilde{B}_{0}\/}})~ = ~-~ 2~\dot{H}~{{\bf \tilde{B}_{0}\/}}^2.$$ Clearly, $~{{\bf \tilde{E}_{0}\/}}\cdot {{\bf \tilde{B}_{0}\/}}~\ne~ 0$,  which implies that the mutual orthogonality of ${{\bf \tilde{E}\/}}$ and ${{\bf \tilde{B}\/}}$ is lost. In the most general case where the KR field depends on both space and time coordinates, ${{\bf \tilde{E}_{0}\/}}$ and ${{\bf \tilde{B}_{0}\/}}$ are also spacetime-dependent and they satisfy the relation $$\partial_{\eta}~({{\bf \tilde{E}_{0}\/}}\cdot {{\bf \tilde{B}_{0}\/}})~ = ~{{\bf \tilde{B}_{0}\/}}\cdot \nabla \times {{\bf \tilde{B}_{0}\/}}~-~ {{\bf \tilde{E}_{0}\/}}\cdot \nabla \times {{\bf \tilde{E}_{0}\/}}~-~ 2~\dot{H}~{{\bf \tilde{B}_{0}\/}}^2 ~+~ 2~\nabla H \cdot {{\bf \tilde{E}_{0}\/}}\times {{\bf \tilde{B}_{0}\/}}$$ Here we find, one cannot ascertain conclusively that ${{\bf \tilde{E}_{0}\/}}\cdot {{\bf \tilde{B}_{0}\/}}$ is manifestly zero. It is, in fact, more justified to think that ${{\bf \tilde{E}_{0}\/}}\cdot {{\bf \tilde{B}_{0}\/}}$ is essentially non-zero in general, at least, by looking at the limiting behaviour when the KR field is stripped off the spatially depending part. The last term in Eq.(33), appearing due to the inclusion of the spatial dependence in $H$ cannot, in general, compensate for the term which actually renders ${{\bf \tilde{E}_{0}\/}}\cdot {{\bf \tilde{B}_{0}\/}}$ non-vanishing in Eq.(32), as the temporal and spatial components of the KR field are completely separate entities. The Poynting equation in the present scenario can be obtained directly from the field equations (14) - (17). It is given by $$\nabla \cdot {\bf S} ~+~ \dot {\omega}_{em} ~=~ - 2 \dot{H}~ {{\bf \tilde{E}\/}}\cdot {{\bf \tilde{B}\/}}$$ where  $ {\bf S} ~=~ ({{\bf \tilde{E}\/}}\times {{\bf \tilde{B}\/}}) $  is the Poynting vector and  $ \omega_{em} ~=~ \frac 1 2 ({{\bf \tilde{E}\/}}^2 ~+~ {{\bf \tilde{B}\/}}^2) $  is the electromagnetic energy density. The distinction of this equation over the Poynting equation in the pure Maxwellian case is the presence of the term on the right hand side which is, in general, non-zero for reasons discussed above. In fact, this term vanishes in the limit  $\dot{H} ~ \rightarrow~ 0$, i.e., when $H$ becomes purely space-dependent. This is quite obvious since spatial inhomogeneity alone cannot bring in any change in the power conservation equation. Polarization States and Duality transformation ---------------------------------------------- Following [@cf] and rearranging terms of the components of the wave equations (27) and (28) (following the procedure in [@cf]), we obtain the following equations for the polarized states $$\begin{aligned} \ddot{b}_{\pm} ~\mp~ 2 i H' \dot{b}_{\pm} ~\mp~ 2 i \dot{H}' b_{\pm} ~&=&~ ~b_{\pm}'' ~\pm~ 2 i \dot{H} b_{\pm}' ~+~ 2 H'' e_{\pm} \\ \ddot{e}_{\pm} ~\mp~ 2 i H' \dot{e}_{\pm} ~\mp~ 2 i \dot{H}' e_{\pm} ~&=&~ ~e_{\pm}'' ~\pm~ 2 i \dot{H} e_{\pm}' ~-~ 2 \ddot{H} b_{\pm}, \end{aligned}$$ where $$\begin{aligned} b_{\pm}(\eta,z) &=& \tilde{B}_{x}(\eta,z) ~\pm~ i \tilde{B}_{y}(\eta,z) \nonumber\\ e_{\pm}(\eta,z) &=& \tilde{E}_{x}(\eta,z) ~\pm~ i \tilde{E}_{y}(\eta,z).\end{aligned}$$ Note that the Eqs.(35) and (36) are converted into each other by the transformation  $e_{\pm} \rightarrow b_{\pm}$, $b_{\pm} \rightarrow - e_{\pm}$  provided the equation  $\Box H ~\equiv~ {\ddot H} ~-~ H'' ~=~0$  is obeyed. This is the usual electro-magnetic [*duality*]{} symmetry of the Maxwell equations. The Maxwell-KR system indeed possess this invariance in a flat spacetime background with cosmplogical scale factor  $R ~=~ 1$,  as is evident from Eq.(20). In a curved spacetime background, however, one does not have this invariance any more. Rewriting the Eqs.(35) and (36) as follows: $$\begin{aligned} \ddot{b}_{\pm} ~\mp~ 2 i H' \dot{b}_{\pm} ~\mp~ 2 i \dot{H}' b_{\pm} ~&=&~ ~b_{\pm}'' ~\pm~ 2 i \dot{H} b_{\pm}' ~+~ 2 H'' e_{\pm} ~=~ \alpha_{\pm}(\eta,z) \\ \ddot{e}_{\pm} ~\mp~ 2 i H' \dot{e}_{\pm} ~\mp~ 2 i \dot{H}' e_{\pm} ~&=&~ ~e_{\pm}'' ~\pm~ 2 i \dot{H} e_{\pm}' ~-~ 2 \ddot{H} b_{\pm} ~=~ \beta_{\pm}(\eta,z), \end{aligned}$$ We seek the appropriate forms of the functions $\alpha(\eta,z)$ and $\beta(\eta,z)$ by examining the limiting forms of the above equations: [**I.  Limit $H' \rightarrow 0$ :** ]{} In this limit, where $H$ becomes purely a function of $\eta$, the equations (37) and (38) reduce to $$\begin{aligned} \ddot{b}_{\pm} &=& ~b_{\pm}'' ~\pm~ 2 i \dot{H} b_{\pm}' ~=~ \alpha_{\pm}(H' \rightarrow 0) \\ \ddot{e}_{\pm} &=& ~e_{\pm}'' ~\pm~ 2 i \dot{H} e_{\pm}' ~-~ 2 \ddot{H} b_{\pm} ~=~ \beta_{\pm}(H' \rightarrow 0) \end{aligned}$$ Assuming solutions of the form $$\begin{aligned} b_{\pm}(\eta,z) &=& ~b_{0}^{\pm}(\eta)~ e^{- i k z} \\ e_{\pm}(\eta,z) &=& ~e_{0}^{\pm}(\eta)~ e^{- i k z}\end{aligned}$$ we find $$\begin{aligned} \ddot{b}_{\pm} &=& ~ - (k^{2} \mp 2 k \dot{H}) b_{\pm} ~=~ \alpha_{\pm}(H' \rightarrow 0) \\ \ddot{e}_{\pm} &=& ~ - (k^{2} \mp 2 k \dot{H}) e_{\pm} ~-~ 2 \ddot{H} b_{\pm} ~=~ \beta_{\pm}(H' \rightarrow 0). \end{aligned}$$ [**II.  Limit $\dot{H} \rightarrow 0$ :** ]{} This limiting case implies $H$ to be a function of $z$ only but as is evident from Eq.(22) $H'$ is merely a constant. Therefore Eqs.(37) and (38) reduce to $$\begin{aligned} \ddot{b}_{\pm} ~\mp~ 2 i H' \dot{b}_{\pm} &=& ~b_{\pm}'' ~=~ \alpha_{\pm}(\dot{H} \rightarrow 0) \\ \ddot{e}_{\pm} ~\mp~ 2 i H' \dot{e}_{\pm} &=& ~e_{\pm}'' ~=~ \beta_{\pm}(\dot{H} \rightarrow 0)\end{aligned}$$ Again assuming solutions of the form $$\begin{aligned} b_{\pm}(\eta,z) &=& ~b_{1}^{\pm}(z)~ e^{ i k \eta} \\ e_{\pm}(\eta,z) &=& ~e_{1}^{\pm}(z)~ e^{ i k \eta}\end{aligned}$$ we get $$\begin{aligned} b_{\pm}'' &=& ~ - (k^{2} \mp 2 k H') b_{\pm} ~=~ \alpha_{\pm}(\dot{H} \rightarrow 0) \\ e_{\pm}'' &=& ~ - (k^{2} \mp 2 k H') e_{\pm} ~=~ \beta_{\pm}(\dot{H} \rightarrow 0). \end{aligned}$$ By looking at these limiting forms of $\alpha_{\pm}$ and $\beta_{\pm}$ given in Eqs.(44),(45) and in Eqs.(50),(51) it seems reasonable to suggest the following simplest possible structures of these functions: $$\begin{aligned} \alpha_{\pm}(\eta,z) &=& ~- \left[ k^{2} \mp 2 k ( \dot{H} + H' ) \right] b_{\pm}(\eta,z) \\ \beta_{\pm}(\eta,z) &=& ~- \left[ k^{2} \mp 2 k ( \dot{H} + H' ) \right] e_{\pm}(\eta,z) ~-~ 2 \ddot{H} b_{\pm}(\eta,z). \\\end{aligned}$$ Moreover, setting $$e_{\pm} (\eta,z) = a_{\pm} (\eta,z)~ b_{\pm} (\eta,z)$$ we write the equations (37) and (38) in a more elegant form $$\begin{aligned} \ddot{b}_{\pm} ~\mp~ 2 i H' \dot{b}_{\pm} ~+~ \left[ k^{2} ~\mp~ 2 k ( \dot{H} + H' ) ~\mp~ 2 i \dot{H}' \right] b_{\pm}(\eta,z) &=& 0 \\ b_{\pm}'' ~\pm~ 2 i \dot{H} b_{\pm}' ~+~ \left[ k^{2} ~\mp~ 2 k ( \dot{H} + H' ) ~+~ 2 H'' a_{\pm} \right] b_{\pm}(\eta,z) &=& 0;\end{aligned}$$ and $$\begin{aligned} \ddot{e}_{\pm} ~\mp~ 2 i H' \dot{e}_{\pm} ~+~ \left[ k^{2} ~\mp~ 2 k ( \dot{H} + H' ) ~\mp~ 2 i \dot{H}' ~+~ 2 \frac{\ddot{H}}{a_{\pm}} \right] e_{\pm}(\eta,z) &=& 0 \\ e_{\pm}'' ~\pm~ 2 i \dot{H} e_{\pm}' ~+~ \left[ k^{2} ~\mp~ 2 k ( \dot{H} + H' ) \right] e_{\pm}(\eta,z) &=& 0.\end{aligned}$$ Flat Spacetime Background ========================= In order to get a preliminary idea as to how the coupling of a spacetime dependent KR field to Einstein-Maxwell theory affects the electromagnetic waves, and thereby effects in an optical activity in the radiation coming from distant galactic sources, we consider the simplest situation — that is, of a flat universe with cosmological scale factor $R(\eta) = 1$. Admittedly, quantitative details of results of this section are cosmologically untenable for obvious reasons. The equation of motion (22) for $H$ can be solved readily to obtain $$H(\eta,z) ~=~ (c_{1} \sin p \eta ~+~ c_{2} \cos p \eta)~ \cos p z$$ where $c_{1}$ and $c_{2}$ are arbitrary integration constants. Demanding that the above solution must reduce to the form $( h \eta ~+~ h_{0} )$ which is the solution of Eq.(22) in the limit $p \rightarrow 0$ we infer $c_{1} = h/p$ and $c_{2} = h_{0}$. Here $h$ and $h_{0}$ are the same arbitrary constants denoted in [@skpm]. Making a Taylor series expansion of the various functions appearing in Eq.(60) around $p = 0$ we write $$H(\eta,z) ~=~ (h \eta ~+~ h_{0}) ~-~ \frac{p^{2}}{2} \left(\frac{h \eta^{3}}{3} ~+~ h_{0} \eta^{2} + h \eta z^{2} + h_{0} z^{2}\right) ~+~ O(p^{4})$$ Substituting this $H$ in Eqs.(55) and (57) and assuming solution of the standard WKB type given by $$b_{\pm} (\eta,z) ~=~ \bar{b} ~ e^{i k~ S_{\pm} (\eta,z)}$$ with $$S_{\pm} ~=~ S_{0}^{\pm} ~+~ \frac{S_{1}^{\pm}}{k} ~+~ \frac{S_{2}^{\pm}}{k^2} ~+~ \cdots$$ we get after partial integrations of Eqs.(56) and (57) with respect to $\eta$ and $z$ respectively $$b_{\pm} (\eta,z) ~=~ \bar{b}~ \exp \left\{ i k \eta ~\mp~ i \left[ h \eta ~-~ \frac{ p^2}{2} \left( \frac{ h \eta^3}{3} ~+~ h_{0} \eta^2 ~+~ h z^2 \eta \right) ~+~ O(p^4) \right] ~+~ O\left(\frac 1 k \right) ~+~ i k f_{\pm} (z) \right\}$$ and $$b_{\pm} (\eta,z) ~=~ \bar{b}~ \exp \left \{ i k g_{\pm} (\eta) ~-~ i k z ~\mp~ i \left[ \frac{ p^2}{2} ( h \eta ~+~ h_{0} ) z^2 ) ~+~ O(p^4) \right] ~+~ O\left(\frac 1 k \right) \right \}.$$ Comparing these two expressions we assert the forms of the arbitrary functions $f_{\pm} (z)$ and $g_{\pm} (\eta)$ and write $b_{\pm} (\eta,z)$ as follows: $$b_{\pm} (\eta,z) ~=~ \bar{b}~ \exp \left \{ i k ( \eta - z ) ~\mp~ i \left[ h \eta ~-~ \frac{ p^2}{2} \left( \frac{ h \eta^3}{3} ~+~ h_{0} \eta^2 ~-~ h_{0} z^2 \right) ~+~ O(p^4) \right] ~+~ O\left(\frac 1 k \right) \right \}$$ A similar approach for $e_{\pm} (\eta,z)$ yields $$e_{\pm} (\eta,z) ~=~ \bar{e}~ \exp \left \{ i k ( \eta - z ) ~\mp~ i \left[ h \eta ~-~ \frac{ p^2}{2} \left( \frac{ h \eta^3}{3} ~+~ h_{0} \eta^2 ~-~ h_{0} z^2 \right) ~+~ O(p^4) \right] ~+~ O\left(\frac 1 k \right) \right \}$$ which differs from Eq.(66) only in the constant coefficient $\bar{e}$ and in the higher order $O\left(\frac 1 k \right)$. However, it should be mentioned here that while using the WKB technique we are assuming that the function $a_{\pm}$ is not increasing rapidly as $k$ increases. In fact, in absence of the KR field, when we have the plane wave solutions of Maxwell’s equations, $a_{+} ~=~ a_{-} ~=~ \bar{e} / \bar{b} ~=~$ constant. The spacetime dependence of $a_{\pm}$ comes only in presence of a spacetime-dependent $H$. Since $H$ is being treated as a small purturbation over the Maxwell field, it is rather plausible to think $a_{\pm}$ to be not much different from the constant $\bar{e} / \bar{b}$ and a very slowly-varying function of spacetime. But the constant $\bar{e} / \bar{b}$ is arbitrary and cannot generically be argued as increasing with $k$. Therefore, the assumption that $a_{\pm}$ remains quite invariant as $k$ increases is fairly justified. Using WKB method, the solutions obtained above involves $a_{\pm}$ only in the higher order $O\left(\frac 1 k \right)$. Now, the circular polarization states are defined by $b_{\pm}$ and $e_{\pm}$ and the extent of the optical birefringence due the presence of the KR field can be estimated directly by calculating the rotation angle of the plane of polarization of the electromagnetic wave, which is given by the phase difference   $\phi_{mag}~ \equiv~ \frac 1 2 [arg ~b_{+} - arg ~b_{-}] $ for the magnetic field and $\phi_{elec}~ \equiv~ \frac 1 2 [arg ~e_{+} - arg ~e_{-}] $ for the electric field. The phase shift is given by $$\phi_{mag} (\eta,z) ~\approx~ \phi_{elec} (\eta,z) ~\approx~ - h \eta ~+~ \frac{p^2}{2} \left( \frac{h \eta^3}{3} ~+~ h_{0} \eta^2 ~-~ h_{0} z^2 \right)$$ for $~ h,p << k$. It is interesting to see that the change in the rotation angle, calculated here, over that found in [@skpm] for flat universe, is primarily given by the $p^2$-dependent part. But as $p$ is considered to be very small we infer that this change is rather insignificant. Spatially Flat FRW Spacetime Background ======================================= We now turn to less trivial background spacetimes. We consider a spatially flat expanding universe dominated by radiation and matter, in turn. Radiation dominated Universe ---------------------------- The scale factor for this model, in real time, is given by $$~[R(\eta)]^{RD} ~=~ \frac{\eta}{\eta_{r}}$$ where $\eta_{r} = \left(8 \pi G \epsilon_{0} /3\right)^{-1/3}$, $\epsilon_{0}$ being the primordial radiant energy density. Substituting this in the equation of motion (22) for $H$ we obtain $$\eta^{2} \ddot{H} + 2\eta \dot{H} + p^{2} \eta^{2} H ~=~ 0$$ which has the form of a transformed Bessel equation with solution $$H (\eta,z) ~=~ \eta^{- \frac 1 2}~\left[ \bar{c}_{1} ~ J_{1/2} (p \eta) ~+~ \bar{c}_{2} ~Y_{1/2} (p \eta) \right].$$ Simplifying the Bessel functions of first and second kinds, viz., $J$ and $Y$, the above solution can be written as $$H (\eta,z) ~=~ \frac 1 \eta ~( c_{1} ~\sin ~p \eta ~+~ c_{2} ~ \cos ~p \eta ) ~ \cos~p z.$$ Imposing again the boundary condition that this must reduce to the limiting form $ ~\left[ -~ \frac {h \eta_{r}^2}{\eta} ~+~ h_{0} \right]~$ — the solution of Eq.(22) —  as $~p \rightarrow 0~$, we set $~c_{1} ~=~ \frac {h_{0}} p~$ and $~c_{2} ~=~ - h_{r}$,  with $~h_{r} ~=~ h~\eta_{r}^2$,  whence $$H (\eta,z) ~=~ \frac 1 \eta ~\left( \frac {h_{0}} p ~\sin ~p \eta ~-~ h_{r} ~ \cos ~p \eta \right)~ \cos~p z.$$ Plugging in the Taylor expanded form of this $H$ in Eqs.(56) - (59) and using the same WKB technique as for the flat universe, we obtain $$b_{\pm} (\eta,z) ~=~ \bar{b}~ \exp \left[ i k ( \eta - z ) ~\pm~ i \left[ \frac {h_{r}} \eta ~+~ \frac{p^{2}}{2} \left( \frac{h_{0} \eta^2}{3} ~-~ h_{r} \eta ~-~ h_{0} z^2 \right) ~+~ O(p^4) \right] ~+~ O\left(\frac 1 k \right) \right]$$ and similar expression for $e_{\pm} (\eta,z)$. The phase shift in this case is given by $$\phi_{mag} (\eta,z) ~\approx~ \phi_{elec} (\eta,z) ~\approx~ \frac {h_{r}} \eta ~-~ \frac{p^2}{2} \left( h_{r} \eta ~-~ \frac{h_{0} \eta^2}{3} ~+~ h_{0} z^2 \right)$$ for $ ~h,p ~<<~ k$ and $~h_{r} ~=~ h~ \eta_{r}^2$. Matter dominated Universe ------------------------- In this case, where the scale factor is governed by $$~[R(\eta)]^{MD} ~=~ \frac{\eta^2}{\eta_{m}^2}$$ with $\eta_{m} = \left(8 \pi G \rho_{0} /3\right)^{-1/3}$  ($\rho_{0}$   – the initial matter density),  Eq.(22) reduces again to a transformed Bessel equation $$\eta^{2} \ddot{H} + 4\eta \dot{H} + p^{2} \eta^{2} H ~=~ 0$$ having simplified solution $$H (\eta,z) ~=~ \frac 1 {\eta^3} ~[ (c_{1} ~-~ c_{2}~p \eta) ~\sin ~p \eta ~-~ (c_{2} ~+~ c_{1}~p \eta) ~ \cos ~p \eta ]~ \cos~p z.$$ Determining the constants $c_{1}$ and $c_{2}$ using, as before, the boundary condition on $H$ in the limit $~p \rightarrow 0~$, we write $$H (\eta,z) ~=~ - ~\frac {h_{m}} {3 \eta^3} ~( \cos ~p \eta ~+~ p \eta ~\sin ~ p \eta)~ \cos~p z.$$ where $~h_{m} ~=~ h \eta^4$. With this $H$, we obtain using the WKB method $$b_{\pm} (\eta,z) ~=~ \bar{b}~ \exp \left[ i k ( \eta - z ) ~\pm~ i \left[ \frac {h_{m}} {3 \eta^3} ~+~ p^{2} h_{m} \left( \frac 1 \eta ~+~ \frac {z^2} 6 \right) ~+~ O(p^4) \right] ~+~ O\left(\frac 1 k \right) \right]$$ and similar expression for $e_{\pm} (\eta,z)$. The phase shift can be calculated $$\phi_{mag} (\eta,z) ~\approx~ \phi_{elec} (\eta,z) ~\approx~ \frac {h_{m}}{3 \eta^3} ~+~ p^2 h_{m} \left( \frac 1 \eta ~+~ \frac {z^2} 6 \right)$$ for $ ~h,p ~<<~ k$. Conclusions =========== One rather surprising aspect of our finding is the loss of orthogonality of the electric and magnetic vectors in the radiation field which exists even in the case of a spatially homogeneous axion field, so long as the axion is time-dependent. This will indeed affect the measurement of the precise rotation of the plane of polarization, although for large redshift sources to which we confine, this effect may be ignored for all practical purposes. In general the electric and magnetic vectors seem to have solutions generically represented as $$(Electric/Magnetic ~field ~combinations) ~=~ (constant) ~e^{i(kz ~-~ \omega \eta)}~e^{i H(\eta, z)}$$ The additional contribution to the angle of rotation arising out of the spatial inhomogeneity introduced in this paper is actually quite small under our assumption that the wavelength of the axion wave is far larger than the radiation. Thus from an observational standpoint this new effect is not too significant, although it is still quite distinct in its behaviour from Faraday rotation. We have used standard WKB type methods to arrive at some solution to the complicated set of equations which arise even at the lowest order. Among other things we have explored the properties of the electric and magnetic fields of the Maxwell–KR system. We also checked the validity of the standard electric–magnetic duality and pointed out the existence of a very simple solution to the complicated equations. Acknowledgements ================ This work is supported by Project grant no. 98/37/16/BRNS cell/676 from The Board of Research in Nuclear Sciences, Department of Atomic Energy, Government of India, and the Council of Scientific and Industrial Research, Government of India. .5in P. Majumdar and S. SenGupta, Class. Quan. Grav. [**16**]{} L89 (1999). M. Green, J. Schwarz and E. Witten, Superstring Theory v.2, 1985 (Cambridge: Cambridge Univ. Press). S. Kar, P. Majumdar, S. SenGupta and A. Sinha, 2000, [*Preprint*]{} gr-qc/0006097. P. Jain and J. Ralston, Mod. Phys. Lett. [**A14**]{}, 417 (1999). B. Nodland and J. P. Ralston, Phys. Rev. Lett. [**78**]{}, 3043 (1997) ; [*ibid*]{} Phys. Rev. Lett. [**79**]{}, 1958 (1997); astro-ph/9708114;astro-ph/9706126. P. Das, P. Jain and S. Mukherji, 2000, [*Preprint*]{} hep-ph/ 0011279. F. Hehl, P. von der Heyde, G. Kerlick and J. Nester, Rev. Mod. Phys.[**48**]{} 393 (1976). S.M. Carroll and G.B.Field, Phys. Rev. [**D 43**]{}, 3789 (1991). [^1]: Electronic address: [*[email protected]*]{} [^2]: Electronic address: [*[email protected]*]{} [^3]: Electronic address: [*[email protected]*]{} [^4]: Electronic address: [*[email protected]*]{}
{ "pile_set_name": "ArXiv" }
--- abstract: 'We performed charge detection on a lateral triple quantum dot with star-like geometry. The setup allows us to interpret the results in terms of two double dots with one common dot. One double dot features weak tunnel coupling and can be understood with atom-like electronic states, the other one is strongly coupled forming molecule-like states. In nonlinear measurements we identified patterns that can be analyzed in terms of the symmetry of tunneling rates. Those patterns strongly depend on the strength of interdot tunnel coupling and are completely different for atomic- or molecule-like coupled quantum dots allowing the non-invasive detection of molecular bonds.' author: - 'M. C. Rogge' - 'R. J. Haug' title: 'Non-invasive detection of molecular bonds in quantum dots' --- Quantum dots are often called artificial atoms [@Kouwenhoven-97] due to their discrete electronic level spectrum. When several quantum dots are connected, they start to interact [@Wiel-03]. If the tunneling rate between the dots is small, the electronic wavefunctions are still constricted to the single quantum dots and the interaction is dominated by electrostatics with sequential interdot tunneling. In contrast, for large tunneling rates electronic states can be found extended over several dots. These extended states introduce covalent bonding as in real molecules [@Blick-96]. Whether or not the interdot tunneling rates are sufficient to form coherent molecule-like states is a crucial information in order to properly describe a quantum dot system. Especially for quantum computing purposes [@Loss-98] coherent states are necessary to form and couple qubits and to implement SWAP gates for qubit manipulation [@Petta-05; @Hanson-07]. With dc-transport experiments there are only a few methods that can give hints for molecular bonds. The width of anticrossings visible in charging diagrams is a measure [@Golden-96], although anticrossings appear for capacitively coupled dots as well. In addition the curvature of the lines forming an anticrossing can be used [@Blick-98] and also the visibility of lines in non-parallel quantum dots [@Rogge-04]. Another alternative is to study excited states. Strongly coupled quantum dots form bonding and antibonding states that are visible in nonlinear measurements [@Huttel-05]. We have studied the impact of the coupling strength on the mean charge of multiple quantum dot systems coupled in series. Using a quantum point contact [@Field-93] we analyzed the mean charge in stability diagrams. We found that depending on the symmetry of tunneling rates and on the coupling strength characteristic patterns are formed in nonlinear measurements. This allows to non-invasively detect the symmetry of tunneling rates and the quality of the interdot coupling. ![(color online) charging diagram of the triple dot device measured with charge detection. Three sets of lines appear from the three dots A, B and C. At the intersections anticrossings are visible due to interdot coupling (circles). Inset: atomic force microscopic image of the device with dots A, B, C and a quantum point contact for charge detection.[]{data-label="fig1"}](fig1.eps) The measurements were performed on a device containing three quantum dots A, B and C (see inset of Fig. \[fig1\]). The device was produced using local anodic oxidation on a GaAs/AlGaAs-heterostructure [@Ishii-95; @Keyser-00]. The three dots are positioned in a star-like geometry with one lead for each dot (Source S at A, Drain1 D1 at B and Drain2 D2 at C). The barriers and the potentials can be tuned with four sidegates G1 to G4. A quantum point contact (QPC) is placed next to the three dots for charge detection. The device is described in detail in Ref. [@Rogge-08]. Figure \[fig1\] shows a charging diagram of the triple dot device as a function of the voltages applied to gates G3 and G1. The derivative of the QPC-current with respect to $V_{\text{G3}}$ is plotted. Dark lines correspond to an increase of charge detected by the QPC (e.g. charging a dot with an electron), bright features appear when the QPC detects a decrease of charge. Three sets of lines are visible that denote charging of the three dots. Lines with a shallow slope correspond to dot B, those with steap slopes appear due to charging of dot C. Intermediate slopes correspond to dot A. Whenever two lines from different sets intersect, an anticrossing appears due to finite coupling between the dots. At these anticrossings, two of the dots are in resonance and can thus be treated as double quantum dot. From transport measurement we know that the double dots A-B and A-C feature finite interdot tunnel coupling, while the double dot B-C is coupled capacitively only (see Ref. [@Rogge-08]). Therefore B-C is not interesting for the purpose of this paper. In the following we concentrate on the analysis of the two double dots A-B and A-C. This analysis is done in the two sections I and II with I showing anticrossings due to resonance of dots A and B, II showing an anticrossing due to resonance of dots A and C (circles). ![(color online) charge measurement in section I of Fig. \[fig1\] for $V_{\text{S}}=-1$ mV, $V_{\text{S}}=0$ mV and $V_{\text{S}}=1$ mV. At $V_{\text{S}}\neq 0$ triangular patterns appear with additional lines of dark and bright features due to excited atomic states.[]{data-label="fig2"}](fig2.eps) Figure \[fig2\] shows three graphs measured in the region of section I. Charge detection is performed in the linear regime with $V_{\text{S}}=0$ (center image) and in the nonlinear regime with $V_{\text{S}}=-1$ mV (left) and $V_{\text{S}}=1$ mV (right). While the center image shows the same pattern as observed in Fig. \[fig1\] with two sets of lines for dots A and B and two anticrossings, the situation changes in the nonlinear regime showing a more complex pattern with ground and excited states. The anticrossings appear shifted to the upper left for $V_{\text{S}}=-1$ mV and to the lower right for $V_{\text{S}}=1$ mV. A triangular shaped pattern with additional lines is connected to the right ($V_{\text{S}}=-1$ mV) and to the left respectively ($V_{\text{S}}=1$ mV). These triangles are familiar from nonlinear transport measurements in weakly coupled quantum dots [@Dixon-96; @Wiel-03] and have recently been measured with charge detection as well [@Johnson-05]. However, triangles with such patterns have not been reported so far. Dark and bright features alternate corresponding to alternating increase and decrease of mean charge measured with the QPC. ![(color online) top left: schematic for the anticrossing of dots A and B for $V_{\text{S}}=0$ with triple points e and h. Top right: schematic for the anticrossing at $V_{\text{S}}>0$. Two triangles $\alpha$ and $\beta$ appear with additional lines. Bottom left: possible configuration of chemical potentials. Bottom right: Section of Fig. \[fig2\], $V_{\text{S}}=$1 mV. The two triangles $\alpha$ and $\beta$ are visible even though they overlap. The line pattern has the opposite order in both triangles. A dark line in $\alpha$ becomes bright in $\beta$.[]{data-label="fig3"}](fig3.eps) The origin of this pattern is explained with the schematics shown in Fig. \[fig3\]. Assuming the total number of electrons to be $N_{\text{A}}-1$ or $N_{\text{A}}$ on dot A and $N_{\text{B}}-1$ or $N_{\text{B}}$ on dot B with ground state energies $E_{N\text{A}-1}$ or $E_{N\text{A}}$ and $E_{N\text{B}-1}$ or $E_{N\text{B}}$, two transitions are possible: $$\begin{aligned} E_{N\text{A}-1}&\leftrightarrow& E_{N\text{A}}\text{ with chemical potential }\\ &&\mu_{N\text{A}}=E_{N\text{A}}-E_{N\text{A}-1},\\ E_{N\text{B}-1}&\leftrightarrow& E_{N\text{B}}\text{ with chemical potential }\\ &&\mu_{N\text{B}}=E_{N\text{B}}-E_{N\text{B}-1}.\end{aligned}$$ If these chemical potentials equal those of the leads ($\mu_{\text{S}}$ and $\mu_{\text{D1}}$) lines are visible in the charging diagram. At $V_{\text{S}}=0$ (left schematic) $\mu_{\text{S}}$ and $\mu_{\text{D1}}$ are degenerate. Therefore each chemical dot potential produces a single line forming the typical hexagonal cells with the so called triple points (marked with e and h) at the edges. At e transport through the serial double quantum dot can be described by sequential tunneling of one electron at a time through the otherwise empty dot states. At h transport occurs by sequential tunneling of one hole at a time through the otherwise filled dot states. In between the two triple points the chemical potentials of both dots are equal ($\mu_{N\text{A}}=\mu_{N\text{B}}$) and an electron can move from dot B to dot A with increasing $V_{\text{G3}}$. As dot A is further away from the QPC, the QPC detects a decrease of charge. Thus a white feature is visible at the anticrossings in the center image of Fig. \[fig2\]. The nonlinear regime is described with the schematic on the right ($V_{\text{S}}>0$). The discussion for $V_{\text{S}}<0$ is analog. At $V_{\text{S}}>0$ the degeneracy of $\mu_{\text{S}}$ and $\mu_{\text{D1}}$ is lifted, $\mu_{\text{D1}}>\mu_{\text{S}}$. Therefore there are two possible resonance conditions for each chemical dot potential. But as each dot does only couple to one lead, only one resonance condition per dot is relevant. Thus still only one line is visible per ground state transition (with $\mu_{N\text{A}}=\mu_{\text{S}}$ and $\mu_{N\text{B}}=\mu_{\text{D1}}$). The other resonance conditions do not appear (dotted lines). However, as the two dots use different chemical lead potentials the anticrossings are shifted to the lower right as observed in the right image of Fig. \[fig2\]. Therefore the exchange of an electron between the dots does not appear at $\mu_{N\text{A}}=\mu_{N\text{B}}$, but at the dashed black line (right schematic). Left to this line there are two triangles (grey) where both chemical potentials, $\mu_{N\text{A}}$ and $\mu_{N\text{B}}$, are between $\mu_{\text{S}}$ and $\mu_{\text{D1}}$ and $\mu_{N\text{A}}<\mu_{N\text{B}}$ (Fig. \[fig3\], bottom left). These are the triangles described above. At the left border of these triangles the resonance condition $\mu_{N\text{A}}=\mu_{N\text{B}}$ is fulfilled opening a transport channel. Further transport channels within the triangles can appear due to excited atomic states. As an example we take into account two excited states per dot with total energies $E_{N-1}^*>E_{N-1}$ and $E_{N}^*>E_{N}$. Now two additional transitions are possible for each dot with new chemical potentials (Fig. \[fig3\], bottom left): $$\begin{aligned} E_{N-1}&\leftrightarrow& E_{N}^*\text{ with chemical potential } \mu_{N}^{*1}>\mu_{N},\\ E_{N-1}^*&\leftrightarrow& E_{N}\text{ with chemical potential } \mu_{N}^{*2}<\mu_{N}.\end{aligned}$$ Additional transport channels form for $V_{\text{S}}>0$, if the following resonance conditions are fulfilled (other channels are forbidden due to trapping): $$\begin{aligned} \mu_{N\text{A}}&=&\mu_{N\text{B}}^{*2},\\ \mu_{N\text{A}}^{*1}&=&\mu_{N\text{B}},\\ \mu_{N\text{A}}^{*1}&=&\mu_{N\text{B}}^{*2}.\end{aligned}$$ Together with the resonance condition $\mu_{N\text{A}}=\mu_{N\text{B}}$ those are the four solid lines drawn in each grey triangle in the schematic. They can appear in conductance measurements with electron-like transport in triangle $\alpha$ and hole-like transport in triangle $\beta$. With charge detection resonances are only visible, if they feature a different mean charge than what is given in the grey regions. Within the grey triangles at $V_{\text{S}}>0$ electrons can enter dot B via Drain1, holes can enter dot A via Source (electrons can leave A). Off resonance no transport between the dots is possible. Therefore the mean charge in the grey regions, added to the charge background of $N_{\text{A}}$ and $N_{\text{B}}$ electrons, is one electron on dot B. On resonance, transport between the dots is possible. Now the mean charge depends on the symmetry of the tunneling rates $\Gamma_{\text{S}}$ between Source and dot A, $\Gamma_{\text{D1}}$ between Drain1 and dot B and the interdot tunneling rate $\Gamma_{\text{AB}}$. Three different symmetries are possible that define the mean charge on resonance within triangles $\alpha$ and $\beta$:\ \ (i) $\Gamma_{\text{AB}}<\Gamma_{\text{S}}$, $\Gamma_{\text{D1}}$:\ $\alpha$: one electron on B.\ $\beta$: one electron on B.\ (ii) $\Gamma_{\text{D1}}<\Gamma_{\text{S}}$, $\Gamma_{\text{AB}}$:\ $\alpha$: no electron on A and B.\ $\beta$: one electron equally occupying both dots.\ (iii) $\Gamma_{\text{S}}<\Gamma_{\text{AB}}$, $\Gamma_{\text{D1}}$:\ $\alpha$: one electron equally occupying both dots.\ $\beta$: one electron on both dots each.\ \ In (i) no lines are visible with charge detection as the mean charge on resonance is identical to the mean charge off resonance in the grey regions. In (ii) and (iii) a resonance changes the mean charge in both triangles and becomes visible. Thus it is much more probable to observe excited states in weakly coupled double dots than in single dots, where excited states can only appear with symmetric tunneling rates [@Rogge-05]. Therefore (ii) or (iii) must be true for the measurements presented in Fig. \[fig2\]. A more detailed analysis reveals the actual ratio of tunneling rates. The bottom of Figure \[fig3\] shows a section of the right image in Fig. \[fig2\] with the triangles $\alpha$ and $\beta$ marked with white and black lines. As the difference between $\mu_{\text{S}}$ and $\mu_{\text{D1}}$ is bigger than the splitting of the anticrossing, both triangles overlap. Within the triangles the additional lines are visible. Following the line marked with arrows from bottom to top, one first observes a dark feature in triangle $\alpha$ and then a bright feature in triangle $\beta$. Thus the effect on the mean charge must be vice versa in both triangles. This is only possible in (iii) with a decrease of mean charge in (iii)$\alpha$ (as dot B is closer to the QPC than dot A) and an increase in (iii)$\beta$. Within the overlapped region an electron can enter the possibly empty dots via Drain1, as the system is within triangle $\alpha$. This electron can now leave the dots again via Source, or another electron can enter via Drain1, as the system is in triangle $\beta$ as well. As $\Gamma_{\text{S}}<\Gamma_{\text{D1}}$, it is much more probable for a second electron to enter the system than for the first one to leave. Therefore the process related to triangle $\beta$ is favored. ![(color online) left: section II from Fig. \[fig1\] at $V_{\text{S}}=-1$ mV. An anticrossing of dots A and C is visible (circle). Instead of triangular shaped patterns a splitting is observed on the left (ellipse). Right: schematic for the measurement on the left. The pattern is formed by a molecular state that is created at the anticrossing.[]{data-label="fig4"}](fig4.eps) Triangular patterns are visible for A-B over a wide range of parameters. They finally fade out with increasing gate voltages as the tunneling rates are changed. In contrast no such patterns appear for A-C, although measured under the same conditions within the same device. Instead a different pattern is found. The left of Figure \[fig4\] shows a measurement at $V_{\text{S}}=-1$ mV, taken within section II (as the lines of dot A appear steeper than those of B, but shallower than those of C, patterns of A-B at $V_{\text{S}}>0$ must be compared with patterns of A-C with $V_{\text{S}}<0$). The measurement shows an anticrossing (circle), that is almost not shifted compared to the one observed at $V_{\text{S}}=0$ (see Fig. \[fig1\]). Another striking feature is the step that appears on the left of the anticrossing (ellipse). The left line of the anticrossing disappears and comes up again with a huge offset to the left. There are no triangular shaped patterns or lines for excited states. The origin of this pattern is described using the schematic at the right of Fig. \[fig4\] assuming molecular bonds. With a relative width of anticrossings of $\approx 0.4$ (with 1 being the maximum [@Golden-96]) the double dot A-C is coupled stronger than A-B, which has a relative width of ca. 0.33. As for the schematics shown before for $V_{\text{S}}\neq 0$ the resonances for ground state transitions split into two resonances as the chemical potentials in the leads differ. Here with $V_{\text{S}}<0$ resonances with $\mu_{\text{S}}$ appear shifted to the lower left compared to those with $\mu_{\text{D2}}$. Far off the anticrossing states of the two dots can be described as atomic with chemical potentials $\mu_{N\text{A}}$ and $\mu_{N\text{C}}$. As dot A is coupled to Source and C to Drain2, the resonances $\mu_{N\text{A}}=\mu_{\text{S}}$ and $\mu_{N\text{C}}=\mu_{\text{D2}}$ must appear (solid straight lines). Due to the strong coupling of A and C the pattern around the anticrossing cannot be described with atomic states any longer. Instead a common symmetric molecular state with energy $E_{\text{s}}$ evolves that is extended over the whole double dot. With the double dot energies $E_0$ for no added electrons and $E_2$ for two electrons added, two new transitions are possible: $$\begin{aligned} E_0&\leftrightarrow& E_{\text{s}} \text{ with chemical potential } \mu_{\text{0,s}}<\mu_{N\text{A}}\text{, }\mu_{N\text{C}},\\ E_{\text{s}}&\leftrightarrow& E_2 \text{ with chemical potential } \mu_{\text{s,2}}>\mu_{N\text{A}}\text{, }\mu_{N\text{C}}.\end{aligned}$$ These new chemical potentials can create two resonances each, one with $\mu_{\text{S}}$, one with $\mu_{\text{D2}}$. Which of those involves a change of the mean charge depends on the tunneling rates again. If $\Gamma_{\text{D2}}<\Gamma_{\text{S}}$, charging appears at resonance with $\mu_{\text{S}}$. If $\Gamma_{\text{S}}<\Gamma_{\text{D2}}$, charging appears at resonance with $\mu_{\text{D2}}$ instead. The latter case results in the two curved solid lines in the schematic, that properly describes the experiment. In the area close to the anticrossing the double dot shows charging at resonance with $\mu_{\text{D2}}$ as well as dot C far off the anticrossing. Dot A shows charging at resonance with $\mu_{\text{S}}$ instead. Therefore a jump must occur when the system changes from the molecular common state to the atomic state of dot A. Two of those jumps are shown in the schematic, but only one is visible in the measurement as the other one is disturbed by a line of dot B. However, the symmetry of tunneling rates is detected for double dot A-C as well. Thus with charge measurements it is possible to detect the symmetry of tunneling rates for weakly and for strongly coupled double quantum dots. For the two double dots in this device the same symmetry was detected: $\Gamma_{\text{S}}<\Gamma_{\text{AB,AC}}$, $\Gamma_{\text{D1,D2}}$. However, depending on the strength of tunnel coupling two completely different patterns were found. Thus non-invasive charge measurement is capable of detecting molecular bonds in quantum dots. For the heterostructure we thank M. Bichler, G. Abstreiter, and W. Wegscheider. This work has been supported by BMBF via nanoQUIT. [20]{} natexlab\#1[\#1]{}bibnamefont \#1[\#1]{}bibfnamefont \#1[\#1]{}citenamefont \#1[\#1]{}url \#1[`#1`]{}urlprefix\[2\][\#2]{} \[2\]\[\][[\#2](#2)]{} , , , , , , in **, edited by , , (, , ), vol. of **, pp. . , , , , , , ****, (). , , , , , , ****, (). , ****, (). , , , , , , , , , ****, (). , , , , , ****, (). , ****, (). , , , , , ****, (). , , , , ****, (). , , , , , ****, (). , , , , , , , ****, (). , ****, (). , , , , , ****, (). , ****, (). , , , , , , ****, (). , , , , , ****, (). , , , , , , , ****, ().
{ "pile_set_name": "ArXiv" }
--- abstract: 'Full counting statistics (FCS) for the transport through a molecular quantum dot magnet is studied theoretically in the incoherent tunneling regime. We consider a model describing a single-level quantum dot, magnetically coupled to an additional local spin, the latter representing the total molecular spin $s$. We also assume that the system is in the strong Coulomb blockade regime, i.e., double occupancy on the dot is forbidden. The master equation approach to FCS introduced in Ref. is applied to derive a generating function yielding the FCS of charge and current. In the master equation approach, Clebsch-Gordan coefficients appear in the transition probabilities, whereas the derivation of generating function reduces to solving the eigenvalue problem of a modified master equation with counting fields. To be more specific, one needs only the eigenstate which collapses smoothly to the zero-eigenvalue stationary state in the limit of vanishing counting fields. We discovered that in our problem with arbitrary spin $s$, some quartic relations among Clebsch-Gordan coefficients allow us to identify the desired eigenspace without solving the whole problem. Thus we find analytically the FCS generating function in the following two cases: i) both spin sectors ($j=s\pm 1/2$) lying in the bias window, ii) only one of such spin sectors lying in the bias window. Based on the obtained analytic expressions, we also developed a numerical analysis in order to perform a similar contour-plot of the joint charge-current distribution function, which have recently been introduced in Ref. , here in the case of molecular quantum dot magnet problem.' author: - 'K.-I. Imura$^1$, Y. Utsumi$^1$, and T. Martin$^{1,2,3}$' title: Full counting statistics for transport through a molecular quantum dot magnet --- Introduction ============ Molecular electronics has emerged as one of the promising subfields of nanophysics. Individual nano objects such as molecules and carbon nanotubes can be connected to leads of different nature, and the degrees of freedom of molecules could be used in principle to modulate quantum transport between the leads [@nitzan; @book; @tsukada]: it is hoped that these degrees of freedom could allow specific functions of nanoelectronics. A large body of work has focused on the potential applications of the specific energy spectrum of molecules, as well as their vibrational degrees of freedom. [@some_reviews] Recently, there has been some focus on transport through molecules which have their own spin, for instance because a specific atom of this molecular system possesses a magnetic moment. This is unlike “usual” molecular quantum dots where the molecular spin corresponds to the effective spin of the itinerant electron, giving rise for example to Kondo physics. [@utsumi_martinek] There are many experimental proposals for such molecular magnets. Mn$_{12}$ acetates or carbon fullerenes with one or more atoms trapped inside are examples. Depending on the sample, and on the structure of the molecule, the spin orientation can have a preferred axis or it can have a more isotropic nature. There have been recent experiments on similar systems, with applications to spin valves. [@heersche; @romeike_incoh] In this paper we focus on the isotropic regime. As in Ref. [@elste_timm], where the current through an endohedral nitrogen doped fullerene was computed, we describe transport in the incoherent tunneling regime, and proceed to compute the full counting statistics for transport through a molecular quantum dot where the dot electrons have an exchange coupling with an impurity spin. Indeed, in mesoscopic physics, [@nazarov; @fazio] the current voltage characteristics alone is in general not sufficient to fully characterize transport. The current fluctuations can provide additional information on the correlations in molecular quantum dot, but so do the higher current moments, or cumulants, such as the “skewness” (third order) and the “sharpness” (fourth order). Such higher order cumulants add up to the so-called full counting statistics (FCS). [@levitov] Yet here we are interested in obtaining the FCS of such devices using the master equation approach. [@bagrets] We will also take the opportunity to present in detail the general framework for studying joint distributions of electron occupation of the dot and of current. A recent work by one of the authors [@utsumi_condmat] applied to a single level quantum dot shows that there are clear correlations between current and occupancy of the dot: there, the quantum fluctuations caused by increasing the tunnel coupling to the leads has a definite impact on the number distribution. Here we will extend these concepts to the case of a molecular quantum dot with an impurity spin. The FCS has remained until recently a highly theoretical concept, with few experimental applications. The third moment has been measured for a tunnel junction, [@reulet; @reznikov] as the effects of the measuring apparatus have been pointed out. [@kindermann] For the regime of incoherent transport through quantum dots two outstanding experiments have been performed: [@fujisawa; @gustavsson_ensslin] a point contact is placed in the close vicinity of the quantum dot, and by monitoring the current in this point contact, it is possible to obtain information on the time series of the current in the dot. From this time series, higher moments of the current can be computed. These recent experiments thus provide additional motivation to study the FCS in this novel system of a molecular quantum dot with an external spin. This paper thus aims at calculating the FCS of transport through a molecular quantum dot magnet, under the following circumstances: 1. The quantum dot is in the strong Coulomb blockade regime, i.e., the system is subject to strong electronic correlation. 2. The coupling between the dot and the reservoirs is relatively weak, and tunneling of electrons between them can be considered incoherent (incoherent tunneling regime). 3. The molecular spin is isotropic, and coupled to the spin of conduction electron through a simple exchange interaction. Assumption (2) allows us to employ the master equation approach for studying the transport through molecular quantum dot. As assumption (1) forbids double occupancy on the dot, we consider the cases where the dot is either empty, or occupied by only one conduction electron. When the dot is empty, the molecular spin $s$, together with its $z$-component $s_z$, is naturally a good quantum number, whereas if the dot is occupied by one conduction electron, then it becomes the total angular momentum $j=s+\sigma$ which is a good quantum number on the dot. Let us now follow the time evolution of the dot state, e.g., using a master equation. The number $n$ of conduction electrons on the dot fluctuates in time between $n=0$ and $n=1$. Such transitions are controlled by the so-called rate matrix $L$, which will be more carefully defined in Sec. II. This rate matrix $L$ can be decomposed into $2\times 2=4$ charge blocks , corresponding to two charge sectors $n=0,1$. Then each block is further characterized by either the molecular spin $s_z$ or the total angular momentum $j$ (and $j_z$) depending on whether $n=0$ or $n=1$. Naturally, off-diagonal blocks of the rate matrix describe transition between different angular momentum states. The rate of such transitions is proportional to the square of a Clebsch-Gordan coefficient between the initial and the final angular momentum states. In the master equation approach to FCS, the derivation of FCS generating function reduces to solving the eigenvalue problem of a modified master equation with counting fields. By using an algebraic identity, this eigenvalue problem can be projected down to a slightly more complicated problem within the $n=0$ charge sector, which can be fully characterized by the molecular spin $s_z$. Our main discovery is that some identities among Clebsch-Gordan coefficients allow us to identify the desired eigenspace without solving the whole problem. In this reduced problem Clebsch-Gordan coefficients appear at quartic order, since off-diagonal blocks are projected, together with the $n=1$ charge sector, down to the $n=0$ subspace. Based on the analytic expressions for the FCS generating function thus obtained, we also developed numerical analysis to perform a contour-plot of the joint charge-current distribution function. The paper is organized as follows. The first two sections after Introduction are devoted to the preparatory stage, i.e., Section II describes our model for the isotropic molecular quantum dot magnet, while Section III reviews the master equation approach to FCS. Then, in Section IV the analytic expressions for the FCS generating function will be obtained, with the help of quartic identities among Clebsch-Gordan coefficients. Section V describes the contour plot of joint distribution functions, before the paper is concluded in Section VI. Molecular quantum dot magnet ============================ We consider a molecule connected to two electrodes, which are specified by an index $\chi=\pm 1$. $\chi=1$ corresponds, e.g., to the left ($\chi=-1$ to the right) electrode. We focus on the strong Coulomb blockade regime, assuming that the number of electrons which can be added to the dot is restricted to one. Because the extra-charge electron state on the dot carries a spin, it is locally coupled to the impurity spin on the molecule via an exchange coupling Hamiltonian. Model Hamiltonian ----------------- The total Hamiltonian $H$ of the system consists of two parts: (i) the Hamiltonian $H_0$ of the isolated parts, and (ii) the tunneling Hamiltonian $H_{\rm tun}$, i.e., $H=H_0+H_{\rm tun}$. $H_0$ can be decomposed further into the dot and lead parts: $H_0=\sum_{\chi=\pm 1}H_{\rm lead}^{(\chi)}+H_{\rm dot}$. For normal metal leads, the Hamiltonian of left ($\chi=1$) and right ($\chi=-1$) leads reads, $H_{\rm lead}^{(\chi)}= \sum_{k\sigma} \epsilon_{k}^{(\chi)}c_{k\sigma}^{(\chi)\dagger}c_{k\sigma}^{(\chi)}$, where $c_{k\sigma}^{(\chi)}$ annihilates an electron with momentum $k$, spin $\sigma$ and energy $\epsilon_{k}^{(\chi)}$ in the lead $\chi$. The tunnel Hamiltonian reads: $H_{\rm tun}=\sum_{k\sigma\chi} T_{k\sigma}^{(\chi)} c_{k\sigma}^{(\chi)\dagger}d_\sigma + {\mathrm h.c.}$, where the operator $d$ ($d^\dagger$) annihilates (creates) an electron on the single dot level. The dot Hamiltonian reads: $$H_{\rm dot}= \epsilon_{\rm dot} \sum_{\sigma} n_{\sigma} + U n_{\uparrow}n_{\downarrow} + H_{\rm spin}, %\nonumber$$ where $n_{\sigma}=d^\dagger_{\sigma}d_{\sigma}$ is the number of electrons on the dot level with spin $\sigma$, and $U$ is a charging energy contribution related to the capacitance with respect to the leads and to the gate. It will be assumed to be large in the present work, in order to prohibit double occupancy. Finally, $H_{\rm spin}$ is the coupling of the dot electron spin $\vec{\sigma}$ with the impurity spin $s$. This is finite, by definition, only when there is an electron on the dot: $$H_{\rm spin}=\left\{ \begin{array}{l} 0\ \ \ (n=0) \\ g \vec{\sigma}\cdot\vec{s}\ \ \ (n=1) \end{array} \right., %\nonumber$$ where $g$ is the exchange coupling constant. As expected the eigenstates of this contribution to the Hamiltonian are the total angular momentum states of $\vec{j} \equiv \vec{\sigma}+\vec{s}$. This is seen by rewriting the exchange term as $\vec{\sigma}\cdot\vec{s}=(j^2-s^2-\sigma^2)/2$. The $n=0$ sector is characterized simply by the spin $s$ of the magnetic impurity, and has $2s+1$ degenerate states, $|n=0,s,s_z\rangle$. When $n=1$, the coupling $g$ between the impurity and the dot electron spins requires to diagonalize the interaction with the total angular momentum states, i.e., $|n=1,j=s\pm 1/2,j_z\rangle$. The two spin subsectors, $j=s+1/2$ and $j=s-1/2$ correspond to an energy eigenvalue of, respectively, $\epsilon_{s+1/2}=\epsilon_{\rm dot} + g s/2$ and $\epsilon_{s-1/2}=\epsilon_{\rm dot} - g (s+1)/2$. The two subsectors each have a degeneracy of $2s+2$ and $2s$, respectively. In our model Hamiltonian, the charging effects are taken into account via the bias-voltage dependence of the position of the dot level $\epsilon_{\rm dot}$ with respect to the chemical potentials of the leads, $\mu_{L,R}$. A gate voltage $V_g$ and gate capacitance $C_g$ can be included in order to move the dot levels up and down. Here we focus on the case where the bias voltage always dominates on the temperature, so for later purpose it is sufficient to specify which level and how many of these are included in the bias window. Incoherent tunneling regime - the master equation ------------------------------------------------- The electron dynamics of transport through a molecular quantum dot in the incoherent tunneling regime can be described by a master equation. [@kampen] In such systems as our molecular quantum dot Hamiltonian, the master equation describes a stochastic evolution in the space spanned by the eigenstates of the non-perturbed Hamiltonian $H_0$. Such states are, of course, of quantum mechanical nature, but the evolution of the system is governed by a deterministic stochastic process. Jumps between different states are to be stochastic events, and are also to be Markovian (no correlation between successive tunneling events). Such a situation is justified in the so-called incoherent tunneling regime, where the the non-perturbed Hamiltonian $H_0$ is only weakly perturbed by the tunneling Hamiltonian $H_{\rm tun}$, and the tunneling of electrons through the dot is still considered to be sequential. A simplification specific to our model is that the lead degrees of freedom can be traced out, and the master equation describes actually the time evolution of the dot states. We will often denote a (quantum mechanical) dot state $|n,j,j_z\rangle$, simply as $\alpha$ in the master equation. The probability $p_\alpha (t)$ with which the system, i.e., our molecular quantum dot, finds itself in a state $\alpha$ at time $t$ satisfies, $${d p_\alpha (t) \over dt} =\gamma_\alpha p_\alpha (t) +\sum_\beta \Gamma_{\alpha\beta}p_\beta (t). \label{mas}$$ $\Gamma_{\alpha\beta}$ is the transition probability for a jump, $\alpha\leftarrow\beta$, which can be calculated, e.g., from Fermi’s golden rule. The subscripts $\alpha$ and $\beta$ take all the possible values, $\alpha_1,\alpha_2,\cdots$ in the space of physically relevant states. Conservation of probability imposes that the sum of all the components [*in a given column*]{} $\beta$ of matrix $L$ vanishes. Thus $\gamma_\alpha=-\sum_\beta\Gamma_{\beta\alpha}$ denotes the decay rate of a state $\alpha$. In matrix notation, this reads $dp(t)/dt=Lp(t)$, where $L=\gamma+\Gamma$, is called the rate matrix. $\gamma$ and $\Gamma$ are, respectively, the diagonal and off-diagonal parts of the rate matrix $L$. In order to apply Fermi’s golden rule for evaluating $\Gamma_{\alpha\beta}$, one must go back to the Hamiltonian, $H_0$ of the isolated parts, and one consider the transition rate $\Gamma_{fi}$ associated with a tunneling perturbation $H_{\rm tun}$, from an initial state $|i\rangle$ to a final state $|f\rangle$: $$\Gamma_{fi}={2\pi\over\hbar} \big|\langle i|H_{\rm tun}|f\rangle\big|^2 \delta(E_i-E_f)~.. \label{gold}$$ The eigenstates $|i\rangle,|f\rangle$ are many electron states, but they factorize into a direct product of many-electron lead states and a dot state. Therefore, after taking into account all the lead degrees of freedom, the above transition rates are rewritten in the form of transition rates between different dot states $\alpha,\beta$, i.e., $\Gamma_{\alpha\beta}$ for a transition, $\alpha\leftarrow\beta$. We also make the standard assumption that the hopping amplitude $T_{k\sigma}^{(\chi)}$ depends weakly on energy for states close to the Fermi levels of the leads. For normal leads it is also spin independent. Thus Fermi’s golden rule gives basically a bare tunneling rates, defined as $\Gamma_{L,R}=2\pi \nu T_{L,R}^2$, weighted by a Clebsch-Gordan coefficient squared, which we will discuss later. $\nu$ is the (constant) density of states at the Fermi level in the leads. We also use the explicit notation $L,R$ rather than $\chi$ for specifying the leads. Before giving explicit matrix elements of the rate matrix $L$, it is convenient to decompose it into four blocks, corresponding to the two charge sectors $n=0,1$ as $$L= \left(\begin{array}{cc} L_{00} & L_{01} \\ L_{10} & L_{11} \end{array}\right) \equiv \left(\begin{array}{cc} A & B \\ C & D \end{array}\right). \label{ratemat}$$ The off-diagonal block matrices $B$ and $C$ are rectangular matrices of size, (i) $(2s+1)\times (4s+2)$ and (ii) $(4s+2)\times (2s+1)$, respectively. They describe tunneling of an electron either (i) out of the dot into one of the reservoirs, or (ii) onto the dot from one of the reservoirs. Both have two spin sectors, corresponding to two possible values of the total angular momentum $j=s\pm 1/2$, i.e., $B=(B_{s-1/2},B_{s+1/2})$ and $C=(C_{s-1/2},C_{s+1/2})^T$, where the $C^T$ is, for example, a transposed matrix of $C$. The $(s_z,j_z)$-component of the submatrices $B_{s\pm1/2}$ and the $(j_z,s_z)$-component of $C_{s\pm1/2}$ are given, respectively, as $$\begin{aligned} &&B_j(s_z,j_z)= \nonumber \\ &&\big[\Gamma_L\{1-f(\epsilon_j-\mu_L)\}+\Gamma_R\{1-f(\epsilon_j-\mu_R)\}\big] \nonumber \\ &&\times|\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 \label{b}\\ &&C_j(j_z,s_z)= \nonumber \\ &&\big[\Gamma_L f(\epsilon_j-\mu_L)+\Gamma_R f(\epsilon_j-\mu_R)\big] \nonumber \\ &&\times|\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2, \label{c}\end{aligned}$$ where $j=s\pm 1/2$, $j_z=-j,\cdots,j$. The diagonal blocks $A$ and $D$ are themselves diagonal matrices, because the dot Hamiltonian is diagonal in the chosen basis. Their elements are specified by recalling that the sum of all the components in a given column of $L$ vanishes identically, as a result of the conservation of probability. The matrix $D$ has two spin sectors, $D={\rm diag}(D_{s-1/2},D_{s+1/2})$. The two diagonal blocks $D_j$ are also diagonal matrices, $D_j(j_z,j'_z)=D_j(j_z,j_z)\delta (j_z,j'_z)$. Let us focus on the column of $L$ belonging to the sector $j$ and $j_z$. The conservation of probability associated with this column requires, $$\begin{aligned} D_j(j_z,j_z)&=&-\sum_{s_z=-s,\cdots,s}B_j(s_z,j_z)\nonumber \\ &=&-\Gamma_L\big[1-f(\epsilon_j-\mu_L)\big] -\Gamma_R\big[1-f(\epsilon_j-\mu_R)\big]~,\nonumber\\ \label{d}\end{aligned}$$ i.e., $D_j(j_z,j_z)$ actually does not depend on $j_z$: this simplification is due to the identity (\[jzquad\]). Similar quadratic relations among Clebsch-Gordan coefficients are also summarized in the Appendix. The diagonal elements of matrix $A$ can be determined in the same manner. Recall that $A(s_z,s'_z)=A(s_z,s_z)\delta(s_z,s'_z)$. This time one fixes $s_z$ in order to specify a particular column. Then, in order to apply the conservation law, we sum over the two spin sectors $j$ and their associated $j_z$, to find, $$\begin{aligned} &&A(s_z,s_z)=-\sum_{j=s\pm 1/2}\sum_{j_z=-j,\cdots,j}C_j(j_z,s_z) \nonumber \\ &&=-{2s\over 2s+1} \Big[\Gamma_L f(\epsilon_{s-1/2}-\mu_L)+\Gamma_R f(\epsilon_{s-1/2}-\mu_R)\Big] \nonumber \\ &&-{2s+2\over 2s+1} \Big[\Gamma_L f(\epsilon_{s+1/2}-\mu_L)+\Gamma_R f(\epsilon_{s+1/2}-\mu_R)\Big]. \label{a}\end{aligned}$$ Here we used the identity (\[szquad\]) in the Appendix. The explicit matrix elements of the rate matrix are thus given. The discussion now turns to the description of FCS in the framework of master equation. Master equation approach to FCS =============================== We are interested in the statistical correlations of such observables as charge $N(t)$ or current $I(t)$, measured during a time interval $t_0<t<t_0+\tau$. In the master equation approach, an observable $Q_f(t)$ ($f=1,2,\cdots; Q_1(t)=N(t),Q_2(t)=I(t),\cdots$), is introduced as a stochastic variable which evolves as a function of time through the master equation. On the other hand, what is experimentally relevant is their time average during the measurement time $\tau$, $$Q_f={1\over\tau}\int_{t_0}^{t_0+\tau} dt Q_f(t). %\nonumber \label{average}$$ The quantity of our eventual interest is the joint statistical distribution function $P(q_1,q_2,\cdots)$ of such observables averaged during time $\tau$, i.e., $$P(q_1,q_2,\cdots)=\big\langle \delta(q_1-Q_1)\delta(q_2-Q_2)\cdots \big\rangle, %\nonumber \label{jdf}$$ where $\langle\cdots\rangle$ represents a stochastic average, whose meaning will be more rigorously defined in Sec. III B. The joint distribution function $P(q_1,q_2,\cdots)$ is related to the generating function via a Fourier transformation: $$\begin{aligned} \sum_{q_1,q_2,\cdots} P(q_1,q_2,\cdots)\exp\Big[i\sum_{f=1,2,\cdots}q_f\Phi_f\Big] \nonumber \\ =\exp\big[\Omega(\Phi_1,\Phi_2,\cdots)\big], \nonumber \label{cgf}\end{aligned}$$ where $\Omega(\Phi_1,\Phi_2,\cdots)$ is called the cumulant generating function (CGF). The ($n_1,n_2,\cdots$)-th order cumulant $C_{n_1,n_2,\cdots}$ is deduced from the generating function, $\Omega(\Phi_1,\Phi_2,\cdots)$ as, $$C_{n_1,n_2,\cdots}=(-i)^n {\partial^{n_1+n_2+\cdots}\over\partial\Phi_1^{n_1}\partial\Phi_2^{n_2}\cdots} \Omega(\Phi_1,\Phi_2,\cdots)\Big|_{\Phi_1,\Phi_2,\cdots=0}. \nonumber$$ Note that the CGF, $\Omega(\Phi_1,\Phi_2,\cdots)$ and the full distribution function $P(q_1,q_2,\cdots)$ contain the same information. In the following, we will focus on how to calculate such CGF in the framework of the master equation approach. The purpose of this section is to review the master equation approach to FCS developed in Ref. , and adapt it to a more specific case of our interest. A close comparison with the time-dependent perturbation theory of quantum mechanics reduces the calculation of CGF to solving a modified master equation with counting fields. Two types of observables and their counting statistics ------------------------------------------------------ In Secs. IV and V we discuss counting statistics of charge $N$ and current $I$. Other observables, such as the total angular momentum $J$ on the dot, its $z$-component $J_z$, or the spin current $I_s$ through the dot, can be handled in a similar way, which we outline below. Note that we can make a distinction between two category of observables: “charge-like” and “current-like” observables. The charge $N$ (or $J,\cdots$) is a property of the state of the dot, whereas the charge current, $I_c$ (or $I_s,\cdots$) is associated with transitions between different occupational states. The former has eigenvalues of a quantum mechanical operator that can be simultaneously diagonalized with $H_0$, whereas the latter is related to the nature of the transitions. We may also classify the counting fields $\Phi_1,\Phi_2,\cdots$ into two categories: $\xi_1,\xi_2,\cdots$ and $\eta_1,\eta_2,\cdots$. In order to give an unambiguous meaning to the stochastic average $\langle\cdots\rangle$ in Eq. (\[cgf\]) or (\[jdf\]) in the context of master equation, one has to go one step [*backward*]{} in regard to Eq. (\[average\]), i.e., one has to follow the time evolution of observables $Q_f(t)$. One may rewrite Eqs. (\[jdf\]) and (\[cgf\]) as $$\begin{aligned} &&\exp\Big[\Omega(\xi_1,\xi_2,\cdots; \eta_1,\eta_2,\cdots)\Big]= \nonumber \\ &&\Big\langle\exp\Big[ i\sum_f \xi_f\int dt N_f(t) +i\sum_{f'}\eta_{f'}\int dt I_{f'}(t) \Big]\Big\rangle~,\nonumber\\ \label{cgf+}\end{aligned}$$ where the time integrals should be performed, e.g., from $t_0$ to $t_0+\tau$. The stochastic average can be performed by giving more concrete expressions to the observables $N_f(t)$ and $I_{f'}(t)$ using the language of stochastic process. Let us focus on a particular realization of the stochastic process by denoting $\{t_j\}$ the sequence of times at which jumps between different states occur. $\{\alpha_j\}$ denotes the sequence of states such that the system stays in state $\alpha_j$ during the period of $t_j<t<t_{j+1}$. A state $|\alpha\rangle$ is specified by the quantum numbers $n_1,n_2,\cdots$, (e.g., $n_1=n$, the charge, $n_2=j$, the total angular momentum, $n_3=j_z$, its $z$-component, etc...). On the other way around, one denotes by $n_1(\alpha),n_2(\alpha),\cdots$ the quantum numbers for specifying the state $\alpha$. With this notation, $N_f(t)$ can be written explicitly for a Markovian sequence $(\{t_j\}, \{\alpha_j\})$: $$N_f(t)=\sum_j n_f(\alpha_j)\theta(t-t_j)\theta(t_{j+1}-t), \label{nf}$$ where $\theta(t)$ is the Heaviside function. A little more care is needed to give a similar expression to current-like observables, reflecting the fact that they have a direction. In order to clarify this point, let us consider the case of our molecular quantum dot. The transition of the dot state is caused by tunneling of an electron onto or out of the dot. The change of the dot state before and after the transition is completely specified by the change of the number of electrons $\Delta n$, and of the total spin $\Delta j$ together with its $z$-component $\Delta j_z$ on the dot. However, in order to define a current associated with the transition, we need additionally an information on the direction of flow, i.e., an information on where the electron goes, to which lead, or from which lead. To measure the current, one has to decide also where one measures it. Such ambiguities are usually removed by considering a [*net*]{} current which flows [*out of*]{} a specific lead, say, $\chi$, i.e., by considering $I^{(\chi)}(t)$. A current $I_f^{(\chi)}=\Delta n_f$ is induced by a tunneling, leading to the change of the dot state $\Delta n_f$, such as $\Delta n$, $\Delta j$ and $\Delta j_z$, which occurs also through the lead $\chi$. This suggest in our Markovian sets $(\{t_j\}, \{\alpha_j\})$ to add $\{\chi_j\}$ specifying through which lead the jump at time $t_j$ occurs. Once such Markovian sets $(\{t_j\}, \{\alpha_j\},\{\chi_j\})$ are given, one can provide explicit formula to current-like observables: $$I_f^{(\chi)}(t)=\sum_j \Delta n_f (t_j) \delta (\chi-\chi_j) \delta (t-t_j), \label{if}$$ where $\Delta n_f (t_j)=n_f(\alpha_{j+1})-n_f(\alpha_{j})$ specifies the nature of transition at time $t=t_j$. Let us consider the case of our model again. We separated the rate matrix $L$ into two parts: the diagonal part $\gamma$ and the off-diagonal part $\Gamma$. We also introduced a decomposition into four charge sectors, which can be naturally applied to $\gamma$ and $\Gamma$ as well: $$\begin{aligned} \gamma= \left(\begin{array}{cc} \gamma_0 & 0 \\ 0 & \gamma_1 \end{array}\right) = \left(\begin{array}{cc} L_{00} & 0\\ 0 & L_{11} \end{array}\right), \nonumber \\ \Gamma= \left(\begin{array}{cc} 0 & \Gamma_{01} \\ \Gamma_{10} & 0 \end{array}\right) = \left(\begin{array}{cc} 0 & L_{01} \\ L_{10} & 0 \end{array}\right). \nonumber\end{aligned}$$ Let us focus on the off-diagonal block matrices in $\Gamma$, $\Gamma_{01}$ and $\Gamma_{10}$. These two blocks correspond simply to different $\Delta n$, i.e., $\Gamma_{10}$ to $\Delta n=1$, and $\Gamma_{01}$ to $\Delta n=-1$. Such submatrices of $\Gamma$ still contain contribution from different leads, e.g., one can further decompose $\Gamma_{10}$ into $\Gamma_{10}=\Gamma_{10}^{(L)}+\Gamma_{10}^{(R)}$. The whole $\Gamma$ can be also decomposed with respect to $\chi$, i.e., $\Gamma=\sum_\chi\Gamma^{(\chi)}$, whereas one can divide the matrix $\Gamma$ into smaller subsectors in terms of different $\Delta n_f$. Thus very generally one can decompose the transition matrix $\Gamma$ as $$\Gamma=\sum_{\chi,\{\Delta n_f\}}\Gamma^{(\chi)}_{\{\Delta n_f\}},$$ where $\{\Delta n_f\}=(\Delta n,\Delta j,\cdots)$. Master equation with counting fields — prescription for obtaining the FCS ------------------------------------------------------------------------- Substituting Eqs. (\[nf\]) and (\[if\]) into Eq. (\[cgf+\]) one finds the expression to be evaluated: $$\begin{aligned} \exp\Big[\Omega(\xi_1,\xi_2,\cdots; \eta_1,\eta_2,\cdots)\Big]= \nonumber \\ \Big\langle\exp\Big[ i\sum_f \xi_f \sum_{j=0}^n (t_{j+1}-t_j) n_f(\alpha_j)+ \nonumber \\ i\sum_{f}\eta_{f}\sum_{j=1}^n \Delta n_f (t_j) \delta (\chi-\chi_j) \Big]\Big\rangle. \label{cgf++}\end{aligned}$$ We need a prescription to compute the stochastic average $\langle\cdots\rangle$. This is achieved by performing a perturbative expansion of the master equation with respect to off-diagonal components $\Gamma$. The resulting perturbation series is analogous to that of the time-dependent perturbation theory of quantum mechanics. Note that the diagonal part $\gamma$ and the off-diagonal part $\Gamma$ of the coefficient matrix $L=\gamma+\Gamma$ generally [*do not commute*]{}. Because our observables are series of events which occur successively on the time axis, the master equation is rewritten in a way which is analogous to the “interaction representation” of quantum mechanics, $$\begin{aligned} {d\over dt}\tilde{p}(t)&=&\tilde{\Gamma}(t)\tilde{p}(t), \nonumber \\ \tilde{p}(t)&=&\exp \big[\gamma(t-t_0)\big] p(t_0), \nonumber \\ \tilde{\Gamma}(t)&=&e^{-\gamma(t-t_0)}\Gamma e^{\gamma(t-t_0)}. \label{intrep}\end{aligned}$$ Developing a perturbative expansion with respect to $\Gamma$, one finds the analog of the Dyson series for $\tilde{p}(t)$: $$\begin{aligned} &&\tilde{p}(t)=p(t_0)+\int_{t_0}^{t} dt' \tilde{\Gamma}(t)\tilde{p}(t') \nonumber \\ &&=\sum_{n=0}^\infty \int_{t_0<t_1<t_2<\cdots<t_n<t} dt_1dt_2\cdots dt_n \nonumber \\ &&~~\times\tilde{\Gamma}(t_n)\cdots \tilde{\Gamma}(t_2)\tilde{\Gamma}(t_1) p(t_0), \label{dyson}\end{aligned}$$ where we used $\tilde{p}(t_0)=p(t_0)$. Let us focus on the $n$-th order term in $\tilde{p}(t)$. Because Eq. (\[dyson\]) is written in matrix notation, we can further develop it as: $$\begin{aligned} &&\tilde{\Gamma}(t_n)\cdots \tilde{\Gamma}(t_2)\tilde{\Gamma}(t_1) \Big|_{\alpha_{n}\alpha_{0}}= \nonumber \\ &&\sum_{\alpha_{n-1},\cdots,\alpha_2,\alpha_1} \tilde{\Gamma}_{\alpha\alpha_{n-1}}(t_n)\cdots\tilde{\Gamma}_{\alpha_2\alpha_1}(t_2) \tilde{\Gamma}_{\alpha_1\alpha_0}(t_1). \label{jumps}\end{aligned}$$ One can identify such terms as the realization of the stochastic process characterized by the sets $\{t_j\}$ and $\{\alpha_j\}$. Indeed, each of the $n$-th order term in Eq. (\[dyson\]) corresponds to a sequence, $\alpha_0\rightarrow \alpha_1\rightarrow \alpha_2\rightarrow \cdots$, with $n$ jumps located at times $\{t_j\}$. As the matrix $\gamma$ is the diagonal part of $L$, the system remains in the same state $\alpha_j$ with probability $e^{\gamma_{\alpha_0} (t_{j+1}-t_j)}$ between the two jumps $t_{j}<t<t_{j+1}$, whereas the transition $\alpha_{j-1}\rightarrow \alpha_j$ occurs at time $t_j$. Thus the probability for a sequence of events to occur is found to be, $$\begin{aligned} &&P(\alpha_0\rightarrow \alpha_1\rightarrow \alpha_2\rightarrow \cdots)= \nonumber \\ &&\cdots e^{\gamma_{\alpha_2} (t_3-t_2)} \Gamma_{\alpha_2\alpha_1} e^{\gamma_{\alpha_1} (t_2-t_1)} \Gamma_{\alpha_1\alpha_0} e^{\gamma_{\alpha_0} (t_1-t_0)} p_{\alpha_0}(t_0). \nonumber\end{aligned}$$ The above identification in the framework of Eqs. (\[intrep\]) and (\[dyson\]) allows us to give an unambiguous meaning to the stochastic average of observables in Eq. (\[cgf++\]). A prescription for calculating the average is given below. The first term in the exponential of Eq. (\[cgf++\]) consists of charge-like observables characterizing the state of the system for $t_{j}<t<t_{j+1}$. This contribution can be [*absorbed*]{} in the diagonal components of L, i.e., by the substitution: $$\gamma_\alpha\rightarrow \gamma_\alpha(\xi_1,\xi_2,\cdots)\equiv \gamma_\alpha+\sum_f n_f(\alpha)\xi_f \label{subs1}$$ If one considers the distribution of charge in our molecular quantum dot, this reduces simply to the replacement, $$\gamma= \left(\begin{array}{cc} \gamma_0 & 0 \\ 0 & \gamma_1 \end{array}\right) \rightarrow \gamma(\xi)= \left(\begin{array}{cc} \gamma_0 & 0 \\ 0 & \gamma_1+\xi \end{array}\right). \nonumber$$ The second term in the exponential of Eq. (\[cgf++\]) is, on the other hand, composed of current-like observables which are associated with transitions $\Gamma_{\alpha\beta}$ between different states. In the perturbative expansion with respect to such jumps, these appear only at specific times $t=t_j$. Eq. (\[jumps\]) suggests that these contribution can be absorbed, in the off-diagonal part $\Gamma$ as follows: $$\Gamma^{(\chi)}_{\{\Delta n_f\}}\rightarrow \Gamma^{(\chi)}_{\{\Delta n_f\}}(\eta_1,\eta_2,\cdots)\equiv \Gamma^{(\chi)}_{\{\Delta n_f\}}\exp\Big[i \sum_f\Delta n_f \eta_f\Big]. \label{subs2}$$ Note that only a part of $\Gamma$ associated with the lead $\chi$ which is involved in the measurement, i.e., $\Gamma^{(\chi)}$ is subjective to the shift rule. In our molecular quantum dot, $\Gamma$ has the following simple structure, $$\Gamma= \left(\begin{array}{cc} 0 & \Gamma_{01}^{(L)} \\ \Gamma_{10}^{(L)} & 0 \end{array}\right)+ \left(\begin{array}{cc} 0 & \Gamma_{01}^{(R)} \\ \Gamma_{10}^{(R)} & 0 \end{array}\right). \nonumber$$ One also adopts the convention to measure the [*net*]{} current which flows [*out of*]{} the left reservoir onto the quantum dot. Then, the above replacement leads to, $$\Gamma\rightarrow \Gamma(\eta)= \left(\begin{array}{cc} 0 & \Gamma_{01}^{(L)}e^{-i\eta}\\ \Gamma_{10}^{(L)}e^{i\eta} & 0 \end{array}\right)+ \left(\begin{array}{cc} 0 & \Gamma_{01}^{(R)} \\ \Gamma_{10}^{(R)} & 0 \end{array}\right). \nonumber$$ The above arguments show that calculation of the CGF is actually equivalent to solving a stochastic problem, specified by the following modified master equation, $$\begin{aligned} {d\over dt}p(\xi,\eta;t)&=&L(\xi,\eta)p(\xi,\eta;t), %\nonumber \\ L(\xi,\eta)&=&\gamma (\xi_1,\xi_2,\cdots) +\Gamma (\eta_1,\eta_2,\cdots), \label{masmod}\end{aligned}$$ where $p(\xi,\eta;t)$ is a vector: $$p(\xi,\eta;t)=\left( \begin{array}{c} p_{\alpha_1}(\xi,\eta;t) \\ p_{\alpha_2}(\xi,\eta;t) \\ \vdots \end{array} \right). \nonumber$$ Its $(\xi,\eta)$-dependence is only written symbolically, but its implication would be clear: $(\xi,\eta)=(\xi_1,\xi_2,\cdots; \eta_1,\eta_2,\cdots)$. For a given initial condition $p(\xi,\eta;t_0)=p(t_0)$, the CGF is simply related to $p(\xi,\eta;t)$ as, $$\exp\Big[\Omega(\xi,\eta)\Big] =\sum_\alpha p_\alpha(\xi,\eta;t_0+\tau). %\label{} \nonumber$$ The modified master equation (\[masmod\]) can also be integrated formally to give, $$p(\xi,\eta;t)=\exp\Big[(t-t_0)L(\xi,\eta)\Big]p(t_0). %\label{} \nonumber$$ Thus for a long measurement time $\tau$, the maximal eigenvalue $\lambda_M(\xi,\eta)$ of the modified rate matrix $L(\xi,\eta)$ dominates, i.e., $$\Omega(\xi,\eta)\simeq\tau\lambda_M (\xi,\eta). \label{cgfmax} %\nonumber$$ In the limit of vanishing counting fields: $\xi,\eta \rightarrow 0$, one can verify that $\lambda_M (\xi,\eta) \rightarrow 0$, the latter corresponding to the stationary state solution. We have thus established a recipe for obtaining FCS. We start with a master equation, Eq. (\[mas\]), controlled by a rate matrix with its explicit elements calculated through the Fermi’s golden rule. We then make the replacements (\[subs1\],\[subs2\]) to define a new stochastic problem in terms of a modified master equation with counting fields: Eq. (\[masmod\]). Finally, our task reduces to finding the maximal eigenvalue, $\lambda_M (\xi,\eta)$ of the modified rate matrix $L(\xi,\eta)$ . The eigenspace corresponding to $\lambda_M (\xi,\eta)$ is smoothly connected to the stationary state in the limit of vanishing counting fields. The analytic solution ===================== The master equation describing an isotropic molecular quantum dot magnet in the incoherent tunneling regime was derived in Sec. II. A prescription for calculating FCS of transport through a quantum dot starting from such a master equation was given in Sec. III. The purpose of this section is to apply the formalism developed in Sec. III to the specific case of our molecular quantum dot magnet and obtain an analytic expression for the FCS of charge and current. In practice we have to solve an eigenvalue problem for the [*modified master equation with counting fields*]{}, which can be constructed by making suitable replacements of the diagonal and off-diagonal components as Eqs. (\[subs1\]) and (\[subs2\]). In this work, we will concentrate on the distribution of charge and current in our molecular quantum dot magnet. Our modified rate matrix is, therefore, a function of only two counting fields, say, $\xi$ and $\eta$, i.e., $$L(\xi,\eta)= \left(\begin{array}{cc} \gamma_0 & \Gamma_{01}^{(L)}e^{-i\eta}+\Gamma_{01}^{(R)}\\ \Gamma_{10}^{(L)}e^{i\eta}+\Gamma_{10}^{(R)} & \gamma_1+\xi \end{array}\right) \label{modratemat} %\nonumber$$ Its eigenvalues and the corresponding eigenvectors are also functions of $\xi$ and $\eta$: $$L(\xi,\eta) u(\xi,\eta)=\lambda(\xi,\eta)u(\xi,\eta). \label{eigen} %\nonumber$$ The FCS of our problem is obtained as a cumulant generating function of the charge-current joint correlation functions, which is identified as the maximal eigenvalue $\lambda(\xi,\eta)$ of the modified rate matrix, Eq. (\[modratemat\]). In the limit of vanishing counting fields $\xi,\eta \rightarrow 0$, the eigenstate $u(\xi,\eta)$ corresponding to $\lambda(\xi,\eta)$, collapses smoothly on the stationary state $u_0$, i.e., $$\begin{aligned} \lim_{\xi,\eta \rightarrow 0}u(\xi,\eta)=u(0,0)=u_0, \nonumber \\ \lim_{\xi,\eta \rightarrow 0}\lambda(\xi,\eta)=\lambda(0,0)=0. \label{solstat}\end{aligned}$$ In the same limit, Eq. (\[modratemat\]) reduces to Eq. (\[ratemat\]). Obtaining FCS of a system thus does not require to solve completely the modified master equation, but simply to identify the maximal eigenvalue, corresponding to the stationary state solution in the limit of vanishing counting fields. It is still rather surprising that one has an access to an analytic solution of such a problem for arbitrary $s$, because: 1. The size of the matrix one has to diagonalize is “large”, e.g., $6\times 6$ for $s=1/2$, and $3(2s+1)\times 3(2s+1)$ for a general $s$. 2. For arbitrary $s$, one has to diagonalize such a matrix of that size, and it is generally not intuitive whether the obtained eigenvalues have a closed form as a function of $s$. For $s=1/2$ the $6\times 6$ eigenvalue problem, Eqs. (\[eigen\]) and (\[modratemat\]), can be treated analytically by reducing the size of the problem down to $2\times 2$, using the identity, $$\det \left(\begin{array}{cc} A & B \\ C & D \end{array}\right) = \det (A-BD^{-1}C), \label{identity}$$ which is valid when $\det D\ne 0$. Here the four block matrices correspond to different charge sectors. One can apply the same identity to the case of arbitrary spin $s$, in which the size of the problem is reduced from $3(2s+1)\times 3(2s+1)$ down to $(2s+1)\times (2s+1)$, because the identity (\[identity\]) projects the whole physical space onto the $(0,0)$-charge sector. Of course, one has to pay the “price” for this reduction of size: the appearance of $BD^{-1}C$ term, in which information on the original larger matrix is compressed. The $BD^{-1}C$ term is, as it should be, a square matrix of size $(2s+1)\times (2s+1)$ (though matrices $B$ and $C$ are generally not square). Each row and column of this matrix are characterized by a set of quantum numbers $(s_z,s'_z)$. As a consequence of angular momentum selection rules, the matrix $BD^{-1}C$, or what we will call later $X_s$ or $Y_{s,j}$, (proportional to $BD^{-1}C$), becomes a [*tridiagonal*]{} matrix, i.e., only $(s_z,s_z)$, $(s_z,s_z+1)$ and $(s_z,s_z-1)$ elements are non-zero. The tridiagonal matrices, $X_s$ and $Y_{s,j}$, have a very characteristic structure (see Appendix), which allows us to identify one of its eigenvectors by inspection, Eq. (\[xseigen\]) or (\[yseigen\]). We will see that this eigenvector gives also a solution of Eqs. (\[eigen\]) and (\[modratemat\]), which turns out, [*accidentally*]{}, to be the right solution satisfying the condition (\[solstat\]). The matrices $X_s$ and $Y_{s,j}$ consist of Clebsch-Gordan coefficients at fourth order. The above characteristic feature of those matrices are actually due to some unconventional quartic relations among Clebsch-Gordan coefficients, which are summarized in the Appendix. The explicit matrix elements of Eq. (\[ratemat\]) is given as four constituent block matrices $A-D$ in Eqs. (\[b\])-(\[a\]). These formulas contain Fermi distribution functions, and therefore, depend on the relative positions of dot levels with respect to the chemical potential of two leads. At zero temperature such formulas can be further specified in specific situations which apply when the temperature is small compared to the voltage bias: (i) two spin sectors in the bias window, (ii) only one spin sector in the bias window (see Fig. \[fig:3states\]). Two spin subsectors in the bias window -------------------------------------- We first consider the case of two spin subsectors in the bias window, $\mu_L > \left(\epsilon_{s+1/2},\epsilon_{s-1/2}\right) > \mu_R$. At zero temperature, Eqs. (\[b\])-(\[a\]) become $$\begin{aligned} A(s_z,s_z)&=&-\sum_{j=s\pm 1/2}\sum_{j_z=-j,\cdots,j}C_j(j_z,s_z)=-2\Gamma_L, %\label{a2} \nonumber \\ B_j(s_z,j_z)&=&\Gamma_R|\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2, %\label{b2} \nonumber \\ C_j(j_z,s'_z)&=&\Gamma_L|\langle s,s'_z;1/2,\sigma'_z|j,j_z\rangle|^2, %\label{c2} \nonumber \\ D_j(j_z,j_z)&=&-\sum_{s_z=-s,\cdots,s}B_j(s_z,j_z) =-\Gamma_R. \label{abcd}\end{aligned}$$ The rate matrix (\[ratemat\]) has the simple form: $$L(\xi,\eta)= \left(\begin{array}{cc} A & B \\ Ce^{i\eta} & D+\xi I \end{array}\right). \nonumber$$ With the use of identity (\[identity\]), the eigenvalue problem defined by Eqs. (\[eigen\]) and (\[modratemat\]) reduces to finding $\lambda$ which satisfies, $$\det \big[ A-\lambda I-B\{D+(\xi-\lambda)I\}^{-1}Ce^{i\eta} \big]=0. %\label{detred} \nonumber$$ As mentioned before, the $BD^{-1}C$-like term is a $(2s+1)\times (2s+1)$, tridiagonal matrix with finite matrix elements only at $(s_z,s_z)$, $(s_z,s_z+1)$ and $(s_z,s_z-1)$. Recalling Eq. (\[abcd\]), one finds, $$B\{D+(\xi-\lambda)I\}^{-1}Ce^{i\eta} =-{\Gamma_L \Gamma_R e^{i\eta}\over \Gamma_R+\lambda-\xi} X_s, \nonumber$$ where $X_s$ is a $(2s+1)\times (2s+1)$ matrix, whose explicit matrix elements are, $$\begin{aligned} X_s (s_z,s'_z)&=& \sum_{j=s\pm 1/2}\sum_{j_z=-j,\cdots,j} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 \nonumber \\ &\times& |\langle s,s'_z;1/2,\sigma'_z|j,j_z\rangle|^2, \label{xs}\end{aligned}$$ The reduced eigenvalue equation now reads, $$%\det \big[ A-\lambda I-B \{D- (\xi+\lambda) I\}^{-1}C \big]= \det \big[ (-2\Gamma_L-\lambda) I +{\Gamma_L\Gamma_R e^{i\eta}\over \Gamma_R+\lambda-\xi} X_s\big]=0. \nonumber$$ As is explained in the Appendix, this matrix $X_s$ has the striking property: $$X_s \left(\begin{array}{c} 1 \\ 1 \\ \vdots \\ 1 \end{array}\right) =2 \left(\begin{array}{c} 1 \\ 1 \\ \vdots \\ 1 \end{array}\right), \label{xseigen}$$ valid for arbitrary $s$. The vector $v_0=(1,1,\cdots,1)^T$ is thus an eigenvector of $X_s$ with an eigenvalue $2$ for all $s$, but it is also an eigenvector of $A-\lambda I$. This means that if $\lambda=\lambda(\xi,\eta)$ is a solution of the quadratic equation, $${(2\Gamma_L+\lambda)(\Gamma_R+\lambda-\xi) \over \Gamma_L \Gamma_R e^{i\eta}}=2, \label{quad}$$ then $v_0$ satisfies, $$\big[ A-\lambda I-B \{ D+(\xi-\lambda) I\}^{-1}C e^{i\eta} \big]v_0=0, %\label{} \nonumber$$ Consequently, this solution $\lambda$ is a solution of the original eigenvalue problem defined by Eqs. (\[modratemat\]) and (\[eigen\]). However, at this point it is still only “a” solution of the original eigenvalue problem. The remarkable feature of Eq. (\[quad\]) is that in the limit of $\xi,\eta\rightarrow 0$, if one takes also the limit of vanishing counting fields: $\lambda\rightarrow 0$, then on the left hand side, the numerator and the denominator cancel each other simply to give a numerical factor 2, which coincides the right hand side of the equation. This implies that a solution of the quadratic equation (\[quad\]) is smoothly connected to the stationary state solution, which satisfies Eq. (\[solstat\]). Indeed, one of the solutions of Eq. (\[quad\]): $$\begin{aligned} \lambda(\xi,\eta)&=&-{2\Gamma_L+\Gamma_R-\xi\over 2} \nonumber \\ &+&{1\over 2} \sqrt{(2\Gamma_L-\Gamma_R+\xi)^2+8\Gamma_L\Gamma_R e^{i\eta}} \label{cgf2}\end{aligned}$$ satisfies Eq. (\[solstat\]). This means that Eq. (\[cgf2\]) is actually the desired cumulant generating function yielding the FCS. Thus we are able to [*identify*]{} the suitable eigenvalue of the problem formulated as Eqs. (\[eigen\]) and (\[modratemat\]) simply by inspection. The obtained result (\[cgf2\]) for general $s$ is identical to the $s=1/2$ case, because the corresponding eigenvalue of $X_s$, i.e., Eq. (\[xseigen\]) is not dependent on $s$. Note that Eq. (\[cgf2\]) is also equivalent to the the CGF calculated for spinful conduction electrons in the absence of local (molecular) spin. [@bagrets] Only one spin subsector in the bias window ------------------------------------------ Consider now the case of only one spin subsector, say, $j=s-1/2$ in the bias window, i.e., $\epsilon_{s+1/2} > \mu_L > \epsilon_{s-1/2} > \mu_R$. This corresponds to the case (ii) in Fig. \[fig:3states\]. In this case, Eqs. (\[b\])-(\[a\]) reduce, at zero temperature, to: $$\begin{aligned} A(s_z,s_z)&=&-{2s\over 2s+1}\Gamma_L, \nonumber \\ B_{s-1/2}(s_z,j_z)&=&\Gamma_R|\langle s,s_z;1/2,\sigma_z|j=s-1/2,j_z\rangle|^2, \nonumber \\ B_{s+1/2}(s_z,j_z)&=&(\Gamma_L+\Gamma_R) \nonumber \\ &\times& |\langle s,s_z;1/2,\sigma_z|j=s+1/2,j_z\rangle|^2, \label{b1} \nonumber \\ C_{s-1/2}(j_z,s'_z)&=&\Gamma_L|\langle s,s'_z;1/2,\sigma'_z|j=s-1/2,j_z\rangle|^2, \nonumber \\ C_{s+1/2}(j_z,s'_z)&=&0, \label{c1} \nonumber \\ D_{s-1/2}(j_z,j_z)&=&-\Gamma_R, \nonumber \\ D_{s+1/2}(j_z,j_z)&=&-\Gamma_L-\Gamma_R, \nonumber\end{aligned}$$ where $\sigma_z$ and $\sigma'_z$ are not independent variables, but they are determined automatically by the constraints: $j_z=s_z+\sigma_z$ and $j_z=s'_z+\sigma'_z$, respectively. The modified rate matrix can be obtained as Eq. (\[modratemat\]). Again we use the identity, Eq. (\[identity\]). Note that $C_{s+1/2}=0$. This allows us to rewrite the $BD^{-1}C$ term as, $$-{B_{s-1/2}C_{s-1/2}e^{i\eta}\over \Gamma_R+\lambda-\xi} =-{\Gamma_L \Gamma_R e^{i\eta}\over \Gamma_R+\lambda-\xi} Y_{s,s-1/2}, \nonumber$$ where $Y_{s,j}$ is a $(2s+1)\times (2s+1)$ square matrix, whose $(s_z,s'_z)$-elements are, $$\begin{aligned} Y_{s,j} (s_z,s'_z)= \sum_{j_z=-j,\cdots,j} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 \nonumber \\ \times |\langle s,s'_z;1/2,\sigma'_z|j,j_z\rangle|^2. \label{ys}\end{aligned}$$ The eigenvalue equation reads, $$\det \Bigg[ \left( -{2s\over 2s+1}\Gamma_L-\lambda\right) I +{\Gamma_L\Gamma_R e^{i\eta}\over \Gamma_R+\lambda-\xi} Y_{s,s-1/2}\Bigg]=0. \label{redeigen1} %\nonumber$$ Another situation to be considered in parallel is the case of only $j=s+1/2$ spin sector in the bias window, i.e., $\epsilon_{s-1/2} > \mu_L > \epsilon_{s+1/2} > \mu_R$. This corresponds to the case (iii) in Fig. \[fig:3states\]. In this case, Eqs. (\[b\])-(\[a\]) reduce, at zero temperature, to, $$\begin{aligned} A(s_z,s_z)&=&-{2s+2\over 2s+1}\Gamma_L, \nonumber \\ B_{s-1/2}(s_z,j_z)&=&(\Gamma_L+\Gamma_R) \nonumber \\ &\times& |\langle s,s_z;1/2,\sigma_z|j=s-1/2,j_z\rangle|^2, \nonumber \\ B_{s+1/2}(s_z,j_z)&=&\Gamma_R|\langle s,s_z;1/2,\sigma_z|j=s+1/2,j_z\rangle|^2, \nonumber \\ C_{s-1/2}(j_z,s'_z)&=&0, \nonumber \\ C_{s+1/2}(j_z,s'_z)&=&\Gamma_L|\langle s,s'_z;1/2,\sigma'_z|j=s+1/2,j_z\rangle|^2, \label{c1''} \nonumber \\ D_{s-1/2}(j_z,j_z)&=&-\Gamma_L-\Gamma_R, \nonumber \\ D_{s+1/2}(j_z,j_z)&=&-\Gamma_R. \nonumber\end{aligned}$$ The modified rate matrix can be constructed accordingly. Note that this time $C_{s-1/2}=0$. After the use of identity (\[identity\]), the reduced eigenvalue equation reads, $$\det \Bigg[ \left( -{2s+2\over 2s+1}\Gamma_L-\lambda\right) I +{\Gamma_L\Gamma_R e^{i\eta}\over \Gamma_R+\lambda-\xi} Y_{s,s+1/2}\Bigg]=0. \label{redeigen2} %\nonumber$$ We now attempt to identify the maximal eigenvalue $\lambda(\xi,\eta)$, following the same logic as the case of two spin subsectors in the bias window. As expected, the matrix $Y_{s,j}$ has the following characterizing feature: $$Y_{s,j} \left(\begin{array}{c} 1 \\ 1 \\ \vdots \\ 1 \end{array}\right) ={2j+1\over 2s+1}\left(\begin{array}{c} 1 \\ 1 \\ \vdots \\ 1 \end{array}\right). \label{yseigen}$$ This is explained in the Appendix. Eqs. (\[redeigen1\]),(\[redeigen2\]) and (\[yseigen\]) suggests that $\lambda(\xi,\eta)$ should satisfy a quadratic equation analogous to Eq. (\[quad\]), $${\left({2j+1\over 2s+1}\Gamma_L+\lambda\right)(\Gamma_R+\lambda-\xi) \over \Gamma_L \Gamma_R e^{i\eta}}={2j+1\over 2s+1}, \label{quad1}$$ then $\lambda(\xi,\eta)$ satisfies also Eqs. (\[redeigen1\]) and (\[redeigen2\]). Of course, at this point it is still only “a” solution of the original eigenvalue problem. However, Eq. (\[quad\]) has again the following remarkable property: in the limit of vanishing counting fields, $\xi,\eta\rightarrow 0$, if one takes also the limit $\lambda\rightarrow 0$, then on the left hand side the numerator cancels with the denominator to give simply a factor ${2j+1\over 2s+1}$, which is [*insensitive*]{} to the $\Gamma$’s, and coincides with the right hand side of the equation. This a rather unexpected coincidence because the factor ${2j+1\over 2s+1}$ on the left hand side of the equation comes from a quadratic relation (\[szquad\]), whereas the same ${2j+1\over 2s+1}$ factor on the right hand side originates from a quartic relation (\[quart\]). Indeed, one of the solutions of Eq. (\[quad1\]), $$\begin{aligned} &&\lambda(\xi,\eta)=-{{2j+1\over 2s+1}\Gamma_L+\Gamma_R-\xi\over 2} \nonumber \\ &&+{1\over 2}\sqrt{\left({2j+1\over 2s+1}\Gamma_L-\Gamma_R+\xi\right)^2+ 4{2j+1\over 2s+1}\Gamma_L\Gamma_R e^{i\eta}}, \nonumber \\ \label{cgf1}\end{aligned}$$ satisfies Eq. (\[solstat\]). We are thus able to solve the problem simultaneously for the two cases, i.e., either of the spin subsectors ($j=s\mp 1/2$) in the bias window, corresponding to the cases (ii) and (iii) in Fig. \[fig:3states\]. Contrary to the previous case, i.e., Eq. (\[cgf2\]) for two spin subsectors in the bias window, result which we obtain (\[cgf1\]) depends on $s$. This is because the eigenvalue of Eq. (\[yseigen\]) is dependent on both $s$ and $j$. The generating function $\lambda(\xi,\eta)$ in Eq. (\[cgf1\]) is also asymmetric with respect to the bare amplitudes, $\Gamma_L$ and $\Gamma_R$, whereas this asymmetry tends to disappear for large spin $s$. These characteristic features of the charge-current joint distribution function are further discussed in the next section. Remarks ------- We have thus successfully calculated the FCS for transport through a molecular quantum dot magnet in the incoherent tunneling regime. The results are obtained, in Eqs. (\[cgf2\]) and (\[cgf1\]), together with Eq. (\[cgf\]) in the form of a cumulant generating function (CGF) of charge-current joint correlation functions. Here, we apply the obtained CGF for deducing lowest-order current correlation functions, in order to check consistency with known results. In the case of two spin subsectors in the bias window, i.e., case (i) in Fig. \[fig:3states\], the obtained CGF, i.e., Eq. (\[cgf2\]) has no $s$-dependence. The lowest order cumulants are obtained by taking the derivatives of CGF with respect to counting fields, and have no $s$-dependence. Thus the Fano factor is given for an arbitrary $s$ by, $${C_{02}\over C_{01}}={4\Gamma^2_L+\Gamma^2_R\over (2\Gamma_L+\Gamma_R)^2}, \label{c2}$$ which coincides with the result for spinful conduction electrons and no local spin. [@struben] The skewness normalized by $C_{01}$ reads, $${C_{03}\over C_{01}}= \frac{16\Gamma_L^4-16\Gamma_L^3 \Gamma_R+24\Gamma_L^2 \Gamma_R^2 -4\Gamma_L \Gamma_R^3+\Gamma_R^4} {(2\Gamma_L+\Gamma_R)^4}. \label{c3}$$ Note that Eqs. (\[c2\]) and (\[c3\]) can be rewritten in terms of simple polynomials of a single parameter, $x=(2\Gamma_L-\Gamma_R)^2/(4\Gamma_L\Gamma_R)$ as, $${C_{02}\over C_{01}}={x+1\over x+2},\ \ \ {C_{03}\over C_{01}}={x^2+x+1\over (x+2)^2}.$$ One might expect that the (extended) Fano factors, $C_{04}/C_{01}, C_{05}/C_{01}, \cdots$ could be obtained by developing this apparently systematic series. However, this naive expectation turns out to be an artifact at fourth order. In the case of $j=s\pm1/2$ spin subsector in the bias window, one has to use Eq. (\[cgf2\]), instead of Eq. (\[cgf1\]) as the CGF. Notice that (\[cgf1\]) is obtained by making the following replacement in Eq. (\[cgf2\]): $2\Gamma_L\rightarrow {2j+1\over 2s+1}\Gamma_L$, e.g., the Fano factor becomes in this case, $${C_{02}\over C_{01}}= {\left({2j+1\over 2s+1}\right)^2\Gamma^2_L+\Gamma^2_R \over \left({2j+1\over 2s+1}\Gamma_L+\Gamma_R\right)^2}. \nonumber$$ In the limit of large local spin $s\rightarrow\infty$, this reproduces the result for spinless conduction electrons, $${C_{02}\over C_{01}}\rightarrow {\Gamma^2_L+\Gamma^2_R \over (\Gamma_L+\Gamma_R)^2}.$$ The third order Fano factor (skewness normalized by $C_{01}$) is also obtained by applying the same replacement to Eq. (\[c3\]). Discussions =========== The exact analytic expressions for the CGF, Eqs. (\[cgf2\]) and (\[cgf1\]), contain the full information of the non-equilibrium statistical properties of the current and of the charge. Though the CGF and the joint probability distribution contain the same information as we can see from their definitions (Sec. III), the former is rather a mathematical tool, whereas the latter provides us with a physical picture. We therefore apply an inverse Fourier transformation to the obtained CGF, transforming it back to a joint probability distribution: $$\begin{aligned} P(N,I) = \frac{\tau}{(2 \pi)^2} \! \int_{-\pi}^{\pi} \!\!\!\!\!\! d \eta \int_{-\infty}^{\infty} \!\!\!\!\!\! d \, \xi \; {\rm e}^ { \Omega(\xi,\eta) - i \, \tau N \xi - i \, \tau I \eta } \nonumber.\end{aligned}$$ The asymptotic form for $\tau \! \rightarrow \! \infty$ is obtained within the saddle point approximation, since $\Omega$ is proportional to $\tau$, and thus the exponent is also proportional to $\tau$: $ \ln P(N,I) \approx \Omega(\xi^*,\eta^*) -iN \tau \xi^*-iI \tau \eta^*, $ where $\xi^*$ and $\eta^*$ are determined by the saddle point equations, $N \, \tau \!=\! -i \, \partial_\xi \, \Omega(\xi^*,\eta^*)$ and $I \, \tau \!=\! -i \, \partial_\eta \, \Omega(\xi^*,\eta^*)$ [@bagrets]. Combining these equations, we can produce a contour plot of $\ln P(N,I)$ as a parametric plot in terms of $\eta$ and $\xi$. The probability distribution for only one of the observables, i.e., either charge or current, is obtained further by fixing $\eta^*\!=\!0$ or $\xi^*\!=\!0$, respectively. In this section, we try to visualize the joint probability distribution, and discuss how a molecular spin influences statistical properties of the SET molecular quantum dot. The joint distribution reveals nontrivial correlations: correlations among multiple current components in a multi-terminal chaotic cavity, [@bagrets] or correlations between two different observables, such as current and charge, [@utsumi_condmat] have been investigated. Here, we first consider a dot with symmetric tunnel junctions $\Gamma_L \!=\! \Gamma_R$, and analyze the joint distribution of charge and current under high and small bias voltages. We then consider the case of a large asymmetry in the tunnel coupling, e.g., $\Gamma_L \ll \Gamma_R$. We will argue that much information on the molecular spin and the exchange coupling can be extracted from transport measurements. Finally, we will plot the lowest three cumulants, which are indeed being measured in the state-of-art experiments. Effects of local spin on the statistical properties for symmetric tunnel coupling $\Gamma_L \! = \! \Gamma_R$ {#sec1} ------------------------------------------------------------------------------------------------------------- Let us consider a molecular quantum dot with symmetric tunnel junctions $\Gamma_L=\Gamma_R$. We first consider the case of a large bias voltage, in which the two spin subsectors are both in the bias window, i.e. the case (i) in Fig. \[fig:3states\]. In this case the CGF has been calculated analytically in Eq. (\[cgf2\]). Here, we attempt to uncover basic physical consequences, hidden in this result. In Figs. 2(a) and 2(b), the probability distribution functions of, respectively, charge and current are plotted. They have been evaluated by applying the saddle-point approximation explained above to the solution, Eq. (\[cgf2\]). In the two cases, the distribution function is not symmetric in regard to the peak position, indicating a clear deviation from the Gaussian distribution. Such asymmetric distributions were somewhat expected, since, e.g., in the case of charge distribution, the average value $\langle N \rangle$ is not located at 1/2 but at 2/3. In Fig. 2(c) the two distributions are shown simultaneously as a joint distribution in the form of a contour plot. The distribution possesses a single peak at $I \!=\! \langle I \rangle$ and $N \!=\! \langle N \rangle$. For a given value of current $I$, the peak position of $N$ converges to $N \!=\! \langle N \rangle \!=\! 2/3$ as the current becomes larger ($I \! \gg \! \langle I \rangle$). Note that the probability distributions depend neither on the size $s$ of the molecular spin nor the exchange coupling $g$. The asymmetry in the charge distribution comes from the coefficient 2 in front of $\Gamma_L$ \[Eq. (\[cgf2\])\], reflecting the number of degrees of freedom of a transmitted spin. The truth is that Eq. (\[cgf2\]) is not different from the CGF in the absence of a molecular spin, i.e., the coefficient 2 appears also in the latter case. [@bagrets] We thus conclude that at a high enough bias voltage, all the statistical properties will reproduce the results in the absence of a molecular spin. We then turn our attention to the small bias voltage regime, where only one spin subsector is in the bias window, the cases (ii) or (iii) in Fig. \[fig:3states\]. Contour plots of the joint probability distribution obtained by using Eq. (\[cgf1\]) are plotted in Fig. \[fig:case2\]. Figs. 3(a) and 3(b) correspond to a small molecular spin $s \!=\! 1/2$. In both ferromagnetic ($g\!<\!0$) \[panel (a)\], and antiferromagnetic ($g\!>\!0$) \[panel (b)\] cases, the charge distribution is not symmetric, i.e., typically the peak position is away from $N \!=\! 1/2$. The nature of this asymmetry is similar to the previous case, i.e., the case of large bias voltage, Fig. \[fig:highbias\] (c). The only difference is that here the peak position is located at $\langle N \rangle \!=\! \frac{2j\! + \!1}{2s\!+\!1} \Gamma_L /(\frac{2j\! + \!1}{2s\!+\!1} \Gamma_L \!+\! \Gamma_R)$, which is 3/5 for panel (a) and 1/3 for panel (b). On the other hand, the distribution becomes symmetric w.r.t. a horizontal line $N=1/2$ in the limit of a large spin $s$ \[panel (c)\]. It might be a little surprise that this limit reproduces the joint probability distribution of a spinless fermion [@utsumi_condmat]. In physical terms this can be interpreted as follows: [*a large molecular spin acts as a spin bath for the electron spin transmitted through the quantum dot*]{}. The $n=0$ charge sector is $(2s+1)$-fold degenerate, whereas the degeneracy of $n=1$ charge sector is either $2s$ or $2s+2$. For a large molecular spin $s$, the conduction electron spin does not feel the difference between the two different total angular momentum subsectors. The above result is counterintuitive in the sense that naively we expect that a large molecular spin may act as a large effective Zeeman field for transmitted spins. Asymmetry in the tunnel coupling $\Gamma_L \! \neq \! \Gamma_R$ {#sec2} --------------------------------------------------------------- Here we discuss how an asymmetry in the tunnel coupling may change the statistical properties of a SET molecular quantum dot. In reality, the tunneling coupling to reservoirs is unlikely to be symmetric, because one cannot control the coupling strength of ligands to metallic leads [@heersche]. In the extremely asymmetric limit $\Gamma_R \! \gg \! \Gamma_L$, the CGF of current reduces to a Poissonian form: $$\begin{aligned} \Omega \approx \tau \, z \, \Gamma_L \, ({\rm e}^{i \eta}-1). \label{cgfpoisson}\end{aligned}$$ The factor $z$ is a positive constant, which is smaller than 2: $z \!=\! 2$ for the case (i) and $z \!=\! (2j\!+\!1)/(2s\!+\!1)$ with $j=s\mp 1$, respectively, for the cases (ii) and (iii) in Fig. \[fig:3states\]. When we apply a negative bias voltage, the CGFs are obtained from Eqs. (\[cgf2\]) and (\[cgf1\]) by replacing $\Gamma_L \! \leftrightarrow \! \Gamma_R$ and $\eta \! \rightarrow \! -\eta$ as, $\bar{\Omega} \! \approx \! \tau \, \Gamma_L \, ({\rm e}^{-i \eta}-1)$. The absolute value of the $n$-th cumulant is $|C_{0n}| \! \approx \! \tau z \Gamma_L$ for a positive bias voltage, whereas it becomes $|\bar{C}_{0n}| \! \approx \! \tau \Gamma_L$ for a negative bias voltage. Therefore, from the ratio $|C_{0n}/\bar{C}_{0n}|$, we can estimate the factor $z$, and consequently, obtain some information on the molecular spin and the exchange coupling. Such a method was previously proposed [@Akera] for a multi-orbital quantum dot. Figs. 4(a) and 4(b) are the joint probability distribution for $\Gamma_L/\Gamma_R \!=\! 0.1$. A strong asymmetry around $N \!=\! \langle N \rangle \!=\! z \Gamma_L/(z \Gamma_L \!+\! \Gamma_R)$ is observed both for the antiferromagnetic (a) and the ferromagnetic (b) cases. A longer tail in the panel (b) implies that the correlations among tunneling processes are weaker for antiferromagnetic coupling. The measure of such correlations, the Fano factor $C_{02}/C_{01}$, as a function of the asymmetry ratio $\Gamma_L/\Gamma_R$ is shown in Fig. \[fig:asymmetry\] (c). We can observe that even for a small asymmetry ratio $\Gamma_L/\Gamma_R \!=\! 0.1$, the Fano factor is still smaller than 1, and a fair amount of difference remains between the two cases. For the charge distribution, around $I \! \approx \! 0$, we observe equidistant contours in panels (a) and (b), which implies that the charge is exponentially distributed. Actually the charge distribution for both cases roughly follows $\ln P(N) \! \propto \! - \tau \Gamma_R N$ \[Figure \[fig:asymmetry\] (d)\]. A similar exponential distribution appears as an equilibrium distribution of charge when the dot level is far away from either of the two lead chemical potentials [@utsumi_condmat]. It is shown that the exponent depends simply on the decay rate of the excited state. In the present case, since the stationary state is fairly approximated by the empty state, this reduces to the outgoing tunneling rate $\Gamma_R$ of an electron in the occupied state to the right reservoir. Relation to experiments ----------------------- Up to now, we have shown contour plots of the probability distribution of charge and current. For a semiconducting quantum dot, it has become possible to measure the statistical probability distribution of current in the incoherent tunneling regime [@fujisawa; @gustavsson_ensslin]. Such measurements are based on the real-time counting of single electron jumps using an on-chip quantum point contact charge detectors. At the present stage, it may be rather challenging to apply the same technique for a single-molecular device with the same precision, but it might be still possible to measure the lowest three cumulants, since the skewness for a tunnel junction can be measured by the state-of-art experiments. [@reulet; @reznikov] We have shown explicitly the lowest order cumulants in Sec. IV C. For a large bias voltage, they are given, e.g., in Eqs. (\[c2\]) and (\[c3\]). For a small bias voltage, the (generalized) Fano factors were obtained by making the replacement, $2\Gamma_L\rightarrow z\Gamma_L={2j+1\over 2s+1}\Gamma_L$, in Eqs. (\[c2\]) and (\[c3\]). Fig. 5 shows the bias voltage dependence of the lowest three cumulants for a large molecular spin $s \!=\! 10$, which is approximately the size of the magnetic moment of ${\rm Mn_{12}}$. [@heersche] In this case, as the factor $z={2j+1\over 2s+1}=20/21$ (for $j=s-1/2$) is close to 1, the molecular spin is expected to act as a [*spin bath*]{}, reproducing characteristic features of the transport of spinless fermions. It is assumed that the tunnel coupling is asymmetric, $\Gamma_R \!=\! 5 \Gamma_L$. We also focus on the antiferromagnetic ($g>0$) case, in which the “triplet” level ($j=s+1/2$), $\epsilon_{s+1/2}=\epsilon_{\rm dot}+ g s/2$ is above the “singlet” level ($j=s-1/2$), $\epsilon_{s-1/2}=\epsilon_{\rm dot}- g (s+1)/2$. In equilibrium, the singlet subsector is tuned to be at the lead chemical potential level, $\mu_L=\epsilon_{s-1/2}=\mu_R=0$. We then apply either a positive \[panels (a-1) and (b-1)\] or a negative \[panel (a-2) and (b-2)\] bias voltage. The singlet level always stays in the bias window, i.e., either $\mu_L>\epsilon_{s-1/2}>\mu_R$ \[panels (a-1) and (b-1)\] or $\mu_L<\epsilon_{s-1/2}<\mu_R$ \[panels (a-2) and (b-2)\] is realized. Note that in Fig. 5 the origin of energy is taken, for simplicity, to be at the singlet level, i.e., $\epsilon_{s-1/2}=0$ is always satisfied. The expected cumulant-voltage characteristics for a positive and for a negative bias voltage are depicted in panels (a-1) and (a-2) of Fig. 5. At a small bias voltage, $|\mu_{L,R}| \! < \! \epsilon_{s+1/2} \!-\! \epsilon_{s-1/2} \! =\! g (s \!+\! 1/2)$, the absolute value of various cumulants are insensitive to the sign of bias voltage \[This case realizes the situation of Fig. \[fig:highbias\] (c)\]. On the other hand, when the bias voltage is large, $|\mu_{L,R}| \! > \! g (s \!+\! 1/2)$, each cumulant takes different values depending on whether the bias is either positive $|\mu_L| \! > \! g (s \!+\! 1/2)$, or negative $|\mu_R| \! > \! g (s \!+\! 1/2)$ \[This case realizes the situation of Fig. \[fig:case2\] (c)\]. Such a feature could be a smoking gun of the transport of spinless fermions. Panels (b-1) and (b-2) show Fano factors corresponding, respectively, to the panels (a-1) and (a-2). At a large bias voltage $|\mu_{L,R}| \! > \! g (s \!+\! 1/2)$, the suppression of Fano factors is stronger for a positive \[panel (b-1)\] than for a negative \[panel (b-2)\] bias voltage. This can be understood as follows. If one compares the large bias regime of panels (a-1) and (a-2), one notices that the absolute value of current, i.e., $C_{01}$ is smaller in this regime in panel (a-2). This implies that tunneling events are less correlated when the bias is large and negative, leading to smaller Fano factors in panel (b-2). This is rather a trivial fact, but subjective to a direct experimental check. Finally, it should be noted again that the magnetic anisotropy of the molecule was neglected throughout the present paper. The extension of our theory to account for a large anisotropy comparable to the exchange coupling is beyond our scope here. Conclusions =========== We have thus studied the full counting statistics (FCS) for a molecular quantum dot magnet, a quantum dot with an intrinsic molecular spin ($s$) degrees of freedom, which is coupled to the conduction electron spin ($\sigma$) via an exchange interaction ($s\cdot\sigma$). Such a molecular quantum dot magnet is now experimentally available, and carries generally a large molecular spin $s\sim 10$. [@heersche] We applied the master equation approach to FCS [@bagrets] to our molecular quantum dot problem, by considering the incoherent tunneling regime. In this approach the cumulant generating function (CGF) for FCS is obtained by solving an eigenvalue problem associated with a modified master equation with counting fields. We also assumed that the molecular quantum dot was in the strong Coulomb blockade regime ($U\rightarrow\infty$) so that the dot state has only two charge sectors $n=0,1$. A standard algebraic identity, Eq. (\[identity\]), allowed us to reduce the size of the eigenvalue problem from $3(2s+1)\times3(2s+1)$ down to $(2s+1)\times (2s+1)$ by projecting the relevant physical space onto the charge $(0,0)$-sector. This was, of course, simply a rewriting of the original problem. In the new representation, Clebsch-Gordan coefficients appear at fourth order. We established a few identities among such fourth order coefficients, which allowed us to identify one solution of the eigenvalue problem, which turned out to be the correct eigenvalue, representing the CGF for FCS. Thus we obtained an analytic expression for the CGF [*as a function of the local spin $s$*]{} for different configurations of the dot states and of the two leads in regard to their relative energy levels (Fig. \[fig:3states\]). Based on the obtained analytic expressions, we also developed numerical analysis, e.g., contour plots of the charge-current joint distribution function. For a small local spin, e.g., $s=1/2$, the obtained contour plots show clearly an asymmetry in the distribution of the charge, i.e., between $n=0$ and $n=1$. We then demonstrated that this asymmetry grows differently for different cases in Fig. \[fig:3states\] mentioned above. In particular, this asymmetry tends to disappear as the local spin $s$ becomes large, in cases (ii) and (iii) of Fig. \[fig:3states\], indicating that the local spin plays the role of a [*spin bath*]{} for the conduction electron. Of course, as emphasized in Sec. 5, it is rather counterintuitive to recover such characteristic transport properties of a [*spinless fermion*]{} in this large $s$ limit. We further demonstrated that such characteristic features of the transport of a spinless fermion, which we believe to have a chance to be realized in reality, e.g., in a Mn$_{12}$ molecular quantum dot, are subjective to a direct experimental check through the bias voltage dependence of lowest-order cumulants. It is interesting to extend our analysis to the case of leads of different nature, such as ferromagnetic leads or superconducting leads. The range of validity of this work is limited to the incoherent tunneling regime, whereas one might naturally wonder what happens when the coupling between the dot and the leads become stronger. Kondo type physics in such a regime is particularly interesting, [@romeike_kondo] because of the coexistence of an intrinsic magnetic impurity (molecular spin) and the quantum dot, the latter known to show enhancement of conductance due to Kondo mechanism at low temperatures. Certainly one has to go beyond the master equation approach to FCS in order to treat FCS in such a regime. We are grateful to JSPS Invitation Fellowship for Research in Japan. K.I. and Y.U. are supported by RIKEN as Special Postdoctoral Researcher. T.M. also acknowledges support from French CNRS A.C. Nanosciences grant NR0114. He thanks the members of Condensed Matter Theory Laboratory at RIKEN for their hospitality. [**Appendices**]{} Useful relations among Clebsch-Gordan coefficients ================================================== For the calculation of FCS for a model with an arbitrary spin $s$ we have used extensively the Clebsch-Gordan algebra. We, therefore, summarize in the following useful relations among Clebsch-Gordan coefficients. We start with conventional quadratic relations which are sometimes discussed in standard quantum mechanics textbooks, and then proceed to unconventional quartic relations which we discovered as a byproduct of this research. Quadratic relations ------------------- Let us summarize the quadratic identities on the Clebsch-Gordan coefficients, $\langle s,s_z;1/2,\sigma_z|j,j_z\rangle$, for a given $s$ and $\sigma=1/2$. Angular momentum selection rule allows only two values of $j=s\pm 1/2$. Because of the constraint $j_z=s_z+\sigma_z$, [*only two of the three parameters*]{}, $s_z$, $\sigma_z$ and $j_z$ are independent. Lets us first look at Table I. The two rows correspond to different total angular momentum $j$. Either of the two columns specified by $\sigma_z$ correspond also to the same $j_z=s_z+\sigma_z$. Then summing up the two coefficients squared given in either of the two rows, one finds, $$\begin{aligned} \sum_{\sigma_z=\pm 1/2}|\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2&=& \nonumber \\ \sum_{j_z=-j,\cdots,j}|\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 &=&{2j+1\over 2s+1}. \label{szquad}\end{aligned}$$ The identity (\[szquad\]) applies to [*a given set of $s$, $j$ and $s_z$*]{}. If we release the quantum number $j$ in Eq. (\[szquad\]), and take the summation over two spin sectors, one finds, $$\begin{aligned} \sum_{j=s\pm 1/2}\sum_{j_z=-j,\cdots,j} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 \nonumber \\ ={2s\over 2s+1}+{2s+2\over 2s+1}=2. \label{CTquad}\end{aligned}$$ which is independent of the spin $s$. The identities (\[szquad\]),(\[CTquad\]) are used to obtain the expressions for $A(s_z,s_z)$. $\sigma_z=-1/2$ $\sigma_z=1/2$ ----------- ---------------------------- ---------------------------- $j=s-1/2$ $\sqrt{s+s_z\over 2s+1}$ $\sqrt{s-s_z\over 2s+1}$ $j=s+1/2$ $\sqrt{s-s_z+1\over 2s+1}$ $\sqrt{s+s_z+1\over 2s+1}$ : Clebsch-Gordan coefficients, $\langle s,s_z;1/2,\sigma_z|j,j_z=s_z+\sigma_z\rangle$ for [*given $s$ and $s_z$*]{}. Note the difference between Tables I and II. They are equivalent but represented in different ways for different use. The four elements on the Table correspond one by one, i.e., they are same in number in different Tables, but written in terms of different parameters (either $j_z$ or $s_z$). $\sigma_z=-1/2$ $\sigma_z=1/2$ ----------- ------------------------------ ------------------------------ $j=s-1/2$ $\sqrt{s+j_z+1/2\over 2s+1}$ $\sqrt{s-j_z+1/2\over 2s+1}$ $j=s+1/2$ $\sqrt{s-j_z+1/2\over 2s+1}$ $\sqrt{s+j_z+1/2\over 2s+1}$ : Clebsch-Gordan coefficients, $\langle s,s_z=j_z-\sigma_z;1/2,\sigma_z|j,j_z\rangle$ for [*given $s$ and $j_z$*]{}. Note that for $s$ and $j_z$ given, only four elements shown in the Table remain finite. On the other hand, keeping $s$ and $j$ fixed, one can also consider another situation where [*$j_z$ is fixed*]{} (instead of $s_z$). This corresponds to Table II. If one sums up the two coefficients squared given in either of the two rows in Table II (as we did for Table I), one finds this time, $$\begin{aligned} \sum_{\sigma=\pm 1/2} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2&=& \nonumber \\ \sum_{s_z=-s,\cdots,s} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2&=&1. \label{jzquad}\end{aligned}$$ If we focus on a particular term in the above summation, for which $s_z$ is given, then $\sigma_z$ is determined automatically by the constraint $j_z=s_z+\sigma_z$. Therefore, the summation over $\sigma_z$ and $s_z$ can be replaced by one another. The identity (\[jzquad\]) is relevant for finding the diagonal elements of the matrix $D$ defined by Eq. (\[d\]). It turns out that $D_j(j_z,j_z)$ does not depend on $j_z$. The identity (\[jzquad\]) applies to [*a given set of $s$, $j$ and $j_z$*]{}. The two identities (\[szquad\],\[jzquad\]) are indeed consistent, $$\begin{aligned} \sum_{s_z=-s,\cdots,s}\sum_{j_z=-j,\cdots,j} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2= \nonumber \\ \sum_{s_z=-s,\cdots,s} {2j+1\over 2s+1}=2j+1, \nonumber \\ \sum_{j_z=-j,\cdots,j}\sum_{s_z=-s,\cdots,s} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2= \nonumber \\ \sum_{j_z=-j,\cdots,j} 1 =2j+1. \label{sumquad}\end{aligned}$$ Furthermore, releasing again the quantum number $j$ and taking the summation over two spin sectors, one can verify, $$\begin{aligned} \sum_{s_z=-s,\cdots,s}\sum_{j=s\pm 1/2}\sum_{j_z=-j,\cdots,j} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 \nonumber \\ =\sum_{s_z=-s,\cdots,s}2=2(2s+1), \nonumber \\ \sum_{j=s\pm 1/2}\sum_{j_z=-j,\cdots,j}\sum_{s_z=-s,\cdots,s} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 \nonumber \\ =\sum_{j=s\pm 1/2}(2j+1)=2(2s+1). \nonumber\end{aligned}$$ We used Eq. (\[CTquad\]) for obtaining the second expression, whereas for the third one, we used Eq. (\[sumquad\]), but, of course, whichever path one chooses, one always obtain at the end, the same $2s+1$. The quartic relation -------------------- As a by-product of the calculation of FCS, we discovered some unconventional quartic relations among Clebsch-Gordan coefficients. In the evaluation of a determinant derived from the modified master equation (\[masmod\]) we encountered a $(2s+1)\times (2s+1)$ square matrix, consisting of forth-order Clebsch-Gordan coefficients, defined as Eq. (\[xs\]), whose $(s_z,s'_z)$-elements one can rewrite here as, $$\begin{aligned} X_s (s_z,s'_z)= \sum_{j=s\pm 1/2}\sum_{j_z=-j,\cdots,j} \nonumber \\ |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 |\langle s,s'_z;1/2,\sigma'_z|j,j_z\rangle|^2. \nonumber\end{aligned}$$ $X_s$ is a [*tridiagonal*]{} matrix, i.e., only $(s_z,s_z)$, $(s_z,s_z+1)$ and $(s_z,s_z-1)$ elements are non-zero. They are given explicitly as, $$\begin{aligned} X_s(s_z,s_z)&=& {2\{s^2+(s+1)^2+2s_z^2\}\over (2s+1)^2}, \nonumber \\ X_s(s_z,s_z+1)&=& {2(s-s_z)(s+s_z+1)\over (2s+1)^2}, \nonumber \\ X_s(s_z,s_z-1)&=& {2(s+s_z)(s-s_z+1)\over (2s+1)^2}. \label{xstri}\end{aligned}$$ This matrix $X_s$ has a very characteristic property, Eq. (\[xseigen\]), i.e., $$X_s \left(\begin{array}{c} 1 \\ 1 \\ \vdots \\ 1 \end{array}\right) =2 \left(\begin{array}{c} 1 \\ 1 \\ \vdots \\ 1 \end{array}\right), \label{xseigen'} \nonumber$$ This can be checked explicitly for $s=1/2, 1, 3/2,\cdots$. Indeed, the matrix $X_s$ reads, $$\begin{aligned} X_{1/2}&=&\left(\begin{array}{cc} 3/2 &1/2 \\ 1/2 &3/2 \end{array}\right), X_{1}=\left(\begin{array}{ccc} 14/9 &4/9 &0 \\ 4/9 &10/9 &4/9 \\ 0 &4/9 &14/9 \end{array}\right), \nonumber \\ X_{3/2}&=&\left(\begin{array}{cccc} 13/8 &3/8 &0 &0 \\ 3/8 &9/8 &1/2 &0 \\ 0 &1/2 &9/8 &3/8 \\ 0 &0 &3/8 &13/8 \end{array}\right), \cdots.\end{aligned}$$ Proving the general result (\[xseigen’\]) for arbitrary $s$ is equivalent to establishing the following identity: $$\begin{aligned} \sum_{j=s\pm 1/2}\sum_{s'_z=-s,\cdots,s}\sum_{j_z=-j,\cdots,j} \nonumber \\ |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 |\langle s,s'_z;1/2,\sigma'_z|j,j_z\rangle|^2=2. \label{sumquart}\end{aligned}$$ When $s_z\neq \pm s$, one can verify this by substituting into it explicit matrix elements given in Eq. (\[xstri\]), i.e., $$X_s(s_z,s_z-1)+X_s(s_z,s_z)+X_s(s_z,s_z+1)=2. \label{xssum}$$ Then one can further verify that Eq. (\[xssum\]) still holds when the [*edges*]{} of the $X_s$ matrix are encountered, at $s_z=\pm s$. The same kind of structure as Eqs. (\[xseigen’\]) and (\[sumquart\]) exists actually for each spin sector $j$. $Y_{s,j}$ introduced in Eq. (\[ys\]) is such a matrix, whose $(s_z,s'_z)$-elements one can rewrite here as, $$\begin{aligned} Y_{s,j} (s_z,s'_z)= \sum_{j_z=-j,\cdots,j} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 \nonumber \\ \times |\langle s,s'_z;1/2,\sigma'_z|j,j_z\rangle|^2. \nonumber\end{aligned}$$ $Y_{s,j}$ is related $X_s$ as, $X_s=\sum_{j=s\pm 1/2}Y_{s,j}=Y_{s,s-1/2}+Y_{s,s+1/2}$. Again $Y_{s,j}$ is a [*tridiagonal*]{} matrix, with finite matrix elements only at $(s_z,s_z)$, $(s_z,s_z+1)$ and $(s_z,s_z-1)$. As expected, the matrix $Y_{s,j}$ has the following characterizing feature, Eq. (\[yseigen\]), i.e., $$Y_{s,j} \left(\begin{array}{c} 1 \\ 1 \\ \vdots \\ 1 \end{array}\right) ={2j+1\over 2s+1}\left(\begin{array}{c} 1 \\ 1 \\ \vdots \\ 1 \end{array}\right), \label{yseigen'}$$ which is also equivalent to the following identity: $$\begin{aligned} \sum_{s'_z=-s,\cdots,s}\sum_{j_z=-j,\cdots,j} |\langle s,s_z;1/2,\sigma_z|j,j_z\rangle|^2 \nonumber \\ \times |\langle s,s'_z;1/2,\sigma'_z|j,j_z\rangle|^2= {2j+1\over 2s+1}. \label{quart}\end{aligned}$$ Now in order to prove Eq. (\[yseigen’\]) or (\[quart\]) for general $s$ let us consider the following two cases separately: 1. [$j=s-1/2$ spin sector.]{} One can first directly verify Eq. (\[yseigen\]) by writing down explicitly the matrices $Y_{s,s-1/2}$ for $s=1/2, 1, 3/2, \cdots$: $$\begin{aligned} Y_{1/2,0}&=&\left(\begin{array}{cc} 1/4 &1/4 \\ 1/4 &1/4 \end{array}\right), \nonumber \\ Y_{1,1/2}&=&\left(\begin{array}{ccc} 4/9 &2/9 &0 \\ 2/9 &2/9 &2/9 \\ 0 &2/9 &4/9 \end{array}\right), \nonumber \\ Y_{3/2,1}&=&\left(\begin{array}{cccc} 9/16 &3/16 &0 &0 \\ 3/16 &5/16 &1/4 &0 \\ 0 &1/4 &5/16 &3/16 \\ 0 &0 &3/16 &9/16 \end{array}\right),\cdots. \nonumber\end{aligned}$$ The finite tridiagonal matrix elements are in this case, $$\begin{aligned} Y_{s,s-1/2}(s_z,s_z)&=& {2(s^2+s_z^2)\over (2s+1)^2}, \nonumber \\ Y_{s,s-1/2}(s_z,s_z+1)&=& {(s-s_z)(s+s_z+1)\over (2s+1)^2}, \nonumber \\ Y_{s,s-1/2}(s_z,s_z-1)&=& {(s+s_z)(s-s_z+1)\over (2s+1)^2}. \label{ys1}\end{aligned}$$ 2. [$j=s+1/2$ spin sector.]{} One verifies Eq. (\[yseigen\]) by writing down explicitly the matrices $Y_{s,s+1/2}$ for $s=1/2, 1, 3/2, \cdots$: $$\begin{aligned} Y_{1/2,1}&=&\left(\begin{array}{cc} 5/4 &1/4 \\ 1/4 &5/4 \end{array}\right), \nonumber \\ Y_{1,3/2}&=&\left(\begin{array}{ccc} 10/9 &2/9 &0 \\ 2/9 &8/9 &2/9 \\ 0 &2/9 &10/9 \end{array}\right), \nonumber \\ Y_{3/2,2}&=&\left(\begin{array}{cccc} 17/16 &3/16 &0 &0 \\ 3/16 &13/16 &1/4 &0 \\ 0 &1/4 &13/16 &3/16 \\ 0 &0 &3/16 &17/16 \end{array}\right),\cdots. \nonumber\end{aligned}$$ Notice also that $X_s=Y_{s,s-1/2}+Y_{s,s+1/2}$. The finite tridiagonal matrix elements are, $$\begin{aligned} Y_{s,s+1/2}(s_z,s_z)&=& {2\left\{(s+1)^2+s_z^2\right\}\over (2s+1)^2}, \nonumber \\ Y_{s,s+1/2}(s_z,s_z+1)&=& {(s-s_z)(s+s_z+1)\over (2s+1)^2}, \nonumber \\ Y_{s,s+1/2}(s_z,s_z-1)&=& {(s+s_z)(s-s_z+1)\over (2s+1)^2}. \label{ys2}\end{aligned}$$ When $s_z\neq \pm s$, one can verify Eq. (\[yseigen’\]), using either Eq. (\[ys1\]) or (\[ys2\]), i.e., one finds for a given $s_z$, $$\begin{aligned} Y_{s,j}(s_z,s_z-1)+Y_{s,j}(s_z,s_z) \nonumber \\ +Y_{s,j}(s_z,s_z+1)={2j+1\over 2s+1}. \label{yssum}\end{aligned}$$ Then one can further verify that Eq. (\[yssum\]) still holds when the [*edges*]{} of the $X_s$ matrix are encountered, at $s_z=\pm s$. This establishes Eq. (\[yseigen’\]), or equivalently, (\[quart\]). Note that the quartic summation, Eq. (\[quart\]), or the eigenvalue of Eq. (\[yseigen’\]) [*happens to be*]{} identical to a quadratic summation, (\[szquad\]). This coincidence is a crucial ingredient for the FCS solvability of our model. If one takes the summation over two spin sectors $j=s\pm 1/2$, Eq. (\[quart\]) recovers Eq. (\[sumquart\]). [99]{} A. Nitzan and M.A. Ratner, Science [**300**]{}, 1384 (2003). G. Cuniberti, G. Fagas, and K. Richter (eds.), [*Introducing Molecular Electronics*]{}, Lect. Notes Phys. 680 (Springer 2005). M. Tsukada, K. Tagami, K. Hirose, and N. Kobayashi, J. Phys. Soc. Jpn. [**74**]{}, 1079 (2005). A. Mitra, I. Aleiner and A. J. Millis, Phys. Rev. [**B 69**]{}, 245302 (2004). Y. Utsumi, J. Martinek, G. Schön, H. Imamura, S. Maekawa, Phys. Rev. [**B 71**]{}, 245116 (2005). H. B. Heersche, Z. de Groot, J. A. Folk, H. S. J. van der Zant, C. Romeike, M. R. Wegewijs, L. Zobbi, D. Barreca, E. Tondello, and A. Cornia Phys. Rev. Lett. [**96**]{}, 206801 (2006). C. Romeike, M. R. Wegewijs, and H. Schoeller Phys. Rev. Lett. [**96**]{}, 196805 (2006). F. Elste, C. Timm, Phys. Rev. [**B 71**]{}, 155403 (2005). Yu.V. Nazarov (ed.), [*Quantum noise in Mesoscopic Phyiscs*]{}, NATO Scinece Series II: Mathematics, Physics and Chemistry, Vol. 97, Klüwer Academic Publishers, Dordrecht, 2002. R. Fazio, V.F. Gantmakher and Y. Imry (eds.), [*New Directions in Mesoscopic Physics (Towards Nanoscience)*]{}, NATO Scinece Series II: Mathematics, Physics and Chemistry, Vol. 125, Klüwer Academic Publishers, Dordrecht, 2003. L.S. Levitov and G.B. Lesovik, JETP Lett. [**58**]{}, 230 (1993); L.S. Levitov, H.-W. Lee, and G.B. Lesovik, J. Math. Phys. [**37**]{}, 4845 (1996); L.S. Levitov, in Ref. [@fazio], p.p. 67-91. D. A. Bagrets, Yu. V. Nazarov, Phys. Rev. [**B 67**]{}, 085316 (2003); ibid., in Ref. [@nazarov], p.p. 429-462. Y. Utsumi, Phys. Rev. B, in press (cond-mat/0604082). B. Reulet, J. Senzier, and D.E. Prober, Phys. Rev. Lett. [**91**]{}, 196601 (2003). Yu. Bomze, G. Gershon, D. Shovkun, L. S. Levitov, M. Reznikov, Phys. Rev. Lett. [**95**]{}, 176601 (2005). M. Kindermann, Yu. V. Nazarov, C. W. J. Beenakker, Phys. Rev. [**B 69**]{}, 035336 (2004). S. Gustavsson, R. Leturcq, B. Simovic, R. Schleser, T. Ihn, P. Studerus, K. Ensslin, D. C. Driscoll, and A. C. Gossard, Phys. Rev. Lett. [**96**]{}, 076605 (2006); S. Gustavsson, R. Leturcq, B. Simovic, R. Schleser, P. Studerus, T. Ihn, K. Ensslin, D. C. Driscoll, and A. C. Gossard, Phys. Rev. B 74, 195305 (2006); S. Gustavsson, R. Leturcq, T. Ihn,, K. Ensslin, M. Reinwald, and W. Wegscheider, cond-mat/0607192. T. Fujisawa, T. Hayashi, R. Tomita, Y. Hirayama, Science [**312**]{}, 1634 (2006) D. Boese and H. Schoeller, Europhys. Lett. [**54**]{}, 668 (2001). N.G. van Kampen, [*Stochastic Processes in Physics and Chemistry*]{}, Elsevier Science Publishers, 1992. Yu.V. Nazarov, J.J.R. Struben, [**B 53**]{}, 15466 (1996). H. Akera, Phys. Rev. [**B 60**]{}, 10683 (1999). C. Romeike, M. R. Wegewijs, W. Hofstetter, and H. Schoeller Phys. Rev. Lett. [**96**]{}, 196601 (2006)
{ "pile_set_name": "ArXiv" }
--- abstract: 'The semiclassical mechanics of the Wigner $6j$-symbol is examined from the standpoint of WKB theory for multidimensional, integrable systems, to explore the geometrical issues surrounding the Ponzano-Regge formula. The relations among the methods of Roberts and others for deriving the Ponzano-Regge formula are discussed, and a new approach, based on the recoupling of four angular momenta, is presented. A generalization of the Yutsis-type of spin network is developed for this purpose. Special attention is devoted to symplectic reduction, the reduced phase space of the $6j$-symbol (the 2-sphere of Kapovich and Millson), and the reduction of Poisson bracket expressions for semiclassical amplitudes. General principles for the semiclassical study of arbitrary spin networks are laid down; some of these were used in our recent derivation of the asymptotic formula for the Wigner $9j$-symbol.' address: - 'Dipartimento di Chimica, Università di Perugia, Perugia, Italy 06100' - 'Department of Physics, University of California, Berkeley, California 94720 USA' author: - Vincenzo Aquilanti - 'Hal M. Haggard, Austin Hedeman, Nadir Jeevanjee, Robert G. Littlejohn, and Liang Yu' title: 'Semiclassical Mechanics of the Wigner $6j$-Symbol' --- Introduction ============ The Wigner $6j$-symbol (or Racah $W$-coefficient) is a central object in angular momentum theory, with many applications in atomic, molecular and nuclear physics. These usually involve the recoupling of three angular momenta, that is, the $6j$-symbol contains the unitary matrix elements of the transformation connecting the two bases that arise when three angular momenta are added in two different ways. Such applications and the definition of the $6j$-symbol based on them are described by Edmonds (1960). More recently the $6j$- and other $3nj$-symbols have found applications in quantum computing (Marzuoli and Rasetti 2005) and in algorithms for molecular scattering calculations (De Fazio  2003, Anderson and Aquilanti 2006), which make use of their connection with discrete orthogonal polynomials (Aquilanti  1995, 2001a,b). The $6j$-symbol is an example of a spin network, a graphical representation for contractions between tensors that occur in angular momentum theory. The graphical notation has been developed by Yutsis (1962), El Baz and Castel (1972), Lindgren and Morrison (1986), Varshalovich (1981), Stedman (1990), Danos and Fano (1998), Wormer and Paldus (2006), Balcar and Lovesey (2009) and others. The $6j$-symbol is the simplest, nontrivial, closed spin network (one that represents a rotational invariant). Spin networks are important in lattice QCD and in loop quantum gravity where they provide a gauge-invariant basis for the field. Applications in quantum gravity are described by Rovelli and Smolin (1995), Baez (1996), Carlip (1998), Barrett and Crane (1998), Regge and Williams (2000), Rovelli (2004) and Thiemann (2007), among others. Alongside the Yutsis school of graphical notation and the Clebsch-Gordan school of algebraic manipulation there is a third approach to the evaluation of rotational ($SU(2)$) invariants. The third method, sometimes called chromatic evaluation, grew out of Penrose’s doctoral work on the graphical representation of tensors and is closely related to knot theory. We will not have further occasion to mention this school, see Penrose (1971) for its introduction, Rovelli (2004) for an overview and Kauffman and Lins (1994) for its full development. The asymptotics of spin networks and especially the $6j$-symbol has played an important role in many areas. By “asymptotics” we refer to the asymptotic expansion for the spin network when all $j$’s are large, equivalent to a semiclassical approximation since large $j$ is equivalent to small $\hbar$. The asymptotic expression for the $6j$-symbol (the leading term in the asymptotic series) was first obtained by Ponzano and Regge (1968), or, more precisely, they obtained several formulas, valid inside and outside the classically allowed region and in the neighborhood of the caustics. In the same paper those authors gave the first spin foam model (a discretized path integral) for quantum gravity. The formula of Ponzano and Regge is notable for its high symmetry and the manner in which it is related to the geometry of a tetrahedron in three-dimensional space. It is also remarkable because the phase of the asymptotic expression is identical to the Einstein-Hilbert action for three-dimensional gravity integrated over a tetrahedron, in Regge’s (1961) simplicial approximation to general relativity. The semiclassical limit of the $6j$-symbol thus plays a crucial role in simplicial approaches to the quantization of the gravitational field. For all these reasons, the asymptotic formula of Ponzano and Regge for the $6j$-symbol has attracted a great deal of attention. Ponzano and Regge obtained their formula by inspired guesswork, supporting their conclusion both with numerical evidence and arguments of consistency and plausibility. The formula itself is of the one-dimensional WKB-type, a reflection of the fact that the $6j$-symbol fundamentally represents a dynamical system of one degree of freedom. The Ponzano-Regge formula was first derived by Neville (1971), using the recursion relations satisfied by the $6j$-symbol and a discrete version of WKB theory. Similar techniques were later used by Schulten and Gordon (1975a,b), who also presented stable algorithms for evaluating the $6j$-symbol numerically. A proof of a different sort was later given by Biedenharn and Louck (1981), based on showing that the Ponzano-Regge formula satisfies a set of defining properties of the $6j$-symbol. More recently there have appeared more geometrical treatments of the asymptotics of the $6j$-symbol, that is, those based on geometric quantization (Kirillov 1976, Guillemin and Sternberg 1977, Woodhouse 1991), symplectic geometry and symplectic and Poisson reduction (Abraham and Marsden 1978, Arnold 1989, Marsden and Ratiu 1999) and other techniques. Among these are the works by Roberts (1999) and by Charles (2008). In addition, the $6j$-symbol has been taken as a test case for asymptotic studies of amplitudes that occur in quantum gravity (Barrett and Steele 2003, Freidel and Louapre 2003), in which the authors developed integral representations for the $6j$-symbol as integrals over products of the group manifold. There have also been quite a few other studies of asymptotics of particular spin networks, including Barrett and Williams (1999), Baez (2002), Rovelli and Speziale (2006), Hackett and Speziale (2007), Conrady and Freidel (2008), Alesci (2008), Barrett (2009), among others. We also mention the works of Gurau (2008), which applies standard asymptotic techniques (Stirling’s approximation, etc) directly to Racah’s sum for the $6j$-symbol; of Ragni (2010) on the computation of $6j$-symbols and on the asymptotics of the $6j$-symbol when some quantum numbers are large and others small; and of Littlejohn and Yu (2009) on uniform approximations for the $6j$-symbol. In addition there has been some work on the $q$-deformed $6j$-symbol, important for the regularization of the Ponzano-Regge spin-foam model (Turaev and Viro 1992, Ooguri 1992a,b) and for its possible connection to quantum gravity with cosmological constant. In particular, Taylor and Woodward (2004) applied the recursion and WKB method of Schulten and Gordon to the $q$-deformed $6j$-symbol. The results are geometrically interesting (the tetrahedron of Ponzano and Regge is moved from $\Reals^3$ to $S^3$ when the $q$-deformation is turned on), but it seems that at present there is no geometrical treatment of the asymptotics of the $q$-deformed $6j$-symbol, analogous to what is available for the ordinary $6j$-symbol. There is also the recent work of Van der Veen (2010) on the asymptotics of general $q$-deformed spin networks, which treats the problem from the standpoint of knot theory and representation theory. Among other things, this work creates a broad generalization of the Schwinger-Bargmann generating function of the $6j$-symbol. In Aquilanti (2007) we applied multidimensional WKB theory for integrable systems to the asymptotics of the $3j$-symbol, and in this paper we apply similar techniques to the $6j$-symbol. These methods bear the closest relationship to the works of Roberts (1999) and of Charles (2008). The point of this paper is not another derivation of the Ponzano-Regge formula, although one is provided, but rather to clarify the relationship among some of the methods used in the past, to reveal useful calculational techniques, and to lay the basis for the development of new results. Among the latter we mention our own work on uniform approximations for the $6j$-symbol (Littlejohn and Yu, 2009), our recent derivation of the asymptotic form for the $9j$-symbol (Haggard and Littlejohn, 2010), both of which relied on techniques explained in this paper, and our work on the Bohr-Sommerfeld quantization of the volume operator in loop quantum gravity (Bianchi and Haggard, 2011). Previous and current work on the volume operator includes Chakrabarti (1964), Lévy-Leblond and Lévy-Nahas (1965), Lewandowski (1996), Major and Seifert (2001), Carbone (2002), Neville (2006), Brunnemann and Rideout (2008, 2010) and Ding and Rovelli (2010). In addition, this paper is distinguished by its use of what we call the “$4j$-model” for the $6j$-symbol, in contrast to the “$12j$-model” used by Roberts (1999). The $4j$-model is less symmetrical than the $12$-model, but it is closer to the manner in which the $6j$-symbol is commonly used in recoupling theory. In addition, in loop quantum gravity (Rovelli 2004) angular momenta represent area vectors, which in the case of four-valent nodes correspond by Minkowski’s (1897) theorem to a tetrahedron. In this context the $4j$-model is closer to the applications than the $12j$-model, indeed, it played an important role in the work of Bianchi and Haggard (2011). In this paper we refer to Aquilanti (2007) as I, for example writing eqn. (I.13) for an equation from that paper. We note two errata in I, namely, $\sigma(x)$ in (I.89) should read $\sigma(u)$, and $j_3$ and $j_4$ should be swapped in the $6j$-symbol in (I.112). Spin network notation {#spinnetworknotation} ===================== We begin by explaining our notation for spin networks, which is based on that of Yutsis (1962) with modifications due to Stedman (1990). At the end of this section we compare our conventions for spin networks with others in the Yutsis tradition. The $3j$-symbol and Wigner intertwiner {#3jwigner} -------------------------------------- The $3j$-symbol is a number that can be regarded as the components of an intertwiner $W:\CS_{j_1}\otimes\CS_{j_2}\otimes\CS_{j_3 }\to\Complexes$ with respect to the standard basis $\ket{j_1m_1 }\otimes\ket{j_2m_2}\otimes\ket{j_3m_3}$, as indicated in Fig. \[3jnetwork\]. In this paper $\CS_j$ denotes a carrier space for unitary irrep $j$ of $SU(2)$, so that $\dim\CS_j = 2j+1$. We call $W$ the “Wigner” intertwiner. We will gradually explain the features of Fig. \[3jnetwork\] as we proceed. The standard notation for the $3j$-symbol is on the left of Fig. \[3jnetwork\], while the central diagram is the standard Yutsis spin network for the $3j$-symbol, with small arrows presented as in the Yutsis notation. The indices $(m_1,m_2,m_3)$ in the central diagram are covariant, that is, they transform under rotations as the components of a dual vector (in contrast to an ordinary vector). In a Hilbert space we regard ordinary wave functions or ket vectors as “vectors,” while bra vectors are regarded as “dual vectors.” Thus, contravariant indices are those that transform as the components of a vector. In the Yutsis notation the arrows indicate the transformation properties of the corresponding $m$ index, and there are rules for “raising and lowering” indices, that is, reversing the direction of the arrow. The rules do not, however, make use of the metric as in ordinary tensor analysis. Our definition of the arrow (explained below) is different from that of Yutsis, but designed so that the two notations agree as much as possible. In particular, our notation for the $3j$-symbol is the same as the Yutsis diagram in Fig. \[3jnetwork\]. A trivalent node of a spin network such as those illustrated in Fig. \[3jnetwork\] is assumed to have a positive or counterclockwise orientation, unless otherwise indicated (thus we dispense with the $+$ sign used by Yutsis). Bras, kets and scalar products {#brasketssps} ------------------------------ The diagram on the right of Fig. \[3jnetwork\] makes use of the standard basis vectors $\ket{j_im_i}$ in $\CS_{j_i}$, $i=1,2,3$. The spin networks for these basis vectors and their duals are shown in Fig. \[basisvectors\]. The large, broadly open arrow is a “chevron” (Stedman 1990). When pointing outward (inward), the chevron indicates a ket (bra) vector. Also shown in Fig. \[basisvectors\] is the spin network for an arbitrary vector $\ket{\psi}$ in $\CS_j$, and the dual bra vector $\bra{\psi}$ obtained by Hermitian conjugation. In the Dirac notation it is customary to label bras by the same symbol as kets, it being understood that the two are related by Hermitian conjugation. This convention is so deeply ingrained that we dare not change it. But in spin networks there are two different ways of converting kets into bras and vice versa, and this presents some notational challenges. One can see in Fig. \[basisvectors\] that Hermitian conjugation applied to bras and kets is notationally the changing of bra chevrons to ket chevrons and vice versa, and the starring of identifying symbols, with a double star being removed. Full rules for Hermitian conjugation of any spin network are given in Sec. \[Hermitianconjugation\]. The lines of a spin network will be referred to as “edges,” including cases like those shown in Fig. \[basisvectors\]. An edge of a spin network ending in an unstarred $m$ index represents a contraction with the basis ket $\ket{jm}$, so the $m$ index transforms under rotations as a covariant index. The explicit insertion of basis kets may be seen in Fig. \[3jnetwork\]. An edge ending in a starred $m$ index represents the insertion of a basis bra $\bra{jm}$, so starred indices are contravariant. Thus, the star on an $m$ index indicates its transformation property, and invariant contractions can only take place between a pair of starred and unstarred $m$ indices. As illustrated in Fig. \[orientation\], the orientation of a spin network on the page does not affect its value. The spin network in the figure has been rotated by $180^\circ$. As illustrated in the final diagram in Fig. \[3jnetwork\] and in Fig. \[scalarprod\], when a bra chevron and a ket chevron are juxtaposed, it represents the scalar product or contraction. In effect, the bra chevron acts as a receptacle for the ket chevron, and vice versa. After the contraction the two edges may be joined, with a small arrow remaining to indicate which was the bra and which the ket in the contraction. One might suppose that the star would carry the same information, but, as shown below, it is possible to change the direction of the arrow without changing the stars. Figure \[scalarprod\] also presents another example of Hermitian conjugation (complex conjugation, in this case). Since the small arrow represents the contraction of a bra and a ket chevron, when the chevrons are reversed, the direction of the arrow changes. As a special case of the scalar product, the orthonormality relations of the basis vectors are illustrated in Fig. \[basison\]. The final spin network follows from the symmetry of the Kronecker delta, or, alternatively, by the reality of $\delta_{mm'}$ and the rules for complex conjugation. The $j$ labels on the edges of the spin networks indicate which carrier space $\CS_j$ the ket or bra lies in, or in which carrier space a contraction has taken place. If there are distinct carrier spaces with the same $j$ label, then additional distinguishing information must be supplied. Intertwiners ------------ An $SU(2)$ intertwiner (the only kind we are interested in) is a linear map between two vector spaces that commutes with the action of $SU(2)$ on the two spaces. The only case of interest here is where the target space is $\Complexes$, consisting of scalars, that is, invariants under rotations. The Wigner intertwiner $W:\CS_{j_1}\otimes\CS_{j_2 }\otimes\CS_{j_3} \to \Complexes$ is of this type. But a linear map from a Hilbert space to $\Complexes$ can be thought of as a dual or bra vector, for example, we can associate the map $W$ with a bra vector $\bra{W}$ belonging to $\CS^*_{j_1}\otimes\CS^*_{j_2}\otimes\CS^*_{ j_3}$. The spin network notation for $\bra{W}$ is shown in Fig. \[Wintertwiner\]. The components of the Wigner intertwiner, that is, the $3j$-symbol, are obtained by inserting basis kets, the ket chevron first, into the bra chevrons of the intertwiner, as in the final diagram of Fig. \[3jnetwork\]. More generally, a $\Complexes$-valued intertwiner on a Hilbert space $\HS$ can be regarded as an $SU(2)$-invariant bra vector on this space, that is, a member of $\HS^*$. By Hermitian conjugation we obtain an $SU(2)$-invariant ket vector in $\HS$. Thus, there is a one-to-one correspondence between the subspace of $\HS$ of rotationally invariant vectors and the set of $\Complexes$-valued intertwiners on $\HS$. The subspace $\ZS$ introduced in Sec. \[4jmodel\] below is a subspace of this type, consisting of rotationally invariant vectors. Tensor products and resolution of identity {#tensorprodsroi} ------------------------------------------ The outer product of a ket with a bra is represented in spin network language simply by placing the spin networks for the ket and the bra on the same page, as illustrated in Fig. \[roi\]. The orientations of the bra and ket spin networks is immaterial, but in the figure they have been placed with their $m$ indices adjacent in order to emphasize the summation (a contraction over two indices, one covariant, one contravariant). The final diagram is the spin network for the identity operator on space $\CS_j$. Figure \[roi\] illustrates another technique, the replacement of a sum on $m$ by a joining of edges. The general usage is shown in Fig. \[msum\]. The directions of the arrows and the stars on one of the $m$’s must be coordinated as shown for the identity to be used as shown. The small arrow in the final diagram in Fig. \[msum\] (a fragment of a spin network) is a reminder of the directions of the arrows before the sum. This small arrow is omitted in the final diagram of Fig. \[roi\] (the identity diagram) because the chevrons already indicate the direction of the edge. Recall that edges are also joined on contracting a bra with a ket, as in Fig. \[scalarprod\]. The use of the identity diagram is illustrated in Fig. \[idj\]. The first row illustrates its action on kets (as a map $:\CS_j\to\CS_j$), where bra and ket chevons are combined as in Fig. \[scalarprod\]. The small arrow is omitted after the joining of the bra and ket chevrons because the remaining chevron indicates the direction of the edge. The second row of the figure illustrates its action on bras (the map $:\CS^*_j\to\CS_j^*$), and the third row illustrates its action as a map $:\CS_j^*\otimes\CS_j\to\Complexes$, that is, the scalar product map $:\bra{\phi}\otimes\ket{\psi}\mapsto\braket{\phi} {\psi}$. All of these usages are encompassed by the same spin network. The identity spin network can also be seen as an element of $\CS_j\otimes\CS^*_j$, that is, the space dual to $\CS_j^*\otimes\CS_j$ on which it, viewed as a $\Complexes$-valued linear operator, acts (the third line of Fig. \[idj\]). In general a spin network has some number of edges that terminate in incoming or outgoing chevrons. Examples are the identity diagram in Fig. \[roi\] and the Wigner intertwiner in Fig. \[Wintertwiner\]. In all cases there are multiple interpretations of the spin network as a linear operator mapping one vector space to another, depending on how many of the incoming and outgoing chevrons have ket and bra chevrons plugged into them (specifying the domain), and how many are left free (specifying the range). The domain is the tensor product of some number of $\CS_j$ times some number of $\CS_j^*$, and so is the range. In the extreme case that all incoming and outgoing chevrons on the spin network have kets and bras plugged into them the result is simply a number and the range is $\Complexes$. In that case the spin network, as a $\Complexes$-valued linear operator, can be seen as a vector in the space dual to the domain. This facile identification of closely associated operators, and their reinterpretation as elements of vector spaces, is an important advantage of spin networks. It is difficult and awkward to do something similar with the Dirac notation. In general, a tensor is a multilinear operator acting on a tensor product of some set of vector spaces and their duals. Thus, a spin network is a notation for a tensor on some product of the $\CS_j$ and their duals. The edges of the spin network terminating in ket or bra chevrons indicate the nature of the space the tensor acts on. In general, the tensor product of two tensors in spin network notation is indicated by the placing of the two spin networks together on the page, in any orientation. The outer product of a bra and a ket illustrated in Fig. \[roi\] is a special case. Partial or complete contractions of tensors are indicated by joining some or all ket chevrons with bra chevrons. The $2j$ symbol and intertwiner ------------------------------- We now consider another intertwiner, which leads to an important mapping between kets and bras, alternative to Hermitian conjugation. The intertwiner acts on the Hilbert space $\CS_j\otimes\CS'_j$, the tensor product of two carrier spaces of the same $j$. In general we wish to consider the second carrier space as distinct from the first, which is the purpose of the prime on the second factor. To within a normalization and phase, there is a unique vector in this space that is invariant under rotations; we call it $\ket{K}$, and express it in terms of the Clebsch-Gordan coefficients by $$\ket{K}= \sqrt{2j+1} \sum_{mm'} \ket{jm}\otimes\ket{jm'}\, C^{00}_{jjmm'}. \label{ketKdef}$$ This vector can also be expressed in terms of the “$2j$-symbol,” which we define in terms of the usual $3j$-symbol by $$\fl\left(\begin{array}{cc} j & j \\ m & m' \end{array}\right)= \left(\begin{array}{ccc} j & j & 0 \\ m & m' & 0 \end{array}\right) = C^{00}_{jjmm'} = \frac{(-1)^{j-m}}{\sqrt{2j+1}} \,\delta_{m,-m'}. \label{2jdef}$$ The terminology “$2j$-symbol” is not entirely standard, but it has been used by Stedman (1990). The invariant vector $\ket{K}$ can also be written, $$\fl\ket{K}=\sum_{mm'} \ket{jm}\otimes\ket{jm'}\, (-1)^{j-m} \,\delta_{m,-m'} = \sum_m \ket{jm}\otimes\ket{j,-m}\,(-1)^{j-m}. \label{K2jdef}$$ By Hermitian conjugation we convert $\ket{K}$ into the bra $\bra{K}$, which is otherwise an intertwiner $K:\CS_j\otimes\CS'_j\to\Complexes$. Just as the components of the intertwiner $W$ are the $3j$-symbol, the components of the intertwiner $K$ are the $2j$-symbol, multiplied however by $\sqrt{2j+1}$ because of a normalization convention. Figure \[2jnetwork\] shows first the spin network for the components of $K$, which is conceived of as a standard bivalent node or intertwiner. The small arrows indicate that the components on the first line can be considered the result of plugging basis kets into the intertwiner itself, as seen on the second line. The short line extending above the node is a “stub” (Stedman 1990), whose purpose is to orient the node. The convention is that if we start at the stub and move in a positive (counterclockwise) direction, the first and second edges we encounter are respectively the first and second operands of $K$, conceived of as a map $:\CS_j\otimes\CS'_j\to\Complexes$. We note that if $V$ and $W$ are vector spaces, then $V\times W$ is not the same as $V\otimes W$, but if we have a bilinear map on $V \times W$ it can be extended to a linear map on $V\otimes W$ by linear superposition. For example, in the previous paragraph we have regarded $K$ as a bilinear map $:\CS_j \times \CS'_j\to\Complexes$, and computed its components as $K(\ket{jm},\ket{jm'})$. The first and second operands of this expression correspond to the first and second edges as specified by the stub. Figure \[2jnetwork\] also gives the numerical values of the components of $K$, and in the final diagram, the network for $K$ itself, with two kets inserted. The network for $K$ in isolation is illustrated in Fig. \[2jKdef\], regarded as a bra vector on $\CS_{ j}\otimes\CS'_j$, that is, as an element of $\CS^*_j\otimes\CS^{\prime *}_j$. The stub in Fig. \[2jnetwork\] can be regarded as a vestigial edge of a $3j$-symbol or $W$-intertwiner with the value $j=0$, although one must beware of the normalization convention. This is illustrated in Fig. \[vestigialstub\], which is equivalent to (\[2jdef\]). The value does not depend on the direction of the arrow on the zero edge (see below for rules for reversing arrows). The components of $K:\CS_j\otimes\CS'_j\to\Complexes$, seen in Fig. \[2jnetwork\], acquire a phase of $(-1)^{2j}$ if $m$ and $m'$ are swapped. This is equivalent to the statement $$K(\ket{\phi},\ket{\psi}) = (-1)^{2j} K(\ket{\psi},\ket{\phi}), \label{Kswapphase}$$ for all $\ket{\psi}$, $\ket{\phi}$, which is illustrated in spin network language in Fig. \[Ksymm\]. The final diagram differs from the preceding simply by a $180^\circ$ rotation, so the value is the same. But this leads to the rule, that a spin network acquires a phase of $(-1)^{2j}$ when the stub at a bivalent node is inverted. In particular, the arrow on the null edge in Fig. \[vestigialstub\] can be inverted without changing the value. Kets to bras {#ketstobras} ------------ The intertwiner $K:\CS_j\otimes\CS_j\to\Complexes$ has an alternative interpretation as a map $K_1:\CS_j\to\CS^*_j$, where for simplicity we have dropped the prime on the second factor, and where the $1$-subscript distinguishes the new map from the old. This alternative interpretation is natural in spin network language, as seen in Fig. \[ket2bra\]. The spin network for $\ket{\psi} \in \CS_j$ is plugged into the second operand of the spin network for $K$, resulting in a spin network with one free bra chevron. The choice of the second operand of $K$ for this purpose is conventional. The result is an element of $\CS_j^*$, that is, it is a bra. As indicated in the figure, we abbreviate this spin network by drawing the same spin network for $\ket{\psi}$ that we started with, except the ket chevron is converted into a bra chevron. In other words, when we use $K_1$ to convert a ket into a bra, we just flip the ket chevron, leaving everything else the same. In particular, we do not put a star on the label of the ket. This distinguishes the map $K_1:\CS_j \to \CS_j^*$ from the metric or Hermitian conjugation, which is also a map $:\CS_j \to \CS_j^*$. When the metric is used to convert a ket to a bra, not only is the ket chevron flipped to a bra chevron, but a star is appended to the label. These two maps are quite distinct; in particular, $K_1$ is a linear map, while the metric is an antilinear map. As indicated in Fig. \[ket2bra\], the results are not the same. When a ket is turned into a bra, the components with respect to some basis change from contravariant to covariant. But since there is more than one way to do this, any notation based on the position (upper or lower) of the indices is inadequate to represent the result. The map $K_1$ can be used to turn a ket chevron into a bra chevron on any spin network, not only on kets themselves. A notation for an arbitrary spin network with an edge terminating in a ket chevron is shown in Fig. \[Xket\]. The circle around the $X$ indicates the rest of the spin network, which may include other edges terminating in ket or bra chevrons. To convert the ket chevron in Fig. \[Xket\] into a bra chevron we simply insert it into the second operand of $K:\CS_j\otimes\CS_j\to\Complexes$, as shown in Fig. \[lowerindex\]. The remainder of the spin network, indicated by the $X$, does not change. Bras to kets {#brastokets} ------------ The map $K_1:\CS_j\to\CS_j^*$ has an inverse, the map $K_1^{-1} :\CS_j^*\to\CS_j$, that takes bras into kets. We associate $K_1^{-1}$ with a closely related map $K^{-1}:\CS_j^*\otimes\CS_j^*\to\Complexes$ that is expressed by the spin network in Fig. \[Kinv\]. This spin network and the meaning of the $-1$ on $K^{-1}$ are to be defined, but the spin network represents a linear operator $:\CS_j^ *\otimes\CS_j^*\to\Complexes$ with the ordering of the two operands being specified by the stub. The action of $K^{-1}$ on two bras is illustrated in Fig. \[Kinvsandwich\]. We now define the network in Fig. \[Kinv\] by requiring that $K_1^{-1}$ act on a bra by inserting it into the first operand of that network, as illustrated in Fig. \[K1invaction\], and by requiring that $K_1^{-1}$ actually be the inverse of $K_1$. Figure \[K1invaction\] shows the action of $K_1^{-1}$ on an arbitrary bra $\bra{\phi}$. The use of the first operand of $K^{-1}$ for this purpose is a convention, but one that makes our overall notation for mapping kets to bras and vice versa consistent (see Fig. \[Kconsist\] below). As indicated, we abbreviate the result by taking the original network for the bra $\bra{\phi}$ and simply flipping the direction of the chevron. We do not unstar the identifying symbol. As indicated, the result differs from Hermitian conjugation applied to $\bra{\phi}$, which is the ket $\ket{\phi}$ (without the star). The requirement that $K_1^{-1}$ actually be the inverse of $K_1$ is illustrated in Fig. \[K1invreq\]. The stub on the network for $K^{-1}$ is inverted so that the output of the first step is fed into the first operand of $K^{-1}$. The identity represented by Fig. \[K1invreq\] is usually encountered in practice in the form shown in Fig. \[2stubs\] (a fragment of a spin network). The $2j$-node inverts the direction of the arrow. Two such inversions, with stubs pointing in opposite directions, annihilate one another. We have made an independent definition of the spin network in Fig. \[Kinv\], but it is the same as the spin network for $K$, shown in Fig. \[2jKdef\], with both bra chevrons flipped. Since we now have a convention for flipping bra chevrons (by applying $K_1^{-1}$), for consistency we must show that the two results are the same. This is done in Fig. \[Kconsist\], which uses the identity of Fig. \[2stubs\]. By inserting resolutions of the identity into the diagram in Fig. \[K1invreq\] it is easy to work out the components of $K^{-1}$. These are displayed in Fig. \[Kinvcomponents\]. Notice that they have the same numerical values as the components of $K$ (see Fig. \[2jnetwork\]). By using $K^{-1}$ we can convert a bra chevron into a ket chevron on any spin network, not only on bras and kets themselves. This is illustrated in Fig. \[raiseindex\], which may be compared to Fig. \[lowerindex\]. Finally, by using Figs. \[lowerindex\], \[raiseindex\] and \[2stubs\], it may be shown that when we reverse the arrow on an edge of a spin network, we incur a phase of $(-1)^{2j}$. This is done in Fig. \[reversearrow\]. Raising and lowering indices {#raisingandlowering} ---------------------------- When we convert a ket to a bra by the action of $K_1$, then the bra has components with respect to the standard basis $\ket{jm}$ that are simple functions of the components of the original ket with respect to the standard basis $\bra{jm}$. Mapping the one set of components to the other is “lowering the index.” Using $K_1^{-1}$ to convert a bra to a ket similarly amounts to “raising the index.” More generally the procedure can be applied to an edge of any spin network terminating in a starred (contravariant) or an unstarred (contravariant) index. The index can refer to any basis, not just the standard one. Figure \[raisem\] shows how to the express contravariant components in terms of the covariant components in the standard basis. By plugging in the numerical values of the components of $K$, we obtain the first line of Fig. \[raiseandlower\]. Similarly we derive the second line of Fig. \[raiseandlower\] for expressing covariant components in terms of contravariant components. If these rules are used to raise all three covariant components of the $3j$-intertwiner (Fig. \[3jnetwork\]), then we find that the completely contravariant components have the same values, namely, the $3j$-symbol. This is illustrated in Fig. \[3jversions\]. To show this it is necessary to use the symmetry of the $3j$-symbol (see Varshalovich  1981, Eq. (8.2.4.6)). Then by setting one of the $j$’s to zero and using Fig. \[vestigialstub\], we find that the same is true for the completely covariant and completely contravariant components of the $2j$-intertwiner, as shown in Fig. \[2jversions\]. The same result is obtained by comparing Figs. \[2jnetwork\] and \[Kinvcomponents\]. Hermitian conjugation of spin networks {#Hermitianconjugation} -------------------------------------- Consider a spin network of arbitrary complexity involving only $2j$- and $3j$-nodes. The network is allowed to have any number of edges terminating in bra or ket chevrons, or in starred or unstarred labels such as $m$ indices. By using the identities above, possibly with the insertion or removal of $2j$-nodes and the extraction of phases of the form $(-1)^{2j}$, it is possible to bring the spin network into a standard form, in which all edges joining $3j$-nodes have arrows pointing toward the $3j$-node, all edges joining $2j$-nodes have arrows pointing away from the $2j$-node, all edges terminating in a starred symbol have arrows pointing toward that symbol, and all edges terminating in an unstarred symbol have arrows pointing away from that symbol. Next, by inserting resolutions of the identity, which involve $m$-sums, it is possible to express the spin network as sum over the completely covariant components of $3j$-symbols and completely contravariant components of $2j$-symbols, times a tensor product of bras and kets. In this form it is easy to take the Hermitian conjugate. Under Hermitian conjugation, bras go to kets and vice versa, while the covariant components of $3j$-symbols and contravariant components of $2j$-symbols do not change, since they are real. By using Figs. \[3jversions\] and \[2jversions\], however, these components can be rewritten as the completely contravariant components of $3j$-symbols and the completely covariant components of $2j$-symbols. The $m$-sums can now be done, reversing the earlier insertions of resolutions of the identity. Then the other steps leading to the standard form can be reversed. The result is a simple rule for the Hermitian conjugation of any spin network of the given form: All ket chevrons are changed to bra chevrons and vice versa, the directions of all arrows are reversed, and all edges terminating in a symbol have a star added to the symbol, with a double star being removed. Discussion of spin network rules {#spinnetworkdiscussion} -------------------------------- Our rules for spin networks differ from those of Yutsis  and most of the literature in the Yutsis tradition primarily by our ability to express abstract vectors (kets), dual vectors (bras) and tensors in addition to the components of those objects. Also, we indicate the nature of an $m$ index (covariant or contravariant) by the presence or absence of a star, rather than the direction of the arrow. One result is that our rules for reversing the direction of the arrow are more uniform than in the Yutsis tradition, where such a reversal picks up a phase $(-1)^{2j}$ only on internal edges. In our approach, the rule applies everywhere, including edges terminating in an $m$ index. In addition, our rules for Hermitian conjugation are simpler than those in the Yutsis tradition, where phase factors must be introduced. The simplification is due to the explicit introduction of $2j$-symbols, and the use of stubs. To translate a Yutsis spin network into one of ours, it is necessary only to put stars on $m$ indices terminating edges with outward pointing arrows. The standard form of a spin network discussed in Sec. \[Hermitianconjugation\] has all $3j$-nodes with inward pointing arrows, and all $2j$-nodes with outward pointing ones. If we assume the standard form, then the arrows become superfluous and can be dropped. Only the stubs and $2j$-nodes remain, in comparison to a Yutsis-style spin network. This is the procedure advocated by Stedman (1990). For the purposes of this paper we will keep the arrows, since we wish to have finer control on the spin network than that offered by the standard form. Models for the $6j$-symbol {#models} ========================== For given values of the six $j$’s, the $6j$-symbol is just a number, but to study its semiclassical limit it is useful to write it as a scalar product $\braket{B}{A}$ of wave functions in some Hilbert space. This can be done in many different ways, corresponding to what we call different “models” of the $6j$-symbol. In this section we describe a class of such models that are related to one another. We begin by summarizing our notation for the Schwinger representation of angular momentum operators. Then we present what we call the “$12j$-model,” which was used by Roberts (1999) in his derivation of the Ponzano-Regge formula. Next we describe the “$4j$-model” which we will use for the semiclassical analysis of this paper. We also mention an $8j$-model for the $6j$-symbol. The Schwinger Representation {#schwingerrep} ---------------------------- Our notation for the Schwinger representation of angular momentum operators is similar to that used in I. We denote the Schwinger Hilbert space by $\SS=L^2(\Reals^2)$; it is the space of wave functions $\psi(x_1,x_2)$ for two harmonic oscillators of unit frequency and mass. The usual annihilation and creation operators are $\ahat_\mu=(\xhat_\mu + i\phat_\mu) /\sqrt{2}$, $\ahat_\mu^\dagger= (\xhat_\mu-i\phat_\mu) /\sqrt{2}$, for $\mu=1,2$; we use hats on operators to distinguish them from their classical counterparts. We define operators $$\Ihat = \frac{1}{2}\ahat^\dagger \ahat, \qquad \Jhat_i = \frac{1}{2}\ahat^\dagger \sigma_i \ahat, \label{IhatJhatdef}$$ where $i=1,2,3$ and where $\ahat$ (without the $\mu$ index) is a 2-vector (or column spinor) of operators, with $\ahat^\dagger$ the adjoint (or row spinor) and with obvious contractions against the Pauli matrices $\sigma_i$. These operators satisfy the commutation relations $[\Ihat,\Jhat_i]=0$, $[\Jhat_i,\Jhat_j] = i\epsilon_{ijk} \, \Jhat_k$. We set $\hbar=1$. Note that $\Ihat = (\Hhat_1+\Hhat_2-1)/2$, where $\Hhat_\mu$, $\mu=1,2$, are the two harmonic oscillators. There is also the operator relation $\Jvechat^2 = \Ihat(\Ihat+1)$. We denote the squares of 3-vectors in bold face. The operators $\Jvechat$ generate an $SU(2)$ action on $\SS$, which carries one copy of each irrep $j=0,1/2,1,\ldots$, that is, $$\SS = \sum_j \oplus \,\CS_j. \label{SSdirectsum}$$ The irreducible subspace $\CS_j$ is an eigenspace of $\Ihat$ with eigenvalue $j$. In the semiclassical analysis of spin networks, the spaces $\CS_j$ that the spin networks refer to are interpreted as one of the irreducible subspaces of a Schwinger Hilbert space $\SS$. Similarly, $\CS_j^*$ is interpreted as a subspace of a space $\SS^*$. In this way the bra and ket vectors referred to by the spin network are interpreted as wave functions on $\Reals^2$, and the spin network itself can be interpreted as a wave function in $\Reals^{2N}$. In the various $nj$-models we take tensor products of the Schwinger Hilbert space, writing $\SS_r$ for the $r$-th copy. Similarly, we put an $r$ index on various operators, for example, $\ahat_r$, $\Ihat_r$, $\Jvechat_r$, $r=1,\ldots,n$, or, with two indices, $\ahat_{r\mu}$, $\mu=1,2$. The $12j$-model of the $6j$-symbol {#12jmodel} ---------------------------------- We begin with the standard spin network (Yutsis 1962) of the $6j$-symbol, shown in Fig. \[symm6j\]. According to the remarks in Sec. \[spinnetworkdiscussion\], this spin network can be reinterpreted according to our conventions, presented in Sec. \[spinnetworknotation\], without modification. We will refer to the labeling of the six $j$’s in the $6j$-symbol shown in Fig. \[symm6j\] as the “symmetric” labeling. We perform an operation on each edge of the spin network that is illustrated for edge 1 in Fig. \[2jinsert\]. After the second equality edges are labeled by $1$ and $1'$. These refer to two distinct carrier spaces, say, $\CS_{j_1}$ and $\CS'_{j_1}$, with the same value of $j$ (that is, $j_1$). The introduction of such distinct carrier spaces does not change the value of the $6j$-symbol, which is just a number. In the second equality we have expressed the lower ket chevron as a bra chevron transformed by $K_1^{-1}$, as in Fig. \[raiseindex\]. In the final diagram the arrows are directed toward both $3j$-nodes connected by the original edge, and a $2j$-node has been inserted. We do this on all six edges of the spin network in Fig. \[symm6j\]. The resulting diagram is somewhat busy so we do not attempt to draw it, but each edge of the original diagram now looks like the final diagram in Fig. \[2jinsert\]. We now break up the final diagram in Fig. \[2jinsert\] in two different ways. The first way is illustrated in Fig. \[edmonds\], in which the primed and unprimed lines are broken into summations over primed and unprimed quantum numbers $m$. This is done for all six edges of the original spin network. The $6j$-symbol is represented as a product of six copies of a $2j$-symbol and four of a $3j$-symbol. Using the definition (\[2jdef\]) for the $2j$-symbol, the result is $$\fl\eqalign{ &\left\{\begin{array}{ccc} j_1 & j_2 & j_3 \\ j_4 & j_5 & j_6 \end{array}\right\} = \left(\prod_{r=1}^6 \sqrt{2j_r+1}\right) \sum_{\hbox{all $m$'s}} \\ &\quad\times \left(\begin{array}{ccc} j_1 & j_2 & j_3 \\ m_1 & m_2 & m_3 \end{array}\right) \left(\begin{array}{ccc} j_1 & j_5 & j_6 \\ m'_1 & m'_5 & m_6 \end{array}\right) \left(\begin{array}{ccc} j_2 & j_6 & j_4 \\ m'_2 & m'_6 & m_4 \end{array}\right) \left(\begin{array}{ccc} j_3 & j_4 & j_5 \\ m'_3 & m'_4 & m_5 \end{array}\right) \\ &\quad\times \left(\begin{array}{cc} j_1 & j_1 \\ m_1 & m'_1 \end{array}\right) \cdots \left(\begin{array}{cc} j_6 & j_6 \\ m_6 & m'_6 \end{array}\right)} \label{edmonds6j}$$ This formula may be compared to Edmonds (1960), eq. (6.2.3). Edmonds uses what he calls a “metric tensor” (really the components of $K$ or $K^{-1}$, multiplied by $(-1)^{2j}$, see his eq. (3.7.1)), which relative to our $2j$-symbol introduces an overall phase of $\prod_{r=1}^6 (-1)^{2j_r}$. He also swaps $m_r$ and $m'_r$ for $r=4,5,6$ relative to our definitions, which introduces further phases. The product of these phases is 1, showing that the formulas agree. Another way of breaking up the $6j$-symbol is to stop with the third diagram of Fig. \[2jinsert\] in the transformation of the six edges of the $6j$-symbol. Again the resulting diagram is too busy to draw, but it can be regarded as the complete contraction of two tensors, one the tensor product of six $2j$-intertwiners, all terminating in ket chevrons, the other the tensor product of four $3j$-intertwiners, all terminating in bra chevrons. There are twelve ket chevrons and twelve bra chevrons altogether, which we think of as living in twelve carrier spaces $\CS_{j_r}$ and $\CS'_{j_r}$ and their duals, where $r=1,\ldots,6$. These can be viewed as subspaces of twelve Schwinger Hilbert spaces, $\SS_r$, $\SS'_r$, $r=1,\ldots,6$, and their duals. Then the $6j$-symbol takes on the form $\braket{B}{A}$, where states $\ket{A}$, $\ket{B}$ belong to the total Hilbert space $\HS_{12j} = (\prod_r\otimes \SS_r)\otimes (\prod_r\otimes\SS'_r)$. These states are illustrated in Fig. \[robertsAB\], where the state $\ket{B}$ has been turned into ket form by Hermitian conjugation. The usual custom in physics is to specify a state by the operators and quantum numbers of which the state is a simultaneous eigenstate. This requires that the eigenstate be nondegenerate, so that it is determined to within a normalization and phase. This in turn requires, in a certain sense, that the number of independent operators should be equal to the number of degrees of freedom of the system. We will not attempt to be precise about this statement, but will illustrate the principle in several examples. One example is the first $3j$-state appearing in Fig. \[robertsAB\], which lies in the Hilbert space $\SS_1 \otimes \SS_2 \otimes \SS_3$ and is a simultaneous eigenstate of $\Ihat_r$, $r=1,2,3$ with eigenvalues $j_r$, $r=1,2,3$. It is also an eigenstate of the vector of operators $$\Jvechat_{123} = \Jvechat_1 + \Jvechat_2 + \Jvechat_3, \label{Jvechat123def}$$ the total angular momentum on this Hilbert space, with eigenvalue $\zerovec$. That this simultaneous eigenstate is nondegenerate follows from standard angular momentum theory; and the number of operators (six) equals the number of degrees of freedom in the Hilbert space (two for each $\SS_r$, $r=1,2,3$). Thus we write this state as illustrated in Fig. \[3jket\], indicating both operators and eigenvalues. This is otherwise the state $\ket{W}$, illustrated in bra form in Fig. \[Wintertwiner\]. As for the normalization and phase, these must be supplied by context. For the $3j$-state illustrated in Fig. \[3jket\], these are given in terms of the $3j$-symbol by Fig. \[3jnetwork\]; in particular, the state is normalized. Similarly, the first $2j$-state in Fig. \[robertsAB\] lies in the Hilbert space $\SS_1\otimes\SS'_1$ and is a simultaneous eigenstate of the operators $\Ihat_1$ and $\Ihat'_1$ with eigenvalues $j_1$ and $j_1$, as well as of the total angular momentum operator on this space, $$\Jvechat_{11'} = \Jvechat_1 + \Jvechat'_1, \label{Jvechat11'}$$ with eigenvalue $\zerovec$. This state is a simultaneous eigenstate of five operators, but in a sense only two of the three components of $\Jvechat_{11'}$ are independent, so we should count only four independent operators, which agrees with the number of degrees of freedom in the Hilbert space (again, two each for $\SS_1$ and $\SS'_1$). We write this state as illustrated in Fig. \[2jket\]; the normalization and phase are given by the components of $K^{-1}$ shown in Fig. \[Kinvcomponents\]. In particular, with the square root factor in Fig. \[2jket\], this state is normalized. In this notation we can write the equations of Fig. \[robertsAB\] in the form $$\fl\ket{B} = \Wket{\Ihat_1}{\Ihat_2}{\Ihat_3}{\Jvechat_{123}} {j_1}{j_2}{j_3}{\zerovec} \Wket{\Ihat'_1}{\Ihat'_5}{\Ihat_6}{\Jvechat_{1'5'6}} {j_1}{j_5}{j_6}{\zerovec} \Wket{\Ihat'_2}{\Ihat'_6}{\Ihat_4}{\Jvechat_{2'6'4}} {j_2}{j_5}{j_6}{\zerovec} \Wket{\Ihat'_3}{\Ihat'_4}{\Ihat_5}{\Jvechat_{3'4'5}} {j_3}{j_4}{j_5}{\zerovec}, \label{roberts3jstate}$$ and $$\fl\ket{A} = \left(\prod_{r=1}^6 \sqrt{2j_r+1}\right) \Kket{\Ihat_1}{\Ihat'_1}{\Jvechat_{11'}} {j_1}{j_1}{\zerovec} \cdots \Kket{\Ihat_6}{\Ihat'_6}{\Jvechat_{66'}} {j_6}{j_6}{\zerovec}. \label{roberts2jstate}$$ One can see that Edmonds’ form of the $6j$-symbol (\[edmonds6j\]) is equal to $\braket{B}{A}$. This scalar product was the starting point for Roberts’ (1999) analysis of the asymptotics of the $6j$-symbol. We shall comment below on further aspects of Roberts’ calculation. The Triangle and Polygon Inequalities {#polygoninequals} ------------------------------------- We make a remark on a generalization of the triangle inequalities before presenting the $4j$-model of the $6j$-symbol. If $(\ell_1,\ell_2,\ell_3)$ are three nonnegative lengths, the usual triangle inequalities are $|\ell_i-\ell_j| \le \ell_k \le \ell_i+\ell_j$, where $(i,j,k)=(1,2,3)$ and cyclic permutations. We generalize these as follows. Let $\{\ell_i,i=1,\ldots,n\}$ be a set of lengths, $\ell_i\ge0$, $i=1,\ldots,n$. Then this set is said to satisfy the “polygon inequality” if $$\max\{\ell_i\} \le \frac{1}{2}\sum_{i=1}^n \ell_i. \label{polygoninequal}$$ This is equivalent to the triangle inequalities when $n=3$. In general, it represents the necessary and sufficient condition that line segments of the given, nonnegative lengths can be fitted together to form a polygon with $n$ sides (in $\Reals^N$, $N>0$). The $4j$-model of the $6j$-symbol {#4jmodel} --------------------------------- A different way of writing the $6j$-symbol as a scalar product begins with Fig. \[asym6j\], in which the $j$’s in the $6j$-symbol of Fig. \[symm6j\] have been relabeled. We will refer to the labeling in Fig. \[asym6j\] as the “asymmetric” labeling, which is more appropriate for the $4j$-model. After the relabeling, we have reversed the arrow on the edge $j_3$ of the spin network (labeled simply by 3), incurring a phase $(-1)^{2j_3}$, and broken four edges into scalar products of a bra and a ket. Then, on the second line of Fig. \[asym6j\], we have unfolded the bra and the ket, written the bra as the Hermitian conjugate of a ket, and adjusted phases. The result expresses the $6j$-symbol as a phase times a scalar product of two states lying in the Hilbert space $\HS_{4j} =\prod_{r=1}^4 \otimes \SS_r$. It is understood that the ket terminating a line labeled $r$ lies in the $j_r$-irreducible subspace of $\SS_r$. These same states arise in the recoupling of four angular momenta with a resultant of zero. Let $\ZS$ be the subspace of $\CS_{j_1} \otimes \CS_{j_2} \otimes\CS_{j_3} \otimes \CS_{j_4}$ upon which $$\Jvechat_{\rm tot} = \sum_{r=1}^4 \Jvechat_r = 0. \label{Jtotdef}$$ This is the subspace of rotational invariants, that is, $\ZS$ is the set of states $\ket{\psi }\in \CS_{j_1}\otimes \CS_{j_2} \otimes\CS_{j_3} \otimes \CS_{j_4}$ that are invariant under rotations. According to the rules for addition of angular momenta, subspace $\ZS$ is nontrivial ($\dim\ZS>0$) if $\sum_{r=1}^4 j_r={\rm integer}$ and if the set $\{j_r\}$ satisfies the polygon inequality (\[polygoninequal\]). In accordance with the remarks in Sec. \[intertwiners\], the subspace $\ZS$ can also be interpreted as the space of 4-valent intertwiners, that is, $SU(2)$-invariant maps $:\CS_{j_1}\otimes \CS_{j_2} \otimes\CS_{j_3} \otimes \CS_{j_4}\to\Complexes$. The notation $\ZS$ is a mnemonic for “zero” (the eigenvalue of $\Jvechat_{\rm tot}$). As far as recoupling theory is concerned the spaces $\CS_{j_r}$ can be any carrier spaces of $SU(2)$ for the given values of $j_r$, but in our application we shall interpret the space $\CS_{j_r}$ as the irreducible subspace $j_r$ of the $r$-th copy of the Schwinger Hilbert space $\SS_r$. Then $\ZS$ becomes a subspace of $\HS_{4j}$. We shall assume that the fixed values of the four $j_r$, $r=1,\ldots,4$ are chosen such that $\dim\ZS>0$. We should properly label $\ZS$ by the four $j_r$ values since $\HS_{4j}$ contains as many subspaces of the type $\ZS$ as there are choices of the four $j$’s. For simplicity, however, we will suppress this dependence in the notation, it being understood that $j_r$, $r=1,\ldots,4$ are given. Standard recoupling theory gives three ways of constructing an orthonormal basis in $\ZS$. One uses Clebsch-Gordan coefficients to couple angular momenta according to the pattern $1+2=12$, $12+3=123$, $123+4=0$, resulting in the normalized state $\ket{B}$ lying in $\ZS$, expressed in terms of a spin network in Fig. \[4jABstates\]. A second way couples according to the pattern $2+3=23$, $1+23=123$, $123+4=0$, producing the normalized state $\ket{A}$ illustrated in Fig. \[4jABstates\]. A third way, which we will not consider further, uses the intermediate coupling $1+3=13$. The quantum numbers $j_{12}$ and $j_{23}$ of the intermediate angular momenta range in integer steps between the bounds $$\eqalign{ j_{12,{\rm min}} &\le j_{12} \le j_{12,{\rm max}}, \\ j_{23,{\rm min}} &\le j_{23} \le j_{23,{\rm max}},} \label{j12j23range}$$ where the maximum and minimum values are given in terms of the four fixed $j_r$, $r=1,\ldots,4$ by $$\eqalign{ j_{12,{\rm min}}&=\max(|j_1-j_2|,|j_3-j_4|), \quad j_{12,{\rm max}}=\min(j_1+j_2,j_3+j_4), \\ j_{23,{\rm min}}&=\max(|j_2-j_3|,|j_1-j_4|), \quad j_{23,{\rm max}}=\min(j_2+j_3,j_1+j_4).} \label{j12j23bounds}$$ Then the dimension of $\ZS$ is given by $$D=\dim\ZS = j_{12,{\rm max}}-j_{12,{\rm min}}+1 = j_{23,{\rm max}}-j_{23,{\rm min}}+1. \label{dimZS}$$ An expression for $\dim\ZS$ can be given that is symmetrical in $(j_1,j_2,j_3,j_4)$. Using the fact that $|x|=\max(x,-x)$, the difference between $j_{12,{\rm max}}$ and $j_{12,\rm{min}}$ becomes the shortest distance between one set of four numbers, $\{j_1-j_2, j_2-j_1, j_3-j_4, j_4-j_3\}$, and another set of two numbers, $\{j_1+j_2, j_3+j_4\}$. But this is the minimum of the distance between all eight possible pairs taken from the two sets. Thus $$D=\dim\ZS = 2\min(j_1,j_2,j_3,j_4,s-j_1,s-j_2,s-j_3,s-j_4) +1, \label{symmdimZS}$$ where $s$ is the semiperimeter, $$s=\frac{1}{2}(j_1+j_2+j_3+j_4). \label{sdef}$$ More precisely, if $D$ computed by (\[symmdimZS\]) is $\le0$, then subspace $\ZS$ is trivial ($\dim\ZS=0$); otherwise $D=\dim\ZS$. The formula (\[symmdimZS\]) bears an interesting relationship to the Regge symmetries of the $6j$-symbol (Varshalovich  1981, Eq. (9.4.2.4)). Now by comparing Figs. \[asym6j\] and \[4jABstates\] we see that the $6j$-symbol is proportional to a scalar product, $$\braket{B}{A}=(-1)^{j_1+j_2+j_3+j_4}\, \sqrt{(2j_{12}+1)(2j_{23}+1)} \left\{\begin{array}{ccc} j_1 & j_2 & j_{12} \\ j_3 & j_4 & j_{23} \end{array}\right\}. \label{6jme}$$ This is the scalar product that we shall use for the semiclassical analysis of the $6j$-symbol in this paper. In a different notation, we can write $\braket{j_{23}}{j_{12}}$ instead of $\braket{B}{A}$, which emphasizes the fact that this is a unitary matrix element connecting two bases on $\ZS$. The usual orthonormality relations satisfied by the $6j$-symbol (see Edmonds eq. (6.2.9)) are equivalent to the unitarity of $\braket{j_{23}}{j_{12}}$. Notice that in this $4j$-model, the angular momenta $\Jvechat_r$, $r=1,2,3,4$ are independent operators, while the two remaining angular momenta, $$\Jvechat_{12} = \Jvechat_1 + \Jvechat_2, \qquad \Jvechat_{23} = \Jvechat_2 + \Jvechat_3, \label{Jvechat1223defs}$$ are not, rather they are functions of the first four. As usual in the Schwinger representation, the quantum number $j_r$, $r=1,\ldots,4$ specifies the eigenvalues of both $\Ihat_r$ and $\Jvechat_r^2$, that is, $j_r$ and $j_r(j_r+1)$, respectively. And the quantum numbers $j_{12}$ and $j_{23}$ specify the eigenvalues of the operators $\Jvechat_{12}^2$ and $\Jvechat_{23}^2$, that is, $j_{12}(j_{12}+1)$ and $j_{23}(j_{23}+1)$, respectively. But there are no operators $\Ihat_{12}$ or $\Ihat_{23}$. The states $\ket{A}$ and $\ket{B}$ in Fig. \[4jABstates\] can be expressed as eigenstates of complete sets of operators, $$\ket{A}=\left| \begin{array}{@{\;}c@{\,}c@{\,}c@{\,}c@{\;}c@{\;}c@{}} \Ihat_1 & \Ihat_2 & \Ihat_3 & \Ihat_4 & \Jvechat_{12}^2 & \Jvechat_{\rm tot} \\ j_1 & j_2 & j_3 & j_4 & j_{12} & \zerovec \end{array}\right>, \qquad \ket{B}=\left| \begin{array}{@{\;}c@{\,}c@{\,}c@{\,}c@{\;}c@{\;}c@{}} \Ihat_1 & \Ihat_2 & \Ihat_3 & \Ihat_4 & \Jvechat_{23}^2 & \Jvechat_{\rm tot} \\ j_1 & j_2 & j_3 & j_4 & j_{23} & \zerovec \end{array}\right>, \label{4jABdefs}$$ in a notation like that used in (\[roberts3jstate\]) and (\[roberts2jstate\]). As mentioned, these states are normalized, and their phases are specified by Fig. \[4jABstates\]. Notice that each state $\ket{A}$ and $\ket{B}$ has a list of eight independent operators (counting the three components of $\Jvechat_{\rm tot}$), corresponding to the eight degrees of freedom in the $4j$-model. We will call these lists of operators the $A$-list and $B$-list, and write them collectively as $$\eqalign{ \Ahat &= (\Ihat_1,\Ihat_2,\Ihat_3,\Ihat_4, \Jvechat_{12}^2,\Jvechat_{\rm tot}), \\ \Bhat &= (\Ihat_1,\Ihat_2,\Ihat_3,\Ihat_4, \Jvechat_{23}^2,\Jvechat_{\rm tot}),} \label{ABoplists}$$ We denote elements of these lists with subscripts, for example, $\Ahat_i$ or $\Bhat_i$, $i=1,\ldots,8$. The operators in the either of the lists (\[ABoplists\]) do not commute with one another (because the components of $\Jvechat_{\rm tot}$ do not commute; otherwise all commutators are zero), but they do possess simultaneous eigenstates in $\ZS$, which are unique to within a phase, namely, the states (\[4jABdefs\]). The $8j$-model of the $6j$-symbol {#8jmodel} --------------------------------- We obtain an $8j$-model of the $6j$-symbol by inserting $2j$-symbols into edges 12 and 23 of the spin network of Fig. \[asym6j\] and then treating them in the same way as the $2j$-symbols in the $12j$-model. The result is the Hilbert space $\HS_{8j}= (\prod_{r=1}^4 \otimes\SS_r) \otimes\SS_{12} \otimes\SS'_{12} \otimes\SS_{23} \otimes\SS'_{23}$. This model has more symmetry than the $4j$-model but less than the $12j$-model. Operators $\Ihat_{12}$ and $\Ihat_{23}$ exist in this model (as well as operators $\Ihat'_{12}$ and $\Ihat'_{23}$), unlike the $4j$-model. This gives the $8j$-model certain advantages over the $4j$-model. We shall not consider the $8j$-model further in this paper. The Classical Manifolds {#classmanifolds} ======================= In this section we study the classical mechanics that will be relevant for the semiclassical analysis of the $6j$-symbol in a $4j$-model. We begin by presenting our notation for the Schwinger phase space and products of it that are used to represent coupled, classical angular momenta. Other spaces that will be important are obtained by Poisson and symplectic reduction. Then we examine the geometry of the $A$- and $B$-manifolds in phase space that support the states (\[4jABdefs\]), including a rather general analysis of why they are Lagrangian. Finally we discuss the Bohr-Sommerfeld quantization of these manifolds. The Schwinger Phase Space and Other Spaces {#schwingerphasespace} ------------------------------------------ The classical phase space for two harmonic oscillators (the Schwinger phase space) is $\Phi=(\Reals^4,dp\wedge dx)$, with coordinates $(x,p) \in \Reals^4$, for $x,p \in \Reals^2$. See Sec. I. for more details on our use of the Schwinger representation, as well as Cushman and Bates (1997) for a more detailed discussion of the geometry of two harmonic oscillators and the role played by the Hopf map. Here $dp\wedge dx$ means $\sum_{\mu=1}^2 dp_\mu \wedge dx_\mu$, and similarly for other obvious contractions over $\mu=1,2$. Complex coordinates $z_\mu= (x_\mu+ip_\mu)/\sqrt{2}$, $\zbar_\mu = (x_\mu-ip_\mu)/\sqrt{2}$ are also useful, allowing us to write $\Phi=(\Complexes^2,id\zbar\wedge dz)$. Interesting functions on $\Phi$ (classical observables) include $$I = \frac{1}{2} \zbar z, \qquad J_i = \frac{1}{2} \zbar \sigma_i z, \label{IJidef}$$ where $i=1,2,3$, obviously the classical analogs of (\[IhatJhatdef\]), and where $z$ or $\zbar$ without indices indicates a 2-component “spinor,” that is, an element of $\Complexes^2$. In comparison to (\[IhatJhatdef\]) notice the absence of the hats, indicating that these are classical observables. These satisfy the Poisson bracket relations $\{I,J_i\}=0$ and $\{J_i,J_j\} = \epsilon_{ijk}\,J_k$, as well as the identity $\Jvec^2 = I^2$. The Hamiltonian flow of $I$ is a $U(1)$ action on $\Phi$, while the flows of $J_i$, $i=1,2,3$, generate an $SU(2)$ action on $\Phi$. Both actions are easily expressed in the complex coordinates $z$: that of $U(1)$ is $z\mapsto \exp(-i\psi/2) z$, where $\psi$ is the variable conjugate to $I$, while that of $SU(2)$ is $z \mapsto uz$ for $u\in SU(2)$. The orbit of the $U(1)$ action generated by $I$ passing through any point $z\ne0$ on $\Phi$ is a circle on which $\psi$ is a coordinate, covered once when $0\le \psi < 4\pi$. The level set $I=J$ for $J>0$, where $I:\Phi\to\Reals$ is the function and $J\in\Reals$ is the contour value, is a 3-sphere to which the orbits of $I$ are confined. The definition (\[IJidef\]) of $J_i$ is interpreted as a map $\pi :\Phi \to \Reals^3$, where the three components $J_i$ of $\Jvec$ are coordinates on $\Reals^3$ (thus, this space is “angular momentum space”). This map is a Poisson map (Marsden and Ratiu 1999), giving $\Reals^3$ the Poisson structure $\{J_i,J_j\} = \epsilon_{ijk}\,J_k$. We denote $\Reals^3$ with this Poisson structure by $\Lambda$, as indicated by the diagram, $$\begin{diagram} (I=J) & \subset & \Phi \\ \dTo>{\pi_H} & & \dTo>\pi \\ \Sigma & \subset & \Lambda \\ \end{diagram} \label{hopfmaps}$$ The map $\pi$ can also be interpreted as the momentum map (Abraham and Marsden 1978) of the $SU(2)$ action on $\Phi$, so that $\Lambda$ or angular momentum space is identified with ${\mathfrak s}{\mathfrak u}(2)^*$. When $\pi$ is restricted to a level set $I=J>0$ in $\Phi$ (a 3-sphere), it projects onto the 2-sphere $|\Jvec| = J$ in $\Lambda$. This is the projection map $\pi_H$ of the Hopf fibration, in which the orbits of $I$ are the fibers (or “Hopf circles”). It is also the map of symplectic reduction (Abraham and Marsden 1978) of the level set $I=J$ in $\Phi$ by the $U(1)$ action, so that the 2-spheres $|\Jvec|=J$ in $\Lambda$ are symplectic manifolds. The symplectic form on one of these 2-spheres is $J\,d\Omega$, where $d\Omega$ is the element of solid angle (to within a sign). We denote one of these spheres with its symplectic structure (for some value $J>0$) by $\Sigma$; these spheres are also the symplectic leaves of the Poisson structure in $\Lambda$. In an $nj$-model of a spin network we take Cartesian products of the Schwinger phase space $\Phi$ to obtain the phase space for $n$ independent classical angular momenta. We will illustrate the notation for the $4j$-model. We write $\Phi_{4j} = \Phi^4$ for the entire phase space; apart from the symplectic structure, this is $\Complexes^8 = \Reals^{16}$. Coordinates on $\Phi_{4j}$ are $x_{\mu r}$, $p_{\mu r}$, $z_{\mu r}$, etc, $\mu=1,2$, $r=1,\ldots,4$. We denote the $r$-th copy of $\Phi$ by $\Phi_r$, and define functions $I_r$, $J_{ri}$ on $\Phi_r$ on the pattern of (\[IJidef\]), that is, just by adding $r$ subscripts to all the variables in those equations. Naturally these can also be viewed as functions on $\Phi_{4j}$. The vector $\Jvec_r$ is the $r$-th classical angular momentum. We generalize the diagram (\[hopfmaps\]) to the $4j$-model as follows, $$\begin{diagram} & & (I_r=J_r) & \subset & \Phi_{4j} \\ & & \dTo>{\pi_H} & & \dTo>\pi \\ {\rm CL} & \subset & \Sigma_{4j} & \subset & \Lambda_{4j} \\ \dTo>{\pi_{KM}} & & & & \\ \Gamma & & & & \\ \end{diagram} \label{4jhopfmaps}$$ where $\pi$ means dividing $\Phi_{4j}$ by the $U(1)^4$ action generated by $I_r$, $r=1,\ldots,4$. Thus $\Jvec_r$, $r=1,\ldots,4$ are coordinates on $\Lambda_{4j}=(\Reals^3)^4$, in which the Poisson bracket of two functions $f$ and $g$ is given by $$\{f,g\} = \sum_{r=1}^4 \Jvec_r\cdot\left( \frac{\partial f}{\partial \Jvec_r} \times \frac{\partial g}{\partial \Jvec_r}\right). \label{Lambda4jPB}$$ Similarly, for four positive contour values $J_r>0$, $r=1,\ldots,4$, the level set $I_r = |\Jvec_r| = J_r$, $r=1,\ldots,4$, in $\Phi_{4j}$ is $(S^3)^4$ (indicated simply by $I_r=J_r$ in the diagram). The map $\pi$ restricted to this space is a power of the Hopf map (simply denoted $\pi_H$ in the diagram), which projects $(S^3)^4$ onto the space $\Sigma_{4j} = (S^2)^4$, a symplectic manifold in $\Lambda_{4j}$. Notice that the radii of the spheres (the 3-spheres in $\Phi_{4j}$ and the 2-spheres in $\Lambda_{4j}$) need not be equal for different $r$. The spaces CL and $\Gamma$ in (\[4jhopfmaps\]) will be explained later. Other important classical observables on $\Phi_{4j}$ are $\Jvec_{12} =\Jvec_1 +\Jvec_2$ and $\Jvec_{23} =\Jvec_2 +\Jvec_3$, their squares, $\Jvec_{12}^2$ and $\Jvec_{23}^2$, and the total angular momentum $\Jvec_{\rm tot} = \sum_{r=1}^4 \Jvec_r$. Level Sets and Contour Values {#levelsets} ----------------------------- The $A$- and $B$-lists of operators in (\[ABoplists\]) correspond to lists of classical observables (without the hats), $$\eqalign{ A &= (I_1,I_2,I_3,I_4,\Jvec_{12}^2,\Jvec_{\rm tot}), \\ B &= (I_1,I_2,I_3,I_4,\Jvec_{23}^2,\Jvec_{\rm tot}).} \label{ABfunctionlists}$$ We will denote the members of these lists by $A_i$, $B_i$, $i=1,\ldots,8$. As discussed in I and in Littlejohn (1990), the Lagrangian manifolds that support the semiclassical approximations to the states $\ket{A}$ and $\ket{B}$ are the level sets in $\Phi_{4j}$ of these lists of classical observables, with quantized values of the contour values. We will defer the question of quantization to Sec. \[BSquant\], and for now just examine these level sets for some suitable contour values. We will call the level sets of the $A$- and $B$-lists the $A$- and $B$-manifolds. It will be convenient to distinguish notationally the functions in the two lists, regarded as maps $:\Phi_{4j} \to \Reals$, from the contour values, which are real numbers. Our notation is summarized in Table \[fvtable\]. The vector of functions $\Jvec_{\rm tot}$ is given the value $\zerovec$ because that is the only contour value we will consider. The other contour values are variable. The conventions in the table solve some notational problems in I, but the notation requires care. For example, the magnitude of the vector $\Jvec_r$, regarded as a function $:\Phi_{4j}\to\Reals$, is not $J_r$ (because $J_r$ is a number, not a function), but on the level set $I_r=J_r$ it is true that that $|\Jvec_r| = J_r$, in view of the identity between functions, $\Jvec_r^2 = I_r^2$. If we wish to refer to the magnitude or the square of the vector $\Jvec_r$, regarded as a function on $\Phi$, we will write $|\Jvec_r|$ or $\Jvec_r^2$ (in bold face), not $J_r$ or $J_r^2$. We will denote the lists of contour values collectively by $a$ or $b$, so that $$\eqalign{ a&=(J_1,J_2,J_3,J_4,J_{12}^2,\zerovec), \\ b&=(J_1,J_2,J_3,J_4,J_{23}^2,\zerovec),} \label{abdefs}$$ with components $a_i$, $b_i$, $i=1,\ldots,8$. function value ------------------- ------------ $I_r$ $J_r$ $\Jvec_{12}^2$ $J_{12}^2$ $\Jvec_{23}^2$ $J_{23}^2$ $\Jvec_{\rm tot}$ $\zerovec$ : \[fvtable\] Notation for functions $:\Phi_{4j} \to \Reals$ and values (real numbers) of those functions. In the first row, $r=1,\ldots,4$. Consider now the conditions on the contour values $a$, $b$ such that the $A$- and $B$-manifolds should exist (as nonempty sets of points). The question of the dimensionality of these manifolds will be postponed to Sec. \[ABproperties\], but it turns out that their maximum (for any contour values) and generic dimensionality is 8, one half the dimension of the $4j$ Schwinger phase space. We work with the $A$-manifold, since the conditions for the $B$-manifold are completely analogous. First, the $A$-manifold clearly does not exist unless $J_r\ge0$, $r=1,\ldots,4$, so let us assume this condition. Then if the $A$-manifold exists, we can pick a point on it and evaluate the vector functions $\Jvec_r$, $r=1,\ldots,4$, to determine the projected point in $\Lambda_{4j}$. We visualize this point as four vectors in a single copy of $\Reals^3$. In addition we compute the vector $\Jvec_{12} =\Jvec_1 +\Jvec_2$ and plot it along with the others in $\Reals^3$. Next, we move the vectors $(\Jvec_1,\Jvec_2,-\Jvec_{12})$ end-to-end to create a triangle with sides $(J_1,J_2,J_{12})$. We move the vectors by parallel transport in $\Reals^3$, that is, without rotating them. Thus, the triangle inequalities are satisfied in the triplet of lengths $(J_1,J_2,J_{12})$. Also, since $\Jvec_{\rm tot}=0$, we can form a second triangle out of vectors $(\Jvec_3,\Jvec_4,\Jvec_{12})$, so the triangle inequalities are satisfied in the triplet $(J_3,J_4,J_{12})$. The two triangles have the edge $J_{12}$ in common, creating a figure like Fig. \[fourJ\]. Conversely, suppose that $J_r\ge0$, $r=1,\ldots,4$ and that the triangle inequalities are satisfied in triplets $(J_1,J_2,J_{12})$ and $(J_3,J_4,J_{12})$. This means that two triangles can be constructed in $\Reals^3$ with the given lengths. By translating and/or rotating the two triangles, we can bring the 12 edges into coincidence, thereby creating a figure like Fig. \[fourJ\], and hence a point of $\Lambda_{4j}$. But since the map $\pi:\Phi_{4j} \to \Lambda_{4j}$ is onto, there exists an inverse image of this point in $\Phi_{4j}$, hence the $A$-manifold exists. In summary, we have shown that the $A$-manifold exists iff $J_r\ge0$, $r=1,\ldots,4$, and the triangle inequalities are satisfied in the triplets $(J_1,J_2,J_{12})$ and $(J_3,J_4,J_{12})$. A similar result holds for the $B$-manifold, where the triplets are $(J_2,J_3,J_{23})$ and $(J_1,J_4,J_{23})$. The triangle inequalities imply that $J_{12}$ and $J_{23}$ lie in the ranges, $$\begin{aligned} J_{12,{\rm min}} &\le J_{12} \le J_{12,{\rm max}}, \label{J12range} \\ J_{23,{\rm min}} &\le J_{23} \le J_{23,{\rm max}}, \label{J23range} \end{aligned}$$ where $$\begin{aligned} \fl J_{12,{\rm min}} &= \max(|J_1-J_2|,|J_3-J_4|), \qquad J_{12,{\rm max}} = \min(J_1+J_2,J_3+J_4), \label{J12maxmin} \\ \fl J_{23,{\rm min}} &= \max(|J_2-J_3|,|J_1-J_4|), \qquad J_{23,{\rm max}} = \min(J_2+J_3,J_1+J_4). \label{J23maxmin} \end{aligned}$$ These inequalities imply $J_{12}, J_{23 }\ge 0$. Notice that the value 0 is not always excluded, for example, if $J_1=J_2$ and $J_3=J_4$ then $J_{12}=0$ is allowed. It turns out that the $A$- and $B$-manifolds, when 8-dimensional, are Lagrangian. According to the Liouville-Arnold theorem (Arnold 1989) the compact level sets of complete sets of Poisson commuting observables are generically Lagrangian tori. The sets of observables $A$ or $B$ of interest in this paper are not commuting, however, so the Liouville-Arnold theorem does not apply. In a similar situation in I we showed that the manifold in question (what we called the “Wigner manifold”) was nevertheless Lagrangian. In the following section we present a general set of circumstances in which Lagrangian manifolds are obtained, which cover not only the cases considered in the Liouville-Arnold theorem but also all cases we know of in the asymptotics of spin networks, including the manifolds studied in I and the $A$- and $B$-manifolds of this paper. Level Sets, Orbits, and Lagrangian Manifolds {#lsolm} -------------------------------------------- We use general notation in this section that differs somewhat from that of the application to the $6j$-symbol in the rest of the paper. The basic conclusion of this section is the following. Let $\{A_i, i=1,\ldots,m\}$ be a collection of classical observables on a phase space (symplectic manifold) $P$ of dimension $2N$, that is, $A_i: P\to\Reals$. Let the set $\{A_i\}$ form a Lie algebra, that is, the Poisson brackets $\{A_i,A_j\}$ are linear combinations of the $A_i$. Let $L$ be the level set $A_i=a_i$, for some contour values $a_i$, and suppose that all Poisson brackets $\{A_i,A_j\}$ vanish on $L$. Finally, suppose $L$ is a smooth manifold of dimension $N$ (thus, $m\ge N$). Then $L$ is Lagrangian. The reader who is willing to accept this conclusion can skip the remainder of this section. We deal with classical Hamiltonian systems with symmetry. Basic references on this subject are Abraham and Marsden (1978), Marsden and Ratiu (1999) and Cushman and Bates (1997). We make quite a few assumptions in this section, but most of them are generic. The most important one that is not is the assumption that the momentum, in the general sense of the value of the momentum map, is a fixed point of the coadjoint action of the group, that is, it is $G$-invariant. Let $P$ be a symplectic manifold of dimension $2N$, let $G$ be a connected Lie group of dimension $m$ with Lie algebra ${\Liealgebra{g}}$, dual ${\Liealgebra{g}}^*$, and symplectic action on $P$. Suppose the momentum map $M:P \to {\Liealgebra{g}}^*$ exists, that is, the action of $G$ is generated by Hamiltonian flows of a set of Hamiltonian functions. Let $\{\xi_i, i=1,\ldots,m\}$ be a basis in ${\Liealgebra{g}}$, and let $c^k{}_{ij}$ be the structure constants, so that $[\xi_i,\xi_j] = c^k{}_{ij}\, \xi_k$ (summation convention). Define functions $A_i:P \to {\mathbb R}$ by $A_i(x) = \langle M(x), \xi_i \rangle$ for all $x\in P$. These functions form a Lie algebra under the Poisson bracket, $\{A_i,A_j\} = c^k{}_{ij} \, A_k$, with the same structure constants as the $\xi_i$ in ${\Liealgebra{g}}$. Let $X_i = \omega^{-1} dA_i$ be the Hamiltonian vector fields associated with the $A_i$, where $\omega$ is the symplectic form on $P$, regarded as a map from vector fields to 1-forms. For some point $x_0 \in P$ let $a=M(x_0)$, so the level set $L$ of the $A$’s passing through $x_0$ is given by $A=a$ (that is, $L=M^{-1}(a)$). For simplicity we assume that $L$ has only one connected component, or else we restrict consideration to the connected component passing through $x_0$. Let $n$ be the rank of the set of differential forms $\{dA_i, i=1,\ldots,m\}$, assumed to be constant over $L$. Then $n\le m$ and $\dim L = 2N-n$. Also let the orbit of the $G$-action through $x_0$ be $B$. The vectors $\{X_i, i=1,\ldots,m\}$ are tangent to $B$ and span its tangent space at each point. Also, $\rank \{X_i\} = \rank\{dA_i\} = n$, since $\omega$ is nonsingular. Thus, $\dim B=n$. At this point we have the basic geometry for symplectic reduction, as illustrated in Fig. \[symred\]. Shown in the figure is the intersection $I$ of $L$ and $B$, which is the orbit of $x_0$ under the isotropy subgroup of $a$ under the coadjoint action of the group. Dividing $L$ by the isotropy subgroup produces the reduced symplectic manifold. Now suppose that $a \in {\Liealgebra{g}}^*$ (the generalized momentum) is a fixed point of the coadjoint action of the group, that is, $\Ad^*_g a =a$, $\forall g\in G$. By differentiation and contraction with an arbitrary element of ${\Liealgebra{g}}$ this implies $\langle a, [\eta,\zeta]\rangle =0$, $\forall \, \eta,\zeta \in {\Liealgebra{g}}$, or, by setting $\eta=\xi_i$, $\zeta=\xi_j$, $c^k{}_{ij} a_k =0$. But this implies $\{ A_i, A_j \}=c^k{}_{ij} \, A_k =0$ on the level set $L$, where $A_k=a_k$. In other words, the functions $A_i$, which may form a nontrivial Lie algebra on $P$, have vanishing Poisson brackets among themselves on $L$. This in turn implies that the $A_i$’s are constant along each other’s flows on $B$, that is, $X_i A_j = -\{A_j,A_i\} =0$, so $B \subset L$. Therefore $\dim B \le \dim L$, or $n \le 2N-n$, or $n\le N$. Finally, let us assume that $n$ takes its maximum value $n=N$. Then $\dim B = \dim L = N$. If $L$ is compact, then $B=L$, that is, the level set and the orbit coincide. Moreover $L$ is Lagrangian, since the vectors $X_i$ span the tangent space to $L$ and $\omega(X_i,X_j) = -\{A_i,A_j\}=0$ on $L$. This in turn implies that $L$ supports locally a solution of the simultaneous Hamilton-Jacobi equations involved in finding a semiclassical eigenfunction of the operators with principal symbols $A_i$. In the case of an integrable system the group is Abelian, $G={\mathbb R}^N$, the $N$ classical observables $A_i$ Poisson commute everywhere in phase space, and on a generic level set $A=a$ the differentials $dA_i$ are linearly independent everywhere. Since the group is Abelian, $a$ is automatically a fixed point of the coadjoint action, and thus the level set is Lagrangian. Moreover, if it is compact, it is an $N$-torus. In the case of the manifolds explored in I (the $jm$- and Wigner manifolds) and the $A$- and $B$-manifolds of this paper, the isotropy subgroup of the action of $G$ on $x_0\in P$ is zero-dimensional (a discrete set of points), so $\dim G = \dim B = m = n$. Moreover, $n=N$, so the manifolds are Lagrangian. Properties of the $A$- and $B$-manifolds {#ABproperties} ---------------------------------------- We now apply the analysis of Sec. \[lsolm\] to the $A$-manifold. In the following we visualize a point of $\Phi_{4j}$ as a $2\times4$ matrix of complex coordinates $z_{\mu r}$, referring to the $r$-th column as the $r$-th “spinor.” Here the symplectic manifold is $P=\Phi_{4j}$, and we take the group to be $G= U(1)^5\times SU(2)$, where the action of $U(1)^5$ is generated by $(I_1,I_2,I_3,I_4 ,\Jvec_{12}^2)$, and that of $SU(2)$ is generated by $\Jvec_{\rm tot}$. For given $r$, the observable $I_r$ generates a $U(1)$ action, the multiplication the $r$-th spinor by $e^{-i\psi_r/2}$, where $\psi_r$ is the evolution variable conjugate to $I_r$, with period $4\pi$. See (I.). Also, $\Jvec_{\rm tot}$ generates the diagonal $SU(2)$ action, that is, all four spinors are multiplied by the same element $u\in SU(2)$ (the $2\times 4$ matrix is multiplied on the left by $u$). See (I.) and (I.). As for the observable $\Jvec_{12}^2$, Hamilton’s equations are $$\fl\frac{dz_{r\mu}}{dt} = \{ z_{r\mu}, \Jvec_{12}^2\} = 2 J_{12i} \{ z_{r\mu}, J_{12i} \} = \cases{-i (\Jvec_{12} \cdot \bsigma)_{\mu\nu} \, z_{r\nu}, & if $r=1,2$, \\ 0, & if $r=3,4$,} \label{J12sqflow}$$ where $t$ is the variable of evolution. See (I.) for a similar calculation, and notice that $\Jvec_{12}$ is a constant of the flow generated by $\Jvec_{12}^2$. Thus, $\Jvec_{12}^2$ generates a $U(1)$ action that rotates spinors 1 and 2 about the axis $\jvec_{12} =\Jvec_{12} /|\Jvec_{12}|$, that is, it multiplies them by $u(\jvec_{12} ,\theta) \in SU(2)$ (in axis-angle notation for an element of $SU(2)$, see (I.)) while leaving spinors 3 and 4 invariant. The period is $\theta=4\pi$ or $t=2\pi/J_{12}$. The $U(1)$ action generated by $\Jvec_{12}^2$ is not a rotation about a fixed axis, since $\Jvec_{12}$ depends on where we are in $\Phi_{4j}$, but it does commute with the other $U(1)$ actions generated by the $I_r$, as well as the $SU(2)$ action generated by $\Jvec$. Similar statements can be made about $\Jvec_{23}^2$. Now we determine the maximum dimensionality of the $A$-manifold. Since there are 8 functions in the $A$-list and $\dim\Phi_{4j}=16$, the answer will be 8 if the functions are independent. Functions are independent at a point if their differentials, in this case, $\{dA_i, i=1,\ldots,8\}$, are linearly independent. In any case, the rank of this set of differentials is the same as the rank of the set of Hamiltonian vector fields $X_i = \omega^{-1}dA_i$. The $A$-manifold cannot have its maximum dimension if any $J_r=0$, for $r=1,\ldots,4$, since that implies $dI_r=0$. So we assume $J_r>0$, $i=1,\ldots,4$. This means that all four spinors $z_r$ are nonzero, and hence the orbit of the group $U(1)^4$ generated by the $I_r$, $r=1,\ldots,4$ is a four-torus $T^4$. This is also the fiber of the projection $\pi :\Phi_{4j }\to\Lambda_{4j}$. The condition $J_r>0$, $r=1,\ldots,4$ also means that all four vectors $\Jvec_r$, $r=1,\ldots,4$ are nonzero. We think of these vectors in a single copy of $\Reals^3$. When the group $SU(2)$, whose action is generated by $\Jvec_{\rm tot}$, acts on $\Phi_{4j}$, the effect on the vectors $\Jvec_r$, $r=1,\ldots,4$, is to rotate them by the corresponding element of $SO(3)$ (see (I.)). If any of these vectors is moved by the $SO(3)$ rotation, then in $\Phi_{4j}$ we will have moved off the initial fiber of the projection $\pi$, hence we will have motions that are linearly independent of the $U(1)^4$ action. To make all three independent directions of rotation in $SO(3)$ give rise to linearly independent motions, we require that the isotropy subgroup of the $SO(3)$ action on the set $\{\Jvec_r$, $r=1,\ldots,4\}$ (that is, the diagonal action) be trivial. This is at a point at which $\Jvec_{\rm tot}=\sum_{r=1}^4 \Jvec_r=0$. This requires that the four vectors $\Jvec_r$, $r=1,\ldots,4$ (which are nonzero) be noncollinear (otherwise the isotropy subgroup is $SO(2)$ or the whole group $SO(3)$). This in turn requires that at least one of the triangles, $(J_1,J_2,J_{12})$ or $(J_3,J_4,J_{12})$, have nonzero area. Next there is the $U(1)$-action on $\Phi_{4j}$, generated by $\Jvec_{12}^2$. This has the effect on the vectors $\Jvec_r$, $r=1,\ldots,4$ of a rotation of vectors $\Jvec_1$ and $\Jvec_2$ about axis $\jvec_{12}$, while leaving $\Jvec_3$ and $\Jvec_4$ invariant. In order that this motion move us off the initial fiber, we require that $\Jvec_1$ and $\Jvec_2$ be linearly independent, that is, that triangle $(J_1,J_2,J_{12})$ have nonzero area. And in order that this motion be linearly independent of overall rotation generated by $\Jvec_{\rm tot}$, we require that the triangle $(J_3,J_4,J_{12})$ also have nonzero area, in order that the rotation of $\Jvec_1$, $\Jvec_2$ about the axis $\jvec_{12}$ should change the shape of the figure created by the four vectors. (The overall rotations generated by $\Jvec_{\rm tot}$ do not change the shape of the figure.) We conclude that at a point of the $A$-manifold, $\rank\{dA_i, i=1,\ldots, 8\} = 8$ iff $J_r>0$, $r=1,\ldots,4$ and none of the triangle inequalities in $(J_1,J_2,J_{12})$ or $(J_3,J_4,J_{12})$, or equivalently, the inequalities in (\[J12range\]), is saturated. Notice in particular that this implies $J_{12}>0$. Since these conditions depend only on the contour values and not where we are on the $A$-manifold, the $A$-manifold, under the stated conditions, is a smooth, 8-dimensional manifold. In addition, the $A$-manifold is compact. It is also connected, as follows by consideration of its projection under $\pi$ (see Sec. \[project\]). This means that the action of the group $U(1)^5 \times SU(2)$ is transitive on the $A$-manifold (it is the orbit of any point on it under the group action). It then follows from the discussion of Sec. \[lsolm\] that the $A$-manifold, when 8-dimensional, is Lagrangian. Similar statements apply to the $B$-manifold. To find the topology of the $A$-manifold, we first find the isotropy subgroup of the action of $U(1)^5\times SU(2)$ on a point $x_0$ on this manifold. If we denote coordinates on $U(1)^5 \times SU(2)$ by $(\psi_1, \psi_2, \psi_3, \psi_4, \theta, u)$, where $u \in SU(2)$ and where the five angles are the $4\pi$-periodic evolution variables corresponding to $(I_1,I_2,I_3,I_4,\Jvec_{12}^2)$, respectively, then the isotropy subgroup is generated by two elements, say, $x=(2\pi, 2\pi, 2\pi, 2\pi, 0, -1)$ and $y=(0,0,2\pi,2\pi,2\pi,-1)$. The isotropy subgroup itself is the Viergruppe $\{e,x,y,xy\}=({\mathbb Z}_2)^2$. Thus the $A$-manifold is topologically $U(1)^5 \times SU(2) / ({\mathbb Z}_2)^2$. This is the same logic used in I to find the topology of the “Wigner manifold” of that paper. The analysis is the same for the $B$-manifold, which has the same topology. The isotropy subgroup would be larger in degenerate cases, for example, when some triangle inequalities are saturated. Now it is easy to find the invariant measure on the $A$- or $B$-manifolds. It is $d\psi_1\wedge d\psi_2\wedge d\psi_3\wedge d\psi_4\wedge d\theta\wedge du$, where $du$ is the Haar measure on $SU(2)$ ($du =\sin\beta\, d\alpha\wedge d\beta\wedge d\gamma$ in Euler angles). Thus, the volume of the $A$- or $B$-manifold with respect to this measure is $$V_A = V_B = \frac{1}{4} (4\pi)^5 \times 16\pi^2 = 2^{12} \pi^7, \label{ABvolume}$$ where the $1/4$ compensates for the 4-element isotropy subgroup. Projections and tetrahedra {#project} -------------------------- We now study the projection of the $A$-manifold onto $\Lambda_{4j}$, for contour values such that the manifold is 8-dimensional. The projection consists of the set of four nonzero vectors in $\Reals^3$, $\{\Jvec_r, r=1,\ldots, 4\}$ of given lengths $J_r>0$ with a sum of zero, $\Jvec_{\rm tot}=0$, creating a closed chain of links as in Fig. \[fourJ\], such that $J_{12}$ has a given, positive value and both triangles 1-2-12 and 3-4-12 have nonzero area. We choose to place the four vectors end-to-end in the order $(\Jvec_1 ,\Jvec_2 ,\Jvec_3 ,\Jvec_4)$, as in Fig. \[fourJ\]; this is an arbitrary choice, but convenient for studying observables $\Jvec_{12}^2$ and $\Jvec_{23}^2$ (if we wished to examine $\Jvec_{13}^2$ we would choose a different order). By filling in the lines $\Jvec_{12}$ and $\Jvec_{23}$, the closed chain of links becomes a tetrahedron, as shown in Fig. \[butterfly\]. The set of all such figures, modulo orientation, is the circle $S^1=SO(2)$, whose angle is the dihedral angle about the edge $\Jvec_{12}$. Including the orientation, we see that the projection of the $A$-manifold onto $\Lambda_{4j}$ is topologically $SO(2) \times SO(3)$. One might imagine that this should be described as an $SO(3)$ bundle over $SO(2)$, but in fact the bundle is trivial. This is connected, so the inverse image under $\pi$, which is the $A$-manifold itself, is a 4-torus bundle over $SO(2)\times SO(3)$, and is also connected. An action of the group $SO(2)\times SO(3)$ on $\Lambda_{4j}$ is generated by $\Jvec_{12}^2$ and $\Jvec_{\rm tot}$, regarded as observables on $\Lambda_{4j}$ using the Poisson structure (\[Lambda4jPB\]). This is the projection of the action of $U(1)^5 \times SU(2)$ on $\Phi_{4j}$. The motion generated by $\Jvec_{12}^2$ is the “butterfly” motion illustrated in Fig. \[butterfly\], in which the butterfly flaps one of its wings. As the triangle 1-2-12 rotates around the edge 12, the length $|\Jvec_{23}|$ varies between a maximum and minimum value. The minimum value is reached when triangles 1-2-12 and 3-4-12 lie in the same plane on the same side of line 12, and maximum when on opposite sides of line 12; these are also the minimum and maximum of $|\Jvec_{23}|$ on the $A$-manifold in $\Phi_{4j}$. These extremal values are reached when the tetrahedron is flat (its volume is zero), and can be obtained in terms of the other five $J$’s as the roots of the Cayley-Menger determinant (Berger 1987, Crippen and Havel 1988), $$\det \left(\begin{array}{ccccc} 0 & 1 & 1 & 1 & 1 \\ 1 & 0 & J_1^2 & J_{12}^2 & J_4^2 \\ 1 & J_1^2 & 0 & J_2^2 & x^2 \\ 1 & J_{12}^2 & J_2^2 & 0 & J_3^2 \\ 1 & J_4^2 & x^2 & J_3^2 & 0 \end{array}\right)=0, \label{CMdet}$$ where $x=|\Jvec_{23}|$. Such determinants were an important part of the analysis of Ponzano and Regge (1968). In this case the determinant expresses the condition that the volume of the tetrahedron is zero. Expanding the determinant gives a quadratic equation in $x^2$ in terms of the five $J$’s specifying the $A$-manifold. The same condition can be expressed in terms of a smaller ($3\times3$) Gram matrix of dot products, as discussed in Littlejohn and Yu (2009). A similar analysis applies to the $B$-manifold except that here we fix the length $J_{23}$ of vector $\Jvec_{23}$, and create tetrahedra of different shapes by varying the dihedral angle between triangles 4-1-23 and 2-3-23, that is, by rotating vectors $\Jvec_2$ and $\Jvec_3$ about the axis $\jvec_{23} = \Jvec_{23}/J_{23}$. In this process, the length $|\Jvec_{12}|$ varies from some minimum to some maximum, which can be obtained by replacing $x^2$ in (\[CMdet\]) by $J_{23}^2$, and then $J_{12}^2$ by $x^2$ where now $x=|\Jvec_{12}|$. Quantizing the Manifolds {#BSquant} ------------------------ The $A$- and $B$-manifolds can be subjected to Bohr-Sommerfeld quantization, which selects certain contour values as quantized. The process uses the Weyl symbols or transforms (Weyl 1927, Wigner 1932, Groenewold 1946, Moyal 1949, Berry 1977, Balazs and Jennings 1984, Littlejohn 1986, Ozorio de Almeida 1998) of the lists (\[ABoplists\]) of operators, which are functions on $\Phi_{4j}$. The Weyl symbols of selected operators are summarized in Table \[weylsymbols\]. It is important that the classical manifolds be the level sets of the Weyl symbols of the operators whose simultaneous eigenfunctions we seek (Littlejohn 1990); this holds in the present case because the Weyl symbols in the table are always equal to the corresponding classical observable (without the hat), to within an additive constant. Notationally we could have defined the classical observable (without the hat) as the Weyl symbol of the corresponding quantum observable, but the conventions in the table make it easier to establish the connections with the usual quantum numbers in physics. In particular, the zero point energy has been subtracted from the quantum observables $\Ihat_r$, but not from the classical ones $I_r$, which explains the $1/2$ on the first row of the table. operator Weyl symbol ---------------------- ---------------------- $\Ihat_r$ $I_r-1/2$ $\Jvechat_r$ $\Jvec_r$ $\Jvechat_{\rm tot}$ $\Jvec_{\rm tot}$ $\Jvechat_{12}$ $\Jvec_{12}$ $\Jvechat_{23}$ $\Jvec_{23}$ $\Jvechat_r^2$ $\Jvec_r^2 - 3/8$ $\Jvechat_{12}^2$ $\Jvec_{12}^2 - 3/4$ $\Jvechat_{23}^2$ $\Jvec_{23}^2 - 3/4$ : \[weylsymbols\] Weyl symbols of selected operators. In rows containing operators $\Ihat_r$, $\Jvechat_r$ and $\Jvechat_r^2$, $r=1,\ldots,4$. To quantize the $A$- or $B$-manifold, we first determine the homotopy group, then we compute action integrals and Maslov indices along generators of the group, then we require that the action plus Maslov correction be an integer multiple of $2\pi$. Only manifolds of full dimensionality (8) can be quantized. This is the procedure followed in I, and the analysis is very similar in this case; in particular, in both cases the homotopy group is Abelian. We just summarize the results, speaking of the $A$-manifold. We find that $J_r$, $r=1,\ldots,4$ are quantized in half-integer steps, which we write in terms of the quantum numbers $j_r$ as $$J_r=j_r+1/2, \label{Jrquant}$$ where the allowed values of $j_r$ are $0,\frac{1}{2},1,\ldots$. Smaller values of $j_r$ are not allowed because for $j_r=-1/2$ the manifolds do not have full dimensionality, while for $j_r<-1/2$ they do not exist. We choose the quantum number $j_r$ so that it agrees with the usual notation in physics for the eigenvalues of various operators, but that is not confirmed until we compute the semiclassical eigenvalues in Sec. \[sceigenvalues\]. Similarly, $J_{12}$ must be an integer or half-integer on quantized manifolds, which we write in terms of a conventional quantum number by $J_{12}=j_{12}+\frac{1}{2}$. In addition, there is the condition that $j_1+j_2+j_{12}$ and $j_3+j_4+j_{12}$ be integers, the requirement (with the usual interpretation of the quantum numbers) that the $3j$-symbols in Fig. \[asym6j\] should exist. See also eq. (I.). These imply that $j_1+j_2+j_3+j_4$ must be an integer, part of the conditions that the subspace $\ZS$ defined in Sec. \[4jmodel\] be nontrivial. The range of the quantum number $j_{12}$ is determined by the requirement that the $A$-manifold be 8-dimensional. Looking first at the upper limit, if $J_{12}$ is quantized we have $$J_{12} = j_{12}+\frac{1}{2} < J_{12,{\rm max}} =\min(j_1+j_2+1,j_3+j_4+1). \label{j12upper}$$ Given the other conditions on $j_{12}$, this implies that the maximum quantized value of $J_{12}$ is $J_{12,{\rm max}}-\frac{1}{2}$. Similarly, we find that the minimum quantized value of $J_{12}$ is $J_{12,{\rm min}}+\frac{1}{2}$. The quantized values of $J_{12}$ are separated from the maximum and minimum classical values (for given $J_r$, $r=1,\ldots,4$) by a margin of $\frac{1}{2}$, and are spaced in integer steps. These rules imply $$j_{12,{\rm max}}=J_{12,{\rm max}}-1, \qquad j_{12,{\rm min}}=J_{12,{\rm min}}. \label{J12j12maxmin}$$ They also imply the bounds (\[j12j23bounds\]) on the quantum number $j_{12}$. Similar results apply to the $B$-manifold and the quantized values of $J_{23}$. Figure \[square\] is a numerical example of these quantization rules. The square in the figure is given by the bounds (\[J12range\]) and (\[J23range\]), while the spots are the quantized values of $J_{12}$ and $J_{23}$. The latter are related to the usual quantum numbers by $$J_{12}=j_{12}+\frac{1}{2}, \qquad J_{23}=j_{23}+\frac{1}{2}. \label{J12J23quant}$$ The spots form a square array because $\braket{j_{23}}{j_{12}}$ is a square matrix. The size of the matrix (the number of rows or columns of spots) is $\dim\ZS$, given by (\[dimZS\]). Other features of this figure will be explained later. Note that the minimum value of $J_r$, $r=1,\ldots,4$ and of $J_{12}$ on any quantized $A$-manifold is $\frac{1}{2}$, so the corresponding vectors always have a positive length. In addition, the triangles 1-2-12 and 3-4-12 always have a positive area. Similar conclusions apply to the quantized $B$-manifolds. Semiclassical Eigenvalues {#sceigenvalues} ------------------------- Once the classical manifolds are quantized, we find the semiclassical approximations to the eigenvalues of the operators in the $A$- or $B$-lists by evaluating the Weyl symbols of those operators on the classical manifold. Doing this for the operators $\Ihat_r$, $r=1,\ldots,4$, we find the eigenvalue of $\Ihat_r$ is $j_r$; this is the exact answer, something that was to be expected since $\Ihat_r$ is a quadratic polynomial in the fundamental $\xhat$’s and $\phat$’s of the system, and Weyl quantization of such operators is exact. See Littlejohn (1986) for more on classical and quantum quadratic polynomials. If we now compute the eigenvalue of $\Jvechat_r^2$ by using the operator identity $\Jvechat_r^2 = \Ihat_r(\Ihat_r+1)$, naturally we get the exact answer, $j_r(j_r+1)$. On the other hand, if we evaluate the Weyl symbol of the operator $\Jvechat_r^2$ on the quantized level set, we get (according to the table) the eigenvalue $(j_r+1/2)^2-3/8 = j_r(j_r+1) -1/8$, with an error of $-1/8$. There is an error because $\Jvechat_r^2$ is a quartic polynomial in the fundamental $\xhat$’s and $\phat$’s of the system, so Weyl quantization is not exact. The error is, however, of relative order $\hbar^2$, that is, $1/j^2$, which is the error expected in lowest order semiclassical approximations. The operators $\Jvechat_r^2$ do not appear in the $A$- or $B$-lists, but operators $\Jvechat_{12}^2$ and $\Jvechat_{23}^2$ do, and here again there is an error at order $1/j^2$. Moreover, in this case there are no operators $\Ihat_{12}$ or $\Ihat_{23}$ which could be used to obtain the exact eigenvalues. This is a drawback of the $4j$-model in comparison to the $8j$- or $12j$-models, where such operators exist. It does not, however, change any of the subsequent analysis, which depends only on using the quantized manifolds to carry out the stationary phase calculation. Time Reversal {#trevclass} ------------- The action of the antilinear time reversal operator $\Thetahat$ on a carrier space $\CS_j$ can be defined by $$\Thetahat\ket{jm} = (-1)^{j-m} \ket{j,-m} \label{Thetahatdef}$$ (Messiah 1966). This is equivalent to $\Thetahat=K_1^{-1} \circ G$, where $G$ is the antilinear metric or map of Hermitian conjugation, and $K_1^{-1}$ is defined in Sec. \[brastokets\]. The composition of the antilinear $G$ with the linear $K_1^{-1}$ is the antilinear time reversal map $\Thetahat$. The map $\Thetahat$ is easily extended to the full Schwinger Hilbert space $\SS$ and tensor products thereof such as $\HS_{4j}$. Classically the antilinear $\Thetahat:\HS_{4j}\to\HS_{4j}$ corresponds to an antisymplectic map $\Theta :\Phi_{4j }\to\Phi_{4j}$. In the complex coordinates, its action on all four spinors is given by $$\Theta \left(\begin{array}{c} z_{1r} \\ z_{2r} \end{array}\right) =\left(\begin{array}{c} -\zbar_{2r} \\ \zbar_{1r} \end{array}\right), \label{classThetadef}$$ that is, $\Theta:z_r \mapsto v \zbar_r$, where $z_r\in\Complexes^2$ and where $$v=\exp(-i\sigma_2\pi/2)=\left(\begin{array}{cc} 0 & -1 \cr 1 & 0 \end{array}\right). \label{vdef}$$ We recall that in quantum mechanics, it is time reversal, not parity, that reverses the direction of angular momenta. At the classical level, this means that $\Jvec_r\bigl (\Theta(x )\bigr)= -\Jvec_r(x)$, where $x\in\Phi_{4j}$. Time reversal can be projected via $\pi$ onto $\Lambda_{4j}$, where its effect on the coordinates is $\Theta:\Jvec_r \mapsto -\Jvec_r$. It is an anti-Poisson map on $\Lambda_{4j}$. Intersections and Actions {#intact} ========================= In this section we consider the intersections of the $A$- and $B$-manifolds, assuming that $J_r>0$, $r=1,\ldots,4$ are given. For now we treat this as a classical problem in which the $J$’s (including $J_{12}$ and $J_{23}$) are continuous variables, but we note that if $J_r$, $r=1,\ldots,4$ are quantized then they are automatically positive. The motivation, however, is to find the stationary phase points of the scalar product (\[6jme\]), which are the intersections of the quantized manifolds. Classically allowed and forbidden regions {#classallowforbid} ----------------------------------------- A simple analogy will help to understand the results. Consider a one-dimensional harmonic oscillator, $H=(1/2)(x^2+p^2)$ (classical or quantum). For a given value of the energy $E$, the classically allowed region is the interval of the $x$-axis between the turning points, given by $x=\pm\sqrt{2E}$, while the classically forbidden region is outside this interval. The classically allowed region can also be defined as the region of the $x$-axis where the two curves in phase space, $x={\rm const}$ and $H=E$, have intersections. These curves are level sets of the observables $x$ and $H$, which appear on the two sides of the matrix element when we write the energy eigenfunction as $\psi_E(x) =\braket{x}{E}$. Inside the classically allowed region the intersections between the two curves consist of two points, with opposite momentum values. These are related by time reversal ($p\to -p$). In the classically forbidden region the two curves have no real intersections, but if we complexify phase space and the curves $x={\rm const}$ and $H=E$ (maintaining real contour values $x$ and $E$), then they do have complex intersections that are related to the exponentially decaying wave function in the classically forbidden region. We can view the classically allowed and forbidden regions in the $x$-$E$ plane, in which both $x$ and $E$ are variables. See Fig. \[horegions\]. The darkly shaded region in the figure, labeled $UU$, is the region $E<0$, for which the level set $H=E$ does not exist. Above the line $E=0$, both level sets, $x={\rm const}$ and $H=E$ exist. The unshaded region, labeled $AA$, is the classically allowed region, while the lightly shaded region, labeled $FF$, is the classically forbidden region. The quantized values of the energy ($E=n+\frac{1}{2}$) are indicated as spots on the $E$-axis. There is only a one-dimensional array of spots because the observable $x$ has a continuous spectrum. The parabola $E=x^2/2$ is the caustic curve, separating the classically allowed from the classically forbidden region. Intersections of the $A$- and $B$-manifolds {#ABintersect} ------------------------------------------- Similarly, for positive $J_r$, $r=1,\ldots,4$, either the $A$- or $B$-manifold does not exist if $J_{12}$ or $J_{23}$ lies outside the bounds (\[J12range\]) and (\[J23range\]). Those inequalities define a square region of the $J_{23}$-$J_{12}$ plane, as illustrated in Fig. \[square\]. A figure like this was first given by Neville (1971). The darkly shaded region outside the square, labeled $U$, is where either the $A$- or $B$-manifold does not exist. In the interior of the square both manifolds exist and have full dimensionality. On the boundary of the square they exist, but have less than full dimensionality, since some triangle inequality is saturated. If the $A$- and $B$-manifolds exist and intersect, then we are in the classically allowed (unshaded) region in the interior of the square. If they exist but do not intersect, then we are in the classically forbidden (lightly shaded) region. The caustic curve is the oval curve in the figure, separating the classically allowed from the classically forbidden regions. Notice that it touches the square boundary at four points. If we fix a value of $J_{12}$, we can regard the classically allowed and forbidden regions as intervals of the $J_{23}$-axis. The interval within which the $B$-manifold exists is given by (\[J23range\]); this is the interval of allowed $J_{23}$ values, given $J_r$, $r=1,\ldots,4$. The classically allowed region, on the other hand, is the interval of allowed $J_{23}$ values, given $J_r$, $r=1,\ldots,4$ and $J_{12}$ (equivalent to the statement that the manifolds intersect). Since this is a more restrictive condition, the classically allowed region must be a subset of the interval (\[J23range\]). Moreover, since the $A$-manifold is connected, the subset must be a connected interval, since this subset is the range of $J_{23}$ values that occur on a given $A$-manifold. This is just what we see in Fig. \[square\]: for most values of $J_{12}$, the interval inside the square consists of a classically allowed region, surrounded on both sides by classically forbidden regions, outside of which are the regions in which manifolds of the given $J_{23}$ values do not exist. The bounds of the classically allowed region are given by the roots of (\[CMdet\]), that is, they correspond to tetrahedra of zero volume. The caustic curve in Fig. \[square\] is the contour of $\det M=0$, where $M$ is the Cayley-Menger determinant (\[CMdet\]). The volume vanishes if the tetrahedron is flat. For most values of the parameters, this does not require that any of the triangular faces have zero area. But for certain values of $J_{12}$, the volume vanishes at a place where the face 2-3-23 has zero area (thus, $\Jvec_2$, $\Jvec_3$ and $\Jvec_{23}$ are linearly dependent). This is the point where the caustic curve touches the boundary of the square on the left or right. Similarly, the caustic curve touches the boundary of the square on the top and bottom of the square, where not only does the volume vanish, but also the area of triangle 1-2-12. The caustic curve never lies outside the bounds (\[J23range\]) defined by the triangle inequalities. Biedenharn and Louck (1981) appear to claim the contrary, but there is the question of whether one is talking about the classical or quantum triangle inequalities. That is, the caustic curve does pass outside the bounds given by the square array of quantized spots. It also seems to us that the interpretation of Biedenharn and Louck of Fig. 6 from Ponzano and Regge is incorrect. If the values of $J_r$, $r=1,\ldots,4$ are quantized, then we can plot the quantized values of $J_{12}$ and $J_{23}$ as a square array of spots, as in the figure. See other comments on this array in Sec. \[BSquant\]. In Littlejohn and Yu (2009) we incorrectly stated that the quantized values of $J_{12}$ and $J_{23}$ can fall exactly on a caustic, citing the theory of Brahmagupta quadrilaterals (Sastry 2002). That theory shows that plane quadrilaterals with integer sides and integer diagonals exist, that is, flat tetrahedra with all integer edges. However, to represent a $6j$-symbol, the sums of the integers around the faces of triangles must be odd, and this condition cannot be met. (The integers divided by 2 are the values of $J_r=j_r+\frac{1}{2}$, and the sum of the $j_r$ around the faces must be an integer.) A correct proof of the nonexistence of flat, quantized tetrahedra, credited to Adler, is given in a brief citation by Ponzano and Regge (1968). We now examine the intersections of the $A$- and $B$-manifolds in greater detail. First, the $A$- and $B$-manifolds (assumed to exist) have an intersection in $\Phi_{4j}$ iff their projections onto $\Lambda_{4j}$ intersect. Furthermore, the intersection in $\Phi_{4j}$ is the lift of the intersection of the projections in $\Lambda_{4j}$, with a $T^4$ fiber over every point. These statements follow from the fact that over every point $x\in\Lambda_{4j}$ there is a 4-torus fiber in $\Phi_{4j}$, and that if $x$ lies on the intersection of the projections of the $A$- and $B$-manifolds, then the 4-torus belongs to both the $A$- and $B$-manifolds in $\Phi_{4j}$. This is the same logic used in I under similar circumstances. To find the intersections of the projections in $\Lambda_{4j}$ we require four vectors $\Jvec_r$, $r=1,\ldots,4$ that satisfy $$\eqalign{ &|\Jvec_r| = J_r, \qquad \sum_{r=1}^4 \Jvec_r=0, \\ &|\Jvec_1+\Jvec_2| = J_{12}, \qquad |\Jvec_2+\Jvec_3| = J_{23},} \label{ABconditions}$$ for the given values of $J_r>0$, $r=1,\ldots,4$ and of $J_{12}$ and $J_{23}$. A nice way of constructing these vectors is given in Appendix A of Littlejohn and Yu (2009), which uses the singular value decomposition of the Gram matrix of dot products associated with the Cayley-Menger determinant (\[CMdet\]). This method not only gives an explicit solution for these vectors at any point in the classically allowed region, it also shows that they are unique to within the overall action of $O(3)$. It is obvious in any case that if $\Jvec_r$, $r=1,\ldots,4$ is a solution of (\[ABconditions\]), then so is $S\Jvec_r$ for any $S\in O(3)$. This method was generalized to the $9j$-symbol in Haggard and Littlejohn (2010). The group $O(3)$ is conveniently decomposed into proper rotations in $SO(3)$ and spatial inversion, which is time reversal in the present case. It is a basic fact of geometrical figures in $\Reals^3$ that spatial inversion is not equivalent to any proper rotation unless the figure is planar. Another fact is that the orbit of a geometrical figure under $SO(3)$ is diffeomorphic to $SO(3)$ itself, unless the dimension of the figure is $\le 1$. These issues are discussed by Littlejohn and Reinsch (1995, 1997) in the context of molecular configurations. They imply that except at the caustics, where the tetrahedron is flat, two tetrahedra related by time reversal are not related by any proper rotation. Therefore, except at the caustics, the solution set of (\[ABconditions\]) in $\Lambda_{4j}$ consists of two disconnected subsets, each diffeomorphic to $SO(3)$, related by time reversal. Each subset consists of tetrahedra of nonzero volume related by proper rotations. At a generic point of the caustic curve, where the tetrahedron is flat but still 2-dimensional, the two subsets merge into one, which is still diffeomorphic to $SO(3)$. The intersections in $\Phi_{4j}$ are the lifts of these intersections in $\Lambda_{4j}$. Therefore, except at the caustics, the intersection of the $A$- and $B$-manifolds consists of two disconnected subsets, related by time reversal, where each subset is a $T^4$-bundle over $SO(3)$. These subsets are 7-dimensional, so the $A$- and $B$-manifolds, which are 8-dimensional, intersect in two 7-dimensional submanifolds. The situation can be visualized as in Fig. \[ABintsect\], where $A \cap B = I_1 \cup I_2$ and where $I_1$ and $I_2$ are the connected intersection sets, related by $\Theta$ (see (\[classThetadef\])). Each intersection set is an the orbit of the group $U(1)^4 \times SU(2)$, where $U(1)^4$ represents the phases of the four spinors and $SU(2)$ is the diagonal action (thus, the group is generated by $I_r$, $r=1 ,\ldots,4$ and $\Jvec_{\rm tot}$). The isotropy subgroup of this group is ${\mathbb Z}_2$, generated by element $(2\pi,2\pi, 2\pi,2\pi, -1)$, in coordinates $(\psi_1, \psi_2, \psi_3, \psi_4 ,u)$ for group $T^4\times SU(2)$, where $u\in SU(2)$. The volume of intersection manifold $I_1$ or $I_2$ with respect to the measure $d\psi_1 \wedge d\psi_2 \wedge d\psi_3 \wedge d\psi_4 \wedge du$ is $$V_I = \frac{1}{2} (4\pi)^4 \times 16\pi^2 = 2^{11} \pi^6. \label{volumeI}$$ An interesting aspect of the method of Appendix A of Littlejohn and Yu (2009) for finding the vectors $\Jvec_r$, $r=1,\ldots,4$ is that it also works in the classically forbidden region, where it produces complex 3-vectors that satisfy (\[ABconditions\]). These solutions are determined modulo the action of spatial inversion and $SO(3,\Complexes)$. In fact, as discussed in that reference, the vectors can be chosen so that two components are real and one purely imaginary, so that the symmetry group of the solution set is the Lorentz group $SO(2,1) \subset SO(3,\Complexes)$. Roberts (1999) and others have referred to the tetrahedra in the classically forbidden region as living in $\Reals^3$ with a Minkowski metric, while those in the classically allowed region live in $\Reals^3$ with a Euclidean metric. This is a correct interpretation of the situation for the $6j$-symbol, but we do not think it is appropriate for generalizations to other spin networks. For example, in the $9j$-symbol (Haggard and Littlejohn 2010) the vectors in the classically forbidden region cannot be chosen so that two components are real, instead they belong to $\Complexes^3$, and the symmetry group of the solution set is $SO(3,\Complexes)$, not some Lorentz subgroup thereof. This simply means that to explore the complex Lagrangian manifolds, complex Euler angles must be used when following the Hamiltonian flows generated by $\Jvec_{\rm tot}$. The complexified $A$- and $B$-manifolds in the $6j$-symbol support the asymptotic forms given by Ponzano and Regge (1968) in the classically forbidden region. Actions and phases on the $A$- and $B$-manifolds {#ABactions} ------------------------------------------------ Let the $x$-space wave functions associated with states $\ket{A}$ and $\ket{B}$ of (\[4jABdefs\]) be denoted $\psi_A(x) =\braket{x}{A}$ and $\psi_B(x) =\braket{x}{B}$, where $x\in\Reals^8$. The semiclassical approximations to these wave functions involve phases $e^{iS_A(x)}$ and $e^{iS_B(x)}$, where actions $S_A(x)$ and $S_B(x)$ are integrals of $p\,dx =\sum_{r\mu} p_{r\mu} \, dx_{r\mu}$ from some initial points on the two manifolds to some final point. The initial points on the $A$- and $B$-manifolds are denoted $a_0$ and $b_0$, respectively, in Fig. \[ABintsect\]; they determine the overall phases of the states $\vert A\rangle$ and $\vert B\rangle$. As explained in I, the branches of the stationary phase evaluation of the matrix element $\langle B \vert A \rangle$ are associated with the intersections of the $A$- and $B$-manifolds, in this case the manifolds $I_1$ and $I_2$, so the asymptotic form of the $6j$-symbol has two branches. Moreover, the phase associated with each branch is $S_A - S_B$ evaluated on the corresponding intersection manifold (see, for example, (I.) or (I.)), and is independent of where we evaluate it on that manifold. Figure \[ABintsect\] illustrates two points $p_1$ and $p_2$, on intersection manifolds $I_1$ and $I_2$, respectively, with paths that may be used for computing the actions $S_A$ and $S_B$. In the following we shall be interested in the relative phase between the two branches. Let us define $$S_1 = S_{A1}-S_{B1}, \qquad S_2 = S_{A2} - S_{B2}, \label{S1S2defs}$$ and $$S=S_2-S_1=\int_{a_0}^{p_2} -\int_{b_i}^{p_2} - \int_{a_0}^{p_1} + \int_{b_0}^{p_1} p\,dx = \oint p \, dx, \label{relphase}$$ where the final integral is taken along the path that goes from $p_1$ to $p_2$ along the $A$-manifold and then back to $p_1$ along the $B$-manifold. The relative phase is independent of the initial points $a_0$ and $b_0$, and is moreover a symplectic invariant. The relative phase is easier to determine than the absolute phases of either branch, which are related to the overall phase convention for the $6j$-symbol. In computing the relative phase, we note that the loop integral in (\[relphase\]) can be evaluated with respect to any symplectic 1-form, such as the complex one used in I (see (I.)), $$\oint \sum_{r\mu} p_{r\mu} \, dx_{r\mu} = \Im \oint \sum z_{r\mu} \, d\zbar_{r\mu}. \label{1forms}$$ The loop integral can be transformed by Stokes’ theorem into an integral of the symplectic form over the enclosed area, since on $\Phi_{4j}={\mathbb C}^8$ all cycles are boundaries. Closing the loop in $\Lambda_{4j}$ {#closeloopAMS} ---------------------------------- We shall construct the closed loop giving the relative phase between the branches according to (\[relphase\]) by following the Hamiltonian flows of various observables. Let us define the signed volume of the tetrahedron as $V=(1/6) \Jvec_1 \cdot(\Jvec_2 \times \Jvec_3)$, and let us take manifold $I_1$ to be the one on which $V<0$, so that $V>0$ on $I_2$. Time reversal changes the sign of the volume when mapping $I_1$ into $I_2$ and vice versa. Let us start at a point $p$ of $I_1$, as in Fig. \[ABintsect1\]. Then by following the $\Jvec_{12}^2$-flow we trace out a path that takes us along the $A$-manifold to a point $q$ of $I_2$. We cannot use any of the other seven observables defining the $A$-manifold for this purpose, namely, $(I_1,I_2,I_3,I_4,\Jvec)$, since their flows confine us to the intersection manifold $I_1$. We see that the $\Jvec_{12}^2$-flow actually does take us to manifold $I_2$ by considering the projection of the path $p$-$q$ in Fig. \[ABintsect1\] onto $\Lambda_{4j}$. We visualize the projected path as a transformation applied to a set of four vectors in ${\mathbb R}^3$ that define a tetrahedron. The situation is illustrated in Fig. \[cycle\]. In part (a) of that figure, we have four vectors $\Jvec_r$ that sum to zero, defining a tetrahedron of negative volume. The lengths $J_r>0$, $r=1,\ldots,4$, $J_{12}$ and $J_{23}$ are assumed to have the prescribed values, and vector $\Jvec_{12}$ is drawn (but not $\Jvec_{23}$). We take the point $p$ of Fig. \[ABintsect1\] to lie on the $T^4$ fiber above this tetrahedron. The $\Jvec_{12}^2$-flow rotates the 1-2-12 triangle about the 12-axis by the right-hand rule while leaving the 3-4-12 triangle fixed (see (\[J12sqflow\])), that is, it rotates the 1-2-12 triangle into the foreground. Let the angle of rotation be $2\phi_{12}$, where $\phi_{12}$ is the interior dihedral angle of the tetrahedron along edge 12 in its original configuration. This brings triangle 1-2-12 through triangle 3-4-12 to the opposite side, creating a new tetrahedron with the same lengths (the new $J_{23}$ is the same as the old one), hence the same dihedral angles, but with the opposite signed volume. The result is illustrated in part (b) of Fig. \[cycle\], a tetrahedron that is the projection of a point $q \in I_2$ in $\Phi_{4j}$, as illustrated in Fig. \[ABintsect1\]. Thus we see that the $\Jvec_{12}^2$-flow does take us from $I_1$ to $I_2$ along the $A$-manifold, as claimed. Having reached point $q \in I_2$, we can go back to $I_1$ along the $B$-manifold by following the $\Jvec_{23}^2$-flow, reaching point $p'\in I_1$ as illustrated in Fig. \[ABintsect1\]. The transformation in $\Lambda_{4j}$ is illustrated in Fig. \[cycle\]. Part (c) of that figure is the same as part (b), except that vector $\Jvec_{23}$ is drawn and $\Jvec_{12}$ is suppressed. The $\Jvec_{23}^2$-flow rotates triangle 2-3-23 about the axis $\jvec_{23} =\Jvec_{23} /J_{23}$, while leaving triangle 1-4-23 fixed. Let the angle of rotation be twice the interior dihedral angle along edge 23, that is, $2\phi_{23}$, as illustrated in part (c) of Fig. \[cycle\]. The result is part (d) of that figure, a tetrahedron in which the volume has been inverted a second time, taking us back to the original (negative) volume in part (a). We arrive at point $p' \in I_1$, as in Fig. \[ABintsect1\]. It is clear that $p'$ is not the same as the original point $p$, because if it were, the orientation of the tetrahedron in part (d) of Fig. \[cycle\] would be the same as that in part (a). Thus to create a closed loop in $\Phi_{4j}$, we must follow some path in $I_1$ taking us from $p'$ to $p$, as in Fig. \[ABintsect1\]. We create this path from $p'$ to $p$ in $I_1$ in two steps. First we apply an $SU(2)$ transformation to all four spinors at point $p'$, that is, a diagonal transformation, whose projection onto $\Lambda_{4j}$ is an $SO(3)$ transformation of the tetrahedron in part (d) of Fig. \[cycle\], returning it to the original orientation in part (a). This is a proper rotation of all four vectors $\Jvec_r$, that is, a rigid rotation of the entire tetrahedron, and it is the final step in a cycle of rotations that transform the original tetrahedron in part (a) into itself. That figures (a) and (d) must be related by some proper rotation is clear, since the lengths of the sides are the same and the signed volume is the same. In fact, it is easy to see that the axis of the final rotation is $\Jvec_4$, since this vector is left invariant by both the $\Jvec_{12}^2$- and $\Jvec_{23}^2$-flows, and is the same in parts (a) and (d). As it turns out, the final rotation has axis $-\jvec_4 = -\Jvec_4/j_4$ (notice the minus sign) and angle $2\phi_4$, twice the internal dihedral angle along edge 4. This is illustrated in part (e) of Fig. \[cycle\], which is the same as part (d) except that all vectors are drawn. The effect of the final rotation is illustrated in part (f), which is the same as part (a) except that all vectors are drawn. Angle of the final rotation {#finalangle} --------------------------- To obtain the angle of the final rotation about axis $-\jvec_4$, we use the fact that the product of two reflections is rotation. Let a reflection about a plane $P$ be $Q(P)$. There are four ways to draw the angle between the planes, one of which is denoted by $\alpha$ in Fig. \[2refl\]. For a given choice of dihedral angle $\alpha$, let the outward pointing normals of the two planes be $\nvec$ and $\mvec$, as in the figure. Then $$Q(\nvec)Q(\mvec) = R(\avec,2\alpha), \label{QQR}$$ where we use axis-angle notation for the rotation $R$ and where the axis $\avec$ of the rotation is along the line of intersection of the two planes and is given by $$\avec=\frac{\nvec \times \mvec}{|\nvec\times\mvec|} =\frac{\nvec \times \mvec}{\sin\alpha}. \label{QQRaxis}$$ Although the angle $\alpha$, the normals $\mvec$ and $\nvec$, and the axis $\avec$ depend on which of the four choices is made for the dihedral angle, the resulting rotation does not (although it does depend on the order in which the reflections are applied). Let us denote the first rotation taking us from part (a) to part (b) of Fig. \[cycle\] in axis-angle form by $R_{12}(\jvec_{12}, 2\phi_{12})$, where the $12$-subscript on $R$ indicates that this rotation only affects vectors $\Jvec_1$ and $\Jvec_2$. It is obvious from Fig. \[cycle\] that the effect of this rotation on all the angular momentum vectors is to reflect them in the plane 3-4-12, that is, $$\Jvec'_r = Q(\hbox{\rm 3-4-12})\Jvec_r, \label{firstQ}$$ where the prime refers to the values of the vectors after the first rotation and where $r=1,\ldots,4$. This applies for $r=1,2$ because the reflection in the 3-4-12 plane has the same effect as the rotation, and for $r=3,4$ because the reflection does nothing to these vectors and the rotation does not apply to them. Note that the 3-4-12 plane is the same in parts (a) and (b) of the figure (the plane is not affected by the first rotation). Similarly, we denote the second rotation, taking us from part (c) to part (d) of Fig. \[cycle\], by $R_{23}(\jvec'_{23},2\phi_{23})$, where the prime on $\jvec'_{23}$ indicates that the axis is the $23$-direction after the first rotation. Then the effect of the second rotation on all four $\Jvec'_r$ is the same as a reflection in the plane $1'$-4-$2'3$, that is, $$\Jvec''_r = Q(\hbox{\rm $1'$-4-$2'3$})\Jvec'_r. \label{secondQ}$$ The planes of the two rotations intersect in edge 4 of the tetrahedron, which can be seen more clearly in Fig. \[cyclec\], which is the same as part (c) of Fig. \[cycle\] except that all vectors are drawn. In Fig. \[cyclec\], plane 3-4-12 is the back plane, and is also the plane of the first reflection. Plane $1'$-4-$2'3$ (primes are omitted in the figure) is the plane of the second reflection. In comparison to (\[QQR\]), if we identify $\alpha$ with the interior dihedral angle $\phi_4$ at edge 4, then $\mvec$ is the outward normal to plane 3-4-12, while $\nvec$ is the outward normal to plane $1'$-4-$2'3$. Their cross product is in the direction $\jvec_4$, so we have $$\Jvec''_r = R(\jvec_4,2\phi_4)\Jvec_r, \qquad r=1,\ldots,4. \label{R1R2}$$ This is the rotation taking us from part (a) to part (e) of Fig. \[cycle\]; to undo that rotation, we apply $R(-\jvec_4, 2\phi_4)$ to pass from part (e) to part (f) of that figure. To summarize the rotational history, we have applied the rotations $$R(-\jvec_4,2\phi_4) R_{23}(\jvec'_{23},2\phi_{23}) R_{12}(\jvec_{12},2\phi_{12}) \label{3rots}$$ to the tetrahedron in part (a) of Fig. \[cycle\], taking it through a cycle of tetrahedra and returning it to its original shape and orientation. The corresponding $SU(2)$ rotations, with the same axes and angles, are applied to point $p$ in Fig. \[ABintsect1\], taking us along a path $p \to q \to p' \to p''$. Point $p''$ is not shown in Fig. \[ABintsect1\], but it is a point of $I_1$ that projects onto the same tetrahedron as point $p$, since the projected path in $\Lambda_{4j}$ is closed. Points $p$ and $p''$ differ by the phases of the four spinors, that is, by transformations generated by $(I_1,I_2,I_3,I_4)$. Thus, there is a final segment $p'' \to p$ needed to close the path in $\Phi_{4j}$, which runs along the $T^4$ fiber over the initial configuration in $\Lambda_{4j}$. Closing the loop in $\Phi_{4j}$ {#closeloopLPS} ------------------------------- There are several ways to compute the final four phases, but we will discuss just one. We start with vector $\Jvec_1$. The action of the rotations (\[3rots\]) on this vector can be written $$R(-\jvec_4,2\phi_4)R(\jvec_{12},2\phi_{12})\Jvec_1 = \Jvec_1, \label{J1xfm}$$ where we omit the subscripts on the $R$’s because it is understood that only vector $\Jvec_1$ is being acted upon, and where we omit the middle rotation in (\[3rots\]) since it does not act on $\Jvec_1$. The product of the two rotations in (\[J1xfm\]) is not the identity, but it is a rotation about axis $\jvec_1$ since it leaves $\Jvec_1$ invariant. To find the angle of this rotation, we use the Rodrigues-Hamilton formula (Whittaker 1960) for the product of two rotations in axis-angle form. Let $\avec_i$, $i=1,2,3$ be three unit vectors, which we can plot on the unit sphere as in Fig. \[hrformula\]. We join the three points on the unit sphere by arcs of great circles. On following the path $1\to 2\to 3\to 1$ we regard the region to our right as the interior of the spherical triangle formed by the arcs. This gives meaning to the interior angles of the triangle, labeled $\phi_i$, $i=1,2,3$ in the figure. On going from point $i$ to point $i+1$ we can go either the long way or short way around the great circle; the interior of the triangle and the definitions of the interior angles depend on which way we go, but the formula is valid in any case. If we follow the arcs the short way around, we obtain a spherical triangle such as that shown in Fig. \[hrformula\]. Then the formula of Rodrigues and Hamilton is $$R(\avec_3,2\phi_3)R(\avec_2,2\phi_2)R(\avec_1,2\phi_1)=I. \label{prodrule}$$ The proof is obtained by using (\[QQR\]) to write each rotation as a product of reflections, that is, $R(\avec_1,2\phi_1)=Q(31)Q(12)$ and cyclic permutations, where for example $Q(12)$ means reflection in the plane defined by axes $\avec_1$ and $\avec_2$. To apply (\[prodrule\]) to the rotations in (\[J1xfm\]) we imagine the vertex of the original tetrahedron (part (a) of Fig. \[cycle\]) where edges 1, 4 and 12 meet as at the center of the sphere in Fig. \[hrformula\], and we identify axes $(\avec_1 ,\avec_2 ,\avec_3)$ with $(\jvec_{12} ,-\jvec_4 ,\jvec_1)$. Then the $\phi$’s of Fig. \[hrformula\] become the interior dihedral angles of the tetrahedron, and we have $$R(\jvec_1,2\phi_1)R(-\jvec_4,2\phi_4) R(\jvec_{12},2\phi_{12})=I. \label{RloopJ1}$$ Thus, the product of the two rotations in (\[J1xfm\]) is $R( -\jvec_1, 2\phi_1)$. The third rotation found in this manner can be regarded as a holonomy. As the first two rotations in (\[J1xfm\]) are applied to $\Jvec_1$, that vector traces out a closed curve on $S^2$ which is composed of the arcs of two small circles (the curve is a “small lune”). The sphere in question can be regarded as $\Sigma_1$, the symplectic manifold denoted $\Sigma$ in (\[hopfmaps\]) (the 1 subscript indicates that we are dealing with the first of the four copies of $\Sigma$ in $\Sigma_{4j}$). The two rotations themselves in (\[J1xfm\]) can be regarded as the lift of that curve on $S^2$ into $SO(3)$. The lift is an open curve starting at the identity in $SO(3)$ and ending at the product of the two rotations shown in (\[J1xfm\]). To close this curve in $SO(3)$, we apply the third rotation shown in (\[RloopJ1\]). This closed curve in $SO(3)$ may then be lifted (a second time) into $SU(2)$, by replacing each $SO(3)$ rotation by an $SU(2)$ rotation with the same axis and angle, that is, the lift is specified by the product $u(\jvec_1,2\phi_1) u(-\jvec_4,2\phi_4) u(\jvec_{12},2\phi_{12})$. This product is either $+1$ or $-1$, depending on the homotopy class of the closed loop in $SO(3)$. To find this class, we continuously deform the tetrahedron (see part (a) of Fig. \[cycle\]), bringing dihedral angle $\phi_{12}$ to zero, so that the tetrahedron becomes flat. At the end of this deformation, $\phi_{12}=0$ and one of $\phi_1$ and $\phi_4$ is 0 and the other is $\pi$. Thus, the closed loop in $SO(3)$ becomes an element of the noncontractible homotopy class of the homotopy group ${\mathbb Z}_2$ of $SO(3)$, so the product of the three $SU(2)$ matrices is $-1$. Therefore to close the loop in $SU(2)$ we apply a final rotation by an angle $-2\pi$, to obtain $$u(\jvec_1,-2\pi)u(\jvec_1,2\phi_1) u(-\jvec_4,2\phi_4) u(\jvec_{12},2\phi_{12})=+1. \label{uloopJ1}$$ The final rotation by angle $-2\pi$ could have been chosen about any axis if all we wanted to do was to close the loop in $SU(2)$, but we choose axis $\jvec_1$ for the following reason. The first two spin rotations in (\[uloopJ1\]), when applied to a spinor in ${\mathbb C}^2$ over vector $\Jvec_1$, produce another spinor that projects onto $\Jvec_1$ again, that is, it differs from the initial spinor only by an overall phase. This overall phase is the $U(1)$ holonomy mentioned above. The final step in closing the loop in Fig. \[ABintsect1\] is to follow the $I_r$-flows to adjust the phases of the four spinors, in particular, we must follow the $I_1$-flow to adjust the phase of the first spinor. But the observable $\Jvec_1^2=I_1^2$ is a function of $I_1$, so we can follow its flow just as well. But $\Jvec_1^2$ generates a rotation about the direction $\Jvec_1$, that is, it is equivalent to multiplying the first spinor by an $SU(2)$ transformation with axis $\jvec_1$ and some angle. The angle required is seen in (\[uloopJ1\]): it is $2\phi_1-2\pi$. We remark that the final rotation in (\[uloopJ1\]) could have been any angle $-2\pi+4n\pi$. Adding a multiple of $4\pi$ to this angle is equivalent to going around some closed contour on the $A$-manifold (in fact, within $I_1$), which, when the manifolds are quantized and the Maslov phase is taken into account, changes the relative phase of the two branches by a multiple of $2\pi$. Next we turn to vector $\Jvec_3$ and the phase needed to bring the third spinor back its original value after the rotations (\[3rots\]). The action of these rotations on $\Jvec_3$ is given by $$R(-\jvec_4,2\phi_4) R(\jvec'_{23},2\phi_{23}) \Jvec_3 = \Jvec_3, \label{J3xfm}$$ where we omit the first one since it does not act on $\Jvec_3$. The product of the two rotations in (\[J3xfm\]) is a third rotation about axis $\jvec_3$, which can be computed with the help of (\[prodrule\]). This time we identify the 3-4-23 vertex of part (c) of Fig. \[cycle\], seen more clearly in Fig. \[cyclec\], with the origin of the sphere in Fig. \[hrformula\]. We use the version of the tetrahedron after the first rotation (part (c)) since the middle rotation in (\[J3xfm\]) involves the rotated axis $\jvec'_{23}$. Thus we find $$R(-\jvec'_{23},2\phi_{23})R(\jvec_4,2\phi_4) R(-\jvec_3,2\phi_3)=I. \label{3rotsJ3}$$ Now using $R(-\avec,\alpha)=R(\avec,\alpha)^{-1}$ and taking the inverse of (\[3rotsJ3\]), we obtain $$R(\jvec_3,2\phi_3) R(-\jvec_4,2\phi_4) R(\jvec'_{23},2\phi_{23})=I, \label{RloopJ3}$$ which specifies a closed loop in $SO(3)$. We find the homotopy class of this loop by taking $\phi_{23} \to 0$, which makes one of $\phi_4$ and $\phi_3$ zero and the other $\pi$, so the loop in $SO(3)$ belongs to the noncontractible class. Thus, the lift into $SU(2)$ is not closed, but we can close it by appending a final spin rotation about axis $\jvec_3$ by angle $-2\pi$. Thus the closed loop in $SU(2)$ is specified by $$u(\jvec_3,-2\pi) u(\jvec_3,2\phi_3) u(-\jvec_4,2\phi_4) u(\jvec'_{23},2\phi_{23})=+1, \label{uloopJ3}$$ which when applied to the third spinor at point $p$ in Fig. \[ABintsect1\] returns it to its initial value. The final $U(1)$ holonomy of spinor 3, after the application of the three rotations (\[3rots\]), is $2\phi_3 -2\pi$. As for $\Jvec_2$, its cycle on the sphere is specified by $$R(-\jvec_4,2\phi_4) R(\jvec'_{23},2\phi_{23}) R(\jvec_{12},2\phi_{12}) \Jvec_2 = \Jvec_2. \label{J2xfm}$$ The product of the three rotations shown must be a rotation with axis $\jvec_2$. To find the angle, we first use (\[RloopJ3\]) to obtain $$R(-\jvec_4,2\phi_4) R(\jvec'_{23},2\phi_{23}) = R(-\jvec_3,2\phi_3). \label{RprodJ2}$$ Substituting this into (\[J2xfm\]) we obtain a product that we can evaluate with the help of (\[prodrule\]) and with reference to the 2-3-12 vertex of the original tetrahedron (part (a) or (f) of Fig. \[cycle\]). The result is $$R(\jvec_2,2\phi_2) R(-\jvec_3,2\phi_3) R(\jvec_{12},2\phi_{12}) = I. \label{RprodJ2a}$$ Thus the closed loop in $SO(3)$ associated with the loop traced by $\Jvec_2$ on the sphere is specified by $$R(\jvec_2,2\phi_2) R(-\jvec_4,2\phi_4) R(\jvec'_{23},2\phi_{23}) R(\jvec_{12},2\phi_{12}) = I. \label{Rloop2}$$ To find the homotopy class of this loop we deform the tetrahedron into a planar shape as before, and find that two of the four angles $(\phi_2,\phi_4, \phi_{23}, \phi_{12})$ are 0 and two are $\pi$. The loop in $SO(3)$ thus becomes the product of two rotations with angles $2\pi$, which belongs to the contractible homotopy class. Therefore the lift into $SU(2)$ is closed, $$u(\jvec_2,2\phi_2) u(-\jvec_4,2\phi_4) u(\jvec'_{23},2\phi_{23}) u(\jvec_{12},2\phi_{12}) = 1, \label{uloop2}$$ which when applied to the second spinor at point $p$ in Fig. \[ABintsect1\] returns it to its initial value. The final $U(1)$ holonomy of spinor 2, after the application of the three rotations (\[3rots\]), is $2\phi_2$. Finally, we treat vector $\Jvec_4$ and the corresponding $U(1)$ holonomy. The effect of (\[3rots\]) on $\Jvec_4$ is simply $$R(-\jvec_4,2\phi_4)\Jvec_4 = \Jvec_4, \label{J4xfm}$$ since the first two rotations do not act on $\Jvec_4$. This specifies an open curve in $SO(3)$ that can be closed (trivially) by multiplying by a rotation about axis $\jvec_4$, $$R(\jvec_4,2\phi_4) R(-\jvec_4,2\phi_4) = I. \label{Rloop4}$$ The closed loop in $SO(3)$ belongs to the contractible homotopy class, so its lift into $SU(2)$ is closed and is specified by $$u(\jvec_4,2\phi_4) u(-\jvec_4,2\phi_4) = 1. \label{uloop4}$$ When applied to the fourth spinor at point $p$ in Fig. \[ABintsect1\] this sequence of spin rotations returns it to its initial value. The final $U(1)$ holonomy of spinor 4, after the application of the three rotations (\[3rots\]), is $2\phi_4$. To summarize, we have succeeded in constructing the closed loop $p$-$q$-$p'$-$p$ illustrated in Fig. \[ABintsect1\] as the product of a sequence of seven spin rotations, each one generated by the Hamiltonian flow of one of the observables in the $A$- or $B$-list (in the $A$-list while we move on the $A$-manifold, and in the $B$-list while we move on the $B$-manifold). Most of these can be regarded as being generated by the square of some angular momentum vector; for example, the first and second spin rotations, specified by the axes and angles of the right-most rotations in (\[3rots\]), are generated by $\Jvec_{12}^2$ and $\Jvec_{23}^2$, respectively, while the last four rotations can be regarded as being generated by $\Jvec_r^2$, $r=1,\ldots,4$. The third rotation in (\[3rots\]), about axis $\jvec_4$, is generated by $\jvec_4 \cdot \Jvec$, that is, $u(-\jvec_4,2\phi_4) = \exp(i\phi_4 \jvec_4 \cdot \bsigma)$ is applied to all four spinors. The Ponzano-Regge phase {#PRphase} ----------------------- The actions associated with these spin rotations are easily computed, using the complex 1-form (\[1forms\]) and the methods of Sec. I.. To summarize the results, let $\Jvec_p$ be a partial or total sum of the four angular momentum vectors (some range of $r=1,\ldots,4$ is summed over), with magnitude $|\Jvec_p| = J_p$. Then the action along the path generated by $\nvec \cdot \Jvec_p$ with elapsed angle $\theta$ is simply $(\nvec \cdot \Jvec_p)\theta$. In particular, the third rotation in (\[3rots\]), the overall rotation of the tetrahedron taking us from part (e) to part (f) in Fig. \[cycle\], does not contribute to the action since in this case $\Jvec_p =\Jvec_{\rm tot}$ which vanishes on the $A$- and $B$-manifolds. See also (I.). As for the rotations generated by some $\Jvec_p^2$, in this case $\nvec=\jvec_p = \Jvec_p/J_p$, so the action is simply $J_p\theta$. Thus the first and second spin rotations specified by (\[3rots\]) contribute $J_{12}(2\phi_{12}) + J_{23}(2\phi_{23})$ to the total action. As for the last four rotations, in this case $\Jvec_p$ is one of the $\Jvec_r$, $r=1,\ldots,4$, and the four angles are summarized in Sec. \[closeloopLPS\]. These rotations therefore contribute $J_1(2\phi_1-2\pi) + J_2(2\phi_2) + J_3(2\phi_3-2\pi) + J_4(2\phi_4)$ to the total action. Altogether, we have $$S = \oint p\, dx = 2\sum_{r=1}^6 J_r \phi_r -2\pi(J_1+J_3), \label{totalaction}$$ where index $r=5$ means $r=12$ and $r=6$ means $r=23$. This can be written $$S = -2\Psi + 2\pi(J_2 + J_4 + J_{12} + J_{23}), \label{thephase}$$ where on quantized manifolds the final term is an integer multiple of $2\pi$, and where $$\Psi = \sum_{r=1}^6 J_r(\pi-\phi_r). \label{Psidef}$$ The angle $\pi-\phi_r$ is the exterior dihedral angle, so $\Psi$ is the phase of Ponzano and Regge. The amplitude determinant and reduced phase space {#ampdet} ================================================= Amplitude determinants are notorious for the trouble they cause in semiclassical approximations, for example, Gutzwiller’s amplitude determinant (Gutzwiller 1967, 1969, 1970, 1971) has that reputation and in several studies of asymptotic approximations to spin networks the authors have resorted to numerical calculations for the amplitude determinant. In fact, amplitude determinants can be expressed in terms of Poisson brackets, which aids considerably in their evaluation. For example, Wigner’s (1959) amplitude for the $6j$-symbol is a single Poisson bracket, while the amplitude determinant for the $9j$-symbol is a $2\times2$ matrix of Poisson brackets (Haggard and Littlejohn 2010), the derivation of which was the easiest part in the asymptotic formula for the $9j$-symbol. Similarly it is easy to obtain tractable expressions for the amplitude determinant for the $15j$-symbol and other more complicated cases of interest in quantum gravity. In Sec. I. we presented a coordinate-based discussion of amplitude determinants in the $3j$-symbol. For a more geometrical treatment of some of the issues discussed there we refer to the literature on the “quantization commutes with reduction” theorems of Guillemin and Sternberg (1982). Here we will simply review the results of I and discuss their geometrical content. The discussion involves symplectic reduction, which in the case of the $4j$-model of the $6j$-symbol leads to the reduced phase space of the $6j$-symbol, a 2-sphere denoted by $\Gamma$ in (\[4jhopfmaps\]). This space is essential for understanding the semiclassical mechanics of the $6j$-symbol, for example, it is the phase space that underlies the 1-dimensional WKB methods used by Schulten and Gordon (1975a,b), and it played an important role in the derivation of the uniform asymptotic approximation of Littlejohn and Yu (2009), as well as in the semiclassical studies of the volume operator in quantum gravity by Bianchi and Haggard (2011). Densities and amplitude determinants {#densities} ------------------------------------ In this section we adopt a general notation, as in Sec. \[lsolm\], so that our results can be applied to the $4j$-model of the $6j$-symbol, the $12j$-model of the $6j$-symbol, the $9j$-symbol, and many other examples. The phase space is $P=\Reals^{2N}$ with symplectic form $\omega=dp\wedge dx$. The theory of symplectic reduction usually begins with a symplectic group action on phase space, but a more natural starting point in our applications is the components of the momentum map of the group, which is a list of classical observables. Actually, there are two lists. We assume that there exist on $P$ an $A$- and a $B$-list of observables, $(A_1 ,\ldots, A_N)$ and $(B_1 ,\ldots, B_N)$, each of which forms a Lie algebra under the Poisson bracket. We denote contour values by $a_i$ and $b_i$, so that level sets are specified by $A_i=a_i$ and $B_i=b_i$. We denote these level sets by $L_a$ and $L_b$ (these are the $A$- and $B$-manifolds). We assume that the $A_i$’s and $B_i$’s are functionally independent at most places in phase space, so that the generic dimensionality of $L_a$ and $L_b$ is $N$, although it may be less in exceptional cases. The $A$- and $B$-lists of observables correspond to two groups with symplectic actions on phase space, which we denote by $G_a$ and $G_b$, with Lie algebras $\Liealgebra{g}_a$ and $\Liealgebra{g}_b$. We assume the groups are connected and have dimensionality $N$, as in Sec. \[lsolm\]. We denote the structure constants of one or the other group by $c^k{}_{ij}$, as in Sec. \[lsolm\]. We are exclusively interested in contour values $a$ and $b$ that are fixed points of the coadjoint actions of the respective groups, since this implies that $L_a$ and $L_b$, if $N$-dimensional, are Lagrangian. Only Lagrangian manifolds can support semiclassical wave functions. The restriction on the contour values means, however, that we do not have a Lagrangian foliation of $P$. We will be interested in the WKB or semiclassical $x$-space wave function associated with the $A$- or $B$-manifold. We will work with the $A$-manifold, since the $B$-case is similar. For modern treatments of WKB theory, see for example Martinez (2002) or Mishchenko (1990). The variables $x$ are half of the coordinates $(x,p)$ on $P$; we denote “$x$-space” by $Q=\Reals^N$, which abstractly is best seen as the quotient space when $P$ is divided by the foliation into vertical Lagrangian planes, $x={\rm const}$. We write this wave function in the form $$\psi_a(x) = \braket{x}{a}= K \sum_{\rm br} |\Omega(x)|^{1/2} \exp\{i[S(x)-\mu\pi/2]\}, \label{psiaofx}$$ where $K$ is a normalization; where the sum is over branches of the projection $\pi:L_a\to Q$ from the Lagrangian manifold to $Q$, assumed to be locally invertible; where the branch index is suppressed in the sum; where $S$ is the action computed as in Sec. \[ABactions\]; and where $\mu$ is the Maslov index. We associate the function $\Omega(x)$ with an $N$-form $\Omega$ on $Q$ by $$\Omega = \Omega(x)\, dx^1 \wedge \ldots \wedge dx^N. \label{Omegadef}$$ It follows from WKB theory that $\Omega(x)$ satisfies a set of amplitude-transport equations on $Q$, one for each observable $A_i$. These are conveniently expressed in terms of an $N$-form $\sigma$ on $L_a$, defined by $\sigma = \pi^* \Omega$. Then the amplitude transport equations on $Q$ are equivalent to $\Liederiv_{X_i} \sigma=0$ on $L_a$, where $\Liederiv$ is the Lie derivative and where $X_i$, $i=1 ,\ldots,N$ are the Hamiltonian vector fields on $L_a$ associated with the $A_i$, that is, $X_i =\omega^{-1} dA_i$. This means that $\sigma$ is invariant under the infinitesimal action of the group, hence under any finite action (recall that $L_a$ is an orbit of the group action). These are left actions. We express the solution $\sigma$ as follows. The vector fields $X_i$ may be chosen as a (generally non-coordinate) basis in each tangent space at each point of $L_a$. Let $\lambda^i$, $i=1 ,\ldots, N$ be the dual 1-forms on $L_a$, that is, $\lambda^i(X_j) =\delta^i_j$. Then, in the case of compact groups, the solution $\sigma$ is given by $$\sigma =\lambda^1\wedge\ldots\wedge\lambda^N, \label{sigmadef}$$ to within a multiplicative constant. The reason is that $\sigma$ is a version of the right-invariant Haar measure on the group, and for compact groups the left and right Haar measures are equal. Thus, $\sigma$ is invariant under the (left) action of the group, and $\Liederiv_{X_i }\sigma=0$. The action of $G_a$ on $L_a$ provides two ways in which geometric structures on the group can be transferred to $L_a$. First, the group action implies a linear map $:\Liealgebra{g}_a \to T_xL_a$ for each $x\in L_a$, which, under our assumptions, is invertible. This map can be used to push forward a standard $N$-form on $\Liealgebra{g}_a$ to $L_a$. Let the basis in $\Liealgebra{g}_a$ be $\{\xi_i, i=1 ,\ldots, N\}$, corresponding to the observables $A_i$ and vector fields $X_i$, and let $\{\alpha^i, i=1,\ldots,N\}$ be the dual basis in $\Liealgebra{g }_a^*$, that is, $\alpha^i(\xi_j) =\delta^i_j$. Then if the $N$-form $\alpha^ 1\wedge\ldots\wedge\alpha^N$ is pushed forward to $L_a$ in this manner, we obtain $\sigma$, defined by (\[sigmadef\]). A second way involves picking a point $x_0\in L_a$ to serve as an “origin” in $L_a$, and then identifying points $x$ of $L_a$ by the group element $g$ such that $x=gx_0$. This creates a diffeomorphism between a neighborhood of the identity in $G_a$ and a neighborhood of $x_0$ in $L_a$, which can be used to push forward differential geometric structures from $G_a$ to $L_a$. It then turns out that $\sigma$ given by (\[sigmadef\]) is the push forward of the right-invariant $N$-form on $G_a$ associated with $\alpha^ 1\wedge\ldots\wedge\alpha^N$. Thus, $\sigma$ is naturally invariant only under a right action of the group; the only reason it is also invariant under the left action is that for the groups we consider, the left and right Haar measures are identical. To see this from another standpoint, a short calculation shows that $\Liederiv_{X_i }\lambda^j = c^j{}_{ik }\,\lambda^k$, from which follows $\Liederiv_{X_i }\sigma = c^j{}_{ij }\,\sigma$. To derive this we recall that although the Poisson brackets $\{A_i, A_j\}$ vanish on $L_a$, the Lie brackets of the corresponding Hamiltonian vector fields do not, instead we have $[X_i, X_j] = -X_k\,c^k{}_{ij}$. But if the left and right Haar measures are equal, then the adjoint representation $\Ad_g$ is volume-preserving, so the structure constants are traceless, and $\Liederiv_{X_i }\sigma=0$. In our examples we deal only with compact groups, so this condition is met. That the solution is unique to within a multiplicative constant can be seen by supposing that $\sigma'$ is another $N$-form on $L_a$ such that $\Liederiv_{X_i }\sigma'=0$. Since all $N$-forms are proportional, we must have $\sigma' = f\sigma$ where $f$ is a function on $L_a$. This implies $X_i(f)=0$, or $f={\rm const}$ on $L_a$. Now given that (\[sigmadef\]) is the solution we want, we write $\sigma = \pi^*\Omega$ in the form, $$\Omega(x)\, dx^1 \wedge \ldots \wedge dx^N = \lambda^1 \wedge \ldots \wedge \lambda^N, \label{Omegasigmaeqn}$$ where we write simply $dx^i$ for $\pi^*(dx^i)$, and we evaluate both sides on the set of vectors $(X_1 ,\ldots ,X_N)$. This gives $$\Omega(x) \det \{x^i,A_j\} = 1, \label{OmegaPB}$$ where we use $dx^i(X_j) =\{ x^i,A_j\}$. This reproduces Eq. (I.). Finally, as shown in I, the normalization integral, evaluated in the stationary phase approximation, implies $K=1/\sqrt{V_a}$, where $V_a$ is the volume of $L_a$ with respect to $\sigma$. The amplitude of $\braket{b}{a}$ {#scmeba} -------------------------------- A coordinate-based derivation of the amplitude of the semiclassical matrix element $\braket{b}{a}$ was presented in I for the case of the $3j$-symbol. Here we discuss the results from a geometrical point of view, using the general notation of Sec. \[lsolm\] and the previous section. The main issue is that the stationary phase set $L_a \cap L_b$ in the evaluation of the integral $\braket{b}{a} =\int dx \braket{b}{x} \braket{x}{a}$ is not a set of isolated points, as expected on the basis of a naive dimensionality count, but rather a set of manifolds of dimensionality $\ge1$. These nontrivial intersection manifolds are due to the existence of a common “intersection group” of $G_a$ and $G_b$, which may be defined as follows. The basic idea is that the $A$- and $B$-lists of observables may have some observables in common, which generate the intersection group. This is obviously the case in (\[ABfunctionlists\]), for example. But the specific observables that occur in the $A$- and $B$-lists depend on the bases chosen in the Lie algebras $\Liealgebra{g}_a$ and $\Liealgebra{g}_b$, and if a basis is changed then the $A$- or $B$-observables are replaced by linear combinations of themselves. So we need a precise definition of the observables “in common.” The actions of groups $G_a$ and $G_b$ on $P$ provide Lie algebra anti-homomorphisms between $\Liealgebra{g}_a$ and $\Liealgebra{g}_b$ and the Lie algebra of (globally) Hamiltonian vector fields on $P$. By our assumptions these anti-homomorphisms have full rank, so the images of these two anti-homomorphisms are two $N$-dimensional Lie algebras of Hamiltonian vector fields. These two Lie algebras have an intersection which itself is a Lie algebra. Let the dimension of the intersection be $p$, and let $p+q=N$. The intersection Lie algebra is generated by a set of Hamiltonian functions, call them $(C_1 ,\ldots, C_p)$. These can be regarded as the functions common to the original $A$- and $B$-lists, that is, by a change of basis in $\Liealgebra{g}_a$ and $\Liealgebra{g}_b$ we can bring the $A$- and $B$-lists into the form $A=(C,D)$ and $B=(C,E)$, where $D=(D_1 ,\ldots, D_q)$ and $E=(E_1 ,\ldots, E_q)$ are sets of observables that the $A$- and $B$-lists do not have in common. The common observables $C$ generate the action of an “intersection group” $G_0$, that is, they are the components of the momentum map of the action of $G_0$ on $P$. The group $G_0$ is not uniquely determined by the $C$ observables, only its action on $P$ is. The same applies to $G_a$ and $G_b$, which are generated by the $A$- and $B$-lists of observables. But in practice there are convenient ways of choosing all these groups so that the kernels of their actions are discrete. This means that $\dim G_a = \dim G_b = N$ and $\dim G_0 = p$. We may assume moreover that $G_0$ is connected (as we have already for $G_a$ and $G_b$), since we are free to take the connected identity component of any of these groups. The level set of the momentum map of the intersection group $G_0$ plays an important role in what follows. We denote this level set by $L_0$; its equation is $C_i=c_i$, $i=1 ,\ldots ,p$, for some contour values $c_i$. We choose the $c_i$ such that $L_0$ has its generic dimension $2N-p=N+q>N$. The level set $L_0$ is foliated into $A$- and $B$-manifolds, where the $A$-manifolds are parameterized by the values $d_i$ of the observables $D_i$, and the $B$-manifolds by the values $e_i$ of the observables $E_i$, $i=1 ,\ldots, q$. We will assume that the Poisson brackets $\{A_i, A_j\}$ and $\{B_i, B_j\}$ vanish on $L_0$, so that the generic $A$- and $B$-manifolds in $L_0$ are Lagrangian. This is an extension of our earlier assumption, that these Poisson brackets vanish on a specific pair of $A$- and $B$-manifolds. Thus, after removing exceptional manifolds of less than generic dimensionality, $L_0$ is foliated into Lagrangian manifolds in two different ways. These Poisson bracket relations imply $\{C_i, C_j\}=0$ on $L_0$, so that $c$ is a fixed point of the coadjoint action of $G_0$. This does not mean that $L_0$ is Lagrangian (the dimension is wrong), but it is coisotropic (Abraham and Marsden 1978). It also means that the isotropy subgroup of the coadjoint action is the whole group, so the reduced phase space of level set $L_0$ under the $G_0$ action is the space $L_0/G_0$. This draws attention to the orbits of $G_0$, which generically have dimension $p$. Not only is $L_0$ foliated into orbits of $G_0$, so is each $A$- and $B$-manifold, since $G_0$ is a simultaneous subgroup of $G_a$ and $G_b$. Thus, the intersections of $L_a$ and $L_b$, which are the stationary phase sets, are also foliated into orbits of $G_0$. We will assume that $L_a \cap L_b$ is a union of a discrete set of orbits of $G_0$. Based on a dimensionality count, this is the generic case. It holds for example in the $4j$-model of the $6j$-symbol, where $L_a \cap L_b$ consists of two orbits of $G_0$, the sets $I_1$ and $I_2$ in Fig. \[ABintsect\]. It also holds in the $12j$-model used by Roberts (1999) and in our own work on the $9j$-symbol (Haggard and Littlejohn 2010). The orbits of $G_0$ appear in the stationary phase evaluation of $$\begin{aligned} \braket{b}{a} &= \int dx\,\braket{b}{x}\braket{x}{a} = \frac{1}{\sqrt{V_aV_b}} \int dx \sum_{\rm br} {1\over\sqrt{|\det\{x,A\}\det\{x,B\}|}} \nonumber \\ &\qquad\times\exp\{i[S_A(x)-S_B(x) -\mu\pi/2]\}, \label{abintegral} \end{aligned}$$ where we have inserted the normalized $x$-space wave functions for the states $\ket{a}$ and $\ket{b}$, where the sum is over all pairs of branches of both wave functions, and where $\mu$ is the cumulative Maslov index. The stationary phase set is the projection of $L_a \cap L_b$ onto $Q$ ($x$-space); it is a union of the projections of the discrete set of orbits of $G_0$ that make up $L_a\cap L_b$. The projection of each orbit is a subset of $Q$ that is locally $p$-dimensional. As in I, we introduce a local change of coordinates $x \to (y,z)$, where $(y^1,\ldots,y^p)$ are coordinates along the projected orbits and $(z^1,\ldots,z^q)$ are coordinates transverse, with $z=0$ being the projected orbit itself. Then the integral becomes (suppressing normalization and branch sums) $$\fl\int \frac{d^py}{|\det\{y,C\}|} \int \frac{d^qz}{\sqrt{|\det\{z,D\} \det\{z,E\}|}} \exp\{i[S_A(y,z)-S_B(y,z)-\mu\pi/2]\}. \label{yzintegral}$$ When the exponent is expanded to second order in $z$, the leading term $S_A(y,0) - S_B(y,0)$ is independent of $y$, that is, it is constant along the orbits of $G_0$. The second derivative matrix of $S_A-S_B$ with respect to $z$ does depend on $y$, but after doing the Gaussian integral and combining with the determinants in the denominator of the $z$-integration, the result is expressed purely in terms of the Poisson brackets $\{E_i,D_j\}$, which are independent of $y$. The fact that the $z$-integral is independent of $y$, that is, our location on the orbit of $G_0$, is noteworthy. It means that the integral does not depend on the detailed nature of the $z$-coordinates, for example, it is invariant under a coordinate transformation of the form $z'=z'(z,y)$ such that $z=0$ implies $z'=0$. It suggests that we are dealing with a quotient operation in which we divide by the orbits of $G_0$. Another remark is that there is nothing special about the $x$-representation in which the integral (\[abintegral\]) is carried out. The $x$-coordinates were introduced as half of the $(x,p)$ coordinates on $P=\Reals^{2N}$, but any representation related to this one by a metaplectic transformation (Littlejohn 1986) would work as well. This amounts to foliating phase space, not by the vertical Lagrangian planes $x={\rm const}$, but rather by other Lagrangian planes related to this one by any linear, symplectic map. In this manner one can divide the orbits of $G_0$ into segments and do the integral over each segment in a representation in which the projection of the orbit onto the representation space has full rank. The segments into which the orbit of $G_0$ is divided can even be infinitesimal, effectively making the representation for the integral a function of where we are along those orbits. Similarly, one can avoid the caustics of the wave functions $\psi_a$ and $\psi_b$. This is the old idea underlying the Maslov method (Maslov and Fedoriuk 1981) in WKB theory. In any case, once the $z$-integral is done and is recognized to be independent of $y$, the remaining $y$-integral can be lifted to the orbit of $G_0$ in phase space whereupon it becomes just the integral of the Haar measure of $G_0$, giving the volume of the orbit. This Haar measure is normalized as in the previous section, that is, we start with the observables $C_i$, $i=1 ,\ldots, p$, we associate these with Hamiltonian vector fields $X_i =\omega^{-1} dC_i$ (a change of notation from above, where the $X$’s were associated with the $A$’s), we define form $\lambda^i$ dual to the $X_i$ on the orbits of $G_0$, and then the Haar measure on the orbit is taken to be $\lambda^ 1\wedge\ldots\wedge \lambda^p$. The final result is $$\braket{b}{a} = (2\pi i)^{q/2} \frac{V_I}{\sqrt{V_aV_b}} \sum_{\rm br} \frac{1}{\sqrt{|\det\{D_i,E_j\}|}} \exp[i(S_I -\mu\pi/2)], \label{scabme}$$ where now the branches refer to the discrete set of orbits of $G_0$ that make up $L_a \cap L_b$, where $V_I$ is the volume of intersection manifold $I$ (an orbit of $G_0$), where $S_I$ means $S_A-S_B$ evaluated on intersection manifold $I$, and where $\mu$ is a cumulative Maslov index (not necessarily the same as the previous ones). The branch index could otherwise be written as $I$, a label of the intersection manifold, and $V_I$ is taken out of the sum because it does not depend on which intersection manifold is taken. The volume $V_I$ differs from the volume of $G_0$ because in general there is a discrete isotropy subgroup, as in (\[volumeI\]) (one is really computing the volume of a coset space). The result (\[scabme\]) contains a $q\times q$ matrix of Poisson brackets, $\{E_i, D_j\}$, whose geometrical content may be understood in terms of a variation of the discussion of densities in Sec. \[densities\]. First we recall that the $A$- and $B$-lists are decomposed according to $A=(C,D)$, $B=(C,E)$. Next, we fix the contour values $c_i$, $i=1 ,\ldots, p$, so that we have a definite level set $L_0$ of the momentum map of $G_0$. Then we let “$b$-space” be $\Reals^q$ with coordinates $(e_1 ,\ldots, e_q)$ or the region of $\Reals^q$ that is the projection of $L_0$ onto $\Reals^q$, where coordinates $e_i$ are interpreted as the contour values in $E_i=e_i$. The matrix element $\braket{b}{a}$ can be thought of as a wave function on $b$-space for fixed values of the $a$’s, that is, of the $c$’s and $d$’s. We write the amplitude of the semiclassical approximation to this wave function as $|\Omega(e)|^{1/2}$, where $\Omega =\Omega(e)\, de_ 1\wedge\ldots\wedge de_q$ is the associated density (a $q$-form on $b$-space). This density is the projection of the natural density on the $A$-manifold, in the following sense. Let the $A$- or $(C,D)$-observables be associated with vector fields $X_i =\omega^{-1}dC_i$, $i=1 ,\ldots, p$, and $Y_i =\omega^{-1} dD_i$, $i=1 ,\ldots, q$, with dual 1-forms $\lambda^i$, $i=1 ,\ldots, p$ and $\mu^i$, $i=1 ,\ldots, q$. These induce a density $\sigma_ 0\wedge\mu^ 1\wedge\ldots\wedge\mu^q$ on $L_a$, where $\sigma_0 =\lambda^ 1\wedge\ldots\wedge\lambda ^p$ is the Haar measure on $G_0$. Then $\Omega$ satisfies $$\sigma_0 \wedge \pi^*\Omega = \sigma_0 \wedge \mu^1 \wedge \ldots \wedge \mu^q, \label{Omegaofeeqn}$$ where $\pi$ is the projection from $L_0$ or $L_a \subset L_0$ onto $b$-space. Now evaluating both sides on the set of vectors $(X_1 ,\ldots, X_p, Y_1 ,\ldots, Y_q)$, we obtain $$\Omega(e) \det dE_i(Y_j) =1. \label{Omegaofedet}$$ This gives the amplitude shown in (\[scabme\]), since $dE_i(Y_j) = \{E_i, D_j\}$. This discussion has treated the $A$- and $B$-manifolds asymmetrically, projecting from the $A$-manifold onto $b$-space, but we could have projected the density on the $B$-manifold onto $a$-space with the same result. There are really four densities, two [*on*]{} the $A$- and $B$-manifolds, and two [*of*]{} the $A$- and $B$-manifolds. This discussion leads us to consider the reduced phase space $\Gamma = L_0/G_0$, which is parameterized by the contour values $c_i$, $i=1 ,\ldots, p$. As is standard in symplectic reduction, the symplectic form on $\Gamma$ is obtained by pulling back vectors from $\Gamma$ to $L_ 0\subset P$ and evaluating them on the symplectic form on $P$; this is meaningful because the answer does not depend on where on the orbit of $G_0$ they are pulled back to, nor on the component of the pulled-back vectors along the orbit. A consequence is that the projections of the $A$- and $B$-manifolds onto $\Gamma$, which are $q$-dimensional since the $G_0$ orbits are $p$-dimensional, are Lagrangian on $\Gamma$. Since we are assuming that $L_a \cap L_b$ is a discrete union of $G_0$ orbits, the projected manifolds on $\Gamma$ intersect in a discrete set of isolated points. This would be the generic case on a symplectic manifold of dimension $2q$. Also, $G_0$-invariant functions on $L_0$ project onto functions on $\Gamma$, whose Poisson brackets on $\Gamma$ are the same as the Poisson brackets of the original functions on $P$. Such functions in the present discussion include the observables $D_i$ and $E_i$, $i=1 ,\ldots, q$, so the Poisson brackets $\{E_i,D_j\}$ of (\[scabme\]) are naturally interpreted as living on $\Gamma$. The case of the $6j$-symbol {#caseof6j} --------------------------- It is straightforward to apply the general theory of Secs. \[densities\] and \[scmeba\] to the case of the $6j$-symbol. The phase space is $\Complexes^8 =\Reals^{16}$ so $N=8$. The common observables $C$ are $I_r$, $r=1 ,\ldots,4$, and $\Jvec_{\rm tot}$, so $p=7$ and $q=1$. The group $G_0$ is $T^4 \times SU(2)$. The level set $L_0$ of the momentum map of $G_0$, for $I_r = J_r$, $r=1,\ldots,4$, and $\Jvec_{\rm tot}=\zerovec$, is the subset of $\Phi_{4j}$ upon which the four angular momenta have specified lengths and their vector sum is zero. It is logical that this would be the subset of the classical phase space $\Phi_{4j}$ that corresponds to the space $\ZS$ of four-valent intertwiners, introduced in Sec. \[4jmodel\], on which $\Ihat_r =j_r$, $r=1,\ldots,4$ and $\Jvechat_{\rm tot} =\zerovec$. The contour values of the $C$’s, that is, of $I_r$, $r=1 ,\ldots,4$ and $\Jvec_{\rm tot}$, must be chosen so that $L_0$ has its maximum dimensionality, namely, 9. The condition is $J_r>0$, $r=1 ,\ldots, 4$ and the polygon inequality, $$\max\{J_1,J_2,J_3,J_4\} < \frac{1}{2} \sum_{r=1}^4 J_r, \label{polygonJr}$$ which as indicated must not be saturated. This is the condition that it is possible to make a noncollinear polygon in $\Reals^3$ out of vectors of four given positive lengths. If this condition is satisfied, then $L_0$ is foliated into $A$- and $B$-submanifolds whose generic dimensionality is 8. There is only one observable of the $D$- and $E$-type; according to (\[ABfunctionlists\]) we should make the identifications $D=\Jvec_{12}^2$ and $E=\Jvec_{23}^2$. However, since we used $d\theta$ in the volume form on the $A$- and $B$-manifolds when computing the volume (\[ABvolume\]), we should use instead $D=|\Jvec_{12}|$ and $E=|\Jvec_{23}|$, since these are conjugate to $\theta$ (really $\theta_a$ and $\theta_b$, since there are two of them). Then the Poisson bracket for the amplitude is $$\fl\{E,D\}= \{|\Jvec_2+\Jvec_3|,|\Jvec_1+\Jvec_2|\}= \frac{\Jvec_2\cdot[(\Jvec_2+\Jvec_3)\times(\Jvec_1+\Jvec_2)]} {|\Jvec_{12}||\Jvec_{23}|} = \frac{6V}{J_{12} J_{23}}, \label{6jPB}$$ where $V$ is the signed volume of the tetrahedron, $6V =\Jvec_ 1\cdot (\Jvec_ 2\times\Jvec_3)$, where we have used (\[Lambda4jPB\]) to evaluate the Poisson bracket, and where in the final step we have evaluated the Poisson bracket on $A$- and $B$-manifolds with contour values $J_{12}$ and $J_{23}$. We see the appearance of Wigner’s volume (Wigner 1959). The volume changes sign between the two intersection manifolds $I_1$ and $I_2$, but it appears with an absolute value sign in (\[scabme\]) so both stationary phase points in the $6j$-symbol have the same amplitude. The amplitude $|\Omega|^{1/2}$ contains the factor $$\sqrt{J_{12}J_{23}} = \frac{1}{2}\sqrt{(2j_{12}+1)(2j_{23}+1)}, \label{sqrtfacs}$$ which, when evaluated as shown on quantized manifolds, reproduces the square roots seen in (\[6jme\]). Thus, based on the pieces of the formula we have determined so far, we can write $$\left\{\begin{array}{ccc} j_1 & j_2 & j_{12} \\ j_3 & j_4 & j_{23} \end{array}\right\} =\frac{e^{i\pi/4}}{\sqrt{12\pi|V|}} \frac{1}{2} \left[e^{i(S_1-\mu_1\pi/2)} + e^{i(S_2-\mu_2\pi/2)} \right], \label{PRv1}$$ where $S_1$ and $S_2$ are given by (\[S1S2defs\]) (they are the phases $S_A-S_B$, evaluated on the intersection manifolds $I_1$ and $I_2$ of Fig. \[ABintsect\]). These phases contain the phase conventions for the states $\ket{A}$ and $\ket{B}$ shown in Fig. \[4jABstates\], which appear semiclassically as the origins for action integrals on the Lagrangian manifolds, denoted $a_0$ and $b_0$ in Fig. \[ABintsect\]. Since we have not considered these phase conventions yet, we cannot say what $S_1$ and $S_2$ are, but the difference $S=S_2-S_1$ is given in terms of the Ponzano-Regge phase by (\[thephase\]). To return to the 9-dimensional space $L_0$, it projects onto what was called “$b$-space” in Sec. \[scmeba\], which in the present case is the $J_{23}$-axis, producing the interval $[J_{23,{\rm min}}, J_{23,{\rm max}}]$, given by (\[J23maxmin\]). The inverse image of a point on this interval is a $B$-manifold with a given value of $J_{23}$. Similarly $L_0$ projects onto “$a$-space”, that is, the $J_{12}$-axis, producing the interval (\[J12maxmin\]), the inverse image of a point of which is an $A$-manifold with given $J_{12}$ value. The space $L_0$ projects onto the $J_{12}$-$J_{23}$ plane producing the classically allowed region of Fig. \[square\]; the inverse image of a point of this region is an intersection of an $A$- and a $B$-manifold of given $J_{12}$ and $J_{23}$ values. If the point does not lie on the caustic curve, the intersection consists of two disconnected components $I_1$ and $I_2$, each an orbit of $G_0$, as in Fig. \[ABintsect\]. On the caustic curve these components merge into one. Finally, by dividing by $G_0$, $L_0$ projects onto the reduced phase space $\Gamma$, which has dimensionality $9-7=2$. We now turn to this space. The reduced phase space $\Gamma$ {#evalrps} -------------------------------- Spaces of the type $\Gamma$ seem to have appeared first in the work of Kapovich and Millson (1995, 1996). Those authors showed that the space of polygons of a given number of sides with fixed lengths in $\Reals^3$, modulo overall rotations, is a symplectic manifold. In fact, for quadrilaterals the space is precisely $\Gamma$. This space was recently subjected to direct geometric quantization by Charles (2008), who connected it with the $6j$-symbol and used it for a new derivation of the Ponzano-Regge asymptotic formula. The space of five-sided polygons is the analog of $\Gamma$ for the $9j$-symbol; it was used by Haggard and Littlejohn (2010) in their study of the asymptotics of the $9j$-symbol. To visualize $\Gamma=L_0/G_0$ it helps to carry out the reduction in two steps. In the first step, we choose contour values $J_r>0$, $r=1 ,\ldots,4$, for which the level set $I_r=J_r$ in $\Phi_{4j}$ is the product of 3-spheres $(S^3)^4$, as shown in (\[4jhopfmaps\]). The $I_r$, $r=1 ,\ldots,4$, generate the action of the group $U(1)^4$, corresponding to the phases of the four spinors. Dividing this level set by $U(1)^4$, we obtain the symplectic manifold $\Sigma_{4j}$ shown in (\[4jhopfmaps\]), which topologically is $(S^2)^4$, and which consists of all sets of four vectors $\Jvec_r$ in $\Reals^3$, $r=1 ,\ldots,4$, with fixed lengths, $|\Jvec_r| = J_r$. The space $\Sigma_{4j}$ is 8-dimensional. In the second step, we consider the submanifold in $\Sigma_{4j}$ upon which $\Jvec_{\rm tot}=0$, which is a level set of the momentum map of the action of $SO(3)$ on $\Sigma_{4j}$. This manifold consists of sets of four vectors $\Jvec_r$ in $\Reals^3$ of fixed lengths $J_r$ such that $\sum_r \Jvec_r=\zerovec$. The vectors can be placed end-to-end to form a “closed link,” that is, a four-sided polygon in $\Reals^3$. The set of closed links is denoted CL in (\[4jhopfmaps\]). We assume the polygon inequality (\[polygonJr\]) is satisfied, so the space CL has dimensionality $8-3=5$. Since this is the level set $\Jvec_{\rm tot}=0$, the isotropy subgroup of the $SO(3)$-action is $SO(3)$ itself, so the reduced phase space is ${\rm CL}/SO(3)$, which is the space $\Gamma$ in (\[4jhopfmaps\]). This has $5-3=2$ dimensions. The phase space $\Gamma$ is parameterized by the four fixed, positive values $J_r$, $r=1 ,\ldots,4$ that satisfy the polygon inequality (\[polygonJr\]). A point of this space specifies a quadrilateral in $\Reals^3$ of the given lengths, modulo overall proper rotations. Once the quadrilateral has been determined, one can draw in the two remaining edges, of lengths $J_{12}$ and $J_{23}$, to fill in a tetrahedron. Thus, $\Gamma$ can be thought of as the shape space for a set of tetrahedra, four of whose edges have fixed, positive lengths. The two lengths that are variable are on opposite sides of the tetrahedron. Here we define “shape” as a configuration modulo proper rotations, as in Littlejohn and Reinsch (1997); two shapes related by spatial inversion are generally distinct. The lengths $|\Jvec_{12}|$ and $|\Jvec_{23}|$ are variable functions on $\Gamma$. In fact, any rotationally invariant quantity associated with the tetrahedron, such as the dihedral angles, the areas of the faces, the signed volume, etc, is also a function on $\Gamma$. It is easy to find coordinates on $\Gamma$. We may take one coordinate to be $J_{12}$, which varies between the bounds (\[J12maxmin\]). (From this point on we drop the distinction between $|\Jvec_{12}|$ and $J_{12}$, and similarly for $J_{23}$ and $J_{13}$.) For a fixed value of $J_{12}$, the allowed set of shapes is generated by executing the “butterfly” motion about the axis $\Jvec_{12}$, that is, rotating the triangle 1-2-12 about this axis, relative to the 3-4-12 triangle. We recall this motion is the Hamiltonian flow generated by $\Jvec_{12}^2$ or the magnitude $|\Jvec_{12}|$ (see (\[J12sqflow\])). Thus, coordinates can be taken to be $(J_{12},\phi_{12})$, where $\phi_{12}$ is the interior dihedral angle about the 12-edge. For each value of $J_{12}$ in the interior of the range (\[J12maxmin\]), a circle of shapes is generated as $\phi_{12}$ goes from 0 to $2\pi$; but at the endpoints there is only a single shape. For example, at the lower limit of (\[J12maxmin\]), the case $J_{12,{\rm min}} = |J_1-J_2|$ is illustrated in part (a) of Fig. \[jlimits\]. In this case the rotation of vectors $\Jvec_1$ and $\Jvec_2$ about the axis $\Jvec_{12}$ does not change the shape. Similarly, part (b) illustrates the case $J_{12}= J_{12,{\rm min}} =|J_3-J_4|$. In this case, the $\Jvec_{12}^2$-action rotates the 1-2-12 triangle, but does not change the shape since the new configurations that result are related to the original ones by an overall $SO(3)$ transformation. A similar analysis applies at the upper limit of (\[J12maxmin\]). We see that the set of shapes generated in this manner for $J_{12}$ in the interior of the interval (\[J12maxmin\]) is a cylinder, but the two endpoints are single points that pinch the cylinder at the two ends, creating a topological sphere. Topologically, $\Gamma=S^2$. The symplectic form on $\Gamma$ may be obtained by projecting $\omega =dp\wedge dx$ on $\Phi_{4j}$ as described in Sec. \[scmeba\] , but it is easier just to notice that $\phi_{12}$ is the parameter of evolution along the flow generated by $J_{12}$, so $(\phi_{12},J_{12})$ form a canonically conjugate $(q,p)$ pair on the sphere. The same obviously applies to $J_{23}$ and $J_{31}$ and their conjugate (dihedral) angles, so we have $$dJ_{12} \wedge d\phi_{12} = dJ_{23} \wedge d\phi_{23} = dJ_{31} \wedge d\phi_{31}, \label{RPSomegas}$$ indicating three sets of canonical coordinates on $\Gamma$, related by canonical transformations. These are examples of the action-angle variables discovered by Kapovich and Millson (1995, 1996), which in all cases are closely related to the recoupling schemes used in angular momentum theory. The length $J_{31}$ and the associated dihedral angle $\phi_{31}$ do not appear in the coupling scheme we described in Sec. \[4jmodel\] or in the tetrahedra we have discussed so far, but would appear in a different tetrahedron in which the four vectors are placed end-to-end in a different order. So far we have described the construction of $\Gamma$ as a purely classical problem, but if the $J_r$ are quantized, $J_r=j_r+1/2$, $r=1 ,\ldots 4$, then one can speak of a quantized level set $L_0$ and quotient space $\Gamma$. The area of the sphere $\Gamma$ with respect to the form $dJ_{12 }\wedge d\phi_{12}$ is obviously just $2\pi(J_{12,{\rm max}} - J_{12,{\rm min}})$. If $\Gamma$ is quantized, then by (\[J12j12maxmin\]) and (\[dimZS\]) the area is $2\pi D$, where $D=\dim\ZS$. When quantized, $\Gamma$ contains one Planck cell of area $2\pi$ for every state in the Hilbert space $\ZS$. Obviously we obtain the same area if we use either of the other symplectic forms in (\[RPSomegas\]). Quantized curves in $\Gamma$ {#quantrps} ---------------------------- States in $\ZS$, such as the $A$- and $B$-states given by (\[4jABdefs\]), are represented semiclassically by Lagrangian manifolds in $\Gamma$, which are quantized curves in that space. For example, the $A$-states are represented by quantized levels sets of $J_{12}$. To plot these we map $\Gamma$ into a unit 2-sphere in ${\mathbb R}^3$ with standard coordinates $(x,y,z)$ by associating $(\phi_{12},J_{12})$ with a standard set of spherical angles $(\theta,\phi)$, where $\phi=\phi_{12}$ and $$J_{12} = J_{12,{\rm min}} + \frac{D}{2} (1+\cos\theta), \label{RPSthetadef}$$ where $D=J_{12,{\rm max}} - J_{12,{\rm min}}$ (generally), or $D=\dim\ZS$ (when $\Gamma$ is quantized). Here $x =\sin\theta\cos\phi$, $y =\sin\theta\sin\phi$, $z =\cos\theta$ are understood. Then the symplectic form on the sphere is $(-1/2)D\sin\theta\, d\theta\wedge d\phi$. These coordinates make the $J_{12}$ orbits look nice (they are small circles), but not the orbits of the other variables, so they should not be thought of as having any privileged role. We use them mainly for plotting figures. The $x$-$z$ plane in these coordinates does have an invariant meaning, however; it is the plane upon which the tetrahedra are planar (the volume vanishes). Moreover, time reversal is a reflection in this plane ($y \to -y$). The quantized orbits of $J_{12}$, on a space $\Gamma$ with five Planck cells of area, are illustrated in part (a) of Fig. \[rps\]. The orbits are just small circles, as noted. The orbits are numbered $n=0 ,\ldots,4$, in order of increasing $J_{12}$, and enclose area $(n+\frac{1}{2})2\pi$. The minimum and maximum values of $J_{12}$ are at the south and north poles, respectively, because of the choice (\[RPSthetadef\]) of coordinates, but the minimum and maximum quantized values are separated from the values at the poles by $\frac{1}{2}$, as indicated by (\[J12J23quant\]), since the polar caps defined by the last quantized orbit before the poles must have area $\frac{1}{2}$. The quantized orbits of $J_{23}$ are illustrated in parts (b) and (c) of that figure. The function $J_{23}$ has a minimum on the $x$-$z$ plane with $x>0$, and increases monotonically toward a maximum on the same plane with $x<0$. The orbits are time reversal symmetric, but have no symmetry under inversion through the origin in the coordinates used (there is no reason why they should). Also shown are some quantized orbits of $J_{13}$, in part (d) of Fig. \[rps\]. Another interesting observable on the reduced phase space is the signed volume $V=(1/6)\Jvec_1 \cdot (\Jvec_2 \times \Jvec_3)$. This has been suggested by Chakrabarti (1964) and by Lévy-Leblond and Lévy-Nahas (1965) as a “democratic” alternative to the usual choices $J_{12}^2$, $J_{23}^2$, $J_{13}^2$ for the intermediate variable in the quantum problem of the recoupling of three angular momenta. More recently, Carbone (2002) have reconsidered the use of this observable, and have derived both recursion relations and asymptotic formulas for the matrix elements connecting the $J_{12}$-basis and the $V$-basis. The level sets (orbits) of $V$ are plotted in part (e) of Fig. \[rps\]. The orbits plotted are not quantized, but are evenly spaced in $V$ about $V=0$, and labelled $n=-2 ,\ldots, +2$. The $V=0$ contour is the $x$-$z$ plane, which is also a quantized orbit when $\dim\ZS$ is odd. The orbit $V=v$ is mapped by time reversal into the orbit $V=-v$. The Bohr-Sommerfeld condition for the quantized orbits of $V$ can be expressed in terms of complete elliptic integrals of third kind. We will report on this and other results on the spectrum of $V$ and on the wave functions $\braket{j_{12}}{V}$ in future publications. Preliminary results are reported in Bianchi and Haggard (2011). The $6j$-symbol on the reduced phase space {#6jrps} ------------------------------------------ As noted below (\[6jme\]) the $6j$-symbol is proportional to the matrix element $\braket{j_{23}}{j_{12}}$, so on $\Gamma$ it is represented semiclassically by the quantized curves, $J_{23} =j_{23} +\frac{1}{2}$ and $J_{12} =j_{12} +\frac{1}{2}$. This is illustrated in part (f) of Fig. \[rps\] for the case $\braket{j_{23}} {j_{12}} =\braket {0}{4}$, where the numbers refer to the labeling of the quantized curves in parts (a) and (b) of that figure. These curves are projections of the $A$- and $B$-manifolds in $\Phi_{4j}$, and their intersections, the points labelled $P$ and $Q$ in the figure, are the projections of intersections manifolds $I_1$ and $I_2$ as illustrated in Fig. \[ABintsect\], respectively. This follows because we defined $I_1$ as the intersection manifold upon which $V<0$, and $V<0$ at point $P$ since it lies in the region $y<0$. Also, the path (with direction indicated) surrounding the shaded area in part (f) of Fig. \[rps\] is the projection of the loop (with direction indicated) in $\Phi_{4j}$ illustrated in Fig. \[ABintsect1\]. We assign arrows to the path in $\Gamma$ by following the flows generated by $J_{12}^2$ and $J_{23}^2$ in the direction of increasing $\phi_{12}$ and $\phi_{23}$. Thus, the path The path $P\to Q\to P$ in part (f) of Fig. \[rps\] corresponds to the sequence of rotations carried out on the tetrahedron in Fig. \[cycle\]. On reaching point $Q$, the path turns down, because the rate of change of $J_{12}$ along the $J_{23}$-flow is the Poisson bracket $\{J_{12},J_{23}\}$, which is negative when $V>0$, as follows from (\[6jPB\]). One can also see this geometrically in Fig. \[cycle\]; parts (b) and (c) of that figure correspond to point $Q$, and it is clear that $J_{12}$ at first decreases, then increases again, on carrying out the rotation about $\Jvec_{23}$, that is, on passing from part (c) to part (d) of Fig. \[cycle\]. The symplectic area of the shaded region in part (f) of Fig. \[rps\] is the same as the Ponzano-Regge phase, to within a constant that depends on $j_r$, $r=1,\ldots,4$. On quantized manifolds it will not matter what region on $S^2$ is taken as the interior of the closed loop on the sphere (the choices differ by multiples of the total area of the sphere), but to be precise one must worry about this. The Ponzano-Regge phase is also $-2$ times the $F_4$-type generating function (Goldstein 1980, Miller 1974). Let us write $(q,p)=(\phi_{12},J_{12})$, $(Q,P)=(\phi_{23},J_{23})$ and $F_4(p,P) = -(1/2) S$, where $S$ is given by (\[thephase\]). Then we have $$q = -\frac{\partial F_4}{\partial P} = \frac{\partial}{\partial J_5} \sum_{r=1}^6 J_r \phi_r, \label{F4q}$$ where we use index 5 instead of 12. Here the angles $\phi_r$ must be interpreted as functions of the $J$’s. A more geometrical way of stating the same thing is to express the canonical transformation $(\phi_{12},J_{12}) \to (\phi_{23},J_{23})$ as a map between two spheres with symplectic forms (\[RPSomegas\]). Then the graph of the canonical transformation is a Lagrangian manifold in $S^2\times S^2$ with symplectic form $dJ_{23}\wedge d\phi_{23} - dJ_{12}\wedge d\phi_{12}$. This Lagrangian manifold is itself a sphere, whose projection onto $J_{12}$-$J_{23}$ space is precisely the interior plus boundary of the oval curve in Fig. \[square\]. The Lagrangian manifold has a double-valued projection onto the interior of the oval region, and a single-valued projection at the boundary (the caustics). The two disks fit together to form a sphere. The Ponzano-Regge Formula {#PRformula} ------------------------- If we factor out the phase $$\left(\frac{S_1+S_2}{2}\right) - \left(\frac{\mu_1+\mu_2}{2}\right)\frac{\pi}{2} \label{splitphase}$$ from the quantity in the square bracket in (\[PRv1\]), we obtain a quantity proportional to $$\cos\left[\left(\frac{S_1-S_2}{2}\right) - (\mu_1-\mu_2)\frac{\pi}{4} \right]. \label{PRphasev1}$$ However, in view of (\[S1S2defs\]), (\[relphase\]), (\[thephase\]) and (\[Psidef\]), this can be written as a sign times $$\cos\left(\Psi+\Delta\mu\frac{\pi}{4}\right), \label{PRcosV2}$$ where the sign is $(-1)^{j_2+j_4+j_{12}+j_{23}}$ and where $\Delta\mu=\mu_2-\mu_1$. Thus to obtain the argument of the cosine we must compute the relative Maslov index between the two branches, just as we have already computed the relative action $S=S_2-S_1$. This is the easiest of the Maslov index computations in deriving the Ponzano-Regged formula. We will omit details and simply remark that the calculation can be carried out entirely on the reduced phase space $\Gamma$, where it is not difficult to show that $\Delta\mu=1$. There remains the overall phase, which, in view of the reality of $6j$-symbol, must be a sign $\pm1$. This phase cannot be computed without taking into account the phase conventions for the two states $\ket{A}$ and $\ket{B}$, as well as the absolute Maslov indices on the intersection manifolds $I_1$ and $I_2$. The analogous phase for the $9j$-symbol was the most difficult part of the derivation of the results presented in Haggard and Littlejohn (2010). Since (in the case of the $6j$-symbol) the answer is known, we can see that this overall phase (combined with the $e^{i\pi/4}$ appearing in (\[PRv1\])) must be 1. We had no such luxury in the case of the $9j$-symbol, nor did we find it possible to guess the answer. In any case, the final result is the Ponzano-Regge formula, $$\left\{\begin{array}{ccc} j_1 & j_2 & j_{12} \\ j_3 & j_4 & j_{23} \end{array}\right\} =\frac{1}{\sqrt{12\pi|V|}} \cos\left(\Psi+\frac{\pi}{4}\right). \label{PRv2}$$ For reasons of space we will omit the details in the calculation of these final phases, also because we have made the main points we wanted to make about the $6j$-symbol. These are the importance of the reduced phase space $\Gamma$, the geometry of symplectic reduction connecting it with higher dimensional spaces, and the manner in which elements of the semiclassical calculation (phases, amplitude determinants, etc) can be mapped from one space to another. Conclusions =========== A glance at the calculation of Roberts (1999) shows that it is much easier and more elegant to compute the relative phase $S=S_2-S_1$ of the $6j$-symbol in a symmetrical or $12j$-model than in the $4j$-model presented in this paper. The sum over edges times dihedral angles appears almost immediately. On the other hand, we have given a much easier way of computing the amplitude determinant, reducing it to a single Poisson bracket, whereas Roberts had to evaluate a large determinant. Obviously what is needed is a way of connecting the various models at a semiclassical level, so that actions, amplitude determinants, Maslov indices, etc, can be mapped from one model to another and computed wherever most convenient. We have investigated this question and will report on our results in the future. For now we merely make a few comments. Already in this paper we have mapped amplitude determinants between various spaces, such as the Schwinger phase space $\Phi_{4j}$ and the reduced phase space $\Gamma$, which are connected by symplecic reduction. One might guess, therefore, that the $12j$- or symmetrical model of the $6j$-symbol and the $4j$-model are related by some kind of symplectic reduction. This is not the case, however, at least not in the way that $\Phi_{4j}$ and $\Gamma$ are related. One piece of picture, however, is the following. It is well known that a unitary map on a Hilbert space $U:\HS\to\HS$ corresponds semiclassically to a symplectic map or canonical transformation $T:\Phi\to\Phi$, where $\Phi$ is the symplectic manifold corresponding to $\HS$. Also, the symplectic map $T$ is conveniently viewed via its graph in $\Phi\times\Phi$, regarded as a symplectic manifold in its own right with symplectic form $\omega_1-\omega_2$, where the subscripts refer to the first and second factors of $\Phi\times\Phi$ (Abraham and Marsden 1978). With this understanding, the graph of $T$ is a Lagrangian manifold, one which supports semiclassically the operator $U$ in the same way as a Lagrangian manifold in $\Phi$ supports a vector in $\HS$. That is, $U$, which begins as a map $:\HS\to\HS$, is reinterpreted as a function $:\HS\otimes\HS^*\to\Complexes$, that is, a “wave function” on a doubled space. The basic ideas inherent in this picture were developed by Hörmander (1971) and are also present in Miller (1974), while an elementary explanation of the geometrical situation is given by Littlejohn (1990). It turns out that this picture generalizes to linear maps between Hilbert spaces $M:\HS\to\HS'$, where the two Hilbert spaces need not have the same dimensionality and where the map need not be unitary or even invertbible. The genralization involves again symplectic reduction, but in a different manner than that in which it appears in this paper. This situation arises in the comparison of two models of the $3j$-symbol, which is a simpler version of the comparison between the $12j$- and $4j$-models of the $6j$-symbol. The first is a $2j$- or Clebsch-Gordan model, in which the Hilbert space is $\CS_{j_1}\otimes\CS_{j_2}$, where the operator $\Jvec_3$ is a function of $\Jvec_1$ and $\Jvec_2$, that is, $\Jvec_3 =\Jvec_1 +\Jvec_2$. The second is a symmetrical or $3j$-model, in which the Hilbert space is $\CS_{j_1}\otimes\CS_{j_2}\otimes\CS_{ j_3}$ and the operator $\Jvec_3$ is independent, but in which we are interested only in states satisfying the constraint $\Jvec_1+\Jvec_2+\Jvec_3=0$. We will report on these investigations in the future. Roberts’ (1999) derivation of the Ponzano-Regge formula effectively proceeded by writing the $6j$-symbol as a scalar product $\braket{B}{A}$, where $\ket{A}$ and $\ket{B}$ are given by Fig. \[robertsAB\], then writing the wave functions for $\ket{A}$ and $\ket{B}$ in the Bargmann (1962) or coherent state representation to obtain an integral representation for the $6j$-symbol as an integral over $\Complexes^{24}$. In the coherent state representation there is one copy of $\Complexes$ for each degree of freedom, while in a $12j$-model there are two degrees of freedom for each $j$, hence 24 degrees of freedom total. Roberts then used the stationary phase approximation to evaluate the integral. The Bargmann representation has been used in a similar manner in several recent asymptotic studies in the quantum gravity literature. In I and in this paper, however, we have mostly worked in a represenation-independent manner. Our emphasis on Lagrangian manifolds and other geometrical structures in phase space is part of this approach. From a certain point of view the coherent state representation is just another representation, albeit a complex one, so there is the question of whether it plays any privileged role or offers any advantages. For some purposes it certainly does, for example, the wave functions in the coherent state representation are polynomials that can be written down explicitly, and these in turn are useful for deriving generating functions and other things, as shown by Schwinger (1952) and Bargmann (1962). Another point is that in the case of $SU(2)$, the coherent state representation appears naturally in the method of geometric quantization, that is, the holomorphic sections of Hermitian line bundles over the symplectic manifold $S^2$ are represented explicitly by Bargmann’s entire analytic wave functions. On the other hand, it is not obvious to us that the identification of the stationary phase set or the other aspects of the semiclassical calculation are easier or more transparent in the coherent state representation than in the representation-independent approach of this paper. Moreover, if one focuses too narrowly on the stationary phase evaluation of an integral, one misses the geometrical structures that support the representation-independent approach. We had to assemble and extend several of the ideas presented in this paper for our derivation of the asymptotic form of the $9j$-symbol, and we have studied other extensions such as “$g$-inserted” spin networks, in which a group element or $D$-matrix is inserted in the edges of the spin network. Such $g$-inserted spin networks are basic amplitudes in loop quantum gravity (Rovelli 2004), taking one from the basis of spin connections to the spin network basis. The asymptotics of these amplitudes leads to piecewise flat manifolds similar to those introduced by Regge (1961). We will report on these and other developments in the future. The authors would like to acknowledge a number of stimulating and helpful conversations with Annalisa Marzuoli and Mauro Carfora. References {#references .unnumbered} ========== Abraham R and Marsden J E 1978 [*Foundations of Mechanics*]{} (Reading, Massachusetts: Benjamin/Cummings) Alesci E, Bianchi E, Magliaro E and Perini C 2008 preprint gr-qc 0809.3718 Ališaukas S 2000 [*J Math Phys*]{} [**41**]{} 7589 Anderson R W and Aquilanti V 2006 [*J Chem Phys*]{} [**124**]{} 214104 Aquilanti V, Cavalli S and De Fazio D 1995 [*J. Phys. Chem.*]{} [**99**]{} 15694 Aquilanti V, Cavalli S and Coletti C 2001 [*Chem. Phys. Lett.*]{} [**344**]{} 587 Aquilanti V and Coletti C 2001 [*Chem. Phys. Lett.*]{} [**344**]{} 601 Aquilanti V, Haggard H M, Littlejohn R G and Yu Liang 2007 [*J Phys A*]{} [**40**]{} 5637 Arnold V I 1989 [*Mathematical Methods of Classical Mechanics*]{} (New York: Springer-Verlag) Baez J C 1996 [*Adv. Math.*]{} [**117**]{} 253 Baez J C and Barrett J W 1999 [*Adv. Theor. Math. Phys.*]{} [**3**]{} 815 Baez J C, Christensen J D and Egan G 2002 [*Class. Quantum Grav.*]{} [**19**]{} 6489 Balazs N L and Jennings B K 1984 [*Physics Reports*]{} [**104**]{} 347 Balcar E and Lovesey S W 2009 [*Introduction to the Graphical Theory of Angular Momentum*]{} (Berlin: Springer-Verlag) Bargmann V 1962 [*Rev. Mod. Phys.*]{} [**34**]{} 829 Barrett J W 1998 [*Adv. Theor. Math. Phys.*]{} [**2**]{} 593 Barrett J W and Crane L 1998 [*J. Math. Phys.*]{} [**39**]{} 3296 Barrett J W, Dowdall R J, Fairbairn W J, Gomes H and Hellmann F 2009 preprint gr-qc 0902.1170 Barrett J W and Steele C M 2003 [*Class. Quantum Grav.*]{} [**20**]{} 1341 Barrett J W and Williams R M 1999 [*Adv. Theor. Math. Phys.*]{} [**3**]{} 209 Berger M 1987 [*Geometry I*]{} (Berlin: Springer-Verlag) Berry M V 1977 [*Phil Trans R Soc*]{} [**287**]{} 237 Berry M V 1984 [*Proc Roy Soc Lond A*]{} [**392**]{}, 45 Bianchi E and Haggard H 2011 [*Phys Rev Lett*]{} [**107**]{} 011301 Biedenharn L C and Louck J D 1981 [*The Racah-Wigner Algebra in Quantum Theory*]{} (Reading, Massachusetts: Addison-Wesley) Borodin K S, Kroshilin A E and Tolmachev V V (1978) [*Teoreticheskaya i Matematicheskaya Fizika*]{} [**34**]{} 110, English translation [*Theoretical and Mathematical Physics*]{} [**34**]{} 69 Brink D M and Satchler G R 1993 [*Angular Momentum*]{} (Oxford: Oxford University Press) Brunnemann J and Rideout D 2008 preprint gr-qc 0706.0469 2010 preprint gr-qc 1003.2348 Carbone G, Carfora M and Marzuoli A 2002 [*Class. Quantum Grav.*]{} [**19**]{} 3761 Carlip S 1998 [*Quantum Gravity in $2+1$ Dimensions*]{} (Cambridge: Cambridge University Press) Chakrabarti A 1964 [*Ann. Inst. H Poincaré*]{} [**1**]{} 301 Charles L 2008 preprint math 0806.1585 Conrady F and Freidel L 2008 [*Phys. Rev. D*]{} [**78**]{} 104023 Crippen G M and Havel T F 1988 [*Distance Geometry and Molecular Confirmation*]{} (Baldock, UK: Research Studies Press Ltd) Cushman R H and Bates L 1997 [*Global Aspects of Classical Integrable Systems*]{} (Basel: Birkhäuser Verlag) Cvitanovich P 2008 [*Group Theory: Birdtracks, Lie’s, and Exceptional Groups*]{} (Princeton, Princeton University Press) Danos M and Fano U 1998 [*Phys. Rep.*]{} [**304**]{} 155 De Fazio D, Cavalli S and Aquilanti V 2003 [*Int. J. Quant. Chem.*]{} [**93**]{} 91 Ding Y and Rovelli C 2010 preprint gr-qc 0911.0543 Edmonds A R 1960 [*Angular Momentum in Quantum Mechanics*]{} (Princeton: Princeton University Press) El Baz E and Castel B 1972 [*Graphical Methods of Spin Algebras*]{} (New York: Marcel Dekker) Freidel L and Louapre D 2003 [*Class. Quantum Grav.*]{} [**20**]{} 1267 Goldstein H 1980 [*Classical Mechanics*]{} 2nd ed (Reading, Massachusetts: Addison-Wesley) Groenewold H J 1946 [*Physica*]{} [**12**]{} 405 Guillemin V and Sternberg S 1977 [*Geometric Asymptotics*]{} (Providence, Rhode Island: American Mathematical Society) 1982 [*Invent Math*]{} [**67**]{} 515 Gurau R 2008 The Ponzano-Regge asymptotic of the $6j$-symbol: an elementary proof [*Preprint*]{} math-ph 0808.3533 Gutzwiller M C 1967 [J Math Phys]{} [**8**]{} 1979 1969 [J Math Phys]{} [**10**]{} 1004 1970 [J Math Phys]{} [**11**]{} 1791 1971 [J Math Phys]{} [**12**]{} 343 Hackett J and Speziale S 2007 [*Class. Quantum Grav.*]{} [ **24**]{} 1525 Haggard H M and Littlejohn R G 2010 [*Classical and Quantum Gravity*]{} [**27**]{} 135010 Hörmander L 1971 [*Acta Math*]{} [**127**]{} 79 Kapovich M and Millson J J 1995 [*J. Differential Geometry*]{} [**42**]{} 133 1996 [*J. Differential Geometry*]{} [**44**]{} 479 Kauffman L H and Lins 1994 [*Temperley-Lieb Recoupling Theory and Invariants of 3-Manifolds*]{}, (Princeton, New Jersey: Princeton University Press, New Jersey) Kirillov A A 1976 [*Elements of the Theory of Representations*]{} (New York: Springer-Verlag) 2004 [*Lectures on the Orbit Method*]{} (Providence, Rhode Island: American Mathematical Society) Kurchan, Leboeuf P and Saraceno M 1989 [*Phys. Rev. A*]{} [**40**]{}, 6800 Lévy-Leblond J-M and Lévy-Nahas M 1965 [*J. Math. Phys.*]{} [**6**]{} 1372 Lewandowski J 1996 preprint gr-qc 9602035 Lindgren I and Morrison J 1986 [*Atomic Many-Body Theory*]{} (Springer-Verlag, New York) Littlejohn R G 1986 [*Phys. Reports*]{} [**138**]{} 193 1990 [*J. Math. Phys.*]{} [**31**]{} 2952 Littlejohn R G and Reinsch M 1995 [*Phys. Rev. A*]{} [**52**]{} 2035 1997 [*Rev Mod Phys*]{} [**69**]{} 213 Littlejohn R G and Yu L 2009 [*J. Phys. Chem. A*]{} [**113**]{} 14904 Major S A and Seifert M D 2001 preprint 0109056 Marsden J E and Ratiu T 1999 [*Introduction to Mechanics and Symmetry*]{} (New York: Springer-Verlag) Martinez A 2002 [*An Introduction to Semiclassical and Microlocal Analysis*]{} (New York: Springer-Verlag) Marzuoli A and Rasetti M 2005 [*Ann. Phys.*]{} [**318**]{} 345 Maslov V P and Fedoriuk M V 1981 [*Semi-Classical Approximations in Quantum Mechanics*]{} (Dordrecht: D. Reidel) Messiah A 1966 [*Quantum Mechanics*]{} (New York: John Wiley) Mishchenko A S, Shatalov V E and Sternin B Yu 1990 [*Lagrangian manifolds and the Maslov operator*]{} (Berlin: Springer Verlag) Miller W H 1974 [*Adv. Chem. Phys.*]{} [**25**]{} 69 Minkowski H 1897 [*Nachr. Gess. Wiss.*]{} 198 Moyal J E 1949 [*Proc Camb Phil Soc*]{} [**45**]{} 99 Neville D 1971 [*J Math Phys*]{} [**12**]{} 2438 2006 [*Phys Rev D*]{} [**73**]{} 124004 Nicolai H and Peeters K 2007, in [*Approaches to Fundamental Problems*]{}, Seiler E and Stamatescu I-O eds (Berlin: Springer) (Lecture Notes in Physics v. 721) 151 Ooguri H 1992a [*Nucl. Phys. B*]{} [**382**]{} 276 1992b [*Mod. Phys. Lett. A*]{} [**7**]{} 2799 Ozorio de Almeida A M 1998 [*Phys Rep*]{} [**295**]{} 265 Penrose R 1971 in [*Combinatorial Mathematics and its Applications*]{}, edited by D. Welsh (New York: Academic Press) Ponzano G and Regge T 1968 in [*Spectroscopy and Group Theoretical Methods in Physics*]{} ed F Bloch  (Amsterdam: North-Holland) p 1 Ragni M, Bitencourt A C P, Ferreira C da S, Aquilanti V, Anderson R W and Littlejohn R G 2010 [*Int. J. Quantum Chem.*]{} [**110**]{} 731 Regge T 1961 [*Il Nuovo Cimento*]{} [**19**]{}, 558 Regge T and Williams R M 2000 [*J. Math. Phys.*]{} [**41**]{} 3964 Roberts J 1999 [*Geometry and Topology*]{} [**3**]{} 21 Rovelli C 2004 [*Quantum Gravity*]{} (Cambridge University Press) Rovelli C and Smolin L 1995 [*Phys. Rev. D*]{} [**52**]{}, 5743 Rovelli C and Speziale S [*Class. Quantum Grav.*]{} [**23**]{} 5861 Sastry K R S 2002 [*Forum Geom*]{} [**2**]{} 167 Schulten K and Gordon R G 1975a [*J. Math. Phys.*]{} [**16**]{} 1961 1975b [*J. Math. Phys.*]{} [**16**]{} 1971 Schwinger J 1952 [*On Angular Momentum*]{} U.S. Atomic Energy Commission, NYO-3071, reprinted in Biedenharn L C and van Dam H 1965 [*Quantum Theory of Angular Momentum*]{} (New York: Academic Press) Smolin L 2005 An Invitation to Loop Quantum Gravity [*Preprint*]{} hep-th/0408048 Stedman G E 1990 [*Diagram Techniques in Group Theory*]{} (Cambridge: Cambridge University Press) Taylor Y U and Woodward C T 2004 preprint math 0406.228 Thiemann T 2007 [*Modern Canonical Quantum General Relativity*]{} (Cambridge: Cambridge University Press) Turaev and Viro 1992 [*Topology*]{} [**31**]{} 865 Varshalovich D A, Moskalev A N and Khersonskii V K 1981 [*Quantum Theory of Angular Momentum*]{} (Singapore: World Scientific) Van der Veen R 2010 PhD Dissertation “Asymptotics of quantum spin networks” http://www.science.uva.nl/research/math/Research/Dissertations/Veen2010.text.pdf Voros A 1989 [*Phys. Rev. A*]{} [**40**]{}, 6814 Whittaker E T 1960 [*A Treatise on the Analytical Dynamics of Particles and Rigid Bodies*]{} (Cambridge: Cambridge University Press) Wormer P E S and Paldus J 2006 [*Adv Quantum Chemistry*]{} [**51**]{} 59 Weissman Y 1982 [*J. Chem. Phys.*]{} [**76**]{}, 4067 Weyl H 1927 [*Z Phys*]{} [**46**]{} 1 Wigner E P 1932 [*Phys Rev*]{} [**40**]{} 749 1940 [*On the matrices which reduce the Kronecker products of representations of simply reducible groups*]{} (unpublished), reprinted in Biedenharn L C and van Dam H 1965 [*Quantum Theory of Angular Momentum*]{} (New York: Academic Press) 1959 [*Group Theory*]{} (Academic Press, New York) Woodhouse N M J 1991 [*Geometric Quantization*]{} (Oxford: Oxford University Press) Yutsis A P, Levinson I B and Vanagas V V 1962 [*The Theory of Angular Momentum*]{} (S Monson, Jerusalem)
{ "pile_set_name": "ArXiv" }
--- abstract: 'A $\cU(1)$ gauge theory turns, on physically motivated models of Quantum Spacetime, into a $\cU(\infty)$ gauge theory, hence free classical electrodynamics is no longer free and neutral fields may have electromagnetic interactions. We discuss the last point for scalar fields, possibly describing dark matter; we have in mind the gravitational collapse of binary systems or future applications to self gravitating Bose-Einstein condensates as possible sources of evidence of quantum gravitational phenomena. The effects so far considered, however, seem too faint to be detectable at present.' author: - Sergio Doplicher - Klaus Fredenhagen - Gerardo Morsella - Nicola Pinamonti title: Pale Glares of Dark Matter in Quantum Spacetime --- Introduction ============ One of the main difficulties of present day Physics is the lack of observation of quantum aspects of gravity; Quantum Gravity has to be searched without guide from nature, except the need to explain the observed universe as carrying traces of quantum gravitational phenomena in the only “laboratory” suitable to those effects, the universe itself few instants after the Big Bang. Looking forward to see those traces in the cosmic gravitational waves background (for the analysis of quantum linearized perturbations see the pioneering work [@MFB], see also [@BFHPR]), one can ask whether an expected consequence of Quantum Gravity, the quantum nature of spacetime at the Planck scale, might leave observable traces. Indeed Quantum Spacetime [@DFR] would explain some aspects, as the horizon problem [@DMP], usually explained by inflation, without having to make that hypothesis; but are there effects which $only$ QST would explain? Free classical electromagnetism on Quantum Spacetime would be no longer free: the electromagnetic field and potential $F, A$ would fulfill $$\partial_\mu F^{\mu \nu} -i [A_{\mu}, F^{\mu \nu}] = 0,$$ where the commutator would not vanish due to the quantum nature of spacetime. This fact was noticed [@DF] at the very beginning of searches on Quantum Spacetime, as well as its first consequence: plane waves would be still solutions, but their superpositions would not in general, they would loose energy in favor of mysterious massive modes (see also [@Z]). A naive computation showed that, by that mechanism, a monochromatic wave train passing through a partially reflecting mirror should loose, in favor of those ghost modes, a fraction of its energy - a very small fraction, unfortunately, of the order of one part in $10^{-130}$ [@DF]. This looked too small to be worth a more accurate computation. However QST should reveal itself, as discussed here below, causing an electromagnetic interaction of $neutral$ fields. This was noticed at the beginning as well but looked even less promising of visible consequences (see however [@CJSWW]). But the recent years brought increasing evidence of the role of dark matter, and the possibility of collapse of huge dark matter binary systems; near the collapse, could those systems emit a seizable amount of electromagnetic radiation, and thus show a signature of the quantum nature of spacetime at the Planck scale? In this note we discuss this problem, and show that a primitive, semiclassical evaluation of that emission gives again a very small result: the fraction of the mass of such a system converted into electromagnetic radiation per unit time by the mechanism envisaged here would be less than $10^{-89} s^{-1}$; nothing comparable to the few percents of the total mass converted into the gravitational wave radiation in the recently observed merges of binary black holes, GW150914 and GW151226, which inspire the numerical input of our calculation. Our discussion proceeds as follows. In Sec. 1, after having recalled the main terminology, notations and results about the model of QST that we use, we discuss the action of the gauge group of QST on a neutral scalar field, and derive the interaction of the latter with the electromagnetic field by the covariant derivative prescription. Moreover, we show that such interaction can be described in terms of a magnetic moment associated to the scalar neutral field. Then, in Sec. 2, we evaluate the electromagnetic energy emitted in a state describing the precession of a stellar member of a collapsing binary system; that energy being computed, once the magnetic moment is evaluated, according to classical electrodynamics. We also comment on another manifestation of that magnetic moment, at first glance potentially giving rise to more visible effects, as they would be only quadratic in the Planck length (but hard to be detected anyway, see comments below): the electromagnetic (besides the gravitational) deviations of charged particles by a massive stellar object of dark matter interposed between us and a distant source. But in the case of the previously used data, we find a contribution to the angular deviation of the order $10^{-34}$. The magnetic moment of a neutral scalar field induced by Quantum Spacetime ========================================================================== The model of QST adopted here is suggested by the Principle of $Gravitational$ $Stability$ $against$ $localization$ $of$ $events$  [@DFR],  [@BDMP]. This principle implies Spacetime Uncertainty Relations $$\Delta q_0 \cdot \sum \limits_{j = 1}^3 \Delta q_j \gtrsim 1 ; \sum \limits_{1 \leq j < k \leq 3 } \Delta q_j \Delta q_k \gtrsim 1 .$$ for the coordinates $q_\mu$ of an event, which must be implemented by Spacetime commutation relations $$[q_\mu ,q_\nu ] = i \lambda_P^2 Q_{\mu \nu}$$ where $\lambda_P$ is the Planck length and where $Q_{\mu \nu}$ satisfies appropriate Quantum Conditions. The simplest solution is given by: $$\begin{aligned} [q^\mu , Q^{\nu \lambda}] &= 0,\\ \label{dopleq14} Q_{\mu \nu} Q^{\mu \nu} &= 0,\\ ((1/2) \left[q_0 ,\dots,q_3 \right] )^2 &= I,\end{aligned}$$ where $$\begin{aligned} \left[q_0 ,\dots,q_3 \right] &\equiv \det \left( \begin{array}{ccc} q_0 & \cdots & q_3 \nonumber\\ \vdots & \ddots & \vdots \\ q_0 & \cdots & q_3 \end{array} \right)\\&\equiv \varepsilon^{\mu \nu \lambda \rho} q_\mu q_\nu q_\lambda q_\rho =\nonumber\\&= - (1/2) Q_{\mu \nu} (*Q)^{\mu \nu}\end{aligned}$$ (notice that $Q_{\mu \nu} Q^{\mu \nu}$ is a scalar and $ Q_{\mu \nu} (*Q)^{\mu \nu}$ is a pseudoscalar, hence we square it in the Quantum Conditions). Called for brevity *the Basic Model of Quantum Spacetime*, this model implements exactly the Space Time Uncertainty Relations and is fully Poincaré covariant. The $noncommutative$ C\* algebra $\mathcal E$ of Quantum Spacetime can be associated to the above relations, by a procedure [@DF; @BDMP] which applies to more general cases. Assuming that the $q_\lambda, Q_{\mu \nu}$ are selfadjoint operators and that the $Q_{\mu \nu}$ commute $strongly$ with one another and with the $q_\lambda$, the relations above can be seen as a bundle of Lie Algebra relations based on the joint spectrum of the $Q_{\mu \nu}$. *Regular representations* are described by representations of the C\* group algebra of the unique simply connected Lie group associated to the corresponding Lie algebra, with the condition that $I$ is not an independent generator but is represented by the unit operator. They obey the Weyl relations $$\label{eq:weyl_rels} e^{ih_\mu q^\mu}e^{ik_\nu q^\nu}=e^{-\frac i2h_\mu Q^{\mu\nu} k_\nu}e^{i(h+k)_\mu q^\mu},\quad h,k\in\mathbb R^4.$$ The C\* algebra of Quantum Spacetime is the C\* algebra of a continuous field of group C\* algebras based on the spectrum of a commutative C\* algebra. In our case, that spectrum - the joint spectrum of the $Q_{\mu \nu}$ - is the manifold $\Sigma$ of the real valued antisymmetric 2-tensors fulfilling the same relations as the $Q_{\mu \nu}$ do: a homogeneous space of the proper orthochronous Lorentz group, identified with the coset space of $SL(2,\bC)$ mod the subgroup of diagonal matrices. Each of those tensors can be taken to its rest frame, where the electric and magnetic parts ${\boldsymbol{e}}$, ${\boldsymbol{m}}$ are *parallel unit vectors*, by a boost, and go back with the inverse boost, specified by *a third vector, orthogonal to those unit vectors*; thus $\Sigma$ can be viewed as the tangent bundle to two copies of the unit sphere in 3 space - its *base* $\Sigma_1$. Irreducible representations at a point of $\Sigma_1$ identify with *Schrödinger’s $p, q$ in $2$ degrees of freedom*. The fibers are therefore the C\* algebras of the Heisenberg relations in 2 degrees of freedom - the algebra of all compact operators on a fixed infinite dimensional separable Hilbert space. The continuous field can be shown to be trivial. Thus the C\* algebra $\mathcal E$ of Quantum Spacetime is identified with the tensor product of the continuous functions vanishing at infinity on $\Sigma$ and the algebra of compact operators. The mathematical generalization of points are pure states. *Optimally localized states* minimize $$\Sigma_\mu(\Delta_\omega q_\mu)^2;$$ the minimum being $2$, reached by states concentrated on $\Sigma_1$, at each point coinciding (if optimally localized at the origin) with the *ground state of the harmonic oscillator*. Such states are the proper quantum version of points; the classical limit of Quantum Spacetime is then the product of Minkowski space and $\Sigma_1$. Thus extradimensions, described by the doubled 2-sphere, are predicted by Quantum Spacetime. Optimally localized states are central in the definition of the *Quantum Wick Product*, which removes the UV divergences in the Gell-Mann–Low expansion of the S matrix for polynomial interactions on QST  [@BDFP:2003]. The mentioned minimum, of the order of the squared Planck length in generic units, for the sum of the four squared uncertainties in the coordinates of an event is the first manifestation of a deeper fact: the minimum euclidean distance between two independent events in Quantum Spacetime is of the order of the Planck length in all reference frames. More generally, for each geometric operator like distance, area, three volume, or four volume, the sum of the squares of all spacetime components is, in each reference frame, at least of the order of the appropriate power of the Planck length [@BDFP:2011]. These are mathematical results on the Quantum Geometry of Quantum Minkowski space. But dynamics, already at the level of a semiclassical treatment of Gravity, strongly suggests that the minimal distance between two independent events ought to have a dynamical meaning, $diverging$ when a singularity is approached [@DMP]. This fact allows a possible solution of the horizon problem [@DMP], and will play a role in our discussion in the last Section. Our first task is now to formulate and analyze gauge theories on the model of Quantum Spacetime just described. On ordinary classical spacetime, the gauge group of electromagnetism is the group of (regular) functions from Minkowski spacetime $\bR^4$ to $U(1)$, which can be regarded as (a subgroup of) the group of unitaries of the algebra $C_b(\bR^4) = M(C_0(\bR^4))$. Going to Quantum Spacetime, this should be naturally replaced by $\cG = \cU(M(\cE))$, the unitaries of the multipliers of the Quantum Spacetime algebra $\cE$. It is therefore a rather interesting possibility that the gauge group of electromagnetism could also act nontrivially on a real scalar field $\varphi(q)$ on QST, as $$\label{eq:gaugephi} \varphi(q) \to U\varphi(q) U^*, \qquad U \in \cG.$$ Of course on commutative spacetime the above action is instead trivial, because $U$ and $\varphi$ commute. In order to find a Lagrangian invariant under the above action, we should introduce a covariant derivative $D_\mu$, i.e., a derivation on $\cE$ such that, under the action of $\cG$, $$D_\mu\varphi(q) \to UD_\mu\varphi(q)U^*.$$ This is accomplished by defining $$\label{eq:covariantderivative} D_\mu\varphi(q) := \partial_\mu \varphi(q) -ie[A_\mu(q),\varphi(q)],$$ where $e$ is the electron charge (see below for a discussion of this choice) which describes the coupling with the gauge field, $\partial_\mu$ is the derivation on $\cE$ defined by $$\partial_\mu \varphi(q) = \frac{\partial}{\partial a_\mu} \varphi(q+a {\mathbbm{1}})|_{a=0},$$ and $A_\mu$ is the electromagnetic potential on QST, on which $\cG$ is assumed to act as $$\label{eq:gaugeA} A_\mu(q) \to UA_\mu(q)U^* +ie^{-1}U\partial_\mu U^*,$$ which reduces to the ordinary gauge transformation on commutative spacetime by writing $U = e^{i\Lambda}$. This also explains the choice of $e$ in , as the coupling constant between the electromagnetic potential and the neutral field $\varphi$. In fact, $A_\mu$ will also interact with the electron field $\psi$, which transforms as $\psi(q) \to U\psi(q)$, and therefore the choice $D_\mu\psi(q) = \partial_\mu\psi(q)-ieA_\mu(q)\psi(q)$ for its covariant derivative gives the correct interaction. A potential problem in this respect is represented by the fact that it seems difficult to write the interaction, on Quantum Spacetime, of $A_\mu$ with a field of charge different from $0,\pm e$ (like the quark fields). For a discussion in the framework of formal \*-products and the Seiberg-Witten map see  [@CJSWW]. The fact that  actually gives the correct definition of covariant derivative for the gauge transformation , is easily verified: the transformed $D_\mu\varphi(q)$ reads in fact, $$\begin{split} \partial_\mu (U\varphi(q)U^*) &-ie[UA_\mu(g)U^*,U\varphi(q)U^*] + [U\partial_\mu U^*,U\varphi(q)U^*]\\ &=U D_\mu\varphi(q) U^* + \partial_\mu U \varphi(q) U^* + U\varphi(q)\partial_\mu U^* + [U\partial_\mu U^*,U\varphi(q)U^*] \\ &=U D_\mu\varphi(q) U^*, \end{split}$$ where the last equation follows from the fact that $$[U\partial_\mu U^*,U\varphi(q)U^*] = -\partial_\mu U \varphi(q) U^*-U\varphi(q)\partial_\mu U^*,$$ which is easily verified using the commutativity of $U$ and $U^*$ and the identity $U\partial_\mu U^* = -(\partial_\mu U)U^*$, consequence of the Leibniz rule for $\partial_\mu$. We obtain therefore the following Lagrangian covariant under gauge transformations (which therefore gives rise to an invariant action) $$\cL = \frac12\eta^{\mu\nu}D_\mu \varphi(q)D_\nu \varphi(q) - \frac12m^2 \varphi(q)^2,$$ and then, expanding the covariant derivatives, interaction terms between the neutral scalar field and the electromagnetic potential given by $$\label{eq:interaction} \cL_I = -\frac{ie}{2}\{[A_\mu(q),\varphi(q)],\partial^\mu\varphi(q)\}-\frac{e^2}{2}[A_\mu(q),\varphi(q)][A^\mu(q),\varphi(q)],$$ with curly brackets denoting the anticommutator. We note that on classical spacetime $\cL_I$ vanishes, as it should, since $A_\mu$ and $\varphi$ commute.[^1] Therefore, we will understand $\cL_I$ as defined through the noncommutative product in $\cE$. This has the drawback that it will depend explicitly on the center. If we neglect, as customary, the quadratic term in $A_\mu$ (weak field approximation) we obtain, for the interaction part of the action, $$S_I = -\frac{ie}{2}\ \int d^4q\, \{[A_\mu(q),\varphi(q)],\partial^\mu\varphi(q)\} = -ie \int d^4q\, A_\mu(q)[\varphi(q),\partial^\mu\varphi(q)],$$ where the cyclicity of $\int d^4q$ was used. We obtain therefore the interaction of $A_\mu$ with a current $j^\mu(q) = -ie[\varphi(q),\partial^\mu\varphi(q)]$, and the classical Euler-Lagrange equations for $A_\mu$ take the form $$\partial_\mu F^{\mu \nu} -ie[A_\mu, F^{\mu \nu}] = - j^\nu,$$ with the field strength defined by $F_{\mu \nu} = \partial_\mu A_\nu - \partial_\nu A_\mu -ie[A_\mu,A_\nu]$ and transforming as $F_{\mu\nu} \to UF_{\mu\nu}U^*$. In order to understand the physical meaning of this interaction, we now assume $A_0(q) = 0$, $A_h(q) = \frac12\ep_{hjk} B^jq^k$, the potential corresponding to an external constant magnetic field $\boldsymbol{B}$ in the classical spacetime limit ($\lambda_P \to 0$) and we again neglect in  the quadratic term in $A_\mu$, thus obtaining $$\label{eq:Binteraction}\begin{split} \cL_I &= -\frac{ie}{4}\ep_{hjk}B^j\big\{\big[q^k,\ph(q)\big],\partial^h\ph(q)\big\} =\frac{e}{2}\ep_{jkh}B^jQ^{k\mu}\{\partial^\mu \ph(q), \partial^h \ph(q)\}, \end{split}$$ where in the second equation we used the identity $[q^\nu, f(q)] = iQ^{\nu\mu} \partial_\mu f(q)$. The above term corresponds then to the energy of a total magnetic moment $\boldsymbol{M}$ with components, in generic units, $$\label{eq:magnetic} M_j = (e/2) \lambda _P ^2 \int_{q^0=t}d^3q\,\left[\frac12(\{\partial_l\ph,\partial^l\ph\}\de_{jk}-\{\partial_j\ph,\partial^k\ph\})m_k-\ep_{jkh}\{\partial_0\ph,\partial^h\ph\}e_k\right], \qquad j=1,2,3,$$ where $e_k := Q^{0k}$ and $m_k := \frac 1 2 \ep_{khl} Q^{hl}$ are respectively the electric and magnetic components of the antisymmetric 2-tensor $Q^{\mu\nu}$. In the next section we will give some numerical estimates on the electromagnetic radiation and on the perturbations of the motion of charged particles associated to such a magnetic moment in suitable astrophysical situations. Some potentially observable consequences ======================================== Defining, as usual, the free scalar field on QST as [@DFR] $$\ph(q) = \int_{\bR^3} \frac{d{{\boldsymbol{k}}}}{\sqrt{\om({{\boldsymbol{k}}})}} [a({{\boldsymbol{k}}})\otimes e^{-ikq}+a({{\boldsymbol{k}}})^*\otimes e^{ikq}],$$ and specializing to a point in the spectrum of the $Q$’s, where $\boldsymbol{e} =\boldsymbol{m}$, a computation yields the following expression for the total magnetic moment  $$\begin{split} {\boldsymbol{M}}(t) = \frac{e\la_P^2}{2}\int_{\bR^3} \frac{d{{\boldsymbol{k}}}}{\om({{\boldsymbol{k}}})}&\left\{ \left[a(-{{\boldsymbol{k}}})a({{\boldsymbol{k}}}) e^{-2i\om({{\boldsymbol{k}}})t}+a({{\boldsymbol{k}}})^*a(-{{\boldsymbol{k}}})^* e^{2i\om({{\boldsymbol{k}}})t}\right] \cos(\om({{\boldsymbol{k}}}){\boldsymbol{e}}\cdot{{\boldsymbol{k}}}){{\boldsymbol{k}}}^2 {\boldsymbol{e}}^\perp\right.\\ &\quad+\left.2a({{\boldsymbol{k}}})^*a({{\boldsymbol{k}}})[2\om({{\boldsymbol{k}}}){{\boldsymbol{k}}}\wedge {\boldsymbol{e}}+ {{\boldsymbol{k}}}^2 {\boldsymbol{e}}^\perp]\right\}, \end{split}$$ with ${\boldsymbol{e}}^\perp = {\boldsymbol{e}}- ({{\boldsymbol{k}}}\cdot{\boldsymbol{e}}){{\boldsymbol{k}}}/{{\boldsymbol{k}}}^2$ the component of ${\boldsymbol{e}}$ orthogonal to ${{\boldsymbol{k}}}$. Therefore the effective magnetic moment of a particle with sharp momentum ${{\boldsymbol{k}}}$ is given by $$\label{eq:momentsharp} {\boldsymbol{\mu}}_{{\boldsymbol{e}},{{\boldsymbol{k}}}} = e\la_P^2\left[2{{\boldsymbol{k}}}\wedge {\boldsymbol{e}}+ \frac{{{\boldsymbol{k}}}^2}{\om({{\boldsymbol{k}}})} {\boldsymbol{e}}^\perp\right].$$ Of course detectable effects, if any, of the above interaction can be obtained in situations which give rise to a very large magnetic moment. To this end, it is natural to consider a compact “star” of $\ph$ particles in rapid rotation around a very massive companion, akin to a binary pulsar or black hole. Neglecting the rotation of this star around its axis, a rough estimate of the associated magnetic moment ${\boldsymbol{M}}_S$ can be obtained by treating such an object as composed by classical particles in uniform rotation with a given angular frequency $\om$, and by associating to such particles the magnetic moment obtained from . More in detail, we choose a reference system in which the orbit lies on the $(x,y)$ plane, and we indicate by $(\theta,\phi)$ the spherical coordinates of ${\boldsymbol{e}}$ with respect to this system. Moreover, recalling that for the binary black hole giving rise to the event GW150914 the angular frequency just before the merger was $\omega \cong 471 \,s^{-1}$ and the radius of the orbit was $R \cong 350 \,km$ [@LV], so that the speed in natural units ($\hbar = c = 1$) was $\beta = \om R \cong 0.6$, we may also assume that the motion of the particles is non-relativistic and approximate $\omega({{\boldsymbol{k}}}) \cong m$ in . We obtain then $$\label{eq:starmoment}\begin{split} {\boldsymbol{M}}_S(t) = e\la_P^2M&\left\{2R\om \left(\begin{matrix} \cos \om t \cos \theta\\ \sin \om t \cos \theta\\ -\cos (\om t-\phi) \sin \theta \end{matrix}\right) \right.\\ &\left.+ R^2\om^2\left[ \left(\begin{matrix} \sin\theta(\cos\phi-\sin \omega t \sin(\omega t-\phi)) \\ \sin \theta(\sin \phi + \cos \omega t \sin(\omega t-\phi))\\ \cos \theta \end{matrix}\right) +\frac 1 5 \left(\frac r R\right)^2 \left(\begin{matrix} \sin\theta \cos\phi \\ \sin \theta \sin\phi \\ 2\cos \theta \end{matrix}\right) \right]\right\}, \end{split}$$ with $M$ the mass of the object and $r$ its radius. In the particular case in which ${\boldsymbol{e}}$ is normal to the orbital plane, ${\boldsymbol{M}}_S$ precedes then around it with the same angular frequency $\om$ of the object motion. In the general case, ${\boldsymbol{M}}_S(t)$ can be written as sum of a constant moment, which of course does not give rise to emission of electromagnetic radiation, and of a time-dependent moment of the form $$\sum_{i} {\boldsymbol{M}}_i \cos(\omega_i t -\psi_i),$$ with $\omega_i = \omega$, $2\omega$ and $\psi_i$ are suitable phases. It is then an exercise in classical electromagnetism to verify that the time-averaged (classical) electromagnetic energy radiated (on classical spacetime) per unit time by this variable magnetic moment is given, in natural units, by $$\label{eq:power}\begin{split} \frac{d \cE}{dt} &= \frac{2}{9}\sum_{\omega_i = \omega_j}\omega_i^4 ( {\boldsymbol{M}}_i \cdot {\boldsymbol{M}}_j) \cos(\psi_i-\psi_j)\\ &= \frac2 9 e^2\lambda_P^4 M^2 R^2\omega^6\left(1+\sin^2 \theta +\frac1 2 \omega^2 R^2 \sin^2 \theta\right). \end{split}$$ Therefore averaging over the unknown direction of ${\boldsymbol{e}}$ we get $$\label{eq:power2} \frac{d \cE }{dt} = \frac{8\pi}{27} e^2\lambda_P^4 M^2 R^2\omega^6 (5+ \omega^2 R^2) \simeq e^2 \lambda_P^4 M^2 R^2 \omega^6 \simeq e^2 \left(\frac{\tau _P}{T} \right)^6 \left( \frac{R}{\lambda_P}\right)^2 M^2,$$ where in the second equation we neglected numerical constants of order 1, and took into account that typically $\omega R \simeq 10^{-1}$ or smaller. Taking then $\cE \simeq M \simeq N m$, $N$ of the order of the number of particles in an object of the size of the sun and density of liquid water, i.e., roughly $10^{56}$, $m \simeq 1 \,GeV$, the rotation period $T = 10^{-2} \,s$ and $R = 10^3\, km$ (comparable to the GW150914 parameters), and recalling that the Planck time $\tau_P \simeq 10^{-44}s$, we get that the fraction of energy radiated by the body per unit time is $$\frac{1}{\cE}\frac{d\cE}{dt} \simeq 10^{-89} s^{-1},$$ and it is therefore negligible. To make this sizable $T$ ought rather to be of Planckian order, which would probably mean that our object collapsed into a black hole and no radiation is visible, and anyway the above Minkowskian picture does not apply. This computation is certainly too primitive, but it suggests that the fraction of the total mass emitted as em radiation can be expected to be negligible; by far nothing comparable with the fraction of a few percents emitted as GW in the BBH collapse GW150914. But could a more cautious approach [*reverse*]{} this conclusion? The question is legitimate, since a heuristic argument, whose qualitative consequences are confirmed by a more cautious analysis [@DMP], suggests that near singularities the [*effective Planck length might diverge*]{} as $\lambda_P g_{00} ^{-1/2}$. This might well introduce a metric dependent factor in our formula for the electromagnetic radiation caused by the magnetic moment of neutral matter, making it considerably larger in the last instants before the collapse into a black hole, heuristically as $$\frac{d \cE}{dt} = \frac{1}{g_{00}^{2}} e^2 \lambda_P^4 M^2 R^2 \omega^6,$$ where $g_{00}$ is the time-time component of the background metric. This qualitative conclusion is supported by the results in [@DMP], which mean in particular that in a flat Friedmann- Robertson-Walker (FRW) background (which is spherically symmetric with respect to every point), with metric, in spatial spherical coordinates, $ds^2 = - dt^2 + a(t)^2[dr^2 + r^2dS^2]$, the size of a localisation region centered around an event at cosmological time $t$, measured by the radial coordinate $r$, must be at least of order $\lambda_P a(0) /a(t)$, $t = 0$ being the time of the present epoch. The situation that we have in mind, namely that of a neutral object rotating in the gravitational field of a collapsing one, is of course better described by a Schwarschild metric than by a FRW one. We note that the metric of a collapsing homogenous sphere of dust is given by the Oppenheimer-Snyder solution [@OS; @BS], which is a Schwarzschild metric outside the sphere, matched with a closed FRW metric $$\label{eq:OSmetric} ds^2 = -dt^2 + a(t)^2(d\chi^2 + \sin^2 \chi d\Omega^2)$$ inside it. The scale factor $a(t)$ in the above metric can be expressed parametrically through the conformal time $\eta$ $$\begin{aligned} a(t(\eta)) &= \frac12\sqrt{\frac{R^3_0}{2GM_0}}(1+\cos \eta), \label{eq:OSa}\\ t(\eta) &= \frac12\sqrt{\frac{R_0^3}{2GM_0}}(\eta + \sin \eta), \label{eq:OSt}\end{aligned}$$ with $M_0>0$ the ADM mass of the collapsing sphere and $R_0 \geq 2GM_0$ its initial areal radius. The conformal time at which the sphere is completely inside its Schwarzschild radius is given by $\eta_0 = \cos^{-1}(4GM_0/R_0 -1)$. The continuous match between the exterior Schwarzschild metric and the interior FRW one and the results of [@DMP] recalled above seem therefore to justify the ansazt of replacing $\lambda _P$ in  by $ \lambda _P a(0)/a(t)$. Indeed such an expression for the effective Planck length converges to the usual value $\lambda_P$ in the limit $M \to 0$ in which the FRW metric becomes Minkowski, as one can easily verify by eliminating the conformal time $\eta$ from , . Then, the above formula for the radiation of a precessing *neutral* object would become $$\label{eq:effectivepower} \frac{d \cE}{dt} = e^2 \left(\frac{\lambda_P a(0)}{a(t)} \right)^4 M^2 R^2 \omega^6.$$ This energy has to be emitted of course at the cost of the kinetic energy of the rotating object due to spin, precession and orbital rotation, as well as of its potential energy, causing a faster inspiraling. For simplicity, we will consider here only the orbital kinetic term, and then $$\label{eq:conservation} \frac{d}{dt}\left(\frac12M R^2 \omega^2\right) = - \frac{d \cE}{dt},$$ which entails that the total radiated energy can be estimated as $$\label{eq:totalenergy} \cE = \frac12M R^2 \left[\omega_0^2 - \omega(t_{collapse})^2\right],$$ where the integration cannot be extended beyond the hiding of our object within the event horizon of the other. For the radiated power according to our formulae near the singularity would diverge, but would remain trapped and would not be visible from outside. According to , the total radiated energy can be sizable only if $\omega(t_{collapse}) \ll \omega_0$. In order to check if this is the case in typical situations, we solve , that, inserting , becomes $$\dot \omega = - e^2(\lambda_P a(0))^4M \frac{\omega^5}{a(t)^4},$$ which can be integrated by separation of variables. To this end, we note that, by , $dt/d\eta = a$, and therefore $$\int_{- \infty}^{t_{collapse}} a(t)^{-4} dt = \int_0^{\eta_0} \frac{d\eta}{a(t(\eta))^3}= 8\left(\frac{2GM_0}{R^3_0}\right)^{3/2}\int_0^{\eta_0}\frac{d\eta}{(1+\cos \eta)^3}.$$ Defining then, for $\eta \in [0,\pi)$, $$F(\eta) := \int_0^{\eta}\frac{dx}{(1+\cos x)^3} = \frac{\sin \eta(6\cos \eta+\cos(2\eta)+8)}{15(1+\cos \eta)^3},$$ we obtain, neglecting numerical constants of order 1, $$\label{eq:omegacoll} \omega(t_{collapse})^2 = \left[\frac{1}{\omega_0^4}+e^2 \lambda_P^4M \left(\frac{R_0^3}{2GM_0}\right)^{1/2}F(\eta_0)\right]^{-1/2},$$ which is smaller than $\omega_0^2$, as it should. One can then observe that for $M_0 \to 0$ one has $\eta_0 = \cos^{-1}(4GM_0/R_0 -1) \sim \pi -\sqrt{8GM_0/R_0}$ and therefore $F(\eta_0) \sim \frac85(R_0/8GM_0)^{5/2}$, so that $$\omega(t_{collapse})^2 \sim \frac{G^{3/2}M_0^{3/2}}{e\lambda_P^2 MR_0^2 },$$ would actually be very small with respect to $\omega_0^2$, making the total radiated energy  non negligible. (Note that for ordinary matter the collapse would stop much before that the matter itself is hidden inside the horizon, due to the non vanishing pressure). Moreover, this effect might disappear if one takes properly into account the red-shift of the radiation emitted near to the horizon. This could probably be done by using the general relativistic version of the radiated power by a magnetic dipole instead of . Conversely, for finite values of $M_0$, one can expand  due to the smallness of $\lambda_P^4$, and obtain for the total radiated energy $$\cE \simeq e^2\lambda_P^4M^2 R^2\omega_0^6 \left(\frac{R_0^3}{2GM_0}\right)^{1/2}F(\eta_0).$$ Thus we see that for $2GM_0/R_0 = 1, F(\eta_0)$ vanishes, as it should, since the collapse takes place at the beginning. If instead $2GM_0/R_0$ is smaller than $1$ but of that order, then $F(\eta_0)$ is also of the same order: e.g., if $2GM_0/R_0 = 1/2$ then $F(\eta_0) = 7/15$. Moreover, if we take as before $M_0 \simeq M \simeq 10^{56} GeV = 10^{37} M_P \simeq \cE _0$ and we recall that for $\hbar = c = 1$ we have $G=M_P^{-2}$, we deduce $R_0 = 4GM_0 \simeq 10^{37} M_P^{-1} \simeq 10^{-1}\,km$, so that, assuming again $R \simeq 10^3 km \simeq 10^{41} M_P^{-1}$ and $T \simeq 10^{-2} s \simeq 10^{42} \tau_P$, we get $$\cE \simeq e^2 \lambda_P^4 M^2 R^2R_0 \omega_0^6 \simeq e^2 \lambda_P^4 M R^2R_0 \omega_0^6 \cE _0 \simeq 10^{-96} \cE _0 \simeq 10^{-40}\,GeV.$$ Note that using the same figures and multiplying the fraction of the total energy emitted as electromagnetic interaction per second, as given by the previous more brutal computation, Eq. , by the collapse time $t_{collapse} = t(\eta_0) = \frac1{\sqrt{2}}(\frac\pi 2-1)R_0 \simeq 10^{37} \tau_P \simeq 10^{-7} s$, we get an estimate of exactly the same order of magnitude. Thus, as noticed earlier on in this discussion, the fraction of the mass converted into electromagnetic radiation is negligible, unless the period $T$ is at the Planck scale, which probably means that collapse took place, and that the emitted radiation is not visible to distant observers. As already mentioned, however, a more realistic estimate ought to treat relativistically the electromagnetic emission. Eventually, another possibility, both more and less favorable, would be offered by a compact spinning concentrate of dark matter interposed to some distant source; spin and concentration apart, these objects exist and are revealed to us by gravitational lensing. Which results from the gravitational deflection of photons experimentally known since a century. But if the source emits also charged particles, say electrons, sufficiently energetic to reach us within a reasonable delay after the $\gamma$ rays, their deflection ought to be modified by the magnetic field caused by the moment of our stellar object, due to Quantum Spacetime. A sort of QST - Northern Light phenomenon. One might hope that this is a “more favorable" situation with respect to the one considered above because, while the energy emitted is proportional to the fourth power of the Planck length, the deviation we are mentioning now would be only $quadratic$ in $\lambda_P$. Nevertheless, a rough estimate of the deviation angle $\theta$ of an electron by a compact object of mass $M$ and radius $R$ spinning at angular velocity $\omega$ gives, using , $$\theta \cong \frac{M_S}{m\gamma R^2} \cong \frac{e \lambda_P^2 M \omega}{m \gamma R},$$ with $m$ the electron mass and $\gamma = (1-v^2)^{-1/2}$. Choosing, as above, $M \simeq 10^{56} \,GeV$, $R \simeq 10^{-1}\, km$ and $\omega = 10^2 s^{-1}$ the deviation would be only $\theta \cong 10^{-34}$ for electrons of energy $1\, TeV$, which would reach us with a delay, with respect to photons, of a few hours if the source is $10^9$ light years distant. The delay for protons of $10^3 \,TeV$ (still considered to be lower than the GKW limit) would be the same, but the deviation would be $10^3$ times smaller. Of course the deviation would be more important for softer electrons, which however would reach the Earth when nobody is there any longer. Moreover, a “less favorable” aspect is that electrons are considerably influenced by the much stronger galactic magnetic field, of which a precise knowledge would be needed, together with a nearly exact location in the sky of the sources of electromagnetic radiation and of electrons, as well as a clear recognition of the coincidence of their origin. Concluding remarks ================== Our discussion was based on the choice  of the covariant derivative, with $e$ denoting the electron charge. This choice seems to be dictated by gauge invariance in a theory which includes the electromagnetic interactions of the electron, taking into account the noncommutativity of $\cG = \cU(M(\cE))$, the group of unitaries in the multipliers of the algebra of Quantum Spacetime $\cE$. On an $\cE$ bimodule, only the left (resp. right) action of $U$ (resp. $U^* $), or the trivial action, are allowed. This poses a problem for the Standard Model, apparently excluding quark fields. This problem has been noticed and discussed by several Authors, see e.g. [@CJSWW]. It deserves further discussions to see whether in our context the choice made in  is really the only choice. According to our preceding discussion, so far there seems to be no indication of visible effects of the quantum nature of spacetime at the Planck scale, except its role in solving the horizon problem [@DMP] and justifying from first principles part of the assumptions made in the inflationary scenario. The effects considered in this note are so tiny that it would be instructive to compare them with those due to the graviton mediated dark matter - photon interaction. Furthermore, the electromagnetic radiation emitted by a collapsing binary system due to the mechanism proposed here ought to be compared with the Hawking radiation. But the QST induced electromagnetic interactions of dark matter might be detectable in more exotic hypothetic astrophysical objects, like self gravitating Bose-Einstein condensates of dark matter, consisting of neutral scalar particles. The stability of such objects, with a solar mass and a radius of few dozen of kilometers, has been recently investigated, both in the isotropic and rotating cases; the possible formation of vortices has also been considered (cf., e.g., [@MT]). Smaller object of this nature were excluded in the quoted study by the nonrelativistic approximation used there, but might well be relevant to manifest sizable QST-electromagnetic effects, possibly also in the form of electromagnetic vortex-vortex interactions, which might potentially change the dynamics of these hypothetical objects. These points will be dealt with in subsequent studies. [99]{} D. Bahns, S. Doplicher, G.Morsella, G. Piacitelli, Quantum Spacetime and Algebraic Quantum Field Theory, in *Advances in Algebraic Quantum Field Theory*, R. Brunetti, C. Dappiaggi, K. Fredenhagen, J. Yngvason, Eds., Springer, 2015; arXiv:1501.03298. D. Bahns, S. Doplicher, K. Fredenhagen and G. Piacitelli, Ultraviolet finite quantum field theory on quantum spacetime, [*Commun. Math. Phys.*]{} [**237**]{} (2003), 221. D. Bahns, S. Doplicher, K. Fredenhagen, G. Piacitelli, Quantum geometry on quantum spacetime: distance, area and volume operators, [*Commun. Math. Phys.*]{} [**308**]{} (2011), 567, arXiv:1005.2130. T. W. Baumgarte, S. L. Shapiro, [*Numerical relativity: solving Einstein’s equations on the computer*]{}, Cambridge University Press, Cambridge (2010). R. Brunetti, K. Fredenhagen, T.-P. Hack, N. Pinamonti, K. Rejzner, Cosmological perturbation theory and quantum gravity, [*J. High Energ. Phys.*]{} **2016** (2016), 32, arXiv:1605.02573. X. Calmet, B. Jurco, P. Schupp, J. Wess, M. Wohlgenannt, The standard model on non-commutative space-time, [*Eur. Phys. J. C*]{} **23** (2002), 363, arXiv:hep-ph/0111115. S. Doplicher, K. Fredenhagen, unpublished S. Doplicher, K. Fredenhagen and J. Roberts The quantum structure of spacetime at the Planck scale and quantum fields, [*Commun. Math. Phys.*]{} [**172**]{} (1995), 187. S. Doplicher, G. Morsella, N. Pinamonti, On quantum spacetime and the horizon problem, [*J. Geom. Phys.*]{} **74** (2013), 196. LIGO Scientific Collaboration and Virgo Collaboration, Observation of gravitational waves from a binary black hole merger, *Phys. Rev. Lett.* **116** (2016), 061102. V. F. Mukhanov, H. A. Feldman, R. H. Brandenberger, Theory of cosmological perturbations, [*Phys. Rep.*]{} **215** (1992), 203. E. J. M. Madarassy, V. T. Toth, [*Phys. Rev. D*]{} **91** (2015), 044041, arXiv:1412.7152. J. Madore, S. Schraml, P. Schupp, J. Wess, Gauge theory on noncommutative spaces, [*Eur. Phys. J. C*]{} **16** (2000), 161, arXiv:hep-th/0001203. J. R. Oppenheimer, H. Snyder, On continued gravitational contraction, [*Phys. Rev.*]{} **56** (1939), 455. J. Zahn, [*Dispersion relations in quantum electrodynamics on the noncommutative Minkowski space*]{}, PhD thesis, Hamburg University (2007), arXiv:0707.2149. [^1]: Note that $\cL_I$ would vanish also on Quantum Spacetime if the products appearing were interpreted as quantum Wick products [@BDFP:2003], as $E^{(n)}(f_1(q_1)\dots f_n(q_n))$ is independent from the ordering of the factors, and the tensor factors in $A_\mu(q)$ and $\ph(q)$ acting on Fock space would commute again. But the Quantum Wick Product would anyway violate not only Lorentz invariance, but also Gauge Invariance, hence it could not be applied in the present context without first elaborating some radical modifications.
{ "pile_set_name": "ArXiv" }
--- abstract: 'We find the existence of sub-Planck scale structures in the P[ö]{}schl-Teller potential, which is an exactly solvable potential with both symmetric and asymmetric features. We analyze these structures in both cases by looking at the Wigner distribution of the state evolved from an initial coherent state up to various fractional revival times. We also investigate the sensitivity to perturbations of the P[ö]{}schl-Teller potential and we verify that, similar to the harmonic oscillator, the presence of sub-Planck structure in phase space is responsible for a high sensitivity to phase-space displacements.' author: - 'Utpal Roy,$^{\mathrm{1}}$[^1] Suranjana Ghosh,$^{\mathrm{1}}$[^2] P. K. Panigrahi,$^{\mathrm{2,3}}$[^3] and David Vitali,$^{\mathrm{1}}$[^4]' title: 'Sub-Planck scale structures in the P[ö]{}schl-Teller potential and their sensitivity to perturbations' --- Introduction ============ The Wigner distribution $W(x,p)$ is a useful tool for visualizing quantum interference phenomena in phase space. Non-local superposition states show interference and the presence of phase space regions where $W(x,p)$ is negative is an unambiguous signature of quantum behavior. The oscillatory structures resulting from quantum interference can have a dimension much smaller than the Planck’s constant $\hbar$, and these sub-Planck scale structures were first found by Zurek in quantum chaotic systems [@zurek]. Sub-Planck scale structures usually appear as alternate small ’tiles’ of maxima and minima. The phase space area of these ’tiles’ are given by $\hbar^2$ divided by the effective phase-space area $A$ globally occupied by the state, which can be significantly larger than $\hbar$. These structures are very sensitive to environmental decoherence [@zurek; @ph; @wisn; @pathak1; @pathak2; @ghosh1] and may have important application in Heisenberg-limited measurements and quantum parameter estimation [@toscano; @dalvit]. A classical analogue of sub-Planck scale structures, regarded as sub-Fourier sensitivity, has been studied in Ref. [@praxmeyer], and these structures have also been found in diatomic molecular systems [@ghosh], entangled cat states [@manan], optical fibers [@fibre], and Kirkwood-Rihaczek distribution [@jay]. Recently, a connection between these structures and teleportation fidelity has been also established [@scott]. Sub-Planck structures have been extensively studied in the case of superpositions of harmonic oscillator coherent states and also for superpositions of generalized coherent states which occur in the time evolution of systems with nonlinear potentials at fractional revival times. It is therefore important to extend the study of this phenomenon to other nonlinear potentials in order to investigate which are the relevant parameters affecting this subtle quantum behavior, and if Heisenberg-limited sensitivity can be approached also in other nonlinear systems. Here we focus our attention to the Pöschl-Teller (PT) potential, which has been applied to model various situations in recent years [@tong1; @tong2; @indjin; @tomak1; @tomak2]. In fact, the PT potential has been used in the context of deep quantum wells, as well as for modeling optical systems with changing refractive index. The PT potential depends upon two parameters and by varying them one can switch from a symmetric to an asymmetric situation. Yildirim and Tomak studied in particular the nonlinear optical properties associated to the PT potential [@tomak1], and also the nonlinear changes in the refractive index resulting from its tunable asymmetry [@tomak2]. Another important feature of the PT potential is that it is exactly solvable and characterized by a quadratic spectrum: as a consequence its time evolution shows fractional revivals [@averbukh; @robinett] and it may lead to the generation of linear superpositions of well distinct states in phase space. In this paper we confirm this expectation and find that an initial coherent state of the PT potential shows sub-Planck structures in phase-space at fractional revival times. We shall first study these structures when the PT potential is symmetric and then, by tuning the potential parameters, we switch to the asymmetric case in order to see the effects of the asymmetry on sub-Planck scale structures. In both cases we verify quantitatively that the dimension of the sub-Planck structures scales as $\hbar^2/A$, as expected, also for the PT potential. We then study the sensitivity of sub-Planck scale structures to perturbations, by considering two different situations: i) “internal” perturbations, i.e., a small change of the potential which is turned from symmetric into a slightly asymmetric one; ii) “external” perturbations, as for example a weak force, yielding a small phase space displacement of the state. Both cases will confirm that also in the PT potential, sub-Planck structures are responsible for a very high sensitivity to displacements. The paper is organized as follows. In Sec II, the symmetric and asymmetric properties of PT potential are reviewed and the revival dynamics of a coherent state (CS) wave-packet are discussed. Existence of sub-Planck scale structures in phase space is demonstrated for both symmetric and asymmetric cases in Sec III. In Sec IV we discuss the sensitivity of sub-Planck scale structures due to a small asymmetry of the potential and then the one associated with a small phase-space displacement of the state. We end with some conclusions in Sec V. ![The Pöschl-Teller potential V(x) (in unit of $10^4\;a.u.$) for different parameter values. Dotted line represents the symmetric case ($\rho=\kappa=50$), solid line refers to the slightly asymmetric case ($\rho=50$, $\kappa=46$), while the strongly asymmetric case ($\rho=50$ $\kappa=6$) is shown by the dashed line. Here, the variable $x$ is in atomic units and we have chosen $\alpha=2\;a.u$. []{data-label="potential"}](fig1.eps){width="3.2in"} Pöschl-Teller potential and the CS wave-packet ============================================== The general PT potential [@posh] is given by $$V(x)\;=\;\frac{\hbar^2\alpha^2}{2m}\left[\frac{\rho(\rho-1)}{\cos^2\alpha x}+\frac{\kappa(\kappa-1)}{\sin^2\alpha x} \right] \label{PT}$$ and it represents an array of potential wells in the whole line. The symmetric case can exhibit energy bands [@ranjani]. Since the wells are impenetrable, it suffices to consider only one of them for a full quantum mechanical description of the system. The potential is depicted in Fig. \[potential\] for different values of $\rho$ and $\kappa$ $(\rho,\kappa>1)$ (units are chosen such that $\hbar=m=1$), and we have selected the well within the interval $[0,\pi/2\alpha]$. The dotted line corresponds to the symmetric well $(\rho=\kappa=50)$, centered around $\pi/4\alpha$. This symmetric well smoothly takes the form of an infinite well in the limit of $\rho,\kappa\rightarrow 1.$ The potential loses its symmetry with respect to the axis of the well if $\rho \neq \kappa$. Small and large asymmetries are shown respectively by the solid and dashed curves in Fig. \[potential\]. The minimum of the well is inclined towards the left side when $\rho>\kappa$. The PT potential is one of the exactly solvable quantum mechanical potentials. The corresponding Schrödinger equation can be solved to obtain the energy eigenvalues and eigenfunctions [@klauder] $$\begin{aligned} E_{n}&=&\frac{\hbar^2\alpha^2}{2m} (2n+\rho+\kappa)^2,\quad n=0,1,2,... ~~\nonumber\\ \psi_{n}(x)&=&N \cos^{\rho}(\alpha x) \sin^{\kappa}(\alpha x) P_n^{\rho-1/2,\kappa-1/2} [\cos(2\alpha x)],\nonumber\\ \label{PTeigenfunction}\end{aligned}$$ where $P_n^{\rho-1/2,\kappa-1/2} [\cos(2\alpha x)]$ is the Jacobi polynomial and $N$ is the normalization constant: $$N=\left[\frac{2 \alpha \Gamma{(n+1)}(2n+\rho+\kappa)\Gamma{(n+\kappa+\rho)}}{\Gamma{(n+\rho+1/2)} \Gamma{(n+\kappa+1/2)}}\right]^\frac{1}{2}.$$ It is worth mentioning that the PT potential has an underlying $SU(1,1)$ dynamical symmetry algebra, which admits an infinite number of energy levels. Some of them can be excited to create a CS wave-packet by using an appropriate laser. These states have many classical features and tend to preserve their properties even for long evolution times. Here, we consider a displacement operator CS [@perelomov] of this potential. Construction of this CS makes use of the exponential form for the solution of hypergeometric differential equation and the correct choice of $SU(1,1)$ generators [@guru1]. The CS depends upon a parameter $\beta$ and can be written as [@utpal] $$\chi_{\beta}(x)\;=\;\sum_{n=0}^{\infty}\;\;d_{n}^{\beta}\;\psi_{n}(x),$$ where the amplitudes $d_{n}^{\beta}$ are given by $$d_{n}^{\beta}\;=\;(-\beta)^n \left[\frac{\Gamma{(\rho+n+1/2)}\Gamma{(\kappa+n+1/2)}}{2 \alpha \Gamma{(\kappa+\rho+n)}\Gamma{(n+1)}(2n+\kappa+\rho)}\right]^{1/2}.$$ The eigenenergy $E_{n}$ has both linear and quadratic terms in the quantum number $n$. The time evolution of the CS generated by the PT Hamiltonian, $$\chi_{\beta}(x,t)=\sum_{n=0}^{\infty}d_{n}^{\beta}\psi_{n}(x)e^{-iE_nt}, \label{timeevolution}$$ is such that, at short times, the center of the wave-packet reproduces the classical motion, due to the linear term in $n$ in the energy eigenvalues. The quadratic term in $n$ is instead responsible for revival and fractional revivals [@averbukh; @robinett], taking place on a longer time scale. In fact, it is easy to check that after the revival time $T_{rev}=\pi/\alpha^2$ (or integer multiples of it), the initial coherent state localizes back to its form after spreading. Here however, we are interested in fractional revivals, which instead take place after a time interval $(r/s)T_{rev}$, with $r$ and $s$ mutually prime integers. At fractional revival times the initial wave-packet partially localizes into a superposition of spatially distributed sub-packets, each of which closely resembles the initial wave-packet: when $s$ is even, the wave-packet breaks into $s/2$ wave-packets, when instead $s$ is odd, the state is a superposition of $s$ distinct states. This means that one obtains a Schrödinger cat state, superposition of two CS with opposite phases, at one fourth of the revival time, while four-way break up, or the so called “compass-like” state emerges at $T_{rev}/8$. As shown by Ref. [@zurek], the latter produces sub-Planck scale structures in the phase space Wigner distribution. The behavior at fractional revival times can be derived in a straightforward way from Eq. (\[timeevolution\]) in the particular case when both $\rho$ and $\kappa$ are even. When the PT potential is symmetric ($\rho=\kappa$), the property of the Jacobi polynomial allows us to write $$\begin{aligned} \label{symm} &&\psi_{n}(x)=(-1)^{n}\psi_{n}(\pi/2\alpha-x)\nonumber\\ \quad &or&\;\; \psi_{n}(\pi/4\alpha-x)=(-1)^{n}\psi_{n}(\pi/4\alpha+x).\end{aligned}$$ Using this fact at $T_{rev}/4$, a straightforward calculation yields $$\chi_{\beta}(x,T_{rev}/4)=\frac{1}{\sqrt{2}}\left[e^{-i\pi/4}\chi_{\beta}(x,0) + e^{i\pi/4}\chi_{\beta}(\frac{\pi}{2\alpha}-x,0)\right],$$ that is, the CS is split into two parts at time $T_{rev}/4$ forming the well known Schrödinger cat state. Each element of the superposition is proportional to the initial wave-packet, even though they are situated at the opposite ends of the potential well. In the asymmetric case instead one gets $$\begin{aligned} &&\chi_{\beta}(x,T_{rev}/4)=\chi_{\beta}^{e}(x,0)- i\chi_{\beta}^{o}(x,0);\;\;\quad \eta=\rho+\kappa=\mathrm{multiple\;of}\; 4\nonumber\\ \mathrm{and}\;\;&&\chi_{\beta}(x,T_{rev}/4)=-i\chi_{\beta}^{e}(x,0)+ \chi_{\beta}^{o}(x,0);\;\;\quad\eta=\mathrm{multiple\;of}\; 2,\end{aligned}$$ where $\chi_{\beta}^{e,o}(x,0)=\sum_{n,\; even,odd} d_n \psi_n (x)$. For the analysis at time $T_{rev}/8$, it is instead convenient to define a classical wave-packet [@averbukh; @robinett]: $$\chi_{\beta}^{cl}(x,t)=\sum_{n=0}^{\infty}d_{n}^{\beta}\psi_{n}(x)e^{-2\pi i n t/T_{cl}}, \label{cl}$$ which evolves with the classical periodicity $T_{cl}=\pi/(\rho+\kappa)\alpha^{2}$. It can be easily seen that at $t=T_{rev}/2$, the CS will revive with an extra half a period ($T_{cl}/2$) of phase. In other words, the CS will be situated on the other side of the potential well, but with the same amplitude. The wave-packet splits into two parts at time $t=T_{rev}/4$ with a relative phase of half a period $T_{cl}/2$. Four way break-up occurs instead at time $t=T_{rev}/8$, and one can see that the state at this fractional revival time becomes $$\label{oneeightcl} \chi_{\beta}(x,T_{rev}/8)=\frac{1}{2}\left[e^{-i\pi/4}\chi_{\beta}^{cl}(x,t) +\chi_{\beta}^{cl}(x,t+T_{cl}/4)-e^{-i\pi/4}\chi_{\beta}^{cl}(x,t+T_{cl}/2)+ \chi_{\beta}^{cl}(x,t+3T_{cl}/4)\right].$$ One can generalize in a straightforward way to a generic fractional revival time. In fact, in this case the state can be written as $$\chi_{\beta}(x,t=\frac{r}{s}T_{rev})=\sum_{p=0}^{l-1}a_{p}\chi_{\beta}^{cl}(x,t-\frac{p}{l}T_{cl}),$$ where $a_{p}$ gives the amplitude with probability $|a_{p}|^{2}$ of each $l$ number of sub-wave-packets or clones of the original wave-packet. Each clone differs in phase from the original one by $T_{cl}/l$. As illustrated above, the six way break-up can be observed at the time $T_{rev}/12$. ![image](fig2.eps){width="7.in"} Wigner distribution and sub-Planck scale structure ================================================== The Wigner quasi-probability distribution provides a clear phase-space description of a quantum state [@schleich]. It is particularly suited for displaying quantum interference phenomena, which are associated to oscillations and to the presence of regions where the Wigner function attains negative values. From its definition, the Wigner function of the CS evolved up to time $t$ of Eq. (\[timeevolution\]) is given by $$W(x,p,t)=\frac{1}{\pi\hbar}\int_{-\infty}^{\infty}\chi_{\beta}^{*}(x-z,t)\chi_{\beta}(x+z,t) e^{-2ipz/\hbar}dz.$$ The PT potential is an infinite array of identical impenetrable wells of width $\pi/2 \alpha$ and therefore the wave-function can be taken equal to zero outside the interval $[0,\pi/2\alpha]$. By imposing this fact, one finds that the following limits of integration can be taken in the above integral: in the left half of the well $(0\leq x \leq\pi/4 \alpha)$, one has $-x\leq z \leq+x$, while in the right half $(\pi/4 \alpha\leq x \leq\pi/2 \alpha)$, one has $-(\pi/2 \alpha-x)\leq z \leq(\pi/2 \alpha-x)$. Let us now analyze the Wigner function of the time-evolved CS wave-packet at different fractional revival times. Symmetric case -------------- First we choose the symmetric case with parameter values $\rho=\kappa=50$ and $\alpha=2\;a.u.$ The CS parameter is taken as $\beta=0.6$, which corresponds to an initial wave-packet with an energy distribution peaked around the level $\bar{n}=12$. The Wigner functions of the CS wave-packet at different fractional revival times in phase space are shown in Fig. \[wigSymAll\]. At time $t=T_{rev}/4$ (see Fig. \[wigSymAll\](a)) the state is a cat-like state and its Wigner function can be explicitly written in the following two equivalent ways $$\begin{aligned} W(x,p,T_{rev}/4)&=&\frac{1}{2}\left[W(x,p,0)+W(\frac{\pi}{2\alpha}-x,p,0)\right]-2 Im(W_{oe}(x,p))\label{Woe}\\ &= &\frac{1}{2}\left[W(x,p,0)+W(\frac{\pi}{2\alpha}-x,p,0)\right]+\sum_{m,n}d_m d_n (-1)^m Im(W_{mn}(x,p))\nonumber .\end{aligned}$$ The first two terms in Eq. (\[Woe\]) are the two localized CS Wigner functions, situated at the left and right sides of the well. The interference fringes are associated to the third terms of the two expressions above, where the following quantities appear $$\begin{aligned} \nonumber W_{oe}(x,p)=\frac{1}{\pi\hbar}\int_{-\infty}^{\infty}\chi_{\beta}^{o}(x-z,0)\chi_{\beta}^{e}(x+z,0) e^{-2ipz/\hbar}dz,\\ W_{mn}(x,p)=\frac{1}{\pi\hbar}\int_{-\infty}^{\infty}\psi_{m}(x-z)\psi_{n}(x+z,0) e^{-2ipz/\hbar}dz.\nonumber\end{aligned}$$ The oscillatory interference ripples are parallel to the line joining the two split sub-CSs, i.e., the interference pattern is oscillatory only in $p$ space and also squeezed. The number of ripples decreases as the two sub-CSs come closer. At $T_{rev}/8$ (Fig. \[wigSymAll\](b)), the initial wave-packet splits into four distinct sub-wave-packets (Eq. (\[oneeightcl\])). The Wigner function contains four distinct parts $W_{1},W_{2},W_{3}$ and $W_{4}$ (see Fig. \[wigSymAll\](b)) and six interference terms, among which two diagonal partners ($W_{13}$ and $W_{24}$) overlap at the center of the phase space and generate a smaller chess-board interference pattern. This structure is the signature of sub-Planck scale structures [@zurek; @ghosh] for a compass-like state. The central interference pattern can also be seen as the superposition of interferences of two orthogonally situated cat states, and it is formed by small “tiles” much smaller than the individual CS peaks. ![image](fig3.eps){width="7.in"} Finally, at $t=T_{rev}/12$, the initial wave-packet splits into six CSs (i.e., three pairs of cat-like states), forming a benzene-like structure (Fig. \[wigSymAll\](c)). Sub-Planck structures also appear here in the center due to the superposition of three interference regions. A visual analysis shows that the dimension of the sub-Planck scale “tiles” of the compass-like state (Fig. \[wigSymAll\](b)) is smaller than that of the benzene-like state (Fig. \[wigSymAll\](c)). The reason behind it is that the area of overlap increases with the number of split CSs, but we will perform a more quantitative study in the following subsection. Asymmetric case --------------- We now discuss sub-Planck scale structures for a strongly asymmetric potential. The potential parameter values are taken as $\rho=50$, $\kappa=6$ and $\alpha=2\;a.u.$, as in Fig. \[potential\] (dashed line). For the sake of comparison with the symmetric case, we choose the CS parameter ($\beta=0.88$), such that the energy distribution is again peaked around $\bar{n}=12$, as in the symmetric case. In spite of this, the interference structures are now very different. The features of the case with large asymmetry is illustrated in Fig. \[wigAsym\]. The cat-like, compass-like, and the benzene-like states become asymmetric and inclined towards the left boundary of the well. Since the potential well is of infinite depth, the stiffness of the potential stretches or squeezes the wave-packet near the left side of the well. Moreover it spreads in $x$ space and becomes more squeezed in $p$ space. The asymmetry affects also the interference pattern of the states: the interference ripples are not anymore placed at the center, the diagonal interference patterns overlap partially and destroy the chess-board and benzene-like structures. Moreover, the interference pattern and the localized sub-CSs partially overlap and become much less distinguishable. Quantitative analysis of the sub-Planck scale structures -------------------------------------------------------- Let us now make a quantitative analysis and check if the smallest structures in the interference patterns, e.g., the alternate positive and negative tiles have really sub-Planck scale dimensions. Therefore we have to verify if the phase-space area occupied by the tiles ‘$a$’ scales, as predicted, as $\sim\hbar^2/A$ ($\sim 1/A$ in atomic units), where $A$ is the classical action of the state. $A$ is approximately given by the product of the effective support of its state in position and momentum: $A\sim \Delta x\times \Delta p$, where $\Delta x=\sqrt{\langle x^2 \rangle - \langle x \rangle^2}$ and $\Delta p=\sqrt{\langle p^2 \rangle - \langle p \rangle^2}$. .1in ------------ ------- ------- ------- ------- ------- ------- ------- -------- -------- -------- ------- 0.748 0.867 1.189 1.904 3.175 5.58 9.225 13.123 18.349 23.256 30.03 \[-1.0ex\] 1.41 1.2 0.859 0.57 0.332 0.194 0.11 0.078 0.057 0.045 0.035 \[-1.0ex\] ------------ ------- ------- ------- ------- ------- ------- ------- -------- -------- -------- ------- \[scaling\] In order to verify the scaling law, we first choose a symmetric potential ($\rho=\kappa=50$) and study the interference structures for CSs with different weighting distributions. More precisely, we vary the parameter $\beta$ from $0.8$ to $0.3$, which corresponds to change the position of the peak of the energy distribution from $n=32$ to $n=2$. The classical action progressively decreases for decreasing CS mean energy, while at the same time the size of the sub-Planck interference structure becomes larger. This behavior is quantitatively confirmed by Table I, where the first row gives the total phase space area measured form the uncertainty product, while the second row shows the area of the structures measured from the Wigner plots. These data are plotted in the $log_e$-$log_e$ plot of Fig. \[loglog\], where they are well fitted by a straight line of slope $-1.02$. This latter plot clearly confirms the expected scaling $a\sim 1/A$ with a very good approximation, confirming also that the interference pattern has sub-Planck dimensions. We have then extended the analysis of the size of the sub-Planck structures to situations with different symmetry, by comparing the expected dimension of the tiles from the measured phase-space area $A$ ($a \sim 1/A$), and the actual size of the tiles directly obtained from the plot of the Wigner function, for the case of the asymmetric PT potential, and for the benzene state evolved up to time $T_{rev}/12$. The results are illustrated in Table II. We have compared the compass-like and the benzene-like states for the symmetric PT potential (rows $1-2$) and for two different asymmetric cases (rows $3-6$). We find a good agreement between the expected and the actual dimension of the sub-Planck tiles in all six cases and this fact suggests that the scaling $a \sim \hbar^2/A$ remains valid also in the asymmetric potential and for the benzene-like state, even though the range of values of $a$ considered in Table II is too small to allow us to make a firm quantitative statement [@note]. We find the smallest phase-space sub-Planck structure for the compass-like state of the symmetric potential (row 1 and Fig. \[wigSymAll\](b)). This is a consequence of two facts that can be easily inferred from our analysis: i) the phase-space area of the tiles is always larger in the case of the benzene-like state; ii) the phase-space area of the tiles *increases* for increasing asymmetry of the PT potential. In the next Section, following the results of [@toscano; @dalvit] we will study the sensitivity to perturbations of the PT sub-Planck structures. One expects to have the highest sensitivity with a state possessing the finest structures in phase space and therefore the above results suggest to choose the compass-like state in the symmetric potential case for this study. ![$Log_e$-$log_e$ plot of the classical action $A$ ($a.u.$) against the area of sub-Planck scale structures ($a$ in $a.u.$) in the Wigner function of the compass-like state for the symmetric PT potential with $\rho=\kappa=50$. The expected scaling law, $a\sim 1/A$, is well verified, since the data are well fitted by the line with slope $-1.02$.[]{data-label="loglog"}](fig4.eps){width="3.in"} ![image](fig5.eps){width="5.2in"} -.02in .1in [|c|c|c|c|c|c|]{} & & & & &\ \[-1.0ex\] & $\rho$ & $\kappa$ & & &\ \[-1.0ex\] & & & $T_{r\!e\!v}$)& &\ 1 & 50 & 50 & 1/8 & 0.1084 & 0.110\ 2 & 50 & 50 & 1/12 & 0.1420 & 0.144\ 3 & 50 & 34 & 1/8 & 0.1278 & 0.132\ 4 & 50 & 34 & 1/12 & 0.2463 & 0.250\ 5 & 50 & 22 & 1/8 & 0.2118 & 0.225\ 6 & 50 & 22 & 1/12 & 0.2878 & 0.290\ \[table\] -.07in Sensitivity of sub-Planck scale structures ========================================== An interesting feature of sub-Planck structures is that they signal the fact that the corresponding quantum state is extremely sensitive to perturbations. We now verify if this holds also for the PT sub-Planck structures found above. Here we consider two different kinds of perturbation: i) an “internal” perturbation, corresponding to a slight change of the PT potential, which is changed from symmetric ($\rho=\kappa$) to a slightly asymmetric one ($\kappa \lesssim \rho$); ii) an “external” perturbation, i.e., a phase-space displacement of the state, for example due to the application of a weak force. As discussed in the preceding section, we consider an initially symmetric PT potential and the compass-like state at the fractional revival time $T_{rev}/8$. **I.** We set the potential parameter values $\rho=50$, $\kappa=46$ and $\alpha=2\;a.u.$, which give a very small asymmetry (see the solid line in Fig. \[potential\]). We compare this case with the symmetric case $\rho=\kappa=50$. We observe that the small asymmetry does not cause an appreciable modification on the phase space behavior of the compass-like state as a whole (see the contour plots in Fig. \[planck\](a) and Fig. \[planck\](c) for the symmetric and slight asymmetric case respectively). However, this is no more true if we look at the effect of the slight asymmetry on the interference pattern and we zoom in the center of the phase space plot (see Fig. \[planck\](b) and Fig. \[planck\](d)). The effect of the potential asymmetry on the interference pattern is equivalent to a phase-space shift, because one has a displacement of a single tile such that, approximately, the maximum of a tile comes to coincide with a zero of the unperturbed state. This fact just confirms that one has an extreme sensitivity with respect to small changes of the potential because the two states becomes quasi-orthogonal to a very good approximation ($\sim 0.1$). In fact, one has that $$|\langle\chi_{\beta}^{\mathrm{sym}}|\chi_{\beta}^{\mathrm{asym}}\rangle|^2\!\!=\!\!\int\!\!\!\!\int \!W_{\mathrm{sym}}(x,p) W_{\mathrm{asym}}(x,p)\; dx\; dp;\label{wigoverlap}$$ when the two interference patterns are displaced by a tile, one has a destructive interference effect, the above integration tends to zero and the two states become quasi-orthogonal. Thus, the size of sub-Planck structures sets a sensitivity limit on probing small potential modifications. ![Oscillations of the central sub-Planck region in $x$-space at $p=0$. Solid line refers to the symmetric case ($\rho=\kappa=50$), while the dotted line refers to the slightly asymmetric case ($\rho=50$, $\kappa=46$) as shown in Fig. \[planck\]. Here, $x$ is in atomic unit.[]{data-label="2dsmall"}](fig6.eps){width="3.2in"} This sensitivity is also visualized in Fig. \[2dsmall\], showing the section at $p=0$ of the Wigner function of the states at $T_{rev}/8$ for the two potentials. The Wigner distribution is negative, due to the non-classicality of the state, and a shift of the maximum of one curve to a minimum of the other is clearly visible. This means a sub-Planck scale shift along $x$ ($\sim 1/\Delta p$) of about $0.01247\;a.u.$, consistent with that in Fig. \[planck\]. A small difference in amplitudes due to the asymmetric effect is also captured in this enlarged $2$D view. **II.** The perturbation studied above is difficult to implement in practice because one has to change to potential by keeping its PT form. In typical experimental situations a small perturbation can be applied through a weak constant force, which will physically shift the state in phase space. Therefore we consider what happens if we shift in phase space the compass-like state of a symmetric PT potential. We apply the displacement operator $\exp(\lambda \; K_{+} -\lambda^{*} \; K_{-})$ on the CS of Eq. \[timeevolution\], where $\lambda$ is the displacement parameter and $K_{+}$,$K_{-}$ are the SU(1,1) generators [@utpal]. The overlap between the initial and final states in terms of the Wigner distribution is $$|\langle\chi_{\beta}^{\mathrm{CS}}|\chi_{\beta^{'}}^{\mathrm{DCS}}\rangle|^2=\int\!\!\!\int W_{\mathrm{CS}}(x,p) W_{\mathrm{DCS}}(x,p) dx\;dp,\label{wigoverlap1}$$ where $\chi_{\beta}^{\mathrm{CS}}$ is the initial CS evolved at time $t$ and $\chi_{\beta^{'}}^{\mathrm{DCS}}$ is the displaced coherent state (DCS) at the same time. After a long calculation we find the DCS and the above mentioned overlap can be written as $$\langle\chi_{\beta}^{\mathrm{CS}}|\chi_{\beta^{'}}^{\mathrm{DCS}}\rangle =\sum_{n,m,p}\textbf{D}_{m,n,p}^{\;\beta,\beta^{'}}\;\;e^{-i(E_n-E_{n-m+p})t},\label{overlap}$$ where $$\begin{aligned} \textbf{D}_{m,n,p}^{\;\beta,\beta^{'}}&=&\frac{\beta^{2n-m+p}(-\beta^{'})^{m}(\beta^{'})^{p}}{m!p!} \times \frac{\Gamma(2\rho+2n-m+p)\Gamma(\rho+n+\frac{1}{2})\Gamma(2\rho-m+p+1)\Gamma(\rho+n-m+p+ \frac{1}{2})}{\Gamma(n-m+1)\Gamma(\rho+n-m+\frac{1}{2})\Gamma(2\rho+p+1)\Gamma(2\rho+2n-2m+p)}\nonumber\\ && \times \frac{\Gamma(\rho+n-m+p+\frac{1}{2})}{\Gamma(2\rho+n-m+p)(n-m+p+\rho)}\;\; e^{\eta(n-m+\rho/2+1/4)}.\end{aligned}$$ Here $\lambda\!\!=\!|\lambda|e^{i\theta}$, $\beta^{'}\!\!\!=\!\tanh|\lambda|e^{i\theta}$ and $\eta=-2\log_e (\cosh|\lambda|)$, after using the normal form of the disentanglement formula for SU(1,1) algebra [@perelomov]. We choose a symmetric potential for $\rho=50$, $\alpha=2\;a.u.$ and an initial CS parameter $\beta=0.4$. We have plotted the overlap of Eq. \[overlap\] at time $T_{rev}/8$ (Fig. \[plot\]) versus the displacement parameter $\lambda$. One has damped oscillations which are again caused by the interference pattern of the compass-like state. The two states becomes quickly orthogonal, showing the sensitivity of the state to displacement. This oscillation is similar to the one shown by the overlap between the displaced and the undisplaced compass-like state of the harmonic oscillator (HO) case [@toscano], with the relevant difference that here in the PT case they decay very quickly. The oscillations for larger $\lambda$ are shown in the enlarged view in the lower inset of Fig. \[plot\]. This decay of the overlap oscillations is due to the fact that for physical systems with SU(1,1) or SU(2) algebras as our PT system, a displaced coherent state does not remain a coherent state, due to the higher order terms in the Baker-Campbell-Hausdorff development of the displacement operator. Therefore, the larger the displacement $\lambda$, the more distorted the coherent state, implying that the overlap is always close to zero when the displacement is not small. The HO instead obeys the Heisenberg–Weyl algebra ($[a,a^{\dagger}]=1$) and the displaced coherent state is still a coherent state, so that the overlap oscillations decay much more slowly. However the important point of our result is that, even though only one or two oscillations are visible, the period of oscillation, which gives the scale of quasi-orthogonality, is extremely small. This means that the PT system is very sensitive to phase-space displacement. Moreover this sensitivity is due just to the sub-Planck structures of the compass-state. In fact, the period of the overlap oscillation is $0.079\;a.u.$ (see the dotted arrow in Fig. \[plot\]), which is very close to the $x$-span of a sub-Planck structure of the compass-like state, equal to $0.075\;a.u.$, as shown by zoomed view of the Wigner function in the upper inset of Fig. \[plot\]. This means that shifting a single sub-Planck tile results in the cancellation between the positive and negative contributions of the two Wigner functions, thereby making the states quasi-orthogonal. ![(Color online) Plot of the overlap function of Eq. (\[overlap\]) versus the real part of the displacement parameter $\lambda$ (in $a.u.$) at the fractional revival time $T_{rev}/8$ for $\beta=0.4$. Here $\beta^{'}=\tanh |\lambda|\; e^{i\pi/4}$. The upper inset shows the central interference pattern of the Wigner function and the size of the sub-Planck structures. The lower inset shows an enlarged view of the oscillation for larger $\lambda$. The size of the sub-Planck structure is consistent with the period of oscillation of the overlap.[]{data-label="plot"}](fig7.eps){width="2.7"} Conclusions =========== We have shown that sub-Planck scale structures emerge in the time evolution at fractional revival times also in the PT potential. We have studied in particular the generation of compass-like states at $T_{rev}/8$ and benzene-like states at $T_{rev}/12$. We have studied both the symmetric and strongly asymmetric PT potential and we have verified that in all cases the phase-space area of the sub-Planck structures scales as $\hbar^2/A$, where $A$ is the classical action associated with the state. In particular we have seen the smallest phase-space structure (‘tile’) is associated with that of the compass-like state of the symmetric PT potential. For this reason we have considered this state for analyzing the sensitivity of the PT system to perturbations. We have considered the effect on this state of a slight asymmetry of the potential and of a phase-space displacement. We have seen that in both cases, the sub-Planck structures are responsible for high sensitivity, because as soon as a single tile is displaced in phase space so that its maximum coincides with a minimum of the undisplaced state, one has destructive interference and the two states become approximately orthogonal. [99]{} W. H. Zurek, Nature [**412**]{}, 712 (2001). Ph. Jacquod, I. Adagideli and C. W. J. Beenakker, Phys. Rev. Lett. [**89**]{}, 154103 (2002). D. A. Wisniacki, Phys. Rev. E [**67**]{}, 016205 (2003). G. S. Agarwal and P. K. Pathak, Phys. Rev. A [**70**]{}, 053813 (2004). P. K. Pathak and G. S. Agarwal, Phys. Rev. A [**71**]{}, 043823 (2005). S. Ghosh, U. Roy, C. Genes and D. Vitali, Phys. Rev. A [**79**]{}, 052104 (2009). F. Toscano, D. A. R. Dalvit, L. Davidovich and W. H. Zurek, Phys. Rev. A [**73**]{}, 023803 (2006). D. A. R. Dalvit, R. L. de M. Filho and F. Toscano, New Journal of Physics [**8**]{}, 276 (2006). L. Praxmeyer, P. Wasylczyk, C. Radzewicz and K. Wódkiewicz, Phys. Rev. Lett. [**98**]{}, 063901 (2007). S. Ghosh, A. Chiruvelli, J. Banerji and P. K. Panigrahi, Phys. Rev. A [**73**]{}, 013411 (2006). J. R. Bhatt, P. K. Panigrahi and M. Vyas, Phys. Rev. A [**78**]{}, 034101 (2008). M. Stobinska, G. J. Milburn and K. Wódkiewicz, Phys. Rev. A [**78**]{}, 013810 (2008). J. Banerji, Contemporary Physics [**48**]{}, 157 (2007). A. J. Scott and S. M. Caves, Annals of Physics [**323**]{}, 2685 (2008). B. Y. Tong, Solid State Commun. [**104**]{}, 679 (1997). B. Y. Tong and N. Kiriushcheva, Phys. Lett. A [**229**]{}, 49 (1997). J. Radovanovic, V. Milanavic, Z. Ikonic and D. Indjin, Phys. Lett. A [**269**]{}, 179 (2008). H. Yildirim and M. Tomak, Phys. Rev. B, [**72**]{}, 115340 (2005). H. Yildirim and M. Tomak, J. App. Phys. [**99**]{}, 093103 (2006). I. Sh Averbukh and N. F. Perelman, Phys. Lett. A [**139**]{}, 449 (1989). R. W. Robinett, Phys. Rep. [**392**]{}, 1 (2004) and references therein. G. Pöschl and E. Teller, Z. Phys. [**83**]{}, 143 (1933). S. Sree Ranjani, A. K. Kapoor, and P. K. Panigrahi, Ann. Phys. [**320**]{}, 164 (2005). J. P. Antoine, J. P. Gazeau, P. Monceau, J. R. Klauder and K. A. Penson, J. of Math. Phys. [**42**]{}, 2349 (2001). A. M. Perelomov, [*Generalized Coherent States and Their Applications*]{} (Springer, Berlin, 1986) N. Gurappa, P. K. Panigrahi and T. Shreecharan, J. Comput. Appl. Math. [**160**]{}, 103 (2003). U. Roy, J. Banerji and P. K. Panigrahi, J. Phys. A [**38**]{}, 9115 (2005). W. Schleich and J. A. Wheeler, Nature [**326**]{}, 574 (1987); W. Schleich and J. A. Wheeler, J. Opt. Soc. Am. B [**4**]{}, 1715 (1987); W. Schleich, D. F. Walls and J. A. Wheeler, Phys. Rev. A [**38**]{}, 1177 (1988); W. P. Schleich, *Quantum Optics in Phase Space* (Wiley-VCH, Berlin, 2001) and references therein. The numerical study of a wider interval of values for ‘$a$’ turns out to be problematic in the case of asymmetric PT potentials because for larger asymmetry the direct measurement of ‘$a$’ from the Wigner plots became more and more unreliable. [^1]: e-mail: [email protected] [^2]: e-mail: [email protected] [^3]: e-mail: [email protected] [^4]: e-mail: [email protected]
{ "pile_set_name": "ArXiv" }
16.5cm 23.5cm -1.5cm ----------- ------------------- [**CHAPTER 5**]{} \[3.5cm\] ----------- ------------------- \ ----------- -------------------------------------------------- [**M. Scheffler and C. Stampfl**]{} \[0.5cm\] Fritz-Haber-Institut der Max-Planck-Gesellschaft Faradayweg 4-6 D-14195 Berlin, Germany \[10cm\] [Handbook of Surface Science]{} [Volume 2, edited by K. Horn and M. Scheffler]{} \[2cm\] ----------- -------------------------------------------------- [**Contents**]{} 1. Introduction 1. The nature of the surface chemical bond: indeterminate concepts,\ yet useful ... 2. What will be discussed and why, and what is missing 2. Concepts and definitions 1. Density of states 2. Energies 3. Binding energy at kink sites 4. The surface energy barrier 3. The tight-binding picture of bonding 1. Adsorbate-substrate interaction 2. Adsorbate band structure 4. Adsorption of isolated adatoms 1. Geometry 2. Density of states $\Delta N(\epsilon)$ 3. Electron density: $n({\bf r})$, $\Delta n({\bf r})$, and $n^\Delta ({\bf r})$ 4. Surface dipole moments 5. Alkali-metal adsorption: the traditional picture of [*on-surface*]{} adsorption 1. The Langmuir-Gurney picture 2. Coverage dependence of the work function 3. Ionization of the adsorbate and screening by the substrate electrons 4. Surface band structure 6. Substitutional adsorption and formation of surface alloys 1. Na on Al(001) 2. Na on Al(111) 3. Co on Cu(001) 7. Adsorption of CO on transition-metal surfaces – a model system for a simple molecular adsorbate 8. Co-adsorption \[the example CO plus O on Ru(0001)\] 9. Chemical reactions at metal surfaces 1. The problems with “the” transition state 2. Dissociative adsorption and associative desorption of H$_2$ at transition metals 1. The potential-energy surface of H$_2$ at transition-metal surfaces 2. The dynamics of H$_2$ dissociation at transition-metal surfaces 10. The catalytic oxidation of CO 11. Summary outline of main points References Introduction {#sec:intro} ------------ The theory of adsorption has reached a level where it is possible to calculate free energies, as well as the electronic and atomic structure, of medium-sized systems with predictive accuracy. Such [*ab initio*]{} calculations (i.e., starting from the electronic structure) typically employ complicated methods and significant computational resources. Clearly, the methodological developments of recent years have been impressive, although further developments, enhancements, and speed-ups of such methods are still necessary. Acknowledging the predictive power of density-functional theory calculations, however, we also note that the need remains for finding explanations and developing simple concepts. Today’s concepts are largely based on experience from gas-phase chemistry, as for example the concepts of electronegativity, HOMOs (highest occupied molecular orbitals), LUMOs (lowest unoccupied molecular orbitals), and the reactivity of open-shell systems. However, it is also clear that some concepts, which are powerful in the gas-phase chemistry of molecules, can be quite misleading when it comes to surfaces. One example is that of “the” transition state of a chemical reaction: For molecular chemistry this concept is typically useful, but for molecules at surfaces it turns out that the dimension of phase space is so high that not just one, but many transition states exist, and all of them may play a role (see Section \[sec:reactions\]). What is needed now, and in the years to come, is to perform more predictive simulations of surface chemical reactions; but also the next step can and should be done, which is the development of [*explanations*]{} and [*understanding*]{}. For example, at this point we are not able to rationalize the factors that determine at which transition-metal surface a rotationally excited molecule will dissociate more easily than a molecule that is not rotationally excited (see Section \[sec:H2\]). Also we are just starting to develop an understanding of why some adsorbates occupy a substitutional site, and not just simply adsorb [*on*]{} the surface, and why this geometry may change with coverage. Note that ten years ago substitutional adsorption of single adatoms was unheard of, but now, it is a well known and common phenomenon (see Section \[sec:substitutional\]). The discussion presented in this chapter is based on results from density-functional theory calculations. Rather than discussing details of these calculations, we simply use the results to show where we are on the way to attaining a rationalization of the factors which actuate the surface chemistry for different systems. This goal does not (and cannot) aim at a quantitative description, and it is clear that concepts, e.g., that of electronegativity, often do not withstand a quantitative analysis. However, describing the results in words and in a chemical language, is the basis on which “understanding” is built. Clearly, the quality (i.e., the nature or character) of an effect is linked to the quantitative numbers, and sometimes a small change in the quantity changes the quality. In this sense our goal is insecure. But we strongly believe that capturing the nature of a situation and understanding trends is the essence of “understanding”. As we said above, theoretical surface science is now able to do this, and some examples along this route already exist, some of which will be discussed in this chapter. ### The nature of the surface chemical bond: indeterminate concepts, yet useful ... {#sec:intro-1} A crude classification of adsorption distinguishes two classes, namely that of a very weak (van der Waals type) interaction between adsorbate and substrate (physisorption), where the adsorption energy is typically less than 0.3 eV per adsorbed particle (6.9 kcal mol$^{-1}$), and that of chemisorption, where the adsorption energy is larger. For chemisorption systems there is a further classification of the nature of bonding which is frequently applied, although it is neither unique nor general. Nevertheless, it has some natural advantages and will also be applied in this chapter. It is based on a survey of electronic, electrical, vibrational, and thermal properties. Thus, altogether we will distinguish four different types of bonding: - van der Waals, - covalent, - metallic, - ionic. As little the nature of bonds of types 2, 3, and 4 is well defined, that of van der Waals bonding is equally so. The concept of the van der Waals interaction is valid for large distances where orbitals do not overlap. It is due to electron-density fluctuations at the different atoms and the polarization that the fluctuations induce at the other atoms. However, at the equilibrium geometry of an adsorbate on a surface, the direct interaction of adsorbate and substrate orbitals is significant. This is indeed plausible as equilibrium geometries are determined by the interplay of attractive interactions and Pauli repulsion. As a consequence, at the bonding geometry of adsorbed noble-gas atoms which are regarded as exhibiting a van der Waals like bonding, and even more so at the turning point of noble-gas atom scattering at surfaces, the physics is largely determined by the interaction of orbitals and [*static*]{} polarization. Thus, [*physisorption*]{}, although implying a weak interaction strength, can induce a static dipole moment at the adsorbate (in particular for larger atoms, e.g., Xe), and the electrostatic interaction of this adsorbate dipole with the substrate contributes noticeably to the bond strength. Indeed, density-functional theory (DFT) calculations performed using either the local-density approximation (LDA) or the generalized gradient approximation (GGA) for the exchange-correlation interaction, which both lack a description of the nature of the van der Waals interaction, as this is intrinsically non-local, seem to give a reasonable description of adsorption of noble-gas atoms at surfaces (see, e.g., Brivio and Trioni, 1999). Nevertheless, it is not clear if the acting “overlap-modified dispersion forces” are described by the present exchange-correlation functionals with sufficient accuracy. In any case, what we like to emphasize is that even weak bonding can give rise to noticeable changes in the electrostatic field at the surface and therefore to changes in surface properties. We elaborate here on this discussion more than usual because the use of the above noted four bonding types has often caused confusion when they were interpreted literally. Knowing about the danger of over-interpreting the relevance of these concepts we still feel that the advantage of categorizing systems is important for identifying trends, noting unusual effects, and for [*scientific understanding*]{} in general. In Sections \[sec:tight\], \[sec:nature\], and \[sec:LT-alkali\] we will continue this discussion with respect to ionic and covalent bonding. Metallic bonding, also noted in the above list, is a special case of covalent interaction where the electron density is more delocalized and not (strongly) peaked between the atoms. In principle, the attractive interaction is due to the fact that the electrons act like a structure-less “glue” between the nuclei. This attraction is stabilized by a repulsive term due to the kinetic energy of the electrons, because electrons do not like to become localized, but instead, as a consequence of the Heisenberg uncertainty principle, they like to spread out. For [*isolated*]{} adatoms, the concept of metallic bonding is not very useful because metallic bonding, and the concept of delocalized electrons, requires a high coordination. Therefore we will not allude to this concept in Section \[sec:nature\] where isolated atoms are discussed, but we will in Sections \[sec:LT-bands\] and \[sec:substitutional\] when we compare adlayers of different density and discuss the formation of a surface electronic band structure. ### What will be discussed and why, and what is missing {#sec:intro-2} This chapter deals with various adsorbates on metal surfaces, where we focus on the electronic properties and the nature of the chemical bond. To some extent this also necessitates a discussion of the atomic geometries because the distances and directions between the atoms and their neighbors, determine which, and how, orbitals will hybridize. Just as one example we mention aluminum which is a nearly-free-electron metal. However, when an Al atom is placed in a certain local coordination, its $s$- and $p$-electrons can hybridize and form directional bonds. This is in fact well known even for bulk systems, as for example AlAs. This property of Al, namely being close to a covalent material, contributes to the low formation energy of surface vacancies at Al(111) (see Neugebauer and Scheffler, 1992) and is instrumental in the theoretical result that self diffusion at Al(001) proceeds via the exchange mechanism (see Feibelman, 1990; Yu and Scheffler, 1997). We will keep the discussion in this chapter simple and accentuate the [*qualitative*]{} nature of the various mechanisms. Nevertheless, we emphasize that all results discussed below are based on quantitative calculations performed using density-functional theory (DFT) (Dreizler and Gross, 1990). Indeed, the dialectic relation of quantity and quality will become obvious through several examples discussed below: Often a small quantitative difference in certain values will cause a significantly changed electronic and/or geometric structure and as a consequence a different bonding quality or nature. The exchange-correlation functional employed in many of the studies discussed below is the local-density approximation (LDA), which gives a reliable description of geometries and the nature of the bond. In the more recent studies, which are presented in Sections \[sec:CO\] – \[sec:catalytic\], the generalized gradient approximation (GGA) is employed. For equilibrium geometries the GGA gives results similar to the LDA. For the description of chemical reactions, in particular of transition states, where the breaking of old bonds and making of new bonds occurs, the GGA is in fact mandatory, i.e., the LDA often gives even qualitatively incorrect results (see for example Hammer et al., 1994, and references therein). For simplicity’s sake, we restrict ourselves to close-packed substrate surafces \[mainly fcc (111) and fcc (001)\]. The more open surfaces are typically close to a structural instability and therefore sometimes already reconstruct when clean, or when adsorbates are added. We also limit the number of substrate materials with the view that this will ease the readability of the chapter. However, essentially all important mechanisms that we would like to discuss are represented by the systems presented. Because of their large dipole moments and special role in various industrial applications, the discussion of alkali-metal adsorbates is addressed in particular detail. This presentation also focuses on the general properties; for more complete discussions we refer to some recent review papers, e.g., Stampfl and Scheffler (1995) and Adams (1996). Noble-gas adatoms[^1] and $f$-electron systems are not discussed, and also for metallic substrates not all “classes” are covered: We concentrate on elemental substrates (only a few words are said about alloys) and consider low coverages, i.e., from a single adatom up to a full monolayer.[^2] As the properties of adsorbates on metals are manifold, we are unable to address all features which may be relevant in one or another situation. For example, core-levels and surface core-level shifts (with their interesting initial state and final state effects) are not mentioned (see for example Andersen et al., 1994; Methfessel et al., 1995; Hennig et al., 1996; Ganduglia-Pirovano et al., 1997, and references therein). The same is true for surface electric resistance, magnetism, and more. Otherwise, we trust that the selection of systems is representative. It covers a small fraction of work done by us over the last several years on adsorption, co-adsorption, and chemical reactions. We note that similar work has also been published by other groups (see, e.g., Hammer and N[ø]{}rskov, 1997, and references therein). Concepts and definitions {#sec:definitions} ------------------------ This section starts with some general remarks and then collects definitions of important quantities which are typically calculated in theoretical work on adsorption and will be used later in this chapter. Accurate knowledge of the geometry of the adsorbate and substrate atoms is a prerequisite for any additional analysis of the adsorbate properties, as for example the surface electronic structure, adsorbate-induced work function changes, and the chemical reactivity. The apparent hierarchy expressed in this sentence reflects the fact that the atomic structure is somewhat more directly accessible for experimental studies. Nevertheless, we emphasize that electronic and atomic structure are closely interconnected and do not consider it very useful to analyze “the chicken and the egg problem”, i.e., to discuss whether the geometry creates the electronic structure or vice versa. In fact, we will stress in this chapter (particularly in Sections \[sec:LT-alkali\] and \[sec:substitutional\]) that the same adsorbate can exhibit a different bonding character, depending on the adsorbate coverage and the local adsorbate geometry. We will discuss these aspects for special examples but note that these are [*examples*]{} which represent many systems; they are selected because they demonstrate the effects most clearly. ### Density of states {#sec:DOS} An important quantity accessible in calculations, though not in experiments, is the local density of states $$n({\bf r}, \epsilon) = \sum_{i=1}^\infty |\varphi_i({\bf r})|^2 \delta(\epsilon - \epsilon_i), \label{eq:l-dos}$$ where $\varphi_i({\bf r})$ are the single-particle eigenfunctions of the Kohn-Sham Hamiltonian. For those who like Green functions this is written as $$n({\bf r}, \epsilon) = -\frac{2}{\pi} {\rm Im} \cal{G}({\bf r},{\bf r}, \epsilon),$$ where $\cal{G}(\epsilon)$ is the retarded Green-function operator of the adsorbate system. The electron density is $$n({\bf r}) = \int \limits_{-\infty}\limits^\infty f(\epsilon, T) n({\bf r}, \epsilon) d \epsilon = \sum_{i= 1}^{ \infty } f(\epsilon_i, T) |\varphi_i({\bf r })|^2, \label{eq:n(r)}$$ with the Fermi distribution $f(\epsilon, T)$ at temperature $T$. Often, when the nature of a chemical bond is analyzed, electron density plots are shown, sometimes complemented with a discussion of charge transfer from atom A to atom B. We do not consider this approach so useful and emphasize that a small difference in the electron density sometimes implies a significantly different physical-chemical interaction. This is demonstrated in Fig. \[nacl\] (0,5.0) which shows the electron density of bulk NaCl, that is a system, which is well accepted to have ionic bond character. However, even in this example the self-consistently calculated electron density (right) and that constructed by a superposition of electron densities of the [*neutral*]{} Na and [*neutral*]{} Cl atoms (left) differ only slightly, and just on the basis of inspecting $n({\bf r})$ charge transfer can hardly be identified. The small changes brought about by the self-consistent rearrangement of electron density are that around the Na nucleus the charge density is very slightly decreased; but also at the Cl atom we see that the maximum of the electron density, which is rather close to the nucleus, becomes slightly lower. Thus, when one only inspects the electron density, even NaCl does not present a clear case for a system with ionic bonding. We emphasize that for low-symmetry situations, as for example a surface, this problem is even more pronounced. However, an analysis of the density of states (see, e.g., also Section \[sec:Delta-N\]) shows that it is indeed appropriate to describe the upper valence band in terms of Cl $3p$ and the lower conduction band in terms of Na $3s$ orbitals. The example of Fig. \[nacl\] shows that despite the fact that the electron density can be considered the ruling quantity in density-functional theory, its inspection can be misleading. Sometimes, though not always, plots of density [*differences*]{} $$\Delta n({\bf r}) = n({\bf r}) - n^0({\bf r}) \label{eq:Deltan}$$ give a better impression (see Sections \[sec:nature\] and \[sec:LT-screening\]). Here $n^0({\bf r})$ is the electron density of the clean substrate where the geometry is (typically) chosen to be that of the adsorbate system. Also the [*difference density*]{} $$\label{n-Delta} n^\Delta({\bf r}) = n({\bf r}) - n^{\rm 0}({\bf r}) -n^{{\rm Na},f_{3 s}}({\bf r}) \quad$$ is often a helpful quantity. Here $n^{{\rm Na},f_{3 s}}({\bf r})$ is the (spherical) electron density of the partially ionized atom to be adsorbed, where the occupation of the valence level is given by the parameter $f_{3s}$. Just as an example, our indices here refer to a Na atom, where the valence level is $3s$. A typically more sensitive quantity is the density of states (DOS), $$N(\epsilon) = \int n({\bf r}, \epsilon) d {\bf r} = \sum_{i= 1}^{ \infty } \delta(\epsilon -\epsilon_i), \label{eq:DOS}$$ where the sum goes over all eigenstates of the Kohn-Sham Hamiltonian. The DOS gives a noticeable contribution to the electrons’ kinetic energy $$T_s[n({\bf r})] = \int \limits_{-\infty}\limits^\infty f(T, \epsilon) N(\epsilon)\, \epsilon\, d \epsilon - \int V^{\rm eff}({\bf r}) n({\bf r}) d{\bf r}, \label{eq:T_s}$$ where $V^{\rm eff}$ is the effective potential of the Kohn-Sham Hamiltonian (see Chapter 1 of this book). We also note that a one-to-one correspondence exists between $N(\epsilon) $ and the electron density,[^3] which (in addition to the argument of practicability and clarity) gives a premise to its use. Throughout this chapter we will follow the thinking that the nature of bonding is determined by the interaction of the orbitals at the different atoms involved, and thus, it is more clearly identified by an inspection of the DOS and/or by the adsorbate-induced change ($\Delta$DOS) $$\Delta N(\epsilon) = N(\epsilon) - N^0(\epsilon), \label{eq:Delta-DOS}$$ where $N^0(\epsilon)$ is the DOS of the clean substrate where the geometry is typically chosen to be that of the adsorbate system. Furthermore, we note the state-resolved DOS, also called the projected DOS, as a very useful quantity: $$N_{\alpha}(\epsilon) = \sum_{i= 1}^{\infty} |\langle \phi_\alpha | \varphi_i \rangle|^2 \delta(\epsilon -\epsilon_i), \label{eq:DOS_a}$$ where $\phi_\alpha$ is a properly chosen localized function. The spatial distribution of the electron density is viewed as a consequence (not the origin) of the hybridization of different orbitals. The critic may argue that this is just (another) chicken-egg problem, and we agree with this assessment, i.e., our approach is simply taken because it is useful. We also note that information about $N_{\alpha}(\epsilon)$ can be obtained (though in a somewhat distorted way) from angle-resolved photoemission, inverse photoemission, and scanning tunneling spectroscopy. Unlike the bond formation in molecular chemistry, the case of adsorption is one between very different partners: The adatom comes with discrete energy levels and few occupied states, but the substrate has a near-infinite number ($\sim$$10^{23}$) of electrons. Thus, the valence level of the adparticle will interact with a semi-infinite continuum of levels $\epsilon \geq \epsilon_{0}$, where $\epsilon_{0}$ is the bottom of the substrate valence band. The location of the highest occupied level, the Fermi energy, is determined by the substrate. On a surface, a description of the chemical bonding in terms of an elementary discrete level scheme has therefore to be extended: Levels lying in the band region of the substrate receive a finite width. These broad levels are called adsorbate-induced resonances and they get filled up to the substrate Fermi level. The wave functions of these resonances can be understood to arise from the adsorbate orbitals. In an alternative view, the states of the clean surface are described as standing waves, incident from the bulk and reflected at the potential-energy barrier of the clean surface, with a node at the surface. An adsorbed particle modifies the reflection properties, which can be described by introducing an energy-dependent phase shift, $\delta_{\alpha}(\epsilon)$, describing the scattering of bulk states of a certain representation $\alpha$ by the adparticle. This phase shift is defined by (Callaway, 1964, 1967) $$\begin{aligned} {\rm tan} \delta_{\alpha}(\epsilon) = \frac{-{\rm Im} D_{\alpha} (\epsilon)}{{\rm Re} D_{\alpha}(\epsilon)},\end{aligned}$$ where $D_{\alpha}$ is the determinant $$\begin{aligned} D_{\alpha}(\epsilon) = {\rm det}\{1 - {\cal G}^0 (\epsilon) \Delta V \}_{\alpha}.\end{aligned}$$ ${\cal G}^0$ is the Green’s function of the bare substrate, and $\Delta V$ is the change in the potential due to the adsorbate $$\begin{aligned} \Delta V = V^{\rm eff}[n] - V^{\rm eff}[n^0].\end{aligned}$$ The density of states, induced by the adparticle, is given by the derivative of the phase shift $$\begin{aligned} \Delta N(\epsilon) = \frac{2d_\alpha}{\pi} \frac{d \delta_\alpha(\epsilon)}{d \epsilon},\end{aligned}$$ where $d_\alpha$ is the dimension of the representation $\alpha$. A resonance in $\Delta N(\epsilon)$ will occur at an energy close to that at which the phase shift $\delta_\alpha(\epsilon)$ increases through $\pi/2$ with increasing energy, i.e., where the real part of the determinant $D_\alpha$ vanishes. Close to this energy, $\epsilon_\alpha$, the induced density of states takes a Lorentzian line shape $$\begin{aligned} \Delta N(\epsilon) = \frac{d_\alpha\Delta_\alpha}{\pi} \frac{1}{ (\epsilon - \epsilon_\alpha)^{2} - \left( {\Delta_\alpha}/2 \right)^{2}}. \label{eq:AGN}\end{aligned}$$ The width of the resonance, $\Delta_\alpha$, is $$\begin{aligned} \Delta_\alpha = \left[ \frac{2{\rm Im} D_\alpha(\epsilon)}{({d {\rm Re} D}_\alpha (\epsilon)/d \epsilon)} \right]_{\epsilon = \epsilon_\alpha}.\end{aligned}$$ We note the close similarity between Eq. (\[eq:AGN\]) and the Anderson-Grimley-Newns model of chemisorption (Grimley, 1975; Muscat and Newns, 1978, 1979). On the lower energy side of the resonance, the phase of the reflected wave is shifted such that electron density is accumulated in the region of the adparticle-substrate bond, indicating that these states are bonding in character. On the higher energy tail of the resonance, the electron density in the bond region is reduced, indicating that these states are antibonding in character (Lang and Williams, 1978; Liebsch, 1978). However, if the interaction with the substrate is very strong, bonding and antibonding states will split apart: A bound state (or resonance) is then formed below the substrate band, which is bonding in character, and a broad (antibonding) resonance will appear in the valence band (see also the discussion of Fig. \[tb-DOS\] in Section \[adsorbate-substrate-interaction\]). ### Energies {#sec:energies} We will assume that the dynamics of the electrons and the nuclei can be decoupled and that whatever the dynamics of the nuclei are, the electrons are in the electronic ground state of the instantaneous geometry. This is the Born-Oppenheimer approximation (Born and Oppenheimer, 1927; Born and Huang, 1954), which for adsorbates, and often also for chemical reactions, is well justified; for some reactions, and in particular for photo-chemistry, important violations of the Born-Oppenheimer approximation occur, but this is not the subject of this chapter. The [*DFT total energy*]{}, $E^{\rm total}(T, V, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots,\{{\bf R}_I\})$, at temperature $T$, volume $V$, and composition $N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots$, when studied as a function of the atomic coordinates, is often called the potential-energy surface (PES) because it defines the potential-energy landscape on which the nuclei $A, B, \cdots$ travel. It is related to an experimentally accessible quantity only in a restricted way: If, as is typically done, the self-consistent calculations are performed at constant volume, the [*DFT total energy*]{} corresponds to the Helmholtz free energy at zero temperature and neglecting zero-point vibrations. In general, the Helmholtz free energy is $$\begin{aligned} F(T, V, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots, \{ {\bf R}_I \} ) & = & E^{\rm total}(T, V, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots, \{{\bf R}_I \}) \nonumber \\ & & + E^{\rm vib.}(T, V, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots, \{ {\bf R}_I \}) \nonumber \\ & & - T\, S(T, V, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots, \{ {\bf R}_I \} ) \label{helmholtz}\end{aligned}$$ with the vibrational contribution noted as $E^{\rm vib.}$ and $S$ is the entropy. At a given volume $V$, the atomic geometry of stable or metastable configurations is determined by $$\left( \frac{\partial F(T, V, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots, \{ {\bf R}_I \} ) } {\partial {\bf R}_I } \right)_{T,\, V, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots} = 0,$$ and for a given pressure $p$ it is determined by $$\left( \frac{\partial G(T, p, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots, \{ {\bf R}_I \} ) } { \partial {\bf R}_I } \right)_{T,\, p, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots} = 0,$$ where $$\begin{aligned} G(T, p, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots, \{ {\bf R}_I\}) & = & F(T, V, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots, \{ {\bf R}_I\}) \nonumber \\ & & + p\, V(T, p, N^{\rm nuc.}_A, N^{\rm nuc.}_B, \cdots, \{ {\bf R}_I\}) \label{gibbs}\end{aligned}$$ is the Gibbs free energy. If the system is in contact with a particle reservoir, for example, the sample is held in some gas phase, particles can be exchanged between the system and the reservoir. Then we have to add to Eq. (\[helmholtz\]) a term $- \sum_X \mu_X N^{\rm nuc.}_X$, where $\mu_X$ is the atom chemical potential of atom-type $X$, which can be controlled by external reservoirs, i.e., by the evironmental conditions (partial pressure and temperatur). Later in this chapter we will study the [*adsorption energy per adatom*]{}. This is the difference of the total energy of the adsorbate system and the total energy of the clean substrate together with a corresponding number of free, neutral atoms. For [*on-surface*]{} adsorption this reads $$E_{\rm ad}^{\rm Na/Al(001)} = - \left( E^{\rm Na/Al(001)} - E^{\rm Al(001)} - N^{\rm nuc.}_{\rm Na} E^{\rm Na-atom} \right)/ N^{\rm nuc.}_{\rm Na}, \label{eq3_1}$$ where $E^{\rm Na/Al(001)}$ is the total energy per adatom of the adsorbate system, $E^{\rm Al(001)}$ is the total energy of the clean Al(001) substrate, and $N^{\rm nuc.}_{\rm Na} E^{\rm Na-atom}$ is the total energy of $N^{\rm nuc.}_{\rm Na}$ free Na atoms that take part in the adsorption. We have used here as an example indices which refer to the adsorption of Na on Al(001), but translation to other systems is obvious. Often adatoms are [*not*]{} adsorbed on the surface with only slight modification of the original surface structure, but instead adsorption may occur substitutionally. In this case the adatom kicks out an atom from the surface and takes its site. In thermal equilibrium the kicked out atom is then re-bound at a kink site at a step (the binding energy at a kink site equals the bulk cohesive energy, which is shown in the next section). For substitutional adsorption, the adsorption energy is defined essentially the same way as above, only that the kicked off surface atoms that are re-bound at kink sites, have to be accounted for as well. Each rebound kicked out atom contributes an energy equal to that of a bulk atom (see Section \[sec:kink\]). Thus, for the substitutional adsorption the adsorption energy is $$\begin{aligned} E_{\rm ad}^{\rm Na/Al(001)-sub} & = & - ( E^{\rm Na/Al(001)-sub} + N^{\rm nuc.}_{\rm Na} E^{\rm Al-bulk} \nonumber \\ & &- E^{\rm Al(001)} - N^{\rm nuc.}_{\rm Na} E^{\rm Na-atom} )/N^{\rm nuc.}_{\rm Na}. \label{eq3_2}\end{aligned}$$ $E^{\rm Al-bulk}$ is the energy of an atom in a bulk crystal of aluminum and the quantity $E^{\rm Na/Al(001)-sub}$ in Eq. (\[eq3\_2\]) is the total energy of the slab with the adatoms adsorbed in substitutional sites. We end this section with a warning that may be relevant for [*Surface Science*]{} studies more often than typically appreciated. The energy quantities defined in this section are relevant if thermal equilibrium conditions are attained, but often this is not the case. Instead, a surface is studied which may be in a metastable state and this state hopefully corresponds to an equilibrium situation of the sample’s history (at best). This somewhat unclear situation will hopefully improve in the future, when not just UHV studies are performed but also experiments under well controlled atmospheres. We also note that structures in [*local*]{} thermodynamic equilibrium, or metastable geometries can have a very long life time – just consider diamond (a metastable structure of carbon) for which at room temperature the phase transition to its ground-state crystal structure (graphite) is known to be rather slow. In particular, for multi-component materials, the stoichiometry at the surface requires exchange of atoms or molecules with some reservoirs and this can be hampered by significant energy barriers. Metal oxides represent an example which comes to mind, as in UHV, oxygen can desorb into the chamber, but with respect to the metal content at the surface, attaining the thermal-equilibrium stoichiometry is hampered (see, e.g., Wang et al., 1998). ### Binding energy at kink sites {#sec:kink} As mentioned above, kicked out surface atoms may be re-bound at kink sites at steps. We will see that this is in particular relevant when we discuss below in this chapter the substitutional adsorption of alkali metals atoms and of cobalt. Therefore we now present a simple discussion of the energetics of substrate atoms at kink sites. Although we use a simplified model, we note that the conclusions are in fact valid in general. The total energy of a many-atom system can be written as the sum over contributions assigned to the individual atoms, $$E = \sum_{I=1}^{N^{\rm nuc.}} E_I. \label{bond-cut}$$ $E_I$ is the energy contribution due to atom $I$. In an approximate way we can write $E_I = E_I(C_I)$, where $C_I$ is the number of nearest-neighbor atoms of atom $I$. This is a simplified presentation of an approach which has many names (e.g., effective-medium theory, embedded-atom method, Finnis-Sinclair potentials, glue model) and which we all call [*bond-cutting models*]{}. The various implementations are similar but differ in the way the function $E(C)$ is represented. In the simplest, yet physically meaningful approach, the function $E(C)$ is roughly proportional to $\sqrt{C}$ (see Spanjaard and Desjonquères, 1990; Robertson et al., 1994; Payne et al., 1996; Methfessel et al., 1992b; Christensen and Jacobsen, 1992). If we consider a single-element system and place an atom at a kink site at a step of an fcc (001) surface, the changes of the local coordination are as noted in Fig. \[fig:kink\]. (0,3.5) Putting this information into Eq. (\[bond-cut\]) and calculating the [*change*]{} in the total energy by placing an atom ($I=1$) at a kink site (i.e., $E^{\rm kink}_{\rm ad} = -E^{\rm after} + E^{\rm before}$) gives $$\begin{aligned} E_{\rm ad}^{\rm kink} & = & - \left( E_{1}(6) +E_{2}(7) +E_{3}(8) +E_{4}(10) +E_{5}(12) +E_{6}(11) +E_{7}(9) \right) \nonumber\\ & & + E_{1}(0)+E_{2}(6)+E_{3}(7)+E_{4}(9) +E_{5}(11) +E_{6}(10) +E_{7}(8) \nonumber\\ & = & -E_{5}(12)+E_{1}(0). \label{kink}\end{aligned}$$ Thus, the adsorption energy at a kink site equals the cohesive energy. This result is in fact plausible if one considers that the adsorption at a kink site leaves the system essentially unchanged, because the kink site is simply moved by one atom and thus, the situation at the surface is physically not altered. Therefore the difference between the original and final situations is simply (and rigorously) the addition of one bulk atom, which has the energy $E(12)$. ### The surface energy barrier {#sec:barrier} We add some remarks on the behavior of the effective Kohn-Sham potential, $V^{\rm eff}$ (see Eq. (\[eq:T\_s\]) and Fig. \[effective-potential\]). The increase in $V^{\rm eff}$ (0,8.0) at the surface from its average bulk value up to the vacuum level is termed the surface barrier. Inside the solid, the potential $V^{\rm eff}$ seen by an electron becomes attractive, due to the electrostatic potential of the ion cores, due to the electrostatic field of the surface-dipole layer, and due to the lowering of the electron energy by the formation of an exchange-correlation hole. For an electron in the vacuum region, the potential is described by the classical image effect $$\begin{aligned} V^{\rm eff}(z) = - \frac{1}{4 \pi \varepsilon_0} \frac{(e^{-})^{2}}{4(z - z_{0})} \mbox{\quad \quad for $(z - z_0) \, \raisebox{.6ex}[-.6ex]{$>\!\!\!\!$} \raisebox{-.6ex}[.6ex]{$\sim$}\, 2$ \AA}, \label{eq:image-potential}\end{aligned}$$ where $z$ is the position of the electron, $z_{0}$ is the position of the reference plane of the image effect, and $\varepsilon_0$ is the vacuum dielectric function. Figure \[effective-potential\] shows the electrostatic potential and the effective one-particle potential at an adsorbate-covered surface. Here, $\Phi$ is the work function, $V^{\rm es}$ is the electrostatic potential due to the electron density $n({\bf r})$ and the nuclei, and $V^{\rm xc}$ is the exchange-correlation potential. The latter is attractive and varies roughly as $n({\bf r})^{1/3}$. The only surface contribution to the total height of the surface barrier is due to the electrostatic potential; the contribution of the exchange-correlation term to the height is a property of the bulk and not of the surface. Nevertheless, the behavior of the potential at the surface is largely determined by the distortion of the exchange-correlation hole which stays behind as an electron passes through the surface region into the vacuum. In the local-density approximation, the exchange-correlation hole is assumed to be spherically symmetric and centered on the electron. Although both assumptions are, in general, incorrect, this approximation affects the potential $V^{\rm eff}$ significantly only in the surface region. The local-density approximation therefore yields an exponential decrease of the effective potential near the surface rather than the l/$z$ behavior of the image effect. This inaccuracy of the local-density approximation appears to become noticeable well outside the surface region in the vacuum. Properties such as the ground-state electron density, the work function, or the surface energy seem to be relatively little affected (see, e.g., Methfessel et al., 1992a, b). The detailed shape of the barrier thus appears to be of less importance than its position and height. The barrier affects individual electron wave functions. For example, reflection of electrons from the inner side of the barrier can give rise to surface resonances when the electron is trapped between the barrier and the rest of the crystal. In particular, the barrier position influences the energies of adsorbate states. For excited states, there are additional effects. Virtual surface resonances will appear, which in LEED, for example, show up as narrow peaks in the intensity versus voltage curves (McRae, 1971; Jennings, 1979). Furthermore, the wavelength of an electron is longer outside the crystal than inside. This yields the well-known refraction effect which broadens the angular range of emitted electrons in the vacuum region such that, at high polar angles, the electron current vanishes (see, for example, Scheffler et al., 1978). Another effect is found in the interaction of light with the surface: Photoabsorption, and hence photoemission, require a gradient in the potential. The surface barrier thus yields a special contribution to this excitation, called the surface photoeffect. For photoemission from adsorbates on transition metals, this contribution appears to be very small compared to that due to the ion core potentials. However, it can be important in systems where the valence electrons are nearly free electron-like, i.e., where their interaction with the ion cores is small. For excited states, the inner potential will be modified (also becoming complex) due to inelastic electron-electron interactions such as the excitation of electron-hole pairs, plasmons or surface plasmons, and dynamical corrections (e.g., a delay in the response of the electrons). The tight-binding picture of bonding {#sec:tight} ------------------------------------ ### Adsorbate-substrate interaction When an atom and a surface start to interact, the respective states mix and new states are created which have energy levels usually broadened and shifted with respect to the energy levels found in the uncoupled systems. Typically, the new states can still be related to the original ones, and Fig. \[tb-DOS\] shows a schematic tight-binding description of the interaction of an atom with a transition metal surface. The free atom’s electronic structure is noted in panel (b) of the figure, and the electronic structure of the substrate is sketched in panel (a). (0,9.0) In principle we could choose any adatom we like for this discussion, but for ease we take hydrogen. Therefore the two levels of relevance, which result from solving the Kohn-Sham equation are the hydrogen $1s$ and hydrogen $2s$ levels; for simplicity we neglect contributions from the H $2p$ and other higher-lying states. We note that the highest occupied DFT-LDA (and DFT-GGA) Kohn-Sham eigenvalue should not be confused with ionization energy (see the discussion of Fig. \[Na3s\] and Eq. (\[eq-janak\]) below). Instead, for partially occupied valence states it is roughly at the mid value of the ionization and affinity energies. Different transition-metal substrates mainly differ in the width of the $d$-band, which increases from $3d$ to $4d$ to $5d$; and they differ in the position of the Fermi level, which varies from the left to the right of the periodic table as follows: For the $4d$ series it is at the lower edge of the $d$-band for strontium, just below the top of the $d$-band for palladium, and about 3 eV above the upper edge of the $d$-band for silver. Thus, in the example of Fig. \[tb-DOS\] we use the Fermi level $\epsilon_{\rm F}$ corresponding to palladium. At first we consider the role of the substrate $s$-band. When the adsorbate and the substrate interact, the hybridization of the adsorbate wave functions and the states of the substrate $s$-band gives rise to a broadening of the adsorbate levels, and the atomic levels will shift because the substrate Fermi level and the electron chemical potential of the atom become aligned. The latter will result in a fractional electron transfer (see the discussion of Fig. \[Na3s\] below). An analysis of the wave-function character in such a broadened peak shows that the low-energy part of the peak belongs to states which have an increased electron density between the adsorbate and the substrate (such states are called “bonding”), and the high-energy part of the peak belongs to states which have a node between the adsorbate and the substrate (such states are called “antibonding”). The substrate $s$-electrons spill out most into the vacuum, and that is why the broadening (and shifting) of electronic levels is the first change that happens when an atom is brought toward a surface. Broadening implies a coupling of the formerly localized electrons of the adatom to the substrate, thus a delocalization. In fact, still neglecting (for the moment) the interaction of the adatom with the $d$-band, there are three contributions which affect the adatom levels: - A shift which can go in either direction and which is caused by charge transfer or charge redistribution (or polarization) at the adatom. For partially occupied valence states it largely reflects an alignment of the adsorbate DOS with respect to the substrate Fermi level. - A shift toward lower[^4] energies because the potential at the surface is lower than that in vacuum (cf. Fig. \[effective-potential\] and/or Chapter 1 of this book). - A contribution which implies a shift to lower$^{\ref{higher-lower}}$ energies because the self-interaction (an artifact in DFT-LDA and DFT-GGA calculations) is smaller for the more extended states of the adsorbate than for the states in the free atom. These broadened and shifted energies are called “renormalized atomic levels”. The three contributions are indeed significant. As a consequence, a self-consistent treatment is crucial for calculating the adsorbate-substrate interaction. We note that the self-interaction effect is indeed strong when DFT-LDA energy levels are studied. But in total-energy [*differences*]{} it nearly cancels out. This is demonstrated in Fig. \[Na3s\], which shows the DFT-LDA (0,6.5) eigenvalue for the atomic Na $3s$-state as function of its occupation. For the neutral Na atom the occupation is 1 and the energy level is at $-$2.8 eV. The extent to which a Kohn-Sham energy eigenvalue $\epsilon_k$ reflects the ionization energy, which is the minimum energy to remove an electron from the $k$-th level, depends on how strongly the eigenvalue depends on the occupation number $f(\epsilon_k, T)$. If this dependence is negligible, then the negative of the energy eigenvalue equals the excitation energy. This is the density-functional theory analogue of Koopmans’ theorem in Hartree-Fock theory. In general, the ionization energy ($I_{k}$) is defined by the total-energy difference of the neutral atom and the positively charged ion, and this can be read off from Fig. \[Na3s\] as the energy at occupation 0.5. This approach only employs the mean-value theorem of integration $$I_k = E^{N-1} - E^N = \int_{N}^{N-1} \frac{ d E^{N'}}{d N'} dN' = - \int_{0}^{1} \epsilon_k(f_k) d f_k \approx -\epsilon_k(f_k=0.5). \label{eq-janak}$$ It is called the Slater-Janak transition-state approach of evaluating total-energy differences (Janak, 1978), which works very well in LDA and GGA calculations (for a discussion of the general proof see Perdew and Levy, 1997; Kleinman, 1997; and references therein). According to Fig. \[Na3s\] the transition-state gives $I_k \approx 5.2$ eV, which agrees well with the experimental result for the ionization energy of 5.14 eV. Analogously we note that the highest occupied Kohn-Sham eigenvalue at occupation 1.0 agrees well with the mean value of the ionization and affinity energies. The result shown in Fig. \[Na3s\] is typical for all atoms. It demonstrates the importance of electron-electron correlation, though it is also largely due to self-interaction, an LDA artifact. Clearly, the measurable ionization energies (level at occupation 0.5) and the theoretical Kohn-Sham eigenvalue (level at occupation 1.0) are very different and should not be confused. Figure \[Na3s\] also shows that typically only partial electron transfer is to be expected upon adsorption. If the substrate mainly plays the role of providing an electron reservoir, thus, the electron chemical potential is fixed by the substrate Fermi level, which for Al(001) is at 4.4 eV below the vacuum level, Fig. \[Na3s\] then shows that the occupancy of the Na $3s$-level will be adjusted to 0.65. Thus we obtain a partial ionization and a shift of the Na $3s$-level, already from a study of the free-atom eigenvalue. This is fully in accord with what is suggested by the electronegativities of the atoms: $\kappa_{\rm Na}=0.93$ and $\kappa_{\rm Al}=1.61$. We continue the discussion of Fig \[tb-DOS\]. At close distances of the adsorbate to the surface the [*renormalized atomic levels*]{} \[i.e., the levels which result after the effects $(i), (ii)$, and $(iii)$\] will interact with the more localized $d$-states. Because the $d$-band is rather narrow and its width comparable to the interaction strength, the interaction will result in a splitting into the bonding states, $ \psi_{\rm b} \approx (\varphi_{{\rm H}\,1s} + \varphi_{{\rm Pd}\,4d} )$ and the antibonding states, $ \psi_{\rm a} \approx (\varphi_{{\rm H}\,1s} - \varphi_{{\rm Pd}\,4d} )$. The resulting adsorbate-induced density of states (DOS) for the H $1s$-state is shown in panel (c) of Fig. \[tb-DOS\]. Panel (d) shows the corresponding result induced by the H $2s$ state. We note that the resulting peaks are close to the lower and upper edge of the $d$-band and inside the $d$-band the density is reduced, which simply reflects the fact that these states are shifted from inside the $d$-band to higher and lower energy upon hybridization with the adsorbate states. (0,7.0) Obviously, the bonding is strongest when bonding states are occupied and antibonding states remain empty. According to Fig. \[tb-DOS\] this will happen when the Fermi level is in the middle of the $d$-band, as for example for Mo and Ru substrates. Figure \[tb-levels\] describes the same physics as Fig. \[tb-DOS\], but with an additional simplification: Now the broadening due to the substrate $s$-band is ignored and the $d$-band is replaced by just a single level, the $d$-band center. The energy levels of the “free” adatom are in fact the “renormalized levels”, and therefore we tagged them in Fig. \[tb-levels\] with a tilde: $1\tilde{s}, 2\tilde{s}$. Deepening this discussion we show in Fig. \[gurney\] results of actual calculations where the distance dependence of the broadening and shift of the energy level is displayed for two qualitatively different systems: Na on Al(001) and O on Ru(0001). (0,5.0) For both adatoms, Na and O, the ionization energy of the free atoms is below the Fermi level. In fact, the relevant “energy level” to be considered is not the ionization energy and also not the electron affinity, but the mean value of those two (or alternatively, the Kohn-Sham eigenvalue of the highest occupied state; cf. the discussion above). This mean value is plotted in Fig. \[gurney\]: For the Na atom it is at (5.1+0.5)/2 eV =2.8 eV and for the O atom it is at (13.6+1.4)/2 eV = 7.5 eV below vacuum. When the adatoms approach the surface, the substrate Fermi level acts as an electron reservoir. For Na this implies a partial charge transfer from the adsorbate to the substrate (see the discussion of Fig. \[Na3s\] above). But for oxygen, with its initially rather deep level, this implies a partial charge transfer from the substrate to the adatom. Upon close approach these partially ionized atoms interact with the substrate orbitals. For Na/Al(001) the DFT-LDA calculations predict a double peak structure, which has its main weight above the Fermi level, but with a tail reaching into the occupied states regime. For O/Ru(0001) the interaction is stronger because the adatom’s equilibrium position is very close to the surface and the substrate $d$-states impart a strong covalent bond and a splitting into bonding and antibonding states. The left panel of Fig. \[gurney\] nicely reflects the Gurney description of alkali-metal adsorption, and the right panel of Fig. \[gurney\] displays the corresponding result for an “opposite” adsorbate, namely one with high electronegativity and a strong covalent adsorbate-substrate interaction. The adsorbate-induced DOS (as shown in Fig. \[tb-DOS\]c, d) enables us to decide on the nature of the bond. Assuming that the antibonding states are at least partially empty we have chemical bonding of either covalent or ionic character. When the states of the occupied peak are predominantly derived either from adsorbate or substrate orbitals the bonding character is called “ionic”. When the wave-function character of the occupied states is derived by about the same amount from the substrate and the adsorbate the bond is called “covalent”. The problem with this classification is in the assumption that a substrate and an adsorbate region can be defined and separated. However, for strong chemisorption situations it is typically not obvious how to assign electron density to the different atoms. Still the concepts are valuable and we will come back to them in Section \[sec:nature\]. We note in passing that an intriguing way to assign electron density to individual atoms in a many-atom system was developed by Bader (1990, 1994, and references therein), but it is only now that these concepts are being used in solid-state calculations. The concept behind the qualitative discussion presented in this section is essentially that of the Anderson-Grimley-Newns model for chemisorption (see, for example, N[ø]{}rskov, 1990; Spanjaard and Desjonquères, 1990). The main difference is that free-atom levels are replaced by [*renormalized*]{} levels, that we have not introduced a restrictive assumption about electron correlation and the localization of the adsorbate-substrate interactions, and that we kept the discussion at a qualitative level. We believe that the approach sketched in Figs. \[tb-DOS\], \[tb-levels\] is useful but should not be overinterpreted. It was used, for example, for a qualitative discussion of H$_2$ dissociation at transition metal surfaces (see Hammer and Scheffler, 1995; and Section \[sec:reactions\]). N[ø]{}rskov et al. (see Hammer and N[ø]{}rskov, 1997, and references therein) refined the approach and used it for a (semi-) quantitative discussion of adsorption at various transition metal surfaces. The results confirm the simple chemical picture that adsorption energies decrease with the $d$-band filling of the substrate (for $N_d \ge 5$). And when the $d$-band is full, the $d$-electrons only play little role; thus the adsorption energy is small, chemical activity is low, and the substrate is called noble. In fact, Hammer and N[ø]{}rskov recently performed extensive DFT calculations of atomic and molecular adsorption at several metal surfaces and then used the above described approach to explain the main trends of bonding to metal surfaces in terms of the $d$-band filling. The crucial term in their approach is the adsorbate-substrate coupling matrix element. The latter depends on the adsorbate geometry (site and interatomic distances); unfortunately, but also quite obviously, this cannot be obtained from a simple Anderson-Gimley-Newns type of description. As a warning remark we add that energy levels and the density of states, $N(\epsilon)$, are important ingredients of the total energy, but from the total-energy contribution $$E^{\rm bands}= \int\limits^{+ \infty}_{- \infty} f(\epsilon) N(\epsilon) \, \epsilon \, d\epsilon, \label{bands}$$ it is typically not possible to decide on the strength of chemical bonding. The situation is different, when different geometries are compared and a frozen-potential approximation applied (the Andersen-Pettifor force theorem), as then (and only then) the other terms in the DFT total-energy expression cancel and [*energy differences*]{} based on Eq. (\[bands\]) may give an approximate description (see Skriver, 1985, and references therein). In this context we note that many-atom systems tend to assume a geometry with an electron density as hard as possible, which means that either a band gap is opened or the density of states at the Fermi level is reduced. This may be viewed as a generalization of the Jahn-Teller theorem. In this sense $N(\epsilon)$, in particular its behavior at the Fermi level, is instructive. ### Adsorbate band structure {#adsorbate-bands} To discuss covalent interactions between adparticles on a surface, we consider an ordered adlayer which is in registry with a crystalline substrate. Such a system has two-dimensional translational symmetry and the eigenfunctions are two-dimensional Bloch states which are defined by their reduced ${\bf k}_{\parallel}$-vector (in the surface Brillouin zone) as well as by their energy. (0,6.0) Figure \[fig:SBZ-1\] shows the surface Brillouin zones for the clean surface and for a c$(2 \times 2)$ adsorbate overlayer on the (001) face of an fcc metal. To obtain the bulk band structure projected onto the surface of a clean semi-infinite system, it is necessary to integrate over $k_{z}$ inside the bulk Brillouin zone and to shift regions outside the first two-dimensional Brillouin zone into its interior by a reciprocal lattice vector of the (two-dimensional) surface structure. This folding back of the ${\bf k}_{\parallel}$-resolved density of states of a clean, unreconstructed surface is indicated in Fig. \[fig:SBZ-2\]a. (0,5.5) The broken lines show the projection of the bulk Brillouin zone. Even for the clean surface, it is found that the surface Brillouin zone is usually smaller than the projection of the three-dimensional bulk Brillouin zone on the surface. For a periodic adlayer at fractional coverage, the surface Brillouin zone is further reduced compared with the clean surface, which again requires a shift of the ${\bf k}_{\parallel}$-resolved density of states of regions outside the new surface Brillouin zone into its interior, as shown in Fig. \[fig:SBZ-2\]b. Because the back-folded density of states adds to the original one, the change in the ${\bf k}_{\parallel}$-resolved density of states is obviously quite large. We note, however, that this folding back is a purely mathematical effect. It takes into account the increase of the periodicity parallel to the surface but it is independent of the strength and the type of the physical mechanism causing this change in periodicity. The physical importance of this effect, i.e., the degree of the mixing of the wave functions involved in the surface region and in turn the question how strongly this effect will affect, for example, photoemission spectra, depends on the strength of the adsorbate-substrate interaction. We return to this back-folding effect in the discussion of Figs. \[bands-1\] and \[bands-2\] of Section \[sec:LT-bands\] below. The dispersion $\epsilon({\bf k}_{\parallel})$ of the chemisorption-induced bands can be described in a simple tight-binding picture. It is appropriate to use a LCMO (linear combination of molecular orbitals) description where the MOs contain the interaction with the substrate as well as the intra-adparticle mixing of wave functions. Because this method is very simple and gives a reasonable qualitative description, we consider it in more detail. Two-dimensional Bloch states are used as a basis $$\begin{aligned} \chi_{\alpha}({\bf r}, {\bf k}_{\parallel}) = \frac{1} {\sqrt{N^{\rm nuc.}}} \sum^{N^{\rm nuc.}}_{{\bf R}_I} e^{i {\bf k}_{\parallel} {\bf R}_{I}} \phi_{\alpha}({\bf r} - {\bf R}_I), \label{eq:basis}\end{aligned}$$ where $\alpha$ labels the different (molecular) orbitals $\phi$ in the unit cell. The ${\bf R}_{I}$ are two-dimensional lattice vectors. The wave functions of the system are thus given by $$\begin{aligned} \varphi (\epsilon,{\bf k}_{\parallel},{\bf r}) = \sum_{\alpha} C_{\alpha}(\epsilon, {\bf k}_{\parallel}) \chi_{\alpha}({\bf r},{\bf k}_{\parallel}).\end{aligned}$$ For the Hamilton matrix we find $$\begin{aligned} H_{\alpha, \alpha^{'}}({\bf k}_{\parallel}) = \langle \chi_{\alpha}({\bf k}_{\parallel})|H|\chi_{\alpha^{'}} ({\bf k}_{\parallel}) \rangle = \sum^{N^{\rm nuc.}}_{{\bf R}_I} e^{i{\bf k}_{\parallel}{\bf R}_{I}} H_{\alpha,\alpha^{'}}({\bf R}_{I})\end{aligned}$$ with $$\begin{aligned} H_{\alpha, \alpha^{'}}({\bf R}_{I}) = \langle \phi_{\alpha} ({\bf r})|H| \phi_{\alpha^{'}}({\bf r} - {\bf R}_{I}) \rangle .\end{aligned}$$ Because the basis \[eq:basis\] is usually not orthogonal, we get an overlap matrix $$\begin{aligned} S_{\alpha, \alpha^{'}}({\bf k}_{\parallel}) = \langle \chi_{\alpha}({\bf k}_{\parallel}) | \chi_{\alpha^{'}} ({\bf k}_{\parallel}) \rangle = \sum^{N^{\rm nuc.}}_{{\bf R}_I} e^{i{\bf k}_{\parallel}{\bf R}_{I}} S_{\alpha, \alpha^{'}}({\bf R}_{I}) \quad\end{aligned}$$ with $$\begin{aligned} S_{\alpha, \alpha^{'}}({\bf R}_{I}) = \langle \phi_{\alpha} ({\bf r})|\phi_{\alpha^{'}}({\bf r} - {\bf R}_{I}) \rangle .\end{aligned}$$ The Schr[ö]{}dinger equation is then $$\begin{aligned} \sum_{\alpha^{'}} \left\{ \sum^{N^{\rm nuc.}}_{{\bf R}_I} e^{i{\bf k}_{\parallel} {\bf R}_{I}} [H_{\alpha,\alpha^{'}}({\bf R}_{I}) - \epsilon({\bf k}_{\parallel})S_{\alpha, \alpha^{'}}({\bf R}_{I})] \right\} C_{\alpha^{'}}(\epsilon, {\bf k}_{\parallel}) = 0 \label{eq:matrix-SG}\end{aligned}$$ and must be solved at each ${\bf k}_{\parallel}$. The zeros of the determinant of the matrix in the curly brackets in Eq. (\[eq:matrix-SG\]) give the dispersion $\epsilon({\bf k}_{\parallel})$. The required matrix elements $H_{\alpha, \alpha^{'}} ({\bf R}_{I})$ and $S_{\alpha, \alpha^{'}}({\bf R}_{I})$ can be calculated numerically in some approximation (Bradshaw and Scheffler, 1979; Horn et al., 1978; Scheffler et al., 1979; Jacobi et al., 1980). Often, an empirical tight-binding calculation might be sufficient, which introduces the following three assumptions: (1) Orbitals $\phi_{\alpha}$ at different centers are orthogonal, the overlap matrix $S_{\alpha, \alpha^{'}}({\bf R}_{I})$ is thus equal to one for $\alpha = \alpha^{'}$ and ${\bf R}_{I}$ = (0, 0) and zero otherwise. (2) Only nearest neighbors (sometimes also second nearest neighbors) are taken into account in the sum over ${\bf R}_{I}$ in Eq. (\[eq:matrix-SG\]). (3) The matrix elements are fitted to experimental results; usually there are only very few remaining. We will illustrate this for the example of an adsorbate with $s$-like states in a c$(2 \times 2)$ overlayer on an fcc (001) surface ($p_z$-like states would behave the same way). The corresponding adsorption-induced MOs $\phi_{\alpha}$ belong to the $a_1$ representations of the $C_{4v}$ point group (we assume that the adsorbate occupies a fourfold symmetric site). We then have two parameters in Eq. (\[eq:matrix-SG\]): $ H_{a_{1}, a_{1}}(0)$ and $H_{a_{1}, a_{1}}({\bf R}_{I})$. Figure \[fig:s-like-Bloch-states\] (0,4.0) shows schematically the Bloch states according to Eq. (\[eq:basis\]) for the three high symmetry points of the surface Brillouin zone (see Fig. \[fig:SBZ-1\]) $$\begin{aligned} \overline{\Gamma} : {\bf k}_{\parallel} = (0,0), \quad \overline{\rm X} : {\bf k}_{\parallel} = \frac{g}{2}(1,0), \quad \overline{\rm M} : {\bf k}_{\parallel} = \frac{g}{2}(1,1), \label{eq:k-points}\end{aligned}$$ where $g$ is the length of the first two-dimensional reciprocal lattice vector of the adlayer. Figure \[fig:s-like-Bloch-states\] has been constructed by placing orbitals of $s$-like symmetry at all the atoms (or molecules) of the adlayer (only 9 of them are shown in the figure). The phase factors of orbitals at different sites is $e^{i{\bf k}_{\parallel}{\bf R}_{I}}$ corresponding to the appropriate ${\bf k}_{\parallel}$-point in the surface Brillouin zone. An analysis of this figure already yields the qualitative band structure. At $\overline{\Gamma}$ the $a_{1}$-derived two-dimensional Bloch state is completely bonding in the adlayer, giving this state the lowest energy. At $\overline{\rm M}$ it is completely antibonding (highest energy), and at $\overline{\rm X}$, it is of mixed character. Thus, $s$-like adsorbate states give rise to an energy band which has the lowest energy at $\overline{\Gamma}$ and the highest energy at $\overline{\rm M}$. The analogous discussion for $p_x$-, $p_y$-like states can be found in an earlier review article by Scheffler and Bradshaw (1983). To discuss the interaction of an adsorbate with the $s$-band of the substrate, we consider in (0,8.0) Fig. \[fig:O-on-jellium-bands\] as an example the adsorption of oxygen on jellium corresponding to a ($1 \times 1$) overlayer on Al(111) (Hoffmann et al., 1979), which has been calculated using the two-dimensional KKR method (Kambe and Scheffler, 1979). The crystallographic point group is $C_{6v}$ because the jellium model neglects the atomic structure of the substrate. Full lines show the center of the peaks; hatched regions indicate the width of the peaks. The strong dispersion of the levels is quite apparent. The width of the band structure (the dispersion of the full lines) indicates the extent of the splitting between states which are bonding with respect to neighboring adparticles and those which are antibonding, just as in the case of the tight-binding scheme discussed above. At $\overline{\Gamma}$ the $p_{x}$- and the $p_{y}$-derived Bloch states are degenerate because of the $C_{6v}$ symmetry and the splitting between the $p_{z}$- and $p_{x}, p_{y}$-derived states is very small. The O $2p_{z}$-induced level is broad but that derived from $p_{x}, p_{y}$ is sharp. This different behavior of the two levels can be explained in the following way. For an ordered overlayer in registry with the substrate, two-dimensional Bloch states of the overlayer can only hybridize with those of the substrate which have the same ${\bf k}_{\parallel}$ vector. The resulting wave functions inside the jellium can thus be given as a sum over plane waves propagating toward the surface and the reflected waves. $$\begin{aligned} \varphi(\epsilon, {\bf k}_{\parallel},{\bf r}) = \sum_{\bf g} U^{+}_{\bf g} e^{i{\bf k}^{+}_{\bf g} {\bf r}} + U^{-}_{\bf g} e^{i{\bf k}^{-}_{\bf g} {\bf r}}.\end{aligned}$$ Because the potential is constant inside the jellium, the ${\bf k}^{\pm}_{\bf g}$-vectors of these plane waves are given by $$\begin{aligned} {\bf k}^{\pm}_{\bf g} = \left( {\bf k}_{\parallel} + {\bf g} , \pm \sqrt{\frac{2m(\epsilon - V_{0})}{\hbar^2}} - ({\bf k}_{\parallel} + {\bf g})^2 \right),\end{aligned}$$ where ${\bf g}$ are reciprocal lattice vectors of the (two-dimensional) surface Brillouin zone. Obviously, for ${\bf k}_{\parallel}$ = (0, 0) and a wave function which is antisymmetric with respect to a mirror plane of the system, the coefficients $U^{+}_{(0, 0)}$ and $U^{-}_{(0, 0)}$ vanish. At energies close to the bottom of the conduction band of the substrate, all the remaining terms have an imaginary value for $k_z$, which implies that these states decay exponentially normal to the surface. These are true surface states and have a sharp energy. In other words, we could say that at the bottom of the band the substrate states are $s$-like and thus do not have the correct symmetry in order to hybridize with $p_{x}, p_{y}$-like adsorbate levels at ${\bf k}_{\parallel}$ = (0, 0). Only at higher energies could these states also couple to bulk bands and broaden. Figure \[fig:O-on-jellium-bands\] also shows that the levels change their width with ${\bf k}_{\parallel}$ and become discrete when they lie outside the metal conduction band. We note the behavior of the symmetric $p_{x}, p_{y}$-derived band at the point at which it begins to hybridize with the substrate wave functions: just outside the metal conduction band region it bends slightly to lower$^4$ energies. Here it is purely bonding in character. At the same ${\bf k}_{\parallel}$ value, we note that the corresponding antibonding (broad) level occurs in the conduction band. When the band enters the conduction band, the bonding and the antibonding levels form one broad peak as mentioned above. Adsorption of isolated adatoms {#sec:nature} ------------------------------ Calculations of isolated adatoms afford an analysis of the nature of the adsorbate-substrate bond without it being obscured by the influence of other adsorbates. In this section we will therefore summarize characteristic results of the adsorption of isolated atoms. The calculations presented in the remaining part of this section were performed with the Green-function method, which provides the most accurate and efficient approach for calculating properties of truly isolated adatoms on extended substrates. For technical details of the method we refer the interested reader to the original papers (see in particular Bormet et al., 1994a; Lang and Williams, 1978; Bormet et al., 1994b; Wenzien et al., 1995; Scheffler et al., 1991; and references therein). We consider group I, group IV, and group VII adsorbates, namely Na, Si, and Cl. By such a trend study of atoms from the left to the right side of the periodic table, a classification of the nature of the bond becomes rather clear, but we will also emphasize (again) the limitation and/or danger of “nature-of-bond” concepts. Three different substrates will be considered, namely, - jellium with an electron density corresponding to aluminum, - Al(111), - Cu(111), which enables us to identify the role of substrate $s$-, $p$-, and $d$-states. For the Al(111) and Cu(111) systems the adsorbates were placed in the fcc-hollow site. In this section we discuss on-surface adsorption; substitutional adsorption, where the adatom replaces a surface atom, will be treated in Section \[sec:substitutional\], and subsurface adsorption will be briefly touched upon in Section \[sec:catalytic\]. ### Geometry Table \[tab:radii\] gives the adsorbate heights $Z$ and effective ----------- --------- --------- --------------------- --------- --------- --------------------- --------- --------- --------------------- Substrate $Z$ (Å) $R$ (Å) $\Delta R/R_{\rm J} $Z$ (Å) $R$ (Å) $\Delta R/R_{\rm J} $Z$ (Å) $R$ (Å) $\Delta R/R_{\rm J} $ $ $ Jellium 2.79 1.82 0 2.37 1.46 0 2.52 1.59 0 Al(111) 2.69 1.73 –5% 1.95 1.13 –22% 2.09 1.24 –22 % Cu(111) 2.43 1.56 –14% 1.68 0.96 –34% 1.78 1.03 –35 % ----------- --------- --------- --------------------- --------- --------- --------------------- --------- --------- --------------------- : Calculated geometrical parameters for adsorbed atoms from the left to the right side of the periodic table, and for different substrates. The height $Z$ is defined with respect to the center of the top substrate layer. The effective radii of the adatoms $R$ are evaluated by subtracting from the calculated bond lengths the radius of a substrate atom (as given by the inter-atomic distances in the bulk). Thus, we use $R_{\rm J} = R_{\rm Al} = 1.41$ Å and $R_{\rm Cu} = 1.27$ Å. For the jellium substrate we assume the geometry of Al(111). Also noted as a percentage is the difference of the adatom radii with respect to that of the jellium calculation: $( R_{\rm J}-R ) / R_{\rm J}$. The jellium results are from Lang and Williams (1978), the Al(111) results are from Bormet et al. (1994a), and the Cu(111) results are from Yang et al. (1994). []{data-label="tab:radii"} radii $R$ of the adsorbates, as obtained from the distances between the adatom and its nearest-neighbor substrate atoms, using the substrate atomic radii from the bulk. Both the calculated heights and the effective radii (the bond length is given by the sum $R + R_{\rm Al}$ or $R + R_{\rm Cu}$) are found to be noticeably smaller for Al(111) than for the jellium substrate. This is due to the fact that jellium is devoid of any information concerning the atomic structure. Thus, the shell-like property of electrons, i.e., their $s$-, $p$-, $d$-like character is missing which hinders the formation of directional orbitals in the substrate; and aluminum, despite its reputation of being a jellium-like system, is in fact a material with noticeable ability for covalent bond formation. On comparison of the results for adsorption on the Cu and Al substrates, it is found that the trend from jellium to aluminum continues: The effective radius of the studied adsorbates on Cu(111) is smaller than on Al(111). The smaller values reflect that bonding of the adsorbates is stronger on Cu(111) than on Al(111), which was identified as being due to the Cu $d$-electrons (Yang et al., 1994); although the top of the Cu $d$-band is about 2 eV below the Fermi level, the $d$-states play a noticeable role. The results demonstrate that stronger bonds go together with shorter bondlengths. As noted above in Table \[tab:radii\] we considered a threefold coordinated adsorption site. For other sites with lower coordination the strength per individual bond will typically increase, because the same number of adsorbate electrons have to be distributed (by tunneling or hopping) over fewer bonds. As a consequence, the bond length will typically decrease. This does not mean that the binding energy will increase as well, because this is determined by [*all*]{} bonds. This correlation between local coordination and bond strength, and the correlation between bond strength and bond length is well known (see, e.g., Pauling, 1960; Methfessel et al., 1992a). But we also emphasize that when significant changes in hybridization occur for different geometries, and/or when the system cannot attain the geometry of optimum bond angles, this simple picture breaks down. ### Density of states $\Delta N(\epsilon)$ {#sec:Delta-N} As emphasized in Sections \[sec:intro-1\], \[sec:DOS\], and \[sec:tight\], inspection of the adsorbate-induced change in the density of states $\Delta N(\epsilon)$ is particularly informative. (0,11.0) Together with the knowledge of the position of the Fermi energy, it tells us whether the electronic states which are formed upon adsorption are occupied, unoccupied, or partially occupied, and this enables us to discern the nature of the chemical bond. Figures \[Na-Si-Cl-DOS\]a and b show such results for adsorbates on Al(111) and on jellium. The adsorbates investigated in the work of Bormet et al. (1994a) (Fig. \[Na-Si-Cl-DOS\]a) were Na, Si, and Cl, and in that of Lang and Williams (1978) (Fig. \[Na-Si-Cl-DOS\]b) they were Li, Si, and Cl. The results of both of these calculations agree qualitatively and show the following: For both alkalis (Na and Li) the adsorbate resonance lies well above the Fermi level and is thus largely unoccupied. This indicates that the valence electron of the alkali metal atom (or part thereof) has been transferred to the substrate and the adatom is partially positively charged. In an opposite manner, on adsorption of chlorine, the resonance in the curve corresponding to the Cl $3p$ resonance lies 5eV below the Fermi level, and the Cl 3$s$ peak is positioned even below the substrate valence band at about $-18$eV, i.e., outside the energy range displayed in Fig. \[Na-Si-Cl-DOS\]. This result implies that a transfer of electron density from the substrate to the Cl adatom has taken place; the adsorbed Cl atom is partially negatively charged. Thus, Li and Na constitute examples of positive ionic chemisorption and Cl is an example of negative ionic chemisorption. For the adsorption of an isolated Si atom it can be seen from the jellium calculations that the Si $3p$ resonance lies just at the Fermi level, which implies that it is about half occupied. As noted in Sections \[sec:DOS\] and \[adsorbate-substrate-interaction\], the states of the energetically lower half of the resonance are bonding between the adatom and the substrate and the energetically higher states are antibonding. Because the Fermi level cuts the $p$-resonance approximately at its maximum, the bonding nature of Si is covalent, i.e., the bonding states are filled and the antibonding states are empty. The results for atomistic (Al and Cu) substrates also show that the bond is covalent. In this case, however, more structure occurs in the DOS than in the jellium calculations. This arises because the atomic structure of the substrate leads to band-structure effects, clearly reflected by the structure of the bulk DOS at $\epsilon_{\rm F}$ in Fig. \[Na-Si-Cl-DOS\]a. As a consequence, adsorption of a covalent atom, such as Si, results in a splitting of the bonding and antibonding states, and the adatom density of states exhibits a minimum at the Fermi level (Bormet et al., 1994a). Similarly to the jellium substrate, also for Si/Al(111) the Fermi level cuts the Si $3p$-induced DOS roughly in the middle. We note in passing that for Si on Al(111) (cf. Fig. \[Na-Si-Cl-DOS\]a) the structure of the $p$-like adsorbate density of states is largely determined by the clean-substrate density of states: With the Green function ${\cal G}^0$ of the clean substrate, we find at maxima of ${\rm Im}\left\{ {\rm Tr}\left( {\cal G}^0(\epsilon)\right)\right\}$ minima of $\Delta N(\epsilon)$, and at minima of ${\rm Im}\left\{ {\rm Tr}\left( {\cal G}^0(\epsilon)\right)\right\}$ we find maxima (see Bormet et al., 1994a, for more details). Despite the differences, which clearly exist between jellium and Al(111), we find for both systems that the adsorbate-induced change in the DOS confirms the expected picture for a chemisorbed adsorbate on a metal surface: - Atomic levels of the adsorbate are broadened due to the hybridization with the extended substrate states (in particular the substrate $s$-states). - Compared to the free atom DFT-LDA level[^5] the adsorbate-induced peak is found at higher$^{\ref{higher-lower}}$ energy for adsorbed Na, at about the same energy for adsorbed Si, and at lower energy for adsorbed Cl. The calculated shifts are $\Delta \epsilon({\rm Na}\, 3s) \approx +0.6$ eV, $\Delta \epsilon({\rm Si}\, 3p) \approx +0$ eV, $\Delta \epsilon({\rm Cl}\, 3p) \approx -0.5$ eV. This trend also conforms to the finding that the ionic character of the adatoms changes from plus to minus when going from Na to Si to Cl. It reflects that partially occupied levels have to align with respect to the substrate Fermi level (contribution $(i)$ discussed in Section \[sec:definitions\]). But fully occupied levels (such as the $3p$ states of the Cl$^-$ ion) are more affected by the surface potential (contribution $(ii)$ discussed in Section \[sec:definitions\]). For a core (or semi-core) state the shift of an adsorbate level upon adsorption mainly reflects the effects $(i)$ (due to the changed electrostatic potential caused by transfer or redistribution of valence electron density) and $(ii)$ (due to the substrate surface potential). With respect to the latter, we note that core levels are affected more by the electrostatic part of the potential than by the full effective potential, because the change in the exchange-correlation potential for core electrons due to the substrate electrons is negligible: At the high electron density in the core region, the exchange-correlation potential is only weakly affected by the relatively slight increase in electron density due to the substrate valence electrons. The shift of core levels due to effects $(i)$ and $(ii)$ of Section \[adsorbate-substrate-interaction\] is the initial-state contribution of X-ray spectroscopy of the adsorbate core-level shift. The results of Fig. \[Na-Si-Cl-DOS\] reveal that the Si $3s$ semi-core level is shifted by 3.4 eV and the Cl $3s$-level by only 1.67 eV toward lower$^{\ref{higher-lower}}$ energy. For Si the shift largely reflects how the adatom core states feel the substrate potential. The shift of Cl is smaller because the Cl is positioned further away from the surface than Si, and because the electron transfer toward the Cl adatom implies a repulsive potential for the core and semi-core states implying a contribution shifting their energy levels to higher$^{\ref{higher-lower}}$ energies. ### Electron density: $n({\bf r})$, $\Delta n({\bf r})$, and $n^\Delta ({\bf r})$ The trend seen in the results of Fig. \[Na-Si-Cl-DOS\] is in accordance with what is expected from electronegativity considerations: Na is electropositive with respect to the neighboring Al atoms, i.e., it gives up an electron more readily than Al. Cl, on the other hand, is strongly electronegative on Al and electron transfer from Al to Cl should occur. Silicon has nearly the same (or slightly higher) electronegativity as Al. (0,5.0) The formation of bonding and antibonding levels, together with the position of the Fermi level, will be reflected in the electron density $n(\bf r)$. With this hierarchy in mind it is useful to inspect [*in addition*]{} to the DOS (Fig. \[Na-Si-Cl-DOS\]) the electron density and, what is more instructive, the electron-density [*change*]{}, i.e., comparing the adsorbate system and the uncoupled systems. Figure \[Na-Si-Cl-e-density\] shows the electron density of the valence states. In the case of sodium, the charge transfer from the adsorbate toward the substrate is clearly visible. From the vacuum side the sodium looks practically naked. Figure \[Na-Si-Cl-e-density\] may overemphasize this impression because it shows a wide range of electron density in order to be able to compare atoms from the left to the right of the periodic table. The first displayed contour has a very low value which supports (again) the description of Na as being a (partially) ionized adatom and that particles which approach the adsorbed Na from the vacuum side will experience the “naked” side of the adsorbate. Figure \[Na-Si-Cl-e-density\] also shows that the electron density between the Na adatom and the Al substrate is increased. Thus, charge has been displaced from the vacuum side of the Na atom toward the substrate side. The details of this charge transfer are more clearly visible in the electron [*difference density*]{} $n^\Delta ({\bf r})$, which is the difference of the density of the adsorbate system, displayed in Fig. \[Na-Si-Cl-e-density\], of the density of the clean surface, and of the free atoms (cf. Eq. (\[n-Delta\])). This [*difference density*]{} is shown in Fig. \[Na-Si-Cl-delta-e-density\]. (0,9.0) The maximum of $n^\Delta({\bf r})$ is located between the adsorbate and the substrate. In a more detailed discussion given in Section \[sec:LT-screening\] we show that the shape of this induced charge density is the quantum-mechanical realization of the classical image effect which is actuated by the partially ionized adsorbate. For Si, as expected from the adsorbate-induced DOS, a directional covalent bond is present between the Si adatom and the nearest-neighbor substrate atom (see Figs. \[Na-Si-Cl-e-density\] and \[Na-Si-Cl-delta-e-density\]). Furthermore, it can be seen that the maximum of the charge density of the chemisorption bond is closer to the more electronegative Si atom. We also see a typical increase of electron density on the opposite side of the bonding hybrid. In the case of Cl on Al the charge density distribution around the adsorbate is almost spherical, again supporting the picture we had derived from the DOS in Fig. \[Na-Si-Cl-DOS\] of a (partially) negatively charged adatom. The results for the Al and Cu substrates show a number of similar features, and some interesting differences. Firstly, it can be noted that in each case the perturbation to the system caused by the adsorbates does not reach very far into the metal substrate. The interior is essentially identical to that of the clean surface for layers deeper than the second. We emphasize that this localization holds for the electron density [*perturbation*]{} but not for individual wave functions. For Na adsorption the results of Fig. \[Na-Si-Cl-delta-e-density\] imply the building up of a surface dipole which locally decreases the work function. The opposite situation is found for Cl, which is negatively charged and sits on an adsorption site that is positively charged. In this case charge has moved from the substrate toward the Cl atom, and the local work function therefore will be increased relative to the value of the clean surface. For both substrates, Si appears covalently bound with a slight electron transfer toward the adatom. Comparing in more detail the results for Al and Cu, we see that there is more charge between the Na adatom and the top layer of the Cu substrate. Also, there is a greater depletion of charge at the vacuum side. These effects are consistent with the fact that the electronegativity difference between Na and Cu $(0.93 - 1.9 = -0.97)$ is larger than that between Na and Al $(0.93 - 1.61 = -0.68)$. The Si atom, which clearly forms a covalent bond with both substrates, is slightly more electronegative than Al (by 0.27). In this respect, it can be seen that more charge resides on the Si atom when adsorbed on Al than when on Cu for which the electronegativity difference is zero. In all cases for the adsorbates studied, the adsorption on Cu exhibits some structure in the valence-electron density change near the nucleus of the Cu atom closest to the adsorbate, which reflects the participation of the Cu $d$-electrons in the bonding. ### Surface dipole moments A further interesting quantity obtainable in adsorption calculations is the change in the adsorbate induced dipole moment as a function ----------- ------ -------- ---------- Substrate Na Si Cl Jellium +0.4 +0.0 $ -0.5$ Al(111) +0.4 $-0.1$ $ -0.5$ \[0.1cm\] ----------- ------ -------- ---------- : Dynamic charge $d \mu/ d Z_{\rm ad}$, as obtained for isolated adatoms on jellium, and from supercell calculations for adsorbates on Al(111) at a low coverage of $\Theta=1/16$. The units are electron charges. The results are from Bormet et al. (1994a).[]{data-label="tab:dyn_mu"} of adsorbate height. In a naive picture one would expect that for ionic bonding the dipole moment $\mu$ changes linearly with the adatom height $Z_{\rm ad}$. For a covalent bond the dipole moment should be approximately constant (for small variations of the adsorbate height). Indeed we find that this picture applies. For Na we obtain a nearly linear increase of the dipole moment with increasing adsorbate height, and for Cl we obtain a decrease. The dynamical charge, which is the slope of $\mu(Z_{\rm ad})$, is given in Table \[tab:dyn\_mu\] and is in good agreement with the jellium calculations of Lang and Williams. These results support again the usefulness of the ionic pictures for Na and for Cl, and of the mainly covalent description of Si adatoms on Al(111). A homogeneously distributed layer of adatoms induces an electric field due to the adatom induced dipoles. The dipole strength is related to the adsorbate induced change in the work function, $\Delta\Phi_{\rm ad}$, by the Helmholtz equation: $$\begin{aligned} \Delta\Phi_{\rm ad} = \Theta \, \mu(\Theta)/ \varepsilon_0. \label{eq:type-3}\end{aligned}$$ We have seen (cf. Figs. \[Na-Si-Cl-e-density\] and \[Na-Si-Cl-delta-e-density\] ) that adsorption on a metal surface only significantly affects the electron density of the bare metal substrate in the outermost layers. The spatial distribution of the electronic charge with respect to the adsorbate nuclei gives rise to a dipole moment which changes the work function. Furthermore, charge transfer toward (or away from) the adparticle, as well as a permanent dipole moment give rise to a long range repulsive electrostatic interaction between different adparticles. If this is the dominant lateral interaction, which is true in some cases at low coverage, it will further result in a depolarization of the adsorption-induced dipole moment with increasing coverage (Antoniewicz, 1978; Topping, 1927; Kohn and Lau 1976). It is convenient to illustrate this depolarization effect by considering the physisorption of rare gases. For these systems there is practically no charge transfer, and we only have to consider the static adsorbate dipole. For not too small adsorbate-substrate separation, we can calculate, using classical electrostatics, the dipole moment $\mu$ which enters Eq. (\[eq:type-3\]). The dipole moment on each atom as a function of coverage is written as $$\begin{aligned} \mu(\Theta) = \mu_{\rm st} + \alpha {\cal E}_z (\Theta), \label{eq:p-theta}\end{aligned}$$ where $\mu_{\rm st}$ is the static dipole moment of the adparticle alone, i.e., without considering the effect of screening by the metal conduction electrons and for $\Theta \rightarrow 0$. ${\cal E}_z(\Theta)$ is the component normal to the surface of the microscopic electric field at the adparticle under consideration and $\alpha$ is its polarisability. To calculate the field ${\cal E}_z(\Theta)$, it is necessary to take the image dipoles into account. Thus, as well as the direct dipole-dipole interaction, the dipole-image and image-image interactions must be considered. The actual dipole moment at coverage $\Theta$ then is (0,8.0) $$\begin{aligned} \mu(\Theta) = \frac{\mu_{\rm st}}{1 + \alpha [T + V - 1/4 D^3]}, \label{eq:p-theta-2}\end{aligned}$$ where $D$ is the distance of the adparticle dipole to the effective image plane, and $T$ and $V$ are lattice sums of the direct and the indirect in teractions (see Scheffler and Bradshaw, 1983, for details). The dipole moment of an adparticle at zero coverage is thus given by $$\begin{aligned} \mu_0 = \frac{\mu_{\rm st}}{(1 - \alpha/4 D^3)}. \label{eq:p-0}\end{aligned}$$ Due to the screening by the substrate, the adatom-induced dipole moment ist thus enhanced from the value $\mu_{\rm st}$ to $\mu_0$. Equation (\[eq:p-theta-2\]) shows that with increasing coverage, i.e., increasing value of $T + V$, the particle is depolarized. Expressions similar to Eq. (\[eq:p-theta-2\]), but without the image terms $(V - 1/4 D^3)$ were proposed by Topping (1927) and Miller (1946). Equation (\[eq:p-theta-2\]) has been applied to the physisorption system Xe/Pd(001) (Bradshaw and Scheffler, 1979). Figure \[fig:Xe-Pd\] shows measured and calculated changes in work function. $D$ has been taken to be 1.5 Å, and for the polarizability the gas phase value, $\alpha$ = 4 Å$^3$, was used. For the monolayer obtained at $\Theta$ = 0.44, the lattice sums are given by $T + V$ = 0.17 Å$^{-3}$. The only remaining parameter is the dipole moment of the single adsorbed Xe atom, $\mu_{0}$. For the best fit to the experimental data, this is given by 0.93 Debye with the positive end away from the surface. At coverages between zero and $\Theta_{\rm max}$, the lattice sums are either proportional to $\Theta^{3/2}$ (Topping, 1927), if an ordered array is present at every coverage, or proportional to $\Theta$ (Miller, 1946), if the adlayer is disordered. Both cases yield similar behavior and are shown in Fig. \[fig:Xe-Pd\]. We thus see that the work function does not change linearly with coverage but that depolarization causes a less rapid change at higher coverages. Alkali-metal adsorption: the traditional picture of [***on-surface***]{} adsorption {#sec:LT-alkali} ----------------------------------------------------------------------------------- In this section we discuss the traditional picture of alkali-metal adsorption. This implies that we consider situations where alkali-metal atoms adsorb [*on the surface*]{} without disrupting the substrate very much, which is in contrast to [*substitutional*]{} adsorption (see Section \[sec:substitutional\]) where the adatom kicks out an atom from the substrate and takes its place. We recall that we discuss only the closed-packed substrate surfaces, namely fcc (111) and fcc (001). For more open surfaces, adsorbates, in particular alkali-metal atoms, can induce significant surface reconstructions. That substitutional adsorption at close-packed surfaces (and in more general terms, surface alloy formation) can happen even with adsorbates that do not mix in the substrate bulk, was first found by Schmalz et al. (1991) in a combined study of surface extended X-ray absorption fine structure (SEXAFS) experiments and DFT total-energy calculations. This subject will be discussed in Section \[sec:substitutional\] below. In this section we discuss [*on-surface*]{} adsorption and note only in passing that in the past, several experiments have been interpreted inappropriately by incorrectly assuming an on-surface adsorbate geometry. We will take care that only those experiments are discussed below for which the adatom indeed sits on the surface. In fact, attaining the substitutional adsorbate geometry is hampered by an energy barrier, and therefore for low-temperature adsorption, adatoms are likely to stay on the surface. For the example discussed below, in which Na is at Al(001), the transition temperature is $T = 160$ K. Substitutional adsorption has been observed for Na, K, and Rb on Al(111), where the tendency toward the substitutional adsorption seems to decrease with increasing adatom radius. Thus, for Cs substitutional adsorption has not been found. ### The Langmuir-Gurney picture {#sec:gurney} In the seminal work of Taylor and Langmuir in (1933) for the Cs/W system, it was shown that the work function of the clean surface is reduced by several electronvolts on Cs adsorption and that for thermal desorption at low coverage, almost all the Cs adatoms leave the surface as positive ions. In the light of such results, the alkali metal bond to the substrate was viewed as a spectacular example of ionic bonding akin to that in alkali halides (see also Naumovets, 1994). As discussed above (see Fig. \[nacl\]), even in NaCl the nature of bonding is debatable, and it is therefore not a surprise that the question whether or not alkali-metal adsorption should be described in terms of ionic bonding has sometimes been raised. With all the warnings mentioned in Section \[sec:intro\] in mind, we argue (and already did so in Section \[sec:nature\]) that the ionic picture is appropriate. For a more detailed analysis of the controversy about the ionic nature of the bond we refer to the publications of Scheffler et al. (1991) and Bormet et al. (1994a). A simple picture of the interaction between alkali-metal atoms and a metal surface, and of the resulting chemisorption bond was proposed by Langmuir (1932). He assumed that the alkali-metal atom transfers its valence $s$-electron completely to the substrate. In a more rigorous theoretical description of alkali-metal adsorption Gurney, in 1935, proposed a quantum-mechanical picture applicable at low coverages, where the discrete $s$-level of the free alkali metal atom broadens and shifts, becoming partially emptied as a result of the interaction with the substrate states as it approaches the surface. Figure \[gurney\] (left) summarizes this view. The partially positively charged adatom then induces a negative screening charge density in the substrate surface giving rise to an adsorbate-induced dipole moment which naturally explains the reduction in the work function of clean surfaces. The building up of such adsorbate-induced dipole moments also leads to the understanding that the dominant interaction between the adsorbates is repulsive and that the adatoms should form a structure with the largest possible interatomic distances compatible with the coverage. In this description it is expected that with increasing coverage the adsorbate-adsorbate distance gradually decreases, and the electrostatic repulsion between the adatoms increases. To weaken this repulsion, i.e., to lower the total energy, some fraction of the valence electrons flows back from the Fermi level of the metal to the adsorbate. Thus, a reduction of the adsorbate-induced dipole moment, i.e., depolarization takes place. This picture of Langmuir and Gurney is fully supported by systematic studies of Lang and Williams, Bormet et al., and others, some of which were discussed in Section \[sec:nature\] above. If the coverage-dependent depolarization is strong and/or the nature of the adsorbate-substrate binding changes with the local coverage, it is also possible (and observed) that with increasing coverage a phase separation occurs into close-packed islands (also called “condensation”) (Neugebauer and Scheffler, 1993; Over et al., 1995). ### Coverage dependence of the work function {#sec:LT-work-function} (0,6.5) In his earlier work, Lang (1971 and 1973) studied the alkali metal-induced work-function change using the jellium model for the substrate and for the adsorbate layer. Despite the extremely approximate nature of this approach, the results demonstrated that many ground-state properties of metal surfaces and metal-adatom systems can be described in a physically transparent manner. As an example, we show in Fig. \[fig:workfunction\] the change in work function with increasing coverage (cf. footnote \[coverage\]) for parameters corresponding to Na on aluminum (in the jellium approach; left side). The right side of the figure shows a calculation for Na on Al(001), which takes the atomic structure into account. We note that for coverages $ \Theta_{\rm Na}\,\, \raisebox{.6ex}[-.6ex]{$>\!\!\!\!$}\raisebox{-.6ex}[.6ex]{$\sim$}\,\, 0.15$ the on-surface geometry is a metastable structure which only exists at temperatures below 160 K (Andersen et al. 1992; Aminpirooz et al., 1992). The coverage dependence of the work function possesses a form similar to that often observed experimentally. It is explained as a consequence of the above-mentioned depolarization of the alkali metal-induced surface dipole-moment induced by continuous reduction of the adsorbate-adsorbate distance and corresponds to a rapidly decreasing work function at low coverage, reaching a minimum at about $\Theta=0.1$ (Fig. \[fig:workfunction\], left) or $\Theta=0.25$ (Fig. \[fig:workfunction\], right), and subsequently rising toward the value of the pure alkali metal. ### Ionization of the adsorbate and screening by the substrate electrons {#sec:LT-screening} We will now, for the moment, assume that we can neglect any effect due to the chemical interaction (i.e., hybridization of orbitals) upon adsorption and that the substrate mainly plays the role of providing an electron reservoir, i.e., fixing the electron chemical potential to its Fermi level. Using the Fermi-level position of Al(001), Fig. \[Na3s\] shows that then the occupancy of the Na $3s$-level will be 65%. Thus 35% of the Na $3s$-electron is now in the substrate. A partially ionized atom in front of a metal surface will not leave the metal electrons unaffected. To discuss this, we will at first sketch the quantum mechanical concept of screening at a metal surface which asymptotically, for large distances, includes the classical image effect. When a charged particle approaches a metal surface a screening charge will be created at the surface. The quantum-mechanical realization of the classical image effect is an induced electron density which has its center of gravity at the “effective metal surface”, i.e., the image plane (cf. Lang and Williams, 1978). Thus, whereas in the classical description the electrostatic screening of a [*point charge*]{} at a distance $D$ in front of a metal surface is modeled by an image point charge inside the metal at a distance $-D$ away from the surface, the truth is that there is no image point charge inside the substrate but a charge density right at the surface (see Fig. \[screening\]b), which creates the same electrostatic field in the vacuum as the (fictitious) point charge of classical electrostatics. (0,6.0) Considering close-packed surfaces, the image-plane position (see Eq. (\[eq:image-potential\])) is on the vacuum side of the center of the outermost atomic layer, because electrons spill out into the vacuum to lower their kinetic energy. At small distances (typically $ d < 2$[Å]{}) the justification of the image-plane concept gradually breaks down. However, it is still approximately valid, however, because metallic screening is still important. The concept can be approximately justified by assuming that for a charged system close to the surface, the image plane ($z_0$ in Eq. (\[eq:image-potential\])) is displaced toward the substrate. This is what we have in mind when we use the term “image effect” in the following discussion. For a deeper discussion of this problem we refer to Scheffler and Bradshaw (1983), Finnis et al. (1995), and references therein. Figure \[Na-Al(001)\] shows that the description of partial ionization of the Na adatom and building up a screening charge in the substrate remains [*qualitatively*]{} valid when the alkali-metal atom is adsorbed. Figure \[Na-Al(001)\] displays the [*difference density*]{} $n^\Delta({\bf r})$ (cf. Eq. (\[n-Delta\])) using a neutral atom ($f_{3s} = 1$). It can be seen that the upper half of the adsorbed Na atom has a negative $n^{\Delta}$ thus less electron density than the free, neutral alkali atom, and on the substrate side of the adatom there is an increase of electron density. Figure \[screening\] a shows a plot of $n^{\Delta}(f_{3s}=0.1)$, where $n^{\rm Na}$ (cf. Eq. (\[n-Delta\])) is now taken from a self-consistent calculation for the partially ionized free atom. (0,6.0) This picture looks very similar to the pure screening charge density which we obtain for an external point charge (see Fig. \[screening\]b). The shape of this screening charge density is very similar for a negative point charge and for a positive point charge, as long as the latter is sufficiently weak so that it cannot bind an electron. The magnitude of the contour lines nearly scale linearly with the magnitude of the external point charge. The similarity between Figs. \[screening\] a and \[screening\] b is quite surprising, because a Na-ion is – of course – different from a point charge. Nevertheless, the response of the metal is similar for both cases. The slight differences between Figs. \[screening\] a and b may be interpreted as an indication that some covalency and some $s$-$p_{z}$ mixing are present in the Na-Al(001) interaction, but these contributions are not very large. From the similarity of Figs. \[screening\] a and b we conclude that the “charge transfer picture” is indeed useful and appropriate to describe the physics of alkali adsorption at low coverage. However, charge transfer alone is not sufficient to understand the adsorbate-substrate interaction because the charge transfer actuates a significant change in the substrate surface electron density. We also note that the metal screening charge is largely located in front of the metal, i.e., between the adsorbate and the substrate (see Fig. \[screening\]b). As a consequence, the adsorbate-induced electron-density change cannot be divided [*directly*]{} into adsorbate and metal contributions. ### Surface band structure {#sec:LT-bands} At low coverages the interaction between adsorbed alkali-metal atoms will originate from their strong dipole moments. As the coverage increases, wave functions will overlap and a surface electronic band structure will form. This band is of $s$-character and its dispersion was discussed in Section \[adsorbate-bands\] above. In Fig. \[bands-1\] we show experimental (left panels) and calculated (right panels) surface states/resonances for (a) the clean Al(001) surface, (b) the on-surface c$(2 \times 2)$-Na/Al(001) structure, and (c) the substitutional c$(2 \times 2)$-Na/Al(001) adsorbate structure (discussed in more detail in Section \[sec:Na-Al(001)\] below). Firstly, we note that there is good overall agreement between theory and experiment. The main, lower-lying band in each case is Al-derived, as indicated by the circles in the theoretical plot. The bands represented by squares are Na-derived. (0,13.0) Compared to the position of the surface-state band of the clean surface, it can be seen that the Al-derived band of the on-surface adsorbate phase lies around half an eV lower in energy. The mechanism giving rise to this downward shift is discussed below. Interestingly, the experimental results show clearly that this state does not have the c$(2 \times 2)$ periodicity of the adsorbates, but rather has kept the $(1 \times 1)$ periodicity of the clean surface (cf. Fig. \[bands-1\]a). Indeed, inspection of the electron density reveals that the on-surface adsorbate represents only a modest perturbation of the surface electronic structure, which cannot be resolved in the photoemission experiments (see the discussion of Fig. \[fig:SBZ-2\] above). From the calculated bands of the on-surface adsorbate phase (right middle panel of Fig. \[bands-1\]), it can be seen that there is also a surface-state/resonance with an energy of about 0.7 eV below the Fermi level at $\overline{\Gamma}$. As indicated by the open squares, it is a Na-derived band (of Na 3$s$ character). It crosses the Fermi level and is partly occupied. From Fig. \[bands-2\], which shows the calculated band structure also in the region [*above*]{} the Fermi level, it can be seen that the part of the Na-derived band above $\epsilon_{\rm F}$ exhibits a band structure with c$(2 \times 2)$ periodicity of the adsorbed Na layer. Thus the higher coverage of Na present in this phase, as compared to the situation of an isolated adatom, has given rise to the development of significant adsorbate-adsorbate bonding and band formation. From construction of electron “difference density” plots (cf. Eq. (\[n-Delta\])), that is, subtracting off from the total valence electron density of the system, the superposition of the electron densities of the clean Al substrate and the Na atoms (arranged in the same c$(2 \times 2)$ periodicity), it is found that due to Na adsorption there is an increase of electron density at the position of the surface-state band (the band of parabolic shape which has a lower energy at 2.8 eV in Fig. \[bands-1\]a, left and 3.2 eV in Fig. \[bands-1\]b, left), and from the region between the Na atoms electron density is depleted. These results, together with comparison of the band structure of a [*free*]{} c$(2 \times 2)$-Na monolayer which has an electron occupancy [*larger*]{} than the Na-derived band seen in Fig. \[bands-1\], suggests electron transfer occurs from the Na atoms directly into the pre-existing surface states of Al(001). Thus, the Na-induced structures seen in Fig. \[bands-1\] b, left (see the dark regions at $\overline{\Gamma}$) originate from a coupling between the 3[*s*]{}-derived band of the free c$(2 \times 2)$ monolayer of Na and the surface-state band of the clean Al(001) surface. This may be understood as being due to the formation of bonding and antibonding states, leading to the downward shift of the Al-derived band with an increase in population and the upward shift of the Na-derived band with a decrease in population. Nevertheless, the electron density of the Al-derived band lies well below the smeared-out density of the Na-derived band, and also below the position of the Na atoms. Therefore the traditional picture of a thin metallic film covering a metallic substrate, as in the jellium model, remains qualitatively valid. We will see in the next section that this is [*not*]{} the case for the substitutional-adsorbate phase. Substitutional adsorption and formation of surface alloys {#sec:substitutional} --------------------------------------------------------- Recent experimental and theoretical studies of the adsorption and co-adsorption of alkali metals on Al surfaces have shown that the traditional view of alkali-metal adsorption outlined in Section \[sec:LT-alkali\] above, is in fact, only part of the whole picture and that phenomena such as the following may occur: 1. The alkali-metal atoms may not necessarily assume highly coordinated sites on the surface. 2. The alkali-metal adatom may kick out surface substrate atoms and adsorb substitutionally. Substitutional adsorption has been shown in three cases to occur as the result of an irreversible phase transition from an “on-surface” site by warming to room temperature, without change in the periodicity of the surface unit cell or of the coverage, i.e., order-preserving phase transformations between metastable and stable structures occur. 3. The alkali-metal atom may switch site on variation of coverage. 4. Island formation may occur. 5. There may be a strong intermixing of the alkali-metal atom with the substrate surface, as for example the formation of a four-layer ordered binary surface alloy which was identified for Na on Al(111) at a coverage $\Theta_{\rm Na} = 0.5$ (Stampfl and Scheffler, 1994c). Also ternary surface alloys form for the co-adsorption of Na ($\Theta_{\rm Na}$=0.25) with either of K, Rb, or Cs on Al(111) (Christensen et al., 1996). Only some of these aspects will be discussed below. A comprehensive description of the updated, modern view of alkali-metal adsorption can be found in a special issue of Surface Reviews and Letters (SRL, 1995) which collected review papers of the key groups which were working on this subject. There, and also in Adams (1996), it is shown in greater detail than possible here that the adsorption and co-adsorption of alkali metals on Al surfaces exhibit a wealth of previously unexpected phenomena. (0,4.0) Although it is well known that many, though not all, materials mix and form alloys, up until recently intermixing was not considered to be very relevant for adsorption on close-packed surfaces. Thus, adsorbates on such surfaces were assumed to occupy on-surface sites. Figure \[Na-Al(001)geometry\]a gives an example for Na on Al(001). It was also assumed that this view is justified, in particular, for systems where the adatom is insoluble in the bulk substrate. This standard picture was questioned in 1991 by a combined study of surface-extended X-ray adsorption fine-structure (SEXAFS) experiments and DFT total-energy calculations for the system Na on Al(111) (Schmalz et al., 1991); by now it is well established that often adatoms do not adsorb on the surface, but instead, it can be energetically favorable that they kick out atoms from the substrate and take their sites. The kicked out atoms then diffuse to a step (see Fig. \[diffusion-and-steps\]a). The process may be (0.0,4.0) kinetically hindered, but this hindrance may be overcome even at rather low temperatures, e.g., for Na on Al(001) for $T \ge 160$ K (Andersen et al. 1992; Aminpirooz et al., 1992) and/or by the heat of adsorption. Whether the solubility in the substrate bulk is low or even zero is of no relevance at all for substitutional adsorption. Atoms, which are too big to fit into a bulk vacancy, can still prefer to take a substitutional site, because at the surface bigger atoms can simply sit somewhat above the center of the created surface vacancy. Since 1991 many examples of substitutional adsorption have been reported, as for example K on Al(111) (Neugebauer and Scheffler, 1992, 1993; Stampfl et al., 1992, 1994a), Na on Al(001) (Stampfl et al. 1994b), Au on Ni(110) (Pleth Nielsen, 1993), Sb on Ag(111) (Oppo et al., 1993), Co on Cu(111) (Pedersen et al., 1997), Mn on Cu(001) (Rader et al., 1997), Co on Cu(001) (Nouvertné et al., 1999), to name a few. Thus, the phenomenon is not at all exotic, but rather general. In this section we will discuss in particular three systems, Na on Al(001), Na on Al(111), and Co on Cu(001), as these exhibit qualitatively different behavior. Alkali metal-induced surface reconstructions are well known on the more open surfaces (Somorjai and Van Hove, 1989; Barnes, 1994; Behm, 1989) as these clean surfaces are close to a structural instability (Heine and Marks, 1986). But that significant reconstruction can occur on close-packed fcc (111) and (001) surfaces had not been expected previously.[^6] As a warning for future work we mention that for alkali-metal adsorbates, several studies, had found “good agreement” between LEED experiments and theory as well as band-structure calculations and photoemission experiments (this also was the case for the example discussed in this section). However, it is now known that the reported agreement was purely coincidental, and the concluded physical and chemical properties were grossly incorrect because the alkali-metal atoms were not sitting on the surface (cf. Fig. \[Na-Al(001)geometry\]a), but in substitutional sites (cf. Fig. \[Na-Al(001)geometry\]b). Consequently, the nature of the adsorbate chemical bond, the surface electronic structure, and the origin of the coverage dependence of the work function are in fact qualitatively different to what was assumed before 1992 (Andersen et al., 1992; Aminpirooz et al., 1992; Stampfl et al., 1994b; Berndt et al., 1995). Thus, good agreement between theory and experiment and/or a convincing physical picture are no guarantee that the description and trusted understanding is indeed correct. The associated electronic properties of these surface atomic arrangements, perhaps not surprisingly, also exhibit behavior deviating from expectations based on early ideas. For example, experimental measurements of the change in work function as a function of alkali-metal coverage can be quite different to the “expected” form of Fig. \[fig:workfunction\] and the density of states induced by alkali-metal adsorbates may not correspond to that expected from the model of Gurney. ### Na on Al(001) {#sec:Na-Al(001)} (0,6.0) For Na on Al(001), the adsorption energy for the on-surface hollow site and the substitutional site as a function of adsorbate coverage is shown in Fig. \[Na-Al(001)-energy\]. It can be seen that the on-surface hollow site is clearly preferred over the substitutional site at low coverages. Its adsorption energy rapidly decreases with increasing coverage indicating a strong repulsive interaction between the Na atoms. The adsorption energy for Na in the surface substitutional site depends much more weakly on coverage than for the on-surface site, and, in fact, the adsorption energy for the $\Theta=0.5$ substitutional structure is more favorable than even that of the lower coverage substitutional structures. These results are interpreted as follows: At low coverages, the on-surface adsorption in a homogeneous adlayer is the stable structure for low as well as high temperature, but for higher coverages, on-surface adsorption becomes metastable. For high temperature adsorption, or warming the substrate, the adatoms then switch to substitutional sites, forming islands with a c$(2 \times 2)$ structure. The phase transition from the on-surface hollow site to substitutional adsorption is indicated in Fig. \[Na-Al(001)-energy\] by the dashed line. This predicted behavior is in good accord with experimental studies, see e.g., Fasel et al. (1996). The widely differing adsorbate geometries and the strong dependence of them on coverage and temperature, as described above, means that the type of bonding and chemical properties of the adsorption system will vary significantly depending on these factors. In this section we examine the electronic structure and bonding nature of the c$(2 \times 2)$ substitutional structure of Na on Al(001); that of the c$(2 \times 2)$ on-surface hollow structure was briefly touched upon already in Section \[sec:LT-bands\]. Firstly, it can be noted from Fig. \[bands-1\]c for substitutional adsorption that the main band in the experimental results clearly exhibits a c$(2 \times 2)$ periodicity, in contrast to the on-surface hollow structure. This is due to the significantly reconstructed Al(001) surface. It can also be noticed that this band is higher in energy than the surface state band of the clean Al(001) surface. Furthermore, the Na-derived band, as clearly observed for the on-surface phase in Fig. \[bands-1\]b in a greater region around $\overline{\Gamma}$ below $\epsilon_{\rm F}$, is nearly absent in the substitutional phase. However, from Fig. \[bands-2\]b which shows the same calculated band structure as in Fig. \[bands-1\]c, but where the energy region extends higher into the positive range, it can be seen that a significant Na-derived band (tagged by square symbols) lies well above the Fermi level, i.e., it is unoccupied. Figure \[bands-2\]c also shows the calculated surface state bands of the artificial c$(2 \times 2)$ vacancy structure. Here a strong Al-derived band can clearly be seen, its position being higher in energy than that of the clean surface. Similarly to the clean Al(001) surface, the maxima of charge density of the surface states for the vacancy structure lie on top of the uppermost surface Al atoms, having this time the c$(2 \times 2)$ periodicity. On adsorption of Na, as for the on-surface phase, this band shifts down in energy. The character of this Al-derived band is only slightly changed compared to that of the vacancy structure, which explains the well developed c$(2 \times 2)$ character of the band found in Fig. \[bands-1\]c. In this case electron transfer occurs from the Na atoms into the surface state/resonance of the [*vacancy structure*]{}. Figure \[den-diff-ht\] (right panel) shows that electron charge has been transferred mainly from the region on top of the Na atoms into the region above the Al atoms, corresponding to the position of the surface states. Compared to the Na-Al bond length of Na in the hollow site, that in the substitutional geometry is approximately 4% (from DFT-LDA) and 11% (from LEED) shorter, indicating a larger ionicity of the bonding. The results discussed above, together with Fig. \[den-diff-ht\] (middle panel), which shows the regions where electron density has been increased due to Na adsorption, demonstrates that for the [*substitutional*]{} adsorption phase the Na adlayer [*cannot*]{} be regarded as a simple metal film on a metallic substrate, and the jellium model is not valid. Another interesting property which yields insight into the bonding nature are the work function change and surface dipole moment. As discussed in Sections \[sec:gurney\] and \[sec:LT-work-function\], for [*on-surface*]{} adsorption the characteristic change in the work function $\Delta \Phi (\Theta)$ with coverage is typically explained in terms of the Gurney picture and the “usual form” can be described reasonably well by assuming a jellium model. (0,6.0) The work function change $\Delta \Phi$ and surface dipole moment $\mu $ are shown in Figs. \[na-al-sub-wf\]a and \[na-al-sub-wf\]b, respectively, as a function of coverage. For comparison, results for Na in the on-surface hollow structures are also displayed. As is consistent with the traditional picture of alkali-metal adsorption, there is a significant decrease of the surface dipole moment with increasing coverage for the on-surface site and the typical form of $\Delta \Phi (\Theta)$ is observed. The substitutional adsorption, on the other hand, exhibits a much weaker dependence. This reflects the fact that in the substitutional site the Na atoms sit lower in the surface and the repulsive adsorbate-adsorbate interaction is screened better. As discussed above (cf. Fig. \[Na-Al(001)-energy\]), under not too low temperature conditions a phase transition to c$(2 \times 2)$ islands occurs at coverage $\Theta \approx 0.15$; thus the local dipole moment will be fixed at the value of the c$(2 \times 2)$ phase and the work function change will vary linearly as $\Delta \Phi = \Theta \mu/\epsilon_{0}$. At $\Theta$=0.5 it is found that the values of $\mu$ (and $\Delta \Phi$) are the same for the on-surface and substitutional adsorption. This is in agreement with experiment, i.e., in all cases (theory and experiment) the value of $\Delta \Phi$ is 1.6 eV  (Porteus, 1974; Paul, 1987). ### Na on Al(111) {#sec:Na-Al(111)} The adsorption of Na on Al(111) was the first alkali metal on close-packed metal system that was discovered to assume substitutional adsorption. Further interesting and unanticipated phenomena were found to occur for higher Na coverages on this surface, in particular, the formation of a “four layer” surface alloy; the complex atomic geometry of which had foiled initial experimental attempts at its determination. This difficulty was related to its relative complexity, but was also due to conceptual barriers since such a structure was not expected to occur. In Fig. \[na-al111-ad-en\] the adsorption energy of Na on Al(111) is displayed for various structures and coverages of Na in on-surface geometries (Fig. \[na-al111-ad-en\]a) and in the substitutional site (Fig. \[na-al111-ad-en\]b). For the latter, the adsorption energy is split into its constituents, namely, the binding energy and the surface vacancy formation energy. (0,7.0) For on-surface adsorption, the theory indicates that for low coverages the hollow site is energetically most favorable and strong repulsive adsorbate-adsorbate interactions exist. A condensed $(4 \times 4)$ structure is seen to be energetically more favorable than homogeneous adlayers of Na for coverages larger than approximately $\Theta_{\rm Na}=0.1$. In the condensed phase, the coverage is $\Theta_{\rm Na}=9/16$ and the structure represents a densely packed hexagonal adlayer with nine Na adatoms per surface unit cell, all occupying different (mostly low symmetry) on-surface sites. Thus, the picture is that for very low coverages, the adsorbates occupy on-surface hollow sites and are uniformly distributed over the surface (homogeneous adlayers) but for coverages greater than about $\Theta_{\rm Na}=0.1$, island formation with the condensed structure and a $(4 \times 4)$ periodicity occurs as indicated by the dashed line. In this case it is energetically favorable to build up a metallic-like bonding between the adatoms and to reduce the ionic character of the adatom-substrate bonding. For the case of substitutional adsorption it can be seen from Fig. \[na-al111-ad-en\]b that the adsorption energy of the substitutional geometry is the most favorable for [*all*]{} structures investigated (note the different energy scales of Figs. \[na-al111-ad-en\]a and \[na-al111-ad-en\]b). In particular, substitutional adsorption with a $(\sqrt{3} \times \sqrt{3})R30^{\circ}$ periodicity has [*the*]{} most favorable adsorption energy. It is therefore expected that a condensation into $(\sqrt{3} \times \sqrt{3})R30^{\circ}$ islands with the Na atoms in substitutional sites occurs, beginning at very low coverages. The atomic structure is displayed in Fig. \[na-al111-geom\]. The repulsive character of the binding energy for the $(\sqrt{3} \times \sqrt{3})R30^{\circ}$ phase can be seen (cf. Fig. \[na-al111-ad-en\]b) to be over-compensated for by the attractive interaction of the surface vacancies. (0,4.5) (0,5.5) In Fig. \[na-sub-ddiff\] the difference between the electron density of the $(\sqrt{3} \times \sqrt{3})R30^{\circ}$-Na/Al(111) phase and that of the underlying vacancy structure plus a free standing Na layer with also $(\sqrt{3} \times \sqrt{3})R30^{\circ}$ periodicity is shown. It can be seen that due to Na adsorption, electron density has been displaced from the Na atoms toward the Al atoms of the substrate. As noted above, the reason for the favorable adsorption energy of Na in this structure is due to the particularly low vacancy formation energy; the reason for this has been attributed to the formation of covalent-like, in-plane bonding between the remaining top-layer Al atoms, the honey-comb arrangement of which is similar to that of graphite. The electronic structure of the reconstructed surface is found to be largely responsible for that of the $(\sqrt{3} \times \sqrt{3})R30^{\circ}$-Na/Al(111) phase. In particular, new states close to the bottom of the Al valence band are found as well as broad unoccupied bands. In this case, the role of Na is apparently mainly to create the vacancy but not to modify very much the electronic structure of the vacancies (Wenzien et al., 1993). On further deposition of 1/6 of a monolayer of Na onto the substitutional $(\sqrt{3} \times \sqrt{3})R30^{\circ}$-Na surface, a $(2 \times 2)$ structure forms with two Na atoms per unit cell. As discussed above, the atomic geometry of this phase proved initially difficult to determine. Its correct structure was first proposed on the basis of DFT-LDA calculations (Stampfl and Scheffler, 1994c) and was subsequently confirmed by a LEED intensity analysis (Burchhardt et al., 1995). From consideration of the atomic structure of the $(2 \times 2)$ phase, it would seem that no mass transport is necessary in its formation, i.e., there is an Al atom missing due to the substitutional adsorption of Na, but there is an additional Al atom embedded between the Na atoms. However, the lower coverage $(\sqrt{3} \times \sqrt{3})R30^{\circ}$ substitutional structure involves displacement of 1/3 of a monolayer of Al atoms, which are assumed to diffuse across the surface to be re-bound at steps. The results indicate, therefore, that formation of the $(2 \times 2)$ structure from the $(\sqrt{3} \times \sqrt{3})R30^{\circ}$ structure involves the reverse process, that is, diffusion of 1/3 of a monolayer of Al atoms back from the steps which are used in the formation of the $(2 \times 2)$ structure. This process is depicted in Fig. \[diffusion-and-steps\]b. Interestingly, in a manner similar to that discussed above for the two substitutional structures of Na on Al(001) and Al(111), the [*occupied*]{} part of the surface band structure of the ($2 \times 2$) Na-Al surface alloy can be explained largely in terms of the underlying Al structure. The latter corresponds to the reconstructed Al $(2 \times 2)$ vacancy layer [*plus*]{} the Al atom in the hcp-hollow site on this structure. In particular, a significant peak is observed at approximately 2 eV below the Fermi level. Analyzing the wave functions of this state at $\overline{\Gamma}$ shows that it is localized on top of the uppermost hcp-hollow Al atoms. For the surface alloy, in addition, unoccupied features are identified which are associated with the Na atoms, and at $\overline{\Gamma}$, they are centered above the uppermost Na atoms (Stampfl and Scheffler, 1994c). ### Co on Cu(001) {#sec:Co-Cu(001)} In the above sections it was shown that the substitutional adsorption of alkali-metal atoms is driven by the strong dipole moment of adatoms, a rather low formation energy of surface vacancies, and the fact that the adsorbate-adsorbate repulsion is reduced in the substitutional geometry. Thus, substitutional adsorption does not occur for single adatoms, but only when the coverage has reached a critical value where the adatom-adatom interaction becomes significant (cf. Fig. \[Na-Al(001)-energy\]). For cobalt adsorbed on Cu(001) the situation is different. Here the adatoms can assume a substitutional geometry at the lowest coverages and with increasing coverage a structural phase transition (0,6.5) occurs toward the formation of close-packed islands. The difference compared to alkali-metal adsorption is due to the fact that cobalt, as a transition metal from the middle of the $3d$ series, likes to involve its $d$-electrons in the chemical binding. This implies that Co likes to assume a highly coordinated site. A single Co adatom on a Cu(001) surface in the on-surface hollow site has four Cu neighbors. However, in the substitutional site it is embedded in the electron density provided by eight Cu neighbors. Thus adsorption in a substitutional site (i.e., in a surface vacancy) is clearly more favorable than adsorption in an on-surface site. In fact, the energy difference between on-surface adsorption and into-surface-vacancy adsorption is larger than the energy to create a surface vacancy. Thus, single Co adatoms tend to kick out Cu atoms from the surface and assume their sites. We note that the Cu atoms taken out of the surface are re-bound at kink sites at steps where they attain the Cu-bulk cohesive energy (cf. Section \[sec:kink\]). In this context it is also relevant that Cu adatoms on Cu(001) have a higher mobility than Co adatoms, which implies that thermal equilibrium with kink sites can be attained easily for Cu adatoms, but is hindered for Co adatoms. Figure \[Co-Cu-energy\] shows the adsorption energies for on-surface and substitutional adsorption of Co. For low coverages substitutional adsorption is energetically favorable. For higher coverage it is still more energetically favorable for an open adlayer to adsorb substitutionally than on-surface, but the energetically lowest configuration is that of a close-packed Co islands. In other words, Fig. \[Co-Cu-energy\] tells us that substitutionally adsorbed Co atoms form strong bonds with their eight Cu neighbors. However, the strongest bonds are achieved when Co adatoms form close-packed Co islands. Thus, for higher coverages, and/or higher temperature, when Co adatoms perceive the existence of other Co atoms on the surface, isolated, substitutionally adsorbed Co atoms are predicted to leave their site and close-packed Co islands will be formed. In fact, recent DFT calculations predict that these islands preferentially will have a thickness of 2-3 Co layers and are capped by a Cu layer (Pentcheva and Scheffler, 2000). Adsorption of CO at transition metal surfaces\ – a model system for a simple molecular adsorbate {#sec:CO} ------------------------------------------------- The adsorption of a diatomic molecule on a surface represents the next degree of complexity with respect to the adsorption of a single atom and serves as a link to understanding the behavior of more complex molecular adsorbates, as well as to the important area of carbonyl chemistry. (0,5.5) As such, CO adsorption has become a paradigm for the study of a simple molecular adsorbate on a surface and has been extensively studied both experimentally and theoretically, see, e.g., Hermann et al. (1987), Hoffmann (1988), Campuzano (1990), and references therein. This interest in CO adsorption also originates from the technological importance of oxidation catalysis (e.g., the car exhaust catalytic converter). (0,5.0) (0,5.5) The three outer valence orbitals of a free CO molecule are sketched in Fig. \[co-free-states\]. With decreasing ionization energies, these are the 5$\sigma$ orbital (largely C $2s$, C $2p_{z}$), the doubly degenerate 1$\pi$ orbital (largely C $2p_{x}$, $p_{y}$, O $2p_{x}$, $p_{y}$) and the 4$\sigma$ orbital (largely O $2s$, O $2p_{z}$). The first unoccupied state, shown at the far right in the figure, is the antibonding C $2p_{x}, p_{y}$, O $2p_{x}, p_{y} (2\pi^{*})$ orbital. The two most important orbitals are the $5\sigma$ and the $2\pi^*$ orbitals which correspond to the HOMO and LUMO, respectively (see Fig. \[co-free-states\]). The notation here is that “$\sigma$” indicates orbitals that are rotationally invariant with respect to the inter-nuclear axis and “$\pi$” represents orbitals that are lacking this symmetry. When CO is brought toward a metal surface the CO 5$\sigma$ is significantly perturbed by the hybridization with the substrate $d$-electrons. The energy of the 5$\sigma$ orbital changes most strongly because the bonding to the substrate is governed by the interaction of this orbital. This gives rise to charge transfer from the CO 5$\sigma$ orbital to the metal, but the metal gives charge back into the antibonding 2$\pi^{*}$-CO orbital. This is called the donor-acceptor model for CO bonding (Blyholder, 1964, 1975), which is known from the metal carbonyls and is similar to the results for H$_{2}$ adsorption (see the discussion of Fig. \[fig:H2-tight-binding\] in Section \[sec:H2\] below). The back donation from the substrate into the 2$\pi^*$-CO orbital weakens the bonding within the CO molecule and strengthens the bond to the substrate. At close distances the ordering of the 1$\pi$- and the 5$\sigma$-derived levels is reversed compared with the gas phase (see also the discussion of CO on Ni by Hermann and Bagus (1977), and of CO on Cu by Hermann et al. (1987)). In the Blyholder model the lower lying 4$\sigma$ and 1$\pi$ MOs (as well as the 3$\sigma$ and of course the core states) are assumed not to play an important role in the CO-metal bond formation. In the following we will use Ru(0001) as the substrate for the discussion of CO adsorption but note, that the basic results are valid valid for other transition-metal substrates as well. In the left panel of Fig. \[co-on-ru\] the valence electron density of CO in the on-top site on the Ru(0001) surface is shown, and in the right panel of Fig. \[co-on-ru\] is the difference between the electron density of the CO/Ru(0001) system and the superposition of Ru(0001) and free molecular CO. From the latter, the electron redistribution can be seen to be in good general agreement with the Blyholder donor-acceptor model: Depletion is clearly seen from the $\sigma$ orbitals of CO and an increase in electron density into the 2$\pi^*$ orbitals. Depletion can also be clearly noted from Ru states with $d_{z^{2}}$-like character, as well as a significant increase in electron density in the region of the adsorbate-substrate bond, i.e., between the C and Ru atoms. A similar behavior was found for CO on other substrates (see, e.g., Wimmer et al., 1985 and Bagus et al., 1986). It is pointed out that in addition, there is participation to the CO-metal bonding by the Ru atoms in the [*second*]{} layer. The Kohn-Sham eigenvalues of the free CO molecule shift noticeably upon adsorption on the Ru(0001) surface; in particular, a significant downward shift of the $5\sigma$ orbital energy due to hybridization with Ru states, and also a small downward shift of the $4\sigma$ level is observed. The $1\pi$-level is changed only little, and the $2\pi^{*}$-level moves up in energy reflecting the increased occupation. Also, correspondingly, the development of antibonding states occurs. Thus, the behavior follows that of Section \[adsorbate-substrate-interaction\]; and we realize that the effects $(ii)$ and $(iii)$ are apparently small for the on-top adsorbed CO. In Fig. \[co-states\] the spatial distribution of some of these CO-derived states are displayed. It can be seen that the $3\sigma$ orbital remains unperturbed (compare with Fig. \[co-free-states\]) by CO adsorption on the substrate since it lies significantly lower in energy and away from the surface. The $4\sigma$ orbital interacts with Ru $d_{z^{2}}$-like states, as does the strongly interacting $5\sigma$ orbital. The $1\pi$ orbital on the other hand interacts only weakly with the substrate. The unoccupied $2\pi^*$-orbital hybridizes with Ru states of $d_{xz}$ and $d_{yz}$-character. Certain of these adsorbate-substrate bonding states also have an antibonding partner (not shown), the weight of which resides predominately in the substrate. We see therefore that the first-principles calculations support in general the Blyholder model, but that the details of the bonding are somewhat more complicated; similar observations have been pointed out and discussed in more detail by Hu et al. (1995) for first-principles studies of CO on Pd(110) and from experiments by Nilsson et al. (1997). Co-adsorption \[the example CO plus O on Ru(0001)\] {#sec:co-adsorption} --------------------------------------------------- (0,10.5) A prerequisite to understanding heterogeneous catalytic reactions is knowledge of the behavior of the various reactants (e.g., the adsorption sites and binding energies), as well as their mutual interactions. The co-adsorption system ($m$CO+$n$O)/Ru(0001) represents a well-studied model system, not least due to the drive aimed at obtaining an understanding of the catalytic oxidation of CO by O$_{2}$, but also as a “simple” model system for oxidation catalysis in general. Despite the considerable interest, it is only recently that the detailed atomic structure of some of the phases of ($m$CO+$n$O) on Ru(0001) have been determined, and new ones discovered. Depending on the experimental conditions, co-adsorption of CO and O on Ru(0001) can form the following phases: $(2 \times 2)$-(1O + 1CO) \[Kostov et al., 1992; Narloch et al., 1995\], $(2 \times 2)$-(2O + 1CO) \[Narloch et al., 1994\], and $(2 \times 2)$-(1O + 2CO) \[Schiffer et al., 1997\]. These mixed ($m$CO+$n$O)/Ru(0001) surface structures are depicted in the top panel of Fig. \[co+o\]. For the first two structures, low-energy electron diffraction (LEED) intensity analyses have been performed: In the first phase, the O atoms occupy hcp sites and the CO molecule adsorbs in the on-top site. In the second phase, a restructuring induced by CO adsorption of the O atoms of the $(2 \times 1)$ (Lindroos et al., 1989) phase occurs: Half of the O atoms, initially occupying the hcp sites, switch to fcc sites and CO adsorbs again in the favored on-top site. For the (1O + 2CO) structure, there has been no LEED intensity analysis, but infrared absorption spectroscopy (IRAS) and X-ray photoelectron spectroscopy (XPS) experiments (Schiffer et al., 1997) indicate that the O atoms occupy hcp sites and the CO molecules occupy on-top and fcc sites. To obtain insight into the behavior of these co-adsorption systems, DFT calculations have been carried out. The calculated atomic geometries are displayed in the middle section of Fig. \[co+o\]. Good agreement with the LEED determined geometry was found for the first two of these phases for which comparison is possible (Stampfl and Scheffler, 1998). The calculations show that for the (2O + 1CO) structure, it is indeed energetically more favorable (by 0.59 eV) for half of the O atoms to occupy the less favorable fcc sites and CO to adsorb in the on-top site rather than maintaining the $(2 \times 1)$-O arrangement and CO adsorbing in a hollow site. In addition to those phases that have been experimentally identified, the theory predicts the stability of another phase (Stampfl and Scheffler, 1998), namely $(2 \times 2)$-(3O + 1CO)/Ru(0001), seen in the far right-hand-side of Fig. \[co+o\]. The adsorption energy of CO in this structure is notably weaker than for CO in the on-top site; it is, however, still appreciably exothermic with a value of 0.85 eV. An important consideration concerning whether a structure can in fact form, is the kinetics. The possibility of kinetic hindering due to energy barriers induced by the adsorbed O atoms was investigated by calculating the total energy of CO at various distances above the hcp-hollow adsorption site, i.e., above the vacant O site of the $(2 \times 2)$-3O/Ru(0001) structure. Incidentally, this structure has recently been shown to represent a new stable phase of O on Ru(0001) (Kostov et al., 1997; Kim et al., 1998; Gsell et al., 1998). The calculations show that there is an energy barrier to adsorption of $\approx$0.35 eV. This implies that rather high CO pressures would be required in order to realize this structure experimentally. Similar calculations were carried out for the (1O + 2CO) phase for the CO molecule above the fcc site. It was found in this case there is also an energy barrier, but slightly smaller of about 0.2 eV, thus (at least partially) explaining the low sticking coefficient and the necessary high exposures found experimentally in order to create this phase (Schiffer et al., 1997). The valence electron density of the various phases are also shown in Fig. \[co+o\]. The oxygen atoms appear as the red (i.e., high electron density), almost spherical features. Both CO and O can be seen to induce a significant redistribution of the electron density of the top-layer Ru atoms. For CO in the hollow sites, it is apparent that the bond strength (per bond) is weaker than that for CO in the on-top site. In the hollow sites, however, CO forms [*three*]{} bonds with the metal surface so it is expected that they be longer and weaker. The calculations show nevertheless that the adsorption energy is significantly weaker in the hollow sites than in the on-top site for these structures; this is not the case for the clean surface where the energy difference is only about 0.04 eV. Thus, this significant energy difference is a consequence of the co-adsorbates. These ($m$CO+$n$O)/Ru structures depicted in Fig. \[co+o\], each posessing the same periodicity but with varying numbers of species and adsorption sites, represent an ideal model co-adsorption series for study by first-principles calculations. From analysis of the results of such calculations much can be learnt about the various interaction mechanisms at play. Chemical reactions at metal surfaces {#sec:reactions} ------------------------------------ This section summarizes some basic aspects of the present understanding of the reactivity of surfaces. Here the term “reactivity” usually refers to the surfaces’ ability to break bonds of an approaching molecule and to adsorb the fragments, which is often the rate limiting step in catalytic reactions. For example, in the ammonia synthesis it is the dissociation of N$_2$, and for various examples of oxidation catalysis (e.g., the catalytic oxidation of CO) it is the dissociation of O$_2$ (see Section \[sec:catalytic\] below). Just having referred to “catalysis” a word of warning is appropriate because industrial catalysis involves many more aspects than just dissociation of a certain molecule. Other aspects are “selectivity”, which means that only the desired reaction should take place, and competing reactions yielding unwanted products are suppressed. Also the buffering of intermediate chemical products is important, as is the self-maintenance of the catalyst, the possible role of the catalyst’s support and of promotors. And (often) it is important that no poisonous by-products are released. ### The problem with “the” transition state {#sec:transition-state} We will discuss the surface reactivity in terms of molecules approaching the surface considering all relevant atomic coordinates. The total energy as function of the atomic coordinates is called the “potential-energy surface” (PES), cf. Section \[sec:energies\]. It represents the energy surface on which the atoms will move. Whereas the electrons are assumed to be in their ground state of the instantaneous geometry, the wave functions of the nuclei describe the details of the atoms’ dynamics, i.e., the vibrations, rotations, center-of-mass translation, the scattering at the surface, the dissociation, and the surface diffusion of the fragments. In order to keep the discussion simple we will discuss the situation of a molecular beam which is sent toward the surface, and in which the molecules have a certain center-of-mass kinetic energy, and are in a well defined vibrationally and rotationally excited state or the ground state. The probability of dissociation then is considered to be a measure for the surface reactivity. It contains - information concerning the surface electronic structure, i.e., on the relevance of so-called frontier orbitals (Wilke et al., 1996, and references therein), - information about the high-dimensional PES on which the approaching molecule travels toward the surface, - the statistical average over many trajectories which finally determines with what probability the active sites at the surface are found and the molecule will dissociate, or if it will be reflected into the gas phase. To discuss the dissociation probability of incoming molecules, it is often assumed that the reaction proceeds along (0,6.0) a one-dimensional (or low-dimensional) reaction coordinate and that it will cross a well defined “transition state” (cf. Fig. \[transition-state\]). Then the reaction probability is given by an Arrhenius behavior with the energy barrier given by this transition state. While the concept behind Fig. \[transition-state\] has proven useful (at least often) in gas-phase chemistry, it may be misleading for the description of surface chemical reactions. In particular, we note that the phase space for a molecule-surface reaction has very high dimensionality. For example, even in the simplest surface reaction, i.e., when H$_2$ molecules are sent toward the surface and if the substrate atoms do not move during the scattering event, the translations, vibrations and rotations of the two H atoms take place in a 6-dimensional configuration space, i.e., a 12-dimensional phase space. As a consequence, the assumption of “the” transition state, as depicted in Fig. \[transition-state\] can be grossly misleading. Instead, many transition states exist, and which of them is taken with what probability depends on the details of the incident H$_2$ dynamics, i.e., the H$_2$ translational kinetic energy, and the vibrations and rotations. Consequently, a good treatment of the statistics of the many possible trajectories is mandatory. As noted in Section \[sec:energies\] we will restrict the discussion in this chapter to situations where the Born-Oppenheimer approximation is justified. Still, a severe problem remains, namely that knowledge of the PES barely exists. Up until recently only rough, and as we now know, often incorrect semi-empirical models were used, and these earlier studies of the dissociation dynamics were restricted in their dimensionality. For example, in the past the dissociation of H$_2$ was described in terms of only 2 or 3 coordinates (out of the six important coordinates), and the dependence of the PES on the other coordinates was simply neglected. The only reason for this simplification was that evaluating a PES using good-quality electronic structure theory is elaborate. Since about 1994 the situation has changed, i.e., several groups started to take into account the higher dimensionality (see Gross and Scheffler, 1998, and references therein). Though involved, even treating the six-dimensions of the two hydrogen atoms is not yet complete because in general it could happen that electronic excitations play a role in the scattering event which is outside the Born-Oppenheimer approximation, and furthermore, it is well possible that the dynamics of the substrate atoms will play a role. These concerns may not be very important for the systems studied so far, but for other systems they may well be relevant. With respect to the validity of the Born-Oppenheimer approximation we are not aware of a serious breakdown, although sometimes it has been speculated. For adsorbates at metal substrates, levels are typically broadened, which implies that excited states have a short life time and thus relaxation into the ground state configuration will be (nearly) instantaneous on the time scale of the nuclear motion. Obviously, for insulators and semiconductors the situation is different (see, e.g., the discussion in Gross et al., 1997). Also, the situation would be different if laser excitations are involved (i.e., photo-chemistry), but this is not the topic of this chapter. ### Dissociative adsorption and associative desorption of H$_2$ at transition metals {#sec:H2} Dissociative adsorption (or the time-reversed process, which is associative desorption) is a dynamical process, and because of the high dimensionality of the PES, a proper treatment of the dynamics is indeed crucial. Typically the dynamics of atoms is treated classically, i.e., by Newton’s equation of motions, but the underlying PES and the forces acting on the atoms are (or could be and should be) calculated by DFT. This is what is called “[*ab initio*]{} molecular dynamics”. It started with the seminal work of Car and Parrinello (1985). The elegance of their approach appears to imply that typically it is not very efficient, and since their original paper several alternative (and numerically more efficient) formulations have been developed (see, e.g., Payne et al., 1992; Kresse and Furthm[ü]{}ller, 1996; Bockstedte et al., 1997; and references therein). When the moving nuclei are hydrogen atoms, it is often necessary to treat also the nuclei as quantum particles. For such problems a rather involved, high-dimensional “[*ab initio*]{} quantum dynamics” method has been implemented (Gross et al., 1995, 1998; Kroes et al., 1997; and references therein). This is probably the most advanced approach and for heavy particles (obviously) it becomes identical to [*ab initio*]{} molecular dynamics. In the following section we describe some general features of the PES and then we show examples which demonstrate the importance of a quantum dynamical treatment of scattering and dissociation of molecules at surfaces. The dissociative adsorption of H$_2$ appears to be a very simple reaction. However, due to the quantum nature of the hydrogen nuclei, the actual processes are rather complex. Again, as always in this chapter, we will keep the discussion simple, and for more details we refer to papers by Gross and Scheffler (1998), Gross (1998) and Kroes (1999).\ [**5.9.2.1  The potential-energy surface of H$_2$ at transition-metal surfaces**]{}\ A good knowledge of the high-dimensional PES of the molecule-surface system is mandatory for a detailed understanding, as the PES rules the scattering and the dissociation. The high-dimensionality implies that the dynamics of the problem will be complex and therefore typically it will be impossible to analyze the PES by simply looking at it. In fact, looking at a PES with dimensionality equal to or higher than six is only possible in terms of cuts along planes in configuration space. Figure \[fig:H2-Pd-elbows\] shows (0,10.0) three examples of what is usually called an “elbow plot” because it often looks like an elbow. The planes in configuration space are identified by the insets in the figure. Along these planes the height of the molecule $Z$ and the H–H distance $d_{\rm H-H}$ is varied, and lines of constant potential energy are displayed. Obviously, the cut through the PES shown in Fig. \[fig:H2-Pd-elbows\]a looks very different to those in Figs. \[fig:H2-Pd-elbows\]b, c. If the H$_2$-surface distance (i.e., $Z$) is large the contour lines in Fig. \[fig:H2-Pd-elbows\] reflect the energetics of a free H$_2$ molecule, which has the equilibrium separation of 0.75 Å. We can also read off the nearly harmonic potential underlying the H-H vibration, which implies an H$_2$ zero-point energy of 0.26 eV, and a vibrational excitation energy of 0.52 eV. Although the knowledge of one elbow plot is already much better than just knowing the transition state, as the latter is just one point in the whole configurations space, restricting the world to only one elbow ignores the fact that two H-atoms have six degrees of freedom, not just two. It is obvious from Fig. \[fig:H2-Pd-elbows\] that neglecting the high dimensionality, i.e., assuming that the elbows for different choices of $(X, Y, \theta_{{\rm H}_2}, \phi_{{\rm H}_2})$ are similar, is by no means justified. We will now discuss one aspect of the PES which explains the general trend of the reactivity of different transition-metal surfaces. However we already like to add the warning that an appropriate (and reliable) analysis of chemical reactivity has to include a good [*statistical treatment*]{} of the [*dynamics*]{} of the atoms, which will be discussed in the next section. Figure \[fig:H2-Pd-elbows\]a shows that on Pd(001) there is at least one pathway toward dissociative adsorption which is not hindered by an energy barrier. In fact, if we inspect the PES for Rh (the left neighbor of Pd in the periodic table) \[see Eichler et al., (1999a)\] one sees that here the PES offers even more pathways without barriers: For molecules with their axis parallel to the surface almost all dissociation paths are non-activated. On the other hand, for Ag (the right neighbor of Pd) we find that all pathways toward dissociative adsorption are hindered by a significant energy barrier (Eichler et al., 1999a). This result simply reflects the higher chemical activity of the true transition metals compared to the noble metal silver. (0,8.5) Figure \[fig:H2-Ag-elbows\] shows the PES of H$_2$ at Ag(001). The important aspect in comparing the (a) panels of Figs. \[fig:H2-Pd-elbows\] and \[fig:H2-Ag-elbows\] is not just that there is an energy barrier for the silver substrate, but [*where this barrier is located*]{} in configuration space: The lowest barrier (see Fig. \[fig:H2-Ag-elbows\]) is found very close to the surface and at a H–H distance which is significantly stretched (by about 100%) compared to the free molecule. We also note in passing that the adsorption state of H on Ag(001) is only metastable, i.e., at a higher energy than the free H$_2$ molecule. Keeping the problems with “the” transition state in mind (we will come back to it in the next section) we now analyze its properties. The fact that the transition state in Fig. \[fig:H2-Ag-elbows\] is found at small $Z$ and significantly stretched $d_{\rm H-H}$ values implies that the H–H bond is nearly broken when the molecule has reached the top of the barrier. In fact, the detailed analysis of the wave functions of the H$_2$–surface system at the geometry of the barrier shows that the differences between different metals should be described in a covalent (or tight-binding) picture. Earlier attempts, which applied a description in terms of Pauli repulsion, and/or frontier orbitals of the unperturbed constituents do not account properly for the character and strength of the interaction: At the barrier we find that the interaction between the molecule and the substrate is already significant and the electronic states ruling the energetics are very different from those of the clean surface. (0,9.0) Figure \[fig:H2-tight-binding\] summarizes the view developed by Hammer et al. (1994, 1995) in their analysis of H$_2$ at Cu(111) and H$_2$ at NiAl(110) \[see also the earlier study by Hjelmberg et al. (1979) for H$_2$ at jellium\]. Because of obvious reasons, this figure looks similar to that in Section \[adsorbate-substrate-interaction\], but now the interacting particle is an H$_2$ molecule and the geometry is that of the transition state, and not of an adsorbate equilibrium configuration. At the transition state the interaction of the molecule with the surface has already produced a clear splitting into states which are bonding between the molecule and the substrate and which are antibonding. Assuming a substrate from the middle of the transition-metal series (e.g., Ru or Rh) implies that the low energy resonances, which are $\sigma_{\rm g}$ and $\sigma_{\rm u}$ derived, are filled with electrons. These states are bonding with respect to the molecule-substrate interaction and thus their filling implies an attraction of the molecule to the surface. But the filling of the $\sigma_{\rm u}$ resonance also implies a weakening of the H–H bond. Thus, when the substrate Fermi-level is in the middle of the $d$-band, we understand that molecules are strongly attracted to the surface and at the same time the molecular bond is broken. On the other hand, when the substrate Fermi level is well above the $d$-band, as for a noble metal, also the states which are antibonding with respect to the molecule-surface interaction get filled (the high energy DOS in panels (c) and (d) of Fig. \[fig:H2-tight-binding\]). This implies that the net interaction between the molecule and the substrate is repulsive. Thus, an energy barrier is built up which hinders the dissociation.\ [**5.9.2.2  The dynamics of H$_2$ dissociation at transition-metal surfaces**]{}\ We have seen that the dissociative adsorption of H$_2$ at Pd(001) can proceed without an energy barrier. However, this holds only for few pathways; in particular, it is necessary that the molecule reaches the surface with an orientation parallel to the surface in order to be able to dissociate. Molecules arriving with an orientation of their axis perpendicular to the surface will be reflected. In order to understand in more detail what goes on and to determine the probability of dissociation it is necessary to calculate the dynamics along the high-dimensional PES. This has been done by Gross et al. (Gross et al., 1995, 1998). They treat also the hydrogen nuclei as quantum particles, and a comparison with a classical treatment of the dynamics showed in great detail, which quantum effects are important (mainly zero-point vibrations, and only little tunneling). (0,6.0) Figure \[fig:sticking\] displays the sticking probability $S$ for two different substrates. The sticking probability is the probability that an incoming H$_2$ molecule dissociates and that the atoms then adsorb at the surface. One could also say that $(1-S)$ is the probability that an incoming H$_2$ molecule gets reflected back into the vacuum. The calculated sticking curve exhibits many oscillations. This is not noise, but reflects the quantum nature of the dissociative adsorption and the scattering event: The reflected H$_2$ beams, which differ by reciprocal lattice vectors of the surface, and the rotationally and vibrationally excited beams are subject to quantum interferences. Thus, the oscillations have a quantum mechanical origin, and when the hydrogen nuclei were treated classically, they were absent. Similar oscillations are well known for other quantum mechanical scattering studies, such as the scattering of electrons (LEED) or that of He. We trust that at some time they will be also observed experimentally for the H$_2$ scattering at metal surfaces, but clearly such experiments are very demanding. More details on the nature of the quantum interference effects of H$_2$/Pd scattering can be found in the paper by Gross and Scheffler (1998). When the theoretical results were broadened with the energy resolution of typical experiments, most of the oscillations were averaged out. One unexpected and surprising result of Fig. \[fig:sticking\] is that for low kinetic energies ($E_i < 0.05$ eV) the sticking probabilities for the Pd and Rh substrates are very similar (we recall that the PES of Rh offers many pathways with vanishing energy barriers toward dissociation but Pd only a few). Despite their differences in electronic structure (Pd is more noble than Rh) and the clear differences in the PESs, both substrates give an $S$ value as high as 75% for low $E_i$. In fact, this low energy range corresponds to the typical thermal kinetic energies. While for Rh the sticking probability always remains high, it decreases for Pd to about 25%. The sticking probability for H$_2$ at Pd(001) starts at a high value and then it goes down; this is found in the theory as well as in the experimental data. Earlier, it had been suggested that this behavior is due to the existence of a precursor adsorption, where H$_2$ is trapped close to the surface from where it can undergo many attempts to dissociate. Whereas this concept is valid for some systems, the theoretical PES does not exhibit such a precursor state and thus, the explanation for this behavior of $S(E_i)$ must have a different origin. The analysis of the H$_2$ dynamics revealed that the effect is best described by the word “steering”, which means the same as on the road. If one is going slowly, one will make it well along a curvy street. However, when one is going too fast one is pulled out of the curve. Thus, molecules which are approaching slowly will be able to follow a pathway along curvy valleys and find (along the high-dimensional PES) the way toward the point where they can dissociate without an energy barrier. They will dissociate even if their initial orientation is unfavorable (e.g., perpendicular to the surface) because they are steered toward a more favorable transition state geometry. On the other hand, fast molecules will not be able to make it around all the curves. They will bump against an energy barrier and be reflected back into the gas phase. Thus, the behavior we see for the Pd substrate which increases the sticking to nearly 75% at low energies of the H$_2$ beam, is truly a dynamical effect. In a static picture the high reactivity of Pd cannot be understood. The efficiency of steering depends on the speed of the incoming molecule and on the shape of the PES. Therefore, to evaluate the sticking probability, which we consider a good measure of the surface reactivity, it is important to consider all degrees of freedom and the dynamics of the nuclei. Obviously, as much as “steering” is important to understand the high reactivity of Pd at low $E_i$, for other systems, which on the grounds of the electronic structure alone may be expected to exhibit a high reactivity, an “anti-steering” may occur, which drives approaching molecules not toward the best transition state but against an energy barrier. We add some words on the influence of rotations of the incoming molecule. For rapidly rotating molecules steering is suppressed because the molecules will quickly rotate out of favorable orientations for dissociation (so-called rotational hindering) (Darling and Holloway, 1994; Gross et al., 1996a, b, c). But, there is also a steric effect (Gross et al., 1995): molecules rotating in the so-called helicopter fashion (i.e., parallel to the surface) with $m=j$ dissociate more easily than molecules rotating in the so-called cartwheel fashion (i.e., perpendicular to the surface) with $m=0$, because the helicoptering molecules have their axis already oriented preferentially parallel to the surface which is the most favorable orientation for dissociation. This orientation effect can be so strong that it even over-compensates the rotational hindering. Furthermore we note that, for example, the hindering effect of the rotational motion may be counterbalanced by the transfer of rotational energy to translational energy (see Eichler et al., 1999b, and Gross and Scheffler, 1996b, for more details). The catalytic oxidation of CO {#sec:catalytic} ----------------------------- (0,5.0) Just as the dissociative adsorption of H$_{2}$ discussed above in Section \[sec:H2\] may be regarded as the model system for studying the dynamics of the dissociative adsorption and associative desorption of small molecules, the CO oxidation reaction may be regarded as the most simple prototype model system of a surface heterogeneous catalytic reaction – a process involving molecular adsorption (chemisorption of CO is non-dissociative under the conditions of catalytic oxidation) and dissociative (atomic) adsorption, surface diffusion, surface reaction, and desorption of products. The oxidation of carbon monoxide has been extensively studied which is largely due to its technological importance (e.g., in car exhaust catalytic converters where the active components are transition metals such as Pt, Pd, and Rh) but is also related to its (relative)“simplicity” (Engel and Ertl, 1979, 1982; Bielánski and Haber, 1991; Peden, 1992, and references therein). Therefore there exists a large data-base concerning its macroscopic behavior and in these terms it is reasonably well understood. On a [*microscopic*]{} level, however, an understanding is still lacking. Steps in this direction have recently been made via first-principles calculations (Stampfl and Scheffler, 1997, 1999; Alavi et al., 1998; Eichler and Hafner, 1999). (0,6.5) The results of one of these studies is briefly discussed below. We first like to mention that the investigation of this type of surface process is more complex than that described above for H$_{2}$ dissociation; firstly significant substrate relaxations may be induced by CO and O adsorption, and the surface should not be treated as rigid. Secondly, the nature of the problem is different; here a “real” surface chemical reaction takes place where a product is formed that desorbs from the surface. So not only the coordinates of one adparticle need to be considered, but also those of the reacting partner, i.e., the problem has an even higher dimensionality. Catalytic oxidation of CO at the ruthenium transition metal surface has been reported to exhibit unusual behavior compared to other transition metals when studied at high pressures (Peden, 1992, and references therein). In particular, the reaction rate over Ru for oxidizing conditions (i.e., at CO/O$_{2}$ pressure ratios $ < 1$) is the highest of the transition metals considered. In contrast, under ultra-high vacuum conditions (UHV), the rate is by far the lowest. In addition, the kinetic data for the reaction over Ru (e.g., the temperature and pressure dependencies of the rate) deviate compared to the other metals and highest reaction rates occur for high oxygen concentrations at the surface. Interestingly, under these conditions, it was speculated that the reaction mechanism may proceed via an Eley-Rideal interaction (Peden et al., 1986, 1991), that is, a scattering reaction where a particle from the gas-phase reacts with an adsorbed species without adsorbing on the surface first. The usual mechanism by which reactions at surfaces take place is via the Langmuir-Hinshelwood mechanism where both species are adsorbed on the surface prior to reaction. (0,6.0) Under standard UHV conditions using O$_{2}$, the saturation coverage has been reported to be approximately half a monolayer. Recent studies employing NO$_{2}$ or very high exposures of O$_{2}$ have shown, however, that Ru(0001) can support higher coverages; namely ordered structures $(2 \times 2)$-3O (Kostov et al., 1997; Kim et al., 1998; Gsell et al., 1998) and $(1 \times 1)$-O (Stampfl et al., 1996), as had initially been predicted by DFT-GGA calculations (Stampfl and Scheffler, 1996), and that subsurface adsorption occurs after completion of the monolayer structure at elevated temperatures ($\approx$600 K) (Stampfl et al., 1996; Mitchell and Weinberg, 1996; Böttcher and Niehus, 1999). Subsurface oxygen can apparently occur as well as the formation of surface oxides[^7], and surfaces with different domains of high oxygen concentration but different stoichiometry are assumed to actuate the high catalytic reactivity of Ru. Thus, the (apparent) oxygen saturation coverage noted above for low (or room) temperature UHV conditions and typical exposure by O$_2$ is solely due to kinetic hindering for O$_{2}$ dissociation. The adsorption energy of O on Ru(0001) decreases notably with increasing coverage; in particular, for concentrations $\geq$ 1 ML the bond strength is atypically weak compared to the lower coverage structures. This can be seen from Fig. \[oadsorp\]. A good catalyst should actuate dissociation of O$_{2}$ but at the same time should not bind the dissociated entities too strongly, as then they have a good capability to diffuse and react. Too strongly bound constituents would have little reason to react at all. It is expected therefore that these weaker adsorption energies for $\Theta \geq 1$ will lead to an enhanced reaction rate of CO$_2$ formation. Indeed, recent experimental studies of CO oxidation over Ru(0001) surfaces loaded with such high oxygen concentrations, have found even higher rates (Böttcher et al., 1997, 1999). (0,10.5) When a complete $(1 \times 1)$-O structure is present, the calculations show that CO cannot adsorb. However, a scattering reaction of gas-phase CO with adsorbed oxygen (i.e., an Eley-Rideal mechanism) is possible, and the minimum energy barrier was found to be about 1.1 eV with a corresponding bent transition state geometry (Stampfl and Scheffler, 1997). Figure \[pes-co2\] shows an appropriate cut through the high-dimensional PES. The minimum barrier corresponds to a tilted CO molecular axis of about 131$^{\circ}$ as depicted in the inset of Fig. \[en-dig-er\] which shows the corresponding energy diagram. It can be seen that due to the surface reaction there would be a significant energy gain of about 1.95 eV so that the produced CO$_{2}$ molecules would be highly energetic. Assuming this energy barrier would be the rate-limiting step of CO$_{2}$ formation via this mechanism, then an estimate of the reaction rate can be made using an Arrhenius-like equation. The prefactor being taken as the number of CO molecules hitting the surface per site per second at a given temperature and pressure. The rate obtained in this way was found to be significantly lower than that measured experimentally (Peden and Goodman, 1986) indicating that this mechanism alone cannot explain the enhanced CO$_{2}$ turnover frequency. (0,5.0) To investigate other possible reaction channels, it is conceivable that there may be vacancies in the $(1 \times 1)$-O adlayer (see Stampfl and Scheffler, 1997), e.g., created by the above mentioned scattering reaction. CO molecules may then adsorb at these vacant sites and react via a Langmuir-Hinshelwood mechanism (see Fig. \[lh-chden\]). In Section \[sec:co-adsorption\] it was seen that there is an energy barrier for CO to adsorb in such a vacancy; however, at the high pressures and elevated temperatures employed in catalytic reactor experiments, it is expected that a barrier of such size can be readily overcome. In view of the weaker CO-metal bond strength compared to that of the O-metal bond strength at this coverage, i.e., 0.85 eV compared to 2.12 eV[^8] (with respect to gas phase 1/2 O$_{2}$), it may be expected that the energetically favorable reaction pathway is via movement of CO toward the O atom. Indeed, among the various pathways investigated, this turns out to be the case where the determined energy barrier is about 1.5 eV. Some selected geometries along the reaction path are shown in Fig. \[react-path-geoms\]. We refer to Stampfl and Scheffler (1999) for further details. (0,7.0) It can be seen that at the transition state (also shown in the inset of Fig. \[en-dig-lh\]), the C-O(a) bond[^9] is almost parallel to the surface. The CO axis is bent away from O(a) yielding a bent CO-O(a) complex with a bond angle of 125$^{\circ}$; similar to that found for the Eley-Rideal mechanism. The C-O(a) bond length is 1.59 Å (about 29% stretched compared to that in CO$_{2}$) and the CO bond length is 1.18 Å. At the transition state CO and O begin to lift-off the surface as they break their metal bonds in favor of developing a C-O(a) bond. This behavior can also be seen from the corresponding valence electron density distribution shown in Fig. \[lh-chden\]. The corresponding energy diagram is given in Fig. \[en-dig-lh\]. In the work of Alavi et al. (1998) and Eichler and Hafner (1999) bent transition state geometries were also identified for the CO oxidation reaction over the Pt(111) surface for the case of low ($\Theta = 0.25$) oxygen coverages. Alavi et al. (1998) attributed the main contribution to the activation barrier to the weakening of the O-metal bond strength, which supports the general understanding in this respect. We end this section by noting that despite its high reactivity for CO oxidation, elemental Ru will not be used in automotive catalysts because a volatile Ru oxide exists which is highly poisonous. Nevertheless, trying to understand why Ru is so much more effective than other transition metals may help to design materials with similar properties as Ru. Furthermore, as noted above, there is an interesting and likely conceptionally important aspect of Ru studies: Ru can exist in many oxidation states and the result (see Fig. \[oadsorp\]) that its surface region can be loaded with a high concentration of oxygen, poses the question whether it is appropriate to call the reactive surface an adsorbate system, or if it more appropriate to call this a surface oxide (a RuO$_2$-like system in the present case). The highly reactive surface of O/Ru(0001) apparently is one with oxygen concentration much higher than $\Theta = 1$ and a coexistence of several domains of different stoichiometry. Summary outline of main points ------------------------------ A deeper understanding of chemisorption, surface chemical reactions, and heterogeneous catalysis is currently one of the main aims of surface science (other important topics are, e.g., crystal growth, surface magnetism). In the context of this complex problem, the nature of the chemisorption bond, the energetics of adsorption, and adsorbate geometries and bond-lengths receive considerable attention. While several simple rules have been found to be useful as in the area of molecular chemistry, others should be applied with much caution. Below we summarize some of the findings discussed in this chapter in terms of a brief list: - Density-functional theory has evolved into an important tool for analyzing surface geometries. For example, the four-layer surface alloy of Na on Al(111) was first predicted and analyzed by DFT calculations (Stampfl and Scheffler, 1994c) and subsequently confirmed by a LEED intensity analysis (Burchhardt et al., 1995); in view of the high number of structural parameters, it was indeed important to start with the DFT analysis. Similarly, the existence of a $(1 \times 1)$ ordered adlayer of O on Ru(0001) and of the $(2 \times 2)$-3O adlayer was first predicted by DFT calculations (Stampfl and Scheffler, 1996) with detailed specification of the geometry. And more examples exist. - In addition to providing accurate atomic geometries, DFT calculations also (and in particular) offer the potential of analysis of the underlying mechanisms which determine if and how a certain geometry can be attained and what the nature of the chemical bond is. - Whereas the geometry is (typically) well described by DFT-LDA and DFT-GGA calculations, the resulting [*energies*]{} must be taken with some more caution. Adsorption energies are typically not more accurate than about 0.2 eV per adatom. And we do not expect that new exchange-correlation functionals will improve this to better than 0.1 eV per adatom in the near future (the letter “A”, as in LDA and GGA, will remain part of exchange-correlation functionals employed in actual DFT calculations). However, energy [*differences*]{} of chemically similar bonding situations (in particular energies of small distortions and phonons) are described with very well (possibly even meV) accuracy. - The influence of thermally induced vibrations of the substrate surface and adatoms is often ignored, even today. Still, it is quite clear that some geometries (e.g., the unreconstructed surface of Au(111), or the in-registry adsorption of Xe on fcc(111) surfaces) are stabilized by vibrational energy and entropy. We also mention that surface stress can be noticeably affected by the surface thermal expansion. Thus, sometimes it may be necessary to include the influence of thermal expansion in theoretical studies (see, e.g., Cho and Scheffler, 1997; Xie et al., 1999). - Unlike a chemical reaction in molecular chemistry, adsorption on a surface involves very unequal partners. For example, the substrate surface gives rise to broadening of adsorbate levels, has an infinite number of electrons, and fixes the electron chemical potential. - Despite item 5), electronegativity differences between adsorbate and substrate species appear to give a qualitatively correct description of the nature of the chemical bond. - Item 6) may hold less true for higher coverages when the adsorbate-adsorbate interaction becomes noticeable. - The correlation between local-coordination and bond-strength, as noted for molecules by Pauling, appears to be (typically) fulfilled for adsorbates. The energy per atom scales roughly proportional to the square root of the coordination. For covalent systems, some geometries may have a more favorable energy than that of this simple proportionality, e.g., when coordination numbers and/or bond-angles conform with the number of available valence electrons.\ In this context we also re-emphasize that unexpected behavior of atoms can occur at surfaces (i.e., the discussed substitutional adsorption and alloy formation). At this time knowledge about the energetics of the underlying processes, as, e.g., the formation of surface vacancies and the adsorption of adatoms at step sites, is not sufficiently well developed that a prediction of the adsorbate site (in particular at low coverage) is possible without performing a DFT calculation. - The validity of a correlation between local coordination and bond-length (in principle a consequence of the local-coordination–bond-strength correlation), appears to be rarely fulfilled for adsorbates. This is due to the fact that geometries are quite constrained, as the atoms of the substrate surface are well bonded to the rest of the substrate, and thus the adsorbate bonds may not succeed to attain their optimum bond angle. Thus, surface bonds my be subject to a “frustrated” hybridization of the adatom orbitals or the substrate-surface orbitals. - A metal surface attempts to reach charge neutrality on a rather short length scale. In fact, typically a perturbation is even slightly over-screened on the length scale of the nearest-neighbor distance and then the induced electron density is slowly decaying in an oscillatory manner. This highly localized screening is achieved by locally shifting valence electron density of states to higher or lower energies and thus changing the occupancy. - Reactivity concepts of molecular chemistry, which explore the electron density of the [*unperturbed*]{} partners, are only of limited value for dissociation at surfaces. It appears that the lowest energy transition states are (often) so close to the surface where the interaction is already so strong, that “new states” are formed, which differ considerably from the clean surface states and the free adatom orbitals. The filling of these “new states” (determined by the substrate Fermi level) rules (at least partially) the reactivity. - In addition to item 11) it is important to include the [*dynamics*]{} of the approaching molecules: It is not enough to know just the energy of the lowest-energy transition state, but it is important to know whether or not (or with what probability) the transition state will be found by an approaching molecule. This stresses the importance of the high dimensionality of the potential-energy surface on which an approaching molecule travels and of [*the statistics*]{}. - Some often used concepts still need qualification. For example, it is not clear how a term like that of a “surface oxide” or “surface hydride” describes a situation (or phase) which is different from an “adsorbate”. Clearly, this problem only arises for high-coverage adsorption ($\Theta > 1$). - We trust that the future will bring more studies which are performed under finite pressure of a well defined atmosphere (getting out of the vacuum) and in which the temperature is changed systematically (warming up and cooling down). This will show which of the previous studies were concerned with a thermal equilibrium geometry and which dealt with a (possibly very special) metastable state.\ \ We are grateful for the collaborations with many colleagues and friends, the results of whom have been discussed in this chapter. In particular we mention J[ö]{}rg Bormet, Axel Gross, and J[ö]{}rg Neugbauer, but this list could easily be extended. We thank Kristen Fichthorn, Klaus Hermann, and Karsten Reuter for helpful comments on the manuscript.\ [**References**]{}\ Adams, D.L., 1996, Appl. Phys. A [**62**]{}, 123.\ Alavi, A., P. Hu, T. Deutsch, P.L. Silvestrelli and J. Hutter, 1998, Phys. Rev. Lett. [**80**]{}, 3650.\ Aminpirooz, S., A. Schmalz, N. Pangher, J. Haase, M.M. Nielsen, D.R. Batchelor, E. B[ø]{}gh and D.L. Adams, 1992, Phys. Rev. B [**46**]{}, 15594.\ Andersen, J.N., M. Qvarford, R. Nyholm, J.F. van Acker and E. Lundgren, 1992, Phys. Rev. Lett. [**68**]{}, 94.\ Andersen, J.N., D. Hennig, E. Lundgren, M. Methfessel, R. Nyholm and M. Scheffler, 1994, Phys. Rev. B [**50**]{}, 17525.\ Antoniewicz, P.R., 1978, Phys. Status Solidi (B) [**86**]{}, 645.\ Bader, R.F.W., 1990, Atoms in Molecules. A Quantum Theory, International Series of Monographs on Chemistry, Vol. 22, Oxford University Press, Oxford.\ Bader, R.F.W., 1994, Phys. Rev. B [**49**]{}, 13348.\ Bagus, P.S., K. Hermann, W. Müller, and C.J. Nelin, 1986, Phys. Rev. Lett. [**57**]{}, 1496.\ Barnes, C.J., 1994, in: The Chemical Physics of Solid Surfaces, Vol. 7, Phase Transitions and Adsorbate Restructuring at Metal Surfaces, eds. D.A. King and D.P. Woodruff. Elsevier, Amsterdam, p. 501.\ Behm, R.J., 1989, in: Physics and Chemistry of Alkali Metal Adsorption, eds. H.P. Bonzel, A.M. Bradshaw and G. Ertl. Elsevier, Amsterdam, p. 111.\ Berndt, W., D. Weick, C. Stampfl, A.M. Bradshaw and M. Scheffler, 1995, Surf. Sci. [**330**]{}, 182.\ Bielański, A. and J. Haber, 1991, Oxygen in Catalysis. Dekker, New York, p. 472.\ Blyholder, G., 1964, J. Phys. Chem. [**68**]{}, 2772.\ Blyholder, G., 1975, J. Vac. Sci. Technol. [**11**]{}, 865.\ Bockstedte, M., A. Kley, J. Neugebauer and M. Scheffler, 1997, Comput. Phys. Commun. [**107**]{}, 187.\ Bormet, J., J. Neugebauer and M. Scheffler, 1994a, Phys. Rev. B [**49**]{}, 17242.\ Bormet, J., B. Wenzien, J. Neugebauer and M. Scheffler, 1994b, Comput. Phys. Commun. [**79**]{}, 124.\ Born, M. and R. Oppenheimer, 1927, Ann. Phys. [**84**]{}, 457.\ Born, M. and K. Huang, 1954, in: Dynamical Theory of Crystal Lattices, eds. N.F. Mott and E. C. Bullard. Clarendon, Oxford, p. 420.\ Böttcher, A. and Niehus, 1999, J. Chem. Phys. [**110**]{}, 3186.\ Böttcher, A., Niehus, S. Schwegmann, H. Over and G. Ertl, 1997, J. Phys. Chem. [**101**]{}, 11185.\ Böttcher, A., M. Rogozia, H. Niehus, H. Over and G. Ertl, 1999, J. Chem. Phys., accepted.\ Bradshaw, A.M. and M. Scheffler, 1979, J. Vac. Sci. Technol. [**16**]{}, 447.\ Brivio, G.P. and M.I. Trioni, 1999, Rev. Mod. Phys. [**71**]{}, 231.\ Bruch, L.W., M.W. Cole and E. Zaremba, 1997, Physical Adsorption: Forces and Phenomena, Clarendon, Oxford.\ Burchhardt, J., M.M. Nielsen, D.L. Adams, E. Lundgren, J.N. Andersen, C. Stampfl, M. Scheffler, A. Schmalz, S. Aminpirooz and J. Haase, 1995, Phys. Rev. Lett. [**74**]{}, 1617.\ Callaway, J., 1964, J. Math. Phys. [**5**]{}, 783.\ Callaway, J., 1967, Phys. Rev. [**154**]{}, 515.\ Campuzano, J.C., 1990, in: The Chemical Physics of Solid Surfaces and Heterogeneous Catalysis, Vol. 3a: Chemisorption Systems, eds. D.A. King and D.P. Woodruff. Elsevier, Amsterdam, p. 389.\ Car, R. and M. Parrinello, 1985, Phys. Rev. Lett. [**55**]{}, 2471.\ Cho, J.-H. and M. Scheffler, 1997, Phys. Rev. Lett. [**78**]{}, 1299.\ Christensen, O.B. and K.W. Jacobsen, 1992, Phys. Rev. B [**45**]{}, 6893.\ Christensen, S.V., J. Nerlov, K. Nielsen, J. Burchhardt, M.M. Nielsen and D.L. Adams, 1996, Phys. Rev. Lett. [**76**]{}, 1892.\ Darling, G.R. and S. Holloway, 1994, J. Chem. Phys. [**101**]{}, 3268.\ Dreizler, R.M. and E.K.U. Gross, 1990, Density Functional Theory. Springer, Berlin.\ Echenique, P.M. and J.B. Pendry, 1975, J. Phys. C [**8**]{}, 2936.\ Eichler, A., J. Hafner, A. Gross and M. Scheffler, 1999a, Phys. Rev. B [**59**]{}, 13297.\ Eichler, A., J. Hafner, A. Gross and M. Scheffler, 1999b, Chem. Phys. Lett. [**311**]{}, 1.\ Eichler A. and J. Hafner, 1999, Phys. Rev. B [**59**]{}, 5960.\ Engel, T. and G. Ertl, 1979, J. Chem. Phys. [**69**]{}, 1267; Adv. Catal. [**28**]{}, 1.\ Engel, T. and G. Ertl, 1982, in: The Chemical Physics of Solid Surfaces and Heterogeneous Catalysis, Vol. 4, Fundamental Studies of Heterogenous Catalysis, eds. D.A. King and D.P. Woodruff. Elsevier, Amsterdam, p. 73.\ Fasel, R., P. Aebi, R.G. Agostino, L. Schlapbach and J. Osterwalder, 1996, Phys. Rev. B [**54**]{}, 5893.\ Feibelman, P.J., 1990, Phys. Rev. Lett. [**65**]{}, 729.\ Finnis, M.W., R. Kaschner, C. Kruse, J. Furthm[ü]{}ller and M. Scheffler, 1995, J. Phys.: Condens. Matter [**7**]{}, 2001.\ Fiorentini, V., M. Methfessel and M. Scheffler, 1993, Phys. Rev. Lett. [**71**]{}, 1051; 1998, Phys. Rev. Lett. [**81**]{}, 2184.\ Ganduglia-Pirovano, M.V., J. Kudrnovský and M. Scheffler, 1997, Phys. Rev. Lett. [**78**]{}, 1807.\ Grimley, T.B., 1975, Prog. Surf. Membr. Sci. [**9**]{}, 71.\ Gross, A., S. Wilke and M. Scheffler, 1995, Phys. Rev. Lett. [**75**]{}, 2718.\ Gross, A. and M. Scheffler, 1996a, Chem. Phys. Lett. [**256**]{}, 417.\ Gross, A. and M. Scheffler, 1996b, Prog. Surf. Sci. [**53**]{}, 187.\ Gross, A., S. Wilke and M. Scheffler, 1996c, Surf. Sci. [**357/358**]{}, 614.\ Gross, A., M. Bockstedte and M. Scheffler, 1997, Phys. Rev. Lett. [**79**]{}, 701.\ Gross, A. and M. Scheffler, 1998, Phys. Rev. B [**57**]{}, 2493.\ Gross, A., 1998, Surf. Sci. Rep. [**32**]{}, 291.\ Gsell, M., M. Stichler, P. Jacob and D. Menzel, 1998, Israel J. Chem. [**38**]{}, 339.\ Gurney, R.W., 1935, Phys. Rev. [**47**]{}, 479.\ Hammer, B., M. Scheffler, K.W. Jacobsen and J.K. N[ø]{}rskov, 1994, Phys. Rev. Lett. [**73**]{}, 1400.\ Hammer, B. and M. Scheffler, 1995, Phys. Rev. Lett. [**74**]{}, 3487.\ Hammer, B. and J.K. N[ø]{}rskov, 1997, in: Chemisorption and Reactivity on Supported Clusters and Thin Films, eds. R.M. Lambert and G. Pacchioni. Kluwer, Dordrecht, p. 285.\ Heine, V. and D. Marks, 1986, Surf. Sci. [**115**]{}, 65.\ Hennig, D., M.V. Ganduglia-Pirovano and M. Scheffler, 1996, Phys. Rev. B [**53**]{}, 10344.\ Hermann, K. and P.S. Bagus, 1977, Phys. Rev. B [**16**]{}, 4195.\ Hermann, K., P.S. Bagus, and C.J. Nelin, 1987, Phys. Rev. B [**35**]{}, 9467.\ Hjelmberg, H., B.I. Lundqvist and J.K. N[ø]{}rskov, 1979, Phys. Scr. [**20**]{}, 192.\ Hoffmann, P., C. von Muschwitz, K. Horn, K. Jacobi, A.M. Bradshaw, K. Kambe and M. Scheffler, 1979, Surf. Sci. [**89**]{}, 327.\ Hoffmann, R., 1988, Rev. Mod. Phys. [**60**]{}, 601.\ Horn, K., M. Scheffler and A.M. Bradshaw, 1978, Phys. Rev. Lett. [**41**]{}, 822.\ Hu, P., D.A. King, M.-H. Lee and M.C. Payne, 1995, Chem. Phys. Lett. [**246**]{}, 73.\ Jacobi, K., C. von Muschwitz and K. Kambe, 1980, Surf. Sci. [**93**]{}, 310.\ Janak, J.F., 1978, Phys. Rev. B [**18**]{}, 7165.\ Jennings, P.J., 1979, Surf. Sci. [**88**]{}, L25.\ Kambe, K. and M. Scheffler, 1979, Surf. Sci. [**89**]{}, 262.\ Kim, Y.D., S. Wendt, S. Schwegmann, H. Over and G. Ertl, 1998, Surf. Sci. [**418**]{}, 267.\ Kleinman, L., 1997, Phys. Rev. B [**56**]{}, 16029.\ Kohn, W. and K.H. Lau, 1976, Solid State Commun. [**18**]{}, 553.\ Kostov, K.L., H. Rauscher and D. Menzel, 1992, Surf. Sci. [**278**]{}, 62.\ Kostov, K.L., M. Gsell, P. Jakob, T. Moritz, W. Widdra and D. Menzel, 1997, Surf. Sci. [**394**]{}, L138.\ Kresse, G. and J. Furthm[ü]{}ller, 1996, Phys. Rev. B [**54**]{}, 11169.\ Kroes, G.J., E.J. Baerends and R.C. Mowrey, 1997, Phys. Rev. Lett. [**78**]{}, 3583; 1998, Phys. Rev. Lett. [**81**]{}, 4781.\ Kroes, G.J., 1999, Prog. Surf. Sci. [**60**]{}, 1.\ Lang, N.D. and W. Kohn, 1971, Phys. Rev. B [**3**]{}, 1215.\ Lang, N.D., 1971, Phys. Rev. B [**4**]{}, 4234.\ Lang, N.D., 1973, Solid State Phys. [**28**]{}, 225.\ Lang, N.D. and A.R. Williams, 1977, Phys. Rev. B [**16**]{}, 2408.\ Lang, N.D. and A.R. Williams, 1978, Phys. Rev. B [**18**]{}, 616.\ Langmuir, I., 1932, J. Am. Chem. Soc. [**54**]{}, 2798.\ Liebsch, A., 1978, Phys. Rev. B [**17**]{}, 1653.\ Lindroos, M., H. Pfnür, G. Held and D. Menzel, 1989, Surf. Sci. [**222**]{}, 451.\ McRae, E.G., 1971, Surf. Sci. [**25**]{}, 491.\ Methfessel, M., D. Hennig and M. Scheffler, 1992a, Phys. Rev. B [**46**]{}, 4816.\ Methfessel, M., D. Hennig and M. Scheffler, 1992b, Appl. Phys. A [**55**]{}, 442.\ Methfessel, M., D. Hennig and M. Scheffler, 1995, Surf. Rev. Lett. [**2**]{}, 197.\ Miller, A.R., 1946, Proc. Cambridge Philos. Soc. [**42**]{}, 492.\ Mitchell, W.J. and W.H. Weinberg, 1996, J. Chem. Phys. [**104**]{}, 9127.\ Muscat, J.P. and D.M. Newns, 1978, Prog. Surf. Sci. [**9**]{}, 1.\ Muscat, J.P. and D.M. Newns, 1979, Phys. Rev. B [**19**]{}, 1270.\ Narloch, B., G. Held and D. Menzel, 1994, Surf. Sci. [**317**]{}, 131.\ Narloch, B., G. Held and D. Menzel, 1995, Surf. Sci. [**340**]{}, 159.\ Naumovets, A.G., 1994, in: The Chemical Physics of Solid Surfaces, Vol. 7, Phase Transitions and Adsorbate Restructuring at Metal Surfaces, eds. D.A. King and D.P. Woodruff. Elsevier, Amsterdam, p. 163.\ Neugebauer, J. and M. Scheffler, 1992, Phys. Rev. B [**46**]{}, 16067.\ Neugebauer, J. and M. Scheffler, 1993, Phys. Rev. Lett. [**71**]{}, 577.\ Nilsson, A., N. Wassdahl, M. Weinelt, O. Karis, T. Weill, P. Bennich, J. Hasselström, A. Föhlisch, J. Stöhr and M. Samant, 1997, Appl. Phys. A [**65**]{}, 147.\ N[ø]{}rskov, J.K., 1990, Rep. Prog. Phys. [**53**]{}, 1253.\ Nouvertné, F., U. May, M. Bamming, A. Rampe, U. Korte, G. G[ü]{}ntherodt, R. Pentcheva and M. Scheffler, 1999, Phys. Rev. B [**60**]{},14382.\ Oppo, S., V. Fiorentini and M. Scheffler, 1993, Phys. Rev. Lett. [**71**]{}, 2437.\ Over, H., H. Bludau, M. Gierer, and G. Ertl, 1995, Surf. Rev. Lett. [**2**]{}, 409.\ Palmberg, P.W., 1971, Surf. Sci. [**25**]{}, 598.\ Paul, J., 1987, J. Vac. Sci. Technol. A [**5**]{}, 664.\ Pauling, L., 1960, The Nature of the Chemical Bond and the Structure of Molecules and Crystals: An Introduction to Modern Structural Chemistry. Cornell University Press.\ Payne, M.C., M.P. Teter, D.C. Allan, T.A. Arias and J.D. Joannopulos, 1992, Rev. Mod. Phys. [**64**]{}, 1045.\ Payne, M.C., I.J. Robertson, D. Thomson and V. Heine, 1996, Philos. Mag. B [**73**]{}, 191.\ Peden, C.H.F., D.W. Goodman, M.D. Weisel and F.M. Hoffmann, 1991, Surf. Sci. [**253**]{}, 44.\ Peden, C.H.F. and D.W. Goodman, 1986, J. Phys. Chem. [**90**]{}, 1360.\ Peden, C.H.F., 1992, in: Surface Science of Catalysis: In Situ Probes and Reaction Kinetics, eds. D.J. Dwyer and F.M. Hoffmann. Am. Chem. Soc., Washington DC.\ Pedersen, M.O., I.A. Bonicke, E. Laegsgaard, I. Stensgaard, A. Ruban, J.K. N[ø]{}rskov and F. Besenbacher, 1997, Surf. Sci. [**387**]{}, 86.\ Pentcheva, R. and M. Scheffler, 2000, Phys. Rev. B, in print.\ Perdew, J.P. and M. Levy, 1997, Phys. Rev. B [**56**]{}, 16021.\ Petersen, M., P. Ruggerone and M. Scheffler, 1996, Phys. Rev. Lett. [**76**]{}, 995; 2000, Phys. Rev. B, submitted.\ Pleth Nielsen, L., F. Besenbacher, I. Stensgaard, E. L[æ]{}gsgaard, C. Engdahl, P. Stoltze, K.W. Jacobsen and J.K. N[ø]{}rskov, 1993, Phys. Rev. Lett. [**71**]{}, 754.\ Porteus, J.O., 1974, Surf. Sci. [**41**]{}, 515.\ Rader O., W. Gudat, C. Carbone, E. Vescovo, S. Blügel, R. Kläsges, W. Eberhardt, M. Wuttig, J. Redinger and F.J. Himpsel, 1997, Phys. Rev. B [**55**]{}, 5404.\ Rendulic, K.D., G. Anger and A. Winkler, 1989, Surf. Sci. [**208**]{}, 404.\ Robertson, I.J., D.I. Thomson, V. Heine and M.C. Payne, 1994, J. Phys. Condens. Matter [**6**]{}, 9963.\ Scheffler, M., K. Kambe and F. Forstmann, 1978, Solid State Commun. [**25**]{}, 93.\ Scheffler, M., K. Horn, A.M. Bradshaw and K. Kambe, 1979, Surf. Sci. [**80**]{}, 69.\ Scheffler, M. and A.M. Bradshaw, 1983, in: The Chemical Physics of Solid Surfaces and Heterogeneous Catalysis, Vol. 2: Adsorption at Solid Surfaces, eds. D.A. King and D.P. Woodruff. Elsevier, Amsterdam, p. 165.\ Scheffler, M., Ch. Droste, A. Fleszar, F. Máca, G. Wachutka and G. Barzel, 1991, Physica B [**172**]{}, 143.\ Schiffer, A., P. Jakob and D. Menzel, 1997, Surf. Sci. [**389**]{}, 116.\ Schmalz, A., S. Aminpirooz, L. Becker, J. Haase, J. Neugebauer, M. Scheffler, D.R. Batchelor, D.L. Adams and E. B[ø]{}gh, 1991, Phys. Rev. Lett. [**67**]{}, 2163.\ Seyller T., M. Caragiu, R.D. Diehl, P. Kaukasoina and M. Lindroos, 1998, Chem. Phys. Lett. [**291**]{}, 567.\ Skriver, H., 1985, Phys. Rev. B [**31**]{}, 1909.\ Somorjai, G.A. and M.A. Van Hove, 1989, Prog. Surf. Sci. [**30**]{}, 201.\ Spanjaard, D. and M.C. Desjonquères, 1990, in: Interaction of Atoms and Molecules with Surfaces, eds. V. Bortolani, N.H. March and N.P. Tosi. Plenum, New York, London, p. 255.\ SRL, 1995, Surf. Rev. Lett. [**2**]{}, 315. A special issue of [*Surface Review and Letters*]{} devoted to alkali-metal adsorption, containing review articles by several groups.\ Stampfl, C., M. Scheffler, H. Over, J. Burchhardt, M. Nielsen, D.L. Adams and W. Moritz, 1992, Phys. Rev. Lett. [**69**]{}, 1532.\ Stampfl, C., M. Scheffler, H. Over, J. Burchhardt, M. Nielsen, D.L. Adams and W. Moritz, 1994a, Phys. Rev. B [**49**]{}, 4959.\ Stampfl, C., J. Neugebauer and M. Scheffler, 1994b, Surf. Sci. [**307/309**]{}, 8.\ Stampfl, C. and M. Scheffler, 1994c, Surf. Sci. [**319**]{}, L23.\ Stampfl, C. and M. Scheffler, 1995, Surf. Rev. Lett. [**2**]{}, 317.\ Stampfl, C. and M. Scheffler, 1996, Phys. Rev. B [**54**]{}, 2868.\ Stampfl, C., 1996, Surf. Rev. Lett. [**3**]{}, 1567.\ Stampfl, C., S. Schwegmann, H. Over, M. Scheffler and G. Ertl, 1996, Phys. Rev. Lett. [**77**]{}, 3371.\ Stampfl, C. and M. Scheffler, 1997, Phys. Rev. Lett. [**78**]{}, 1500; J. Vac. Sci. Technol. A [**15**]{}, 1635; Surf. Sci. [**377-379**]{}, 808.\ Stampfl, C., K. Kambe, R. Fasel, P. Aebi and M. Scheffler, 1998, Phys. Rev. B [**57**]{}, 15251.\ Stampfl, C. and M. Scheffler, 1998, Israel J. Chem. [**38**]{}, 409.\ Stampfl, C. and M. Scheffler, 1999, Surf. Sci. [**433-435**]{}, 119.\ Taylor, J.B. and I. Langmuir, 1933, Phys. Rev. [**44**]{}, 423.\ Topping, J., 1927, Proc. Roy. Soc. London Ser. A [**114**]{}, 67.\ Wang, X.-G., W. Weiss, Sh.K. Shaikhutdinov, M. Ritter, M. Petersen, F. Wagner, R. Schl[ö]{}gl and M. Scheffler, 1998, Phys. Rev. Lett. [**81**]{}, 1038.\ Wilke, S. and M. Scheffler, 1996, Phys. Rev. B [**53**]{}, 4926.\ Wilke, S., M.H. Cohen and M. Scheffler, 1996, Phys. Rev. Lett. [**77**]{}, 1560.\ Wenzien, B., J. Bormet, J. Neugebauer and M. Scheffler, 1993, Surf. Sci. [**287/288**]{}, 559.\ Wenzien, B., J. Bormet and M. Scheffler, 1995, Comput. Phys. Commun. [**88**]{}, 230.\ Wimmer, E., C.L. Fu and A.J. Freeman, 1985, Phys. Rev. Lett. [**55**]{}, 2618.\ Xie, J., S. de Gironcoli, S. Baroni and M. Scheffler, 1999, Phys. Rev. B [**59**]{}, 970.\ Yang, L., G. Vielsack and M. Scheffler, 1994, unpublished.\ Yu, B.D. and M. Scheffler, 1997, Phys. Rev. B [**56**]{}, R15569.\ [^1]: For a discussion of recent work on noble-gas atom adsorption see Bruch et al., 1997; Brivio and Trioni, 1999; Seyller et al., 1998; Petersen et al., 1996, 2000. [^2]: The coverage is defined in this chapter such that for $\Theta=1$, the number of adatoms is the same as the number of atoms in the clean, unreconstructed surface. \[coverage\] [^3]: Obviously, as $N(\epsilon)$ defines the Kohn-Sham Hamiltonian, it also defines all ground-state and all excited-state properties. [^4]: Our wording is chosen as such that “higher energy” of a bound electronic state means closer to the vacuum level, i.e., with respect to Fig. \[tb-DOS\] or \[Na-Si-Cl-DOS\] the energy is more to the right. Obviously, “lower energy” then refers to an energy more to the left in these figures.\[higher-lower\] [^5]: For open-shell systems, which are discussed here, the DFT-LDA Kohn-Sham eigenvalue of the highest occupied level is a good estimate of the mean value of the ionization energy and the electron affinity (cf. Figs. \[Na3s\], \[gurney\] and their discussion). [^6]: The late $5d$ transition metals (Ir, Pt, Au) are exceptions to this rule. In these systems relativistic effects give rise to a substantial lowering of the $s$-band and a rather high surface stress. For the (001) surfaces, the desire to achieve a higher coordination in the top layer is indeed stronger than the cost of breaking (or stretching) some bonds between the first and second layer (see Fiorentini et al., 1993). [^7]: We note that the term “surface oxide” is not well defined, and in particular for Ru, which can exist in many oxidation states, it may be difficult, if not impossible, to distinguish between a 2-3 layer thick “surface oxide” and an on-surface plus sub-surface adsorbate phase. [^8]: Here 2.12 eV is the average adsorption energy of O at coverage $\Theta = 0.75$ (cf. Fig. 5.39), i.e., corresponding to the (2 $\times$ 2)-3O structure. [^9]: O(a) labels the oxygen atom which is adsorbed to the metal substrate
{ "pile_set_name": "ArXiv" }
--- abstract: | At non-equilibrium phase transitions into absorbing (trapped) states, it is well known that the directed percolation (DP) critical scaling is shared by two classes of models with a single (S) absorbing state and with infinitely many (IM) absorbing states. We study the crossover behavior in one dimension, arising from a considerable reduction of the number of absorbing states (typically from the IM-type to the S-type DP models), by following two different (excitatory or inhibitory) routes which make the auxiliary field density abruptly jump at the crossover. Along the excitatory route, the system becomes overly activated even for an infinitesimal perturbation and its crossover becomes discontinuous. Along the inhibitory route, we find continuous crossover with the universal crossover exponent $\phi\simeq 1.78(6)$, which is argued to be equal to $\nu_\|$, the relaxation time exponent of the DP universality class on a general footing. This conjecture is also confirmed in the case of the directed Ising (parity-conserving) class. Finally, we discuss the effect of diffusion to the IM-type models and suggest an argument why diffusive models with some hybrid-type reactions should belong to the DP class. author: - 'Su-Chan Park' - Hyunggyu Park title: 'Nontrivial critical crossover between directed percolation models: Effect of infinitely many absorbing states' --- \[Sec:intro\]introduction ========================= The directed percolation (DP) has been studied extensively as one of typical dynamic critical phenomenon far from equilibrium [@H00]. Nonequilibrium phase transitions of systems with a unique absorbing state are found to belong to the DP class if no symmetry or conservation of the order parameter is present [@DPC]. Even systems with infinitely many (IM) absorbing states such as the pair contact process (PCP) [@J93] are believed to share the same critical behavior with the DP, but the theoretical understanding of the IM-type DP (DP${}_\textrm{IM}$) models is still lacking. For example, the phenomenological field theory introduced in Ref. [@FT_PCP] and elaborated in Ref. [@W02] shows inconsistency with the numerical studies. In fact, the field theory based on the phenomenological Langevin equation predicts that the PCP should belong to the dynamic percolation rather than the DP class [@W02]. Besides the universality issue, its spreading dynamics is also not fully understood [@GCR; @DH]. One may reduce the number of absorbing states significantly by introducing particle diffusion (PCP with diffusion or PCPD) [@HH04]. Surprisingly, the PCPD has brought up serious turmoil, which could not be settled down in spite of extensive numerical [@KC03; @NP04; @H06] and analytical [@JvWOT04] studies. The answer seems to be one of two possibilities: The PCPD belongs to the DP class with a long transient, or forms a new universality class distinct from known universality classes to date. As an attempt to resolve the issue, the present authors suggested two different approaches to the PCPD. First, introducing the dynamic perturbation which is implemented by the biased hopping, we showed that the one-dimensional PCPD with biased diffusion (driven PCPD or DPCPD), exhibits critical scaling distinct from the unbiased PCPD, but instead shares the critical behavior with the two-dimensional PCPD without bias (dimensional reduction) [@PP05a; @PP05b]. Since all known DP models are robust against the biased diffusion [@PP05a], a DPCPD-type variant can serve as a litmus test for the PCPD scaling [@P06]. Second, we studied the crossover behavior from the PCPD to the DP by introducing single-particle annihilation/branching reactions and showed that there are diverging crossover scales with the universal nontrivial crossover exponent [@PP06]. These results provided another evidence that the PCPD is distinct from the DP. In this paper, we study the crossover behavior from the DP${}_\textrm{IM}$ to the DP in order to understand better the difference between these two “equivalent” (DP) universality classes. Actually, it would be absurd to talk about the crossover between two models belonging to the identical universality class (see Sec. \[Sec:single\_to\_single\]). However, the DP${}_\textrm{IM}$ models differ from the DP models in regard to the “non-order” parameter at the transitions, which shows a discontinuous singularity at the crossover. This singularity induces a well-defined and nontrivial crossover from the DP${}_\textrm{IM}$ to the DP, more generally a crossover arising from a considerable reduction of the number of absorbing states (between two different DP${}_\textrm{IM}$). We find two distinct crossover behaviors depending on the routes to reduce the number of absorbing states. Infinitesimal inclusion of an “excitatory” process (like single-particle branching) makes the system overly active, which gives rise to a discontinuous crossover (see Sec. \[Sec:dpima\]). While, the opposite “inhibitory” route (like single-particle annihilation) reveals a continuous crossover with the nontrivial crossover exponent $\phi\simeq 1.78(6)$. We find that this crossover exponent is universal for various kinds of models including the PCP and the triplet contact process (TCP). We argue that the crossover exponent is not independent but equal to $\nu_\|$, the relaxation time exponent of the DP universality class on a general footing. This conjecture is also confirmed in the case of the directed Ising (parity-conserving) class (see Sec. \[Sec:discussion\]). The PCPD [*per se*]{} can be considered one of crossover models from the DP${}_\textrm{IM}$ by allowing diffusion, which reduces the number of absorbing states considerably (from exponentially many to linearly many absorbing states with respect to system size). However, the crossover study from the PCP to the PCPD does not give any useful information on the DP${}_\textrm{IM}$, because the particle diffusion makes the system more active (excitatory) and the crossover turns out to be discontinuous. Finally, we study the crossover from the diffusive reaction models to the DP. Such an example is the crossover from the PCPD to the DP studied in Ref. [@PP06]. In Sec. \[Sec:dptodp\], the hybrid-type models with diffusion ([tp12]{}: $2A\rightarrow\emptyset, 3A\rightarrow 4A$) [@KC03] are perturbed by adding a single-particle annihilation process ($A\rightarrow\emptyset$) and its crossover to the DP is investigated. These hybrid-type models (where the branching process is of higher order than the annihilation process) are numerically known to belong to the DP class. As expected, the critical line emanates “linearly” from the hybrid-type model point. We suggest an argument why these models should belong to the DP rather than a PCPD-type nontrivial class. \[Sec:single\_to\_single\] Crossover between the identical universality class? ============================================================================== This section considers a $d$-dimensional stochastic system of hard core particles with dynamics summarized in Table \[Table:DP\]. In Ref. [@PP05c], it is shown that two different stochastic systems modeled by tilded and untilded rates are equivalent if transition rates satisfy the relations $$\begin{aligned} \tilde \sigma = \mu \sigma&,\qquad 2 \tilde D + \tilde \xi = 2 D + \xi,\\ \tilde D + \frac{\tilde \sigma}{2} &= D + \frac{\sigma}{2},\\ \tilde \lambda + \frac{\tilde \eta}{2} + \frac{\tilde \sigma}{2} &= \frac{1}{\mu}\left (\lambda + \frac{\eta}{2}+\frac{\sigma}{2}\right ),\\ \tilde \lambda + \tilde \eta + \tilde \sigma - 2 \tilde D &= \lambda + \eta+\sigma - 2 D , \end{aligned} \label{Eq:prime_unprime}$$ where any positive number $\mu$ which renders all tilded and untilded transition rates be nonnegative is physically meaningful. By equivalence is meant that all correlation functions of the model with tilded parameters can be deduced from those of the model with untilded parameters and vice versa. For instance, the particle density $\rho$ at time $t$ of two different stochastic many body systems becomes [@PP05c] $$\rho(D,\lambda,\eta,\xi,\sigma;\rho_0,t) = \frac{1}{\mu} \tilde \rho(\tilde D, \tilde \lambda,\tilde \eta,\tilde \xi,\tilde \sigma;\tilde \rho_0,t), \label{Eq:rho_relation}$$ if the initial density of two systems has the relation $\tilde \rho_0 = \mu \rho_0$. Needless to say, both $\rho_0$ and $\tilde \rho_0$ should lie between 0 and 1. Although the equivalence is shown only for the one-dimensional systems in Ref. [@PP05c], Eq.  is generally true for any $d$-dimensional systems, which can be easily shown by the same technique developed in Ref. [@PP05c]. -- ------------------------------------------ ----------------- $A\emptyset \leftrightarrow \emptyset A$ $D/d$ $ A A \rightarrow \emptyset \emptyset$ $\lambda/d$ $ A A \rightarrow A \emptyset$ $ {\eta}/(2d)$ $ A A \rightarrow \emptyset A$ $ {\eta}/(2d) $ $A \rightarrow \emptyset$ $\xi$ $\emptyset A \rightarrow A A $ ${\sigma}/(2d)$ $ A \emptyset \rightarrow A A $ $\sigma/(2d) $ -- ------------------------------------------ ----------------- : \[Table:DP\] $d$-dimensional reaction-diffusion processes of single species with hard core exclusion and their rates. Now consider the branching annihilating random walks with one offspring (BAW1) [@TT92] which corresponds to the model with $\xi = 0$ in Table \[Table:DP\]. The parameters used in Ref. [@TT92] in one dimension are $D = p/2$, $\lambda = p$, and $\eta = \sigma = 1-p$ with the tuning parameter $p$. If $w \equiv \mu - 1$ is very small and nonnegative, the solution of Eq.  up to the order of $w$ is $$\begin{aligned} &\tilde D -D= - \frac{w}{2} \sigma,\quad \tilde \sigma -\sigma = w \sigma,\quad \tilde \xi = w \sigma,\\ &\tilde \lambda - \lambda \simeq - w (\eta+ 2 \lambda),\quad \tilde \eta - \eta \simeq w (\eta + 2 \lambda - 2 \sigma). \end{aligned} \label{Eq:crossBAW}$$ If $w$ is sufficiently small, it is always possible to associate the BAW1 with a stochastic process with spontaneous death in an equivalent way with all nonnegative rates. Since the BAW1 in high dimensions is also known to have a nontrivial transition point [@CCD04], the following discussion is valid in any spatial dimension. The transition points for stochastic systems with small $w$ can be always calculated exactly from Eq.  (approximately from Eq. ), if the transition point of the BAW1 is given. The conclusions from the above analysis are two-fold. First, it is clear from Eq.  that the phase boundary (critical line) should meet the BAW1 transition point [*linearly*]{} with finite slope, as $w$ vanishes. This implies that there is no additional singularity involved near $w=0$, which is fully expected from the crossover between models with the identical universality class. If one defines the crossover exponent $\phi$ from the shape of the critical line near $w=0$ (see Sec. \[Sec:dpima\]), one can say that $\phi=1$. Second, the critical decay of the density is given by $\tilde \rho_c(t) = (1+w) \rho_c(t)$ from Eq. , which implies that there is no diverging crossover time scale for small $w$. Since the introduction of the spontaneous death does not change the structure of the absorbing phase space (single absorbing state) let alone the universality class, the above analysis is in good harmony with the naive expectation as to the “crossover” between two models belonging to the identical class. In the next section, however, we will show that the substantial change of the absorbing phase space without affecting the universality class will trigger a nontrivial crossover. \[Sec:dpima\]Crossover from the DP${}_\textrm{IM}$ to the DP ============================================================ Unlike the BAW1, the pair contact process (PCP) is the prototype of the DP${}_\textrm{IM}$ models with exponentially many absorbing states. By introducing single-particle reactions to the PCP, the number of absorbing states changes drastically from infinity to one. This section shows that this qualitative change is reflected into the singular behavior of the phase boundary close to the PCP transition point in one dimension. The dynamics of the model is summarized as \[Eq:cross\_PCP\] $$\begin{aligned} \label{Eq:cross_PCP_a} &&AA \stackrel{p}{\longrightarrow} \emptyset \emptyset,\quad \left .\begin{matrix} AA\emptyset\\ \emptyset AA \end{matrix} \right \} \begin{CD} @>(1-p)/2>> AAA, \end{CD} \\ &&\begin{CD} \left . \begin{matrix} A\emptyset\\ \emptyset A \end{matrix} \right \} @>(1-q) w/2>> \emptyset \emptyset,\quad \left . \begin{matrix} A\emptyset\\ \emptyset A \end{matrix} \right \} @>q w/2>> AA, \end{CD} \label{Eq:cross_PCP_b}\end{aligned}$$ where $0\le q \le 1$. For the PCP at $w=0$, any configuration without a pair of neighboring particles (a mixture of isolated particles and vacant sites) is absorbing and its number grows exponentially with system size. The order parameter of the PCP is the pair density (the number density of $AA$ pairs) and the particle density field is [*auxiliary*]{} which is finite even in the absorbing phase. At nonzero $w$, an isolated particle becomes active and only the vacuum becomes the true absorbing state. In this case, the particle density is usually adopted as the order parameter and the pair density scales in the same way. Figure \[Fig:PCP\_pc\] locates the transition point of the PCP ($w=0$) at $p_0 = 0.077\;0905(5)$ by exploiting the critical decay of the pair density as $\rho_p(t) \sim t^{-\delta}$ with $\delta$ to be the critical exponent of the DP class whose accurate value can be found in [@J99]. In numerical simulations, the system size is $L=2^{18}$ and the number of independent samples are 750, 1500, and 400 for the data in the active, critical, and absorbing phases, respectively. The flatness of $\rho_p(t) t^{\delta}$ over four log decades in time confirms the solid DP critical scaling of the PCP. At finite $w$, the model still belongs to the DP class irrespective of $q$. Unlike the PCPD to the DP crossover model in Ref. [@PP06], however, the critical lines show two completely different singular behaviors, depending on the value of $q$. For large $q$, the activity of the system is enhanced by additional single-particle reaction processes (excitatory process) and the system becomes overly activated even with infinitesimal $w$. The critical line does not converge to the PCP critical point as $w$ decreases to zero ($w=0^+$) and shows a discontinuous jump. On the other hand, for small $q$, the system activity is suppressed (inhibitory process) and the system becomes more inactive. The critical line nicely converges to the PCP critical point and shows a continuous crossover with a nontrivial crossover exponent. ![\[Fig:PCP\_pc\] (Color online) Semilogarithmic plot of $\rho_p(t) t^{\delta}$ vs $t$ of the PCP near criticality with $\delta = 0.1595$ to be the exponent of the DP. Since the upper (lower) curve veers up (down), we estimate the critical point as $p_0 = 0.077\;0905(5)$ with the error in the last digit by 5.](ParkFig1.eps){width="45.00000%"} ![\[Fig:A2A\] (Color online) Plots of $\rho(t) t^\delta$ vs $t$ for $w=10^{-3}$, $10^{-4}$, and $w=10^{-5}$ at $q=1$ close to the critical points in semi-logarithmic scales. Again, $\delta$ assumes the DP value. The curves in the middle correspond to $p=0.1451$, $0.1448$, and $0.1448$ (from bottom to top), respectively and the value of $p$’s of other two curves are $\pm 0.0001$ off from the middle value. As $w$ becomes smaller, the relaxation time becomes larger though the critical point does not change much and approaches $p \simeq 0.1448$. ](ParkFig2.eps){width="48.00000%"} First, we choose the $q=1$ case as a typical excitatory route of the crossover from the DP${}_\textrm{IM}$ to the DP. As shown in Fig. \[Fig:A2A\], the critical line approaches $p\simeq 0.1448$ as $w$ approaches zero, which is by far above the critical value of the PCP ($p_0\simeq0.077$). So there is a big jump of the critical line at $w=0$. The discontinuity can be understood as follows: Consider a system with $p$ slightly above the PCP critical point $p_0$ and $0<w\ll\tau^{-1}$ where $\tau$ is the relaxation time which is finite off criticality. Then the single-particle branching event ($A\rightarrow 2A$) with the characteristic time of $w^{-1}$ occurs effectively after the system falls into one of the PCP absorbing states in which the isolated particle density is finite. Since the branching event creates a new pair, the system is reactivated and performs the damage-spreading-type “defect dynamics” for some time proportional to $\tau$ and again falls into one of the PCP absorbing states. This defect dynamics continues forever with the period of time $w^{-1}$. As the particle density is finite (and quite large) even in the PCP absorbing states, the time-averaged particle density in this iterated process should be finite in this region of the phase diagram. This implies that the continuous absorbing phase transition at infinitesimal $w$ into vacuum should occur way above $p_0$, which is consistent with our finding. Note that the discontinuity in the auxiliary field (particle) density is crucial in this crossover. Actually, the same argument can be applied to the crossover from the PCP to the PCPD. We can introduce the diffusion rather than the single-particle reactions and again consider $p$ slightly above $p_0$. Let $\rho_0$ denote the isolated particle density at the PCP absorbing states, then the characteristic length scale between isolated particles is $1/\rho_0$. If $0< D\rho_0^2 \ll \tau^{-1}$ with the diffusion constant $D$, the “defect dynamics” will continue again indefinitely for small $D$. So the phase boundary in the $D-p$ plane should have a discontinuity at $D=0$. $w$ $p_c(w)$ ------------------- ------------------ $0 $ $0.077~0905 (5)$ $10^{-5} $ $0.077~002 (3)$ $5\times 10^{-5}$ $0.076~885 (3)$ $10^{-4} $ $0.076~784 (4)$ $2\times 10^{-4}$ $0.076~642 (2)$ $3\times 10^{-4}$ $0.076~530 (5)$ $4\times 10^{-4}$ $0.076~432 (2)$ $5\times 10^{-4}$ $0.076~345 (2)$ $6\times 10^{-4}$ $0.076~264 (1)$ $10^{-3} $ $0.075~988 (4)$ : \[Table:PCP\_pc\] Critical points of the model with dynamics of Eq.  for some values of $w$’s at $q=0$. The numbers in the parentheses indicate the error of the last digits. Let us turn to the crossover model with $q=0$, which should represent a typical inhibitory route. Table \[Table:PCP\_pc\] summarizes the critical points of the model for some $w$’s at $q=0$ and the corresponding phase boundary is plotted in Fig. \[Fig:PCPphase\]. Unlike the previous case, the reactive phase shrinks continuously with the rate of additional single-particle annihilation process and the phase boundary is continuous. The usual analysis method can be applied to this case [@DL9]. If we define $\Delta = (p_0 - p)/p_0$ and $\Delta_c(w) = (p_0 - p_c(w))/p_0$, the phase boundary is well fitted by $\Delta_c \sim w^{1/\phi}$ with $\phi^{-1} = 0.56(2)$ or $\phi = 1.78(6)$; see the inset of Fig. \[Fig:PCPphase\]. Let us assume the existence of the well-defined crossover scaling which is described by the scaling function [@DL9] $$\rho_p(\Delta,w;t) = t^{-\delta} {\cal F}(\Delta^{\nu_\|} t, w^{\mu_\|} t), \label{Eq:scaling}$$ where $\rho_p$ is the pair density and $\mu_\| = \nu_\|/\phi$ with $\phi$ estimated in the above. We examine whether the scaling function in Eq.  correctly describe the crossover near the PCP critical point. ![\[Fig:PCPphase\] (Color online) Phase boundary of the model of Eq.  at $q=0$ in $w-p$ plane. Symbols locate the numerically estimated critical points. The error of the critical point is smaller than the symbol size. The curve shows the least-square-fit result of the phase boundary. The absorbing (active) phase is above (below) the curve. Inset: the same but the vertical axis is $\Delta_c(w)$ in log-log scale. The slope corresponds to the inverse of the crossover exponent which is estimated as $\phi = 1.78(6)$. ](ParkFig3.eps){width="45.00000%"} ![\[Fig:mucollapse\] (Color online) Log-log plot of the scaling function Eq.  for the PCP crossover model using $\delta = 0.1595$ and $\mu = 0.97$. All curves are collapsed into a single curve. ](ParkFig4.eps){width="45.00000%"} First, we measure the pair density for various $w$’s at $\Delta=0$. From the scaling ansatz , the pair density at $\Delta = 0$ should collapse as $$t^{\delta} \rho_p(t) = {\cal G}(w^{\mu_\|} t ). \label{Eq:mucoll}$$ With $\mu_\| \simeq 0.97$, all curves for the pair density are collapsed into a single curve as Fig. \[Fig:mucollapse\] shows. Next, we take $\Delta = \Delta_c(w)$ along the critical line. Since $\Delta_c(w) \simeq w^{1/\phi}$ and $\nu_\|/\phi = \mu_\|$, the scaling function should take the form $$t^{\delta} \rho_p(\Delta_c(w);t) = {\cal H}(w^{\mu_\|} t ), \label{Eq:DCcoll}$$ where ${\cal H}(x)$ approaches a constant as $x \rightarrow \infty$. In Fig. \[Fig:pair\], all curves at different critical points collapse well into a single curve. Hence we conclude that the scaling function Eq.  correctly describes the crossover behavior from the DP${}_\textrm{IM}$ to the the DP. ![\[Fig:pair\] (Color online) Scaling plot of $\rho_p(t) t^{\delta}$ vs $w^{\mu_\|}t$ with $\delta = 0.1595$ and $\mu_\|=0.97$ at criticality in semi-log scales. Inset: Scaling function ${\cal H}$ at finite $w$ and $w=0$. The asymptotic value of $w=0$ is different from the $w\rightarrow 0$ limit. ](ParkFig5.eps){width="45.00000%"} ![\[Fig:single\] (Color online) Log-log plot of the particle density for some finite $w$’s at $\Delta = 0$. For any finite value of $w$, the particle density $\rho(t)$ will go to zero, which is clearly different from the $w=0$ case. ](ParkFig6.eps){width="45.00000%"} Since the models at $w=0$ and at $w\neq 0$ belong to the same DP universality class, it is natural to ask what is the origin of such a nontrivial singularity near the PCP critical point. The inset of Fig. \[Fig:pair\] gives a hint to this question, which shows that the scaling function ${\cal H}$ (the amplitude of the critical decay) does not approach the PCP value as $w$ goes to zero, i.e., it is not continuous at $w=0$. The discontinuity in this amplitude must originate again from the discontinuity in the auxiliary field (particle) density. One can see it directly from the behavior of the particle density ($\rho$). Unlike the pair density, $\rho$ can not be described by the scaling function . Consider again the case at $\Delta = 0$ and nonzero $w$. For any finite value of $w$ in the thermodynamic limit, $\rho(t)$ approaches to zero as $t\rightarrow \infty$; see Fig. \[Fig:single\]. On the other hand, the model at $\Delta = w=0$ (the critical PCP), has a nonzero density of $\rho(t)$ as $t\rightarrow \infty$. In other words, the $w\rightarrow 0$ limiting process is different from the $w=0$ model itself in regard to the auxiliary field density. To check the universality of the crossover exponent, we study the modified PCP with the replacement of $2A \rightarrow 3A$ with $3A \rightarrow 4A$ in Eq.  which is the model of Eq.  with no diffusion ($D=0$). This model also has infinitely many absorbing states and belongs to the DP${}_\textrm{IM}$ class. By introducing single-particle reactions ($w\neq 0$), the same crossover behavior is found as the above (data not shown). ![\[Fig:TCPcollapse\] (Color online) Scaling plot of $\rho_t(t) t^{\delta}$ vs $w^{\mu_\|} t$ with the same exponents in Fig. \[Fig:mucollapse\] for the crossover TCP model of Eq.  in semi-log scales. As in Fig. \[Fig:mucollapse\], all curves are collapsed into a single curve.](ParkFig7.eps){width="45.00000%"} We also study more general crossover behavior from one DP${}_\textrm{IM}$ to another DP${}_\textrm{IM}$ with the considerably reduced number of absorbing states. To be specific, we consider the triplet contact process (TCP) and its crossover model by introducing the $2A \rightarrow A$ process without spontaneous death. The TCP with pair dynamics is defined as $$\label{Eq:TCPcross} \begin{aligned} &AAA \stackrel{p}{\longrightarrow} \emptyset \emptyset\emptyset,\quad \left .\begin{matrix} AAA\emptyset\\ \emptyset AAA \end{matrix} \right \} \begin{CD} @>(1-p)/2>> AAAA, \end{CD}\\ &\begin{CD} AA\emptyset @>w/2>> A\emptyset \emptyset,\quad \emptyset AA @>w/2>> \emptyset \emptyset A. \end{CD} \end{aligned}$$ The above model has infinitely many absorbing states, but with nonzero $w$ the number of absorbing states is greatly reduced. At $w=0$, there is again a jump in the auxiliary field density (here, the pair density). We found that the critical point for the TCP at $w=0$ is $p_c = 0.036~865(5)$, exploiting the DP critical scaling (data not shown). Figure \[Fig:TCPcollapse\] shows the scaling plot of the triplet density $\rho_t$ in the same way as in Fig. \[Fig:mucollapse\]. We also measured the crossover exponent from the phase boundary and found the same exponent (data not shown). Hence we conclude that there is the well-defined and universal crossover scaling from the DP${}_\textrm{IM}$ to the DP which is mediated by the significant reduction of the number of absorbing states. The discontinuity in the auxiliary field plays a crucial role in this nontrivial crossover. \[Sec:discussion\] Conjecture on the crossover exponent $\phi$ ============================================================== The crossover exponent from the DP${}_\textrm{IM}$ to the DP is estimated as $\phi = 1.78(6)$. Since this crossover occurs between the same universality class, we are suspicious that $\phi$ may not be independent but related to the well-known DP critical exponents. Actually, we argue that the crossover exponent is given by the DP relaxation time exponent: $\phi = \nu_\|\simeq 1.733$ (or $\mu_\|=1$) which is compatible with the numerical estimation within error. The reason is as follows: Take the model of Eq.  at $q=0$. The critical line should be determined by the competition between the single-particle annihilation process ($A\rightarrow\emptyset$) parameterized by $w$ and the multi-particle (pair) reaction process ($2A\rightarrow\emptyset$ or $3A$) parameterized by $\Delta\sim(p_0-p)$. We expect both events should appear at the same time scale along the critical line to balance off each other. Since the single-particle (auxiliary field) density is finite at the PCP (DP${}_\textrm{IM}$) point, the time scale for $A\rightarrow\emptyset$ should be simply proportional to $w^{-1}$. The time scale for the pair reaction process should be given by the relaxation time scale $\tau\sim \Delta^{-\nu_\|}$. Consequently, the critical line is determined as $\Delta_c\sim(p_0-p_c(w)) \sim w^{1/\nu_\|}$, which yields $\phi=\nu_\|$ and equivalently $\mu_\|=1$. Considering the crossover between the DP models with a unique absorbing state, the time scale for the process parameterized by $w$ is proportional to $w^{-\nu_\|}$ like the other competing process because the auxiliary field (particle) density is also vanishing critically as $w$ decreases to zero. In this case, we get $\phi=1$ and equivalently $\mu_\|=\nu_\|$, which is consistent with our result in Sec. \[Sec:single\_to\_single\]. $w$ $p_c(w)$ ------------------- ------------------ $0 $ $0.4518 (1)$ $10^{-4} $ $0.4443 (2)$ $3\times 10^{-4}$ $0.4405 (2)$ $10^{-3} $ $0.4353 (1)$ $3\times 10^{-3}$ $0.4276 (1)$ : \[Table:DI\_pc\] Critical points of the DI crossover model for some values of $w$’s (see the text). The numbers in the parentheses indicate the error of the last digits. Since our argument for the crossover exponent is generally applicable to any universality class, we can check its validity through studying the similar type crossover between the directed Ising (DI) class models [@DIPark]. Consider a one-dimensional system with two species, say $A$ and $B$. Between the same species, hard core exclusion is applied, but different species can reside at the same site. The dynamic rules are as follows: The dynamics always starts with an $A$ particle. A randomly chosen $A$ particle can hop to one of nearest neighbors with probability $p$. If two $A$ particles meet at the same site by hopping, both particles are removed with probability $\frac{1}{2}$. If this annihilation attempt fails, the particle goes back to the original site. With probability $1-p$, a $B$ particle is generated at the same site occupied by the chosen $A$ particle. If that site is already occupied by another $B$ particle, the two $B$ particles transmute to two $A$ particles which will be placed at two nearest neighbor sites. If any of transmuted $A$ particles is placed at the site already occupied by another $A$ particle, both particles are annihilated immediately. In summary, $2A\rightarrow\emptyset$, $A\rightarrow A+B$, and $2B\rightarrow 2A$ processes are allowed with $A$ particle diffusion. Since $B$ particles are not allowed to hop, the system is inactive without an $A$ particle but only with $B$ particles. The number of the absorbing states grows exponentially with system size and the auxiliary field ($B$ particle) density is finite at the absorbing transition. Besides, the number of $A$ particles is conserved modulo 2, which is the characteristic of the DI (or parity-conserving) class. As in Sec. \[Sec:dpima\], we study the crossover by introducing spontaneous annihilation of $B$ particles (inhibitory route) with rate $w$. As expected, we find the DI critical scaling for both $w=0$ and $w\neq 0$ cases. We locate the critical line by exploiting the known DI critical exponents [@DIPark], which is summarized in Table \[Table:DI\_pc\] for some $w$’s. Since the critical exponent $\nu_\|$ of the DI class ($\simeq 3.25$) is much larger than that of the DP ($\simeq 1.733$), the accuracy of the critical points in Table \[Table:DI\_pc\] is worse than that for the DP cases. From these data, one can estimate the crossover exponent $1/\phi_\text{DI}$ as $0.34(4)$ which should be compared with $1/\nu_\|$ of the DI class ($\simeq 0.31$). Hence we conclude that our argument for $\phi=\nu_\|$ also applies to the crossover from the IM-type DI to the DI models. \[Sec:dptodp\]Diffusion effect ============================== This section studies how the crossover scaling is affected if particles are allowed to hop in models considered in Sec. \[Sec:dpima\]. With single-particle diffusion, the PCP becomes the PCPD where the particle (auxiliary field) density as well as the pair density vanishes at criticality even without any single-particle reaction process. The crossover from the PCPD to the DP caused by including single-particle reactions has been studied previously by the present authors [@PP06] where the value of the crossover exponent is reported as $1/\phi = 0.58(3)$. This value is quite close to that obtained for the crossover from the DP${}_\textrm{IM}$ to the DP in Sec. \[Sec:dpima\]. This similarity may mislead one to jump to the wrong conclusion that the critical nature of the PCP and the PCPD is equivalent. However, one should remember that the origin of the nontrivial crossover from the PCP to the DP lies in the finiteness of the auxiliary field density at the PCP critical point, while it vanishes at the PCPD critical point. So, if the PCPD belongs to the DP class, then one should expect a trivial crossover with $\phi=1$. Our finding of the nontrivial crossover in [@PP06] implies that the PCPD class is distinct from the DP class. Hence it is likely that the similarity of two crossover exponents be a mere coincidence. The distinction of the PCPD from the DP can also be evidenced by the study of the crossover model with the hybrid-type reaction dynamics $2A \rightarrow 0$ and $3A \rightarrow 4A$. Without diffusion, this model belongs to the DP${}_\textrm{IM}$ and its crossover to the DP was studied in Sec. \[Sec:dpima\]. Even if particles are allowed to diffuse, this model (called as [tp12]{}) is numerically known to belong to the DP class [@KC03]. ![\[Fig:tp12\] (Color online) Semi-log plot of $\rho(t) t^{\delta}$ vs $t$ with $\delta = 0.1595$ at $p_c =0.058~427,~0.058~723,~0.058~966$, and $0.059~031$ for $w=10^{-3},~5\times 10^{-4},~10^{-4}$, and $0$ (from left to right), respectively. There is no diverging time scale as $w$ approaches zero. Inset: Phase boundary near $w=0$. Each symbol corresponds to the critical point used in the main figure. The straight line is drawn between two consecutive points, independently. ](ParkFig8.eps){width="45.00000%"} It will be illuminating to see how the diffusion in the [tp12]{} can change the crossover behavior to the DP. The dynamics of the model is summarized as $$\label{Eq:tp12} \begin{aligned} &AA \stackrel{p}{\longrightarrow} \emptyset \emptyset,\quad \left .\begin{matrix} AAA\emptyset\\ \emptyset AAA \end{matrix} \right \} \begin{CD} @>(1-p)/2>> AAAA, \end{CD} \\ &\begin{CD} \left . \begin{matrix} A\emptyset\\ \emptyset A \end{matrix} \right \} @>w/2>> \emptyset \emptyset,\quad A\emptyset \stackrel{D}{\longleftrightarrow} \emptyset A, \end{CD} \end{aligned}$$ where $D = (1-w)/2$. The model at $w=0$ is the [tp12]{}. Our numerical results in Fig. \[Fig:tp12\] show the typical “crossover” behavior between the identical universality class discussed in Sec. \[Sec:single\_to\_single\]. The critical line converges to the $w=0$ point [*linearly*]{} ($\phi=1$) and there is no diverging time scale as $w$ becomes smaller with almost perfect collapse of all critical density decay curves near $w\approx 0$. Our crossover study provides another strong numerical evidence that the [tp12]{} belongs to the DP class. ![\[Fig:Spacetime\] (Color online) Space-time configurations of (a) the [tp12]{} and (b) the PCPD at criticality. The initial density is (a) $\frac{1}{16}$ for the [tp12]{} and (b) $\frac{1}{32}$ for the PCPD with size $L = 2^{10}$. Isolated particles (particles in a cluster) are designated by the blue (red) color.](ParkFig9.eps){width="45.00000%"} If the PCPD does not belong to the DP class, why should the [tp12]{} belong to the DP class? The PCPD and the [tp12]{} seem quite similar in the sense of multi-particle nature in reaction dynamics and also the absorbing space structure (vacuum and a single-particle state). However, they are quite different in the role of diffusing isolated particles. See Fig. \[Fig:Spacetime\] for the space-time configurations for the [tp12]{} and the PCPD at criticality, starting from the low density initial condition without pairs. The diffusing isolated particles of the [tp12]{} cannot increase the number of particles in most cases, because $2 A \rightarrow \emptyset$ dynamics dominates over $3 A \rightarrow 4 A$ dynamics: Pairs generated by collisions of two isolated particles evaporate before greeting another isolated particle to become “active” triplets. Consequently there is effectively no feedback mechanism from isolated particles to increase the particle density or make the system more active. Therefore the region of isolated particles can be regarded as absorbing like in the PCP model. This case may correspond to the [*no-feedback*]{} point ($r=0$) for the generalized PCPD (GPCPD) studied in [@NP04], which is the DP fixed point. This argument can be generalized to systems with hybrid-type reaction dynamics $ m A \rightarrow (m+k)A$ and $ n A \rightarrow (n-l) A$ with $m>n$ and $k,l>0$, which are numerically shown to belong to the DP class [@KC03]. The isolated particles of the PCPD, however, cannot be regarded as absorbing as Fig. \[Fig:Spacetime\] shows; the isolated particles may affect the critical spreading actively, because both dynamics of $2 A \rightarrow \emptyset$ and $2 A \rightarrow 3A$ compete each other and consequently there is an effective feedback mechanism from diffusing particles to make the system active. This corresponds to the GPCPD with [*long-term memory*]{} effects at $r\neq 0$ [@NP04]. \[Sec:sum\]Summary and conclusion ================================= In summary, we studied the crossover from the model belonging to the directed percolation (DP) class with infinitely many absorbing states (DP${}_\textrm{IM}$) to the DP model by reducing the number of absorbing states significantly. The crossover is found to be well described by the usual crossover scaling function for the order parameter. The crossover exponent $\phi$ is argued to be related to one of the DP critical exponent, [*i.e.*]{}, $\phi = \nu_\|$, which is further evidenced by the similar crossover model belonging to the directed Ising class. The origin of the diverging scale and the nontrivial crossover comes from the discontinuity of the auxiliary field density at the DP${}_\textrm{IM}$ critical point. Our study for the first time presents the existence of the nontrivial scaling in the DP${}_\textrm{IM}$, which is compared with the study on the spreading exponents. We also studied how the crossover behavior from the DP${}_\textrm{IM}$ to the DP is affected by particle diffusion. The crossover from the pair contact process with diffusion (PCPD) to the DP studied in Ref. [@PP06] is well classified by the nontrivial crossover exponent. On the other hand, the [tp12]{} which is known to belong to the DP class is characterized by the trivial “crossover” between the identical class detailed in Sec. \[Sec:single\_to\_single\]. This provides an additional evidence supporting that the PCPD is distinct from the DP. In addition, we suggest an argument based on the role of diffusing isolated particles why the [tp12]{} should belong to the DP class, but the PCPD does not need to be. It will be a challenging problem to see if the crossover scaling from the DP${}_\textrm{IM}$ to the DP can be anticipated in the framework of the field theory [@FT_PCP]. [99]{} H. Hinrichsen, Adv. Phys. [**49**]{}, 815 (2000); G. Ódor, Rev. Mod. Phys. [**76**]{}, 663 (2004). H.K. Janssen, Z. Phys. B [**42**]{}, 151 (1981); P. Grassberger, [*ibid*]{}. [**47**]{}, 364 (1982). I. Jensen, Phys. Rev. Lett. [**70**]{}, 1465 (1993). M. A. Muñoz, G. Grinstein, R. Dickman, and R. Livi, Phys. Rev. Lett. [**76**]{}, 451 (1996); Physica D [**103**]{}, 485 (1997). F. van Wijland, Phys. Rev. Lett. [**89**]{}, 190602 (2002). P. Grassberger, H. Chaté, and G. Rousseau, Phys. Rev. E [**55**]{}, 2488 (1997). A. Jiménez-Dalmaroni and H. Hinrichsen, Phys. Rev. E [**68**]{}, 036103 (2003). For a review, see M. Henkel and H. Hinrichsen, J. Phys. A: Math. Gen. [**37**]{}, R117 (2004). J. Kockelkoren and H. Chaté, Phys. Rev. Lett. [**90**]{}, 125701 (2003). J. D. Noh and H. Park, Phys. Rev. E [**69**]{}, 016122 (2004). H. Hinrichsen, Physica A [**361**]{}, 457 (2006). H.-K. Janssen, F. van Wijland, O. Deloubriere, and U. C. Täuber, Phys. Rev. E [**70**]{}, 056114 (2004). S.-C. Park and H. Park, Phys. Rev. Lett. [**94**]{}, 065701 (2005) S.-C. Park and H. Park, Phys. Rev. E [**71**]{}, 016137 (2005). S.-C. Park, Eur. Phys. J. B [**50**]{}, 327 (2006). S.-C. Park and H. Park, Phys. Rev. E [**73**]{}, 025105 (2006). S.-C. Park and J.-M. Park, Phys. Rev. E [**71**]{}, 026113 (2005). H. Takayasu and A.Y. Tretyakov, Phys. Rev. Lett. [**68**]{}, 3060 (1992). L. Canet, H. Chaté, and B. Delamotte, Phys. Rev. Lett. [**92**]{}, 255703 (2004). I. Jensen, J. Phys. A: Math. Gen. [**32**]{}, 5233 (1999). See, e.g., I.D. Lawrie and S. Sarbach, in [*Phase Transitions and Critical Phenomena*]{}, edited by C. Domb and J.L. Lebowitz (Academic Press, London, 1984), Vol. 9. M. H. Kim and H. Park, Phys. Rev. Lett. [**73**]{}, 2579 (1994); W. Hwang, S. Kwon, H. Park, and H. Park, Phys. Rev. E [**57**]{}, 6438 (1998); W. Hwang and H. Park, [*ibid*]{}. [**59**]{}, 4683 (1999).
{ "pile_set_name": "ArXiv" }
--- abstract: 'During development, the mammalian brain differentiates into specialized regions with distinct functional abilities. While many factors contribute to functional specialization, we explore the effect of neuronal density on the development of neuronal interactions [*in vitro*]{}. Two types of cortical networks, dense and sparse, with $50,000$ and $12,000$ total cells respectively, are studied. Activation graphs that represent pairwise neuronal interactions are constructed using a competitive first response model. These graphs reveal that, during development [*in vitro*]{}, dense networks form activation connections earlier than sparse networks. Link entropy analysis of dense network activation graphs suggests that the majority of connections between electrodes are reciprocal in nature. Information theoretic measures reveal that early functional information interactions (among 3 cells) are synergetic in both dense and sparse networks. However, during later stages of development, previously synergetic relationships become primarily redundant in dense, but not in sparse networks. Large link entropy values in the activation graph are related to the domination of redundant ensembles in late stages of development in dense networks. Results demonstrate differences between dense and sparse networks in terms of informational groups, pairwise relationships, and activation graphs. These differences suggest that variations in cell density may result in different functional specialization of nervous system tissue [*in vivo*]{}.' author: - 'Michael I. Ham' - Vadas Gintautas - 'Marko A. Rodriguez' - 'Ryan A. Bennett' - 'Cara L. Santa Maria' - 'Luìs M.A. Bettencourt' bibliography: - 'JCNSfile.bib' title: 'Density-dependence of functional development in spiking cortical networks grown [*in vitro*]{}' --- Introduction {#intro} ============ The mammalian brain is a remarkable structure composed of many specialized regions and types of cells. Despite organizational differences in neural tissue, the basic functional units of the nervous system, neurons, are generally similar across tissues, as are methods of forming and modifying synaptic connections between them (e.g. spike-timing dependent plasticity [@markram_1997], [@bi_1998]). Thus, functional specialization of brain regions is a function of neuron specialization, (e.g. excitatory or inhibitory), ratios of neurons to neuroglia, synaptic density, learning, and many other factors that are not yet well understood. Though the functional role of neurons in different types of tissues can be similar, neuronal density can vary dramatically. For example, neuronal density in human fascia dentata is $\approx3.2 X 10^5$ neurons/mm$^3$ [@kim_1990] while cortical tissue density is $\approx3.4 X 10^4$ neurons/mm$^3$ [@anderson_1996], nearly an order of magnitude difference. A previous study of network development [*in vitro*]{} demonstrated that network bursting (when a large majority of neurons fire in a coordinated pattern) and spiking patterns are affected by neuronal density [@wagenaar_2006]. Since spike and burst activity are observable results of neuronal interaction (functional connectivity) they can be used to infer neuronal relationships [@pfister_2006], [@bcourt_2007], [@tang_2008], [@bettencourt_2008]. In this work the effect of density on functional units is explored using dissociated cortical tissue developing [*in vitro*]{} on microelectrode arrays. These are well established models of neuronal interaction [@paninski_2003], [@beggs_2004], [@feinerman_2007], though obvious constraints and limitations must be considered when attempting to extrapolate between [*in vitro*]{} and [*in vivo*]{} structures [@corner_2002], [@marom_2002]. To gain new insight into functional connectivity in developing networks, we analyze the coordinated electrophysiological activity of groups of two and three spike trains, each representing an integration of all of the action potentials recorded at a single electrode. Each electrode may capture the activity of a single neuron, or less frequently, incorporate signals from several cells. Activation between pairs of electrodes is inferred using a competitive first response model, whereby directional links are derived from spike train data. These pairwise links estimate the probability that activity at one electrode causes activity at another. The set of all such links forms a network activation graph where vertices represent recording electrodes and weighted edges represent a dependent activation probability. We use entropy based link analysis (link entropy) to characterize connectivity between electrodes. Information theoretic measures applied to ensembles of $3$ electrodes reveal functional information structures. Such structures are characterized using an information overlap method that reveals whether interactions are synergetic (more information is obtained from the measurement of two electrodes together conditioned on the third compared to measuring them separately), redundant (less information is obtained from the measurement of two electrodes together conditioned on the third compared to measuring them separately) or independent. Results show that functional structures in networks are strongly influenced by neuronal density and suggest that varying cell density is a potential strategy for differentiating tissue functionality. Methods {#Methods} ======= We analyze developmental activity patterns in eight cultured cortical networks growing on microelectrode arrays. Networks were recorded [@wagenaar_2006] on days ranging from $3$ to $36$ days [*in vitro*]{} (DIV, days after plating). Half the networks studied are sparsely seeded ($625$ cells/ml) the other half densely seeded ($2,500$ cells/ml). Cells (neurons and support glia) are removed from embryonic mice and their existing tissue structure is dissociated mechanically and enzymatically. After cells are seeded on a microelectrode array, they form new connections and self-organize into spontaneously active networks [@grossbook_1994], [@maeda_1995]. Figure \[fig:f1\] depicts a representative dense and sparse network at $1$, $15$ and $32$ DIV. Experiments and data collection were performed by Wagenaar et al. [@wagenaar_2006] and made publicly available for analysis. Detailed information about culturing, plating, and feeding techniques are provided in this reference. ![image](Fig1.eps){width=".95\textwidth"} First response model {#frmodel} -------------------- We build a model of network connectivity using spike timing correlations to study the evolution of pairwise neural interactions during development. The model is based on the assumption that, within a small network, all activity preceding a spike within a biologically plausible time window contributes to its firing. Any given spike is assumed to be correlated with the ignition of the next spike to fire within $1-10$ ms. This temporal window is selected such that the first spike has enough time to influence the production of the second, but not so long that its effect will have faded [@nakanishi_1998], [@bi_1998]. All spikes collected at a single electrode form a single spike train. Therefore, the number of spike trains is equal to the number of active channels (Table \[table:t1\]) and functional connections discussed here are made between electrodes. Network Electrodes Recording days --------- ------------ ----------------------------------------------- D1 56 4-7,9-26,28,31-35 D2 56 4-26,28,31-35 D3 56 4-26,28,31-35 D4 56 4-26,28,31-35 S1 58 4-7,10,11,13,14,17,19,21,22,24-26,28,31,33,34 S2 58 4-7,10,11,13,14,17,19,21,22,24-26,28,31,33,34 S3 58 4-7,10,11,13,14,17,19,21,22,24-26,28,31,33,34 S4 58 4-8,10,12-14,17,18,20,21,31-35 : Network: D1-D4 are densely seeded and S1-S4 are sparsely seeded. Electrodes: number of electrodes with action potential activity. Recording days: days [*in vitro*]{} a network was recorded. \[table:t1\] Recent analysis of similar recordings in which individual neurons at an electrode were discriminated (spike sorted) shows that, for plating densities comparable to those in the data presented here, typically no more than 2 neurons are observed at any given electrode [@ham_2008]. The first response model provides a method for estimating whether a spike produced at one electrode initiates a spike at another. Each time a spike at an electrode is the first to fire within $1-10$ ms after a spike at another electrode (Fig. \[fig:f2\]a), the link (edge) from the first to the second is incrementally increased. This results in a weighted and directed edge between electrodes (Fig. \[fig:f2\]b). ![a) Activation links are made when activity at an electrode is the first to occur within a $1-10$ ms window after another electrode. Note that no link is made from the second spike in $X$ since the time window between it and $Z$ is less than $1$ ms and it and $Y$ is greater than $10$ ms. b.) Pairwise activation graph constructed from data in part a. Edges are directed and weights increase as connections are observed. Edges are normalized by the lesser of the total spikes observed at either of the connected electrodes. Activation networks represent activation flow and are not representative of the physical connections between two electrodes \[fig:f2\]](Fig2.eps){width=".45\textwidth"} Edge weights are calculated for all electrode pairs. Edges represent inferred activation links in the network and are not necessarily representative of actual synaptic connection between neurons recorded at any two electrodes. Edges are denoted by $x_{ij}$ , where $i$ is the first electrode to fire and $j$ the second. When pairwise edges are studied individually, they are normalized by the lesser of the total spikes recorded at $i$ or $j$, giving an estimate of the probability that $i$ leads $j$. Normalized frequency edges $x_{ij}$ and $x_{ji}$ are denoted $X_n$ and $Y_n$ respectively. Activation graph {#acgraph} ---------------- The combination of all weighted directed edges between electrodes forms an activation graph, which is a weighted directed graph, with each edge normalized between zero and unity. This graph estimates the activation probability between all electrodes. New activation graphs are created for each network on each recording day. Information ensembles of order 2 {#io2} -------------------------------- Neuronal activation along pathways in the brain is generally a sequential process. Activity in one area can ignite activity in another on large scales (tissue structure, e.g. [@spitsyna_2006]) as well as small scales (neuronal networks, e.g. [@ham_2008]). To characterize the influence a target electrode has on all other electrodes, we apply a link entropy analysis of the activation graphs. Link entropy is calculated for each electrode and measures the uncertainty of which electrodes will be activated by the target electrode. For example, in Fig. \[fig:f2\]b, electrode $X$ activates only $Z$. Since there is no uncertainty about which electrode $X$ activates, the link entropy is $0$. If $X$ activated more than one other electrode, uncertainty in activation and link entropy increase. In a fully connected network with $N$ electrodes, there are $N-1$ potential pathways (edges) originating from each electrode. Edges ($x_{ij}$) originating from a target electrode ($i$) in the weighted directed activation graph are used to estimate the probability $p(x_{ij})$: ![ a) Synergetic relationship where $X$ and $Y$ together provide more information about the state of $Z$ than they do individually, as in the case of a logic gate. b) Redundant relationship, where $X$ and $Y$ together provide less information about the state of $Z$ than they do individually, as in the case of a Markov chain \[fig:f3\]](Fig3.eps){width=".45\textwidth"} $$p(x_{ij})=\frac{x_{ij}}{\sum_{j=1}^{N-1} x_{ij}}$$ The link entropy of the target electrode ($i$) in an activation graph with N active electrodes is: $$H_i = -\sum_{j=1}^{N-1} p(x_{ij}) log_2(p(x_{ij}))$$ This measure represents the uncertainty (in bits) of the potential activation paths originating from a target electrode. Link entropy values range from $0$ (no uncertainty; activation flows along only one path with probability of $100\%$) to a maximal value of $log_2(N-1)$. The latter occurs only when (non-zero) probabilistic connections between the target and all other $N-1$ electrodes in the network are equal. Informational ensembles of order $3$ {#order 3} ------------------------------------ Information theoretic measures applied to ensembles of $3$ electrodes are used to estimate functional relationships, beyond simple directed connectivity, in developing networks. At this point, probabilities are no longer based on activation graph links between electrodes. Rather, each electrode can assume two states – spiking or not spiking. In the convention established in Reike et al. [@reike_1997], see also [@bcourt_2007], information theoretic quantities are measured only during network bursts, where the vast majority of all coordinated interactions occur. More details about network burst detection are given in [@ham_2008]. The average number of detected network bursts per minute across all dense and sparse networks on all recorded DIV is shown in Fig. \[fig:f4\]a. Bursts are digitized by dividing time into $10$ ms bins. If activity occurs during a bin, the bin is assigned a $1$. Otherwise the bin is assigned a $0$. Probabilities are computed by normalizing the number of spiking and not spiking bins by the total number of bins. Joint probabilities can also be computed in the same way for multiple electrodes. ![image](Fig4.eps){width=".95\textwidth"} After digitization, all groups of spike train triplets are examined for redundant or synergetic relationships per the convention established in [@bcourt_2007], [@schneidman_2003]. Specifically, for three electrodes ${X,Y,Z}$, we consider the conditional mutual information between $X$ and the state of two other electrodes $Y$ and $Z$ ([@cover_1991]): $$I(X;\{Y,Z\})=\sum_{x,y,z} p(x,y,z) log_2 \frac{p(x,y,z)}{p(x)p(y,z)}.$$ Subtracting this quantity from the sum of $I(X;Y)$ and $I(X;Z)$ reveals the informational nature of the ensemble [@cover_1991]. For example, in a synergetic ensemble (see Fig. \[fig:f3\]a), more information is gained when considering $Y$ and $Z$ together rather than separately: $I(X;{Y,Z}) > I(X;Y) + I(X;Z)$. However, in a redundant ensemble (see Fig. \[fig:f3\]b), less information is gained when considering $Y$ and $Z$ together rather than separately: $I(X;{Y,Z}) < I(X;Y) + I(X;Z)$. If $Y$ and $Z$ are independent, $I(X;{Y,Z}) = I(X;Y) + I(X;Z)$. Therefore, we can use the measure $R(X,Y,Z)=I(X;Y) + I(X;Z)-I(X;{Y,Z})$ to characterize the functional informational nature of an ensemble. $R$ can be positive (redundant), negative (synergetic) or zero (independent). To test for statistical significance, experimentally obtained $R$ values were compared to values obtained from a Poisson null model with the same spiking rates as in the experiment [@bcourt_2007]. In this model, experimentally obtained spike trains are used to form a randomly distributed Poisson spike train for each electrode. This eliminates all correlations between spike trains, so that non zero values of $R$ are the result of finite sampling. The Poisson spike train contains the same number of spikes as the experimentally observed spike train. In practice, $R$ values obtained from this model are very small since coincidences are minimal between random Poisson spike trains. The Poisson model establishes a background noise level below which $R$ values were considered to be effectively 0. This null model was used to establish the background noise level below which $R$ values were considered to be effectively 0. The mean Poisson $R$ values over 10 instances of the model were on the order $10^{-5}$ while typical values from the experimental data ranged between $-10^{-3}$ and $10^{-1}$. Results ------- We analyze eight developing networks, four dense ($2.5\pm1.5$ cells/mm$^2$) and four sparse ($0.6\pm0.24$ cells/mm$^2$). For each network on each recording day (\[table:t1\]), new activation graphs are generated by the competitive first response model (see Methods) and directional activation links between pairs are established. Thus, pairs can be represented by the coordinate $[X_n,Y_n]$, where $X_n$ is the estimated probability that electrode $X$ activates $Y$ and $Y_n$ is the estimated probability $Y$ activates $X$ (Fig \[fig:f4\]b). If $X_n$=$Y_n$, the point falls on the line $y=x$ (solid line, Fig. \[fig:f4\]b). Each coordinate has a magnitude (distance from zero) and a deviation angle from the line $y=x$. If $X_n$ = $Y_n$ , then the deviation angle = $0$ and the magnitude = $\sqrt{X^2+Y^2}$ . In figures \[fig:f4\]c & \[fig:f4\]d we examine the magnitude and deviation of all non-zero ($X_n \ne 0$ & $Y_n \ne 0$) coordinates in all networks. In figure \[fig:f4\]c, the average coordinate magnitude for dense and sparse networks on each recorded DIV is plotted. At early ages, the average coordinate magnitude in dense networks is greater than in their sparse counterparts. This indicates that pairwise connections form faster in dense networks. However, in sparse networks, the average coordinate magnitude stabilizes at (slightly) higher values than in dense networks (Fig. \[fig:f4\]c). In Fig \[fig:f4\]d, we examine the average deviation from the line $y=x$ in dense and sparse networks, for all active pairs. Pairwise activation is relatively balanced during dense network maturation ($X_n\approx Y_n$), as demonstrated by low deviation angles. In sparse networks, deviation angles decrease, but remain larger than dense deviation angles throughout maturation. Wagenaar et al. [@wagenaar_2006] observed that sparse networks are slower to develop network bursting than their dense counterparts (Fig. \[fig:f4\]a). Additionally, activation graphs reveal that late onset bursting, a function of pairwise interactions [@ham_2008], is highly correlated to the slow development of pairwise activation connections. Therefore, early development of pairwise interactions in dense networks appears to be related to the onset of multiple network bursts per minute around $10$ DIV. Note that multiple bursts per minute are not a present in sparse networks until about $22$ DIV (Fig \[fig:f4\]a). link entropy {#link entropy} ------------ Figure \[fig:f5\] shows the average daily link entropy for all electrodes in each of the eight networks studied. Note that a similar number of active electrodes were observed in all networks (Table \[table:t1\]). As dense networks mature, average individual electrode link entropy values approach the maximum possible value, which indicates nearly uniform probabilistic links between any target electrode and all other active electrodes. In other words, given activation at a particular dense network electrode, the next electrode to activate is almost completely random. Mature sparse networks exhibit midrange link entropy values, indicating that probabilistic connections from a target to all other electrodes are unequally distributed and therefore more predictable. This corresponds to the larger deviation angles observed in sparse networks (Fig. \[fig:f4\]d). Note that sparse network S2 has larger link-entropy values than S1, S3 or S4 though not as large as in the dense networks. ![image](Fig5.eps){width=".95\textwidth"} Informational relationships of order 3 {#ir3} -------------------------------------- Groups of three electrodes (triplets), are the smallest (and most abundant) arrangement capable of providing information about functional connectivity in a network [@bettencourt_2008]. As described in Methods, informational structures are determined by comparing the information from two electrodes together conditional on the third with information gained from the two electrodes separately conditioned on the third. The resulting value ($R$) categorizes activity between electrodes as redundant, synergetic or independent. On each recording day, $R$ values for all unique electrode triplet ensembles are calculated (see Methods). Note that $R(X,Y,Z)=R(X,Z,Y)$. In Figure \[fig:f6\], distributions of $R$ values from each network on each recorded day are shown. Synergetic values ($R<0$) are grouped in bin sizes of $.001$ bits while redundant values ($R>0$) are grouped in bins of size $.01$ bits. The asymmetry in relative sizes of positive and negative $R$ values has been demonstrated previously [@bcourt_2007]. Independent values ($R=0$) account for 35% of all $R$ values and are not shown. As described in Methods, all values are obtained during network burst events to provide a comparison with earlier studies. In all instances, young network bursts ($10-20$ DIV) are dominated by synergetic triplets. However, around $18$ DIV, all networks show a shift towards redundant triplets. By $30$ DIV, matured dense networks are dominated by redundant triplet ensembles. This finding is commensurate with our previous study of a mature $45$ DIV network [@bcourt_2007]. Sparse networks are observed to switch back and forth between primarily synergetic and primarily redundant from $18-25$ DIV. Unlike their dense counterparts, sparse network triplets became primarily synergetic after $25$ DIV. While three out of the four sparse networks remain primarily synergetic (Fig \[fig:f6\], S1, S3, S4), in one network (S2) redundant groups reemerge around $30$ DIV and are present throughout the remainder of the experiment. Note that synergetic triplets still maintain a strong presence during this time period. This is in clear contrast to dense networks where synergetic groups essentially disappear altogether in favor of redundant ensembles. This shift is primarily due to the evolution of individual triplet $R$ values changing from synergetic to redundant. In Figure \[fig:f7\], several randomly selected representative triplets and their $R$ values on different recording days are shown from network D4. We observe that the vast majority of dense network triplets (98%) that are synergetic at a young age become redundant during maturation. At around $21$ DIV, redundant ensembles begin emerging in both dense and sparse networks (Fig. \[fig:f6\]). This correlates to the time period when cortical networks [*in vitro*]{} are generally considered mature [@jimbo_1999] and when Wagenaar et al. [@wagenaar_2006] observed that spike rates stopped increasing and leveled off. Triplet interactions -------------------- ![image](Fig6.eps){width=".95\textwidth"} ![image](Fig7.eps){width=".50\textwidth"} Large link-entropy values and the presence of redundant ensembles do not appear to be directly correlated to one another. However, large values of link entropy seem to be an indicator that redundant ensembles will dominate the informational relationships. It should be noted that in D1, the rise in average link-entropy and the increase in redundant triplets are well correlated, but this is not observed in the other three dense networks where link-entropy maximizes before redundant ensembles fully dominate. Regardless, results suggest that high values of link entropy are necessary for the emergence and dominance of redundant neuronal ensembles. Figure \[fig:f5\] S2 indicates that, with moderate link entropy values, synergetic and redundant groups coexist. Relatively low link entropy, like that seen in S1, S3, and S4, appears indicative of synergetic relationships (Fig. \[fig:f5\]). Discussion {#discuss} ========== In this manuscript, we explored quantitative patterns of neuronal interactions in developing dense and sparse cultured cortical networks. Such networks are formed when dissociated prenatal cortical tissue is plated on microelectrode arrays and new connections are spontaneously formed. Multisite electrophysiological data from these networks provides unique access to the development of cortical cultures. It should be noted that [*in vitro*]{} network formation is guided by similar mechanisms (synaptogenesis; [@herndon_1981]) used by intact nervous systems. Pairwise interactions {#pwi} --------------------- At the most basic level, network formation can be thought of as a pairwise phenomenon where two neurons form synaptic connections with one another through electrical and chemical means. We extrapolate functional pairwise connections through a competitive first response model which is based on the assumption that all previous activity (within a biologically plausible time window) contributes to action potential initiation. This model is used to create directed, weighted activation graphs. Mature activation graphs of dense networks reveal that connections between electrodes are mostly reciprocal ($X_n = Y_n$, Fig \[fig:f4\]d) as indicated by relatively low deviation angles. In contrast, electrodes in sparse networks tend to develop activation connection strengths that are skewed in one direction or the other (larger deviation angle, $X_n \ne Y_n$, Fig. \[fig:f4\]d). Additionally, link entropy in sparse networks is smaller than link entropy in dense networks (S1-S4, Fig. \[fig:f5\]). Together, these results indicate that electrodes show activation pathway biases (low link entropy) in sparse networks, but these biases are not reciprocal between pairs. In dense networks, the opposite appears true; probabilistic activation pathways between electrodes are nearly uniform (large values of link entropy, Fig. \[fig:f5\]) and reciprocal between pairs. In networks processing information, each type of connection structure has advantages and disadvantages. Dense networks appear to have more redundancy (high link entropy) which may allow for higher fault tolerance: if a path via a given electrode becomes unresponsive to activation attempts by another, many other pathways could easily be activated. However, such a system may be detrimental to rapid information processing where fewer strong pathways, like those seen in sparse networks, could help an animal reach a fast decision state with greater precision. The presence of favored activation pathways in sparse networks may indicate that they are better platforms for performing training studies that seek to change network interactions [*in vitro*]{} [@shahaf_2001]. Since there are fewer probabilistic pathways for modification, changes to these connections will likely produce greater differences in existing efferent and afferent activation pathways. Conversely, shutting down or building pathways in dense networks would likely produce a less noticeable change in directional activation patterns as many detours are available. Ensembles of $3$ nodes are the minimum required to form computational groups [@bettencourt_2007]. Sequential chains of redundant cells, shown in Figure \[fig:f3\]b, can be used to relay information, and similar arrangements have been shown to play a role in short-term memory in the brain [@abeles_1991] [@diesmann_1999]. In the synergetic configuration depicted in Fig \[fig:f2\]a, a neuron receives inputs from many other cells in such a way that spiking activity is a nonlinear function of the inputs. Both systems (redundant and synergetic) allow neurons to integrate activity from many sources and process information. We demonstrated that plating density has an important effect on informational relationships. Dense and sparse networks develop similar distributions of synergetic informational structures between $10$ and $20$ DIV (Fig. \[fig:f6\]). However, as dense networks mature, most triplets change from synergetic to redundant functional informational connections (Fig. \[fig:f7\]). By about $30$ DIV very few synergetic triplets remain in dense networks. Conversely, individual triplets in sparse networks remain primarily synergetic from early to late stages of development. These results suggest that sparse networks may be better suited for processing information since synergetic relationships are necessary to perform these tasks [@bcourt_2007]. Of course synergetic relationships also exist in denser networks, but in a smaller relative proportion. They may also occur in larger groups of cells, not analyzed here [@bettencourt_2008]. One mature sparse network, S2 (Fig. \[fig:f6\]), showed a relatively even mix of both synergetic and redundant ensembles during late stages of maturation. This was not seen in any of the other networks. Furthermore, this network exhibited link entropy values lying between the other sparse networks and the dense networks (Fig \[fig:f5\]). These findings suggest a link between activation connections and informational relationships. Specifically, when activation connections are balanced throughout a network (larger link entropy values), redundant ensembles result. However, if activation connection are not quite balanced (mid range link entropy values), but are close, both synergetic and redundant ensembles appear. The observed density dependent functional organization of cortical tissue [*in vitro*]{} raises interesting questions for future research. To elucidate how network density affects connectivity structures, experiments that span a broader range of plating densities are needed. Additionally, it would be interesting to apply these analytical techniques to recordings where a network is monitored continuously during development. Conclusion {#conc} ---------- The development of functional connectivity in neural networks [*in vitro*]{} comprising similar cells is greatly affected by plating density. In a comparison of dense and sparsely prepared networks, we demonstrate that strong pairwise connections occur earlier in dense networks. Link entropy analysis of pairwise activation graphs reveal that nodal connections tend to be biased in sparse and nearly equal in dense networks. Connections between electrode triplets in both dense and sparse networks are primarily synergetic during early stages of development. However, in dense networks, they become primarily redundant by $30$ DIV. Conversely, in sparse networks triplets are primarily synergetic by $30$ DIV. Due to the fact that [*in vitro*]{} cultures have many similarities with [*in vivo*]{} tissue, we believe that the developmental features identified here may also play a role in living organisms. These findings should be applicable to future research seeking to replicate or emulate brain functions and provide useful constraints for experiments of neural function. acknowledgements ================ We are grateful to Daniel Wagenaar, Jerome Pine and Steve Potter for making the data analyzed in this manuscript available to us. This work was supported by the Los Alamos National Laboratory Directed Research and Development (LDRD) 2009006DR synthetic cognition through petascale models of the primate visual cortex project.
{ "pile_set_name": "ArXiv" }
--- abstract: 'The de Gennes free energy functional of an infinite smectic-A liquid crystal at the dual point is shown to be topological and to depend only on the number of screw dislocations and the anisotropy. This result generalizes the existence of a Bogomol’nyi bound to an anisotropic system. The role of the boundary of a finite mesoscopic smectic is to provide a mechanism for the existence of thermodynamically stable screw dislocations. We obtain a closed expression for the corresponding free energy and a relation between the applied twist and the number of screw dislocations.' author: - 'Eric Akkermans, Sankalpa Ghosh and Amos Schtalheim' title: 'Bogomol’nyi bound and screw dislocations in a mesoscopic smectic-A' --- ¶ A successful Êand striking analogy between different phases of a liquid crystal and those of a superconductor has been proposed by de Gennes back in 1972 [@DG72]. It provides an example of how geometric and topological considerations lead to similar behaviours in two different physical systems [@Kamien02]. This analogy has been further used by Renn and Lubensky to predict a new phase called the Twist Grain Boundary (TGB) phase [@RL88], whose existence has been experimentally confirmed [@Good88; @GS90; @KIhn92; @NPGN93; @NPGN98; @RPRYN01; @ASSTS03]. This phase is the liquid crystal analogue of the Abrikosov vortex lattice. It consists of screw dislocations that are topological defects. The TGB phase lies in between the smectic-A (SmA) and the cholesteric phase. In this letter, we show that the de Gennes free energy of a SmA has an integrability point at which it is topological. This leads to the existence of screw dislocation solutions, whose thermodynamic stability is obtained in the mesoscopic limit yet to be specified and for the geometry of an infinitely long cylinder of radius $R$. We characterize the smectic-A phase by the following set of length scales: the transverse and the longitudinal coherence lengths $\xi_{\perp}$ and $\xi_{\parallel}$, the twist penetration depth $\lambda_{T}$, and the wavenumber $q_s=\frac{2\pi}{d}$ of the density modulation of the nematogen molecules along the $z$-axis of the cylinder. The cholesteric pitch length defined by $L_p= \frac{2\pi}{k_0}$ [@RL88] is generally larger than $\xi_{\perp}$. Here $d$ is the separation between smectic layers and $k_0$ is the wavenumber of the Frank director in the cholesteric phase. We first show that when written in a dimensionless Êform, the de Gennes free energy density ${\cal F}$ of an effective two-dimensional system is solely characterized by two dimensionless parameters, namely the twist Ginzburg parameter $\kappa= \frac{\lambda_{T}}{\xi_{\perp}}$ and the anisotropy parameter $\alpha= \sqrt{2} \kappa \xi_{\parallel} q_s \propto \xi_{\parallel}/\xi_{\perp}$. The dual point [@SJ69] is defined by $\kappa = 1/\sqrt{2}$. At this point and for a large system, the de Gennes free energy in the presence of a finite twist is a topological quantity that can be written = 2l(1-\^2), \[inenergy\]where the integer $l$ accounts for the number of screw dislocations. This expression corresponds to the Bogomol’nyi bound [@Bogo77] for the liquid crystal. However, from (\[inenergy\]) it appears that there is no mechanism to select the number $l$ of defects, and the free energy is minimum for $l=0$. A way to thermodynamically stabilize solutions with $l \ne 0$ is to consider a mesoscopic finite size smectic, just like the unconventional vortex patterns obtained in mesoscopic superconductors [@Geim1]. A mesoscopic smectic-A for the geometry of a cylinder of radius $R$ is characterized by $R \ll L_p$. Moreover, for $R > \xi_{\perp}$, we show that for a wide range of parameters $k_0$ and $\alpha$, a mesoscopic SmA Êcan accommodate screw dislocations and the selection mechanism for the equilibrium value of $l \ne 0$ is provided by the boundary conditions imposed on the total twist through the sample, in a way quite similar to the Little-Parks setup in superconducting rings. We first Êconsider some of the features of the SmA phase and the conditions of Êits description through the de Gennes free energy [@DG72]. The order parameter that characterizes the SmA phase is the first harmonic of the density of the nematogen molecules. It can be written as $\P(\bs{r})=|\P| e^{i \Phi}$ where $\Phi= q_s \bs{n} \cdot \bs{r}$. The unit vector $\bs{n}$ is the Frank director that specifies the molecular alignment. For a perfect SmA phase we have $\bs{n}=\{0,0,1\}$. The smectic planes are defined by $z= ld, (l \in \mathbb{Z})$, so that the phase $\Phi=2\pi l$ [@DGP93; @CL; @luben1]. A screw dislocation is a spiral staircase structure of the smectic planes [@luben1; @CL] characterized by a finite winding number. The free energy of a SmA is given by the sum of the Landau ($F_L$) and the Frank ($F_F$) contributions [@DGP93] which are, respectively: F\_[L]{}=r|¶|\^2 + |¶|\^4 + C\_ ||\^2 + C\_ (||\^2 + ||\^2) \[BGLE\]F\_[F]{}= K\_1 ()\^2 + K\_2 (. )\^2 + K\_3( )\^2. The Frank term $F_F$ is the sum of energies associated to bend, twist and splay of $\bs{n}$. For a perfect SmA, the Frank energy vanishes. Hereafter we consider only the twist part in $F_F$. The de Gennes free energy [@DG72; @DGP93; @LBG98] is obtained as the limit of the total free energy $F = F_L + F_F$, when the Frank director $\bs{n}$ slightly deviates from the $z$-axis, namely $n_z \approx 1$ and both $n_{x}$ and $n_{y}$ are much smaller than $n_{z}$. This approximation together with the condition that $R \ll L_p$ defines the mesoscopic limit. The de Gennes free energy is thus given by [@DGP93] F\_[dG]{} &=& r|¶|\^2 + |¶|\^4 + C\_ ||\^2 Ê Ê\ &+ & C\_ |(\_ - iq\_s \_)¶|\^2 + (\_ \_)\^2 Ê\[CVE1\], where $\bs{\nabla}_{\perp}=\{\partial_x,\partial_y\}$. We have introduced the two-dimensional vector $\delta \bs{n}_{\perp}=\bs{n}-\hat{\bs{z}}$. Since ${\bs n}^2=1$, at the lowest order $\delta \bs{n}_{\perp}$ lies in the $x$-$y$ plane. The lengths $\xi_{\perp}$ and $\xi_{\parallel}$ over which the SmA order parameter changes Êrespectively in the $x$-$y$ plane and along the $z$-axis are obtained by comparing the corresponding gradient terms with the first term in the energy functional (3). This sets $\xi_{\perp}=\sqrt{C_{\perp}/ |r|}$ and $\xi_{\parallel}=\sqrt{C_{\parallel} / |r|}$. The twist penetration depth Êis $\lambda_{T} = \sqrt{\frac{g K_2}{C_{\perp} q_s^2 |r|}}$ [@DG72; @RL88]. In the mean field approximation, only the coefficient $r$ depends on temperature [@DG72; @RL88; @LBG98], so that the Ginzburg parameter $\kappa=\frac{\lambda_T}{\xi_{\perp}}$ is temperature independent. By rescaling all the lengths in units of $\sqrt{2} \lambda_T$, namely defining the scaling relations $\frac{r^2}{2g}{\cal F} = F_{dG}$, $\frac{-r}{g}|f|^2 = |\P|^2$, $\sqrt{2}\lambda_T\overline{x}= x$, $\sqrt{2}\lambda_T q_s=\overline{q}_s$ and $\overline{\bs{\nabla}}_{\perp}= \sqrt{2}\lambda_T {\bs{\nabla}}_{\perp}$, the de Gennes free energy (3) can be rewritten, up to a constant, in the dimensionless form &=& \^2(|f|\^2 -1)\^2 + \^2|f|\^2 +\ &+& |(\_ - i\_s \_)f|\^2 + (\_ \_s\_)\^2, \[dlfreeen1\] where both $f$ and $\delta \bs{n}_{\perp}$ do not depend on $z$. The anisotropy parameter $\alpha^2 = 2\kappa^2 q_s^2 \xi_{\parallel}^2 = \overline{q}^2_s C_{\parallel}/C_{\perp}$ contains $|r|$ and it Êis therefore temperature dependent. The expression (\[dlfreeen1\]) is analogous to the dimensionless Ginzburg-Landau free energy density of a two-dimensional superconductor in a magnetic field [@dgsuper; @eric99], except for the additional term $ \alpha^2|f|^2 $. This term is proportional to $1/ d^2$, and it accounts for the interaction between the nematogen density in SmA layers. The equilibrium equations are obtained by minimizing the functional (\[dlfreeen1\]) with respect to $f^*$ and $\delta\bs{n}_{\perp}$. It gives |\_ - iq\_[s]{}\_|\^2 f &=& 2\^2 f(1-|f|\^2) - \^2 f\ \_ (\_ \_s \_)&=&2. \[MA\] The second relation is the liquid crystal analogue of the Maxwell-Ampère equation. It is written in terms of the two-dimensional liquid crystal current density, defined by $\bs{j} = Im(f^{\ast} \overline{\bs{\nabla}}_{\perp} f) - |f|^2 \overline{q}_s\delta\bs{n}_{\perp}$. This current density vanishes for a perfect smectic. Using the phase of the SmA order parameter, it can also be written as $\bs{j}=|f|^2 (\overline{\bs{\nabla}}_{\perp} \Phi - \overline{q}_s \delta\bs{n}_{\perp})$. In a superconductor, the vanishing of the in-plane circulation of the current density around a closed contour $\Gamma$ implies the quantization of the magnetic flux. Similarly, in a SmA the vanishing of the circulation of the two-dimensional current density $\bs{j}$ around a closed contour implies that the smectic planes are lifted by an integer multiple of the layer separation $d$. We define Êthe liquid crystal analogue of the London fluxoid Êby \_ \_.d = Ê \_ ( + \_s \_).d = 2 l, Ê\[fluxoid\]where $l$ is an integer. We now make use of the following identity which holds in two dimensions only |(\_ - i\_[s]{}\_)f|\^2 =|f|\^2 + (\_ )+ |\_ \_s \_||f|\^2.\[BGI\] The scalar operator $\overline{{\cal D}}$ is defined by $\overline{{\cal D}}=\partial_{\overline{x}} + i\partial _{\overline{y}} -i\overline{q}_s(\delta n_x + i \delta n_y)$. We denote by $\Omega$ the cross-sectional area of the cylinder and by $\partial \Omega$ its boundary. We set boundary conditions by imposing that the system behaves as a perfect smectic at large distances, namely $|f| \rightarrow 1$ and $\bs{j} \rightarrow 0$ on $\partial \Omega$. Making use of those boundary conditions and of the identity (\[BGI\]), the free energy (\[dlfreeen1\]) at the dual point $\kappa = 1/ \sqrt{2}$ can be written, up to a constant, as $$\begin{aligned} \int_{\Omega} {\cal F} &=& \int_{\Omega} {1 \over 2} \left( 1 - |f|^2 - \alpha^2 - |\overline{\bs{\nabla}}_{\perp} \times \overline{q}_s \delta \bs{n}_{\perp}| \right)^2 + |\overline{{\cal D}}f|^2 \nonumber \\ &+& (1 - \alpha^2) \oint_{{\partial \Omega}} (\frac{\bs{j}}{|f|^2} + \overline{q}_s \delta \bs{n}_{\perp}).d\overline{\bs{l}}. \label{dggbogo}\end{aligned}$$ The boundary integral results from both the Stokes theorem and the boundary conditions. Making use of the relation (\[fluxoid\]), it is reduced to the constant term $2 \pi l (1 - \alpha^2)$. Then the free energy (\[dggbogo\]) is reduced to the sum of two positive terms. It is minimum when each of them vanishes, so that the minimum free energy is given by the so-called Bogomol’nyi bound (\[inenergy\]) and the corresponding field configurations (\[MA\]) are now solutions of the Bogomol’nyi like equations [@Bogo77] f = 0,    |\_ \_s \_| + \^2 = 1 - |f|\^2. \[bogo\] It is important at this point to notice the role played by the nonlinear term in the free energy (\[dlfreeen1\]) in order to reach the Bogomol’nyi bound. The second equation in (\[bogo\]) expresses the dual relation between the twist and $|f|$. By eliminating the twist between these two equations, we obtain for $|f|$ the nonlinear equation $\overline{\nabla}_{\perp}^2 ln|f|^2 = 2(\alpha^2+ |f|^2 -1)$. For $\alpha=0$, it is a Liouville equation which is known [@Liouville] to admit families of vortex solutions characterized by the winding number $l$ that appears in the free energy (\[inenergy\]). The existence of a Bogomol’nyi bound for the de Gennes free energy of a SmA and the equations (\[bogo\]) are one of the main results of this letter. For an infinite system there is no mechanism to select the number $l$ of screw dislocations. The boundary of a finite Êmesoscopic SmA Êintroduces such a mechanism that determines the value of $l$ corresponding to stable screw dislocation solutions as a function of the twist. In a finite system, the order parameter is not equal to $1$ at the boundary, so that the boundary conditions used to derive (\[bogo\]) are not generally satisfied. Hence, the identification of the boundary integral in (\[dggbogo\]) with the fluxoid (\[fluxoid\]) is no longer possible and the free energy cannot be minimized just by using the Bogomol’nyi equations (\[bogo\]). To fix the boundary conditions for a finite system, Êwe impose that the flux of the twist of the SmA through the cylinder is equal to the flux of a cholesteric in the same geometry. To implement it, we consider first the expression of the director $\bs{n}$ in the cholesteric phase [@RL88; @DGP93] given by $ \bs{n}_{ch} (\bs{x})=(0, sin k_0 x, cos k_0 x)$ and obtained by minimizing the cholesteric free energy $F_{ch}=F_{F} - K_2 k_0 \int (\bs{n} \cdot \bs{\nabla} \times \bs{n})$. The flux of the twist through a cylinder of radius $R$ is given by $\pi R^2 \bs{n}_{ch} \cdot \bs{\nabla} \times \bs{n}_{ch} = \pi R^2 k_0$. And because the total flux of the twist of the SmA is $2 \pi R \delta n_{\perp} (R)$, we have \_ (R)= n\_(R) = = . \[bclq\] This boundary condition disregards edge dislocations [@BKL01; @Kamien97]. For the geometry of a cylinder, the current density $\bs{j}$ has only an azimuthal component. We note that on the boundary, $\overline{\bs{\nabla}}_{\perp}\Phi|_{\partial \Omega} = l / \overline{R}$, so that for l = \_[ch]{} \[zerocurrent\]the current density $j_{\theta}(R) $ is negative. Moreover, since the current density is positive at the core of a screw dislocation, there exists a circle along which the current density vanishes. It allows to split the domain $\Omega$ into two concentric subdomains $\Omega = {\Omega_1} \cup {\Omega_2}$, so that the boundary $\partial \Omega_1$ is the zero current density line. The system can now be divided into a bulk ($\Omega_1$) and an edge region ($\Omega_2$). We neglect the interaction between the screw dislocation and the edge current, so that the total free energy density of the system is now given by ${\cal F}= {\cal F}(\Omega_1) + {\cal F}(\Omega_2)$. Since the boundary $\partial \Omega_1$ is the zero current line, we can approximate ${\cal F}(\Omega_1)$ by the Bogomol’nyi bound (\[inenergy\]) [@eric99]. For the edge free energy, we use expression (\[dlfreeen1\]) integrated over $\Omega_2$. The twist ${\cal T}(\overline{r}) = \overline{\bs{\nabla}}_{\perp} \times \overline{q}_s \delta \bs{n}_{\perp}$ enters the system up to a distance of order 1 in units of $\lambda_{T} \sqrt2$ [@DG72]. Therefore, it decreases from the boundary at $r=R$ with a behaviour well described by () = \_ \_s \_ = [\_[ch]{} - l ]{}e\^[- (- )]{}, \[twistom2\] where $\Phi_{ch}$ is defined by (\[zerocurrent\]). The value at the boundary ${\cal T}(\overline{R})= (\Phi_{ch} - l)/ \overline{R}$ results from the expression of the total twist $ 2 \pi \Phi_{ch} $ as a sum $2 \pi l + \int_{\Omega_2} {\cal T}(\overline{r})$ of the twists associated respectively to the dislocations in the subdomain $\Omega_1$ and to the contribution of the subdomain $\Omega_2$. The twist contribution to the free energy in expression (\[dlfreeen1\]) is thus given by $\int_{\Omega_2} \frac{1}{2}(\overline{\bs{\nabla}}_{\perp} \times \overline{q}_s \delta \bs{n}_{\perp})^2 = (l- \Phi_{ch})^2 / 4 \overline{R}$. Along the same line of arguments, the quantity $\bs{j}/|f|^2 = \overline{\bs{\nabla}}_{\perp} \Phi - \overline{q}_s \delta\bs{n}_{\perp}$ is such that $\int_{\Omega_2} \left( (\overline{\bs{\nabla}}_{\perp} \times \bs{j})/|f|^2 + {\cal T}(\overline{r})\right) =0$. Then, $-\bs{j}/|f|^2$ restricted to the subdomain $\Omega_2$ is also given by (\[twistom2\]). At the dual point we have $\lambda_T = \xi_{\perp}/ \sqrt2$, so that the amplitude of the nematogen density modulation $f$ saturates to its equilibrium value (undistorted SmA) over a distance $\lambda_T \sqrt{2}$. With the help of these observations, we estimate ${\cal F}(\Omega_2)$ using a variational ansatz, namely that the modulus $|f|$ has a constant value $f_0$ over a ring of width $\lambda_T \sqrt{2}$, included in $\Omega_2$, so that the term $\overline{\bs{\nabla}}_{\perp}|f||_{\partial \Omega}=0$ [@eric99]. This leads to an edge energy of the form: \_[\_2]{}[F]{} &=& [(l- \_[ch]{})\^2 4 ]{} +\ &+& ( [(l- \_[ch]{})\^2 \^2]{} f\_0\^2 + (1 - f\_0\^2)\^2 + 2 \^2 f\_0\^2 ) . \[edgeen\] We then minimize with respect to $f_0$, which yields $2(1 - f_0^2) ={(l- \Phi_{ch})^2 \over \overline{R}^2} + 2 \alpha^2$. The corresponding expression for the extremum value of the edge free energy (\[edgeen\]) together with the bulk term (\[inenergy\]) gives for the total free energy F\_[l]{}(\_[ch]{}) &= & 2 Êl(1-\^2) + [ ]{} \[[3 2]{} (l- \_[ch]{})\^2 + 2 \^2\^2\]\ &- & [4 \^3]{} \[(l - \_[ch]{})\^2 + 2 \^2 \^2 \]\^2 . \[r14\] This relation constitutes the second main result of this letter. In the limit of a radius $R \gg \lambda_T$, the quartic term becomes negligible compared to the quadratic one, so that the $\Phi_{ch} $-dependence of the total free energy appears as the envelop of a family of parabolae each indexed by the integer $l$. The system chooses its winding number $l$ in order to minimize the free energy. This provides a relation between the number $l$ of screw dislocations and the twist $\Phi_{ch}$ given by the integer part $l = [\Phi_{ch} - {2 \over 3} \overline{R} (1 - \alpha^2)]$. The entrance of the $(l+1)$-$th$ screw dislocation occurs for a value of the twist, so that $F_{l}=F_{l+1}$. The entrance of the first dislocation is thus given by $\Phi_{ch} = {1 \over 2} + {2 \over 3} \overline{R} ( 1 - \alpha^2)$, Ê which according to (\[zerocurrent\]) gives for the corresponding twist the expression $k_0 = {1 \over q_s R}\left( {1 \over R} + { 4 (1 - \alpha^2) \over 3 \sqrt{2} \lambda_T }\right)$. The highest value $k_0^{(c)}$ of the twist that allows the entrance of screw dislocations corresponds to the critical value at which the system becomes cholesteric. It is given by $k_0^{(c)}=\sqrt{\frac{r^2}{gK_2}}$ [@RL88] and it corresponds to $\Phi_{ch}^{(c)} = \overline{R}^2/\sqrt 2$. Using this expression in (\[zerocurrent\]), Êwe obtain a bound for the largest number of screw dislocations that a mesoscopic sample can accomodate. All these results are summarized in Figure \[fig3\], which presents the behaviour of the free energy $F_l$ as a function of $\Phi_{ch}$. 0.751.0pc In deriving the expression (\[r14\]), we have implicitely assumed that the parameters $\Phi_{ch} \propto k_0$ and $\alpha^2$ can be changed independently. Nevertheless, it has been shown experimentally [@GS90] that both $k_0$ and $\alpha^2$ depend on temperature. However, it has been argued that Ê$k_0$ might also be changed at fixed temperature [@RL88; @KS95]. Since we do not know the exact temperature dependence of $k_0$ and $\alpha^2$, we have presented in Figure \[fig4\] the phase diagram of a mesoscopic ÊSmA in the $\Phi_{ch}$-$\alpha^2$ plane. It exhibits finite areas where screw dislocations are energetically stable. 0.751.0pc In conclusion, we have shown that the de Gennes free energy of a smectic-A is characterized by two dimensionless parameters $\kappa$ and the anisotropy $\alpha$. At the dual point $\kappa = 1 / \sqrt 2$, the free energy is topological and it saturates the Bogomol’nyi bound. The role of the boundary of a finite size mesoscopic smectic is to provide a mechanism for the existence of stable screw dislocations. The Bogomol’nyi equations (\[bogo\]) have been obtained without any specific assumption on the nature of the twist field. Therefore, they share some similarities with those obtained in other situations like mesoscopic superconductors [@eric99], Chern-Simons [@jackiw] and Yang-Mills [@witten] field theories. These analogies can be pushed further towards other anisotropic layered systems like anisotropic High-$T_c$ superconductors or double layer quantum Hall systems [@DLQH]. This research was supported in part by the Israel Academy of Sciences and by the Fund for Promotion of Research at the Technion. [8.]{} P. G. de Gennes, [*Solid State Commun*]{}, [**10**]{}, 753 (1972) R.D.Kamien, Rev. of Mod. Phys. [**74**]{} 953 (2002) S. R. Renn and T. C. Lubensky, Phys. Rev. A [**38**]{}, 2132 (1988) J. Goodby, M. A. Waugh, S. M. Stein, E. Chin, R. Pindak, and J. S. Patel, Nature [**328**]{}, 408 (1987) G. Srajer [*et al.*]{}, Phys. Rev. Lett [**64**]{}, 1545 (1990) K. Ihn [*et al.*]{}, Science [**258**]{}, 275 (1992) L. Navailles, P. Barios and H. Nguyen, Phys. Rev. Lett, [**71**]{} (1993) L. Navailles, B. Pansu, L. Gorre-Talini and H. T. Nguyen, Phys. Rev. Lett [**81**]{}, 4168 (1998) D. S. S. Rao [*et al.*]{}, Phys. Rev. Lett [**87**]{}, 085504 (2001) A. Aodrjan [*et al.*]{}, Phys. Rev. Lett [**90**]{}, 035503 (2003) D. Saint-James, E.J.Thomas and G. Sarma, [*Type II Super conductivity*]{}, chapter 2, Pergamon (1969) E. B. Bogomol’nyi, Sov. J. of Nucl. Phys. [**24**]{}, 449 (1977) A. K. Geim [*et al.*]{}, Nature, [**390**]{}, 259 (1997) and A.K.Geim [*et al.*]{}, Nature, [**396**]{}, 144 (1998) P. M. Chaikin and T. C. Lubensky, [*Principles of Condensed Matter Physics*]{}, chapter 9, Cambridge University Press (1991) T. C. Lubensky, Physica A [**220**]{}, 99 (1995) P. G. de Gennes and J. Prost. [*The Physics of Liquid Crystals*]{}, chapters 3 and 10, Clarendon Press, Oxford (1993) L. Benguigui, Liquid Crystals [**25**]{}, 505 (1998) P. G. de Gennes [*Superconductivity of Metals and Alloys*]{}, Êchapter 6, Addison-Wesley (1994) E. Akkermans and K. Mallick, J. Phys. A [**32**]{}, 7133 (1999) and E. Akkermans, D. M. Gangardt and K. Mallick, Phys. Rev. B [**62**]{}, 12427 (2000) C. Taubes, Comm. Math. Phys. [**72**]{}, 277 (1980) R. D. Kamien and T. R. Powers, Liquid ÊCrystals Ê[**23**]{}, 213 (1997), arXiv:cond-mat/9612169 I. Bluestein, R. D. Kamien and T. C. Lubensky, Phys. Rev. A [**63**]{}, 061702 (2001) S. Kralj and T. J. Sluckin, Liquid Crystals [**18**]{}, 887 (1995) and S. Kralj and S. Zumer, Phys. Rev. E [**54**]{}, 1610 (1996) R. Jackiw and S.Y. Pi, Phys. Rev. Lett. [**64**]{}, 2969 (1990) and Phys. Rev. D [**42**]{}, 3500 (1990) E. Witten, Phys. Rev. Lett. [**38**]{}, 121 (1977) S.M. Girvin and A.H. MacDonald, in “[*Novel quantum liquids in low-dimensional semiconductor structures*]{}”, eds. S.D. Sarma and A. Piczuk (Wiley, New York, 1995)
{ "pile_set_name": "ArXiv" }
Real time, non-zero temperature ($T$), correlation functions of quantum many body systems are directly related to the observables of many experiments. For strongly interacting systems, there are few quantitative results on the relaxation and transport processes that are believed to occur at long times at any $T\neq 0$. Monte Carlo and perturbative methods work best in imaginary time, but the analytic continuation to real time is most dangerous, and often fails, in the low frequency limit. At special conformally invariant points in dimension $d=1$, real time correlations describing relaxation of an order parameter can be computed exactly. Among systems in arbitrary $d$, which are tuned across a quantum critical point by a variable coupling, there is only one for which reliable results are available: the $d=1$ impenetrable Bose gas, whose correlators were determined by a profound and sophisticated inverse scattering analysis [@korepbook]. In this paper, we study real time, $T\neq 0$ correlations of the $d=1$ Ising model in a transverse field. We use a novel semiclassical method to obtain the exact asymptotics of order parameter correlations in the two low $T$ regions on either side of its quantum critical point. Our main new result, given in Eqs. (\[main\_res\_RC\]), (\[main\_res\]) below, is that in these low $T$ regimes, the spin correlation function can be expressed as a product of two factors. One, arising from quantum effects, gives the $T=0$ value of the correlation function, and the other, which comes from a [*classical*]{} theory, describes the effects of temperature. Combined with earlier results [@perk; @ssising], our new results give a complete description of time-dependent correlations in all the distinct limiting regions of the Ising model, and exhibit, simply and clearly, the crossovers in the roles of quantum and thermal fluctuations in the relaxational dynamics. For the Ising chain with only nearest neighbor exchange, we also use the free fermion representation [@lsm; @mccoy] to obtain accurate, numerical data [@yr; @snm] for real time spin correlations for systems with up to 512 spins: these results are in excellent agreement with the asymptotic theoretical results even at relatively short times and distances. We also use our semiclassical method to obtain correlators of the $d=1$ impenetrable Bose gas in a certain low $T$ regime; our results here are in agreement with earlier work [@korepbook], although our approach is much simpler and more physically transparent. Our method should also apply to other $d=1$ quantum models (like the non-linear sigma or sine-Gordon) with an excitation gap. We consider the Hamiltonian $$H_I = -\sum_i \left( \sum_{\ell >0} J_{\ell} \sigma^z_i \sigma^z_{i+\ell} + g \sigma^x_i \right) \label{ham}$$ where $\sigma^z_i$, $\sigma^x_i$ are Pauli matrices on a chain of sites $i$, $J_{\ell}$ $(> 0)$ are short-ranged exchange constants, and $g$ $(>0)$ is the transverse field. The ground state of $H_I$ is expected to have long range order with $N_0 \equiv \langle \sigma^z \rangle \neq 0$ for all $g < g_c$, and a gap to all excitations for $g \neq g_c$ [@lsm], see Fig \[fig1\]. Away from $g=g_c$, the low lying states consist of a stable particle with energy $\epsilon_p$ at momentum $p$, and $n > 1$ particle continua at higher energy. We also have $\epsilon_{-p} = \epsilon_p$, and expect that $\epsilon_p$ has a minimum at $p=0$. Consequently, for small $p$ we parameterize $\epsilon_p = ((m c^2)^2 + c^2 p^2 + {\cal O}(p^4))^{1/2}$, which defines the ‘velocity’ $c>0$ and the ‘mass’ $m$; we choose $m > 0$ ($m < 0$) for $g < g_c$ ($g > g_c$). The critical point at $g=g_c$ is described by a continuum theory with dynamic exponent $z=1$, and correlation length exponent $\nu=1$; hence $c \sim \mbox{constant}$ for $g$ near $g_c$, while $m$ vanishes linearly, $|m| \sim |g-g_c|$, with possibly a different slope on the two sides. The model with $J_{\ell>1} = 0$ is integrable [@lsm], and all parameters are known exactly: in units where $\hbar = k_B = 1$ we have $g_c = J_1$, $N_0 = (1-(g/J_1)^2)^{1/8}$ and $\epsilon_p = 2 ( J_1^2 + g^2 - 2 J_1 g \cos (pa))^{1/2}$, where $a$ is the lattice spacing, so $mc^2=2(J_1-g)$. However, the existence of these stable particles is [*not*]{} a special feature of this integrable point. Indeed, for $g \ll g_c$, a particle has the simple weak-coupling interpretation as the boundary between domains with opposite orientations in $\langle \sigma^z \rangle$. Conversely, for $g \gg g_c$, the ground states has all the spins oriented in the $+x$ direction, and the particle is a $-x$ spin hopping from site to site. We expect that these interpretations remain correct at the lowest energies, as we approach $g_c$ from either side, but not at $g=g_c$. In the above interpretations, the particles are evidently bosons. They have short-range repulsive interactions, which, by an elementary calculation [@korepbook] implies that the two-particle $S$-matrix $S_{pp'}$ approaches $-1$ for $pa, p'a \ll 1$. The integrable model has $S_{pp'} = -1$ for all $p$, $p'$, and this is often used to obtain a free-fermion description after a non-local gauge transformation: we shall not use this transformation in our analytical computations. Our new analytical results are in the two low $T$ regimes on either side of $g_c$, with $T \ll |m c^2 |$. In these regimes [@ssising], the density, $\rho$, of thermally excited particles $ \sim e^{- |mc^2|/T}$, is exponentially small, their mean spacing is much larger than their thermal de Broglie wavelength, $\sim (2mT)^{-1/2}$. These are the standard conditions for [*classical*]{} behavior in which $\rho = \int dp/(2\pi) e^{-\epsilon_p /T}$. We now derive results in the “renormalized classical” (RC) region $g<g_c$, $T \ll mc^2$. Consider the correlation function $$C(x_i ,t) = \mbox{Tr} \left ( e^{-H_I /T} e^{i H_I t} \sigma^z_i e^{-i H_I t} \sigma^z_0 \right)/Z , \label{keldysh}$$ where $x_i = ia$, $Z = \mbox{Tr} e^{-H_I /T}$. This can be evaluated in terms of the trajectories of the dilute gas of classical particles noting that, since the particles physically represent domain walls, $\sigma^z$ changes sign every time a particle goes by. Observe that the classical trajectories remain straight lines across collisions because the momenta before and after the collision are the same in $d=1$. This implies that the trajectories are simply independently distributed straight lines, placed with a uniform density $\rho$ along the $x$ axis, with an inverse slope $ v_p \equiv d \epsilon_p /dp$, and with their momenta chosen with the Boltzmann probability density $e^{-\epsilon_p /T}/\rho$ (Fig \[fig2\]). This heuristic argument, based on the classical picture, can be justified by taking the semi-classical (stationary phase) limit of the the double-time (‘Keldysh’) path integral [@keldysh], in which each collision appears in both the forward and backward paths (generated by the $e^{-iH_I t}$ and the $e^{i H_I t}$ in (\[keldysh\]) respectively) and therefore contributes the factor $|S_{pp'}|^2 = 1$. Computing $C(x,t)$ is now an exercise in classical probabilities. The value of $\sigma^z (0,0) \sigma^z (x,t)$ is the square of the magnetization renormalized by quantum fluctuations ($N_0^2$), times $(-1)$ if the number of trajectories intersecting the dashed line in Fig. \[fig2\] is odd. Consider an Ising system of size $L \gg |x|$, and let it contain $N$ thermally excited particles; then $\rho = N/L$. Let the probability that any given trajectory intersect the dashed line $= q$; then the probability only a given set of $k$ lines will intersect is $=q^k (1-q)^{N-k}$. Summing over all possibilities, we have $$\begin{aligned} C(x,t) &=& N_0^2 \sum_{k=0}^N (-1)^k q^k (1-q)^{N-k} N! /(k! (N-k)!) \nonumber \\ &=& N_0^2 (1-2q)^N \approx N_0^2 e^{-2 q N}. \label{cxt}\end{aligned}$$ The last step holds because, as we shall now compute, $q \ll 1$. First, for $t=0$, as the density of trajectories along the $t=0$ axis is uniform, $q = |x|/L$. For $t \neq 0$, consider first all trajectories with a fixed momentum $p$: they will intersect the dashed line if their intersection with $t=0$ axis is between ordinates $0$ and $x -v_p t$, and so $q = |x - v_p t|/L$. Averaging over all $p$, and inserting in (\[cxt\]), we get one of our main results [@maki]: $$C(x,t) = N_0^2 R(x,t) \qquad\mbox{(RC region)} , \label{main_res_RC}$$ $$\mbox{where}~~R(x, t) = \exp\left( - \int \frac{dp}{\pi}e^{-\epsilon_p/T} \left| x - v_p t \right| \right). \label{rxt}$$ Notice that $R(x,0) = e^{-|x|/\xi}$ and $R(0,t) = e^{-|t|/\tau}$, but the general behavior is more complicated. The correlation length $\xi = 1/(2 \rho)$. Remarkably, we find from (\[rxt\]) that the correlation time, $\tau$, is independent of the functional form of $\epsilon_p$ and depends only on the gap: $\tau = (\pi/2T) e^{mc^2 /T}$. In the $T \rightarrow 0$ scaling limit, in which $\xi = (\pi/2mT)^{1/2} e^{mc^2/T}$, $R(x,t)$ obeys the scaling form $R(x,t) = \phi \left( |x| /\xi, |t|/\tau \right)$, where the scaling function $\phi$ is given by $$\ln \phi(\bar{x}, \bar{t}) = - \bar{x}~\mbox{erf} \left ( {\bar{x} \over\bar{t} \sqrt{\pi} }\right) - \bar{t}e^{-\bar{x}^2/(\pi \bar{t}^2)} .$$ Similar classical scaling forms have been discussed earlier [@ssising], but it was incorrectly conjectured that the scaling functions would be those of the Glauber model [@glauber]; Glauber dynamics does not conserve total energy and momentum, and these conservation laws have played a crucial role in the kinematic constraints on particle collisions. We have also investigated the nearest neighbor model numerically by the mapping to free fermions [@lsm; @mccoy]. We generalized an earlier study [@yr] of equal time properties to dynamical quantities [@snm] for the case of free boundary conditions; details will be presented elsewhere. We find that in the RC regime the imaginary part of $C(x,t)$ is much smaller than its real part, which suggests that the dynamics is indeed classical as argued above. In Fig. \[fig3\] we show data for $g=0.6, T=0.3$ and $x=20$ for a lattice size $L=256$. =2.75in We took $J_1=1$ so $g_c=1$ and $mc^2 = 2(1-g) = 0.8$. For comparison we show the theoretical prediction from Eqs. (\[main\_res\_RC\]) and (\[rxt\]), in which we used the full lattice dispersion relation. The agreement is remarkably good. It is interesting to note that because there is a maximum velocity, $v_{max}$, (on the lattice this is given by $v_{max}= 2 J_1 g$ for $g<g_c$ and $v_{max}= 2 J_1$ for $g > g_c$, whereas in the continuum $v_{max}=c$), Eq. (\[rxt\]) predicts that $C(x,t)$ should be independent of $t$ for $|t| < |x| / v_{max}$, and the numerical results show this very clearly. The inset, with a much increased vertical scale, gives an idea of how small are the deviations between the theory and numerics. We find that the agreement is also excellent even at $r=0$, for $t>1$. Now we turn to the “quantum disordered” (QD) region $g> g_c$, $m<0$, $T \ll |mc^2|$. The operator $\sigma^z$ flips spins between the $\pm x$ directions, and the large $g$ picture, noted earlier, then suggests that $\sigma^z$ is the sum of a creation and annihilation operator for the particles. As a result, the $T=0$ the spectral density, obtained from the Fourier transform of $C(x,t)$, has a contribution $\sim \delta (\omega - \epsilon_p)$ associated with the stable particle. Higher order corrections in $1/g$, or the form factor expansion [@formfac] on the continuum theory valid close to $g_c$, show that the next contribution to the spectral density is a continuum above the 3 particle threshold. Here we will focus exclusively on how the one particle pole broadens as $T$ becomes non-zero. We define $K(x,t) \equiv C(x,t)$ at $T=0$, and dropping the multi-particle terms, we have $$K(x,t) = \int \frac{dp}{2\pi} D(p) e^{i px - i \epsilon_p t} \label{defk}$$ where $D(p)$ is a form factor. For the general lattice model, $D(p)$ is not known; neglecting the multi-particle terms, $D(p)$ is seen from (\[defk\]) to be the spatial Fourier transform of $C(x,0)$ [*i.e.*]{} the structure factor. For the continuum theory, we have $D(p) = {\cal A} c/(2 \epsilon_p)$ where ${\cal A}$ is the dimensionless quasiparticle amplitude [@formfac] (for an underlying integrable lattice model ${\cal A} = 2 (|mc^2| / J_1)^{1/4}$), whence $K(x,t) = {\cal A} K_0 (mc(x^2 - c^2 t^2)^{1/2})/(2\pi)$, with $K_0$ the modified Bessel function. Now we consider $T\neq 0$ in the semiclassical approximation. A typical set of paths contributing to the Keldysh path integral is still given by Fig \[fig2\], but its physical interpretation is now very different. The dashed line now represents the trajectory of a particle created at $(0,0)$ and annihilated at $(x,t)$, and $\pm$ signs in the domains should be ignored. In the absence of any other particles this dashed line would contribute $K(x,t)$ to $C(x,t)$. The scattering off the background particles (the full lines in Fig \[fig2\]) introduces factors of the $S$-matrix element $S_{pp'}$; as the dashed line only propagates forward in time, the $S$-matrix elements for collisions between the dashed and full lines (and [*only*]{} these) are not neutralized by a complex conjugate partner. Using the low momentum value $S_{pp'} = -1$, we see that the contribution to $C(x,t)$ equals $(-1)^{n_{\ell}} K(x,t)$ where $n_{\ell}$ is the number of full lines intersecting the dashed line. The $(-1)^{n_{\ell}}$ is precisely the term that appeared in the RC region, although for very different reasons. We can carry out the averaging over all trajectories as before, and obtain our main result $$C(x,t) = K(x,t) R(x,t) \qquad \mbox{(QD region)} , \label{main_res}$$ where $K(x,t)$ is given by Eq. (\[defk\]) and $R(x, t)$ by Eq. (\[rxt\]). While (\[defk\]) is valid for all $x, t$, relaxation due to classical particles makes sense only within the ‘light cone’, and so strictly speaking, (\[main\_res\]) requires $ct \gg x$. Outside the light cone, $K(x,t)$ decays exponentially to zero on the short length scale $\sim 1/mc$ at which $R\approx 1$ and $T$-dependent effects are not expected to be large: so it is reasonable to use (\[main\_res\]) except, perhaps, for $x,t$ extremely small. The result (\[main\_res\]) clearly displays the separation in scales at which quantum and thermal effects act. Quantum fluctuations determine the oscillatory, complex function $K(x,t)$, which gives the $T=0$ value of $C(x,t)$. Exponential relaxation of spin correlations occurs at longer scales $\sim \xi, \tau$, and is controlled by the classical motion of particles and a purely real relaxation function $R(x,t)$. In Fig. \[fig4\] we compare the predictions of Eq. (\[main\_res\]) with numerical results on a lattice of size $L=512$. =2.75in The theoretical curve was determined from the continuum expression for $K(x, t)$, but the full lattice form for $\epsilon_p$ was used in Eq. (\[rxt\]) to determine $R(x, t)$. With the parameters used, $|mc^2| = 0.2$, which is greater than $T$ $(=0.1)$ as required to be in the QD region. The theory predicts that $\mbox{Im} C(x, t) = 0$ for $|t| < |x| / v_{max}$, with a singularity in both real and imaginary parts at $ |t| = |x| / v_{max}$. At longer time both parts oscillate and decay. Part of this decay comes from the $1/\sqrt{t}$ decay of the Bessel functions, but for $|t| > \tau$ this is dominated by the exponential decay from $R(x, t)$. The theory agrees well with the numerics; some differences are visible for small $x$, outside the light cone, but this is outside the domain of validity of (\[main\_res\]). Precisely the same semiclassical arguments can also be applied to the $d=1$ impenetrable gas of bosons of mass $m_B$, in a chemical potential $\mu < 0$, with $T \ll |\mu|$. This system has $S$-matrix elements $S_{pp'} = -1$, and its single particle propagator will be given by (\[main\_res\],\[rxt\],\[defk\]), with $\epsilon_p = -\mu + p^2 /2 m_B$ and $D(p) = 1$. This propagator has also been computed by the inverse scattering method [@korepbook], and the $T \ll |\mu|$ limit of their result agrees with (\[main\_res\]). Turning to the “quantum critical” region $|mc^2 |\ll T \ll J_1,g$, the dynamics is now given by finite $T$ correlators of the $g=g_c$ critical point. An explicit expression for $C(x,t)$ was given earlier [@perk; @ssising], and has exponential decay at a single time scale $\sim 1/T$ in both its real and imaginary parts. There is no clear separation between the contributions of thermal and quantum fluctuations since both are effective at this scale. Hence there can be no effective classical model describing relaxation [@CSY]. For completeness, we also note the non-universal lattice high $T$ region $T \gg J_1,g$; time-dependent correlators at $T=\infty$ were obtained earlier [@perk] and show a Gaussian decay in time. Before concluding, we emphasize that although we compared the analytical results with numerical data on an integrable model, the main result does not depend on integrability. The simplification of the integrable model is that the parameters in the theory, $m, c, {\cal A}$ and $N_0$, are known exactly, whereas in general, they would be phenomenological parameters of a low energy theory. The equality of the classical relaxation functions in the RC and QD regions is surely related to the self-duality of the Ising critical theory, and is a special feature of this model. Finally, it is interesting to speculate that Eq. (\[main\_res\]) may also be true for $d>1$ because of the separation of time scales between the quantum and thermal fluctuations. We thank S. Majumdar, K. Damle, and T. Senthil for helpful discussions. This research was supported by NSF Grants No DMR 96–23181 and DMR 94–11964. by V.E. Korepin [*et. al.*]{}, Cambridge University Press, Cambridge (1993), and references therein. H.W. Capel and J.H.H. Perk, Physica [**87A**]{}, 211 (1977); J.H.H. Perk [*et. al.*]{} Physica [**123A**]{}, 1 (1984). S. Sachdev, Nucl. Phys. B [**464**]{}, 576 (1996); A. Leclair [*et. al.*]{} Nucl. Phys. B [**482**]{}, 579 (1996). E. Lieb, T. Schultz, and D. Mattis, Ann. of Phys. [**16**]{}, 406 (1961); P. Pfeuty, Ann. of Phys. [**57**]{}, 79 (1970). B.M. McCoy [*et. al.*]{}, Phys. Rev. A [**4**]{}, 2331 (1971). A.P. Young and H. Rieger, Phys. Rev. B [**53**]{}, 8486 (1996). J. Stolze, A. Nöppert and G. Müller, Phys. Rev. B [**52**]{}, 4319 (1995); H. Asakawa, Physica A [**233**]{}, 39 (1996). J. Rammer and H. Smith, Rev. Mod. Phys. [**58**]{}, 323 (1986). A relaxation function similar to $R$ was also obtained in a phenomenological analysis of related models by exponentiating a short-time expansion which ignored collisions in K. Maki Phys. Rev. B [**24**]{} 335 (1981) and F. Devreux and J.P. Boucher J. de Physique [**48**]{}, 1663 (1987); we thank O. Starykh for pointing this out to us. R.J. Glauber, J. Math. Phys. [**4**]{}, 294 (1963). J.L. Cardy and G. Mussardo, Nucl. Phys. B [**340**]{}, 387 (1990); V.P. Yurov and Al.B. Zamalodchikov, Int. J. Mod. Phys. A [**6**]{}, 3419 (1991). A.V. Chubukov [*et. al.*]{}, Phys. Rev B [**49**]{}, 11919 (1994).
{ "pile_set_name": "ArXiv" }
--- abstract: 'Screened potential, modified by non standard electron cloud distributions responsible for the shielding effect on fusion of reacting nuclei in astrophysical plasmas, is derived. The case of clouds with depleted tails in space coordinates is discussed. The modified screened potential is obtained both from statistical mechanics arguments based on fluctuations of the inverse of the Debye-Hückel radius and from the solution of a Bernoulli equation used in generalized statistical mechanics. Plots and tables useful in evaluating penetration probability at any energy are provided.' author: - 'P. Quarati' - 'A.M. Scarfone' title: | Modified Debye-Hückel Electron Shielding\ and Penetration Factor --- Introduction ============ Nuclear fusion reaction cross sections and rates are sensitive to the screening effect of the electron cloud around reacting nuclei, an effect that has been widely investigated both theoretically and experimentally since the early works of Salpeter [@Salpeter1; @Salpeter2].\ Different situations arise when fusion reactions take place: 1) in laboratory experiments, where a metal or gaseous target of a given element is bombarded by an ionic or charged particle beam, electrons are for the most part bound in atomic orbits and few of them can be considered free; 2) in stellar cores and other space and astrophysical plasmas where ions and nuclei are embedded in an electronic environment made by mainly free electrons; 3) in a deuterated metal and other solid-state matrices where an impinging deuteron beam reacts with implanted deuterons. — In laboratory experiments, penetration through a screened Coulomb potential at center of mass energy $E$ is shown to be equivalent to that of bare nuclei at energy $E+U_{\rm e}$ where $U_{\rm e}=Z_1\,Z_2\,e^2/ R_{\rm a}$ and $R_{\rm a}$ is the atomic radius or the radius of the innermost electrons. $Z_1\,e$ and $Z_2\,e$ are the charges of the two reacting nuclei and $U_{\rm e}$ is usually taken as constant in the evaluation of cross sections and rates at any energy. Very often, at very low energy, fusion cross sections measured in laboratory experiments are higher than the value calculated by means of the usual Debye-Hückel (DH) screening factor. Stopping power of the incoming beam and temperature reached after energy deposition are important quantities for the correct measurement of the cross sections. Modifications of electron distribution can be induced. The screening effect must be evaluated to obtain the correct astrophysical factor at very low energies [@Assenbaum; @Carraro; @Bracci; @Shoppa; @Strieder]. — In astrophysical plasmas, which that can be considered ideal, the free electrons move around the reacting nuclei and occupy a sphere of DH radius $R_{_{\rm DH}}=\sqrt{k\,T/(4\,\pi\,e^2\,n\,Z_\rho)}$ which is taken on the order of $R_{\rm a}$, with $n$ the particle density and $Z_\rho=\sum_i(Z_i^2+Z_i)\,X_i/A_i$, where the sum is over all positive ions and $X_i$ is the mass fraction of nuclei of type $i$. Only with a decreasing radius does the screening effect become important. A screening factor of the rate can be derived when the energy of the Gamow peak $E_{_{\rm G}}>U_{\rm e}$. It is given by the Debye factor $f=\exp( U_{\rm e}/k\,T)$ with, this time, $U_{\rm e}=Z_1\,Z_2\,e^2/R_{_{\rm DH}}$ [@Rolfs; @Ichimaru; @Castellani; @Opher]. Recently, it has been clarified through numerous experimental observations that the velocity distribution function of electrons (and possibly also of ions) in stellar atmospheres, in space and astrophysical plasmas, may deviate from a Maxwell-Boltzmann distribution in the high-energy tail if non-local thermodynamic effects are non-negligible [@Oxenius; @Collins; @Peyraund; @Chevallier]. In stellar atmospheres, atomic processes such as radiative and dielectronic recombination show rates depending on deviations from Maxwellian distributions of electrons [@Maero]. In stellar cores, signals of possible deviations of ion distributions are evident and, although small, should be considered because they are capable of sensibly influencing nuclear fusion reaction rates [@Ferro; @Lissia]. — In deuterated metals or solid-state matter, the strong screening effect has still to be clearly understood and discussed, although a few interesting descriptions have recently been brought forward to reproduce experimental results [@Raiola; @Coraddu1; @Coraddu2; @Coraddu3; @Kim]. The approach we are discussing here is very useful in understanding the fusion rates in this matter which could simulate some high-density astrophysical plasmas. However, this application deserves a separate detailed paper and we will not discuss this case here. An important issue in the shielding of electrostatic potential in plasmas concerns the investigation of non-linear charge screening effects that can induce modifications in the DH potential, usually derived by linearizing the Poisson equation [@Cravens; @Gruzinov]. One of the first studies which, on a microscopic basis, demonstrated deviations from DH factor can be found in [@Johnson]. \[See also [@Shaviv1; @Shaviv2; @Shaviv3; @Shaviv4; @Chitanvis]\]. The different approaches elaborated up to now are based on the assumption that electrons are distributed in space according to a Boltzmann factor. Few authors [@Bryant; @Treuman; @Leubner; @Kim1; @Rubab] have assumed a stationary $k$-Lorentz distribution to describe significant deviations from standard distributions, constructed from experimental distributions, due to the presence of enhanced high-energy tails. In a recent reference [@Rubab], for instance, we can find quotations from many works where such distributions are reported. By analyzing energy profiles the authors derive an effective length smaller than the standard DH radius, depending on the $k$-parameter. This result is obtained by solving a Poisson equation in the linear weak approximation. The consequence must be an increased barrier penetration factor. Our approach differs. We assume that the electron cloud is spatially distributed following a generalized steady state distribution of the $q$-type that reduces to an exponential distribution when the $q$ parameter (also known as the Tsallis entropic parameter) approaches the $q\to1$ limit [@Tsallis1; @Tsallis2]. We refer the reader to Refs. [@Leubner1; @Leubner2; @Burlaga] for a detailed description of Tsallis generalized statistics and some of its applications to astrophysical problems. We then calculate the modified screening potential by considering two different approaches. One concerns the use of a generalized Poisson equation or Bernoulli equation as used by @Tsallis2; the other is based on super-statistics [@Beck1; @Beck2; @Wilk2; @Wilk1] considering fluctuations of an intensive parameter (the inverse of the DH radius). This implies fluctuation in temperature and density of the components of the plasma. Once we obtain the modified potential, we calculate penetration probability through that potential. In this paper we limit ourselves to the value range $0<q<1$ (distribution in spatial coordinates with depleted tail cut at about $\langle1/R_{_{\rm DH}}\rangle$), leaving the case $q>1$ that describes a distribution in space coordinates with enhanced long-distance. In the standard DH shielding approach, after linearization of the Poisson equation, the electrostatic screened potential behaves like $V_{_{\rm DH}}\sim r^{-1}\,\exp(-r/R_{_{\rm DH}})$ where $r$ is the coordinate with which to evaluate the DH potential. The assumptions made to derive the above relation are, among others things, that [@Cravens; @Bellan]: plasmas are collisionless; the induced perturbation is slow and depends slowly on time (slowness); only electrostatic fields are present while induced fields are negligible; temperature is spatially uniform and the plasma remains in equilibrium during a perturbation; a temperature can always be defined; the number of particles inside the DH sphere is large and therefore fluctuations are small; although ions and electrons have a random thermal motion, perturbations induced around the equilibrium, responsible for small spatial variations of electrostatic potential, can be neglected. Of course, if the real situation differs from the one imposed by one or more of the assumptions reported above, the use of DH potential may induce errors in the evaluation of the penetration factor and nuclear reaction rates.\ Deviations from the conditions imposed by the assumptions are taken into account in this work by choosing the inverse of DH length $1/R_{_{\rm DH}}$ as a fluctuating parameter.\ The electrostatic quantity $r\,V(r)$ is asymptotically given, in this case, by a power law instead of an exponential law because it must satisfy a differential equation, the Bernoulli equation (or a special case of it), that has power law functions as solutions.\ We can derive the modified DH potential $V_ q(r)$ and, a posteriori, the charge distribution $\rho_q$ as an asymptotically power law function. Also in the standard DH approach also two equations are needed, one coming from electromagnetism (Poisson equation), the other from statistical mechanics (Boltzmann factor).\ Penetration factor $\Gamma(E)$ can be calculated by means of the WKB approach using the modified DH potential. We obtain results that differ from the ones calculated with standard DH potential and that will be useful in the interpretation of experimental results on atomic and nuclear rates in several astrophysical systems and processes. We also derive equivalent energy $U_q$ that we give by an interpolating analytical expression, a plot and a tabulation. Energy $U_q$ results a function of the variable $D/E$, where $D$ is defined by $D=Z_1\,Z_2\,e^2\,\langle1/R_{_{\rm DH}}\rangle$. It is easy to observe that $U_q$ depends, for a wide range of values of $D/E$, on $1/\sqrt{k\,T}$ and $\sqrt{n}$. In Section 2, we explain how to derive the modified potential on the basis of super-statistics arguments with the inverse DH length as a fluctuating parameter; we then introduce a nonlinear differential equation, to be associated with the Poisson equation, whose solution coincides with the potential derived directly from the super-statistics approach.\ In Section 3, we derive the penetration factor that can be used for evaluation of nuclear fusion rates in astrophysical plasmas and indicate the range of validity of approximations adopted. In Section 4, we discuss some representative examples whilst, in Section 5, we report our conclusions. Modified Debye-Hückel potential =============================== Combining the Gauss law and the relation that links the electrostatic field to the electric potential $V(r)$ of a point test unitary charge at the origin in a vacuum, from the Poisson equation one obtains the pure Coulomb potential. After a sufficiently long time, electrons and ions rearrange themselves as a response to the forces on them. Ion density eventually remains uniform while electron density near the test charge increases. At the new thermal equilibrium, the distribution of electrons in an electrostatic field is assumed to be given by the well-known Boltzmann factor. Assuming the Boltzmann factor for all the particles, after linearization and using neutrality condition, the Poisson equation can be written as $${1\over r}\,{d^2\over d r^2}\Big(r\,V(r)\Big)={1\over R_{_{\rm DH}}^2}\,V(r) \ ,\label{poi}$$ with solution given by the DH potential and the charge density $\rho_{_{\rm DH}}$ expressed as $$\rho_{_{\rm DH}}\sim-{1\over r\,R_{_{\rm DH}}}\,\exp\left(-{r\over R_{_{\rm DH}}}\right) \ .$$ When one or more of the linearity constraints are violated or relaxed, a different description of the screening is required. If we assume that nonlinear effects produce fluctuations on the inverse DH radius, by following the approach usually developed by super-statistics for inverse temperature $\beta=1/k\,T$ [@Beck1; @Beck2], we can describe the plasma around the test charge as made of cells where $R_{_{\rm DH}}$ is approximately constant and the system can be described by ordinary statistical mechanics, in this case by the exponential (Boltzmann) factor $\exp(-r/R_{_{\rm DH}})$. In the long term run the system is described by a spatial average over the mean of fluctuating quantity $1/R_{_{\rm DH}}$. Fluctuation of the inverse DH radius also means fluctuation of the plasma parameter given by $$\gamma={1\over n\,R_{_{\rm DH}}^3}=\left(4\,\pi\,e^2\,Z_\rho\over k\,T \right)^{3/2}\,\sqrt{n} \ .$$ With few changes, we follow the approach by Wilk and W[ł]{}odarczyk for the case of distributions with depleted tails ($q<1$) [@Wilk1; @Wilk2]. Here we focus our attention on $q<1$ distribution because this shows a depleted tail with a cut-off, that is, the spatial distribution we assume for the electrons.\ We assume that a certain variable $r$ of the system is limited between $0$ and $[(1-q)\,\lambda_0]^{-1}$ where $\lambda_0$ is a constant parameter.\ We define the function $${\cal F}_{q<1}(r,\,\lambda_0)=C_q\,\int\limits_0\limits^\infty f_{q<1} \left(r,\,\lambda;\,\lambda_0\right)\,\exp\left(-\lambda\,r\right) \,d\lambda \ ,$$ where $C_q$ is a normalization factor and $f_{q<1}(r,\,\lambda;\,\lambda_0)$ is the probability density to observe a certain value $\lambda$ of the system which is spread around the value $\lambda_0$. The expression we choose for $f_{q<1}(r,\,\lambda;\,\lambda_0)$ is a gamma distribution $$f_{q<1}(r,\,\lambda;\,\lambda_0)={A_q(r;\,\lambda_0)^{1\over1-q}\over \Gamma\left({1\over1-q}\right)}\,\lambda ^{{1\over1-q}-1}\,\exp\Big(-{\lambda\,A_q(r;\,\lambda_0)}\Big) \ ,$$ where $$A_q(r;\,\lambda_0)={2-q\over 1-q}\,\lambda_0^{-1}-r \ ,$$ and $q$ is the entropic Tsallis parameter.\ Inserting the function $f_{q<1}(r,\,\lambda;\,\lambda_0)$ into ${\cal F}_{q<1}(r;\,\lambda_0)$ we obtain the normalized power law distribution $${\cal F}_{q<1}(r;\,\lambda_0)=\lambda_0\, \left(1-{1-q\over2-q}\,\lambda_0\,r\right)^{1\over1-q} \ .$$ Average value and variance of $\lambda$ depend on variable $r$ according to $$\overline\lambda={1\over(1-q)\,A_q(r;\,\lambda_0)} \ ,\hspace{10mm} \overline{\lambda^2}={2-q\over\big[(1-q)\,A_q(r;\,\lambda_0)\big]^2} \ ,$$ where $\overline {x(r)}=\int x(r,\,\lambda)\,f_{q<1}(r,\,\lambda;\,\lambda_0)\,d\lambda$ is evaluated by means of distribution $f_{q<1}(r,\,\lambda;\,\lambda_0)$.\ However, the relative variance depends only on $q$ $$\omega={\overline{(\lambda^2)}-(\overline\lambda)^2\over (\overline\lambda)^2}=1-q \ .$$ We remark that quantity $\lambda_0$ coincides with the spatial average of $\overline\lambda$, that is $$\langle\lambda\rangle=\lambda_0 \ ,$$ where $\langle x\rangle=\int\overline{x(r)}\,{\cal F}_{q<1}(r;\,\lambda_0)\,dr$ is evaluated by means of distribution ${\cal F}_{q<1}(r;\,\lambda_0)$ which is the weighted average of the exponential (or Boltzmann-like) factor $\exp(-\lambda\,r) $ with weight equal to $f_{q<1}(r,\,\lambda;\,\lambda_0)$ and coincides with the Laplace transform of $f_{q<1}(r,\,\lambda;\,\lambda_0)$.\ Some special limiting cases are $$f_{q\to1}(r,\,\lambda;\,\lambda_0)=\delta(\lambda-\lambda_0) \ .$$ and $${\cal F}_{q\to1}(r;\,\lambda_0)=\lambda_0\,\exp(-\lambda_0\,r) \ ,$$ the ordinary Boltzmann-like factor, with $\overline{(\lambda^2)}=(\overline\lambda)^2$ and $$f_{q\to0}(r,\,\lambda;\,\lambda_0)=A_0(r;\,\lambda_0)\, \exp\Big(-\lambda\,A_0(r;\,\lambda_0)\Big) \ ,$$ where $A_0(r;\,\lambda_0)=2/\lambda_0-r$ and $${\cal F}_{q\to0}(r;\,\lambda_0)=\lambda_0\,(1-\lambda_0\,r) \ ,$$ a linear function, with $\overline{(\lambda^2)}=2\,(\overline\lambda)^2$.\ Let us now identify functional ${\cal F}_{q<1}(r;\,\lambda_0)$ with quantity $r\,V_q(r)$ and substitute $\lambda$ with $1/R_{_{\rm DH}}$, so that $\lambda_0=\langle1/R_{_{\rm DH}}\rangle$ coincides with the spatial average of the DH radius fluctuation. By posing $$\zeta_q=(2-q)\,\langle1/R_{_{\rm DH}}\rangle^{-1} \ ,$$ a characteristic length of the system under inspection which reduces to $R_{_{\rm DH}}$ in the $q\to1$ limit, we obtain the following expression $$V_q(r)={1\over r}\,\left(1-(1-q)\,{r\over\zeta_q}\right)^{1\over1-q} \ ,\label{qpot}$$ which, for $q\to1$ reduces to the standard DH potential. We also have charge distribution $$\rho_q(r)\sim-{1\over (2-q)\,r\,\zeta_q^2}\,\left(1-(1-q)\,{r\over\zeta_q}\right)^{1\over1-q} \ ,\label{qcharge}$$ which for $q\to1$ reduces to $\rho_{_{\rm DH}}$, the charge distribution of the DH approximation.\ Therefore, by considering $1/R_{_{\rm DH}}$ subject to fluctuations described by a gamma distribution, quantity $r\,V(r)$ related to the potential energy barrier is modified from the exponential DH expression to a power-like law, typical of generalized $q<1$ distribution. When we can establish the functional relation between density $n$ and temperature $k\,T$, as, for instance in a solar-like star, where quantity $n/(k\,T)^3$ is constant along the star profile [@Ricci], by means of the relative variance we can establish a link between inverse DH radius and temperature fluctuations and parameter $q$ through $${\Delta(1/R_{_{\rm DH}})\over1/R_{_{\rm DH}}}= {\Delta\sqrt{n/k\,T}\over \sqrt{n/k\,T}}={\Delta(k\,T)\over k\,T}=(1-q)^{1/2} \ .$$ Diffusion of matter between layers with different temperatures induces local temperature fluctuations and density perturbations; fluctuations around an equilibrium or a steady-state matter profile induce fluctuations of quantity $1/R_{_{\rm DH}}$ in the DH sphere, particularly in the regions where the number of particles inside the DH sphere is small. Density and temperature fluctuations do not alter the macroscopic plasma parameters and must agree with the requirements of the constraints imposed by the macroscopic observations. In the solar interior, @Gruzinov have evaluated that the DH radius is about $2\cdot10^{-9}cm$, therefore containing a small number of particles with a corresponding non-negligible particle fluctuation. Let us now justify expressions (\[qpot\]) and (\[qcharge\]) by means of another approach that considers a generalized version of the Poisson equation. In fact, in the standard case, the DH potential can be obtained from the solution of the second order differential equation of the type $${dy\over dr}=a\,y \ , \hspace{10mm} {\rm or} \hspace{10mm} {d^2y\over dr^2}=a^2\,y \ ,$$ with $y\equiv r\,V(r)=\exp(a\,r)$.\ We replace the above linear equation with the following one $${dy\over dr}=a_q\,y^q \ , \hspace{10mm} {\rm or} \hspace{10mm} {d^2y\over dr^2}=q\,a_q^2\,y^{2\,q-1} \ ,$$ with $$y=\exp_q(a_q\,r)\equiv\Big[1-(1-q)\,a_q\,r\Big]^{1\over1-q} \ ,$$ where $q$ is a real number parameter and coincides with the Tsallis parameter. For $q\to1$ we obtain $\exp_1(a\,r)\equiv \exp(a\,r)$.\ To be explicit, for our case we generalize Eq. (\[poi\]) into $${d^2\over dr^2}\Big(r\,V_q(r)\Big)={q\over\zeta_q^2}\,\Big(r\,V_q(r)\Big)^{2\,q-1} \ ,$$ which is a generalized Poisson equation whose solution coincides with Eq. (\[qpot\]).\ We report, for completeness, that by considering the Bernoulli equation introduced by @Tsallis2 $${dy\over dr}=a_1\,y+a_q\,y^q \ ,$$ or $${d^2y\over dr^2}=a_1^2\,y +(1+q)\,a_1\,a_q\,y^q+q\,a_q^2\,y^{2\,q-1} \ ,$$ the solution is $$y=\Bigg[e^{(1-q)\,a_1\,r}+{a_q\over a_1}\,\Big(e^{(1-q)\,a_1\,r}-1\Big)\Bigg]^{1\over1-q} \ .$$ By posing $y=r\,V(r)$ we obtain $$r\,V(r)=\Big[1+(1-q)\,a_q\,r\Big]^{1\over1-q}$$ when $a_1=0$ and $r\,V(r)=\exp(a_1\,r)$ when $a_q=0$. Penetration probability ======================= We calculate penetration probability through the repulsive barrier, one of the terms we need for evaluation of fusion reaction rates, using the WKB approach and following @Bahcall.\ The fusion cross section of two isolated reacting nuclei is written as $$\sigma(E)={S(E)\over E}\,e^{-2\,\pi\,\eta(E)} \ ,$$ where $S(E)$ is the astrophysical factor, $E$ the center of mass energy of the fusing nuclei of charge $Z_1\,e$ and $Z_2\,e$ colliding with relative velocity $v/c=\sqrt{2\,E/(\mu\,c^2)}$, reduced mass $\mu$ and $$\eta(E)={1\over\hbar\,c}\,{Z_1\,Z_2\,e^2\over \sqrt{E}}\,{\sqrt{{1\over2}\,\mu\,c^2}} \ .$$ First of all we define the penetration factor when the pure Coulomb electrostatic potential energy barrier $\widehat V_{_{\rm C}}(r)=Z_1\,Z_2\,e^2\,V_{_{\rm C}}(r)$ occurs, that is, when the reacting nuclei are isolated: $$\begin{aligned} \nonumber \Gamma_{_{\rm C}}(E)&=&e^{-2\,\pi\,\eta(E)}\\ \nonumber &=&\exp\left\{-{2\over\hbar\,c}\int\limits_0 \limits^{r_{\rm c}}\left[2\,\mu\,c^2\,\left(\widehat V_{_{\rm C}}(r)-E\right)\right]^{1/2}\,dr\right\} \ ,\\ &&\label{gc}\end{aligned}$$ where $r_{_{\rm C}}$ is the classical turning point whose value is fixed by the relation $\widehat V_{_{\rm C}}(r_{_{\rm C}})=E$. Secondly, by using the standard DH potential and still taking the turning point $r_{_{\rm DH}}=r_{_{\rm C}}$, a relation valid only for $E>D$, for small values of $r_{_{\rm C}}/R_{_{\rm DH}}$ we have $$\Gamma_{_{\rm DH}}(E)=\exp\left[-\pi\,\left(2+{r_{_{\rm C}}\over R_{_{\rm DH}}}\right)\,\eta(E)\right] \ .\label{gdh}$$ If we consider the rates instead of the cross sections, factor $\exp(-\pi\,\eta\,r_{_{\rm C}}/R_{_{\rm DH}})$ can be evaluated at most probable energy $E_0$ in such a way that the rate can be factorized as the product of $\Gamma_{_{\rm C}}(E)$ times a factor fixed at $E=E_0$. Finally, we consider the deformed DH potential energy barrier $$\widehat V_q(r)={D\over r\,\big\langle{1/R_{_{\rm DH}}}\big\rangle}\,\left[1-(1-q)\,{r\over\zeta_q}\right]^{1\over1-q} \ ,$$ and the penetration factor $$\Gamma_q(E)=\exp\left\{-{2\over\hbar\,c}\int\limits_0\limits^{r_q} \left[2\,\mu\,c^2\,\left(\widehat V_q(r)-E\right)\right]^{1/2}\,dr\right\} \ ,\label{gq}$$ where $r_q$ must be derived from relation $\widehat V_q(r_q)=E$.\ Because we consider $q<1$ potential energy $\widehat V_q(r)$ has a cutoff and, as a consequence, $0<r<\zeta_q/(1-q)$. We write Eq. (\[gq\]) as $$\Gamma_q(E)=\exp\Big(-2\,\pi\,\eta(E)\,\tau_q\Big) \ ,$$ where function $\tau_q$, which depends on quantity $D/E$, goes to one for $D/E\to0$.\ The evaluation of $r_q$ and $\Gamma_q(E)$ can be worked out only numerically although, for small deformation $(q\approx1)$, the penetration factor can be worked out analytically and is given as a product of $\Gamma_{_{\rm C}}(E)$ times a correction factor. However we do not report here the expression for simplicity’s sake. In the case $q=0$, which represents the greatest deformation with respect to the exponential function, $\tau_q$ has the simplest analytical solution $$\tau_{q=0}={E\over E+D} \ .$$ ![Log-linear plot of the quantity $r_q/\zeta_q$ as a function of $D/E$ for several values of $q$ (in the insert the region $10^{-1}<D/E<10$ is expanded).](fig1.eps){width="110mm"} In Figure 1, we report the plot of quantity $r_q/\zeta_q$ as a function of $D/E$ for a few values of $q$ between zero and one.\ In Figure 2, quantity $1-\tau_q$ is plotted as a function of $D/E$.\ Penetration of potential energy barrier $\widehat V_q(r)$ at energy $E$ is equivalent to penetration of the pure Coulomb barrier at an effective energy $E+U_q$ where $$U_q=\left({D\over r_q\,\big\langle{1/R_{_{\rm DH}}}\big\rangle}-E\right) \ ,$$ is a function of $D/E$.\ ![Log-linear plot of the quantity $1-\tau_q$ as a function of $D/E$ for several values of $q$ (in the insert the region $10^{-1}<D/E<10$ is expanded).](fig2.eps){width="110mm"} We have also calculated numerically equivalent energy $U_q$. The behavior of $U_q/D$, as a function of $D/E$, for a few values of $q$ is plotted in Figure 3. Finally, a quantitative comparison of quantities $r_q/\zeta_q$, $U_q/D$ and $1-\tau_q$, corresponding to the values of $q$ depicted in the figures can be obtained from the numerical Table 1.\ ![Log-linear plot, in arbitrary unities, of $U_q/D$ as a function of $D/E$ for several values of $q$.](fig3.eps){width="110mm"} For $q=0$ potential $V_{q=0}$ is cut at $r=2\,\langle1/R_{_{\rm DH}}\rangle^{-1}$ and we have $U_q=D$ for any value of energy $E$. In the other cases of $0<q<1$ we have that: at high energy $E>D$, $U_q$ approaches $D$. At low energy, $E<D$, $U_q$ approaches the value $(1-q)\,D$, meaning that $U_1$ (DH exponential potential) approaches zero. Therefore, when the electron distribution is a deformed or generalized distribution of the $q$-type, with a cut-off at $r=(2-q)/[(1-q)\,\langle1/R_{_{\rm DH}}\rangle]$, the penetration factor is enhanced even at very low energy, except for the case $q=1$. Enhancement depends on $U_q$ which is an energy proportional to $1/\sqrt{k\,T}$ and $\sqrt{n}$.\ [lccc||ccc||ccc||ccc]{} & & $q=0.25$ & & & $q=0.50$ & & & $q=0.75$ & & & $q=0.099$ &\ 0.10 & 0.091 & 0.988 & 0.046 & 0.091 & 0.977 & 0.045 & 0.091 & 0.966 & 0.045 & 0.091 & 0.956 & 0.045\ 0.20 & 0.167 & 0.978 & 0.085 & 0.167 & 0.958 & 0.084 & 0.168 & 0.938 & 0.083 & 0.168 & 0.920 & 0.082\ 0.30 & 0.232 & 0.969 & 0.120 & 0.233 & 0.941 & 0.118 & 0.235 & 0.915 & 0.116 & 0.236 & 0.891 & 0.114\ 0.40 & 0.288 & 0.961 & 0.151 & 0.291 & 0.927 & 0.148 & 0.294 & 0.894 & 0.145 & 0.297 & 0.866 & 0.142\ 0.50 & 0.338 & 0.955 & 0.178 & 0.343 & 0.914 & 0.174 & 0.347 & 0.877 & 0.170 & 0.351 & 0.844 & 0.166\ 0.60 & 0.382 & 0.948 & 0.203 & 0.389 & 0.902 & 0.198 & 0.395 & 0.861 & 0.193 & 0.401 & 0.825 & 0.188\ 0.70 & 0.421 & 0.942 & 0.226 & 0.430 & 0.892 & 0.220 & 0.439 & 0.846 & 0.214 & 0.447 & 0.807 & 0.208\ 0.80 & 0.457 & 0.937 & 0.246 & 0.468 & 0.882 & 0.239 & 0.479 & 0.834 & 0.232 & 0.489 & 0.792 & 0.226\ 0.90 & 0.489 & 0.932 & 0.265 & 0.503 & 0.874 & 0.257 & 0.517 & 0.822 & 0.250 & 0.529 & 0.778 & 0.243\ 1.00 & 0.518 & 0.928 & 0.283 & 0.535 & 0.866 & 0.274 & 0.552 & 0.811 & 0.266 & 0.566 & 0.765 & 0.258\ 2.00 & 0.716 & 0.896 & 0.407 & 0.763 & 0.809 & 0.393 & 0.809 & 0.735 & 0.380 & 0.850 & 0.675 & 0.367\ 3.00 & 0.826 & 0.876 & 0.482 & 0.902 & 0.774 & 0.465 & 0.977 & 0.689 & 0.449 & 1.047 & 0.621 & 0.434\ 4.00 & 0.898 & 0.863 & 0.534 & 1.000 & 0.750 & 0.516 & 1.102 & 0.657 & 0.498 & 1.198 & 0.584 & 0.482\ 5.00 & 0.949 & 0.852 & 0.572 & 1.073 & 0.731 & 0.554 & 1.200 & 0.633 & 0.535 & 1.321 & 0.556 & 0.518\ 6.00 & 0.988 & 0.844 & 0.603 & 1.131 & 0.717 & 0.583 & 1.281 & 0.613 & 0.564 & 1.426 & 0.534 & 0.546\ 7.00 & 1.019 & 0.838 & 0.627 & 1.179 & 0.705 & 0.608 & 1.349 & 0.598 & 0.588 & 1.517 & 0.516 & 0.570\ 8.00 & 1.043 & 0.832 & 0.648 & 1.219 & 0.695 & 0.628 & 1.408 & 0.584 & 0.608 & 1.597 & 0.500 & 0.590\ 9.00 & 1.064 & 0.828 & 0.665 & 1.253 & 0.686 & 0.646 & 1.461 & 0.573 & 0.626 & 1.670 & 0.487 & 0.607\ 10.00 & 1.081 & 0.824 & 0.680 & 1.283 & 0.679 & 0.661 & 1.507 & 0.563 & 0.641 & 1.735 & 0.476 & 0.622\ 15.00 & 1.140 & 0.810 & 0.733 & 1.390 & 0.652 & 0.715 & 1.684 & 0.526 & 0.695 & 1.996 & 0.434 & 0.676\ 20.00 & 1.174 & 0.801 & 0.766 & 1.459 & 0.635 & 0.749 & 1.806 & 0.503 & 0.730 & 2.188 & 0.406 & 0.711\ We have also interpolated the function $U_q/D$ (for values $0<q<0.8$) which is well described by the following analytic function $${U_q\over D}=1-q+q\,\left[1-(1-q)\,a\,\left(D/E\right)^b\right]^{-{c\over1-q}} \ ,\label{interpo}$$ where parameters $a(q)$, $b(q)$ and $c(q)$ are given by $$\begin{aligned} \nonumber &&a(q)=0.484012+1.984961\,q+42.933001\,q^8 \ ,\\ \nonumber &&b(q)=0.958079-0.129717\,q-0.038993\,q^5+\\ &&\hspace{10.5mm}0.804107\,q^6 \ ,\\ \nonumber &&c(q)=0.760938-1.272798\,q+0.493549\,q^2 \ .\end{aligned}$$ The interpolating function allows us to find the value of $U_q(E)$ once we know the temperature and electron density, fixed at $Z_1$, $Z_2$, and energy $E$ for a certain value of $q$. In the range $0<D/E\leq10^2$, for $0.8<q<1$, function (\[interpo\]) still gives a good interpolation of $U_q/D$ but with more complicated relationships of $a(q)$, $b(q)$ and $c(q)$. We omit the details.\ Calculation of nuclear fusion rates (quantities weighted over the reacting nuclei distribution which in many cases is a generalized distribution with a proper ionic parameter $q_{\rm i}$ close to one) requires the insertion into the average integral of the penetration factor that is a function of $E$. In the case of the pure Coulomb barrier the screening factor can be factorized. In our general case this factorization is not possible and the behavior of $U_q$ as a function of $D/E$ must be considered with care, both for non-resonant and resonant reactions. The same consideration is valid if instead of $U_q$ we calculate the rates with the use of the plotted and tabulated function $\tau(E)$. Representative examples of penetration probability ================================================== We report some representative examples on the evaluation of electronic screening factor for non resonant and resonant fusion reactions of interest in solar core and in other dense astrophysical plasmas. Resonant fusions, beside electron screening, are influenced by a resonance screening factor [@Cussons]. Furthermore, Maxwellian rates can be corrected by non standard ionic distributions. In this work, we are interested in modified DH potential and here we limit the discussion to electron screening factor.\ The enhancement of the penetration factor $\Gamma_q(E)$ over the pure Coulomb penetration $\Gamma_{_{\rm C}}(E)$ can be expressed, using Eqs. (\[gc\]) and (\[gq\]), by the ratio $$\begin{aligned} \nonumber &&f_{q,\,_{\rm C}}(E)={\Gamma_q(E)\over\Gamma_{_{\rm C}}(E)}\\ \nonumber &&=\exp\left[-2\,\pi\,{Z_1\,Z_2\,e^2\over\hbar\,c}\,\sqrt{{1\over2}\,\mu\,c^2}\left( {1\over\sqrt{E+U_q}}-{1\over\sqrt{E}}\right)\right] \ ,\\ &&\end{aligned}$$ and the enhancement of $\Gamma_q(E)$ over $\Gamma_{_{\rm DH}}(E)\equiv\Gamma_{q\to1}(E)$ by $$\begin{aligned} \nonumber f_{q,\,_{\rm DH}}(E)&=&{\Gamma_q(E)\over\Gamma_{_{\rm DH}}(E)}\\ &=&\exp\left[-2\,\pi\,\eta(E)\,\Big(\tau_q(E)-\tau_{q\to1}(E)\Big)\right] \ .\end{aligned}$$ It is evident from Figure 2 and from Table 1 that important enhancement over the DH potential comes from the energy range $E\simeq2\cdot10^{-4}\,D\,\div\,D$, with a maximum at about $E\simeq D/15$. The most effective burning energy is at $E_0=(E_{\rm G}\,(k\,T)^2/4)^{2/3}$, where $E_{\rm G}$ is the Gamow energy.\ [lllll]{}$E\,(MeV)$    & $\Gamma_{_{\rm C}}(E)$          &$\Gamma_{_{\rm DH}}(E)$         &$\Gamma_{q=0.50}(E)$      &$\Gamma_{q=0.99}(E)$\ 0.0059 & $8.16\cdot10^{-494}$ & $2.16\cdot10^{-349}$ & $5.20\cdot10^{-285}$ & $1.75\cdot10^{-303}$\ 0.0590 & $1.18\cdot10^{-156}$ & $2.13\cdot10^{-149}$ & $9.70\cdot10^{-141}$ & $2.16\cdot10^{-141}$\ 0.5900 & $4.91\cdot10^{-50}$ & $8.63\cdot10^{-50}$ & $2.00\cdot10^{-49}$ & $1.97\cdot10^{-49}$\ 2.2000 & $2.92\cdot10^{-26}$ & $3.16\cdot10^{-26}$ & $3.56\cdot10^{-26}$ & $3.16\cdot10^{-26}$\ The first example concerns the p $-$ p fusion in solar core. At the mass center energies above $E_0=5.9\cdot10^{-3}\,MeV$, $U_q$ can be taken constant and equal to $D=0.7\cdot10^{-4}\,MeV$. The factor $f_{q,\,_{\rm C}}(E)$ is of the order of few percent above unity at high energies, for ions with energy above $E_0$ and belonging to the distribution tail and having fusion probability greater than those belonging to the head of distribution. However, at these high energies $f_{q,\,_{\rm DH}}(E)$ is practically equal to $1$ also with $q\ll1$. Therefore, the rate is not modified significantly respect to the standard value. At energies below $E_0$ the enhancement is not negligible and $U_q$, which depends on $E$, goes to $(1-q)\,D$ as $E\to0$. We have, for instance, $f_{q,\,_{\rm DH}}(E=D)=1.12$ and $f_{q,\,_{\rm DH}}(E=0.1\,D)=1.65$, when $q=0.5$. However, protons with these energies have a very small probability to fuse and the rate in conclusion cannot change respect to the standard evaluation more than few percent, as also required by luminosity constraints.\ The above discussion is valid for all the other reactions of the hydrogen burning cycle, the effective burning energy being in the range $E_0\simeq15\div30\,KeV$ with $D<E_0$.\ Among the reactions of CNO cycle we consider $^{14}$N (p, $\gamma$) $^{15}$O. The astrophysical factor recently measured [@Runkle; @Imbriani] has important consequences in the evolution of stars, estimation of age of globular clusters and of course in the evaluation of CNO neutrino flux [@Liolios; @Innocenti]. At solar core conditions, we have that in the high energy region $U_q=D\ll E_0=27.0\,KeV$, therefore $\Gamma_q(E)\approx\Gamma_{_{\rm DH}}(E)$. At low energy $U_q\to(1-q)\,D$ and for $E=D/15=0.04\,MeV$ we obtain $\Gamma_q(E)=2.63\cdot10^{-4}$ with $f_{q,\,_{\rm DH}}(E)=3.33$ at $E=0.047\,MeV$ and $q=0.5$. Of course, $f_{q,\,_{\rm DH}}(E)\to1$ in the $q\to1$ limit as we can expect in laboratory experiments.\ Fusion reactions, like $^7$Li (p, $\alpha$) $\alpha$ and $^6$Li (d, $\alpha$) $\alpha$ are of particular interest for their implications in astrophysics. Experimental measurements in laboratory of the screening potential [@Engstler; @Pizzone] have indicated a value of $350\div400\,eV$, greater than the adiabatic theoretical one of $186\,eV$. We should evaluate $D$ at the experimental conditions of target temperature and density. The quantity $U_q$ is a fraction of $D$, decreasing to $(1-q)\,D$ and not to zero as in the standard DH approach. Assuming $q=0.5$ we need $D=800\,eV$ to obtain $U_q\simeq400\,eV$. Unfortunately, we do not know the correct target temperature and density to calculate $D$. This difficulty is a motivation to consider questionable the application of this approach to laboratory experiments.\ We consider now the important non resonant reaction $^{12}$C ($\alpha,\,\gamma$) $^{16}$O at a temperature of $k\,T=17.2\,KeV$ and density $\rho=10^{3.5}\,gr/cm^3$ (as in the core plasma of helium burning red giant stars) with a mass fraction $X($He$)=1/2$ and $X($C$)=1/2$. We have $D=0.573\,KeV$. With $q=0.5$, $U_q$ goes from about $U_q=0.286\,KeV$ at very low energy to $U_q=0.496\,KeV$ at $E=D$ and to $U_q\simeq D$ at high energy. The greatest value of $f_{q,\,_{\rm DH}}(E)$ is at $E\approx4\cdot10^{-2}\,KeV$. At such low energy the screening factor $f_{q,\,_{\rm DH}}(E)\approx10^{55}$, the Coulomb penetration factor being practically zero and $\Gamma_{_{\rm DH}}(E)\approx10^{-458}$.\ Next we consider the example of $^{12}$C $-$ $^{12}$C fusion in laboratory experiments and in massive stars (in the classical thermonuclear regime, without considering, for simplicity’s sake, effects due to the presence of resonances and degeneration [@Cussons; @Itoh; @Ferro1]). In the experimental study of $^{12}$C $-$ $^{12}$C fusion near the Gamow energy, the target temperature is $k\,T\simeq6\cdot10^{-8}\,MeV$ and the graphite density is $\rho\sim1.7\,gr/cm^3$ [@Spillane], therefore we have approximately $R_{_{\rm DH}}\simeq0.31\cdot10^4\,fm$ and $D\simeq0.0168\,MeV$. From [@Assenbaum] we have $U_{\rm e}=5900\,eV$, $R_{\rm a}=0.88\cdot10^4\,fm$, $\Gamma_{_{\rm C}}(E)=\exp(-87.21/\sqrt{E})$ and $\Gamma_{_{\rm DH}}(E)=\exp(-87.21/\sqrt{E+U_{\rm e}})$. We assume $q=0.5$ representing an average deformation of electron distribution. While $U_{\rm e}$ is fixed, $U_q$ depends on energy $E$. Ratios $\Gamma_{_{\rm DH}}(E)/\Gamma_{_{\rm C}}(E)$ coincide with those of Ref. [@Spillane]. In Table 2 we report the values of $\Gamma_{_{\rm C}}(E)$, $\Gamma_{_{\rm DH}}(E)$, $\Gamma_{q=0.50}(E)$ and $\Gamma_{q=0.99}(E)$. Our value $\Gamma_{q\to1}(E)$ differs from $\Gamma_{_{\rm DH}}(E)$ because the first factor is calculated using $U_{q\to1}(E)$ and the second using $U_{\rm e}$ fixed.\ In massive stars [@Gasques], the strong screening effect in a dense plasma can be simulated by assuming $q=0.25$, a value that represents a large electron deformation. At $k\,T=30\cdot10^{-3}\,MeV$ and $\rho\simeq10^9\,gr/cm^3$, we have $R_{_{\rm DH}}\simeq0.5\cdot10^2\,fm$ and $D\simeq1.04\,MeV$. Enhancement of $\Gamma_{q=0.25}(E)$ over $\Gamma_{q=0.99}(E)$ is very evident; at $E=2.2\,MeV$ enhancement is given by a factor of about 2.5. In Table 3 we report, at several energies, $\Gamma_{q=0.25}(E)$ and $\Gamma_{q=0.99}(E)$.\ [lll]{}$E\,(MeV)$       &$\Gamma_{q=0.25}(E)$       &$\Gamma_{q=0.99}(E)$\ 0.0059 & $1.89\cdot10^{-43}$ & $3.07\cdot10^{-82}$\ 0.0590 & $7.50\cdot10^{-41}$ & $2.42\cdot10^{-54}$\ 0.5900 & $2.19\cdot10^{-31}$ & $1.08\cdot10^{-33}$\ 2.2000 & $6.64\cdot10^{-22}$ & $2.62\cdot10^{-22}$\ Finally, in Ia supernova environmental conditions give $k\,T=5\cdot10^{-3}\,MeV$, central density $\rho=3\cdot10^9\,gr/cm^3$ and $X($C$)=1/3$. Therefore, $D=1.5\,MeV$, the greatest deviation is at $E=0.1\,MeV$ and consequently we have at this energy $\Gamma_{_{\rm C}}=7.9\cdot10^{-125}$, $\Gamma_{_{\rm DH}}=6.5\cdot10^{-78}$ and $\Gamma_{q=0.5}=4.5\cdot10^{-73}$ with $f_{q,\,_{\rm DH}}=6.9\cdot10^4$.\ Of course, in all above examples $q$ is an arbitrary parameter whose value should be determined a priori. Rates of fusion reactions will be evaluated using the parametric expression of $U_q(E)$ given in Eq. (\[interpo\]). Conclusions =========== We have shown that in those systems where in the space coordinates stationary electron distributions deviate from the standard exponential one, the shielded electrostatic Coulomb potential is modified with respect to the standard DH potential derived using linear conditions and constraints. We used two different approaches that produced the same results. One consists of associating a Poisson equation having an unknown electron density distribution with a Bernoulli equation for quantity $r\,V_q(r)$ whose solution is asymptotically a power law.\ This result can also be obtained, and justified, by means of the super-statistics approach recently developed within generalized statistical mechanics by considering $1/R_{_{\rm DH}}$, the inverse DH radius, as a fluctuating intensive parameter with relative variance $\omega$ and parameter $q$ given by relation $q=1+\omega$, that characterizes the fluctuation.\ The $q$-modified DH potential has asymptotic power law behavior and we have discussed its meaning for the case $q<1$ which holds when the electrostatic potential has a depleted and cut tail. Fluctuation of the inverse DH radius may be due to temperature and density fluctuation of the electrons surrounding the reacting nuclei in astrophysical plasmas.\ Exact evaluation of the penetration factor has to be carried out numerically. However, in the case of small deformations, one could calculate $\Gamma(E)$ as the pure Coulomb penetration factor $\Gamma_{_{\rm C}}(E)$ times a correction that may sensibly differ from the standard DH correction.\ We have reported detailed plots and numerical tables of several useful quantities that allow evaluation of the penetration factor and the reaction rates which, being weighted integrals, depend sensibly on the behavior of the penetration factor as a function of $E$.\ Penetration factor $\Gamma_q(E)$ can be given as the pure Coulomb penetration factor at an equivalent energy $E+U_q$. Energy $U_q$ is not a constant and can be evaluated only numerically. We have given a useful fit of it. $U_q$ is a function of energy $D$, characteristic of reacting nuclei, their density and fusion temperature, the constant of proportionality depending on $q$ and varying with $D/E$. $U_q$ is proportional to $1/\sqrt{k\,T}$ and to $\sqrt{n}$ as observed experimentally in fusion reactions in metal matrices. The complete and correct expression of the penetration factor or of $U_q$ or $1-\tau_q$ reported in the paper are necessary to evaluate the fusion rates at any energy. In fact, a possible deformed ion distribution together with an electron deformed distribution may sensibly affect the values of the rates. In the evaluations of the rates of fusion reactions, such as in the examples of p $-$ p, $^{14}$N (p, $\gamma$) $^{15}$O, $^7$Li (p, $\alpha$) $\alpha$ and $^6$Li (d, $\alpha$) $\alpha$, $^{12}$C ($\alpha,\,\gamma)$ $^{16}$O and $^{12}$C $-$ $^{12}$C fusion, mentioned in the previous paragraph, important enhancements may occur, in addition to possible enhancements or decrements due to non-Maxwellian energy-momentum ion distributions, if the electron clouds surrounding the reacting nuclei are spatially modified by an assumed $q<1$ distribution. Specific applications of our approach are in progress; they are comprehensive of discussion of the recent experimental results of the LUNA collaboration [@Spillane; @Imbriani; @Gyurky] to derive astrophysical factors at stellar energy range and will be the argument of a further work. We are confident that our approach and numerical tables shown here will be used to study astrophysical process where temperature and density fluctuations cannot be neglected to evaluate DH shielding effect on reaction rates. Assenbaum, H. J., Langanke, K., & Rolfs, C. 1987, Z. Phys. A, 327, 461 Bahcall, J., Chen, X., & Kamionkowski, M. 1998, , 57, 2756 Beck, C. 2001, , 87, 180601 Beck, C. 2004, Continuum Mech. Thermodyn., 16, 293 Bellan, P. M. 2006 Foundations of Plasma Physics (Cambridge University Press, Cambridge) Bracci, L., Fiorentini, G., Melezhik, V. S., Mezzorani, G., & Quarati, P. 1990, , 513, 316 Bryant, D. A. 1996, J. Plasma Phys., 56, 87 Burlaga, L.F., Viñas, A.F., Ness, N.F., & Acuña, M.H. 2006, , 644, L83 Carraro, C., Schafer, A., & Koonin, S. E. 1988, , 331, 565 Castellani, V., Degl’Innocenti, S., Fiorentini, G., Lissia, M., & Ricci, B., 1997, , 281, 309 Chevallier, L. 2006, Thermalisation of electrons in a stellar atmosphere, arXiv:astro-ph/0601459v2 Chitanvis, S. M., , 654, 693 Collins II, G.W. 1989, The fundamentals of stellar astrophysics (Freeman, New York) Coraddu, M., Lissia, M., Mezzorani, G., Petrushevich, Yu. V., Quarati, P., & Starostin, A. N. 2004, Physica A, 340, 490 Coraddu, M., Mezzorani, G., Petrushevich, Yu. V., Quarati, P., & Starostin, A. N. 2004, Physica A, 340, 496 Coraddu, M., Lissia, M., Mezzorani, G., & Quarati, P. 2006, Eur. J. Phys. B, 50, 11 Cravens, T. E. 1997, Physics of Solar System Plasmas (Cambridge University Press, Cambridge) Cussons, R., Langanke, K., & Liolios, T. 2002, Eur. Phys. J. A, 15, 291 Degl’Innocenti, S., et al. 2004, Phys. Lett. B, 590, 13 Engstler, S., et al. 1992, Zeit. Phys. A, 342, 471 Ferro, F., Lavagno, A., & Quarati, P. 2004, Eur. Phys. J. A, 21, 529 Ferro, F., & Quarati, P. 2005, , 71, 026408 Gasques, L. R., et al. 2005, , 72, 025806 Gruzinov, A. V., & Bahcall, J. 1998, , 504, 996 Gyurky, Gy., et al. 2007, $^3$He ($\alpha,\,\gamma$) $^7$Be cross section at low energies, arXiv:nucl-ex/0702003 Ichimaru, S. 1993, Rev. Mod. Phys., 65, 255 Imbriani, G., et al. 2005, Eur. Phys. J. A, 25, 455 Itoh, N., Tomizawa, N., Wanajo, S., & Nozawa, S. 2003, , 586, 1436 Johnson, C. W., Kolbe, E., Koonin, S. E., & Langanke, K. 1992 , 392, 320 Kim, C. -G., & Jung, Y. -D. 2004, Plasma Phys. Contr. Fusion, 46, 1493 Kim, Y. E., & Zubarev, A. L. 2006, Japanese J. Appl. Phys., 45, L552 Leubner, M. P. 2004, , 604, 469 Leubner, M. P. 2005, , 632, L1 Leubner, M.P., & Vörös, Z. 2005, , 618, 547 Liolios, T. E. 2000, , 61, 055802 Lissia, M., & Quarati, P. 2005, Europhysics News, 36, 211 Maero, G., Quarati, P., & Ferro, F. 2006, Eur. Phys. J. B, 50, 23 Opher, M., & Opher, R. 2000, , 535, 473 Oxenius, J. 1986, Kinetic Theory of Particles and Photons, Theoretical foundations of non LTE Plasma Spectroscopy (Sprinter Verlag, Berlin) Peyraud-Cuenca, N. 1992, , 261, 633 Pizzone, R. G., et al. 2003, , 398, 423 Raiola, F., et al. 2004, Eur. Phys. J. A, 19, 283 Ricci, B., Degl’Innocenti, S., & Fiorentini, G. 1995, , 52, 1095 Rolfs, C.E., & Rodney, W.S. 2005, Cauldrons in the Cosmos: Nuclear Astrophysics (University of Chicago Press) Rubab, N., & Murtaza, G. 2006, , 74, 145 Runkle, R. C., et al. 2005, , 94, 082503 Salpeter, E. E. 1953, Australian J. Phys., 7, 373 Salpeter, E. E., & Van Horn, H. M. 1969, , 155, 183 Shaviv, G. 2004, Prog. Theor. Phys. Suppl., 154, 293 Shaviv, G., & Shaviv, N. J. 2000, , 529, 1054 Shaviv, G., & Shaviv, N. J. 2001, , 558, 925 Shaviv, G., & Shaviv, N. J. 2002, AIP Conf. Proceed., 637, 150 Shoppa, T.D., Koonin, S.E., Langanke, K., & Seki, R. 1993, , 48, 837 Spillane, T., et al. 2007, Study of the $^{12}$C - $^{12}$C fusion reactions near the Gamow energy, arXiv:nucl-ex/0702023 Strieder, F., Rolfs, C., Spitaleri, C., & Corvisiero, P. 2001, Naturwissenschaften, 88, 461 Treuman, R. A., Jaroschek, C. H., & Scholer, M. 2004, Phys. Plasmas, 11, 1317 Tsallis, C. 1988, J. Stat. Phys., 52, 479 Tsallis, C., & Borges, E.P. 2003, Nonextensive statistical mechanics - Applications to nuclear and high energy physics, Proceedings of the Xth International Workshop on Multiparticle Production - Correlations and Fluctuations in QCD, Antoniou N., ed. (World Scientific, Singapore); arXiv:cond-mat/0301521 Wilk, G., & W[ł]{}odarczyk, Z. 2001, Chaos, Solitons and Fractals, 13, 581 Wilk, G., & W[ł]{}odarczyk, Z. 2006, Fluctuations, correlations and non-extensivity, arXiv:hep-ph/0610292v3
{ "pile_set_name": "ArXiv" }
--- abstract: | In this work we tackle the reconstruction of discontinuous coefficients in a semilinear elliptic equation from the knowledge of the solution on the boundary of the domain, an inverse problem motivated by biological application in cardiac electrophysiology. We formulate a constraint minimization problem involving a quadratic mismatch functional enhanced with a regularization term which penalizes the perimeter of the inclusion to be identified. We introduce a phase-field relaxation of the problem, replacing the perimeter term with a Ginzburg-Landau-type energy. We prove the $\Gamma$-convergence of the relaxed functional to the original one (which implies the convergence of the minimizers), we compute the optimality conditions of the phase-field problem and define a reconstruction algorithm based on the use of the Frèchet derivative of the functional. After introducing a discrete version of the problem we implement an iterative algorithm and prove convergence properties. Several numerical results are reported, assessing the effectiveness and the robustness of the algorihtm in identifying arbitrarily-shaped inclusions. Finally, we compare our approach to a shape derivative based technique, both from a theoretical point of view (computing the sharp interface limit of the optimality conditions) and from a numerical one. author: - 'Elena Beretta, Luca Ratti,   Marco Verani' date: title: 'A phase-field approach for the interface reconstruction in a nonlinear elliptic problem arising from cardiac electrophysiology' --- Introduction ============ We consider the following Neumann problem, defined over $\Omega \subset \mathbb{R}^2$: $$\left\{ \begin{aligned} -\diverg(\tilde{k}(x) \nabla y) + \chi_{\Omega \setminus \omega}y^3 &= f \qquad \text{in } \Omega \\ \partial_{\normal} y &= 0 \qquad \text{on } \partial \Omega, \end{aligned} \right. \label{eq:probclassic}$$ where $\chi_{\Omega \setminus \omega}$ is the indicator function of $\Omega \setminus \omega$ and $$\tilde{k}(x) = \left\{ \begin{aligned} k & \text{ if } x \in \omega \\ 1 & \text{ if } x \in \Omega \setminus \omega, \end{aligned} \right. \quad k_{in} \neq k_{out},$$ being $ 0 < k \ll 1$ and $f \in L^2(\Omega)$. The boundary value problem consists of a semilinear diffusion-reaction equation with discontinuous coefficients across the interface of an inclusion $\omega \subset \Omega$, in which the conducting properties are different from the background medium. Our goal is the determination of the inclusion from the knowledge of the value of $y$ on the boundary $\partial \Omega$, i.e., given the measured data $y_{meas}$ on the boundary $\partial \Omega$, to find $\omega \subset \Omega$ such that the corresponding solution $y$ of satisfies $$y|_{\partial \Omega} = y_{meas}. \label{eq:invintro}$$ Since at the state of the art very few works tackle similar inverse problems in a nonlinear context, the reconstruction problem to which this work is devoted is particularly interesting from both an analytic and a numerical standpoint. The direct problem can be related to a meaningful application arising in cardiac electrophysiology, up to several simplifications. In that context (see [@book:sundes-lines], [@book:pavarino]), the solution $y$ represents the electric transmembrane potential in the heart tissue, the coefficient $\tilde k$ is the tissue conductivity and the nonlinear reaction term encodes a ionic transmembrane current. An inclusion $\omega$ models the presence of an ischemia, which causes a substantial alteration in the conductivity properties of the tissue. The objective of our work, in the long run, is the identification of ischemic regions through a set of measurements of the electric potential acquired on the surface of the myocardium. Indeed, a map of the potential on the boundary of internal heart cavities can be acquired by means of non-contact electrodes carried by a catheter inside a heart cavity; this is the procedure of the so-called intracardiac electrogram technique, which has become a possible (but invasive) inspection technique for patients showing symptoms of heart failure. We remark that our model is a simplified version of the more complex *monodomain* model (see e.g. [@phd:tung], [@book:sundes-lines]). The monodomain is a continuum model which describes the evolution of the transmembrane potential on the heart tissue according to the conservation law for currents and to a satisfying description of the ionic current, which entails the coupling with a system of ordinary differential equations for the concentration of chemical species. In this preliminary setting, we remove the coupling with the ionic model, adopt instead a phenomenological description of the ionic current, through the introduction of a cubic reaction term. Moreover, we consider the stationary case in presence of a source term which plays the role of the electrical stimulus. Despite the simplifications, the problem we consider in this paper is a mathematical challenge itself. Indeed, here the difficulties include the nonlinearity of both the direct and the inverse problem, as well as the lack of measurements at disposal. In fact, already the linear counterpart of the problem, obtained when the nonlinear reaction term is removed, is strictly related to the *inverse conductivity problem*, also called *Calderón problem*, which has been object of several studies in the last decades. Without additional hypotheses on the geometry of the inclusion, but only assuming a sufficient degree of regularity of the interface, uniqueness from knowledge of infinitely many measurements has been proved in [@art:isakov_discontinuous] and logarithmic-type stability estimates have been derived in [@art:ales]. Finitely many measurements are sufficient to determine uniquenely and in a stable (Lipschitz) way the inclusion introducing additional information either on the shape of the inclusion or on its size, e.g. when the inclusion belongs to a specific class of domains with prescribed shape, such as discs, polygons, spheres, cylinders, polyhedra (see [@art:isa-po], [@book:ammari-kang], [@art:barcelo]) or when the volume of the inclusion is small compared to the volume of the domain (see [@art:fried-vog], [@art:cfmv]). Several reconstruction algorithms has been developed for the solution of the inverse conductivity problem, and it is beyond the purposes of this introduction to provide an exhaustive overview on the topic. Under the assumption that the inclusion to be reconstructed is of small size, we mention the constant current projection algorithm in [@art:amm-seo], the least-squares algorithm proposed in [@art:cfmv], and the linear sampling method in [@art:bhv] for similar problems. Although these algorithms have proved to be effective, they heavily rely on the linearity of the problem. On the contrary, it is possible to overcome the strict dependence on the linearity of the problems by aims of a variational approach, based on the constraint minimization of a quadratic misfit functional, as in [@art:kv1987], [@art:amstutz2005crack] and [@art:ammari2012]. When dealing with the reconstruction of extended inclusions in the linear case, both direct and variational algorithms are available. Among the first ones, we mention [@art:bhv] and [@art:ikehata2000]; instead, from a variational standpoint, a shape-optimization approach to the minimization of the mismatch functional, with suitable regularization, is explored in [@art:kalvk] [@art:hett-run], [@art:afraites] and [@art:abfkl]. In [@art:hintermuller2008] and [@art:cmm], this approach is coupled with topology optimization; whereas the level set technique coupled with shape optimization technique has been applied in [@art:santosa] and in [@art:ito-kunish], [@art:burger2003], [@art:chan-tai], also including a Total Variation regularization of the functional. Recently, total-variation-based schemes have been employed to solve inverse problems: along this line we mention, among the others, the Levenberg-Marquardt and Landweber algorithms in [@art:bb] and the augmented Lagrangian approach in [@art:chen1999], [@book:bartels Chapter 10] and [@art:bartels2012]. Finally, the phase field approach has been explored for the linear inverse conductivity problem e.g in [@art:rondi2001] and recently in [@art:deckelnick], but consists in a novelty for the non-linear problem considered in this paper. Concerning the reconstruction algorithm for inverse problems dealing with non-linear PDEs, we recall some works related to sensitivity analysis for semilinear elliptic problems as [@art:scheid], [@art:amstutzNL], although in different contexts with respect to our application. We remark that the level-set method has been implemented for the reconstruction of extended inclusion in the nonlinear problem of cardiac electrophysiology (see [@art:lyka] and [@chavez2015]), by evaluating the sensitivity of the cost functional with respect to a selected set of parameters involved in the full discretization of the shape of the inclusion. In [@art:BMR] the authors, taking advantage from the results obtained in [@art:bcmp], proposed a reconstruction algorithm for the nonlinear problem based on topological optimization, where a suitable quadratic functional is minimized to detect the position of small inclusions separated from the boundary. In [@art:BCCMR] the results obtained in [@art:BMR] and [@art:bcmp] have been extended to the time-dependent monodomain equation under the same assumptions. Clearly, this type of assumptions on the unknown inclusions are quite restrictive particulary for the application we have in mind. In this paper we propose a reconstruction algorithm of conductivity inclusions of arbitrary shape and position by relying on the minimization of a suitable boundary misfit functional, enhanced with a perimeter penalization term, and, following the approach in [@art:deckelnick], by introducing a relaxed functional obtained by using a suitable phase field approximation, where the discontinuity interface of the inclusion is replaced by a diffuse interface with small thickness expressed in terms of a positive relaxation parameter $\varepsilon$ and the perimeter functional is replaced by the Ginzburg-Landau energy. The outline of the paper is as follows: in Section \[sec:regul\] we introduce and motivate the Total Variation regularization for the optimization problem. In Section \[sec:relax\], after introducing the phase-field regularization of the problem and discussing its well-posedness, we show $\Gamma$-convergence of the relaxed functional to the original one as the relaxation parameter approaches zero. We furthermore derive necessary optimality conditions associated to the relaxed problem, exploiting the Fréchet derivative of the functional. The computational approach proposed in Section \[sec:discr\] is based on a finite element approximation similarly to the one introduced in [@art:deckelnick]. Despite the presence of the nonlinear term in the PDE it is possible to show that the the discretized solution converges to the solution of the phase field problem. We derive an iterative method which is shown to yield an energy decreasing sequence converging to a discrete critical point. The power of this approach is twofold: on one hand it allows to consider conductivity inclusions of arbitrary shape and position which is the case of interest for our application and on the other it leads to remarkable reconstructions as shown in the numerical experiments in Section \[sec:numer\]. Finally, in Section \[sec:shape\] we compare our technique to the shape optimization approach: after showing the optimality conditions derived for the relaxed problem converge to the ones corresponding to the sharp interface one, we show numerical results obtained by applying both the algorithms on the same benchmark cases. Minimization problem and its regularization {#sec:regul} =========================================== In this section, we give a rigorous formulation both of the direct and of the inverse problem in study. The analysis of the well-posedness of the direct problem is reported in details in the Appendix, and consists in an extension of the results previously obtained in [@art:bcmp]. The well-posedness of the inverse problem is analysed in this section: in particular, we formulate an associated constraint minimization problem and investigate the stability of its solution under perturbation of the data, following an approach analogous to [@book:engl Chapter 10], but setting the entire analysis in a non-reflexive Banach space, which entails further complications. The strategy adopted to overcome the instability is the introduction of a Tikhonov regularization, and the properties of the regularized problem are reported and proved in details. We formulate the problems and in terms of the indicator function of the inclusion, $u = \chi_{\omega}$. We assume an *a priori* hypothesis on the inclusion, namely that it is a subset of $\Omega$ of finite perimeter: hence, $u$ belongs to $BV(\Omega)$, i.e. the space of the $L^1(\Omega)$ functions for which the Total Variation is finite, being $$TV(u) = \sup \left\{ \int_\Omega u \text{div}(\phi); \quad \phi \in C^1_0(\Omega;\R^2), \ \norm{L^\infty}{\phi}\leq 1 \right\},$$ endowed with the norm $\norm{BV}{\cdot} = \norm{L^1}{\cdot} + TV(\cdot)$. In particular, $$u \in X_{0,1}= \{ v \in BV(\Omega): v(x) \in \{0,1\} \text{ a.e. in $\Omega$ }\}.$$ The weak formulation of the direct problem in terms of $u$ reads: find $y$ in $H^1(\Omega)$ s.t., $\forall \varphi \in H^1(\Omega)$, $$\int_{\Omega} a(u) \nabla y \nabla \varphi + \int_{\Omega} b(u) y^3 \varphi = \int_{\Omega} f \varphi, \label{eq:prob}$$ being $a(u) = 1-(1-k)u$ and $b(u) = 1-u$. Define $S: X_{0,1} \rightarrow H^1(\Omega)$ the *solution map*: for all $u \in X_{0,1}$, $S(u) = y$ is the solution to problem with indicator function $u$; the inverse problem consists in: $$\textit{ find } u \in X_{0,1} \textit{ s.t. } S(u)|_{\partial \Omega} = y_{meas}. \label{eq:inv}$$ As it is proved in the Appendix, in Proposition \[prop:wellpos\], the solution map $S$ is well defined between the spaces $BV(\Omega;[0,1])$ and $H^1(\Omega)$, thus for each $u \in X_{0,1}$ there exists a unique solution $S(u) \in H^1(\Omega)$. We introduce the following constraint optimization problem: $$\argmin_{u \in X_{0,1}} J(u); \qquad J(u) = \half \norm{L^2(\partial \Omega)}{S(u) - y_{meas}}^2. \label{eq:opt}$$ It is well known that this problem is ill-posed, and in particular instability under the perturbation of the boundary data occurs. A possible way to recover well-posedness for the minimization problem in $BV(\Omega)$ is to introduce a Tikhonov regularization term in the functional to minimize, e.g. a penalization term for the perimeter of the inclusion. The regularized problem reads: $$\argmin_{u \in X_{0,1}} J_{reg}(u); \quad J_{reg}(u) = \half \norm{L^2(\partial\Omega)}{S(u) - y_{meas}}^2 + \alpha TV(u), \label{eq:minreg}$$ Then, we can prove the following properties: - for every $\alpha >0$ there exists at least one solution to ; - the solutions of are stable w.r.t. perturbation of the data $y_{meas}$; - if $\{\alpha_k\}$ is a sequence of penalization parameters suitably chosen, then the sequence of the corresponding minimizers $\{u_k\}$ has a subsequence converging to a minimum-norm solution of . Before proving the listed statements, it is necessary to formulate and prove a continuity result for the solution map with respect to the $L^1$ norm, which consists in an essential property for the following analysis and requires an accurate treatment due to the non-linearity of the direct problem. Let $f \in L^2(\Omega)$ satisfy the hypotheses in Proposition \[prop:geq\] or \[prop:geq2\]. If $\{ u_n \} \subset X_{0,1}$ s.t. $u_n \xrightarrow{L^1} u \in X_{0,1}$, then $S(u_n)|_{\partial \Omega} \xrightarrow{L^2(\partial \Omega)}S(u)|_{\partial \Omega}$. \[prop:continuity\] Define $w_n = S(u_n) - S(u)$; then, subtracting the evaluated in $u_n$ and the same one evaluated in $u$, $w_n$ is the solution of: $$\int_{\Omega} a(u_n) \nabla w_n \nabla \varphi + \int_{\Omega} b(u_n) q_n w_n \varphi = \int_{\Omega} (1-k)(u_n - u) \nabla S(u) \nabla \varphi - \int_{\Omega} (u_n - u) S(u)^3 \varphi, \label{eq:diff}$$ where $q_n = S(u_n)^2 + S(u_n) S(u) + S(u)^2$. Considering $\varphi = w_n$ and taking advantage of the fact that $a(u_n) \geq k$ and (by simple computation) $q_n \geq \frac{3}{4}S(u)^2$, we can show that, via Cauchy-Schwarz inequality, $$\begin{aligned} k \norm{L^2(\Omega)}{\nabla w_n}^2 + \frac{3}{4}\int_\Omega b(u_n) S(u)^2 w_n^2 \leq &(1-k)\norm{L^2(\Omega)}{(u_n-u)\nabla S(u)} \norm{L^2(\Omega)}{\nabla w_n} \\ &+\norm{L^2(\Omega)}{(u_n-u)S(u)^3}\norm{L^2(\Omega)}{w_n}. \end{aligned}$$ We remark that $(u_n - u) S(u)^3 \in L^2(\Omega)$ since $S(u) \in H^1(\Omega) \subset \subset L^6(\Omega)$. Moreover, $$\begin{aligned} k \norm{L^2(\Omega)}{\nabla w_n}^2 + \frac{3}{4}\int_\Omega b(u) S(u)^2 w_n^2 \leq &(1-k)\norm{L^2(\Omega)}{(u_n-u)\nabla S(u)} \norm{L^2(\Omega)}{\nabla w_n} \\&+ \norm{L^2(\Omega)}{(u_n-u)S(u)^3}\norm{L^2(\Omega)}{w_n} + \frac{3}{4}\int_\Omega (u_n-u)S(u)^2 w_n^2 \end{aligned}$$ Thanks to Proposition \[prop:geq\], $\exists Q>0,$ $\Omega^*\subset\Omega$ s.t. $|\Omega^*|\neq 0$ s.t. $$k \norm{L^2(\Omega)}{\nabla w_n}^2 + \frac{3}{4}Q \norm{L^2(\Omega^*)}{w_n}^2 \leq (q_1 + q_2 + q_3) \norm{H^1(\Omega)}{w_n},$$ where $q_1 = \norm{L^2(\Omega)}{(u_n - u)\nabla S(u)}$, $q_2 = \norm{L^2(\Omega)}{(u_n - u) S(u)^3}$ and $q_1 = \frac{3}{4}\norm{L^2(\Omega)}{(u_n - u) S(u)^2}$, which implies, thanks to the Poincarè inequality in Lemma \[Lemma1\], $$\norm{H^1(\Omega)}{w_n} \leq C (q_1 + q_2 + q_3).$$ Consider $$q_1 = \left( \int_\Omega (u_n - u)^2 |\nabla S(u)|^2 \right)^\half;$$ since $u_n \xrightarrow{L^1} u$, then (up to a subsequence) $u_n \rightarrow u$ pointwise almost everywhere, then also the integrand $(u_n - u)^2 |\nabla S(u)|^2$ converges to $0$. Moreover, $|u_n - u| \leq 1$, hence $\forall n$ $(u_n - u)^2 |\nabla S(u)|^2 \leq |\nabla S(u)|^2 \in L^1(\Omega)$, and thanks to Lebesgue convergence theorem, we conclude that $q_1 \rightarrow 0$. Analogously, $q_2 \rightarrow 0$, $q_3 \rightarrow 0$ and eventually $\norm{H^1(\Omega)}{w_n} \rightarrow 0$, i.e. $S(u_n) \xrightarrow{H^1} S(u)$. Thanks to the trace inequality, we can assess that also $S(u_n)|_{\partial \Omega} \xrightarrow{L^2(\partial \Omega)} S(u)|_{\partial \Omega}$. It is now possible to verify the expected properties of the regularized optimization problem. For every $\alpha >0$ there exists a solution of \[prop:exist\] Let $\{u_n\}$ be a minimizing sequence: then $\{S(u_n)|_{\partial \Omega}\}$ is bounded in $L^2(\partial \Omega)$ and $\{u_n\}$ is bounded in $BV(\Omega)$ (since $\{TV(u_n)\}$ is bounded and $\norm{L^1(\Omega)}{u_n}\leq |\Omega|$ for all $u_n \in X_{0,1}$). Thanks to the result of compactness for the $BV$ space (see [@book:ambrosiofuscopallara], Theorem 3.23), there exists a subsequence $u_{n_k}$ weakly converging to an element $\bar{u} \in BV(\Omega)$. Moreover, being $\mathcal{D}(S)$ weakly closed, $\bar{u} \in \mathcal{D}(S)$. Since the weak $BV-$convergence implies the $L^1-$convergence, thanks to Proposition \[prop:continuity\] we can assess that $S(u_{n_k}) \rightarrow S(\bar{u})$ in $H^1(\Omega)$ and in $L^2(\partial \Omega)$. Eventually, this proves that $\norm{L^2(\partial\Omega)}{S(u_{n_k})-y_{meas}} \rightarrow \norm{L^2(\partial\Omega)}{S(u)-y_{meas}}$. Anologously, by semi-continuity of the total variation with respect to the weak convergence in BV, $TV(\bar{u}) \leq \liminf_{k} TV(u_{n_k})$, and it is possible to conclude that $$\half \norm{L^2(\partial\Omega)}{S(u)-y_{meas}}^2 + \alpha TV(u) \leq \liminf_k (\half \norm{L^2(\partial\Omega)}{S(u_{n_k})-y_{meas}}^2 + \alpha TV(u_{n_k})),$$ thus $u$ is a minimum of the functional. Even if the existence of the solution is ensured by the previous result, uniqueness cannot be guaranteed since the functional is neither linear nor convex (in general). We now investigate the stability of the minimizer of the regularized cost functional with respect to small perturbations of the boundary data. We point out that, due to the non-reflexivity of the Banach space $BV$, it is not possible to formulate a stability result with respect to the strong $BV$ convergence; nevertheless, we can perform the analysis with respect to the *intermediate convergence* of $BV$ functions. A sequence $\{u_n\} \subset BV(\Omega)$ tends to $u \in BV(\Omega)$ in the sense of the intermediate convergence iff $ u_n\xrightarrow{L^1} \bar{u}$ and $TV(u_n)\rightarrow TV(\bar{u})$. Fix $\alpha >0$ and consider a sequence $\{y_k\}\subset L^2(\partial \Omega)$ such that $y_k \rightarrow y_{meas}$ in $L^2(\partial \Omega)$. Consider the sequence $\{u_k\}$, where $u_k$ is a solution of with datum $y_k$. Then there exists a subsequence $\{u_{k_n}\}$ which converges to a minimizer $\bar{u}$ of with datum $y_{meas}$ in the sense of the intermediate convergence. \[prop:stability\] For every $u_k$, we have that $$\half \norm{L^2(\partial \Omega)}{S(u_k)-y_k}^2 + \alpha TV(u_k) \leq \half \norm{L^2(\partial \Omega)}{S(u)-y_k}^2 + \alpha TV(u) \quad \forall u \in \mathcal{D}(S).$$ Hence, $\{\norm{L^2(\partial \Omega)}{S(u_k)}\}$ and $\{TV(u_k)\}$ (and therefore $\{\norm{BV(\Omega)}{u_k}\}$) are bounded, and there exists a subsequence $\{u_{k_n}\}$ such that both $u_{k_n} \rightharpoonup \bar{u}$ in $BV(\Omega)$ and $S(u_{k_n}) \rightarrow S(\bar{u})$ in $L^2(\partial \Omega)$. Thanks to the continuity of the map $S$ with respect to the convergence (in $L^1$) of $u_{k_n}$ and to the weak lower semi-continuity of the $BV(\Omega)$ norm, $$\begin{aligned} \half \norm{L^2(\partial \Omega)}{S(\bar{u})-y_{meas}}^2 + \alpha TV(\bar{u}) &\leq \lim \inf_n \left(\half \norm{L^2(\partial \Omega)}{S(u_{k_n})-y_{k_n}}^2 + \alpha TV(u_{k_n}) \right) \\ & \leq \lim_{n} \left(\half \norm{L^2(\partial \Omega)}{S(u)-y_{k_n}}^2 + \alpha TV(u) \right) \quad \forall u \in \mathcal{D}(S)\\& = \half \norm{L^2(\partial \Omega)}{S(u)-y_{meas}}^2 + \alpha TV(u) \quad \forall u \in \mathcal{D}(S). \end{aligned} \label{eq:comput}$$ Hence, $\bar{u}$ is a solution of problem . In order to prove that also $TV(u_{k_n}) \rightarrow TV(\bar{u})$, first consider that, according to , $$\begin{aligned} J_{reg}(\bar{u}) &\leq \liminf_n \left(\half \norm{L^2(\partial \Omega)}{S(u_{k_n})-y_{k_n}}^2 + \alpha TV(u_{k_n}) \right) \\&\leq \lim_n \left(\half \norm{L^2(\partial \Omega)}{S(u_{k_n})-y_{k_n}}^2 + \alpha TV(u_{k_n}) \right) = J_{reg}(\bar{u}), \end{aligned}$$ hence $$\lim_n \left(\half \norm{L^2(\partial \Omega)}{S(u_{k_n})-y_{n_k}}^2 + \alpha TV(u_{k_n}) \right) = \half \norm{L^2(\partial \Omega)}{S(\bar{u})-y_{meas}}^2 + \alpha TV(\bar{u}).$$ In addition, thanks to the continuity of $S$, the first term in the sum admits a limit, i.e.: $$\lim_n \half \norm{L^2(\partial \Omega)}{S(u_{k_n})-y_{k_n}}^2 = \half \norm{L^2(\partial \Omega)}{S(\bar{u})-y_{meas}}^2 ,$$ which eventually implies that also $TV(u_{k_n}) \rightarrow TV(\bar{u})$. We finally state and prove the following result regarding asymptotic behaviour of the minimum of $J_{reg}$ when $\alpha \rightarrow 0$. Consider a sequence $\{ \alpha_k \}$ s.t. $\alpha_k \rightarrow 0$, and define the sequence $\{u_k\}$ of the solutions of with the same datum $y_{meas}$ but different weights $\alpha_k$. Suppose there exists (at least) one solution of the inverse problem . Then, $\{u_k\}$ admits a convergent subsequence with respect to the $L^1(\Omega)$ norm and the limit $u$ is a minimum-variation solution of the inverse problem, i.e. $S(u)|_{\partial \Omega} = y_{meas}$ and $TV(u) \leq TV(\widetilde{u})$ $\forall \widetilde{u}$ s.t. $S(\widetilde{u})|_{\partial \Omega} = y_{meas}$. \[prop:convergence\] Let $u^{\dag}$ be a solution of the inverse problem. By definition of $u_k$, $$\half \norm{L^2(\partial \Omega)}{S(u_k) - y_{meas}}^2 + \alpha_k TV(u_k) \leq \half \norm{L^2(\partial \Omega)}{S(u^\dag) - y_{meas}}^2 + \alpha_k TV(u^\dag) = \alpha_k TV(u^{\dag})$$ Hence, $\{TV(u_k)\}$ is bounded, and since $\norm{L^1(\Omega)}{u_k} \leq |\Omega|$, ${u_k}$ is also bounded in $BV(\Omega)$ and there exists a subsequence (still denoted as $u_k$) and $u \in X_{0,1}$ s.t. $u_k \xrightharpoonup{BV} u$. Moreover, $\norm{L^2(\partial \Omega)}{S(u_k)|_{\partial \Omega} -y_{meas}}\rightarrow 0$, which implies that $u$ is a solution of the inverse problem , and $$TV(u_k) \leq TV(u^{\dag}) \quad \Rightarrow \quad \limsup_k TV(u_k) \leq TV(u^{\dag})$$ The lower semicontinuity of the $BV$ norm with respect to the weak convergence, together with the continuity of the $L^1$ norm, implies that $$TV(u) \leq \liminf_k TV(u_k) \leq \limsup_k TV(u_k) \leq TV(u^{\dag})$$ for each solution $u^\dag$ of the inverse problem, which eventually implies that $u$ is a minimum-variation solution. Notice that, if the minimum-variation solution of problem is unique, then the sequence $\{u_k\}$ converges to it. The latter result can be improved by considering small perturbation of the data. By similar arguments as in proof of Proposition \[prop:convergence\], one can prove the following Let $y^{\delta} \in L^2(\partial \Omega)$ s.t. $\norm{L^2(\partial \Omega)}{y^{\delta}-y_{meas}} \leq \delta$ and let $\alpha(\delta)$ be such that $\alpha(\delta) \rightarrow 0$ and $\frac{\delta^2}{\alpha(\delta)} \rightarrow 0$ as $\delta \rightarrow 0$. Suppose there exists at least one solution of the inverse problem . Then, every sequence $\{u_{\alpha_k}^{\delta_k}\}$, with $\delta_k \rightarrow 0$, $\alpha_k = \alpha(\delta_k)$ and $u_{\alpha_k}^{\delta_k}$ solution of corresponding to $\alpha_k$ and $y^{\delta_k}$, has a converging subsequence with respect to the $L^1(\Omega)$ norm. The limit $u$ of every convergent subsequence is a minimum-variation solution of the inverse problem. \[prop:convergence2\] Consider a solution $u^{\dag}$ of the inverse problem. By definition of $u_{\alpha_k}^{\delta_k}$, $$\half \norm{L^2(\partial \Omega)}{S(u_{\alpha_k}^{\delta_k}) - y^{\delta_k}}^2 + \alpha_k TV(u_{\alpha_k}^{\delta_k}) \leq \half \norm{L^2(\partial \Omega)}{S(u^\dag) - y^{\delta_k}}^2 + \alpha_k TV(u^\dag) \leq \delta_k^2 + \alpha_k TV(u^{\dag}) \label{eq:1.6.1}$$ In particular, $$TV(u_{\alpha_k}^{\delta_k}) \leq \frac{\delta_k^2}{\alpha_k} + TV(u^{\dag}), \label{eq:1.6.2}$$ hence $\{u_{\alpha_k}^{\delta_k}\}$ is bounded in $BV(\Omega)$ and admits a subsequence (denoted by the same index $k$) such that $\exists u \in X_{0,1}$: $u_{\alpha_k}^{\delta_k}\xrightharpoonup{BV} u$. Passing to the limit in as $k \rightarrow +\infty$, $$\norm{L^2(\partial \Omega)}{S(u_{\alpha_k}^{\delta_k}) - y^{\delta_k}}^2 \rightarrow 0,$$ hence also $$\norm{L^2(\partial \Omega)}{S(u_{\alpha_k}^{\delta_k}) - y_{meas}}^2 \leq \norm{L^2(\partial \Omega)}{S(u_{\alpha_k}^{\delta_k}) - y^{\delta_k}}^2 + \norm{L^2(\partial \Omega)}{y^{\delta_k} - y_{meas}}^2 \rightarrow 0$$ and by continuity of the solution map, we have that $S(u)|_{\partial \Omega} = y_{meas}$, which implies that $u$ is a solution of the inverse problem. By lower semi continuity of the BV norm (hence of the total variation) with respect the weak convergence and from inequality , $$TV(u) \leq \liminf_k TV(u_{\alpha_k}^{\delta_k}) \leq \limsup_k TV(u_{\alpha_k}^{\delta_k}) \leq TV(u^{\dag}),$$ which allows to conclude that $u$ is also a minimum-variation solution of the inverse problem. Thanks to the results outlined in this section, it is possible to assess the stability of the regularized inverse problem: $$\argmin_{u \in X_{0,1}} J_{reg}(u); \quad J_{reg}(u) = \half \norm{L^2(\partial \Omega)}{S(u) - y_{meas}}^2 + \alpha TV(u),$$ In principle, one can obtain successive approximations of the solution of the inverse problem by solving the minimization problem with fixed $\alpha > 0$. However, this approach would require to deal with several technical difficulties, namely the non-differentiability of the cost functional $J_{reg}$ and the non-convexity of the space $X_{0,1}$. This will be object of future research by adapting, e.g., the technique in [@art:bartels2012] to the present context. However, in the sequel we follow a different strategy, namely introducing a phase-field relaxation of problem . Relaxation {#sec:relax} ========== In this section, we formulate the phase-field relaxation of the optimization problem . Fixed a relaxation parameter $\varepsilon > 0$, the Total Variation term in the expression of $J_{reg}$ is replaced with a smooth approximation, known as Ginzburg-Landau energy or Modica-Mortola functional; moreover, the minimization is set in a space of more regular functions. We follow a similar strategy as in [@art:deckelnick], with the additional difficulty of the non-linearity of the direct problem. In particular, we prove the main properties of the relaxed problem: existence of a solution is assessed in Proposition \[prop:exist2\] and convergence to the (sharp) initial problem as $\varepsilon \rightarrow 0$ is proved in Proposition \[prop:conv2\]. Moreover, in Proposition \[prop:optcond\], we describe the optimality conditions associated to the minimization problem, and compute the Frechét derivative of the relaxed cost functional, which is useful for the reconstruction purposes. Consider $u \in \mathcal{K} = H^1(\Omega;[0,1]) = \{ v \in H^1(\Omega): \ 0 \leq v \leq 1 \ a.e.\}$ and, for every $\varepsilon >0$, introduce the optimization problem: $$\argmin_{u \in \mathcal{K}} J_{\varepsilon}(u); \quad J_{\varepsilon}(u) = \half \norm{L^2(\partial \Omega)}{S(u)-y_{meas}}^2 + \alpha \int_{\Omega}\left( \varepsilon|\nabla u|^2 + \frac{1}{\varepsilon}u(1-u)\right). \label{eq:minrel}$$ The first theoretical result that is possible to prove guarantees the existence of a solution of the relaxed problem, employing classical techniques of Calculus of Variations. For every fixed $\varepsilon > 0$, the minimization problem has a solution $\ue \in \mathcal{K}$. \[prop:exist2\] Fix $\varepsilon > 0$ and consider a minimizing sequence $\{ u_k \} \subset \mathcal{K}$ (we omit the dependence of $u_k$ on $\varepsilon$). By definition of the minimizing sequence, $J_\varepsilon(u_k) \leq M$ indepentently of $k$, which directly implies that also $\norm{L^2(\Omega)}{\nabla u_k}^2$ is bounded. Moreover, being $u_k \in \mathcal{K}$, $0\leq u_k \leq 1$ a.e., thus $\norm{L^2(\Omega)}{u_k}^2 \leq |\Omega|$ and it is possible to conclude that $\{u_k\}$ is bounded in $H^1(\Omega)$. Thanks to weak compactness of $H^1$, there exist $\ue \in H^1(\Omega)$ and a subsequence $\{ u_{k_n} \}$ s.t. $u_{k_n} \xrightharpoonup{H^1} \ue$, hence $u_{k_n} \xrightarrow{L^2} \ue$. The strong $L^2$ convergence implies (up to a subsequence) pointwise convergence a.e., which allows to conclude (together with the dominated convergence theorem, since $u_{k_n}(1-u_{k_n}) \leq 1/2$) that $$\int_{\Omega} u_{k_n}(1-u_{k_n}) \rightarrow \int_{\Omega} \ue(1-\ue).$$ Moreover, by the lower semicontinuity of the $H^1$ norm with respect to the weak convergence, and by the compact embedding in $L^2$, $$\begin{aligned} \norm{H^1(\Omega)}{\ue}^2 &\leq \liminf_n \norm{H^1(\Omega)}{u_{k_n}}^2 \\ \norm{L^2(\Omega)}{\ue}^2 + \norm{L^2(\Omega)}{\nabla \ue}^2 &\leq \lim_n \norm{L^2(\Omega)}{u_{k_n}}^2 + \liminf_n \norm{L^2(\Omega)}{\nabla u_{k_n}}^2 \\ \norm{L^2(\Omega)}{\nabla \ue}^2 &\leq \liminf_n \norm{L^2(\Omega)}{\nabla u_{k_n}}^2. \end{aligned}$$ Moreover, using the continuity of the solution map $S$ with respect to the $L^1$ convergence, we can conclude that $$J_\varepsilon(\ue) \leq \liminf_n J_\varepsilon(u_{k_n}).$$ Finally, by pointwise convergence, $0 \leq \ue \leq 1$ a.e., hence $\ue$ is a minimum of $\Je$ in $\mathcal{K}$. The asymptotic behaviour of the phase-field problem when $\varepsilon \rightarrow 0$ is investigated in the next two propositions. First, we prove that the relaxed cost functional $J_\varepsilon$ converges to $J_{reg}$ in the sense of the $\Gamma$-convergence, which will naturally entail the convergence of the corresponding minimizers. Before stating the next result, we have to introduce the space $X$ of the Lebesgue-measurable functions over $\Omega$ endowed with the $L^1(\Omega)$ norm and consider the following extension of the cost funtionals: problem is replaced by $$\argmin_{u \in X} \tilde{J}(u); \quad \tilde{J}(u) = \left\{ \begin{aligned} J_{reg}(u) & \quad \textit{if $u \in X_{0,1}=BV(\Omega;\{0,1\})$}\\ \infty & \quad \textit{otherwise} \end{aligned} \right. \label{eq:minregbis}$$ whereas is replaced by $$\argmin_{u \in X} \tilde{J}_{\varepsilon}(u); \quad \tilde{J}_{\varepsilon}(u) = \left\{ \begin{aligned} J_\varepsilon(u) & \quad \textit{if $u \in \mathcal{K}$}\\ \infty & \quad \textit{otherwise} \end{aligned} \right. \label{eq:minrel2}$$ It is now possible to formulate the convergence result, whose proof can be easily obtained by adapting the one of [@art:deckelnick Theorem 6.1]. Consider a sequence $\{\varepsilon_k\}$ s.t. $\varepsilon_k \rightarrow 0$. Then, the functionals $\tilde{J}_{\varepsilon_k}$ converge to $\tilde{J}$ in $X$ in the sense of the $\Gamma-$convergence. Finally, from the compactness result in [@baldo1990 Proposition 4.1] and applying the definition of $\Gamma$-convergence, it is easy to prove the following convergence result for the solutions of . Consider a sequence $\{\varepsilon_k\}$ s.t. $\varepsilon_k \rightarrow 0$ and let $\{u_{\varepsilon_k}\}$ be the sequence of the respective minimizers of the functionals $\{\tilde{J}_{\varepsilon_k}\}$. Then, there exists a subsequence, still denoted as $\{\varepsilon_k\}$ and a function $u \in X_{0,1}$ such that $u_{\varepsilon_k} \rightarrow u$ in $L^1$ and $u$ is a solution of . \[prop:conv2\] Optimality conditions {#sec:oc} --------------------- We can now provide an expression for the optimality condition associated with the minimization problem , which is formulated as a variational inequality involving the Fréchet derivative of $\Je$. Consider the solution map $S:\mathcal{K} \rightarrow H^1(\Omega)$ and let $f \in L^2(\Omega)$ satisfy the hypotheses in Proposition \[prop:geq\] or \[prop:geq2\]: for every $\varepsilon>0$, the operators $S$ and $\Je$ are Fréchet-differentiable on $\mathcal{K} \subset L^{\infty}(\Omega) \cap H^1(\Omega)$ and a minimizer $\ue$ of $\Je$ satisfies the variational inequality: $$J'_\varepsilon(u_\varepsilon)[v - u_\varepsilon] \geq 0 \qquad \forall v \in \mathcal{K}, \label{eq:OC}$$ being $$J'_\varepsilon(u)[\vartheta] = \int_{\Omega}(1-k)\vartheta \nabla S(u) \cdot \nabla p + \int_{\Omega} \vartheta S(u)^3 p + 2 \alpha \varepsilon \int_{\Omega}\nabla u \cdot \nabla \vartheta + \frac{\alpha}{\varepsilon} \int_{\Omega}(1-2u)\vartheta; \label{eq:VI}$$ where $\vartheta \in \mathcal{K}-\mathcal{K}=\{v \ s.t. \ v=u_1-u_2, \ u_1,u_2 \in \mathcal{K}\}$ and $p$ is the solution of the *adjoint problem*: $$\int_{\Omega} a(u)\nabla p \cdot \nabla \psi + \int_{\Omega} 3b(u)S(u)^2 p \psi = \int_{\partial \Omega}{(S(u)-y_{meas})\psi} \qquad \forall \psi \in H^1(\Omega). \label{eq:adjoint}$$ \[prop:optcond\] First of all we need to prove that $S$ is Fréchet differentiable in $L^{\infty}(\Omega)$: in particular, we claim that for $\vartheta \in L^{\infty}(\Omega)\cap (\mathcal{K}-\mathcal{K})$ it holds that $S'(u)[\vartheta] = S_*$, being $S_*$ the solution in $H^1(\Omega)$ of $$\int_{\Omega} a(u)\nabla S_* \nabla \varphi + \int_{\Omega} b(u) 3S(u)^2 S_* \varphi = \int_{\Omega} (1-k)\vartheta \nabla S \nabla \varphi + \int_{\Omega} \vartheta S(u)^3 \varphi \quad \forall \varphi \in H^1(\Omega), \label{eq:derprob}$$ namely, that $$\norm{H^1(\Omega)}{S(u+\vartheta) - S(u) - S_*} = o(\norm{L^{\infty}(\Omega)}{\vartheta}). \label{eq:Frechet}$$ First we show that if $\vartheta \in L^{\infty}(\Omega)\cap (\mathcal{K}-\mathcal{K})$, then $\norm{H^1(\Omega)}{S(u+\vartheta)-S(u)} \leq C \norm{L^{\infty}(\Omega)}{\vartheta}$. Indeed, the difference $w = S(u+\vartheta)-S(u)$ satisfies $$\begin{aligned} \int_{\Omega} a(u+\vartheta)\nabla w \nabla \varphi + \int_{\Omega} b(u+\vartheta) q w \varphi =& -\int_{\Omega} (a(u+\vartheta)-a(u))\nabla S(u) \nabla \varphi \\ &- \int_{\Omega} (b(u+\vartheta)-b(u)) S(u)^3 \varphi \qquad \forall \varphi \in H^1(\Omega), \end{aligned} \label{eq:difference}$$ with $q = S(u+\vartheta)^2 + S(u)S(u+\vartheta) +S(u)^2$. Since $a(u+\vartheta)-a(u)= -(1-k)\vartheta$ and $b(u+\vartheta)-b(u) = -\vartheta$, and arguing as in the proof of Proposition \[prop:continuity\], take $\varphi = w$ in : we obtain $$k\norm{L^2}{\nabla w}^2 + \frac{3}{4}\int_\Omega b(u+\vartheta)S(u)^2w^2 \leq (1-k)\norm{L^\infty}{\vartheta} \norm{L^2}{\nabla S(u)} \norm{L^2}{\nabla w} + \norm{L^2}{S(u)^3} \norm{L^2}{w}\norm{L^{\infty}}{\vartheta}$$ and again by Proposition \[prop:geq\] $$\begin{aligned} k \norm{L^2}{\nabla w}^2 + \frac{3}{4}Q \norm{L^2(\Omega^*)}{w}^2 \leq &(1-k)\norm{L^\infty}{\vartheta} \norm{L^2}{\nabla S(u)} \norm{L^2}{\nabla w} + \norm{L^\infty}{\vartheta} \norm{L^2}{S(u)^3}\norm{L^2}{w} \\&+ \frac{3}{4} \norm{L^\infty}{\vartheta} \norm{L^2}{S(u)^2}\norm{L^2}{w}. \end{aligned}$$ By and by Sobolev inequality, eventually $$\norm{H^1(\Omega)}{w}^2 \leq C \norm{H^1(\Omega)}{S(u)}\norm{H^1(\Omega)}{w}\norm{L^\infty}{\vartheta},$$ hence $\norm{H^1(\Omega)}{S(u+\vartheta)-S(u)} = O(\norm{L^\infty(\Omega)}{\vartheta})$. Take now and subtract . Define $r = S(u+\vartheta)-S(u) - S_*$: it holds that $$\begin{aligned} \int_{\Omega} a(u)\nabla r \nabla \varphi + \int_{\Omega} b(u) 3 S(u)^2 r \varphi =& \int_{\Omega} (a(u + \vartheta) - a(u))\nabla w \cdot \nabla \varphi \\ &+ \int_{\Omega} (b(u+\vartheta)q - 3b(u)S(u)^2 )w \varphi \qquad \forall \varphi \in H^1(\Omega). \end{aligned}$$ The second integral in the latter sum can be split as follows: $$\int_{\Omega} (b(u+\vartheta)q - 3b(u)S(u)^2 )w \varphi = \int_{\Omega} (b(u+\vartheta)-b(u))q w \varphi + \int_{\Omega} (q - 3S(u)^2 )b(u)w \varphi,$$ and in particular $q - 3S(u)^2 = S(u+\vartheta)^2 + S(u+\vartheta)S(u) - 2S(u)^2 = hw$, where $h = S(u+\vartheta)+2S(u) \in H^1(\Omega)$. Hence, chosen $\varphi = r$ and exploiting again the Poincaré inequality in Lemma \[Lemma1\] and the H[ö]{}lder inequality: $$\begin{aligned} \frac{1}{C}\norm{H^1}{r}^2 \leq & k\norm{L^2}{\nabla r}^2 + Q \norm{L^2(\Omega^*)}{r} \leq (1-k)\norm{L^{\infty}}{\vartheta} \norm{L^2}{\nabla w}\norm{L^2}{\nabla r} \\ & + \norm{L^{\infty}}{\vartheta} \norm{L^4}{q}\norm{L^2}{w}\norm{L^4}{r} + \norm{L^4}{h}\norm{L^4}{w}^2 \norm{L^4}{r}\\ & \leq \left( (1-k)\norm{L^{\infty}}{\vartheta}\norm{H^1}{w} + \norm{H^1}{q}\norm{L^{\infty}}{\vartheta}\norm{H^1}{w} + \norm{H^1}{h}\norm{H^1}{w}^2 \right)\norm{H^1}{r}. \end{aligned}$$ It follows eventually that $\norm{H^1(\Omega)}{r} \leq C \norm{L^{\infty}}{\vartheta}^2 = o(\norm{L^{\infty}}{\vartheta})$, which guarantees that $S_*=S'(u)[\vartheta]$. The last step is to provide an expression of the Fréchet derivative of $\Je$. Exploiting the fact that $S$ is differentiable, we can compute the expression of $\Je'(u)$ through the *chain rule*: $$\Je'(u)[\vartheta] = \int_{\partial \Omega} (S(u)-y_0) S'(u)[\vartheta] + \alpha \int_{\Omega}\left( 2\varepsilon \nabla u \nabla \vartheta + \frac{1}{\varepsilon}(1-2u)\vartheta \right). \label{eq:first}$$ Finally, thanks to the expression of the adjoint problem, $$\begin{aligned} \int_{\partial \Omega} (S(u)-y_0) S'(u)[\vartheta] =& \int_{\partial \Omega} (S(u)-y_0)S_* = \int_{\Omega} a(u)\nabla p \cdot \nabla S_* + \int_{\Omega} 3S(u)^2 p S_* = \\ \textit{(by definition of $S_*$)} =& \int_{\Omega}(1-k)\vartheta \nabla S(u) \cdot\nabla p + \int_{\Omega} \vartheta S(u)^3p , \end{aligned}$$ and hence: $$\Je'(u)[\vartheta] = \int_{\Omega}(1-k)\vartheta \nabla S(u)\cdot \nabla p + \int_{\Omega} \vartheta S(u)^3p + \alpha \int_{\Omega}\left( 2\varepsilon \nabla u \cdot \nabla \vartheta + \frac{1}{\varepsilon}(1-2u)\vartheta \right).$$ It is eventually a standard argument that, being $\Je$ a continuous and Frechét differentiable functional on a convex subset $\mathcal{K}$ of the Banach space $H^1(\Omega)$, the optimality conditions for the optimization problem are expressed by the variational inequality . Discretization and reconstruction algorithm {#sec:discr} =========================================== For a fixed $\varepsilon>0$, we now introduce a discrete formulation of problem in order to define a numerical reconstruction algorithm and compute an approximated solution of the inverse problem. In what follows, we consider $\Omega$ to be polygonal, in order to avoid a discretization error involving the geometry of the domain. Let $\Tau_h$ be a shape regular triangulation of $\Omega$ and define $V_h \subset H^1(\Omega)$: $$V_h = \{ v_h \in C(\bar{\Omega}), v_h|_K \in \mathbb{P}_1(K) \text{ } \forall K \in \Tau_h \}; \qquad \mathcal{K}_h = V_h \cap \mathcal{K}.$$ It is well known, from the Clément interpolation theory (see e.g. [@book:brenner2007]), that for every $w \in H^1(\Omega)$ there exists a sequence $\{w_h\}$ such that $$w_h \in V_h \qquad w_h \xrightarrow{H^1} w \qquad \text{as $h \rightarrow 0$}. \label{eq:Clement}$$ For every fixed $h>0$, we define the solution map $S_h: \mathcal{K} \rightarrow V_h$, where $S_h(u)$ solves $$\int_{\Omega} a(u) \nabla S_h(u) \nabla v_h + \int_{\Omega} b(u) S_h(u)^3 v_h = \int_{\Omega} f_h v_h \quad v_h \in V_h,$$ being $f_h$ the Clément interpolator of $f$ in the space $V_h$, hence $\norm{H^1(\Omega)}{f_h - f} \rightarrow 0$. Convergence analysis as $h \rightarrow 0$ ----------------------------------------- The present section is devoted to the numerical analysis of the discretized problem: the convergence of the approximated solution of the direct problem is studied, taking into account the difficulties implied by the non linear term. Moreover, the existence and convergence of minimizers of the discrete cost functional is analysed. The following result, which is preliminary for the proof of the convergence of the approximated solutions to the exact one, can be proved by resorting at the techniques of [@art:ciarlet Theorem 2.1]. For completeness we briefly report the proof. Let $f \in L^2(\Omega)$ satisfy the hypotheses in Proposition \[prop:geq\] or \[prop:geq2\]; then, for every $u \in \mathcal{K}$, $S_h(u) \rightarrow S(u)$ strongly in $H^1(\Omega)$. \[prop:ciarlet\] As in the proof of Proposition \[prop:wellpos\], for a fixed $u \in \mathcal{K}$ we define the operator $T:H^1(\Omega) \rightarrow (H^1(\Omega))^*$ such that $$\langle T(y), \varphi \rangle = \int_{\Omega} a(u)\nabla y \nabla \varphi + \int_{\Omega} b(u) y^3 \varphi;$$ then $y_h = S_h(u)$ and $y = S(u)$ are respectively the solutions of the equations $$\begin{aligned} \langle T(y_h), \varphi \rangle &= \int_\Omega f_h \varphi \quad & \forall \varphi \in V_h; \qquad \qquad \langle T(y), \varphi \rangle &= \int_\Omega f \varphi \quad & \forall \varphi \in H^1(\Omega). \end{aligned}$$ It is easy to prove that $$\langle T(y_h) - T(y), y_h - y \rangle \geq C(k,y) \norm{H^1}{y_h-y}^2; \label{eq:ellipticity}$$ indeed, thanks to Lemma \[Lemma1\] and Proposition \[prop:geq\], $$\langle T(y_h) - T(y), y_h - y \rangle = \int_\Omega a(u) |\nabla (y_h - y)|^2 + \int_\Omega b(u)(y_h - y)^2 (y_h^2 + y_h y + y^2) \geq C \norm{H^1}{y_h - y}^2.$$ Thanks to , consider a generic $w_h \in V_h$, $$\begin{aligned} \norm{H^1}{y_h - y}^2 &\leq \langle T(y_h) - T(y), y_h - y \rangle = \langle T(y_h) - T(y), w_h - y \rangle + \langle T(y_h) - T(y), y_h - w_h \rangle \\ & \leq \langle T(y_h) - T(y), w_h - y \rangle + \int_\Omega (f_h - f)(y_h - w_h) \\ & \leq K \norm{H^1}{w_h - y} \norm{H^1}{y_h - y} + \norm{H^1}{f_h - f}\norm{H^1}{y_h - w_h}, \end{aligned}$$ where $K$ is the (local) Lipschitz constant of $T$ (see Proposition \[prop:wellpos\]). Hence: $$\begin{aligned} \norm{H^1}{y_h - y} &\leq K \norm{H^1}{w_h - y} + \half \sqrt{K^2 \norm{H^1}{w_h - y}^2 + \norm{H^1}{y_h - w_h}^2 \norm{H^1}{f_h - f}^2} \\ &\leq C_1 \norm{H^1}{w_h - y} + C_2 \norm{H^1}{y_h - w_h} \norm{H^1}{f_h - f}\\ &\leq (C_1 + C_2 \norm{H^1}{f_h - f})\norm{H^1}{w_h - y} + C_2 \norm{H^1}{y_h - y} \norm{H^1}{f_h - f}. \end{aligned}$$ Since $f_h \xrightarrow{H^1} f$, we can choose $h$ sufficiently small s.t. $C_2 \norm{H^1}{f_h - f} \leq \half$, hence: $$\norm{H^1}{y_h - y} \leq 2 \left(C_1 + \half \right) \norm{H^1}{w_h - y},$$ and since the latter inequality holds for each $w_h \in H^1(\Omega)$, it holds: $$\norm{H^1(\Omega)}{y_h - y} \leq C \inf_{w_h \in V_h}\norm{H^1(\Omega)}{w_h - y}.$$ Finally, exploiting , we conclude the thesis. The convergence of the solution of the discrete direct problem to the continuous one is an immediate consequence of Lemma \[prop:ciarlet\] and of the continuity of the map $S_h$ in the space $V_h$, which can be assessed analogously to the proof of Proposition \[prop:continuity\]. Let $\{h_k\}, \{u_k\}$ be two sequences such that $h_k \rightarrow 0$, $u_k \in \mathcal{K}_{h_k}$ and $u_k \xrightarrow{L^1} u$ and $u$ is not identically equal to 1. Then $S_{h_k}(u_k) \xrightarrow{H^1} S(u)$. \[cor:1\] Define the discrete cost functional, $J_{\varepsilon,h}: \mathcal{K}_h \rightarrow \R$ $$J_{\varepsilon,h}(u_h) = \half \norm{L^2(\partial \Omega)}{S_h(u_h)-y_{meas,h}}^2 + \alpha \int_{\Omega}\left( \varepsilon|\nabla u_h|^2 + \frac{1}{\varepsilon}u_h(1-u_h)\right), \label{eq:Jeh}$$ being $y_{meas,h}$ the best approximation of the boundary datum $y_{meas}$ in the space of the traces of $V_h$ functions. The existence of minimizers of the discrete functionals $J_{\varepsilon,h}$ is stated in the following proposition, together with an asymptotic analysis as $h \rightarrow 0$. Taking advantage of Proposition \[cor:1\], the proof is analogous to the one of [@art:deckelnick Theorem 3.2]. For each $h>0$, there exists $u_h \in \mathcal{K}_h$ such that $J_{\varepsilon,h}(u_h) = min_{v_h \in \mathcal{K}_h}J_{\varepsilon,h}(v_h)$. Every sequence $\{u_{h_k}\}$ s.t. $\lim_{k\rightarrow \infty}h_k = 0$ admits a subsequence that converges in $H^1(\Omega)$ to a minimum of the cost functional $J_\varepsilon$. \[prop:minJeh\] The strategy we adopt in order to minimize the discrete cost functional $J_{\varepsilon,h}$ is to search for a function $u_h$ satisfying discrete optimality conditions, which can be obtained as in section \[sec:oc\]: $$J'_{\varepsilon,h}(u_h)[v_h - u_h] \geq 0 \quad \forall v_h \in \mathcal{K}_h \label{eq:discrVI}$$ where for each $\theta_h \in \mathcal{K}_h - \mathcal{K}_h := \{\theta_h = w_h - v_h; \ w_h, v_h \in \mathcal{K}_h\}$ it holds $$J'_{\varepsilon,h}(u_h)[\vartheta_h] = \int_{\Omega}(1-k)\vartheta_h \nabla S_h(u_h) \cdot \nabla p_h + \int_{\Omega}\vartheta_h S_h(u_h)^3 p_h + 2 \alpha \varepsilon \int_{\Omega}\nabla u_h \cdot \nabla \vartheta_h + \frac{\alpha}{\varepsilon} \int_{\Omega}(1-2u_h)\vartheta_h, \label{eq:discrJ1}$$ where $p_h$ is the solution in $V_h$ of the adjoint problem associated to $u_h$. It is finally possible to demonstrate the convergence of critical points of the discrete functionals $J_{\varepsilon,h}$ (i.e., functions in $\mathcal{K}_h$ satisfying ) to a critical point of the continuous on, $J_\varepsilon$. The proof can be adapted from the one of [@art:deckelnick Theorem 3.2]. Consider a sequence $\{h_k\}$ s.t. $h_{k} \rightarrow 0$ and for every $k$ denote as $u_k$ a solution of the discrete variational inequality . Then there exists a subsequence of $\{u_k\}$ that converges a.e and in $H^1(\Omega)$ to a solution $u$ of the continuous variational inequality \[prop:discrVI\] Reconstruction algorithm: a Parabolic Obstacle Problem approach {#POP} --------------------------------------------------------------- The necessary optimality conditions that have been stated in Proposition \[prop:optcond\], together with the expression of the Fréchet derivative of the cost functional reported in allow to define a Parabolic Obstacle problem, which consists in a very common strategy in order to search for a solution of optimization problems in a phase-field approach. In this section we give a continuous formulation of the problem, and provide a formal proof of its desired properties. We then introduce a numerical discretization of the problem and rigorously prove the main convergence results. The core of the proposed approach is to rely on a parabolic problem whose solution $u(\cdot,t)$ converges, as the fictitious time variable tends to $+\infty$, to an asymptotic state $u_{\infty}$ satisfying the continuous optimality conidtions . The problem can be formulated as follows, for a fixed $\varepsilon > 0$: let $u$ be the solution of $$\left\{ \begin{aligned} \int_{\Omega} \partial_t u (v-u) + \Je'(u)[v-u] &\geq 0 \qquad \forall v \in \mathcal{K}, \quad t \in(0,+\infty)\\ u(\cdot,0) &= u_0 \in \mathcal{K} \end{aligned} \right. \label{eq:POP}$$ The theoretical analysis of the latter problem is beyond the purposes of this work, and would require to deal with the severe non-linearity of the expression of $\Je'(u)$. We reduce ourselves to formally report the expected properties of the Parabolic Obstacle Problem and then analyse in detail its discretised version. The motivation for the introduction of the Parabolic Obstacle Problem is twofold: - *the evaluation of the cost functional along the solution of problem is a decreasing function of time.* Indeed, $$\begin{aligned} \frac{d}{dt} \Je(u(\cdot,t)) \leq -\norm{L^2(\Omega)}{\partial_t u(\cdot,t)}^2 \leq 0. \end{aligned}$$ - *As $t \rightarrow +\infty$, the solution $u(\cdot,t)$ converges to $u_\infty \in H^1(\Omega)$, which satisfies the optimality conditions* . We now provide a complete discretization of the Parabolic Obstacle Problem by setting in the discrete spaces $\mathcal{K}_h$ and $V_h$, and by considering a semi-implicit one-step scheme for the time updating, as in [@art:deckelnick]: i.e., by treating explicitly the nonlinear terms and implicitly the linear ones. We obtain that the approximate solution $\{u_h^n\}_{n\in \N} \subset V_h,$ $u_h^n\approx u(\cdot,t^n)$ is computed as: $$\left\{ \begin{aligned} u_h^0 &= u_0 \in \mathcal{K}_h \qquad \textit{(a prescribed initial datum)}\\ u_h^{n+1}& \in \mathcal{K}_h: \int_{\Omega}(u_h^{n+1}-u_h^n)(v_h-u_h^{n+1}) + \tau_n \int_{\Omega}(1-k)\nabla S_h(u_h^n)\cdot \nabla p_h^n (v_h - u_h^{n+1}) \\ &\qquad \qquad+\tau_n \int_{\Omega}S_h(u_h^n)^3 p_h^n (v_h - u_h^{n+1}) + 2 \tau_n \alpha \varepsilon \int_{\Omega}\nabla u_h^{n+1} \cdot \nabla(v_h - u_h^{n+1}) \\ &\qquad \qquad+\tau_n \alpha \frac{1}{\varepsilon} \int_{\Omega}(1-2u_h^n)(v_h - u_h^{n+1})\geq 0 \quad \forall v_h \in \mathcal{K}_h, \ n=0,1,\ldots \end{aligned} \right. \label{eq:fulldiscrPOP}$$ For the fully-discretized problem , it is possible to prove rigorously the properties that we have formally stated for the continuous one; in particular, the convergence of the sequence $\{u_h^n\}$ to a critical point of the discrete cost functional $J_{\varepsilon,h}$. The following preliminary result is necessary for the proof of the main property: For each $n > 0$, there exists a positive constant $\mathcal{B}_n = \mathcal{B}_n(\Omega, h, k, \norm{H^1}{p_h^n}, \norm{H^1}{y_h^n}, \norm{H^1}{y_h^{n+1}})$ such that, provided that $\tau_n \leq \mathcal{B}_n$ it holds that: $$\norm{L^2}{u_h^{n+1} - u_h^n}^2 + J_{\varepsilon,h}(u_h^{n+1}) \leq J_{\varepsilon,h}(u_h^n) \quad n>0. \label{eq:Jdecrease}$$ \[prop:decrease\] In the expression of the discrete parabolic obstacle problem , consider $v_h = u_h^n$: via simple computation, we can point out that $$\begin{aligned} \frac{1}{\tau_n}&\norm{L^2}{u_h^{n+1}-u_h^n}^2 + J(u_h^{n+1}) - J(u_h^{n}) + \alpha \varepsilon \norm{L^2}{\nabla (u_h^{n+1} - u_h^n)}^2 + \frac{\alpha}{\varepsilon}\norm{L^2}{u_h^{n+1}-u_h^n}^2 \\ &\leq \int_{\Omega}\left( a(u_h^{n+1}) - a(u_h^{n})\right) \nabla y_h^n \nabla p_h^n + \int_{\Omega}\left( b(u_h^{n+1}) - b(u_h^{n}) \right)(y_h^n)^3 p_h^n \\ &+ \half \norm{L^2(\partial \Omega)}{y_h^{n+1}-y_h^n}^2 + \int_{\partial \Omega} (y_h^{n+1}-y_h^n)(y_h^{n+1} - y_{meas,h}), \end{aligned}$$ where $y_h^n = S_h(u_h^n)$ and $y_h^{n+1} = S_h(u_h^{n+1})$. Moreover, by the expression of the adjoint problem, $$RHS = \half \norm{L^2(\partial \Omega)}{y_h^{n+1}-y_h^n}^2 + {\tikz[baseline=(char.base)]{ \node[shape=circle,draw,inner sep=2pt] (char) {I};}} + {\tikz[baseline=(char.base)]{ \node[shape=circle,draw,inner sep=2pt] (char) {II};}},$$ where $$\begin{aligned} {\tikz[baseline=(char.base)]{ \node[shape=circle,draw,inner sep=2pt] (char) {I};}} &= \int_{\Omega}\left( a(u_h^{n+1}) - a(u_h^{n})\right) \nabla y_h^n \cdot \nabla p_h^n + \int_\Omega a(u_h^n)\nabla p_h^n \cdot \nabla(y_h^{n+1}-y_h^n) \\ &= \int_{\Omega}\left( a(u_h^n) - a(u_h^{n+1})\right) \nabla (y_h^{n+1} - y_h^n) \cdot \nabla p_h^n + \int_\Omega a(u_h^{n+1})\nabla y_h^{n+1}\cdot \nabla p_h^n - \int_\Omega a(u_h^{n})\nabla y_h^{n}\cdot \nabla p_h^n; \end{aligned}$$ $$\begin{aligned} {\tikz[baseline=(char.base)]{ \node[shape=circle,draw,inner sep=2pt] (char) {II};}} &= \int_{\Omega}\left( b(u_h^{n+1}) - b(u_h^{n})\right) (y_h^n)^3 p_h^n + 3\int_\Omega b(u_h^n)(y_h^n)^2 p_h^n (y_h^{n+1}-y_h^n) = \\ &= \int_{\Omega} b(u_h^{n+1}) \left((y_h^n)^3-(y_h^{n+1})^3\right) p_h^n + 3\int_\Omega b(u_h^n)(y_h^n)^2 p_h^n (y_h^{n+1}-y_h^n) + \int_\Omega b(u_h^{n+1})(y_h^{n+1})^3p_h^n \\ & \qquad - \int_\Omega b(u_h^{n})(y_h^{n})^3p_h^n = \qquad \textit{(by the expansion $(y_h^{n+1})^3 = \left( y_h^n + (y_h^{n+1}-y_h^n) \right)^3$)} \\ &= 3\int_\Omega \left(b(u_h^n) - b(u_h^{n+1})\right)(y_h^n)^2 p_h^n (y_h^{n+1}-y_h^n) - 3\int_\Omega b(u_h^{n+1})(y_h^n) p_h^n (y_h^{n+1}-y_h^n)^2 \\ & \qquad - \int_\Omega b(u_h^{n+1}) p_h^n (y_h^{n+1}-y_h^n)^3 + \int_\Omega b(u_h^{n+1})(y_h^{n+1})^3p_h^n - \int_\Omega b(u_h^{n})(y_h^{n})^3p_h^n. \end{aligned}$$ Collecting the terms and taking advantage of the expression of the direct problem, we conclude that $$\begin{aligned} RHS =& \half \norm{L^2(\partial \Omega)}{y_h^{n+1}-y_h^n}^2 + \int_{\Omega}\left( a(u_h^n) - a(u_h^{n+1})\right) \nabla (y_h^{n+1} - y_h^n) \cdot \nabla p_h^n \\ & \qquad + 3\int_\Omega \left(b(u_h^n) - b(u_h^{n+1})\right)(y_h^n)^2 p_h^n (y_h^{n+1}-y_h^n) \\ & \qquad - 3\int_\Omega b(u_h^{n+1})(y_h^n) p_h^n (y_h^{n+1}-y_h^n)^2 - \int_\Omega b(u_h^{n+1})p_h^n (y_h^{n+1}-y_h^n)^3. \end{aligned}$$ We now employ the Cauchy-Schwarz inequality and the regularity of the solutions of the discrete direct and adjoint problems (in particular the equivalence of the $W^{1,\infty}$ and $H^1$ norm in $V_h$: $\norm{W^{1,\infty}}{u_h} \leq C_1 \norm{H^1}{u_h}$, $C_1 = C_1(\Omega, h)$): $$RHS \leq C_2 \norm{L^2}{u_h^{n+1} - u_h^n}\norm{H^1}{y_h^{n+1} - y_h^n} + C_3 \norm{H^1}{y_h^{n+1} - y_h^n}^2$$ with $C_2 = (1-k)C_1\norm{H^1}{p_h^n} + C_1 \norm{H^1}{y_h^n}\norm{H^1}{p_h^n}$ and $C_3 = 3C_1^2\norm{H_1}{y_h^n}\norm{H_1}{p_h^n} + C_1^3\norm{H_1}{p_h^n}(\norm{H_1}{y_h^n}+\norm{H_1}{y_h^{n+1}}) + \half C_{tr}^2$, being $C_{tr}$ the constant of the trace inequality in $H^1(\Omega)$. Eventually, similarly to the computation included in the proof of Proposition \[prop:optcond\], one can assess that $$\norm{H^1}{y_h^{n+1} - y_h^n} \leq C_4 \norm{L^2}{u_h^{n+1} - u_h^n},$$ with $C_4 = C_4(k, C_1,\norm{H^1}{y_h^{n}},\Omega)$. Hence, we can conclude that there exists a positive constant $\mathcal{C}_n = C_2 C_4 + C_3 C_4^2$ such that $$\frac{1}{\tau_n}\norm{L^2}{u_h^{n+1}-u_h^n}^2 + J(u_h^{n+1}) - J(u_h^{n}) \leq \mathcal{C}_n \norm{L^2}{u_h^{n+1} - u_h^n}^2,$$ and choosing $\tau_n < \mathcal{B}_n := \frac{1}{1 + \mathcal{C}_n}$ we can conclude the thesis. We are finally able to prove the following convergence result for the fully discretized Parabolic Obstacle Problem: Consider a starting point $u_h^0 \in \mathcal{K}_h$. Then, there exists a collection of timesteps $\{\tau_n\}$ s.t. $0 <\gamma \leq \tau_n \leq \mathcal{B}_n$ $\forall n>0$. Corresponding to $\{\tau_n\}$, the sequence $\{u_h^n\}$ generated by has a converging subsequence (which we still denote with $u_h^n$) such that $u_h^n \xrightarrow{W^{1,\infty}} u_h \in V_h$, which satisfies the discrete optimality conditions . Consider a generic collection of timesteps $\tilde{\tau}_n$ satisfying $\tilde{\tau}_n \leq \mathcal{B}_n$ $\forall n>0$. Hence, by Lemma \[prop:decrease\], $$\sum_{n=0}^{\infty} \norm{L^2}{u_h^{n+1} - u_h^n}^2 \leq J_{\varepsilon,h}(u_h^0) \quad \text{and} \quad \sup_{n} J_{\varepsilon,h}(u_h^n) \leq J_{\varepsilon,h}(u_h^0)$$ which implies that $\norm{L^2}{u_h^{n+1} - u_h^n} \rightarrow 0$ and hence $u_h^n$ is bounded in $H^1(\Omega)$, and this implies that also $\{y_h^n\}$ and $\{p_h^n\}$ are bounded in $H^1(\Omega)$. According to the definition of the constants $\mathcal{C}_n$ and $\mathcal{B}_n$ reported in the proof of Lemma \[prop:decrease\], this entails that there exists a constant $M>0$ such that $\mathcal{C}_n \leq M$ $\forall n>0$, and equivalently there exists a positive constant $\gamma$ s.t. $\gamma \leq \mathcal{B}_n$. Hence, it is possible to choose, for each $n>0$, $\gamma \leq \tau_n \leq \mathcal{B}_n$.\ Eventually, we conclude that there exists $u_h \in \mathcal{K}_h$ such that, up to a subsequence, $u_h^n \rightarrow u_h$ a.e. and in $W^{1,\infty}(\Omega)$ (and $y_h^n \rightarrow y_h^n := S_h(u_h^n)$, $p_h^n \rightarrow p_h$ in $H^1$ and in $W^{1,\infty}$ as well). We exploit the expression of the discrete Parabolic Obstacle Problem to show that $$\begin{aligned} \int_{\Omega}(1-k)\nabla y_h^n \cdot \nabla p_h^n (v_h - u_h^{n+1}) + \int_{\Omega}(y_h^n)^3 p_h^n (v_h - u_h^{n+1}) + 2 \alpha \varepsilon \int_{\Omega}\nabla u_h^{n+1} \cdot \nabla(v_h - u_h^{n+1}) \\ + \alpha \frac{1}{\varepsilon} \int_{\Omega}(1-2u_h^n)(v_h - u_h^{n+1}) \geq - \frac{1}{\tau_n} \int_{\Omega}(u_h^{n+1}-u_h^n)(v_h-u_h^{n+1}) \quad \forall v_h \in \mathcal{K}_h, \end{aligned}$$ and since $\frac{1}{\tau_n} > \frac{1}{\gamma}$ $\forall n$, when taking the limit as $n \rightarrow \infty$, the right-hand side converges to $0$, which entails that $u_h$ satisfies the discrete optimality conditions . The most remarkable outcome of the analyzed discretization of the Parabolic Obstacle Problem is the implementation of an iterative reconstruction algorithm which requires, at each iteration, the solution of two boundary value problem and of a quadratic constraint minimization problem. Indeed, introducing a basis $\{\phi_i\}_{i=1}^{N_h}$ in the discrete space $V_h$, the variational inequality in can be written in algebraic form. The resulting system of inequalities can be interpreted as the optimality condition of a minimization problem involving a quadratic cost functional in the compact set $[0,1]^{N_h}$, and is efficiently solved by means of the Primal-Dual Active Set method, introduced in [@art:bergounioux1997augemented] and applied in [@art:blank2013nonlocal] on a nonlocal Allen-Cahn equation. The final formulation of the reconstruction algorithm is the following: $u_h^{n}$ It is a common practice to increase the performance of a reconstruction algorithm taking advantage of multiple measurements. In this context, it is possible to suppose the knowledge of $N_f$ different measurements of the electric potential on the boundary, $y_{meas,j}$ $j=1,\cdots,N_f$, associated to different source terms $f_j$. Therefore, instead of tackling the optimization of the mismatch functional $J$ as in , it is possible to introduce the averaged cost functional $J^{TOT}(u) = \frac{1}{N_f}\sum_{j=1}^{N_f}J^j(u)$, where $J^j(u) = \half \norm{L^2(\partial \Omega)}{S_j(u) - y_{meas,j}}^2$, being $S_j(u)$ the solution of the direct problem with source term $f = f_j$. The process of regularization, relaxation and computation of the optimality conditions is exactly the same as for $J$, and yields the same reconstruction algorithm as in Algorithm \[al:POP\], where at each timestep the solution of $N^f$ direct and adjoint problem must be computed. \[multi\] Numerical results {#sec:numer} ================= In this section we report various results obtained applying Algorithm \[al:POP\]. In all the numerical experiments, we consider $\Omega = (-1,1)^2$ and we introduce a shape regular tessellation $\mathcal{T}_h$ of triangles. Due to the lack of experimental measures of the boundary datum $y_{meas}$, we make use of synthetic data, i.e., we simulate the direct problem via the Finite Element method, considering the presence of an ischemic region of prescribed geometry, and extract the value on the boundary of the domain. In order to avoid to incurr an inverse crime (i.e. the performance of the reconstruction algorithm are improved by the fact that the exact data are synthetically generated with the same numerical scheme adopted in the algorithm) we introduce a more refined mesh $\mathcal{T}_h^{ex}$ on which the exact problem is solved, and interpolate the resulting datum $y_{meas}$ on the mesh $\mathcal{T}_h$. The section is organised as follows: in Section \[first\_results\], we describe the performance of Algorithm \[al:POP\] for the minimization of the phase-field relaxed functional , showing that different and rather complicated geometry of inclusion can be satisfactorily reconstructed. In Section \[settings\] we test the robustness of the reconstruction when some of the main parameters involved in the algorithm are modified. Moreover, particular attention is given to the use of a mesh-adaptative strategy. Parabolic Obstacle Problem: main results {#first_results} ---------------------------------------- In the following test cases, we applied Algorithm \[al:POP\] in order to reconstruct inclusions of different geometries, in order to investigate the effectiveness of the introduced strategy. We used the same computational mesh $\mathcal{T}_h$ (mesh size $h_{max} = 0.04$, nearly $6000$ elements) for the numerical solution of the boundary value problems involved in the procedure, whereas the mesh $\mathcal{T}_h^{ex}$ for the generation of each different synthetic datum associated to the different inclusions is specifically refined along the boundary of the respective ischemic region. According to Remark \[multi\], we make use of $N_f = 2$ different measurements, associated to the source terms $f_1(x,y) = x$ and $f_2(x,y) = y$. The main parameters for all the simulations lie in the ranges reported in Table \[tab:param\]. We make use of the same relationship between $\varepsilon$ and $\tau$ as in [@art:deckelnick]. $\alpha$ $\varepsilon$ $\tau$ $tol_{POP}$ ------------------------ --------------- ------------------------------- ------------- $10^{-4} \div 10^{-3}$ $1/(8 \pi)$ $(0.01 \div 0.1)/\varepsilon$ $10^{-4}$ : Range of the main parameters[]{data-label="tab:param"} The initial guess for each simulation is $u_0 \equiv 0$. In Figure \[fig:circle\] we report some of the iterations of Algorithm \[al:POP\] for the reconstruction of a circular inclusion ($\alpha = 0.0001$, $\tau = 0.01/\varepsilon$). The boundary $\partial \omega$ is marked with a black line, which is superimposed to the contour plot of the approximation of the indicator function $u_h^n$ at different timesteps $n$. The algorithm converged after $N_{tot} = 568$ iterations, corresponding to a final (fictitious) time $T_{tot}= 1427.54$. In Figure \[fig:many\] we investigate the ability of the algorithm to reconstruct inclusions of rather complicated geometry. For each test case, we show the contour plot of the final iteration of the reconstruction (the total number of iterations $N$ and the final time $T$ are reported in the caption), and the boundary of the exact inclusion is overlaid in black line. Moreover, each result is equipped with the graphic (in semilogarithmic scale) of the evolution of the cost functional $J_\varepsilon$, split into the components $J_{PDE}(u) = \half \norm{L^2(\partial \Omega)}{S(u)-y_{meas}}^2$ and $J_{regularization}(u) = \alpha \varepsilon \norm{L^2(\Omega)}{\nabla u}^2 + \frac{\alpha}{\varepsilon} \int_{\Omega}u(1-u)$. \ The reported results consist in approximations of minimizers of $J_\varepsilon$ in $\mathcal{K}$: they are smooth function and range between $0$ and $1$. They show large regions in which they attain the limit values $0$ and $1$, and a region of *diffuse interface* between them, whose thickness is about $\varepsilon/2$. As Figures \[fig:circle\] and \[fig:many\] show, the algorithm is capable of reconstructing inclusion of rather complicated geometry. The identification of smooth inclusion is performed with higher precision, whereas it seems that the accuracy is low in presence of sharp corners. We point out that we don’t need to have any *a priori* knowledge on the topology of the inclusion $\omega$, i.e., the number of connected components is correctly identified. Parabolic Obstacle Problem: setting of the parameters {#settings} ----------------------------------------------------- This section is devoted to the description of the performance of Algorithm \[al:POP\] when some of the main parameters and settings are perturbed. In particular, we start assessing that the final result of the reconstruction is independent of the initial guess imposed as a starting point of the Parabolic Obstacle problem. In Figure \[fig:starting\] we compare the behaviour of the algorithm applied to the reconstruction of a circular inclusion (the same as in Figure \[fig:circle\]), where we impose a different initial datum with respect to the constant zero function. In the first experiment, we start from an initial datum which is the indicator function of an arbitrarily chosen region. In the second one, we impose as a starting point the indicator function of a sublevel of the *topological gradient* of the cost functional $J$. As investigated in [@art:BMR], the topological gradient is a powerful tool for the detection of small-size inclusions, which yield a small perturbation in the cost functional with respect to the background (unperturbed) case. The position of a small inclusion is easily identified by searching for the point where the topological gradient of $J$ attains its (negative) minimum. As the information given by the topological gradient $G$ has shown to be remarkable even in the case of large-size inclusions (see, e.g., [@art:BCCMR], [@art:rapun]), we take advantage of it by computing $G$ (see Theorem 3.1 in [@art:BMR]), setting a threshold $G_{thr}$ and defining $u_0 = \chi_{\{G \leq G_{thr}\}}$. \ The results reported in Figure \[fig:starting\] show the starting point of the algorithm, an intermediate iteration and the final reconstruction. In both cases we set $\alpha = 0.001$, $\varepsilon = 1/(8 \pi)$ and $\tau = 0.1/\varepsilon$.We underline that the result in each case is similar to the one depicted in Figure \[fig:circle\], but through the second strategy it was possible to perform a smaller number of iterations. Another interesting investigation is the comparison of the results obtained when perturbing the relaxation parameter $\varepsilon$. In Figure \[fig:eps\] we report the final reconstruction of an ellipse-shaped inclusion when setting $\varepsilon = \frac{1}{4\pi}, \frac{1}{8\pi}, \frac{1}{8\pi}$. As expected, it is possible to remark that the thickness of the diffuse interface region decreases as $\varepsilon$ decreases. Nevertheless, one must take into account the size of the computational mesh $\mathcal{T}_h$: in the last test, the thickness of the region in which the final iteration $u_h^{N_tot}$ increases from $0$ to $1$ is of the same order of magnitude as $h_{max}$. This is rather likely the reason why the edge of the reconstructed inclusion appears to be irregular and jagged. A natural strategy to avoid the problem would be to make use of a finer mesh, e.g., ensuring that $h_{max}<\varepsilon/10$; however, that could result in an extremely high computational effort. It is possible to overcome this drawback by introducing an adaptive mesh refinement strategy, i.e., by locally refining the mesh close to the region of the detected edges. In Figure \[fig:adapt\] we compare the result obtained when approximating a rectangular and a circular inclusion with $\varepsilon = \frac{1}{16\pi}$ on the reference mesh or through a process of mesh adaptation. We invoked a goal-oriented mesh adaptation algorithm each $N_{adapt} = 50$ iterations, requiring for a higher refinement of the grid in proximity of higher values of $|\nabla u_h^n|$ and for a lower refinement in the regions where $u_h^n$ is approximatively constant. This allows to have more precise reconstruction even for small $\varepsilon$, almost without increasing the global number of elements of the mesh. In Figure \[fig:adapt\], we also report the final configuration of the refined computational mesh. \ Comparison with the Shape Derivative approach {#sec:shape} ============================================= In the previous sections, we have analyzed in detail the phase-field relaxation of the minimization problem expressed in . We now aim at describing the relationship between this method and the Shape Derivative approach, which has become a very common strategy to tackle the reconstruction of discontinuous coefficients. The algorithm based on the shape derivative consists in updating the shape of the inclusion to be reconstructed by perturbing its boundary along the directions of the vector field which causes the greatest descent of the cost functional, that can be deduced by computing the shape derivative of the functional itself. In this section, we first theoretically investigate the relationship between the shape derivative of the cost functional $J_{reg}$ and the Fréchet derivative of $\Je$ and then report a comparison between the numerical results of the two algorithms in a set of benchmark cases. Sharp interface limit of the Optimality Conditions {#sec:sharplim} -------------------------------------------------- In order to study the relationship between the optimality conditions in the phase-field approach and the ones derived in the sharp case, we follow an analogous approach as in [@art:garcke2016]. First of all, in Proposition \[prop:sharpOC\] we introduce the necessary optimality condition for the sharp problem , taking advantage of the computation of the material derivative of the cost functional. We then define in Proposition \[prop:pfOC\] similar optimality conditions for the relaxed problem , which are related but not equivalent to the one stated in - through the Fréchet derivative. In Proposition \[prop:SIL\] we finally assess the convergence of the phase-field optimality condition to the sharp one when $\varepsilon \rightarrow 0$. For the sake of simplicity, in this section we will refer to $J_{reg}$ as $J$. Consider the minimization problem (as in ): $$\argmin_{u \in X_{0,1}} J(u); \quad J(u) = \half \norm{L^2(\partial\Omega)}{S(u) - y_{meas}}^2 + \alpha TV(u). \label{eq:minreg2}$$ Since $u \in X_{0,1}$ implies that $u = \chi_\omega$, being $\omega$ a finite-perimeter subset of $\Omega$, we can perturb $u$ by means of a vector field $\phi_t: \Omega \rightarrow \R^2$, $\phi_t(x) = x + tV(x)$, being $$V \in C^1(\Omega) \text{ s.t. } V(x) = 0 \text{ in }\Omega^{d_0} = \{x \in \Omega \ s.t. \ dist(x,\partial \Omega) \leq d_0\}. \label{eq:V}$$ Consider the family of functions $\{u_t\}$: $u_t = u \circ \phi_t^{-1}$: we can compute the shape derivative of the functional $J$ in $u$ along the direction $V$ (see [@book:DZ2011]) as $$DJ(u)[V] := \lim_{t \rightarrow 0} \frac{J(u_t)-J(u)}{t}, \label{eq:shapedersharpDef}$$ where $J(u_t)$ is the cost functional evaluated in the deformed domain $\Omega_t=\phi_t(\Omega)$ but, according to , $\Omega_t$ and $\Omega$ are the same set, thus we do not adopt a different notation. We prove the following result: If $u$ is a solution of and $f \in L^2(\Omega)$ satisfies the hypotheses in Proposition \[prop:geq\] or \[prop:geq2\], then $$DJ(u)[V] = 0 \qquad \textit{ for all the smooth vector fields $V$}, \label{eq:OCmatder}$$ The shape derivative is given by: $$DJ(u)[V] = \int_{\partial \Omega}(S(u) - y_{meas})\dot{S}(u)[V] + \int_\Omega(div V - DV \normal \cdot \normal)d|Du|, \label{eq:matdersharp}$$ where $d|Du|=\delta_{\partial \omega} dx$, $\normal$ is the generalized unit normal vector (see [@book:giusti]) and $\dot{S}(u)[V] =: \dot{S}$, the material derivative of the solution map, solves $$\begin{aligned} \int_\Omega a(u) \nabla \dot{S} \cdot \nabla v + \int_\Omega b(u)3 S(u)^2 \dot{S} v = &-\int_\Omega a(u)\mathcal A \nabla S(u) \cdot \nabla v -\int_\Omega b(u)S(u)^3 v div V + \\ &\int_\Omega div(fV)v \qquad \forall v \in H^1(\Omega), \label{eq:matder} \end{aligned}$$ being $\mathcal{A} = div V - (DV + DV^T)$. \[prop:sharpOC\] We start by deriving the formula of the material derivative of the solution map. Define $S_0 = S(u)$ and $S_t: \Omega \rightarrow \R$, $S_t = S(u_t) \circ \phi_t$. Then, applying the change of variables induced by the map $\phi_t$, it holds that $$\int_{\Omega} a(u) A(t) \nabla S_t \cdot \nabla v + \int_{\Omega} b(u)S_t^3 v |det D\phi_t| = \int_{\Omega} (f\cdot \phi_t) v |det D\phi_t| \qquad \forall v \in H^1(\Omega), \label{eq:2}$$ where $A(t) = D\phi_t^{-T} D\phi_t^{-1} |det D\phi_t|$. By computation, $$\frac{d}{dt} A(t) = \mathcal{A} = (div V)I - (DV^t+DV) \qquad \text{and} \qquad \frac{d}{dt} |det D\phi_t| = div V.$$ Subtract from and divide by $t$: then $w_t = \frac{S_t - S_0}{t}$ is the solution of $$\begin{aligned} \int_\Omega a(u) A(t) \nabla w_t \cdot \nabla v &+ \int_\Omega b(u)q_t w_t v |det(D\phi_t)| = -\int_{\Omega}a(u) \frac{A(t)-I}{t} \nabla S_0 \cdot \nabla v \\ &- \int_{\Omega} \frac{|det(D\phi_t)|-1}{t}b(u)S_0^3v + \int_{\Omega} \frac{1}{t}(f\circ \phi_t) v |det(D\phi_t)| -\int_{\Omega} \frac{1}{t} fv \qquad \forall v \in H^1(\Omega), \end{aligned} \label{eq:4}$$ where the norm of the right-hand side in the dual space of $H^1(\Omega)$ is bounded by $$\begin{aligned} \norm{L^{\infty}(\Omega)}{\frac{A-I}{t}}&\norm{H^1(\Omega)}{S_0} + \norm{L^{\infty}(\Omega)}{\frac{|det(D\phi_t)|-1}{t}}\norm{H^1(\Omega)}{S_0} \\&+ \norm{L^{\infty}(\Omega)}{\frac{|det(D\phi_t)|-1}{t}}\norm{L^2(\Omega)}{f} + C(\norm{C(\Omega)}{V})\norm{H^1(\Omega)}{f} \leq C_F, \end{aligned}$$ being $C_F$ independent of $t$. Moreover, the matrix $A(t)$ is symmetric positive definite: $(A(t) y) \cdot y\geq \half \norm{}{y}^2$ $\forall y \in \R^2, \forall t$. Together with the property that $q_t = u_t^2 + u_t u + u^2 \geq \frac{3}{4}u^2$, and thanks to Proposition \[prop:geq\] and to the Poincaré inequality in Lemma \[Lemma1\], $$\norm{H^1}{w_t}^2 \leq C k\norm{L^2}{\nabla w_t}^2 +\frac{3}{4}Q \norm{L^2(\Omega^*)}{w_t}^2 \leq C_F\norm{H^1}{w_t}.$$ Thus, $\norm{H^1}{w_t}$ is bounded independently of $t$, from which it follows that $\norm{H^1(\Omega)}{S_t - S_0} \leq C t$ and that every sequence $ \{w_n\} = \left\{ w_{t_n}, \ t_n \rightarrow 0 \right\}$ is bounded in $H^1(\Omega)$, thus $w_t \xrightharpoonup{H^1} w \in H^1(\Omega)$. We aim at proving that $w$ is also the limit of $w_t$ in the strong convergence, which entails that $$\dot{S}(u)[V]:= \lim_{t\rightarrow 0} \frac{S_t - S_0}{t} = w.$$ First of all, we show that $w$ is the solution of problem . It follows from , since $q_t w_t = \frac{1}{t}(S_t^3 - S_0^3) = \frac{1}{t}((S_0 +t w_t)^3 - S_0^3) = 3 S_0^2 w_t + 3t S_0 w_t^2 + t^2 w_t^3$, that $$\begin{aligned} \int_{\Omega} a(u) A(t) \nabla w_t \cdot \nabla v &+ \int_{\Omega} b(u) 3 S_0^2 w_t v |det D\phi_t| = -\int_{\Omega}a(u) \frac{A(t)-I}{t} \nabla S_0 \cdot \nabla v \\ &-\int_{\Omega} \frac{|det D\phi_t|-1}{t} b(u)S_0^3v - \int_{\Omega} b(u) 3t S_0 w_t^2 v |det D\phi_t| - \int_{\Omega} b(u) t^2 w_t^3 v |det D\phi_t|\\ &+ \int_{\Omega} (f \circ \phi_t)\frac{|det D\phi_t|-1}{t} v -\int_{\Omega}\frac{(f\circ \phi_t) - f}{t}v \qquad \forall v \in H^1(\Omega). \end{aligned} \label{eq:4.5}$$ Taking the limit as $t \rightarrow 0$ and by the weak convergence of $w_t$ in $H^1$, we recover the same expression as in . One may eventually show that $w_t\xrightarrow{H^1} w$. In order to do this we start proving that $$\int_\Omega a(u) A(t)|\nabla w_t|^2 + \int_\Omega b(u)|det D\phi_t|3S_0^2w_t^2 \rightarrow \int_\Omega a(u) |\nabla w|^2 + \int_\Omega b(u)3S_0^2w^2. \label{eq:6}$$ Indeed, take and substitute $v = w_t$: using the weak convergence of $w_t$ in the right-hand side, we obtain that $$\begin{aligned} \int_\Omega a(u) A(t)|\nabla w_t|^2 &+ \int_\Omega b(u)|det D\phi_t|3S_0^2w_t^2 \rightarrow -\int_{\Omega}a(u) \mathcal{A} \nabla S_0 \cdot \nabla w -\int_{\Omega} divV \ b(u)S_0^3w \\ &+ \int_{\Omega} f w\ divV - \int_{\Omega}\nabla f \cdot V w \overset{\eqref{eq:matder}}{=} \int_{\Omega} a(u) |\nabla w|^2 + \int_{\Omega} b(u) 3 S_0^2 w^2. \end{aligned}$$ We then compute: $$\begin{aligned} \int_\Omega a(u)& A(t) |\nabla(w_t - w)|^2 + \int_\Omega b(u) 3 S_0^2 (w_t - w)^2 |det D\phi_t| = \\ & \int_\Omega a(u) A(t) |\nabla w_t|^2 + \int_\Omega a(u) A(t) |\nabla w|^2 - 2 \int_\Omega a(u) A(t) \nabla w_t \cdot \nabla w \\ & +\int_\Omega b(u)3S_0^2 w_t^2 |det D\phi_t| + \int_\Omega b(u) 3S_0^2 w^2 |det D\phi_t|- 2\int_\Omega b(u)3S_0^2w_t w |det D\phi_t|. \end{aligned} \label{eq:eq7}$$ Using , the convergence of $A$ to $I$ and of $|det D\phi_t|$ to $1$, and the fact that $w_t \xrightharpoonup{H^1} w$, we derive that $$\int_\Omega a(u) |\nabla(w_t - w)|^2 + \int_\Omega b(u) 3S_0^2(w_t - w)^2 \rightarrow 0$$ A combination of the Proposition \[prop:geq\] and of the Poincarè inequality in Lemma \[Lemma1\] allows to conclude that also $\norm{H^1}{w_t - w} \rightarrow 0$. We now prove the necessary optimality conditions for the optimization problem . The derivative of the quadratic part of the cost functional $J$ can be easily computed by means of the material derivative of the solution map: $$\begin{aligned} \lim_{t \rightarrow 0} \half &\int_{\partial \Omega}\frac{(S(u_t) - y_{meas})^2 |det(D\phi_t)| - (S_0-y_{meas})^2}{t} \qquad \textit{(since $S(u_t) = S_t$ on $\partial \Omega$)}\\ &= \lim_{t \rightarrow 0} \half \int_{\partial \Omega}(S_t - y_{meas})^2 \frac{|det(D\phi_t)| - 1}{t} + \lim_{t \rightarrow 0} \half \int_{\partial \Omega}\frac{(S_t - y_{meas})^2 - (S_0-y_{meas})^2}{t} \\&= \half \int_{\partial \Omega}(S_0-y_{meas})^2 div V + \int_{\partial \Omega} \dot{S}(u)[V](S_0-y_{meas}), \end{aligned} \label{eq:7}$$ and the first integral in the latter expression vanishes since $V=0$ on $\Omega_{d_0}$. On the other hand, using Lemma 10.1 of [@book:giusti] and the remark 10.2, we recover the expression for the derivative of the Total Variation of $u$, which is the same reported in . The optimality conditions reported in are, at the best of our knowledge, the most general result which can be obtained in this case, i.e. by simply assuming that $u =\chi_\omega$ and $\omega$ is a set of finite perimeter. We point out that, assuming more *a priori* knowledge on the $u$, it is possible to recover from the expression of the *shape derivative* of the cost functional $J$. The following proposition can be rigorously proved by means of an analogous argument as in [@art:abfkl], except for the derivative of the perimeter penalization, which can be found in Section 9.4.3 in [@book:DZ2011]. Suppose that $\omega \subset \Omega$ is open, connected, well separated from the boundary $\partial \Omega$ and regular (at least of class $C^2$), and $u = \chi_\omega$. Then, the expression of the shape derivative of the cost functional $J$ along a smooth vector field $V$ is: $$DJ(u)[V] = \int_{\partial \omega} \left[ (1-k)\left( \nabla_{\tau}S(u) \cdot \nabla_\tau w + \frac{1}{k} \nabla_\normal S(u)^e\cdot \nabla_\normal w^e \right) + S(u)^3 w + h \right] V\cdot \normal \qquad \forall V, \label{eq:shapeder}$$ where $w$ is the solution of the adjoint problem (see ). The gradients $\nabla S(u)$ and $\nabla w$ are decomposed in the normal and tangential component with respect to the boundary $\partial \omega$, and due to the transmission condition of the direct problem their normal components are discontinuos across $\partial \omega$: the valued assumed in $\Omega \setminus \omega$ is marked as $\nabla_\normal S(u)^e$. The term $h$ is instead the mean curvature of the boundary. \[prop:shapeder\] For the sake of completeness, we point out that the latter result can be easily generalized to the case in which $\omega$ is the union of $N_c$ disjoint, well separated, components, each of them satisfying the expressed hypotheses. Thanks to the results recently obtained in [@art:bfv], we expect formula to be valid also under milder assumption, in particular for polygons. We aim at demonstrating that the expression of the shape derivative reported in is the limit, as $\varepsilon \rightarrow 0$, of a suitable derivative of the relaxed cost functional $\Je$. In order to accomplish this result, we need to introduce necessary optimality conditions for the relaxed problem $\eqref{eq:minrel}$ which are different from the ones reported in Proposition \[prop:optcond\] and can be derived by the same technique as in Proposition \[prop:sharpOC\] as shown in the following result. If $u_\varepsilon$ is a solution of , then $$DJ_\varepsilon(u_\varepsilon)[V] = 0 \qquad \textit{ for all the smooth vector fields $V$}, \label{eq:pfOC}$$ The expression of the derivative is given by: $$\begin{aligned} DJ_\varepsilon(u_\varepsilon)[V] = &\int_{\partial \Omega}(S(u_\varepsilon) - y_{meas})\dot{S}(u_\varepsilon)[V] + \alpha \varepsilon \int_\Omega |\nabla \ue|^2 div V \\ &- 2\alpha \varepsilon \int_\Omega DV \nabla \ue \cdot \nabla \ue + \frac{\alpha}{\varepsilon} \int_{\Omega} \ue(1-\ue)div V \end{aligned} \label{eq:shapederpf}$$ where $\dot{S}(u_\varepsilon)[V]$ solves the same problem as in , replacing $u$ with $u_\varepsilon$. \[prop:pfOC\] The same strategy as in the proof of Proposition \[prop:sharpOC\] can be adapted to compute $\dot{S}(\ue)[V]$ and the derivative of the first term of the cost functional. We now derive with the same computational rules the relaxed penalization term. Recall $$F_\varepsilon(\ue) = \alpha \varepsilon \int_{\Omega}|\nabla \ue|^2 + \frac{\alpha}{\varepsilon} \int_\Omega \psi(\ue),$$ being $\psi: \R \rightarrow \R,$ $\psi(x) = x(1-x)$. After the deformation from $u_\varepsilon$ to $u_\varepsilon \circ \phi_t^{-1}$ and applying the change of variables induced by $\phi_t$, $$F_\varepsilon(\ue \circ \phi_t^{-1}) = \alpha \varepsilon \int_{\Omega} A(t) \nabla \ue \cdot \nabla \ue + \frac{\alpha}{\varepsilon} \int_{\Omega} \psi \circ \ue \circ \phi_t^{-1}.$$ Hence, $$\begin{aligned} \dot{F}_\varepsilon(u_\varepsilon)[V] &= \lim_{t \rightarrow 0}\frac{F_\varepsilon(\ue \circ \phi_t^{-1}) - F_\varepsilon(\ue)}{t} = \alpha \varepsilon \int_{\Omega} \mathcal{A}\nabla \ue \cdot \nabla \ue + \alpha \varepsilon \frac{\alpha}{\varepsilon}\int_{\Omega} \psi(\ue)div V = \\ &= \alpha \varepsilon\int_\Omega |\nabla \ue|^2 div V - \alpha \varepsilon \int_\Omega (DV + DV^T)\nabla \ue \cdot \nabla \ue + \frac{\alpha}{\varepsilon}\int_\Omega \ue(1-\ue)div V, \end{aligned}$$ which is the same expression as in , since $DV^T \nabla v \cdot \nabla v = DV \nabla v \cdot \nabla v$. We point out that the optimality conditions deduced in the latter proposition are not equivalent to the ones expressed in Proposition \[prop:optcond\] via the Fréchet derivative of $\Je$. Nevertheless, if $u_\varepsilon$ satisfies -, then it also satisfies (it is sufficient to consider in $v = \ue \circ \phi_t^{-1}$, which belongs to $\mathcal{K}$ thanks to the regularity of $V$), whereas the contrary is not valid in general. In particular, due to the regularity of the perturbation fields $V$, the optimality conditions do not take into account possible topological changes of the inclusion: for example, the number of connected components of $\omega$ cannot change. We remark that this holds also for the optimality conditions for the sharp problem, and consists in a limitation for the effectiveness of the reconstruction via a shape derivative approach: the initial guess of the reconstruction algorithm and the exact inclusion must be diffeomorphic. We are now able to show the sharp interface limit of the expression of the shape derivative of the relaxed cost functional $\Je$ as $\varepsilon \rightarrow 0$, which is done in the following proposition. Consider a family ${\bar{u}_\varepsilon}$ s.t. $\bar{u}_\varepsilon \in \mathcal{K}$ $\forall \varepsilon > 0$ and $\bar{u}_\varepsilon \xrightarrow{L^1} \bar{u} \in BV(\Omega)$ as $\varepsilon \rightarrow 0$. Then, $$DJ_\varepsilon(\bar{u}_\varepsilon)[V] \rightarrow DJ(\bar{u})[V] \qquad \textit{for every smooth vector field $V$.}$$ \[prop:SIL\] We follow a similar argument as in the proof of [@art:garcke2016 Theorem 21]. Thanks to Proposition \[prop:continuity\], $\bar{\ue} \xrightarrow{L^1} \bar{u}$ $\Rightarrow$ $S(\bar{\ue}) \xrightarrow{H^1} S(\bar{u})$. Also $\dot{S}(\bar{\ue})[V] \xrightarrow{H^1} \dot{S}(\bar{u})[V]$: the proof is done by subtracting the equations of which $\dot{S}(\bar{\ue})[V]$ and $\dot{S}(\bar{u})[V]$ and verifying that the norm of their difference is controlled by the norm of $S(\bar{\ue})-S(\bar{u})$ in $H^1(\Omega)$. Thanks to these results, surely $$\int_\Omega (S(\ue)-y_{meas})\dot{S}(\bar{\ue})[V] \rightarrow \int_\Omega (S(u) -y_{meas})\dot{S}(\bar{u})[V].$$ Eventually, the convergence $$\alpha \varepsilon \int_\Omega |\nabla \bar{\ue}|^2 div V - 2\alpha \varepsilon \int_\Omega DV \nabla \bar{\ue} \cdot \nabla \bar{\ue} + \frac{\alpha}{\varepsilon} \int_{\Omega} \bar{\ue}(1-\bar{\ue})div V \rightarrow \int_\Omega(div V - DV \normal \cdot \normal)d|D\bar{u}|$$ is proved in [@art:garcke2008], Theorem 4.2 (see also annotations in [@art:garcke2016], proof of Theorem 21). In particular, we point out that this implies, together with Proposition \[prop:conv2\], that the expression of the optimality condition for the phase field problem converges, as $\varepsilon \rightarrow 0$, to the one in the sharp case. Comparison with the shape derivative algorithm {#shapeder} ---------------------------------------------- In this section, we report some results of the application of the algorithm based on the shape derivative. In the implementation, we take advantage of the Finite Element method to solve the direct and adjoint problems and compute the shape gradient as in . We consider an initial guess for the inclusion (in all the simulations reported, the initial guess is a disc centered in the origin with radius $0.02$) and discretize its boundary with a finite number of points, which always coincide with vertices of the numerical mesh. We iteratively perturb the inclusion by moving the boundary with a vector field $V$ which is the projection in the finite element space $V_h \times V_h$ of the shape gradient reported in and $\normal_\omega$ the external normal vector of $\partial \omega$ (see e.g. [@art:marco] for more details). After the descent direction is determined, a backtracking scheme is implemented (see [@book:NocedalWright]), in order to guarantee the decrease of the cost functional $J$ at each iteration. As in the case of Algorithm \[al:POP\], we start from the initial guess $u^0 \equiv 0$ and take advantage of $N_f = 2$ measurements, associated to the same source terms. The main parameters of this set of simulations are reported in Table \[tab:shapeder\]. $\alpha$ $max step$ $tol$ ----------- ------------ ----------- $10^{-3}$ $10$ $10^{-6}$ : Values of the main parameters[]{data-label="tab:shapeder"} In Figure \[fig:uzawa\] we report the results of the reconstruction with the shape gradient algorithm compared to the ones of the Parabolic Obstacle problem (with $\varepsilon = \frac{1}{16\pi}$ and with mesh adaptation). Each result is endowed with a plot of the evolution of the cost functional throughout time (in particular, of $J_{PDE}(u) = \half \norm{L^2(\partial \Omega)}{S(u) - y_{meas}}$ \ The reconstruction achieved by the shape gradient algorithm seems to be as accurate as the phase-field one. The sharp method seems to be less expensive in term of computational cost, and it involves a smaller number of iterations. Nevertheless, it requires the knowledge of remarkable *a priori* information, i.e. the topology of the inclusion. Appendix {#appendix .unnumbered} ======== In the proof of various proposition we have used the following generalized Poincarè inequality: Let $\Omega^* \subset \Omega$ be s.t. $|\Omega^*| \neq 0$. Then $\exists C > 0, C = C(\Omega)$ s.t., $\forall u \in H^1(\Omega)$, $$\norm{H^1(\Omega)}{u}^2 \leq C \left( \norm{L^2(\Omega)}{\nabla u}^2 + \norm{L^2(\Omega^*)}{u}^2 \right). \label{eq:Poincare}$$ \[Lemma1\] The proof of the Lemma is given in the Appendix of [@art:bcmp] and easily follows by Theorem 8.11 in [@liebloss]. Thanks to Lemma \[Lemma1\], we can prove the following well-posedness result for the direct problem. We remark that a similar analysis was performed in [@art:bcmp], but here we extend the result for the case of inclusions which have the property of being finite-perimeter sets. Consider $f \in \left(H^1(\Omega)\right)^*$ and a function $u \in BV(\Omega;[0,1])$ s.t. $u$ is not (a.e.) equal to $1$. Then there exists an unique solution $S(u)\in H^1(\Omega)$ of $$\int_\Omega a(u) \nabla S(u) \cdot \nabla v + \int_\Omega b(u)S(u)^3v = \int_\Omega fv \qquad \forall v \in H^1(\Omega),$$ where $a(u) = 1-(1-k)u$ and $b(u) = 1-u$. \[prop:wellpos\] We use the Minty-Browder theorem (see, e.g., Theorem 5.16 in [@book:brezis]), introducing (for a fixed $u$) the operator $T:H^1(\Omega) \rightarrow \left(H^1(\Omega)\right)^*$ s.t. $$\langle T(S),v \rangle_* = \int_\Omega a(u)\nabla S \cdot \nabla v + \int_\Omega b(u) S^3 v.$$ We can easily verify that the nonlinear operator $T$ is continuous, coercive and monotone. - Continuity: we indeed prove that $T$ is locally Lipschitz continuous with respect to $S$. $$\begin{aligned} |\langle T(S)-T(S_0), v \rangle_*| &= \left| \int_\Omega a(u) \nabla (S-S_0) \cdot \nabla v + \int_\Omega b(u)(S-S_0)q\right| \textit{ (being $q = S^2+S S_0 + S_0^2$)} \\ & \leq \norm{L^2}{\nabla (S-S_0)}\norm{L^2}{\nabla v} + \norm{L^6}{S-S_0}\norm{L^3}{ q}\norm{L^2}{v} \end{aligned}$$ If $S$ and $S_0$ belong to a bounded subset of $H^1(\Omega)$, then (thanks to the Sobolev Embedding of $H^1(\Omega)$ in $L^6(\Omega)$) we can assess that $\norm{L^3}{q}\leq M$ and moreover $\exists K>0$ s.t. $$|\langle T(S)-T(S_0), v \rangle_*|\leq K \norm{H^1}{S-S_0} \norm{H^1}{v} \qquad \forall v \in H^1(\Omega).$$ - Coercivity: we show that $\langle T(S),S \rangle_* \rightarrow +\infty $ as $\norm{H^1(\Omega)}{S} \rightarrow + \infty$. Since $u$ is not identically equal to $1$, $\exists Q>0$ and $\Omega^*: |\Omega^*|\neq 0$ s.t. $b(u) = 1-u \geq Q$ a.e. in $\Omega^*$. Then, $$\begin{aligned} \langle T(S),S \rangle_* &\geq k \int_\Omega |\nabla S|^2 + Q\int_{\Omega^*}S^4 \geq k \norm{L^2(\Omega)}{\nabla S}^2 + \frac{Q}{|\Omega|}\norm{L^2(\Omega^*)}{S}^4 \\ &=k \left( \norm{L^2(\Omega)}{\nabla S}^2 + \norm{L^2(\Omega^*)}{S}^2 \right) + R, \end{aligned}$$ where $R = \frac{Q}{|\Omega|}\norm{L^2(\Omega^*)}{S}^4 - k \norm{L^2(\Omega^*)}{S}^2$ can be bounded by below independently of $S$ by considering that $$\left( A\norm{L^2(\Omega^*)}{S}^2 - B \right)^2 \geq 0 \Rightarrow A^2 \norm{L^2(\Omega^*)}{S}^4 - 2AB \norm{L^2(\Omega^*)}{S}^2 \geq - B^2,$$ which implies (chosen $A=\sqrt{\frac{Q}{|\Omega|}}$ and $B = \frac{k \sqrt{|\Omega|}}{2 \sqrt{Q}}$) that $R \geq -\frac{k^2 |\Omega|}{4Q}$. Together with the Poincarè inequality in Lemma \[Lemma1\], we conclude that $$\langle T(S),S \rangle_* \geq \frac{k}{C} \norm{H^1(\Omega)}{S}^2 - \frac{k^2 |\Omega|}{4Q}.$$ - (Strict) monotonicity: we claim that $\langle T(S) - T(R), S-R \rangle_* \geq 0$ and $\langle T(S) - T(R), S-R \rangle_* = 0 \Leftrightarrow S = R$. Indeed, $$\langle T(S) - T(R), S-R \rangle_* \geq \int_\Omega k |\nabla(S-R)|^2 + Q \int_{\Omega^*} (S^2+SR+R^2)(S-R)^2 \geq 0.$$ Moreover, since $S^2+SR+R^2 \geq \frac{1}{4}(S-R)^2$, $$\langle T(S) - T(R), S-R \rangle_* = 0 \Rightarrow \norm{L^2(\Omega)}{\nabla(S-R)}=0 \text{ and } \int_{\Omega^*} (S-R)^4 = 0,$$ and from the latter equality it follows that $S=R$ a.e. in $\Omega^*$, hence also $\norm{L^2(\Omega^*)}{S-R}=0$, and thanks to the Poincaré inequality in Lemma \[Lemma1\], $\norm{H^1(\Omega)}{S-R}=0$. Finally, we prove an estimate which occurs many times in the proof of various results. Suppose that $f\in L^2(\Omega)$ s.t. $\int_\Omega f \neq 0$. Consider $S(u)$ the solution of problem associated to $u \in BV(\Omega;[0,1])$, $u$ not identically equal to $1$. Then, there exists $\Omega^*$ and $Q>0$ s.t. $|\Omega^*|\neq 0$ and $$b(u)S(u)^2 \geq Q \qquad \text{ a.e. in }\Omega^*$$ \[prop:geq\] By contraddiction, suppose the opposite of the thesis: $b(u)S(u)^2 = 0$ a.e. in $\Omega$. Then, this would imply that $\int_\Omega b(u) S(u)^3 = 0$, and then it would hold that $$\int_\Omega a(u)\nabla S(u)\cdot \nabla v = \int_\Omega f v \qquad \forall v \in H^1(\Omega).$$ Taking $v = const.$ we obtain that $\int_\Omega f = 0$, which contraddicts the hypothesis. We remark that the previous result can be extended to class of functions $f$ satisfying more general hypotheses. If, for example, we restrict to the case of inclusions well separated from the boundary ($u = 0$ a.e. in $\Omega^{d_0}$, being $\Omega^{d_0} = \{x \in \Omega: dist(x,\partial \Omega) \leq d_0\}$, $d_0 > 0$), then it is sufficient to require $f \neq 0$ a.e in $\Omega^{d_0}$ to guarantee an estimate equivalent to the one of Proposition \[prop:geq\]. Suppose that $u \in BV(\Omega;[0,1])$ satisfies $u = 0$ a.e. in $\Omega^{d_0}$. If $f\in L^2(\Omega)$ does not vanish in $\Omega^{d_0}$then there exists $Q>0$ s.t. the solution $S(u)$ of satisfies $$b(u)S(u)^2 \geq Q \qquad \text{ a.e. in }\Omega^{d_0}$$ \[prop:geq2\] By contraddiction of the thesis, suppose $S(u) \equiv 0$ in $\Omega^{d_0}$ and recall $\Omega^{in} = \Omega \setminus \Omega^{d_0}$; then it holds $$\int_{\Omega^{in}}a(u) \nabla S(u)\cdot \nabla v + \int_{\Omega^{in}} b(u) S(u)^3 v = \int_{\Omega} fv \quad \forall v \in H^1(\Omega) \label{eq:geq2eq}$$ The space $H^1_0(\Omega^{d_0})$, obtained by closing the space of all the smooth function whose support is compactly contained in $\Omega^{d_0}$ with respect to the $H^1$ norm, is well defined; moreover $H^1_0(\Omega^{d_0})\subset H^1(\Omega)$. Hence, equation holds for all $v \in H^1_0(\Omega^{d_0})$, and this implies that $$\int_\Omega f v = 0 \qquad \forall v \in H^1_0(\Omega^{d_0}),$$ which eventually entails that $f=0$ a.e. in $\Omega^{d_0}$, that is a contraddiction with hypotheses. Acknowledgments {#acknowledgments .unnumbered} =============== E. Beretta and M. Verani thank the New York University in Abu Dhabi for its kind hospitality that permitted a further development of the present research. We acknowledge the use of the MATLAB library redbKIT [@redbKIT] for the numerical simulations presented in this work.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Deep neural networks (DNNs) have demonstrated impressive performance on many challenging machine learning tasks. However, DNNs are vulnerable to adversarial inputs generated by adding maliciously crafted perturbations to the benign inputs. As a growing number of attacks have been reported to generate adversarial inputs of varying sophistication, the defense-attack arms race has been accelerated. In this paper, we present [MODEF]{}, a cross-layer model diversity ensemble framework. [MODEF]{} intelligently combines unsupervised model denoising ensemble with supervised model verification ensemble by quantifying model diversity, aiming to boost the robustness of the target model against adversarial examples. Evaluated using eleven representative attacks on popular benchmark datasets, we show that [MODEF]{} achieves remarkable defense success rates, compared with existing defense methods, and provides a superior capability of repairing adversarial inputs and making correct predictions with high accuracy in the presence of black-box attacks.' author: - bibliography: - 'references.bib' title: | Denoising and Verification Cross-Layer Ensemble\ Against Black-box Adversarial Attacks --- adversarial deep learning, ensemble defense, ensemble diversity, robustness Introduction ============ The recent advances in deep neural networks (DNNs) have powered numerous applications in different domains due to their outstanding performance compared to traditional machine learning techniques. However, it has been shown that DNNs can be easily fooled by adversarial inputs [@goodfellow2014explaining], making them become a double-edged sword as the vulnerability of DNNs to adversarial attacks has posed serious threats to many security-critical applications, such as biometric authentication and autonomous driving. As a number of defenses are being proposed, more attacks of varying sophistication have been put forward, accelerating the defense-attack arms race. Some even argue that designing new attacks requires much less efforts than developing effective defenses. Thus, improving the robustness and defensibility against adversarial attacks is crucial. Adversarial examples are generated by maliciously perturbing benign examples sent to the target DNN model through querying its prediction API, aiming to fool and mislead the target model to misclassify by producing incorrect predictions randomly (untargeted attack) or purposefully (targeted attack). Given a $D$-dimensional benign example $\boldsymbol{x}\in\mathbb{R}^D$ and a $K$-class classification target model $\text{TM}:\mathbb{R}^D\rightarrow\mathbb{R}^K$ such that the prediction is given as $C_{\boldsymbol{x}}=\arg\max_{1\leq i\leq K}\text{TM}_i(\boldsymbol{x})$, the generation process of the adversarial example $\boldsymbol{x}'$ can be formulated as $$\label{eq:adv-perturb} \min||\boldsymbol{x} - \boldsymbol{x}'||_{p} \quad s.t.\;C_{\boldsymbol{x}'} = {y^*},\ C_{\boldsymbol{x}'} \ne C_{\boldsymbol{x}},$$ where $p$ is the distance metric and $y^{*}$ denotes the target class label for targeted attacks and any class other than the correct one for untargeted attacks. For image inputs, the distance metric $p$ can be $0$, $2$, or $\infty$ representing $L_0$, $L_2$, and $L_\infty$ norms where $L_0$ norm counts the number of pixels of $\boldsymbol{x}$ that are changed, $L_2$ norm is the Euclidean distance between $\boldsymbol{x}$ and $\boldsymbol{x}'$, and $L_\infty$ norm denotes the maximum change to any pixel of $\boldsymbol{x}$. [**Related Work.**]{} Existing defense proposals are mostly attack-dependent and can be broadly divided into three categories. [*Adversarial training*]{} counters known attacks by retraining the target model using adversarial examples generated by each attack algorithm [@madry2017towards; @kurakin2016adversarial; @tramer2017ensemble]. [*Input transformation*]{} defenses apply noise reduction techniques to input data to reduce the sensitivity of the target model to small changes in input data due to adversarial perturbations. However, to distinguish adversarial examples from benign ones, they rely on finding a dataset-specific or attack-specific threshold [@ma2018characterizing; @meng2017magnet; @xu2017feature; @samangouei2018defense; @shen2017ape]. [*Gradient masking*]{} defenses aim to harden the process of generating adversarial examples by hiding the gradient information from attackers [@gu2014towards; @papernot2016distillation], such as distillation training techniques [@hinton2015distilling]. But they either significantly reduce the benign accuracy of the target model or are vulnerable to attack transferability [@papernot2017practical; @carlini2017towards]. In this paper, we present [MODEF]{}, a cross-layer MOdel Diversity Ensemble Framework by integrating unsupervised model denoising ensemble with supervised model verification ensemble, aiming to protect the target model against adversarial attacks. From a manifold learning perspective [@chapelle2009semi], natural high-dimensional data concentrate close to a nonlinear low-dimensional manifold, and adversarial attacks can be considered as a malicious process to drag benign examples away from the manifold where they concentrate. [MODEF]{} first exploits denoising autoencoders to learn such manifold and to map adversarial examples back to their corresponding benign form. Different from prior works using denoising autoencoders as an input preprocessing-based defense [@sahay2019combatting; @liao2018defense; @gu2014towards], [MODEF]{} leverages multiple denoisers to boost defensibility by joint force and quantifies model diversity to enable strategic teaming of denoisers with diverse denoising effects. To further improve the robustness and repair those adversarial examples escaped from the denoising autoencoders, the denoised example is sent to our second layer of model verification ensemble, which exploits the weak spots of attack transferability using a team of failure-independent models, for prediction verification and repairing. We further enhance the defensibility of [MODEF]{} with defense structure randomization by creating a pool of models and enabling strategic randomized ensemble teaming. By promoting such uncertainty at runtime, it allows [MODEF]{} to combat adversarial attacks with higher robustness. In summary, [MODEF]{} advances existing works from three perspectives. First, it provides three model diversity ensemble defenses against adversarial examples: the model denoising ensemble, the model verification ensemble, and the denoising-verification cross-layer ensemble, each improves the previous layer by empowering the target model with higher robustness. Second, [MODEF]{} is by design attack-independent and can generalize well over attack algorithms. [MODEF]{} does not use dataset-specific or attack-specific magic parameters, such as the detection thresholds, to distinguish adversarial examples from benign ones. Furthermore, [MODEF]{} enables strategic defense structure randomization to improve its defensibility against adversarial examples and hardens black-box attacks [@papernot2017practical]. Extensive experiments on popular benchmark datasets with eleven representative attacks are used to validate the robustness of [MODEF]{}, showing that [MODEF]{} can significantly improve the robustness of a target DNN model against adversarial attacks. The remainder of this paper is organized as follows. We first briefly review the representative attacks and the benchmark datasets used in this paper in Section \[sec:attacks\]. Then, we introduce the Model Denoising Ensemble Defense in Section \[sec:dnn-denoising\] and the Model Verification Ensemble Defense and the Denoising-Verification Cross-Layer Ensemble Defense in Section \[sec:co-defense\], followed by experimental evaluation in Section \[sec:experimental-results\]. We conclude the paper in Section \[sec:conclusion\]. Adversarial Examples {#sec:attacks} ==================== Adversarial examples are generated over their corresponding benign examples by adding maliciously crafted perturbations that can fool the target model to misclassify during the model prediction (testing) phase. For the attack threat model, if the adversary can generate adversarial examples by only accessing the prediction API of the target model with no knowledge of the target model training process, we call such attack the black-box attack. In this paper, we focus on defense strategies against black-box attacks. We generate adversarial examples using seven representative attack algorithms: FGSM [@goodfellow2014explaining], BIM [@kurakin2016adversarial], CW$_{0}$, CW$_{2}$, CW$_{\infty}$ [@carlini2017towards], DeepFool [@moosavi2016deepfool], and JSMA [@papernot2016limitations] which cover a wide spectrum of techniques including both $L_\infty$, $L_2$, and $L_0$ distortions. We build those attacks on top of EvadeML-Zoo [@xu2017feature] using the same set of hyperparameters. For each targeted attack, we study two attack targets representing two ends of the targeted attack spectrum: the most-likely (ML) attack class in the prediction vector (i.e., $y^* = \arg\max_{1\leq i\leq K,i \ne C_{\boldsymbol{x}}} \text{TM}_i(\boldsymbol{x})$) and the least-likely (LL) attack class (i.e., $y^* = \arg\min_{1\leq i\leq K} \text{TM}_i(\boldsymbol{x})$). Thus we have a total of eleven different attacks. Table \[tab:attack-evaluation\] summarizes the evaluation of the attacks for two popular benchmark datasets: MNIST and CIFAR-10. MNIST consists of 70,000 gray-scale images of ten handwritten digits, each image is $28\times28\times1$ in size with 60,000 images for training and the remaining 10,000 images for testing. CIFAR-10 consists of 60,000 colorful images of 10 classes, each is $32\times32\times3$ in size. Similarly, for CIFAR-10, 50,000 images are used for training with the remaining 10,000 images for testing. For MNIST, the target model is a seven-layer CNN [@carlini2017towards] with an accuracy of $0.9943$. For CIFAR-10, the target model is a DenseNet [@huang2017densely] with an accuracy of $0.9484$. The first $100$ testing examples ($10$ per class) that are correctly classified by the target model are selected to generate adversarial examples. We exclude DeepFool (DF) from MNIST because the generated images are unrecognizable to humans. -- ------ ------------ ---------- --------- -------- -------- ---------- ---------------- ----------- ----------- **** **Attack** **Mode** **L$_\infty$** **L$_2$** **L$_0$** FGSM $0.003$ $0.46$ $0.46$ $0.9475$ $0.302$ $5.931$ $0.563$ BIM $0.01$ $0.92$ $0.92$ $0.9982$ $0.302$ $4.819$ $0.522$ ML $57.2$ $1.00$ $1.00$ $0.9999$ $0.226$ $3.235$ $0.416$ LL $50.1$ $1.00$ $1.00$ $0.9998$ $0.279$ $4.655$ $0.507$ ML $0.3$ $1.00$ $1.00$ $0.9999$ $0.603$ $2.151$ $0.443$ LL $0.3$ $1.00$ $1.00$ $0.9999$ $0.733$ $3.207$ $0.458$ ML $65.1$ $1.00$ $1.00$ $0.9999$ $0.983$ $3.667$ $0.029$ LL $61.8$ $1.00$ $1.00$ $0.9999$ $0.995$ $5.122$ $0.061$ ML $0.6$ $0.93$ $0.93$ $0.7137$ $1.000$ $3.728$ $0.031$ LL $0.7$ $0.43$ $0.53$ $0.6253$ $1.000$ $5.555$ $0.063$ FGSM $0.1$ $0.86$ $0.86$ $0.9694$ $0.016$ $0.864$ $0.998$ BIM $0.4$ $0.92$ $0.92$ $0.9872$ $0.008$ $0.367$ $0.994$ ML $186.0$ $1.00$ $1.00$ $0.9890$ $0.009$ $0.341$ $0.960$ LL $197.6$ $1.00$ $1.00$ $0.9751$ $0.014$ $0.522$ $0.994$ DF UA $0.5$ $1.00$ $1.00$ $0.8382$ $0.028$ $0.028$ $0.993$ ML $6.1$ $1.00$ $1.00$ $0.9866$ $0.024$ $0.204$ $0.641$ LL $7.0$ $1.00$ $1.00$ $0.9730$ $0.041$ $0.353$ $0.847$ ML $459.4$ $1.00$ $1.00$ $0.9898$ $0.581$ $1.547$ $0.010$ LL $461.2$ $1.00$ $1.00$ $0.9764$ $0.694$ $2.517$ $0.024$ ML $7.5$ $1.00$ $1.00$ $0.5326$ $0.843$ $3.681$ $0.046$ LL $12.0$ $0.98$ $1.00$ $0.3984$ $0.904$ $5.563$ $0.098$ -- ------ ------------ ---------- --------- -------- -------- ---------- ---------------- ----------- ----------- : Evaluation of the attacks on MNIST and CIFAR-10.[]{data-label="tab:attack-evaluation"} Attack success rate (ASR) measures the percentage of successful adversarial examples over all attacked inputs while misclassification rate (MR) is defined as the percentage of misclassified adversarial examples over all attacked inputs. For untargeted attacks (UA), MR is the same as ASR, but for targeted attacks, MR may be higher than ASR because an adversarial example may fail the targeted attack but still cause misclassification, resulting in a wrong prediction output other than the true class. Most adversarial attacks evaluated produce a high ASR and MR with high prediction confidence, meaning that the target model being attacked predicts adversarial examples to be a wrong class with a high probability. We utilize the following metrics to evaluate defensibility: - *Prevention Success Rate (PSR)*: The percentage of the adversarial examples that are repaired and correctly classified by the target model under defense. With the benign testing set, PSR denotes the benign accuracy. - *Detection Success Rate (TSR)*: The percentage of adversarial examples that could not be repaired but are correctly flagged as the attack example by the defense system. For the benign testing set, TSR represents the benign false positive rate, i.e., the percentage of the benign examples being flagged as adversarial (BFP) over the total number of benign testing examples. - *Defense Success Rate (DSR)*: The percentage of adversarial examples that are either repaired or detected. $\text{DSR} = \text{PSR} + \text{TSR}$. - *False Positive Rate (FP)*: The percentage of the adversarial examples that can be correctly classified (repaired) but are flagged as adversarial when all inputs are adversarial examples. For the benign testing set, it represents the percentage of the correctly classified benign examples being flagged as adversarial. Note that all evaluations on benign examples are based on the entire testing set with $10,000$ images. Model Denoising Ensemble Defense {#sec:dnn-denoising} ================================ We design the model denoising ensemble defense as the first perimeter of defense to prevent adversarial misclassification. The defense structure consists of multiple DNN denoisers, each performs noise reduction via unsupervised learning using one specific denoising autoencoder, aiming to remove adversarial perturbations as much as possible by joint force. Denoising Autoencoders ---------------------- Denoising autoencoders are initially proposed as a way to extract and compose robust features which can be utilized to replace and optimize the random initialization and bootstrap the efficiency of training deep neural networks [@vincent2008extracting]. From a manifold learning perspective [@chapelle2009semi], natural high-dimensional data concentrate close to a nonlinear low-dimensional manifold. Adversarial attacks can be considered as a malicious process to drag benign examples away from the manifold where they concentrate. DNN denoising training is to learn a function to map a corrupted example, likely to be outside and farther from the manifold, back to its uncorrupted form. Thus, we can utilize denoising autoencoders, trained with uniformly corrupted examples, to reverse the adversarial perturbation process such that the adverse effect can be removed. Several efforts [@xie2012image; @agostinelli2013adaptive] have been proposed to perform image denoising tasks with different DNN models and different design of autoencoder structures and hyperparameters. ![image](figures/dae.pdf){width="0.805\linewidth"} For image data, a denoising autoencoder takes an input image, transforms it into a noisy version, and feeds the noisy image to the autoencoder to perform latent space projection and then reconstructs the image with the goal of generating a clean version of the original image as the output. Figure \[fig:dae-training\] illustrates, by example, the key components of training a denoising autoencoder. Let $\boldsymbol{x}$ be an example from the training set and $\tilde{\boldsymbol{x}}$ be the version corrupted by a stochastic noise mapping $q$ such that $\tilde{\boldsymbol{x}}\sim q(\tilde{\boldsymbol{x}}\vert\boldsymbol{x})$ [@vincent2010stacked]. The encoder is an $L_f$-layer neural network, which projects the corrupted example $\tilde{\boldsymbol{x}}$ from a high-dimensional image space to a low-dimensional latent feature space, producing its latent representation $f_\theta(\tilde{\boldsymbol{x}})=f^{L_f}(\cdots(f^{2}(f^{1}(\tilde{\boldsymbol{x}};\theta_1);\theta_2));\theta_{L_f})$ where $f^i$ is the operation at the $i$-th encoding layer (e.g., convolution) with weights $\theta_i$, $f_\theta$ and $\theta=(\theta_{1},...,\theta_{L_f})$ can be viewed as the composite function of the encoder and the weights respectively. The decoder, an $L_g$-layer neural network, then restores the spatial structure of $f_\theta(\tilde{\boldsymbol{x}})$ by mapping its latent representation back to the original image feature space and produces the reconstructed example $g_{\theta'}(f_\theta(\tilde{\boldsymbol{x}}))=g^{L_g}(\cdots(g^{2}(g^{1}(f_\theta(\tilde{\boldsymbol{x}});\theta'_1);\theta'_2));\theta'_{L_g})$ where $g^i$ is the operation at the $i$-th decoding layer (e.g., deconvolution) with weights $\theta_i'$, $g_{\theta'}$ and $\theta'=(\theta'_{1},...,\theta'_{L_g})$ can be viewed as the composite function of the decoder and the weights respectively. The multilayer encoder and decoder constitute a deep denoising autoencoder. Given $N$ training examples $\{\boldsymbol{x}_1,...,\boldsymbol{x}_N\}$, the denoising autoencoder is trained by backpropagation to minimize the reconstruction loss: $$\label{eq:dae-opt} \begin{split} & \mathcal{L}(\theta, \theta'; \{\boldsymbol{x}_i\}^{N}_{i=1},d, q, \{f^i\}^{L_f}_{i=1}, \{g^i\}^{L_g}_{i=1},\lambda)\\ =&\frac{1}{N}\sum_{i=1}^{N}d(\boldsymbol{x}_i, g_{\theta'}(f_{\theta}(\tilde{\boldsymbol{x}}_i))) + \frac{\lambda}{2}(\vert\vert\theta\vert\vert^2_\text{F}+\vert\vert\theta'\vert\vert^2_\text{F}), \end{split}$$ where $d$ is a distance function and $\lambda$ is a regularization hyperparameter penalizing the Frobenius norm of $\theta$ and $\theta'$. Given a query example $\boldsymbol{x}$ at runtime, the denoising autoencoder produces a denoised version $\mathcal{D}(\boldsymbol{x})=g_{\theta'}(f_\theta(\boldsymbol{x}))$ with fixed weights $(\theta, \theta')$. Strategic Teaming of Multiple DNN Denoisers ------------------------------------------- Recall in Figure \[fig:dae-training\] that the noise injection function $q$ of a denoising autoencoder transforms the original input $\boldsymbol{x}$ to a noisy version of $\boldsymbol{x}$, denoted by $\tilde{\boldsymbol{x}}$. Clearly, different data corruption methods produce different versions of input $\boldsymbol{x}$ and result in denoisers that exhibit different denoising effects. Figure \[fig:denoising-example\] visualizes two testing examples from MNIST and CIFAR-10 and the corresponding adversarial examples generated by six attack methods (the 1st row), their denoised versions produced by two denoisers trained with Gaussian noise (the 2nd row) and salt-and-pepper noise (the 3rd row) respectively. The predicted class label and confidence by the target model are presented for each case. On the first row, it shows that adversarial attacks successfully fool the target model under all six attacks for the examples from both datasets. On the second row, Gaussian denoiser successfully remove malicious perturbations from four out of six attacks for MNIST example and three out of six attacks for CIFAR-10 example, showing the robustness improvement of the target model even with single DNN denoiser. On the third row, the salt-and-pepper denoiser successfully repaired five out of six attacks for CIFAR-10 example and three out of six attacks for MNIST example. ![image](figures/dae-example.pdf){width="0.82\linewidth"} These examples deliver two important messages: First, exploiting denoising autoencoders trained with uniformly corrupted examples can remove meticulously crafted adversarial perturbations and restore the classification capability of the target model. Second, no single denoiser is effective across all attacks and each of them is good at removing some types of noise but not the others. This motivates us to employ a team of diverse denoisers, instead of relying on a single denoiser, such that an adversarial example can be denoised from different perspectives. In our model denoising ensemble defense, we first construct a set of base DNN denoisers by exploiting different approaches to generate denoisers. For example, we could use different input transformation techniques to perform the input corruption process $q$ (e.g., Gaussian noise, salt-and-pepper noise, and masking noise) such that the optimization in Equation \[eq:dae-opt\] attempts to learn different functions to reverse the corruption [@xie2012image]. Alternatively, we can also create different denoisers by altering the model structure (i.e., $\{f^i\}^{L_f}_{i=1}$ and $\{g^i\}^{L_g}_{i=1}$), or the hyperparameter settings (e.g., $\lambda$, training epochs, and random seed for weight initialization), because different ways of generating denoisers can have different effects with respect to the manifold and the convergence of the DNN learning [@wei2018adversarial; @huang2017snapshot; @wu2018comparative]. The third approach to generate different denoisers is to use different optimization objectives, such as changing the distance function $d$ from a simple per-pixel loss to a perceptual loss [@johnson2016perceptual], or using an advanced regularization such as a sparsity constraint [@cho2013simple]. Upon obtaining a set of base denoisers, we next define ensemble diversity metric to identify and select diverse ensembles such that each of the ensemble members has high benign accuracy and the ensemble teaming has high diversity in terms of failure-independence and low negative example correlation. In our first prototype of [MODEF]{}, we use the kappa coefficient ($\kappa$) [@mchugh2012interrater] as a pairwise metric to quantify the diversity of an ensemble by measuring each pair of its member denoisers from a statistical perspective. The top-ranked ensemble teams by kappa diversity from the set of base denoisers will be chosen as the pool of our candidate ensembles. [**Kappa Diversity Metric.**]{} Let $N^-$ denote the number of testing examples that the denoised versions are labeled as different classes by the target model, $N^-_{ij}$ denote the number of instances in the testing set that the target model labels one denoised version as class $i$ and the other as class $j$. The kappa metric can be computed by $$\label{eq:cohen-kappa} \kappa=\frac{\frac{1}{N^-}\sum_{i=1}^{K}N^-_{ii}-\sum_{i=1}^{K}{(\frac{N^-_{i*}}{N^-}-\frac{N^-_{*i}}{N^-})}}{1-\sum_{i=1}^{K}(\frac{N^-_{i*}}{N^-}-\frac{N^-_{*i}}{N^-})}.$$ Based on the kappa value of each pair of base denoisers, we enumerate and rank all possible ensemble teaming combinations according to their average kappa values and maintain a “$\kappa$-ranked list” of denoising ensembles with the low average kappa values below a system-supplied threshold to ensure the denoising ensemble used by [MODEF]{} has high denoising diversity. [**Denoising Ensemble Selection and Output Decision.**]{} We select one denoising ensemble team from the $\kappa$-ranked list of denoising ensembles at runtime. For the chosen denoising ensemble of size $Z_{D}$, we send the query (test) example $\boldsymbol{x}$ to each ensemble member and obtain $Z_{D}$ denoised versions of $\boldsymbol{x}$. There are several ways to produce the ensemble output accordingly. In the first prototype of [MODEF]{}, we use a majority voting method. When the target model makes a consistent prediction on a majority of the denoised versions, we randomly choose one of the majority members and uses its denoised version as the ensemble output. Otherwise, the query example is flagged as adversarial. Figure \[fig:denoising-example\] provides an illustrative example. First, the benign case refers to the target model without attacks (the 1st column for MNIST and the 8th column for CIRAR-10). The 2nd to the 7th columns correspond to six different attacks for MNIST and the 9th to the 14th correspond to six attacks for CIFAR-10. The first row shows that the attacks to the two original test examples (digit $4$ in MNIST and dog in CIFAR-10) are successful with high confidence. However, when we employ the two denoisers (see the 2nd row and the 3rd row), the six adversarial attacks may fail for both test examples with either Gaussian denoiser or salt-and-pepper denoiser. For MNIST digit $4$, the denoising ensemble will survive under CW$_{0}$ and CW$_{2}$. For CIFAR-10 dog example, the denoising ensemble will survive under all three CW attacks. This example also illustrates that some attack algorithms, such as FGSM, BIM and JSMA, can escape from the denoising ensemble defense. One may argue that the reconstruction vectors from a well-trained denoising autoencoder form a vector field which points in the direction of the data manifold. Yet, this may not hold for the examples which are distant from the manifold as they are rarely sampled during training [@alain2014regularized]. Thus, a more comprehensive solution approach should be developed. This motivates us to propose the output model verification ensemble (Section \[sec:multi-model-verification\]) and to integrate the input denoising and output verification to formulate a cross-layer model ensemble defense (Section [\[sec:cross-layer\]]{}). Denoising and Verification Co-Defense {#sec:co-defense} ===================================== In this section, we first describe the model verification ensemble as an alternative defense method and then present our denoising-verification cross-layer ensemble defense framework, which takes the model denoising ensemble as the front-end defense and then combines it with the model verification ensemble as the back-end defense. Model Verification Ensemble Defense {#sec:multi-model-verification} ----------------------------------- One of the intriguing properties of adversarial examples is the attack transferability [@papernot2016transferability]. The main idea of model verification ensemble is to break adversarial attacks by exploiting the weak spots of adversarial transferability. It is known that two diverse models trained on the same dataset for the same learning task, their decision boundaries can be quite different even if they obtain a similar training accuracy. Hence, an adversarial example generated by one attack algorithm through one way of perturbing a benign one may successfully fool the target model but may not succeed in fooling other models trained on the same dataset, especially when such models are diverse with respect to negative examples and thus are failure-independent. By creating model ensembles with high diversity as output verifiers for the target model, we argue that (1) the model verification ensembles can improve the robustness of the target model against adversarial examples, and (2) the model verification ensembles are complementary and alternative to denoising ensembles, and thus a cross-layer diversity ensemble defense that integrates input denoising ensemble with output verification ensemble provides greater potential for higher robustness, because the adversarial examples are unlikely to be transferable to all of the member models of the cross-layer ensemble. Similar to the creation of denoising ensembles, we perform two steps to create verification ensembles. First, we construct a set of base models as the verifiers for examining and repairing the target model prediction output, each is trained on the same dataset for the same task as the target model. Several techniques can be used to produce the base candidate models, such as varying neural network structures [@wu2018comparative], training hyperparameters, or performing data augmentation [@krizhevsky2012imagenet]. One can also conduct snapshot learning [@huang2017snapshot] to obtain a set of model verifiers efficiently in a single run of model training. We employ two pruning criteria to generate model diversity ensembles. The first filter is to use the test accuracy to select only those base candidate models with the test accuracy similar to that of the target model. For MNIST and CIFAR-10, we collect numerous pre-trained DNN models from the public domain, with high training and test accuracy under no attack scenarios. The second filter is to select those model ensemble teams that have high model diversity with respect to high failure-independence and low negative/error correlation. In the first prototype of [MODEF]{}, we utilize the kappa coefficient metric to quantify the disagreement between each pair of base candidate models. Let $N^-$ be the number of testing examples the base candidate models produce inconsistent predictions and $N^-_{ij}$ be the number of instances in the testing set, which are labeled as class $i$ by one model and class $j$ by the other in Equation \[eq:cohen-kappa\]. We then enumerate and rank all possible ensemble teaming combinations with a minimum team size of $3$ and we compute the average kappa value for each ensemble team examined and select a $\kappa$-ranked list of model verification ensembles with low average kappa values using the system-defined kappa-diversity threshold. Any model ensemble from this chosen kappa-ranked list will have its average kappa value below the given threshold and thus high model diversity. ---------- -------------- --------------------- -------------- --------------------- **Name** **Benign Accuracy** **Name** **Benign Accuracy** V$_1$ CNN-5 $0.9919$ CNN-10 $0.9062$ V$_2$ CNN-4 $0.9861$ ResNet20 $0.9205$ V$_3$ CNN-6 $0.9917$ ResNet32 $0.9313$ V$_4$ LeNet $0.9980$ ResNet50 $0.9312$ V$_5$ MLP $0.9761$ ResNet56 $0.9419$ V$_6$ MobileNet $0.9934$ ResNet110 $0.9419$ V$_7$ MobileNet-v2 $0.9939$ ResNet152 $0.9392$ V$_8$ PNASNet $0.9950$ ResNext29 $0.9725$ V$_9$ SVM $0.9832$ VGG $0.9359$ V$_{10}$ VGG $0.9946$ WideResNet28 $0.9712$ ---------- -------------- --------------------- -------------- --------------------- : The ten model verifiers for each dataset with their prediction accuracy on benign examples.[]{data-label="tab:a-verifier-acc"} Table \[tab:a-verifier-acc\] gives the set of ten base models for MNIST and CIFAR-10 respectively, and all ten models have similar benign accuracy under no attack scenarios. Table \[fig:verification-example\] illustrates the model verification ensemble method using the same two query examples: digit $4$ from MNIST and a color image of dog from CIFAR-10. The 2nd row shows how the TM (target model) responds to the query example under benign (no attack) and under the six different attacks for MNIST (columns 2-8) and for CIFAR-10 (columns 9-14). When no defense protection to the target model, all adversarial examples successfully fool the target model for both digit $4$ of MNIST and dog image of CIFAR-10. Interestingly, even with the adversarial example transferability, some of these adversarial examples are correctly classified by some of the ten base model verifiers (denoted by $\text{V}_1,...,\text{V}_{10}$). [|r|c|c|c|c|c|c|c|c|c|c|c|c|c|c|]{} & &\ & **Benign** & **FGSM** & **BIM** & **CW$_\infty$** & **CW$_2$** & **CW$_0$** & **JSMA** & **Benign** & **FGSM** & **BIM** & **CW$_\infty$** & **CW$_2$** & **CW$_0$** & **JSMA**\ & ![image](figures/dae-example-s/m-benign.png) & ![image](figures/dae-example-s/m-fgsm.png) & ![image](figures/dae-example-s/m-bim.png) & ![image](figures/dae-example-s/m-cwi.png) & ![image](figures/dae-example-s/m-cw2.png) & ![image](figures/dae-example-s/m-cw0.png) & ![image](figures/dae-example-s/m-jsma.png) & ![image](figures/dae-example-s/c-benign.png) & ![image](figures/dae-example-s/c-fgsm.png) & ![image](figures/dae-example-s/c-bim.png) & ![image](figures/dae-example-s/c-cwi.png) & ![image](figures/dae-example-s/c-cw2.png) & ![image](figures/dae-example-s/c-cw0.png) & ![image](figures/dae-example-s/c-jsma.png)\ **TM** & --------------------- $\textbf{4}$ $\textbf{[1.0000]}$ --------------------- & ------------------- $\text 9$ $\text{[1.0000]}$ ------------------- & ------------------- $\text{9}$ $\text{[1.0000]}$ ------------------- & ------------------- $\text{3}$ $\text{[0.9998]}$ ------------------- & ------------------- $\text{3}$ $\text{[0.9998]}$ ------------------- & ------------------- $\text{3}$ $\text{[0.9999]}$ ------------------- & ------------------- $\text{3}$ $\text{[0.5537]}$ ------------------- & --------------------- **dog** $\textbf{[0.9949]}$ --------------------- & ---------------- horse [\[0.9978\]]{} ---------------- & ---------------- horse [\[1.0000\]]{} ---------------- & ---------------- airplane [\[0.9724\]]{} ---------------- & ---------------- airplane [\[0.9758\]]{} ---------------- & ---------------- airplane [\[0.9754\]]{} ---------------- & ---------------- airplane [\[0.4943\]]{} ---------------- \ **$\text{V}_1$** & --------------------- $\textbf{4}$ $\textbf{[0.9029]}$ --------------------- & ---------------- 9 [\[0.3287\]]{} ---------------- & ---------------- [9]{} [\[0.6249\]]{} ---------------- & ---------------- [9]{} [\[0.2235\]]{} ---------------- & --------------------- $\textbf{4}$ $\textbf{[0.6314]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.7399]}$ --------------------- & ---------------- [9]{} [\[0.5120\]]{} ---------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9530]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.4945]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.7772]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.8809]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.7491]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9023]}$ --------------------- & ---------------- [truck]{} [\[0.8392\]]{} ---------------- \ **$\text{V}_2$** & --------------------- $\textbf{4}$ $\textbf{[1.0000]}$ --------------------- & ---------------- 9 [\[1.0000\]]{} ---------------- & ---------------- [9]{} [\[1.0000\]]{} ---------------- & --------------------- $\textbf{4}$ $\textbf{[0.9925]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9997]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9934]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.5663]}$ --------------------- \ **$\text{V}_3$** & --------------------- $\textbf{4}$ $\textbf{[1.0000]}$ --------------------- & ---------------- [9]{} [\[0.9994\]]{} ---------------- & ---------------- [9]{} [\[1.0000\]]{} ---------------- & --------------------- $\textbf{4}$ $\textbf{[0.9999]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[1.0000]}$ --------------------- & ---------------- [9]{} [\[0.9525\]]{} ---------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9999]}$ --------------------- & ------------------- $\text{horse}$ $\text{[0.8076]}$ ------------------- & ------------------- $\text{horse}$ $\text{[0.5280]}$ ------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9934]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9996]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9996]}$ --------------------- & ------------------- $\text{cat}$ $\text{[1.0000]}$ ------------------- \ **$\text{V}_4$** & --------------------- $\textbf{4}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9686]}$ --------------------- & ---------------- [9]{} [\[0.8475\]]{} ---------------- & --------------------- $\textbf{4}$ $\textbf{[0.8105]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9985]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9886]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9640]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.7023]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.8187]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9998]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & ------------------- $\text{cat}$ $\text{[0.9584]}$ ------------------- \ **$\text{V}_5$** & --------------------- $\textbf{4}$ $\textbf{[1.0000]}$ --------------------- & ------------------- $\text{9}$ $\text{[0.8796]}$ ------------------- & ------------------- $\text{9}$ $\text{[1.0000]}$ ------------------- & ------------------- $\text{9}$ $\text{[0.5978]}$ ------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9974]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9939]}$ --------------------- & ------------------- $\text{9}$ $\text{[0.9130]}$ ------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9956]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9484]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9881]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9961]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9959]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.7752]}$ --------------------- & ------------------- $\text{cat}$ $\text{[0.7850]}$ ------------------- \ **$\text{V}_6$** & --------------------- $\textbf{4}$ $\textbf{[0.9999]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.8277]}$ --------------------- & ------------------- $\text{8}$ $\text{[0.9915]}$ ------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9949]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9995]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9939]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9994]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9233]}$ --------------------- \ **$\text{V}_7$** & --------------------- $\textbf{4}$ $\textbf{[1.0000]}$ --------------------- & ------------------- $\text{9}$ $\text{[0.6400]}$ ------------------- & --------------------- $\text{9}$ $\textbf{[0.7800]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.5576]}$ --------------------- & ------------------- $\text{3}$ $\text{[0.9908]}$ ------------------- & --------------------- $\textbf{4}$ $\textbf{[0.4516]}$ --------------------- & --------------------- $\textbf{4}$ $\textbf{[0.9903]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[1.0000]}$ --------------------- & ------------------- $\text{horse}$ $\text{[0.9994]}$ ------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.8934]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9958]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9998]}$ --------------------- & ------------------- $\text{horse}$ $\text{[0.8080]}$ ------------------- & ------------------- $\text{frog}$ $\text{[0.9975]}$ ------------------- \ **$\text{V}_8$** & --------------------- $\textbf{4}$ $\textbf{[0.9997]}$ --------------------- & ------------------- 9 $\text{[0.4434]}$ ------------------- & ------------------- 9 $\text{[0.5201]}$ ------------------- & ------------------- 8 $\text{[0.8267]}$ ------------------- & --------------------- **4** $\textbf{[0.8710]}$ --------------------- & ------------------- 3 $\text{[0.7317]}$ ------------------- & --------------------- **4** $\textbf{[0.9998]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.8284]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.6055]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.7699]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.8572]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.8565]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.6601]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.5119]}$ --------------------- \ **$\text{V}_9$** & --------------------- **4** $\textbf{[0.9948]}$ --------------------- & ------------------- 9 $\text{[0.5192]}$ ------------------- & ------------------- 9 $\text{[0.6377]}$ ------------------- & --------------------- **4** $\textbf{[0.8961]}$ --------------------- & --------------------- **4** $\textbf{[0.9885]}$ --------------------- & --------------------- **4** $\textbf{[0.9612]}$ --------------------- & --------------------- **4** $\textbf{[0.3684]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9996]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9987]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9995]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9996]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9994]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9975]}$ --------------------- & ------------------- $\text{cat}$ $\text{[0.9879]}$ ------------------- \ **$\text{V}_{10}$** & --------------------- **4** $\textbf{[0.9997]}$ --------------------- & [@c@]{}\ $\text{[0.9166]}$ & [@c@]{}\ $\text{[0.9994]}$ & --------------------- **4** $\textbf{[0.3515]}$ --------------------- & --------------------- **4** $\textbf{[0.5573]}$ --------------------- & [@c@]{}\ $\text{[0.4868]}$ & [@c@]{}\ $\text{[0.6144]}$ & --------------------- $\textbf{dog}$ $\textbf{[0.9990]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9985]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9990]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9989]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9988]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9989]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9869]}$ --------------------- \ & --------------------- **4** $\textbf{[0.9982]}$ --------------------- & [@c@]{}\ $\text{[0.4663]}$ & [@c@]{}\ $\text{[0.5459]}$ & --------------------- **4** $\textbf{[0.6303]}$ --------------------- & --------------------- **4** $\textbf{[0.9952]}$ --------------------- & --------------------- **4** $\textbf{[0.9830]}$ --------------------- & --------------------- **4** $\textbf{[0.4559]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9775]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.6748]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.8246]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9722]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9599]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.8334]}$ --------------------- & ------------------- $\text{cat}$ $\text{[0.3731]}$ ------------------- \ & --------------------- **4** $\textbf{[1.0000]}$ --------------------- & --------------------- **4** $\textbf{[0.9825]}$ --------------------- & --------------------- **4** $\textbf{[0.8452]}$ --------------------- & --------------------- **4** $\textbf{[0.9994]}$ --------------------- & --------------------- **4** $\textbf{[0.9996]}$ --------------------- & --------------------- **4** $\textbf{[0.9996]}$ --------------------- & --------------------- **4** $\textbf{[0.9586]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9844]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9576]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9775]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9809]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9798]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9742]}$ --------------------- & --------------------- $\textbf{dog}$ $\textbf{[0.9306]}$ --------------------- \ Given a query example $\boldsymbol{x}$, a model verification ensemble of size $Z_\mathcal{V}$ is selected randomly at runtime from the $\kappa$-ranked list of ensembles as the team of verifiers for the target model, each member, denoted by $\text{V}_i$, takes the query example as the input and produces a prediction probability distribution, $\text{V}_i(\boldsymbol{x}):\mathbb{R}^D\rightarrow\mathbb{R}^K$, over the output space. Then, the [MODEF]{}-approved prediction output $\hat{y}$ will be produced based on a soft-voting scheme. In addition to the unweighted majority voting (recall the denoising ensemble in Section \[sec:dnn-denoising\]), another example of such voting scheme could be weighted averaging based on the confidence of the prediction made by each verifier in the chosen verification ensemble: $$\hat{y}=\arg\max{}_{1\leq j\leq K}\frac{1}{Z_\mathcal{V}}\sum_{i=1}^{Z_\mathcal{V}}\text{V}_{i,j}(\boldsymbol{x}),$$ where $\text{V}_{i,j}(\boldsymbol{x})$ denotes the confidence of $\boldsymbol{x}$ being from class $j$ predicted by verifier $\text{V}_i$. The 2nd row from the bottom of Table \[fig:verification-example\] shows the results of using model verification ensemble defense with the $3$-model team $\text{V}_5$,$\text{V}_6$,$\text{V}_9$ for MNIST and the $10$-model team $\text{V}_1, ..., \text{V}_{10}$ for CIFAR-10. Without using the model denoising ensemble, we observe that for these two query inputs (digit $4$ and dog), most of the adversarial examples can be correctly classified, including JSMA and CW$_{\infty}$ attacks on the MNIST example, the FGSM and BIM attacks on the CIFAR-10 example, which failed by both the denoising ensemble and any of the two denoisers in Figure \[fig:denoising-example\]. However, the verification ensemble selected from the ten base models still fails to repair the adversarial examples generated by FGSM and BIM for digit $4$ of MNIST and adversarial example generated by JSMA for the dog example of CIFAR-10. We are motivated to combine denoising ensemble with verification ensemble, which has a high probability to outperform denoising ensemble or verification ensemble alone. Denoising-Verification Cross-Layer Ensemble {#sec:cross-layer} ------------------------------------------- There are a number of ways that we can create cross-layer ensembles. Due to the space constraint, we below describe two approaches representing two ends of the spectrum. The first approach sends an ensemble voted denoising output to the model verification ensemble. The second approach sends to the model verification ensemble all denoised versions of the query input produced by every member of the denoising ensemble. Let the back-end defense to conduct the cross-layer ensemble. This approach ensures that no errors from the front-end defense will be propagated to the back-end defense phase, at the cost of sending every denoised version without any consolidation effort. Other solutions could be in between of these two spectrums, such as sending only those voted denoising outputs that are ranked higher than the system-defined consensus threshold. [**One-to-Many Denoising-Verification Cross-Layer Ensemble.**]{} In the front-end defense phase, [MODEF]{} will employ denoising ensemble to produce one transformed input example by removing adversarial noises through unsupervised denoising and multiple model denoising consensus, such as unweighted majority voting, simple averaging or weighted averaging. In the back-end defense phase, it takes the voted denoised version of the query example, performs the target model prediction first and then performs the model verification ensemble over the target prediction outcome. This process serves two purposes: (1) It aims to ensure the correctness of the target model prediction outcome by repairing the prediction error through the cross-layer ensemble. (2) It also provides the detection capability to flag those adversarial examples that escape from or cannot be repaired by the cross-layer model ensemble defense framework. [**Many-to-Many Denoising-Verification Cross-Layer Ensemble.**]{} When the denoising ensemble as the front-end defense chooses not to filter out any denoised versions of the query input. The back-end defense phase will need to consider each denoised version. -- ------------------------- ------------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- **FGSM** **BIM** **ML** **LL** **ML** **LL** **ML** **LL** **ML** **LL** MODEF $0.9855$ $0.89$ $\mathbf{0.86}$ $\mathbf{0.99}$ $\mathbf{0.97}$ $\mathbf{0.98}$ $\mathbf{0.95}$ $\mathbf{0.94}$ $\mathbf{0.92}$ $\mathbf{0.98}$ $\mathbf{0.90}$ $\mathbf{0.94}$ $\mathbf{0.04}$ Adversarial Training $\mathbf{0.9884}$ $\mathbf{0.91}$ $0.81$ $0.97$ $0.88$ $0.92$ $0.84$ $0.67$ $0.64$ $0.73$ $0.69$ $0.81$ $0.12$ Defensive Distillation $0.9784$ $0.68$ $0.57$ $0.91$ $0.85$ $0.91$ $0.84$ $0.78$ $0.72$ $0.85$ $0.75$ $0.79$ $0.11$ Ensemble Transformation $0.9820$ $0.60$ $0.22$ $0.64$ $0.51$ $0.37$ $0.33$ $0.21$ $0.21$ $0.57$ $0.64$ $0.43$ $0.18$ MODEF $\mathbf{0.9631}$ $\mathbf{0.91}$ $\mathbf{0.96}$ $\mathbf{0.98}$ $\mathbf{0.98}$ $\mathbf{0.98}$ $\mathbf{0.98}$ $\mathbf{0.94}$ $\mathbf{0.93}$ $\mathbf{0.92}$ $\mathbf{0.80}$ $\mathbf{0.94}$ $\mathbf{0.06}$ Adversarial Training $0.8790$ $0.64$ $0.58$ $0.68$ $0.77$ $0.75$ $0.79$ $0.44$ $0.48$ $0.50$ $0.45$ $0.61$ $0.14$ Defensive Distillation $0.9118$ $0.60$ $0.65$ $0.79$ $0.88$ $0.86$ $0.90$ $0.60$ $0.69$ $0.70$ $0.47$ $0.71$ $0.14$ Ensemble Transformation $0.8014$ $0.23$ $0.40$ $0.56$ $0.61$ $0.57$ $0.61$ $0.19$ $0.34$ $0.45$ $0.41$ $0.44$ $0.15$ -- ------------------------- ------------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- In this case, we may choose to produce one cross-layer verified prediction outcome for each denoised version and then use an ensemble ranking algorithm, such as unweighted majority voting, simple averaging or weighted averaging, to determine the [MODEF]{} defense verified prediction outcome for the target model. Alternatively, we could also pool all the cross-layer verification results for all denoised versions together and run the ensemble consensus algorithm to rank and select the top-$1$ or top-$k$ prediction outcomes. Concretely, given a query example $\boldsymbol{x}$ at runtime, we first exercise the front-end defense by employing a model denoising ensemble of size $Z_\mathcal{D}$ to produce $Z_\mathcal{D}$ denoised versions of $\boldsymbol{x}$, denoted by $\mathcal{D}_1(\boldsymbol{x}),...\mathcal{D}_{Z_D}(\boldsymbol{x})$. Then, we select a model verification ensemble team of size $Z_\mathcal{V}$ from the [|l|l|lll]{} & &\ & & & &\ Convolution + Sigmoid & $3\times3\times3$ & & &\ Average Pooling & $2\times2$ & & &\ Convolution + Sigmoid & $3\times3\times3$ & & &\ Convolution + Sigmoid & $3\times3\times3$ & & &\ Upsampling & $2\times2$ & & &\ Convolution + Sigmoid & $3\times3\times3$ & & &\ Convolution + Sigmoid & $3\times3\times1$ & & &\ $\kappa$-ranked list of model verification ensembles. For each of the denoised versions, we will produce $Z_\mathcal{D}$ probability distributions $P_1(\boldsymbol{x}),...,P_{Z_{\mathcal{D}}}(\boldsymbol{x})$ where $P_k(\boldsymbol{x})=\frac{1}{Z_\mathcal{V}}\sum_{i=1}^{Z_\mathcal{V}}\text{V}_i(\mathcal{D}_k(\boldsymbol{x}))$. We select the one, denoted as the $c$-th denoiser, producing the most confident prediction and take the corresponding predicted class label as the final defense-approved prediction. Similar to the one-to-many approach, detection capability can be offered to flag irreparable adversarial examples. Table \[fig:verification-example\] shows the result of the cross-layer ensemble approach in the last row using the above many-to-many cross-layer ensemble defense. Compared with model verification ensemble only defense (the 2nd row from the bottom), clearly, the denoising-verification cross-layer ensemble can successfully verify and repair all the adversarial examples. In this experiment, the cross-layer ensemble defense method uses the two denoisers as shown in Figure \[fig:denoising-example\]. Experimental Evaluation {#sec:experimental-results} ======================= Experimental Setup ------------------ We evaluate [MODEF]{} on two popular benchmark datasets: MNIST and CIFAR-10. Each of them is associated with one denoiser trained with Gaussian noise and another trained with salt-and-pepper noise. Their neural architectures are reported in Table \[tab:a-denoiser-architectures\] where we use [\[]{}kernel width[\]]{}$\times$[\[]{}output channel[\]]{} to denote parameters of convolution layers and [\[]{}kernel width[\]]{} to denote parameters of average pooling and upsampling layers. The volume of Gaussian noise is set to be $0.3$ for MNIST and $0.01$ for CIFAR-10 while the ratio of pixels randomly modified by the salt-and-pepper corruption process is set to be $0.1$ for both datasets. All denoisers are trained with Adam optimizer with a learning rate of $0.001$ and a batch size of $256$. Euclidean distance is employed to measure the difference between two examples (i.e., the function $d$ in Equation \[eq:dae-opt\]) while the regularization hyperparameter $\lambda$ is set to be $10^{-9}$. The number of training epochs for MNIST is $100$ and that for CIFAR-10 is $400$. For each dataset, a set of ten model verifiers is collected from the public domain. As reported in Table \[tab:a-verifier-acc\], each of them provides a competitive accuracy on benign examples as the target model. In our experimental studies, we consider a fixed team of DNN denoisers and demonstrate different teaming options through model verifiers. All experiments were run on Google Colab[^1] operated under Ubuntu 18.04.2 LTS with an Intel Xeon CPU (two cores @2.3 GHz), an NVIDIA Tesla K80 (12 GB) GPU, and 13 GB RAM. Comparing [MODEF]{} with Existing Representative Defenses --------------------------------------------------------- This set of experiments compares the [MODEF]{} cross-layer ensemble defense with three representative defense methods: adversarial training [@szegedy2013intriguing], defensive distillation [@papernot2016distillation], and ensemble transformation of input examples [@guo2017countering]. We select the ensemble team to be $\text{V}_5$,$\text{V}_6$,$\text{V}_9$ for MNIST and $\text{V}_1,...,\text{V}_{10}$ for CIFAR-10 from the $\kappa$-ranked list based on their model diversity. They are exploited by default in the following studies unless further specified. Table \[tab:exp-comparison\] reports the results in benign accuracy under no attack and DSR under ten attacks for MNIST and CIFAR-10 introduced in Section \[sec:attacks\]. Clearly, [MODEF]{} outperforms existing schemes for all attacks on CIFAR-10 and nine out of ten attacks on MNIST. The one exception on MNIST is the FGSM attack, under which [MODEF]{} has a DSR of $0.89$, slightly lower than the DSR of $0.91$ by adversarial training. However, adversarial training is attack-dependent and does not generalize over attack algorithms. Indeed, existing defense methods tend to perform inconsistently across different attacks. For instance, defensive distillation reaches a high DSR of $0.90$ under the CW$_2$ LL attack on CIFAR-10 but achieves only a low DSR of $0.47$ under JSMA LL. This can also be observed in their high standard deviation of DSRs over attacks. In contrast, [MODEF]{} achieves a competitive benign accuracy with a high and stable DSR, showing that it can generalize well across all attacks. Comparison of Three MODEF Ensemble Defenses ------------------------------------------- Given that [MODEF]{} outperforms existing defenses, the next set of experiments was conducted to compare the performance of three [MODEF]{} ensemble defense approaches and understand how denoising ensemble and verification ensemble complement one another in delivering higher robustness for the target model in the cross-layer ensemble defense. Table \[tab:exp-denoiser\] reports the results with upper wide table for MNIST and lower wide table for CIFAR-10. We highlight the best benign accuracy (PSR) under no attack and defense accuracy (DSR) against each of the attacks in boldface. [@ @ plus 1fil minus @skip]{} [|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|]{} & & & & &\ & **PSR** & **TSR** & **DSR** & **FP** & **PSR** & **TSR** & **DSR** & **FP** & **PSR** & **TSR** & **DSR** & **FP** & **PSR** & **TSR** & **DSR** & **FP** & **PSR** & **TSR** & **DSR** & **FP**\ & $\mathbf{0.9943}$ & $0.00$ & ${0.9943}$ & $0.00$ & $\mathbf{0.9943}$ & $0.00$ & ${0.9943}$ & $0.00$ & $0.9911$ & $0.0038$ & $0.9949$ & $0.0024$ & $0.9888$ & $0.00$ & $0.9888$ & $0.00$ & $0.9855$ & $0.0035$ & $0.9890$ & $0.0030$\ FGSM & & $0.82$ & $0.00$ & $0.82$ & $0.00$ & $0.40$ & $0.00$ & $0.40$ & $0.00$ & $0.40$ & $0.43$ & $0.83$ & $0.28$ & $0.71$ & $0.00$ & $0.71$ & $0.00$ & $0.87$ & $0.02$ & $\mathbf{0.89}$ & $0.00$\ BIM & & $0.71$ & $0.00$ & $0.71$ & $0.00$ & $0.06$ & $0.00$ & $0.06$ & $0.00$ & $0.06$ & $0.65$ & $0.71$ & $0.17$ & $0.78$ & $0.00$ & $0.78$ & $0.00$ & $0.84$ & $0.02$ & $\mathbf{0.86}$ & $0.00$\ & ML & $0.94$ & $0.00$ & $0.94$ & $0.00$ & $0.03$ & $0.00$ & $0.03$ & $0.00$ & $0.02$ & $0.93$ & $0.95$ & $0.55$ & $0.93$ & $0.00$ & $0.93$ & $0.00$ & $0.99$ & $0.00$ & $\mathbf{0.99}$ & $0.00$\ & LL & $0.99$ & $0.00$ & $\mathbf{0.99}$ & $0.00$ & $0.02$ & $0.00$ & $0.02$ & $0.00$ & $0.02$ & $0.97$ & $\mathbf{0.99}$ & $0.62$ & $0.85$ & $0.00$ & $0.85$ & $0.00$ & $0.97$ & $0.00$ & ${0.97}$ & $0.00$\ & ML & $0.89$ & $0.00$ & $0.89$ & $0.00$ & $0.14$ & $0.00$ & $0.14$ & $0.00$ & $0.10$ & $0.83$ & $0.93$ & $0.44$ & $0.96$ & $0.00$ & $0.96$ & $0.00$ & $0.98$ & $0.00$ & $\mathbf{0.98}$ & $0.00$\ & LL & $0.83$ & $0.00$ & $0.83$ & $0.00$ & $0.12$ & $0.00$ & $0.12$ & $0.00$ & $0.12$ & $0.83$ & $\mathbf{0.95}$ & $0.50$ & $0.94$ & $0.00$ & $0.94$ & $0.00$ & $0.95$ & $0.00$ & $\mathbf{0.95}$ & $0.00$\ & ML & $0.28$ & $0.00$ & $0.28$ & $0.00$ & $0.63$ & $0.00$ & $0.63$ & $0.00$ & $0.26$ & $0.40$ & $0.66$ & $0.18$ & $0.94$ & $0.00$ & $\mathbf{0.94}$ & $0.00$ & $0.94$ & $0.00$ & $\mathbf{0.94}$ & $0.00$\ & LL & $0.41$ & $0.00$ & $0.41$ & $0.00$ & $0.68$ & $0.00$ & $0.68$ & $0.00$ & $0.41$ & $0.41$ & $0.82$ & $0.15$ & $0.89$ & $0.00$ & $0.89$ & $0.00$ & $0.92$ & $0.00$ & $\mathbf{0.92}$ & $0.00$\ & ML & $0.82$ & $0.00$ & $0.82$ & $0.00$ & $0.94$ & $0.00$ & $0.94$ & $0.00$ & $0.80$ & $0.16$ & $0.96$ & $0.12$ & $0.98$ & $0.00$ & $\mathbf{0.98}$ & $0.00$ & $0.98$ & $0.00$ & $\mathbf{0.98}$ & $0.00$\ & LL & $0.65$ & $0.00$ & $0.65$ & $0.00$ & $0.81$ & $0.00$ & $0.81$ & $0.00$ & $0.64$ & $0.22$ & $0.86$ & $0.14$ & $0.88$ & $0.00$ & $0.88$ & $0.00$ & $0.90$ & $0.00$ & $\mathbf{0.90}$ & $0.00$\ & $0.73$ & $0.00$ & $0.73$ & $0.00$ & $0.38$ & $0.00$ & $0.38$ & $0.00$ & $0.28$ & $0.58$ & $0.86$ & $0.32$ & $0.89$ & $0.00$ & $0.89$ & $0.00$ & $0.93$ & $0.00$ & $\mathbf{0.94}$ & $0.00$\ & $0.23$ & $0.00$ & $0.23$ & $0.00$ & $0.35$ & $0.00$ & $0.35$ & $0.00$ & $0.26$ & $0.28$ & $0.11$ & $0.18$ & $0.09$ & $0.00$ & $0.09$ & $0.00$ & $0.05$ & $0.01$ & $\mathbf{0.04}$ & $0.00$\ \ & **PSR** & **TSR** & **DSR** & **FP** & **PSR** & **TSR** & **DSR** & **FP** & **PSR** & **TSR** & **DSR** & **FP** & **PSR** & **TSR** & **DSR** & **FP** & **PSR** & **TSR** & **DSR** & **FP**\ & $0.9304$ & $0.00$ & $0.9304$ & $0.00$ & $0.8835$ & $0.00$ & $0.8835$ & $0.00$ & $0.8772$ & $0.0656$ & $0.9428$ & $0.0369$ & ${0.9608}$ & $0.00$ & ${0.9608}$ & $0.00$ & $\mathbf{0.9631}$ & $0.0002$ & $0.9633$ & $0.0002$\ FGSM & & $0.19$ & $0.00$ & $0.19$ & $0.00$ & $0.34$ & $0.00$ & $0.34$ & $0.00$ & $0.17$ & $0.24$ & $0.41$ & $0.12$ & $0.85$ & $0.00$ & $0.85$ & $0.00$ & $0.83$ & $0.08$ & $\mathbf{0.91}$ & $0.00$\ BIM & & $0.22$ & $0.00$ & $0.22$ & $0.00$ & $0.50$ & $0.00$ & $0.50$ & $0.00$ & $0.21$ & $0.33$ & $0.54$ & $0.13$ & $0.92$ & $0.00$ & $0.92$ & $0.00$ & $0.95$ & $0.01$ & $\mathbf{0.96}$ & $0.00$\ & ML & $0.21$ & $0.00$ & $0.21$ & $0.00$ & $0.67$ & $0.00$ & $0.67$ & $0.00$ & $0.21$ & $0.47$ & $0.68$ & $0.22$ & $0.95$ & $0.00$ & $0.95$ & $0.00$ & $0.98$ & $0.00$ & $\mathbf{0.98}$ & $0.00$\ & LL & $0.83$ & $0.00$ & $0.83$ & $0.00$ & $0.85$ & $0.00$ & $0.85$ & $0.00$ & $0.77$ & $0.17$ & $0.94$ & $0.11$ & $0.98$ & $0.00$ & $\mathbf{0.98}$ & $0.00$ & $0.98$ & $0.00$ & $\mathbf{0.98}$ & $0.00$\ DF & UA & $0.64$ & $0.00$ & $0.64$ & $0.00$ & $0.78$ & $0.00$ & $0.78$ & $0.00$ & $0.59$ & $0.25$ & $0.84$ & $0.15$ & $0.97$ & $0.00$ & $0.97$ & $0.00$ & $0.98$ & $0.00$ & $\mathbf{0.98}$ & $0.00$\ & ML & $0.42$ & $0.00$ & $0.42$ & $0.00$ & $0.75$ & $0.00$ & $0.75$ & $0.00$ & $0.42$ & $0.34$ & $0.76$ & $0.22$ & $0.95$ & $0.00$ & $0.95$ & $0.00$ & $0.98$ & $0.00$ & $\mathbf{0.98}$ & $0.00$\ & LL & $0.87$ & $0.00$ & $0.87$ & $0.00$ & $0.90$ & $0.00$ & $0.90$ & $0.00$ & $0.82$ & $0.14$ & $0.96$ & $0.09$ & $0.97$ & $0.00$ & ${0.97}$ & $0.00$ & $0.98$ & $0.00$ & $\mathbf{0.98}$ & $0.00$\ & ML & $0.03$ & $0.00$ & $0.03$ & $0.00$ & $0.33$ & $0.00$ & $0.33$ & $0.00$ & $0.03$ & $0.33$ & $0.36$ & $0.10$ & $0.88$ & $0.00$ & $0.88$ & $0.00$ & $0.86$ & $0.08$ & $\mathbf{0.94}$ & $0.00$\ & LL & $0.21$ & $0.00$ & $0.21$ & $0.00$ & $0.70$ & $0.00$ & $0.70$ & $0.00$ & $0.18$ & $0.76$ & $\mathbf{0.94}$ & $0.34$ & $0.91$ & $0.00$ & $0.91$ & $0.00$ & $0.93$ & $0.00$ & ${0.93}$ & $0.00$\ & ML & $0.40$ & $0.00$ & $0.40$ & $0.00$ & $0.77$ & $0.00$ & $0.77$ & $0.00$ & $0.39$ & $0.45$ & $0.84$ & $0.31$ & $0.83$ & $0.00$ & $0.83$ & $0.00$ & $0.92$ & $0.00$ & $\mathbf{0.92}$ & $0.00$\ & LL & $0.32$ & $0.00$ & $0.32$ & $0.00$ & $0.71$ & $0.00$ & $0.71$ & $0.00$ & $0.31$ & $0.58$ & $\mathbf{0.89}$ & $0.38$ & $0.54$ & $0.00$ & $0.54$ & $0.00$ & $0.79$ & $0.01$ & ${0.80}$ & $0.00$\ & $0.39$ & $0.00$ & $0.39$ & $0.00$ & $0.66$ & $0.00$ & $0.66$ & $0.00$ & $0.37$ & $0.37$ & $0.74$ & $0.20$ & $0.89$ & $0.00$ & $0.89$ & $0.00$ & $0.93$ & $0.02$ & $\mathbf{0.94}$ & $0.00$\ & $0.27$ & $0.00$ & $0.27$ & $0.00$ & $0.19$ & $0.00$ & $0.19$ & $0.00$ & $0.25$ & $0.18$ & $0.21$ & $0.10$ & $0.12$ & $0.00$ & $0.12$ & $0.00$ & $0.07$ & $0.03$ & $\mathbf{0.05}$ & $0.00$\ [**Denoising Ensemble Defense.**]{} Recall Table \[tab:attack-evaluation\] discussed in Section \[sec:attacks\], we have shown that most of the attacks successfully fool the target model by $100\%$ attack success rate (ASR) or by reducing its test accuracy significantly to zero (i.e., the target model is completely defeated with $100\%$ misclassification rate, $\text{MR}=1.00$). Table \[tab:exp-denoiser\] shows three important results: First, by employing an unsupervised DNN denoiser, either trained with Gaussian or salt-and-pepper noise, one can improve the robustness of the target model, as shown in the columns under “Denoiser (Gaussian Noise)” and “Denoiser (Salt-and-Pepper Noise)”. Second, even though some attacks are mitigated with a high DSR, the effectiveness of one denoiser varies wildly across different attacks and different datasets. For instance, the Gaussian denoiser on MNIST achieves a high DSR of $0.99$ in defending the CW$_\infty$ LL attack but only gets a low DSR of $0.28$ under the CW$_0$ ML attack. In fact, we have performed experiments on multiple denoisers, and we found consistently that no DNN denoiser can remove adversarial perturbations from all eleven attacks examined and each denoiser fails to generalize well over all attacks. Third, the DSR of model denoising ensemble is consistently better than using one denoiser as shown in the columns under “Model Denoising Ensemble Defense”. This validates our argument that a robust defense method should not rely on a single defense strategy, such as one denoiser as adopted in prior works [@sahay2019combatting; @liao2018defense; @gu2014towards] but employ a denoising ensemble. Also, a simple denoising ensemble as those advocated in [@meng2017magnet] may not outperform a diversity optimized denoising ensemble, as an ensemble team of diverse denoisers offers different denoising effects and hence generalizes better over different adversarial attacks. [**Verification Ensemble and Cross-Layer Ensemble Defense.**]{} Table \[tab:exp-denoiser\] also reports the defensibility of model verification ensemble under the column “Model Verification Ensemble Defense”. Compared with the denoising ensemble, although the verification ensemble achieves a higher average DSR over all eleven attacks in both datasets (i.e., $0.89$ v.s. $0.86$ for MNIST and $0.89$ v.s. $0.74$ for CIFAR-10 shown on their respective “Average” row in Table \[tab:exp-denoiser\]), it performs much worse than denoising ensemble under the JSMA LL attack ($0.54$ v.s. $0.89$) for CIFAR-10 and CW$_\infty$ LL attack ($0.85$ v.s. $0.99$) for MNIST. The model verification ensemble is hence a competitive alternative to the denoising ensemble, since neither of them is the winner over all attacks on both datasets, and yet each of them is the winner for some attacks. Compared with denoising or verification ensemble only defense, the cross-layer ensemble defense consistently outperforms across almost all attacks with a remarkably small standard deviation of DSRs (see the columns of “Denoising-Verification Cross-Layer Ensemble Defense” in Table \[tab:exp-denoiser\]). In other words, the cross-layer ensemble can generalize well over the eleven attacks examined. We make another interesting observation that although the model verification ensemble and denoising-verification cross-layer ensemble may have slightly lower benign accuracy ($0.9888$ and $0.9855$) than the target model on MNIST that they set to protect ($0.9943$) under no attack scenario, their benign accuracy on CIFAR-10 are $0.9608$ and $0.9631$, which are higher than the target model accuracy of $0.9484$. This is consistent with the theoretical proof in [@Holloway2007-ensemble] that ensemble can reduce uncorrelated errors. For example, if we have $5$ completely independent classifiers and we use unweighted majority voting as a consensus algorithm, by the probability formula, with accuracy of $70\%$ for each classifier, we have $83.7\%$ ensemble accuracy with majority voting and if we have $101$ such classifiers, we reach $99.9\%$ ensemble accuracy by majority voting. Equivalently, let $\delta$ be the classification accuracy of each member in the ensemble $E$ of size $Z$, the accuracy can be boosted to $P(E\geq\left\lceil Z/2\right\rceil)=\sum_{i=\left\lceil Z/2\right\rceil}^{Z}{Z \choose i}\delta^i(1-\delta)^{Z-i}$. In [MODEF]{}, we consider three options to create an ensemble team of base models: (1) randomly selecting a team from the entire pool of ensemble combinations (**Rand**); (2) randomly selecting an ensemble team from the $\kappa$-ranked list (**Rand$\boldsymbol{\kappa}$**), which filters out those ensemble team combos that have high kappa values above the system-defined threshold (due to the space, the factors impact on the decision for the threshold setting is omitted here); and (3) selecting the ensemble team from the $\kappa$-ranked list with the highest prediction accuracy (**Best$\boldsymbol{\kappa}$**). The $\kappa$-ranked list is built on the kappa value of each pair of model verifiers. Different levels of disagreement occur between a pair of model verifiers. Table \[fig:kappa-mnist\] gives an example of the pairwise kappa values of the model verifiers on MNIST where the pair of $\text{V}_4$ and $\text{V}_5$ has a low kappa value of $0.25$, showing their high prediction disagreement diversity (high failure-independence and low error correlation). **TM** **$\text{V}_1$** **$\text{V}_2$** **$\text{V}_3$** **$\text{V}_4$** **$\text{V}_5$** **$\text{V}_6$** **$\text{V}_7$** **$\text{V}_8$** **$\text{V}_9$** **$\text{V}_{10}$** --------------------- -------- ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ --------------------- **TM** $0.79$ $0.65$ $0.78$ $0.66$ $0.37$ $0.79$ $0.78$ $0.81$ $0.58$ $0.78$ **$\text{V}_1$** $0.67$ $0.81$ $0.65$ $0.47$ $0.77$ $0.77$ $0.76$ $0.69$ $0.75$ **$\text{V}_2$** $0.67$ $0.50$ $0.39$ $0.62$ $0.62$ $0.62$ $0.56$ $0.60$ **$\text{V}_3$** $0.64$ $0.48$ $0.76$ $0.77$ $0.75$ $0.66$ $0.74$ **$\text{V}_4$** $0.25$ $0.69$ $0.71$ $0.67$ $0.47$ $0.71$ **$\text{V}_5$** $0.35$ $0.36$ $0.36$ $0.43$ $0.31$ **$\text{V}_6$** $0.84$ $0.83$ $0.56$ $0.82$ **$\text{V}_7$** $0.86$ $0.58$ $0.84$ **$\text{V}_8$** $0.54$ $0.82$ **$\text{V}_9$** $0.54$ **$\text{V}_{10}$** : The pairwise kappa value of model verifiers for MNIST.[]{data-label="fig:kappa-mnist"} To understand the effectiveness of three different ensemble teaming options, we report in Table \[fig:exp-teaming-mnist\] the defensibility with the ensembles randomly sampled according to the above options. We make three observations. First, the target model has benign accuracy of $0.9943$ under no attack scenario and suffers partial failure under FGSM and JSMA LL attacks, close to zero accuracy under BIM and JSMA ML attacks, and zero accuracy under all six CW attacks. In comparison, each of the model verifiers consistently provides better robustness against most of the attacks than that of the target model. For instance, each verifier has higher test accuracy (i.e., higher DSR) under CW attacks, JSMA attacks, and BIM attacks and comparable performance under FGSM, and comparable benign accuracy under no attack. Second, verification ensemble defense has better robustness compared to each of the individual verifiers. For each verification ensemble, its average DSR is higher than that of individual ensemble member. Third, using an ensemble team from the $\kappa$-ranked list performs better in many cases than random ensemble selection because the ensembles in the $\kappa$-ranked list have low kappa value and thus high model diversity with respect to failure-independence or error correlation. With the cross-layer ensemble defense, the above observations also hold and the denoising-verification cross-layer ensembles significantly outperform their corresponding verification ensembles with the same team and performs much better than individual ensemble member. One example is the Rand$\kappa$ team with $\text{V}_3, \text{V}_4, \text{V}_{10}$ having an average DSR over all attacks of $0.72$, $0.62$, and $0.48$ respectively while the cross-layer ensemble method makes a significant improvement to $0.89$, demonstrating the complementary property of denoising and verification ensembles. [@ @ plus 1fil minus @skip]{} -- ------------------------------------------------------------------------------------ ---------- ----------------- ----------------- ----------------- ----------------- ----------------- -------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- **FGSM** **BIM** **ML** **LL** **ML** **LL** **ML** **LL** **ML** **LL** $0.9943$ $0.54$ $0.08$ $0.00$ $0.00$ $0.00$ $0.00$ $0.00$ $0.00$ $0.07$ $0.47$ $0.12$ $0.21$ $0.9919$ $0.58$ $0.27$ $0.59$ $0.43$ $0.74$ $0.78$ $0.72$ $0.72$ $0.90$ $0.81$ $0.65$ $0.19$ $0.9861$ $0.65$ $0.44$ $0.85$ $0.68$ $0.85$ $0.74$ $0.60$ $0.57$ $0.76$ $0.63$ $0.68$ $0.13$ $0.9917$ $0.65$ $0.37$ $0.81$ $0.71$ $0.82$ $0.89$ $0.73$ $0.65$ $0.86$ $0.73$ $0.72$ $0.15$ $\mathbf{0.9980}$ $0.49$ $0.44$ $0.72$ $0.31$ $0.72$ $0.63$ $0.70$ $0.57$ $0.86$ $0.74$ $0.62$ $0.17$ $0.9761$ $0.51$ $0.44$ $0.74$ $0.61$ $0.88$ $0.87$ $0.83$ $0.81$ $0.85$ $0.74$ $0.73$ $0.16$ $0.9934$ $0.19$ $0.18$ $0.28$ $0.16$ $0.49$ $0.36$ $0.76$ $0.55$ $0.95$ $0.87$ $0.48$ $0.30$ $0.9939$ $0.22$ $0.23$ $0.36$ $0.11$ $0.34$ $0.14$ $0.70$ $0.44$ $0.96$ $0.86$ $0.44$ $0.30$ $0.9950$ $0.15$ $0.15$ $0.19$ $0.08$ $0.20$ $0.10$ $0.68$ $0.43$ $0.94$ $0.80$ $0.37$ $0.32$ $0.9832$ $0.67$ $0.72$ $0.98$ $0.92$ $\mathbf{0.98}$ $\mathbf{0.97}$ $0.93$ $0.86$ $0.96$ $0.80$ $0.88$ $0.11$ $0.9946$ $0.40$ $0.32$ $0.59$ $0.03$ $0.49$ $0.35$ $0.62$ $0.32$ $0.89$ $0.80$ $0.48$ $0.25$ Rand = $\text{V}_{1}$,$\text{V}_{2}$,$\text{V}_{4}$,$\text{V}_{7}$,$\text{V}_{10}$ $0.9951$ $0.61$ $0.39$ $0.78$ $0.51$ $0.76$ $0.69$ $0.79$ $0.68$ $0.94$ $0.80$ $0.70$ $0.16$ Rand$\kappa$ = $\text{V}_{3}$,$\text{V}_{4}$,$\text{V}_{10}$ $0.9919$ $0.63$ $0.42$ $0.84$ $0.65$ $0.88$ $0.88$ $0.80$ $0.77$ $0.92$ $0.80$ $0.76$ $0.15$ Best$\kappa$ = $\text{V}_{5}$,$\text{V}_{6}$,$\text{V}_{9}$ $0.9888$ $0.71$ $0.78$ $0.93$ $0.85$ $0.96$ $0.94$ $\mathbf{0.94}$ $0.89$ $\mathbf{0.98}$ $0.88$ $0.89$ $0.09$ Rand = $\text{V}_{1}$,$\text{V}_{2}$,$\text{V}_{4}$,$\text{V}_{7}$,$\text{V}_{10}$ $0.9912$ $\mathbf{0.90}$ $0.76$ $0.95$ $\mathbf{0.98}$ $0.86$ $0.81$ $0.83$ $0.84$ $0.95$ $0.85$ $0.87$ $0.07$ Rand$\kappa$ = $\text{V}_{3}$,$\text{V}_{4}$,$\text{V}_{10}$ $0.9873$ $0.84$ $0.71$ $0.98$ $0.94$ $0.96$ $0.92$ $0.88$ $0.88$ $0.95$ $0.88$ $0.89$ $0.08$ Best$\kappa$ = $\text{V}_{5}$,$\text{V}_{6}$,$\text{V}_{9}$ $0.9855$ $0.89$ $\mathbf{0.86}$ $\mathbf{0.99}$ $0.97$ $\mathbf{0.98}$ $0.95$ $\mathbf{0.94}$ $\mathbf{0.92}$ $\mathbf{0.98}$ $\mathbf{0.90}$ $\mathbf{0.94}$ $\mathbf{0.04}$ -- ------------------------------------------------------------------------------------ ---------- ----------------- ----------------- ----------------- ----------------- ----------------- -------- ----------------- ----------------- ----------------- ----------------- ----------------- ----------------- **TM** **$\text{V}_1$** **$\text{V}_2$** **$\text{V}_3$** **$\text{V}_4$** **$\text{V}_5$** **$\text{V}_6$** **$\text{V}_7$** **$\text{V}_8$** **$\text{V}_9$** **$\text{V}_{10}$** ------ -------- ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ ------------------ --------------------- -------- ----------------- ----------------- FGSM $1.00$ $0.72$ $0.63$ $0.74$ $0.74$ $0.67$ $0.76$ $0.76$ $0.83$ $0.57$ $0.70$ $0.63$ $\mathbf{0.24}$ BIM $1.00$ $0.79$ $0.61$ $0.68$ $0.61$ $0.61$ $0.83$ $0.78$ $0.86$ $0.30$ $0.71$ $0.24$ $\mathbf{0.15}$ ML $1.00$ $0.34$ $0.13$ $0.19$ $0.21$ $0.14$ $0.21$ $0.35$ $0.36$ $0.02$ $0.23$ $0.06$ $\mathbf{0.01}$ LL $1.00$ $0.28$ $0.06$ $0.10$ $0.15$ $0.07$ $0.04$ $0.30$ $0.16$ $0.00$ $0.15$ $0.03$ $\mathbf{0.00}$ ML $1.00$ $0.24$ $0.10$ $0.18$ $0.22$ $0.08$ $0.35$ $0.55$ $0.61$ $0.02$ $0.39$ $0.04$ $\mathbf{0.01}$ LL $1.00$ $0.03$ $0.04$ $0.01$ $0.06$ $0.02$ $0.07$ $0.45$ $0.45$ $0.00$ $0.21$ $0.02$ $\mathbf{0.00}$ ML $1.00$ $0.25$ $0.22$ $0.24$ $0.27$ $0.11$ $0.20$ $0.20$ $0.25$ $0.03$ $0.34$ $0.05$ $\mathbf{0.03}$ LL $1.00$ $0.06$ $0.12$ $0.14$ $0.06$ $0.02$ $0.11$ $0.16$ $0.19$ $0.01$ $0.16$ $\mathbf{0.00}$ $\mathbf{0.00}$ ML $1.00$ $0.06$ $0.13$ $0.13$ $0.10$ $0.08$ $0.05$ $0.02$ $0.02$ $0.01$ $0.08$ $\mathbf{0.01}$ $\mathbf{0.01}$ LL $1.00$ $0.07$ $0.19$ $0.16$ $0.05$ $0.09$ $0.00$ $0.00$ $0.00$ $0.02$ $0.00$ $0.02$ $\mathbf{0.00}$ $1.00$ $0.28$ $0.22$ $0.26$ $0.25$ $0.19$ $0.26$ $0.36$ $0.37$ $0.10$ $0.30$ $0.10$ $\mathbf{0.05}$ $0.00$ $0.27$ $0.22$ $0.25$ $0.24$ $0.24$ $0.30$ $0.28$ $0.31$ $0.19$ $0.24$ $0.20$ $\mathbf{0.08}$ While the above shows that all three MODEF ensembles (denoising ensemble, verification ensemble and denoising-verification cross-layer ensemble) improve the test accuracy of the target model under MODEF protection for all eleven attacks on the two benchmark datasets, this indicates that ensembles with good diversity hold great potential for improving robustness of DNN models in the presence of adversarial inputs. We would like to note that DSR is computed by the sum of the prevention success rate (PSR) and the detection success rate (TSR) as defined in Section \[sec:attacks\]. PSR indicates the ability of a defense to repair adversarial examples and TSR indicates the capability of a defense to flag those adversarial examples that it cannot repair successfully. We notice that model verification ensemble offers a stronger ability to repair adversarial examples with correct prediction outcomes while model denoising ensemble tends to be more robust by flagging and rejecting adversarial examples. The cross-layer ensemble advocates high robustness with a high DSR by achieving a high PSR, which is in stark contrast to existing detection-only methods [@grosse2017statistical; @metzen2017detecting]. The final remark is related to the transferability of adversarial examples. To gain a better understanding of how the transferability of adversarial examples behaves under [MODEF]{}, we measure the attack transferability in ASR of the target model and the ASR of the ten model verifiers. Due to the space limit, we show only the results on MNIST in Table \[fig:exp-transfer-mnist\]. Given that the adversarial examples are generated over the TM, thus the TM column shows the transferability is $1.00$. The adversarial examples generated over the target model (TM) may not be transferable to other independently trained models consistently. For instance, the adversarial examples generated by the CW$_2$ ML attack can be effectively transferred to $\text{V}_7$ and $\text{V}_8$ with a transferability of $0.55$ and $0.61$ respectively but achieve only a low transferability of $0.10$ for $\text{V}_2$, $0.08$ for $\text{V}_5$, and $0.02$ for $\text{V}_9$. Such divergence in transferability of adversarial examples is the main motivation of MODEF to exploit the weak spots of attack transferability using the denoising ensemble, the verification ensemble and the denoising-verification cross-layer ensemble as effective defense methods. As shown in the last two columns in Table \[fig:exp-transfer-mnist\], the model verification ensemble for MNIST has a very low transferability under all attacks. With cross-layer ensemble that combines the denoising ensemble of two denoisers and verification ensemble of three model verifiers ($\text{V}_5$,$\text{V}_6$,$\text{V}_9$), its attack transferability becomes further reduced. Conclusion {#sec:conclusion} ========== We have presented [MODEF]{}, a diversity ensemble defense framework against adversarial deception in deep learning. Driven by quantifying ensemble diversity based on classification failure-independence, [MODEF]{} provides three diversity ensemble methods: (1) the model denoising ensemble, leveraging multiple diverse denoising autoencoders to remove adversarial perturbations, (2) the model verification ensemble that exploits the weak spots of attack transferability to verify and repair prediction output of the target model, and (3) the denoising-verification cross-layer ensemble, which can guard input and output of the target model through intelligently combining input denoising ensemble and output verification ensemble. Furthermore, [MODEF]{} is by design attack-independent and can generalize over different attacks. Due to the space constraint, we could not include additional experimental results on the effectiveness of [MODEF]{} under new attack algorithms including PGD [@madry2017towards] and EAD [@chen2018ead]. Our ongoing work continues along two threads: First, we will explore different ensemble diversity measures and understand their roles in protecting a target DNN model from adversarial attacks under MODEF [@ling2019ensemble]. Second, we are working on defenses against deceptions against data modality beyond images and machine learning models beyond CNN classifiers. Acknowledgement {#acknowledgement .unnumbered} =============== This work was a part of XDefense umbrella project for defining diversity ensemble against deception and the MODEF system development has benefited from the strategic teaming of input ensemble and output ensemble system, led by Wenqi Wei [@wei2018adversarial; @wei2019demystifying; @wei2019strategic]. This research is partially sponsored by National Science Foundation under CISE SaTC (1564097), an IBM Faculty Award, and a Georgia Tech IISP grant. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the National Science Foundation or other funding agencies and companies mentioned above. [^1]: https://colab.research.google.com/
{ "pile_set_name": "ArXiv" }
--- abstract: 'We study a portfolio optimization problem for competitive agents with CRRA utilities and a common finite time horizon. The utility of an agent depends not only on her absolute wealth and consumption but also on her relative wealth and consumption when compared to the averages among the other agents. We derive a closed form solution for the $n$-player game and the corresponding mean field game. This solution is unique in the class of equilibria with constant investment and continuous time-dependent consumption, both independent of the wealth of the agent. Compared to the classical Merton problem with one agent, the competitive model exhibits a wide range of highly nonlinear and non-monotone dependence on the agents’ risk tolerance and competitiveness parameters. Counter-intuitively, competitive agents with high risk tolerance may behave like non-competitive agents with low risk tolerance.' address: - 'Industrial Engineering & Operations Research, Columbia University, New York, NY.' - 'Industrial Engineering & Operations Research, Columbia University, New York, NY.' author: - Daniel Lacker - Agathe Soret bibliography: - 'biblio.bib' title: 'Many-player games of optimal consumption and investment under relative performance criteria' --- Introduction ============ In this paper, we extend the CRRA model for the optimal investment problem recently developed by Lacker and Zariphopoulou [@lacker2017mean] to include consumption. Our model can be loosely described as follows, with full details given in Section \[finite population\]. Each agent chooses consumption and investment policies, with access to a riskless bond and a lognormal stock. The stocks in which the different agents specialize can be correlated, and we cover the extreme case of perfect correlation, i.e., a single stock in which all agents trade. Each agent has a CRRA utility depending on both absolute and relative wealth at the (common) time horizon $T$ as well as absolute and relative consumption, the latter in a time-integrated sense. Agents have different levels of risk aversion and different preferences toward absolute versus relative performance. The relative performance criteria couple these $n$ optimization problems, and we find the (unique, in a sense to be clarified later) Nash equilibrium in terms of the various model parameters. As is natural in light of the classical Merton problem [@merton1969lifetime], in equilibrium each agent invests a constant fraction of wealth in the stock, and the consumption strategy is time-dependent but independent of the agent’s wealth. The equilibrium behavior fits with Samuelson’s result [@samuelson1975lifetime]: the investment strategy is independent of the consumption strategy. That is, the investment strategy is exactly the same as in the model without consumption studied in [@lacker2017mean]. The equilibrium consumption policy, as a function of the various model parameters, displays even more highly nonlinear and non-monotone behavior than the investment policy, and we study this in detail in Section \[se:discussion\]. Notably, each agent’s rate of consumption $c_t$ changes monotonically with time $t$ over the entire horizon $[0,T]$; however, whether an agent increases or decreases consumption over time depends in a complex manner on her own risk preferences as well as certain aggregates of the other agents’ parameters. Three key features of our model are relative consumption concerns, relative wealth concerns, and asset specialization. We defer to the introduction of [@lacker2017mean] for a thorough discussion of the latter two topics and further references, but we stress the particularly important and by now well-established point that mutual fund choice is highly influenced by relative performance [@sirri1998costly]. That is, out-performing other fund managers tends to attract greater future investment in one’s own fund. The most closely related works to ours, after [@lacker2017mean], are [@anthropelos-geng-zariphopoulou; @basak2015competition; @bielagk2017equilibrium; @dosreis-zariphopoulou; @espinosa2015optimal; @frei2011financial]. These papers study continuous-time models of optimal investment under relative performance concerns in various settings, including different kinds of utilities, equilibrium pricing, and state constraints, but none incorporate consumption. There are two natural arguments for studying relative consumption concerns. On the one hand, interpreting agents as fund managers, we may think of consumption as capital accumulation, in the form of equipment, technology, or benefits for employees. A high relative consumption in this sense would naturally attract clientele or better employees, and more generally it should lead to similar benefits as a high relative wealth. On the other hand, if we interpret the agents in our model as household investors, then relative consumption concerns fit naturally with models of *keeping up with the Joneses*; this line of literature directly incorporates the social aspects of investment and consumption decisions [@abel1990asset; @chan2002catching; @demarzo2004diversification; @gali1994keeping]. Our paper contributes to the literatures on optimal consumption and investment as well as the application of mean field games. The dynamic problem of lifetime consumption and investment planning with one player was formalized and studied in the landmark papers of Merton [@merton1969lifetime; @merton1975optimum] and Samuelson [@samuelson1975lifetime]. Later work incorporated more complex features into the models, such as general price processes, bankruptcy, etc. [@karatzas1987optimal; @karatzas2003optimal]. Those that incorporated *multiple agents* into the model, such as [@karatzas1997explicit; @sethi1992infinite], did so in an equilibrium context; each agent’s behavior depends on the others only through the price, which is determined in equilibrium. Agents are price-takers in our model, and we do not attempt to incorporate price equilibrium, as this would severely strain tractability. In another direction, our work provides a new explicitly solvable mean field game model. Mean field games, introduced in [@huang2006large; @lasry-lions], are rarely explicitly solvable outside of linear-quadratic examples. See [@bayraktar2018large; @carmona2018dyson; @gueant-lasry-lions; @lacker2017mean; @sun2016systemic] for some notable exceptions and the book [@carmona-delarue-book] for further background on the active area of mean field games. From a mean field game perspective, our model is rather complex: It involves common noise, degenerate volatility coefficients, singular objective functions, and a mean field interaction through both the states and controls (i.e., an *extended mean field game* [@carmona-delarue-book Chapter I.4.6]). Nevertheless, the precise structure of the problem lends itself to an explicit solution. Our argument follows along the lines of [@lacker2017mean], treating the mean field term (geometric mean of wealth) as a state variable, which leads to a fixed point problem involving a single Hamilton-Jacobi-Bellman (HJB) equation as opposed to the $n$-dimensional HJB system often used for stochastic differential games. After showing that this equation admits a unique and separable classical solution, the fixed point is resolved via a system of non-linear ordinary differential equations. Despite many similarities with [@lacker2017mean], the consumption renders the arguments substantially more involved. The paper is organized as follows. In Section \[finite population\], we formulate and solve the $n$-agent model described above. Then, in Section $3$, we study the infinite population counterpart of this problem, arguing that the $n\to\infty$ limit results in a simpler form of the equilibrium. Finally, Section $4$ discusses and interprets the form of the equilibrium and its dependence on the model parameters. The n-agent game {#finite population} ================ In this section, we consider the $n$-player game, where each agent trades in a common investment horizon $[0,T]$. Agents may invest in their own specific stocks or in a common riskless bond which offers zero interest rate. The price of stock $i$, in which only agent $i$ trades, is given by the dynamics $$\label{prices} \frac{dS^i_t}{S^i_t} = \mu_i dt + \nu_i dW^i_t + \sigma_i dB_t,$$ where the Brownian motions $W^1, ..., W^n, B$ are independent and defined on a probability space $(\Omega, \F, \PP)$, which we endow with the natural filtration $(\F_t)_{t \in [0,T]}$ generated by these $n+1$ Brownian motions, and where the market parameters are constants $\mu_i > 0$, $\sigma_i \ge 0$, and $\nu_i \ge 0$, with $\sigma_i + \nu_i > 0$. The prices $S^i_t$ are assumed to be one-dimensional for simplicity, although we could easily extend the results to $k$-dimensional prices. This setup covers the important special case of a *single stock*, corresponding to the situation where all the stocks are identical. That is, $\mu_i = \mu$, $\nu_i = 0$, and $\sigma_i = \sigma$, for all $i = 1, ..., n$ and for some $\mu, \sigma > 0$ independent of $i$, and so $S^i \equiv S^j$ for each $i,j$ (assuming the initial values agree). In the single stock case, the agents face the same market opportunities rather than specializing in different assets, though their risk preferences still differ. Each agent $i$ chooses a self-financing strategy, $(\pi^i_t)_{t \in [0,T]}$, denoting the proportion of wealth invested in the stock $i$, and a consumption policy, $(c_t^i)_{t \in [0,T]}$. The wealth process of agent $i$ is then given by $$\label{processX} dX^i_t = \pi^i_t X_t^i \left(\mu_i dt + \nu_i dW^i_t + \sigma_i dB_t\right) - c^i_t X^i_t dt, \quad\quad X_0^i = x_0^i.$$ Note that $c^i_tX^i_t$ represents the instantaneous rate of consumption of agent $i$, so that $c^i_t$ is the rate per unit wealth. We say that a portfolio strategy is admissible if it belongs to the set $\A$ of $\FF$-progressively measurable $\R \times \R_+$-valued process $(\pi_t, c_t)_{t \in [0,T]}$ satisfying $\E \int_0^T (\pi_t^2 + c_t^2)dt < \infty$. Throughout the paper, $\R_+ :=(0,\infty)$ denotes the *strictly* positive reals, and we do not allow a consumption rate of zero. This is reasonable and less restrictive than it may at first appear because the form of our utility functions, introduced in the next paragraph, will ensure that agents’ marginal utilities approach $+\infty$ as consumption approaches zero. Note also that for any admissible portfolio strategy we have $X^i_t > 0$ for all $t \in [0,T]$. Indeed, we parametrized our consumption process as we did in in part to avoid the possibility of bankruptcy and in part to avoid imposing any state constraints. The utility function of each agent belongs to the family of power (CRRA) utilities, $$U(x; \delta) = \begin{cases} \frac{1}{1 - 1 / \delta}x^{1 - 1 / \delta} & \text{ if } \delta \ne 1 \\ \log x & \text{ if } \delta = 1, \end{cases}$$ defined for $x,\delta > 0$. Agent $i$ seeks to maximize the expected utility $$\begin{aligned} J_i((\pi^i,c^i)_{i=1}^n) = \E\left[\int_0^T U\left(c^i_tX^i_t (\overline{cX}_t)^{- \theta_i}; \delta_i \right)dt + \epsilon_i U\left(X^i_T \overline{X}_T^{- \theta_i}; \delta_i \right)\right], \label{def:utility}\end{aligned}$$ defined for any vector of admissible strategies $(\pi^i,c^i)_{i=1}^n$ where $(\pi^i,c^i) \in \A$ for each $i=1,\ldots,n$. Here $\overline{X}_T = \left( \prod_{k = 1}^{n} X_T^k \right)^{1/n}$ and $\overline{cX}_t = \left( \prod_{k = 1}^{n} (c_tX_t)^k \right)^{1/n}$ are the population (geometric) average wealth and consumption rate, respectively. The parameters $\delta_i > 0$ and $\theta_i \in [0,1]$ represent respectively the $i^{\mathrm{th}}$ agent’s risk tolerance and competition weight. We apply the same utility function to both wealth and consumption for tractability reasons, but we scale the utility of wealth with the parameter $\epsilon_i > 0$ to capture the relative importance that the agent assigns to terminal wealth compared to consumption. Our choice to work with geometric averages instead of arithmetic averages is also motivated by tractability: Just as the (arithmetic) average of independent Brownian motions is again a Brownian motion, the geometric average of independent geometric Brownian motions is again a geometric Brownian motion. It is worth emphasizing another form of the utility function, revealed by writing the terms inside the utility function as $$\begin{aligned} c^i_tX^i_t (\overline{cX}_t)^{- \theta_i} = (c^i_tX^i_t)^{1-\theta_i}\left(\frac{c^i_tX_t^i}{\overline{cX}_t}\right)^{\theta_i}, \quad\quad X^i_T \overline{X}_T^{- \theta_i} = (X^i_T)^{1-\theta_i}\left(\frac{X_T^i}{\overline{X}_T}\right)^{\theta_i}.\end{aligned}$$ The ratios $c^i_tX_t^i/\overline{cX}_t$ and $X^i_T/\overline{X}_T$ measure the relative consumption rate and relative terminal wealth, respectively. In particular, the utility function in is applied to the log-convex combination between absolute and relative consumption rate and terminal wealth, with $\theta_i$ controlling the tradeoff between absolute and relative performance. For $\theta_i$ close to $1$, agent $i$ is more concerned with relative performance than absolute performance, and for $\theta_i=0$ agent $i$ is not at all competitive and ignores the rest of the population. The goal is to find a Nash equilibrium, an investment strategy $(\vec{\pi}^*_t, \vec{c}^*_t)_{t \in [0,T]}$ such that $\pi_t^{i,*}$ and $c_t^{i,*}$ are respectively the optimal stock and consumption allocation exercised by agent $i$ in response to the strategy of all the other agents. With Merton’s problem and the recent findings of [@lacker2017mean] in mind, we might expect to find an equilibrium where the investment strategies $\vec{\pi}^*_t$ are constant and the consumption strategies $c_t^{i,*}$ are only time-dependent. We say that a vector $(\pi^{i,*}, c^{i,*})_{i=1}^n$ of admissible strategies (i.e., $(\pi^{i,^*}, c^{i,*}) \in \A$ for each $i$) is an *equilibrium* if for each $i = 1, ..., n$ and each $(\pi, c) \in \A$ we have $$J_i((\pi^{i,*},c^{i,*})_{i=1}^n) \ge J_i\Big( \ldots, (\pi^{i-1,*},c^{i-1,*}), (\pi, c), (\pi^{i+1,*},c^{i+1,*}), \ldots\Big).$$ An equilibrium $(\pi^{i,*}, c^{i,*})_{i=1}^n$ is called a *strong equilibrium* if, for each $i$, the process $c^{i,*}$ is deterministic and continuous, and the process $\pi^{i,*}$ is deterministic and constant.[^1] The main result is the following, which gives the explicit form of an equilibrium: \[thmN\] Let $n \ge 2$. Assume that for all $i = 1, ..., n$, we have $x_0^i > 0$, $\delta_i > 0$, $\theta_i \in [0,T]$, $\epsilon_i > 0$, $\mu_i > 0$, $\sigma_i \ge 0$, $\nu_i \ge 0$, and $\sigma_i + \nu_i > 0$. Then there is a unique strong equilibrium $(\pi^{i,*}, c^{i,*})_{i=1}^n$, and it takes the following form: $$\begin{aligned} \pi^{i,*} &= \frac{\delta_i \mu_i }{\sigma_i^2 + \nu^2_i(1 + (\delta_i -1)\theta_i/n)} - \frac{\theta_i (\delta_i -1)\sigma_i}{\sigma_i^2 + \nu_i^2(1 + (\delta_i -1)\theta_i/n)} \frac{\phi}{1 + \psi} \label{optcontroleq} \\ c^{i,*}_t &= \begin{cases} \left(\frac{1}{\newomega_i} + \left(\frac{1}{\newC_i} - \frac{1}{\newomega_i} \right)e^{- \newomega_i (T - t)} \right)^{-1} & \mathrm{if } \ \beta_i \neq 0 \\ (T - t + \newC_i^{-1})^{-1} & \mathrm{if } \ \beta_i = 0. \end{cases} \label{optconsumption}\end{aligned}$$ The constants $(\phi,\psi)$ and $(\newomega_i,\newC_i)_{i=1}^n$ are given by $$\begin{aligned} \phi &= \frac{1}{n} \sum_{k = 1}^n \delta_k \frac{\mu_k \sigma_k}{\sigma_k^2 + \nu_k^2(1 + (\delta_k-1)\theta_k/n)}, \nonumber \\ \psi &= \frac{1}{n} \sum_{k = 1}^n \theta_k (\delta_k - 1)\frac{\sigma_k^2}{\sigma_k^2 + \nu_k^2(1 + (\delta_k-1)\theta_k/n)}, \nonumber \\ \newomega_i &= \theta_i (\delta_i - 1)\frac{\frac{1}{n} \sum_{k = 1}^n \delta_k \rho_k}{1 + \frac{1}{n} \sum_{k = 1}^n \theta_k(\delta_k -1) }- \delta_i \rho_i, \nonumber \\ \newC_i &= \epsilon_i^{-\delta_i} \left(\left(\prod_{k = 1}^n \epsilon_k^{\delta_k} \right)^{1/n}\right)^{\theta_i(\delta_i - 1) / (1 + \frac{1}{n} \sum_{k = 1}^n \theta_k(\delta_k-1))}, \label{def:C_i}\end{aligned}$$ where we define also $(\rho_i)_{i=1}^n$ by $$\begin{aligned} \rho_i = (1 - 1/\delta_i) \Bigg\{ & \frac{(1 - \theta_i/n)(\mu_i - \sigma_i\theta_i(1 - 1/\delta_i)\frac{1}{n}\sum_{k \ne i} \sigma_k \pi^{k,*})^2}{2 (\sigma_i^2 + \nu_i^2)(1-(1 -\theta_i/n)(1 - 1/\delta_i) )} \\ &+ \frac{1}{2}\Big(\Big(\frac{1}{n} \sum_{k\ne i} \sigma_k \pi^{k,*} \Big)^2 + \frac{1}{n^2} \sum_{k \ne i} (\nu_k \pi^{k,*})^2\Big) \theta_i^2(1 - 1/\delta_i) \\ &- \theta_i \frac{1}{n} \sum_{k \ne i} \mu_k \pi^{k,*} + \frac{\theta_i}{2n} \sum_{k \ne i} ( \sigma^2_k + \nu_k^2)(\pi^{k,*})^2 \Bigg\},\end{aligned}$$ Moreover, we have the identity $$\begin{aligned} \frac{1}{n} \sum_{k=1}^{n} \sigma_k \pi^{k,*} = \frac{\phi}{1 + \psi}. \label{def:vol-identity}\end{aligned}$$ Note that $\delta_i=1$ implies $\beta_i=0$, which means that log-investors always use the second form of $c^{i,*}_t$ given in . The form of the equilibrium does not seem to simplify much further, except in the single stock case: (Single stock) Assume that for all $i = 1, ..., n$ we have $\mu_i = \mu > 0$, $\sigma_i = \sigma > 0$, and $\nu_i =0$. Then there is a unique strong equilibrium $(\pi^{i,*}, c^{i,*})_{i=1}^n$, and it takes the following form: $$\begin{aligned} \pi^{i,*} &= \frac{\mu}{\sigma^2}\left(\delta_i - \frac{\theta_i}{\theta_{\mathrm{crit}}}(\delta_i-1)\right) \label{optcontroleq-singlestock} \\ c^{i,*}_t &= \begin{cases} \left(\frac{1}{\newomega_i} + \left(\frac{1}{\newC_i} - \frac{1}{\newomega_i} \right)e^{- \newomega_i (T - t)} \right)^{-1} & \mathrm{if } \ \beta_i \ne 0, \\ (T - t + \newC_i^{-1})^{-1} & \mathrm{if } \ \beta_i = 0. \end{cases} \label{optconsumption-singlestock}\end{aligned}$$ For each $i$, the constant $\newC_i$ is given by , and $\newomega_i$ and $\theta_{\mathrm{crit}}$ are given by $$\begin{aligned} \newomega_i &= \frac{\mu^2}{2\sigma^2}(1 - \delta_i)\left(1 - \frac{\theta_i}{\theta_{\mathrm{crit}}}\right)\left(\delta_i - \frac{\theta_i}{\theta_{\mathrm{crit}}}(\delta_i-1)\right), \\ \theta_{\mathrm{crit}} &= \frac{1 + \frac{1}{n}\sum_{k = 1}^{n} \theta_k(\delta_k -1)}{\frac{1}{n}\sum_{k = 1}^{n} \delta_k }.\end{aligned}$$ In Section \[se:mfg\] we simplify this further by sending $n\to\infty$. Then, in Section \[se:discussion\], we analyze in detail how the equilibrium behavior depends on the various parameters. First, we fix an agent $i \in \{1,\ldots,n\}$ and suppose that all other players follow given strategies. That is, for $k \neq i$, let $\pi_k \in \R$ and $c_k : [0,T] \to \R_+$ denote fixed admissible strategies for the other agents, in which the investment policy $\pi_k$ is constant and the consumption policy $c_k$ is a deterministic continuous function. We will solve the optimization problem for agent $i$, determining the agent’s best response to the competitors’ strategies. Then, we will resolved the resulting fixed point problem. Define $Y_t := (\prod_{k \ne i} X_t^k)^{1/n}$, where $X_t^k$ solves (\[processX\]) subject to the strategies $(\pi_k,c_k)$, with $X_0^k = x^k_0$. We use the following abbreviations: $$\begin{array}{ccc} \Sigma_k = \sigma_k^2 + \nu_k^2 & \quad \widehat{\mu \pi}_{-i} = \frac{1}{n} \sum_{k \ne i} \mu_k \pi_k, & \quad \widehat{\sigma \pi}_{-i} = \frac{1}{n} \sum_{k \ne i} \sigma_k \pi_k, \\ \widehat{\Sigma \pi^2}_{-i} = \frac{1}{n} \sum_{k \ne i} \Sigma_k \pi_k^2, & \quad \widehat{(\nu \pi)^2}_{-i} = \frac{1}{n} \sum_{k \ne i} \nu_k^2 \pi_k^2, & \quad \widehat{c}_{-i}(t) = \frac{1}{n} \sum_{k \ne i} c_k(t). \\ \end{array}$$ A straightforward calculation with Itô’s formula (cf. the proof of Theorem 14 in [@lacker2017mean]) shows that the process $Y_t$ satisfies $$\label{processY} \frac{dY_t}{Y_t} = (\eta_i - \widehat{c}_{-i}(t)) dt + \frac{1}{n} \sum_{k \ne i} \nu_k \pi_k dW^k_t + \widehat{\sigma \pi}_{-i} dB_t,$$ where we define also $$\eta_i = \widehat{\mu \pi}_{-i} - \frac{1}{2}\left( \widehat{\Sigma \pi^2}_{-i} - \widehat{\sigma \pi}_{-i}^2 - \frac{1}{n} \widehat{(\nu \pi)^2}_{-i}\right).$$ The $i^{\mathrm{th}}$ agent then solves the optimization problem $$\label{opt} \sup_{(\pi^i, c^i) \in \A} \E\left[\int_0^T U\left((c^i_t X^i_t)^{1 - \theta_i /n} (\bar{c}_{-i}(t)Y_t)^{- \theta_i}; \delta_i \right)dt + \epsilon_i U\left(\left(X^i_T\right)^{1 - \theta_i / n} Y_T^{- \theta_i}; \delta_i \right) \right],$$ where $\bar{c}_{-i}(t) = \left( \prod_{k \ne i} c_k(t) \right)^{1/n}$ and[^2] $$dX^i_t = \pi^i_t X_t^i (\mu_i dt + \nu_i dW^i_t + \sigma_i dB_t) - c^i_t X^i_t dt, \hspace{0.5 cm} X_0^i = x_0^i,$$ with $(Y_t)_{t \in [0,T]}$ solving (\[processY\]). Treating $(X^i,Y)$ as the state process, we solve this stochastic optimal control problem by noting that the value (\[opt\]) should equal $v(X_0^i, Y_0, 0)$, where $v(x,y,t)$ solves the HJB equation $$\begin{split} 0=v_t &+ \sup_{\pi \in \R} \left[ \pi(\mu_i x v_x + \sigma_i \widehat{\sigma \pi}_{-i} xy v_{xy}) + \frac{1}{2} \pi^2\Sigma_i x^2 v_{xx}\right] \\ &+ \sup_{c \in \R_+} \left[- c x v_x + U\left((cx)^{(1 - \theta_i/n)}(\bar{c}_{-i}(t)y)^{- \theta_i}; \delta_i \right) \right] \\ &+ (\eta_i - \widehat{c}_{-i}(t))y v_y + \frac{1}{2}\left(\frac{1}{n} \widehat{(\nu \pi)^2}_{-i} + \widehat{\sigma \pi}_{-i}^2 \right) y^2 v_{yy}, \end{split} \label{pf:HJB}$$ for $(x,y,t) \in \R_+ \times \R_+ \times [0,T)$, with terminal condition $$\begin{aligned} v(x,y,T) = \epsilon_i U(x^{1 - \theta_i/n}y^{- \theta_i}; \delta_i). \label{pf:HJB-boundary}\end{aligned}$$ Notice that the two suprema in are finite if $v_{xx} < 0$ and $v_x > 0$, so we assume for the moment that this is the case, and we will ultimately check that our solution does satisfy these constraints. If follows from a standard verification theorem that there can be at most one classical solution of this PDE. Since the utility takes a different form depending on whether or not $\delta_i = 1$, we treat these two cases separately in the next part of the proof.\ [ ]{} **The case** $\bm{\delta_i \neq 1}$: The utility function takes the form $$U\left((cx)^{\left(1 - \theta_i/n\right)}(\bar{c}_{-i}(t)y)^{-\theta_i}; \delta_i \right) = \left(1 - \frac{1}{\delta_i} \right)^{-1} (cx)^{(1 - \theta_i/n)(1 - 1 / \delta_i)}(\bar{c}_{-i}(t)y)^{- \theta_i(1- 1 / \delta_i)}.$$ Applying the first order conditions, the suprema in are attained by $$\label{optcontrolNot} \pi^{i,*}(x,y,t) = - \frac{\mu_i x v_x(x,y,t) + \sigma_i \widehat{\sigma \pi}_{-i} xy v_{xy}(x,y,t)}{\Sigma_i x^2 v_{xx}(x,y,t)},$$ and $$c^{i,*}(x,y,t) = \frac{1}{x}\left(\frac{(1 - \frac{\theta_i}{n})(\bar{c}_{-i}(t)y)^{- \theta_i(1- 1 / \delta_i)}}{ v_x(x,y,t)} \right)^{\frac{1}{1-(1 - \theta_i/n)(1- 1/\delta_i)}}. \tag{\theequation'}$$ Let us introduce the following constants: $$\begin{aligned} \label{gammai} \gamma_i &= \frac{1}{1-(1 - \theta_i/n)(1- 1/\delta_i)}, \\ \Gamma_i &= \left(1 - \frac{\theta_i}{n}\right)^{ \gamma_i}\left(\frac{1}{ \gamma_i - 1} \right). \nonumber \end{aligned}$$ Using these constants and the expressions and (\[optcontrolNot\]’), the HJB equation becomes $$\label{HJBnot} \begin{split} 0 = v_t &- \frac{1}{2}\frac{(\mu_i x v_x + \sigma_i \widehat{\sigma \pi}_{-i} xy v_{xy})^2}{\Sigma_i x^2 v_{xx}} + \frac{1}{2}\big(\widehat{\sigma \pi}_{-i}^2 + \frac{1}{n}\widehat{(\nu \pi)^2}_{-i}\big)y^2 v_{yy} + (\eta_i - \widehat{c}_{-i}(t)) y v_y \\ &+ (v_x)^{1 - \gamma_i}(\bar{c}_{-i}(t)y)^{- \gamma_i \theta_i(1- 1 / \delta_i)}\Gamma_i. \end{split}$$ We now make the ansatz $$\begin{aligned} v(x,y,t) = \epsilon_i \left( 1 - \frac{1}{\delta_i} \right)^{-1} x^{(1 - \theta_i/n)(1 - 1 / \delta_i)} y^{-\theta_i(1 - 1 / \delta_i)}f_i(t), \label{ansatz-Not}\end{aligned}$$ for a differentiable function $f_i : [0,T] \to \R$ to be determined. Note that the boundary condition requires $f_i(T)=1$. Plugging this into the HJB equation , we find that $v(x,y,t)/f_i(t)$ factors out of each term, and we get $$\label{eqf} 0 = f_i'(t) + \left(\rho_i + \theta_i \left(1 - \frac{1}{\delta_i}\right)\widehat{c}_{-i}(t)\right) f_i(t) + \frac{\epsilon_i^{-\gamma_i}}{\gamma_i} \bar{c}_{-i}(t)^{- \gamma_i \theta_i(1- 1 / \delta_i)} f_i(t)^{1 - \gamma_i},$$ where we define $$\label{rhoi} \begin{split} \rho_i = \left(1 - \frac{1}{\delta_i}\right) & \left( \frac{\gamma_i (1 - \theta_i/n)(\mu_i - \sigma_i \widehat{\sigma \pi}_{-i}\theta_i (1 - 1 / \delta_i))^2}{2\Sigma_i} \right. \\ &\quad + \left. \frac{1}{2}(\widehat{\sigma \pi}_{-i}^2 + \frac{1}{n}\widehat{(\nu \pi)^2}_{-i})\theta_i^2(1 - 1 / \delta_i) - \theta_i \widehat{\mu \pi}_{-i} + \frac{\theta_i}{2}\widehat{\Sigma \pi^2}_{-i} \right). \end{split}$$ Indeed, the last term in comes from the identity $$\begin{split} v_x(x,y,t)^{1 - \gamma_i}y^{- \gamma_i \theta_i(1- 1 / \delta_i)} = \epsilon_i^{-\gamma_i} \left( 1 - \frac{\theta_i}{n} \right)^{1 - \gamma_i} \left( 1 - \frac{1}{\delta_i}\right) f_i(t)^{- \gamma_i} v(x,y,t). \end{split}$$ To solve , let us for the moment abbreviate $$\begin{aligned} a_i(t) &:= \rho_i + \theta_i(1 - 1/\delta_i)\widehat{c}_{-i}(t), \quad\qquad b_i(t) := \frac{\epsilon_i^{-\gamma_i}}{\gamma_i} \bar{c}_{-i}(t)^{- \gamma_i \theta_i(1- 1 / \delta_i)}. \label{def:a-b-for-f}\end{aligned}$$ Then (\[eqf\]) rewrites as $$f_i' + a_i f_i + b_i f_i^{1 - \gamma_i} = 0.$$ This is an example of a *Bernoulli equation*, and a well known change of variables leads to the solution. Indeed, and divide by $f_i^{1-\gamma_i}$ (after noting that $\gamma_i > 0$) and use the substitution $u_i(t) = f_i^{\gamma_i}(t)$, so that (\[eqf\]) becomes the linear differential equation $$\frac{1}{\gamma_i} u_i' + a_i u_i + b_i = 0,$$ with terminal condition $u_i(T) = 1$. This linear equation admits the unique solution $$u_i(t) = e^{\gamma_i \int_t^T a_i(s)ds} + \int_t^T \gamma_i b_i(s) e^{- \gamma_i \int_t^s a_i(r)dr}ds.$$ Note that $b_i$ is positive everywhere, and thus so is $u_i$. Hence, $u_i^{1/\gamma_i}$ is well defined, and the unique solution to (\[eqf\]) is given by $$\label{solBern} f_i(t) = \left(e^{\gamma_i \int_t^T a_i(s)ds} + \int_t^T \gamma_i b_i(s) e^{- \gamma_i \int_t^s a_i(r)dr}ds \right)^{1/\gamma_i}.$$ Substituting this solution into the ansatz yields the solution $v(x,y,t)$ of the HJB equation, as long as we check that $v_{xx} < 0$ and $v_x > 0$. But this is straightforward: $$\begin{aligned} v_x(x,y,t) &= \epsilon_i (1 - \theta_i/n) x^{- 1 / \gamma_i} y^{- \theta_i( 1 - 1/\delta_i)} f_i(t) > 0, \\ v_{xx}(x,y,t) &= - \frac{\epsilon_i}{\gamma_i} (1 - \theta_i/n) x^{- 1 / \gamma_i - 1} y^{- \theta_i( 1 - 1/\delta_i)} f_i(t) < 0,\end{aligned}$$ where we again use $\gamma_i > 0$. Therefore, in terms of $f_i$, we may express the optimal controls from and (\[optcontrolNot\]’) as $$\label{optcontrolNotOne} \begin{split} \pi^{i,*} &= \frac{\gamma_i(\mu_i - \sigma_i \widehat{\sigma \pi}_{-i}\theta_i (1 - 1 / \delta_i))}{\Sigma_i}, \\ c^{i,*}_t &= \epsilon_i^{-\gamma_i} (\bar{c}_{-i}(t))^{- \gamma_i \theta_i(1- 1 / \delta_i)} f_i^{- \gamma_i}(t). \end{split}$$ [  ]{}\ **The case** $\bm{\delta_i=1}$: In the case $\delta_i = 1$ we must proceed differently, but we will ultimately derive optimal controls that are consistent with the formulas in . Note first that we may greatly simplify the form of , because the logarithmic utility function implies in particular that the other players no longer influence player $i$’s optimization. That is, player $i$ maximizes the simplified objective $$\label{valuefOne} (1 - \theta_i /n) \E \left[ \int_0^T \log (c^i_t X^i_t) dt + \epsilon_i \log X^i_T \right].$$ Noting that $1-\theta_i/n > 0$, the value (\[valuefOne\]) is equal to $(1-\theta_i/n)\val2(X_0^i,0)$, where $\val2(x,t)$ solves the HJB equation $$\begin{split} 0 = \val2_t &+ \max_{\pi \in \R} \left[ \pi \mu_i x \val2_x + \pi^2 \frac{1}{2} \Sigma_i x^2 \val2_{xx}\right] + \max_{c \in \R_+} \left[- c x \val2_x + \log(cx) \right], \end{split} \label{HJB-One-simplified}$$ for $(x,t) \in \R_+ \times [0,T)$, with terminal condition $\val2(x,T) = \epsilon_i \log x$. The maximum is attained by $$\label{optconsone} \begin{split} \pi^{i,*}(x,t) &= - \frac{\mu_i x \val2_x(x,t)}{\Sigma_i x^2 \val2_{xx}(x,t)}, \quad\qquad c^{i,*}(x,t) = \frac{1}{x \val2_x(x,t)}. \end{split}$$ The HJB equation then becomes $$\label{HJBone} 0 = \val2_t - \frac{1}{2}\frac{(\mu_i x \val2_x)^2}{\Sigma_i x^2 \val2_{xx}} - 1 - \log \val2_x.$$ Make the ansatz $$\val2(x,t) = f_i(t)\epsilon_i\log x + g_i(t),$$ where $f_i$ and $g_i$ are to be determined and satisfy $f_i(T)=1$ and $g_i(T)=0$. Plug this into (\[HJBone\]), defining the constant $\widehat\rho_i = \mu_i^2\epsilon_i / 2\Sigma_i$, to get $$\left( \epsilon_i f_i'(t) + 1 \right) \log x + g_i'(t) + \widehat\rho_i f_i(t) - 1 - \log \epsilon_i - \log f_i(t) = 0.$$ Since there is only one term depending on $x$, we must have $$\epsilon_i f_i'(t) + 1 = 0, \quad\quad f_i(T)=1.$$ This yields $f_i(t) = \epsilon_i^{-1}(T - t) +1$. Then $g_i$ must solve $$g'_i(t) = - \widehat\rho_i (\epsilon_i^{-1}(T - t) +1) + 1 + \log ((T - t) + \epsilon_i),$$ which is easily integrated using $g_i(T)=0$ to get the solution $g_i$, though we will not need to use it explicitly. Note also that $f_i > 0$, and thus $w_x > 0$ and $w_{xx} < 0$ everywhere. We have therefore solved the HJB when $\delta_i = 1$. Recalling (\[optconsone\]), we deduce that the optimal controls are $$\begin{aligned} \label{optcontrolOne} \pi^{i,*} &= \frac{\mu_i}{\Sigma_i}, \quad\qquad c^{i,*}_t = \frac{1}{T - t + \epsilon_i}.\end{aligned}$$ Note that the result obtained above specializes to when $\delta_i=1$. Indeed, if $\delta_i = 1$, then $\rho_i = 0$ and $\gamma_i = 1$, and the functions $a_i$ and $b_i$ defined in reduce to $a_i \equiv 0$ and $b_i \equiv 1/\epsilon_i$. Thus becomes $f_i(t) = \epsilon_i^{-1}(T- t)+1$, and becomes . [  ]{} **Completing the proof:** We now complete the proof, using the form we found above for the optimal control of player $i$ in response to the other players’ choices. Namely, the best response of player $i$ is given by the controls in , where $f_i$ is defined as in , and we have seen that these formulas are valid for both cases $\delta_i=1$ and $\delta_i \neq 1$. Note that if we assume that the other consumption functions are positive and continuous, then the optimal feedback consumption that we have found is also positive and continuous on $[0,T]$. Now, to conclude the proof, note that the original choice of $(\pi_i, c_i)_{i=1}^n$ is a strong equilibrium if and only if for each $i = 1, ..., n$, we have $$\pi^{i,*} = \pi_i \hspace{0,5cm} \text{ and } \hspace{0,5cm} c^{i,*}_t = c_i(t) \hspace{0,5cm} \forall t \in [0,T],$$ where $(\pi^{i,*},c^{i,*})$ were given in . We first address the investment policy. Note that we obtained exactly the same optimal control $\pi^{i,*}$ as in the problem without consumption, and we can conclude as in the proof of [@lacker2017mean Theorem 14] that $\pi^{i,*} = \pi_i$ for all $i = 1, ..., n$ if and only if $\pi^{i,*}$ is as in . In addition, we prove the identity just as in [@lacker2017mean Theorem 14]. Recall that $\rho_i$ defined in (\[rhoi\]) depends on the investment policies $\pi_k$ of the other agents but not on the consumption policies; in particular, in equilibrium we have $$\begin{split} \rho_i := \Big(1 - \frac{1}{\delta_i}\Big) &\Bigg\{ \frac{(1 - \theta_i/n)(\mu_i - \sigma_i\theta_i(1 - 1/\delta_i)\frac{1}{n}\sum_{k \ne i} \sigma_k \pi^{k,*})^2}{2 (\sigma_i^2 + \nu_i^2)(1 - (1 -\theta_i/n)(1 - 1/\delta_i))} \\ &\quad+ \frac{1}{2}\theta_i^2\Big(1 - \frac{1}{\delta_i}\Big)\Bigg(\Big(\frac{1}{n} \sum_{k\ne i} \sigma_k \pi^{k,*} \Big)^2 + \frac{1}{n} \sum_{k \ne i} (\nu_k \pi^{k,*})^2\Bigg) \\ &\quad - \theta_i \frac{1}{n} \sum_{k \ne i} \mu_k \pi^{k,*} + \frac{\theta_i}{2n} \sum_{k \ne i} ( \sigma^2_k + \nu_k^2)(\pi^{k,*})^2 \Bigg\} \end{split} \label{pf:rhoi-def}$$ Next, we address the consumption policies. In light of our arguments above, in order to have an equilibrium, we must simultaneously solve the following system of equations, for $i=1,\ldots,n$: $$\begin{aligned} c_i(t) &= \epsilon_i^{-\gamma_i} (\bar{c}_{-i}(t))^{- \gamma_i \theta_i(1- 1 / \delta_i)} (f_i(t))^{- \gamma_i} \label{systeq1} \\ 0 &= f_i'(t) + (\rho_i + \theta_i (1 - 1 / \delta_i)\widehat{c}_{-i}(t)) f_i(t) + \frac{\epsilon_i^{-\gamma_i}}{\gamma_i} \bar{c}_{-i}(t)^{- \gamma_i \theta_i(1- 1 / \delta_i)} f_i(t)^{1 - \gamma_i}, \label{systeq2}\end{aligned}$$ with $f_i(T) = 1$. Indeed, the first equation gives the best response of agent $i$ in terms of the other agents’ strategies (computed in ) and the function $f_i$. The second equation is exactly the differential equation which determined $f_i$, which we solved explicitly in terms of the other agents’ strategies in . However, now that we have verified the validity of the ansatz for $v_i(x,y,t)$, to resolve the equilibrium it is more convenient to abandon the explicit form for $f_i$ and instead solve the equations and simultaneously. To do this, first plug into the last term of to find $$f_i'(t) + \left(\rho_i + \theta_i \left(1 - \frac{1}{\delta_i}\right)\widehat{c}_{-i}(t) + \frac{1}{\gamma_i} c_i(t)\right) f_i(t) = 0.$$ Defining the full average $\widehat{c}(t) = \frac{1}{n} \sum_{k = 1}^n c_k(t)$, note that $\widehat{c}_{-i}(t) = \widehat{c}(t) - c_i(t)/n$. Recalling the definition of $\gamma_i$ in , we deduce that $$f_i'(t) + \left(\rho_i + \theta_i \left(1 - \frac{1}{\delta_i}\right)\widehat{c}(t) + \frac{1}{\delta_i} c_i(t)\right) f_i(t) = 0.$$ Hence, with $f_i(T) = 1$ this leads to $$\begin{aligned} f_i(t) = \exp \left( \int_t^T \left(\rho_i + \theta_i \Big(1 - \frac{1}{\delta_i}\Big)\widehat{c}(s) + \frac{1}{\delta_i} c_i(s)\right) ds \right). \label{pf:f-expression}\end{aligned}$$ Now notice that is equivalent to $$c_i(t)^{1- \gamma_i(\theta_i/n)(1 - 1/\delta_i)} = \epsilon_i^{-\gamma_i} \bar{c}(t)^{- \gamma_i \theta_i(1 - 1/\delta_i)} f_i(t)^{-\gamma_i},$$ where $\bar{c}(t)$ denotes the full geometric average, $\bar{c}(t) := \left(\prod_{k = 1}^n c_k(t) \right)^{1/n}$. Hence, recalling the definition of $\gamma_i$ in , $$c_i(t) = \left(\epsilon_i f_i(t)\right)^{- \frac{\gamma_i}{1 - \gamma_i(\theta_i/n)(1 - 1/\delta_i)}}\bar{c}(t)^{- \frac{ \gamma_i \theta_i(1 - 1/\delta_i)}{1 - \gamma_i(\theta_i/n)(1 - 1/\delta_i)}} = \left(\epsilon_i f_i(t)\right)^{-\delta_i}\bar{c}(t)^{- \theta_i (\delta_i - 1)} .$$ Next, plug in the expression for $f_i$ from to get $$c_i(t) = \epsilon_i^{-\delta_i} \bar{c}(t)^{- \theta_i (\delta_i - 1)} \exp \left( - \delta_i \int_t^T \left(\rho_i + \theta_i \Big(1 - \frac{1}{\delta_i}\Big)\widehat{c}(s) + \frac{1}{\delta_i} c_i(s)\right) ds \right),$$ which is equivalent to $$\label{ci0} c_i(t) \exp \left( \int_t^T c_i(s) ds \right) = \epsilon_i^{-\delta_i} \bar{c}(t)^{- \theta_i (\delta_i - 1)} e^{- \delta_i \rho_i (T-t)} \exp \left(- \theta_i (\delta_i - 1) \int_t^T \widehat{c}(s) ds \right).$$ Taking the geometric mean over $i=1,\ldots,n$, we get $$\bar{c}(t) \exp \left( \int_t^T \widehat{c}(s) ds \right) = \left(\overline{\epsilon^{\delta}}\right)^{-1} \bar{c}(t)^{- \widehat{\theta(\delta -1)}} e^{- \widehat{\delta \rho} (T-t)}\exp \left(- \widehat{\theta (\delta - 1)}\int_t^T \widehat{c}(s) ds \right),$$ where we defined $$\begin{array}{lll} \widehat{\theta (\delta - 1)} := \frac{1}{n} \sum_{k = 1}^n \theta_k (\delta_k -1), & \widehat{\delta \rho} := \frac{1}{n} \sum_{k = 1}^n \delta_k \rho_k, \text{ and} & \overline{\epsilon^{\delta}} := \left(\prod_{k = 1}^n \epsilon_k^{\delta_k} \right)^{1/n}. \\ \end{array}$$ Thus, $$\bar{c}(t) \exp \left( \int_t^T \widehat{c}(s) ds \right) = \left(\overline{\epsilon^{\delta}}\right)^{-\frac{1}{1 +\widehat{\theta(\delta-1)}}} e^{- \frac{\widehat{\delta \rho}}{1 + \widehat{\theta(\delta -1)} } (T - t)}.$$ Plugging this expression into , we get $$\label{derivci} c_i(t) \exp \left( \int_t^T c_i(s) ds \right) = \newC_i e^{ \newomega_i (T - t)},$$ where we define $$\begin{aligned} \newomega_i &:= \theta_i (\delta_i - 1)\frac{\widehat{\delta \rho}}{1 + \widehat{\theta(\delta -1)} }- \delta_i \rho_i, \\ \newC_i &:= \epsilon_i^{- \delta_i} \left(\overline{\epsilon^{\delta}}\right)^{\frac{\theta_i(\delta_i -1 )}{1 +\widehat{\theta(\delta-1)}}} > 0.\end{aligned}$$ Integrate (\[derivci\]) from $t$ to $T$ and take the logarithm to get $$\int_t^T c_i(s) ds = \begin{cases} \log \left( 1 + \frac{\newC_i}{ \newomega_i} \left( e^{ \newomega_i (T - t)} - 1 \right) \right) &\text{if } \beta_i\neq 0 \\ \log\left(\newC_i(T-t) + 1\right) &\text{if } \beta_i = 0. \end{cases} \label{intci}$$ This is indeed well defined because, when $\beta_i \neq 0$, the function $t \mapsto 1 + \frac{\newC_i}{ \newomega_i} \left( e^{ \newomega_i (T - t)} - 1 \right)$ is decreasing on $[0,T]$ and equal to $1$ at $t=T$. Differentiating both sides, we finally obtain $$c_i(t) = \begin{cases} \left(\frac{1}{\newomega_i} + \left(\frac{1}{\newC_i} - \frac{1}{\newomega_i} \right)e^{- \newomega_i (T - t)} \right)^{-1} &\text{if } \beta_i\neq 0 \\ (T - t + \newC_i^{-1})^{-1} &\text{if } \beta_i = 0. \end{cases}$$ In summary, we have found the unique solution of the system of equations given in and , justifying our ansatz for the HJB equation . With a classical solution of the HJB equation in hand, by a standard verification argument [@fleming2006controlled; @pham2009continuous] we conclude that the portfolio and consumption policies identified above do indeed provide the unique best responses and thus the unique strong equilibrium. The mean field game {#se:mfg} =================== We study in this section the limit as $n \rightarrow \infty$ of the $n$-player game analyzed previously, and we explain how the limit can be viewed as the equilibrium outcome of a (mean field) game with a continuum of agents. For each agent $i$, define the *type vector* $$\zeta_i := (x_0^i, \delta_i, \theta_i, \epsilon_i, \mu_i, \nu_i, \sigma_i).$$ We now allow these parameters to depend also on $n$, though we will not burden the notation with an additional index. These type vectors induce an empirical measure, the [*type distribution*]{}, which is the probability measure on the [*type space*]{} $$\ZZ := (0, \infty) \times (0, \infty) \times [0,1] \times (0, \infty) \times (0, \infty) \times [0, \infty) \times [0, \infty),$$ given by $m_n = \frac{1}{n}\sum_{k=1}^n\delta_{\zeta_k}$. Now note that for each agent $i$, the equilibrium strategy for the consumption as well as the investment only depends on the agent’s own type vector and on the distribution $m_n$ of the type vectors. Hence, if we assume $m_n$ converges weakly to some limiting probability measure, then we expect the equilibrium outcome to converge in a certain sense. In order to pass to the limit, let us now denote by $(x_0,\delta,\theta,\epsilon,\mu,\nu,\sigma)$ a $\ZZ$-valued random variable, with $\nu+\sigma > 0$ a.s. This law of this random type vector represents the distribution of type vectors of a continuum of agents, and a single realization of this random type vector is to be interpreted as the type assigned to a single representative agent. We assume that all expectations appearing in this paragraph are finite. The $n\to\infty$ limiting forms of the constants $\phi$ and $\psi$ defined in Theorem \[thmN\] are as follows: $$\begin{aligned} \phi &= \E\left[\frac{\delta \mu \sigma}{\sigma^2 + \nu^2} \right], \qquad \psi = \E\left[\frac{\theta (\delta -1) \sigma^2}{\sigma^2 + \nu^2} \right]. \label{MF-phi}\end{aligned}$$ To identify limiting forms of the remaining quantities in Theorem \[thmN\], we additionally remove the $i$ subscript, letting the randomness of the type vector play the role of the names of the agents. This gives $$\begin{aligned} \newomega &= \theta(\delta -1) \frac{\E \left[\delta \rho \right]}{1 + \E \left[ \theta(\delta -1)\right]} - \delta \rho, \label{MF-omega}\end{aligned}$$ and $$\begin{split} \rho = \left(1 - \frac{1}{\delta}\right) &\left\{ \frac{\delta}{2(\sigma^2+\nu^2)} \left(\mu - \sigma \frac{\phi}{1 + \psi}\theta (1 - 1 / \delta)\right)^2 + \frac{1}{2}\left(\frac{\phi}{1 + \psi} \right)^2 \theta^2(1 - 1 / \delta)\right.\\ &- \theta\frac{\phi}{1 + \psi} \E\left[\frac{\delta \mu^2 - \theta (\delta -1)\sigma \mu}{\sigma^2 + \nu^2} \right] + \left. \frac{\theta}{2}\E \left[\frac{(\delta \mu - \theta (\delta -1)\sigma \frac{\phi}{1+\psi})^2}{\sigma^2 + \nu^2} \right] \right\}. \label{MF-rho} \end{split}$$ The limiting form of $\newC_i$ is given by $$\newC = \epsilon^{-\delta} \left( e^{\E\left[\log(\epsilon^{-\delta})\right]}\right)^{-\frac{\theta(\delta-1)}{1 + \E[\theta(\delta-1)]}}. \label{MF-C}$$ Indeed, this is determined by noting that $$\left(\prod_{k=1}^n\epsilon_k^{\delta_k}\right)^{1/n} = \exp\left(\frac{1}{n}\sum_{k=1}^n \log(\epsilon_k^{\delta_k})\right).$$ The equilibrium investment policy of Theorem \[thmN\] of the representative agent then becomes $$\begin{aligned} \pi^* = \frac{\delta \mu}{\sigma^2 + \nu^2} - \frac{\theta(\delta -1) \sigma}{\sigma^2 + \nu^2}\frac{\phi}{1 + \psi} \label{MF-pi*}\end{aligned}$$ and the consumption policy becomes $$\begin{aligned} \label{MF-c*} c^*_t = \begin{cases} \left(\frac{1}{\newomega} + \left(\frac{1}{\newC} - \frac{1}{\newomega} \right)e^{- \newomega (T - t)} \right)^{-1} & \text{ if } \beta \neq 0 \\ (T - t + \newC^{-1})^{-1} & \text{ if } \beta=0. \end{cases}\end{aligned}$$ We next illustrate how this strategy arises as the equilibrium of a mean field game. Let $(\Omega,\F,\FF=(\F_t)_{t \in [0,T]},\PP)$ be a filtered probability space supporting independent Brownian motions $B$ and $W$ as well as a random type vector $\zeta= (\xi, \delta, \theta, \epsilon, \mu, \nu, \sigma)$ as above. Assume that $\FF$ is the minimal complete filtration with respect to which $\zeta$ is $\F_0$-measurable and $W$ and $B$ are $\FF$-Brownian motions. The representative agent’s wealth process is determined by $$\label{wealth} dX_t = \pi_t X_t (\mu dt + \nu dW_t + \sigma dB_t) - c_t X_t dt.$$ As before, admissible strategies are given by $\FF$-progressively measurable $\R \times \R_+$-valued processes $(\pi,c)$ satisfying $\E\int_0^T(\pi_t^2 + c_t^2)dt < \infty$, and every admissible strategy results in a strictly positive wealth process. Because this is a mean field game with common noise $B$, the mean field equilibrium condition will involve *conditional* means given $B$. Intuitively, because the interaction between the agents occurs through the (geometric) average over the whole population, we expect some kind of a law of large numbers and asymptotic independence between the agents as $n\to\infty$. Due to the presence of common noise, any asymptotic independence between the agents must be conditional on the common noise $B$, and we refer to [@carmona-delarue-book; @carmona-delarue-lacker] for more thorough and precise treatments of mean field games with common noise. In other words, the population average wealth and consumption processes should be adapted to the complete filtration $\FF^B=(\F^B_t)_{t \in [0,T]}$ generated by the common noise $B$. Now, suppose that the representative agent knows that the geometric mean wealth and consumption of the (continuum of) other agents are governed by some $\FF^B$-adapted processes $\overline{X}$ and $\overline{\Gamma}$, respectively. Then, the objective of the representative agent is to maximize the expected payoff $$\label{optpb} \sup_{(\pi, c) \in \mathcal{A}_{MF}} \E \left[\int_0^T U\left(c_t X_t (\overline{\Gamma}_t \overline{X}_t)^{-\theta}; \delta \right)dt + \epsilon U \left(X_T \overline{X}^{-\theta}_T; \delta \right)\right].$$ In equilibrium, the optimal $(\pi^*,c^*)$ for this problem should lead to $\exp\E[\log X_t \, | \, \F^B_t] = \overline{X}_t$ and $\exp\E[\log c^*_t \, | \, \F^B_t] = \overline{\Gamma}_t$, where we note that $\exp\E[\log(\cdot)]$ is the continuous analogue of geometric mean. We formalize this discussion in the following definition: Let $(\pi^*, c^*)$ be admissible strategies, and consider the $\FF^{B}$-adapted processes $\overline{X}_t := \exp \E[\log X^*_t | \mathcal{F}_{t}^B]$ and $\overline{\Gamma}_t = \exp \E[ \log c^*_t | \mathcal{F}_t^B]$, where $(X_t^*)_{t \in [0,T]}$ is the wealth process in (\[wealth\]) corresponding to the strategy $(\pi^*, c^*)$. We say that $(\pi^*, c^*)$ is a *mean field equilibrium* if $(\pi^*, c^*)$ is optimal for the optimization problem (\[optpb\]) corresponding to this choice of $\overline{X}$ and $\overline{\Gamma}$. We call $(\pi^*, c^*)$ a *strong equilibrium* if $\pi^*$ is constant (and thus $\F_0$-measurable) and if $(c^*_t)_{t \in [0,T]}$ is continuous and $\F_0$-measurable. Because $\F_0$ is the $\sigma$-field generated by the type vector, to say that a strategy is $\F_0$-measurable simply means that it depends on the type vector only, not on the Brownian motions or wealth process. We may now state a theorem which explains the precise sense in which the $n\to\infty$ limiting strategies computed above can be viewed as the equilibrium outcome of a mean field game. \[thmMF\] Assume that a.s. $\delta > 0$, $\theta \in [0,1]$, $\epsilon > 0$, $\mu > 0$, $\sigma \ge 0$, $\nu \ge 0$, and $\sigma + \nu > 0$. Define $(\phi,\psi,\newomega,\rho,\newC)$ as in –, and assume all of the expectations therein are finite. Then there is a unique strong equilibrium $(\pi^*,c^*)$, and it takes the form given by and . (Single Stock) \[co:MFG-singlestock\] Assume that $(\mu, \nu, \sigma)$ is deterministic with $\nu = 0$ and $\mu, \sigma > 0$. Then $\newomega$ defined in can be simplified to $$\newomega = \frac{\mu^2}{2\sigma^2}(1 - \delta)\left(1 - \frac{\theta}{\theta_{\mathrm{crit}}} \right) \left( \delta - \frac{\theta}{\theta_{\mathrm{crit}}}(\delta -1) \right),$$ where $$\theta_{\mathrm{crit}} := \frac{1 + \E[\theta(\delta-1)]}{\E[\delta]},$$ and the optimal investment simplifies to $$\pi^* = \left( \delta - \frac{\theta}{\theta_{\mathrm{crit}}}(\delta - 1) \right)\frac{\mu}{\sigma^2}.$$ We omit the proof, because it closely parallels the proofs of Theorem \[thmN\] and [@lacker2017mean Theorem 3.6]. The main idea, as in the proof of Theorem \[thmN\], is to identify the dynamics of the process $\overline{X}_t=\exp\E[\log X_t \, | \, \F^B_t]$, when $X$ is subject to $\F_0$-measurable strategies $(\pi,c)$, with $\pi$ time-independent. The representative agent’s optimization problem can then be cast as a (tractable) stochastic control problem over the two-dimensional state process $(X,\overline{X})$. Discussion of the equilibrium {#se:discussion} ============================= We now discuss the interpretation of the equilibria computed in the previous sections and the nature of the dependence on the various model parameters. First, notice that our result is consistent with Samuelson’s [@samuelson1975lifetime], in the sense that the investment strategy we obtain is the same as in the model without consumption, derived in the previous work [@lacker2017mean]. More generally, the investment strategy $\pi^*$ does not depend on the relative importance that the agents give to terminal wealth versus consumption, quantified by $\epsilon$ in our model. With this in mind, we refer to [@lacker2017mean] for the discussion on the investment strategy, and we focus the rest of the discussion here on the consumption strategy. We further limit the discussion of the equilibrium to the mean field case, for which the equilibrium consumption policy is given by , as the equilibrium in the $n$-agent game has essentially the same structure but more complicated formulas. Moreover, we restrict our attention mostly to the single stock case of Corollary \[co:MFG-singlestock\], which is again more tractable but already quite rich. From the expression for $c^*_t$, we can distinguish three regimes of consumption behavior. The optimal consumption is necessarily a monotone function of time $t$, and a quick computation shows that it is increasing when $\newomega < \newC$, decreasing when $\newomega > \newC$ and constant when $\newomega = \newC$. Recalling the form of the wealth process $X$ in , we see that the expected rate of return of wealth, $\frac{d}{dt}\E[\log X_t \, | \, \F_0]$, is also a monotone function of time, with the opposite monotonicity of the consumption policy. (Note that conditioning on $\F_0$ is equivalent to conditioning on the representative agent’s type.) See Figure \[fig:consumption-policies\] for some typical consumption policies. Recall that $\epsilon$ captures the relative importance that an agent gives to terminal wealth compared to consumption. Note that $\newC \to \infty$ as $\epsilon \to 0$, and in particular we have $\newomega < \newC$ for small $\epsilon$. As discussed above, this means the agent aims for a decreasing rate of growth of wealth and an increasing rate of consumption. This is natural, because a small $\epsilon$ indicates the agent’s lack of interest in terminal wealth, which drives $X(t)$ toward zero as $t \rightarrow T$ (as there is no bequest in our model). That is, for $\epsilon$ sufficiently small, the agent dis-invests after some time in order to consume more. In fact, the agent may even begin dis-investing immediately if $\pi^* < c^*(0)$. On the contrary, if $\epsilon$ is large, the agent is more concerned with terminal wealth than consumption and will thus decrease her consumption over time. Indeed, $\newC$ is decreasing in $\epsilon$, so for large $\epsilon$ we have $\newomega > \newC$. We now turn to the key question of the impact of an agent’s competitiveness and risk tolerance on her consumption behavior. To simplify the discussion, let us assume henceforth that no agent has a preference between her utility of wealth or utility of consumption; that is, $\epsilon = 1$ and $\E[\log(\epsilon^{-\delta})] = 0$, which in particular implies $\newC = 1$. Note that if $\theta=0$, then we recover the classical Merton solution without competition, with $\newomega= \frac{\mu^2}{2 \sigma^2}\delta(1 - \delta)$ and $\pi^* = \delta\mu/\sigma^2$. For general $\theta$, we may still rewrite $\newomega$ and $\pi^*$ in an analogous manner as $$\newomega = \frac{\mu^2}{2 \sigma^2}\delta_{\mathrm{eff}}(1 - \delta_{\mathrm{eff}}), \qquad\quad \pi^* = \delta_{\mathrm{eff}}\mu/\sigma^2,$$ where we define the *effective risk tolerance parameter* $$\begin{aligned} \delta_{\mathrm{eff}} := \left(1 - \frac{\theta}{\theta_{\mathrm{crit}}} \right) \delta + \frac{\theta}{\theta_{\mathrm{crit}}}.\end{aligned}$$ In other words, in the face of competition, the agent behaves like a Merton investor/consumer but with a different risk tolerance parameter. We can interpret $\delta_{\mathrm{eff}}$ as a weighted average of the agent’s own risk tolerance $\delta$ and the critical log-investor case $\delta_{\mathrm{log}} = 1$, with the weight determined by the agent’s competitiveness. Take note, however, that the range of the weight $\theta/\theta_{\mathrm{crit}}$ is $[0,\infty)$ and $\delta_{\mathrm{eff}}$ can be negative, so we should avoid interpreting $\delta_{\mathrm{eff}}$ too literally as a risk tolerance parameter. Let us investigate in more detail what distinguishes between agents who decrease versus increase their consumption over time, and let us continue to assume that $\epsilon \equiv 1$ (and thus $\newC \equiv 1$) for all agents. As discussed above, an agent increases her consumption over time if and only if $\newomega < \newC=1$. If $8\sigma^2 > \mu^2$, then we always have $\newomega < 1$, because $\delta_{\mathrm{eff}}(1-\delta_{\mathrm{eff}}) \le 1/4$ for any $\delta_{\mathrm{eff}}$. So assume instead that $8\sigma^2 < \mu^2$. Then, because $\newomega$ is a quadratic function of $\delta$ (if all other parameters are held fixed), we may find $\delta^*_{\pm}$ such that $$\newomega > 1 \ \ \Longleftrightarrow \ \ \delta \in (\delta_-^*, \delta_+^*).$$ That is, the agent decreases consumption over time if $\delta_-^* < \delta < \delta_+^*$, increases consumption over time if $\delta \notin (\delta_-^*, \delta_+^*)$, and consumes at a constant rate if $\delta \in \{\delta_-^*, \delta_+^*\}$. Precisely, these two values are $$\delta_{\pm}^* := 1 + \frac{1}{2} \left( \frac{1}{\theta / \overline{\theta}_{\mathrm{crit}}- 1} \pm \frac{\sqrt{1 - 8 \sigma^2 / \mu^2}}{\left| \theta/ \overline{\theta}_{\mathrm{crit}} -1\right|} \right).$$ (If $\theta=\theta_{\mathrm{crit}}$, then $\newomega=0$, so let us assume $\theta \neq \theta_{\mathrm{crit}}$.) This explains the non-monotonicity in $\delta$ of the equilibrium consumption strategy, as well as the wave-like shape of the curve of $c^*_t$ versus $\delta$ and $\theta$ pictured in Figure \[fig:opt\_c\_vs\_delta\_theta\]. In the classical Merton problem with no competition, recovered by setting $\theta=0$, the endpoints become $\delta_{\pm}^* = \tfrac12(1 \pm \sqrt{1 - 8 \sigma^2 / \mu^2})$, both of which are less than $1$; in this case only risk averse ($\delta < 1$) agents decrease their consumption over time. In contrast, in the competitive case $\theta > 0$, the interval $(\delta_-, \delta_+)$ may lie above or below $1$ depending on the sign of $1 - \theta /\overline{\theta}_{\mathrm{crit}}$. On the one hand, suppose the agent is less competitive than the critical value, or $\theta < \overline{\theta}_{\mathrm{crit}}$. Then $\delta_-^* < \delta_+^* < 1$, and we have seen that the agent will decrease her rate of consumption over time only if her risk tolerance lies within the range $(\delta_-^*, \delta_+^*)$. As we would expect, a relatively uncompetitive agent behaves similarly to a Merton investor in this respect. On the other hand, if the agent is more competitive than the critical value, or $\theta > \overline{\theta}_{\mathrm{crit}}$, then $1 < \delta_-^* < \delta_+^*$. This means that even highly risk tolerant agents may decrease consumption over time. Noting that $\delta_{\mathrm{eff}}$ decreases with $\theta$, one interpretation is as follows: Increasing $\theta$ exposes an agent to relative performance pressures, which is itself a source of risk, and to offset this additional risk the agent behaves like a Merton investor with a smaller risk tolerance parameter. Note that a highly competitive agent, with $\theta > \theta_{\mathrm{crit}}$, behaves in a sense *opposite* to how they would if $\theta=0$. Indeed, when $\theta > \overline{\theta}_{\mathrm{crit}}$, the effective risk tolerance $\delta_{\mathrm{eff}}$ is less than $1$ if $\delta > 1$ and greater than $1$ if $\delta < 1$. The agent effectively switches to the other side of the critical risk tolerance $\delta_{\mathrm{log}}=1$. We have seen by now how, with other parameters held fixed, the consumption policy may depend non-monotonically on the risk tolerance $\delta$, with an intermediate range of risk tolerance parameters $(\delta_-^*,\delta_+^*)$ in which agents decrease consumption over time. Similarly, with other parameters held fixed, consumption can exhibit the same non-monotonicities as a function of $\theta$, with an intermediate range $(\theta_-^*,\theta_+^*)$ in which agents decrease consumption over time. See Figure \[fig:region\] for a depiction of the range of $(\delta,\theta)$ parameters leading agents to decrease versus increase consumption over time. [^1]: This definition of equilibrium is more specifically of *open-loop* type, but a strong equilibrium, being nonrandom, can be shown to also provide an equilibrium over closed-loop or Markovian controls. [^2]: We use a bar $\overline{c}$ to denote a geometric average and a hat $\widehat{c}$ to denote an arithmetic average.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Dans cet article on construit l’isomorphisme entre les tours de Lubin-Tate et de Drinfeld au niveau des points de ces espaces. Les points considérés sont ceux intervenant dans la théorie des espaces analytiques de Berkovich.' address: 'CNRS-IHES-université Paris-Sud Orsay' author: - Laurent Fargues bibliography: - 'biblio.bib' title: 'L’isomorphisme entre les tours de Lubin-Tate et de Drinfeld au niveau des points' --- In this article we construct the isomorphism between Lubin-Tate and Drinfeld towers at the level of points. The points we consider are the one of the theory of analytic spaces in the sens of Berkovich. Introduction {#introduction .unnumbered} ============ Le but de cet article est de démontrer l’existence de l’isomorphisme entre les tours de Lubin-Tate et de Drinfeld au niveau des points c’est à dire de décrire la bijection correspondante. Nous nous inspirons bien sûr de [@Faltings8] (cependant l’auteur ne garantit pas que la bijection décrite coïncide avec celle de [@Faltings8] qu’il n’a pu complètement comprendre). Cet isomorphisme n’est pas algébrique : un $\Qpb$-point peut s’envoyer sur un $\Cp$-point ne provenant pas d’un $\Qpb$-point. C’est pourquoi nous devons travailler avec des points à valeurs dans de “gros corps” du type $\Cp$. Par points nous entendons donc ceux intervenant par exemple dans la théorie des espaces analytiques de Berkovich : ce sont ceux à valeurs dans un corps valué complet pour une valuation de rang $1$ extension de $\Qp$. La simplification par rapport à la construction de l’isomorphisme dans le cas général provient de ce que l’on n’a pas à introduire d’éclatements admissibles ou de modèles entiers particuliers des espaces de Lubin-Tate et de Drinfeld : si $K|\Qp$ est valué complet pour une valuation de rang $1$ les points à valeurs dans $K$ de la tour de Lubin-Tate et de Drinfeld sont définis indépendamment du modèle entier. En particulier cet article est indépendant de la construction du schéma formel de [@Cellulaire]. L’utilisation de corps valués pour des valuations non-discrètes à corps résiduel non forcément parfait introduit des problèmes de théorie de Hodge $p$-adique non disponibles dans la littérature. Les résultats associés peuvent se déduire de l’approche utilisée dans [@Faltings1]. Nous ne les démontrons pas dans cet article mais nous y consacrons un appendice dans lequel nous y exposons les résultats auxquels on peut parvenir en utilisant les méthodes de [@Faltings1]. L’auteur y consacrera un article plus général ([@Periodes]). L’un des autres problèmes auquel on est confronté est le fait que l’on travaille avec des $\O$-modules formels, le cas usuel correspondant à $\O=\Zp$. Pour y remédier on applique la théorie des $\O$-extensions vectorielles universelles développée dans l’appendice B de [@Cellulaire]. Très peu est nécessaire : seules les sections un à quatre de ce cet appendice sont utilisées, nous n’utiliserons pas le fait que le cristal algèbre de Lie de la $\O$-extension vectorielle universelle s’étend aux $\O$-puissances $\pi$-divisées. qui est la partie délicate de l’appendice B de [@Cellulaire]. L’utilisation de cette théorie relative à $\O$ rend la rédaction des problèmes reliés purement formels. Seul le cas de l’utilisation des théorèmes de comparaison cristallins n’est pas rédigé dans ce cadre relatif : l’auteur n’a pas eu le courage de développer la théorie des anneaux de Fontaine obtenus en remplaçant les vecteurs de Witt par les vecteurs de Witt ramifiés... Dans cet article nous donnons deux descriptions différentes de l’isomorphisme. La première dans le chapitre 9 est la plus simple. La seconde dans le chapitre 12 est plus alambiquée mais s’adapte mieux au cas relatif (on renvoie à l’introduction du chapitre \[ptovue\] pour plus de détails).\ Avant de décrire succinctement la bijection rappelons quelques faits sur les espaces de Lubin-Tate et de Drinfeld associés à $F|\Qp$. Plaçons nous au niveau des points de ces espaces. Soit $D$ l’algèbre à division d’invariant $\frac{1}{n}$ sur $F$. La tour de Lubin-Tate $(\LT_K)_{K\subset GL_n (\O_F)}$ est une tour indexée par les sous-groupes ouverts $K$ de $\GL_n (\O_F)$ munie d’une action “horizontale” de $D^\times$ (sur chaque élément de la tour) et verticale (par correspondances de Hecke) de $\GL_n (F)$. Soit $\LT_\infty =\underset{K}{\limp} \LT_K$ (cela a bien un sens au niveau des points) qui est muni d’une action de $\GL_n(F)\times D^\times$. Elle forme un pro-revêtement de groupe $\GL_n (\O_F)$ $$\xymatrix@C=5mm{ \ar@/_1.2pc/@{-}[d]_{K} \LT_\infty \ar[d] \ar@/^2pc/@{-}[dd]^{GL_n (\O_F)} \\ \LT_K \ar[d] \ar@(l,ld)_{D^\times} \\ \LT_{GL_n (\O_F)} \ar@{=}[r] & \coprod_\Z {\mathring{\mathbb{B}}^{n-1}} }$$ où $\LT_{GL_n (\O_F)}$ est l’espace de Lubin-Tate sans niveau (une union disjointes de boules $p$-adiques ouvertes de dimension $n-1$ car on travaille en fait avec des espaces de Rapoport-Zink et non les espaces de Lubin-Tate classiques) qui classifie des déformations de groupes formels de dimension $1$ et $\LT_K$ est obtenu en trivialisant partiellement le module de Tate de la déformation universelle au dessus de $\LT_{GL_n (\O_F)}$ (on force la monodromie sur le module de Tate à vivre dans $K$). On a donc $\LT_\infty/\GL_n (\O_F) = \coprod_\Z\mathring{\mathbb{B}}^{n-1}$. Il y a de plus une application des périodes de Hodge-De-Rham $\LT_K \ldrt \mathbb{P}^{n-1}$ donnée par la filtration de Hodge du module filtré définissant la déformation et dont les fibres sont les orbites de Hecke. Et donc $\LT_\infty/\GL_n (F)= \mathbb{P}^{n-1}$. $$\xymatrix@R=8mm{ \LT_\infty \ar[d] \ar@/^2pc/@{-}[d]^{GL_n (\O_F)} \ar@/_2pc/@{-}[dd]_{GL_n (F)} \\ {\coprod_\Z \mathring{\mathbb{B}}^{n-1}} \ar[d]^{\text{Périodes}} \\ \mathbb{P}^{n-1} \ar@(l,ld)_{D^\times} }$$ Il y a une description du même type pour la tour de Drinfeld. L’espace de Drinfeld en niveau infini $\D_\infty$ est muni d’une action de $\GL_n (F)\times D^\times$, l’action de $\GL_n (F)$ est “horizontale” sur chaque élément de la tour. Il y a un diagramme pour $K\subset \O_D^\times$ un sous-groupe ouvert $$\xymatrix{ \ar@/_1.2pc/@{-}[d]|{K} \ar@/_5pc/@{-}[ddd]_{D^\times} \D_\infty \ar[d] \ar@/^2pc/@{-}[dd]^{\O_D^\times} \\ \D_K \ar[d] \ar@(l,ld)_{GL_n (F)} \\ \D_{GL_n (\O_F)} \ar@{=}[r]\ar[d]^{\text{Périodes}} & \coprod_\Z \Omega \\ \Omega\ar@(l,ld)_{GL_n (F)} }$$ L’application des périodes de Hodge-De-Rham n’est rien d’autre que la projection de $\coprod_\Z \Omega$ sur $\Omega$, c’est à dire le quotient par $D^\times/\O_D^\times = \Pi^\Z$.\ Les deux applications inverses l’une de l’autre entre $\LT_\infty$ et $\D_\infty$ sont construites en deux étapes. $$\xymatrix@R=20mm{ \LT_\infty \ar[d]_{\text{Périodes de} \atop{\text{Hodge-De-Rham}}}\ar@<.8ex>[rr] \ar[rrd] && \D_\infty \ar[d]^{\text{Périodes de} \atop{\text{Hodge-De-Rham}}} \ar[lld] \ar@<.8ex>[ll] \\ \LT_\infty/\GL_n (F) = \mathbb{P}^{n-1} & & \D_\infty/D^\times =\Omega }$$ On commence par construire une application de périodes de Hodge-Tate d’un des espaces en niveau infini vers l’espace des périodes de Hodge-De-Rham de l’autre espace (les flèches diagonales dans la figure précédente). Les périodes de Hodge-Tate sont données par la rigidification du module de Tate (le fait qu’on ai fixé une base de ce module au sommet de la tour) et la suite de Hodge-Tate, contrairement aux périodes de Hodge-De-Rham données elles par la rigidification de la cohomologie cristalline (l’isocristal du groupe $p$-divisible) et la filtration de Hodge. Puis on relève la flèche en niveau infini en utilisant la théorie de Messing qui permet de construire des éléments dans le module de Tate en les construisant modulo $p$. Nous n’expliquerons pas d’avantage ce dernier point dans l’introduction, mais afin d’en donner un avant-goût citons le point clef suivant : soit $K|\Qp$ une extension valuée complète et $H$ un groupe $p$-divisible sur $\O_K$ . Notons $\bar{\eta}=\spec (\overline{K})$. Alors le module de Tate de $H$ admet deux définitions : l’une en fibre générique, la définition usuelle $$T_p (H) =\underset{n}{\limp} H[p^n] (\overline{K}) = \Hom (\Qp/\Zp, H_{\bar{\eta}})$$ l’autre modulo $p$ $$T_p (H) = \{\; f \in \Hom (\Qp/\Zp, H\otimes \O_{\overline{K}}/p \O_{\overline{K}})\;|\; f_* (1) \in \Fil\, \Lie (E(H))\otimes \O_{\widehat{\overline{K}}}\; \}$$ où $\Lie\, E(H)$ désigne l’algèbre de Lie de l’extension vectorielle universelle de $H$ muni de sa filtration de Hodge et $f_* : \O_{\widehat{\overline{K}}} \ldrt \Lie \, E(H)\otimes \O_{\widehat{\overline{K}}}$ est le morphisme induit sur les cristaux de Messing évalués sur l’épaississement à puissances divisées $\O_{\widehat{\overline{K}}} \twoheadrightarrow \O_{\overline{K}}/p\O_{\overline{K}}$. La démonstration est un jeu basé sur ces deux aspects : fibre générique et modulo $p$. Les deux groupes $p$-divisibles sur $\overline{\mathbb{F}}_p$ que l’on déforme pour définir les espaces de Lubin-Tate et de Drinfeld sont reliés modulo $p$, l’un est isogéne à un produit d’un nombre fini de copies de l’autre, et l’on peut ainsi “transférer” des éléments du module de Tate sur une des tours vers l’autre.\ Une autre interprétation de cet isomorphisme est donnée dans la section \[kfyygpzr5rrppn\]. On peut associer à un point de l’une des deux tours en niveau infini une matrice de périodes dans $M_n (B^+_{cris})$. En effet, pour la tour de Lubin-Tate le théorème de comparaison entre cohomologie cristalline et cohomologie étale $p$-adique pour les groupes $p$-divisibles fournit une matrice de périodes $X_{cris}$ à coefficients dans $B^+_{cris}$ une fois que l’on a fixé une base du module de Tate (la rigidification en niveau infini) et une base du module de Dieudonné (la rigidification définissant la déformation) $$(B^+_{cris})^n\iso V_p\otimes B^+_{cris} \iso \mathbb{D}^{\LT}\otimes {B^+_{cris}} \xleftarrow{\; \sim\;} (B^+_{cris})^n$$ Pour la tour de Drinfeld il y a un isomorphisme de $D\otimes_{\Qp} B^+_{cris}$-modules en niveau infini $$D\otimes_{\Qp} B^+_{cris} \iso V_p \otimes B^+_{cris} \iso \mathbb{D}^\D\otimes B^+_{cris} \xleftarrow{\; \sim\;} D\otimes_{\Qp} {B^+_{cris}}$$ utilisant la $D$-équivariance de l’isomorphisme des périodes cristalline on en déduit une matrice $Y_{cris}\in M_n (B^+_{cris})$ comme précédemment. L’isomorphisme entre les deux tours consiste alors à prendre la transposé de ces matrices de périodes $$Y_{cris} = \,^t X_{cris}$$ Si $X_{cris}$ désigne la matrice de périodes sur $\LT_\infty$ sa réduction modulo l’idéal d’augmentation $B^+_{cris}\ldrt \O_{\widehat{\overline{K}}}$, $X$, est telle que les colonnes de $X$ engendrent les périodes de Hodge-De-Rham et ses lignes les périodes de Hodge-Tate ! $$\xymatrix@C=24mm@R=12mm{ \LT_\infty \ar[dd]_{\text{Périodes}\atop\text{de Hodge-De-Rham}} \ar[r]^{\text{Périodes}\atop \text{cristallines}} \ar@/_2pc/@{..>}[rrdd]|(.25){\text{Périodes}\atop\text{de Hodge-Tate}} & M_n (B^+_{cris}) \ar[d] & \ar[l]_{\text{Transposé des périodes}\atop \text{cristallines}} \D_\infty \ar[dd]^{\text{Périodes}\atop\text{de Hodge-De-Rham}} \ar@/^2pc/@{..>}[lldd]|(.25){\text{Périodes}\atop\text{de Hodge-Tate}} \\ & M_n ({\widehat{\overline{K}}}) \ar[ld]_{\text{Im}\,^t X} \ar[rd]^{\text{Im X}} \\ \mathbb{P}^{n-1} && \Omega }$$ (les deux flêches $M_n (\widehat{\overline{K}})\ldrt \mathbb{P}^{n-1}$ et $M_n (\widehat{\overline{K}})\ldrt \Omega$ ne sont définies que sur le sous-ensemble des matrices de rang $n-1$). Dans le diagramme précédent l’image des deux flèches “Périodes cristallines” sont les matrices $X_{cris}\in M_n (B^+_{cris})$ vérifiant une équation fonctionnelle “à la Fontaine” du type $$X_{cris}^{\ph} \Psi = p X_{cris}$$ où $\Psi \in \GL_n ( W(\overline{\mathbb{F}}_p)\unp)$ est la matrice d’un Frobenius sur un certain isocristal, $X_{cris}^\ph$ est obtenue en appliquant le Frobenius cristallin aux termes de la matrice et la réduite via $B^+_{cris}\ldrt \widehat{\overline{K}}$, $X\in M_n (\widehat{\overline{K}})$ vérifie $$\text{rang} (X) =n-1$$\ [ *Prérequis : La chapitre 3 fournit les rappels nécessaires concernant les espaces de Lubin-Tate et de Drinfeld. Le seul prérequis non rappelé est la théorie de la déformation de Grothendieck-Messing ([@Messing1]).* ]{}\ [*Avertissements : Dans tout l’article on supposera $p\neq 2$. Il est conseillé au lecteur de supposer en première lecture que $\O=\Zp$.* ]{}\ [*Remerciement : L’auteur remercie Jean François Dat qui, à partir d’une version préliminaire de cet article, a donné un exposé sur le sujet au groupe de travail de l’IHES au printemps 2004 et ainsi suggéré des améliorations.* ]{} Décomposition de Hodge-Tate des groupes $p$-divisibles dans le cas infiniment ramifié {#HoTa} ===================================================================================== Soit $K|\Qp$ un corps valué complet pour une valuation à valeurs dans $\R$ étendant celle de $\Qp$.\ Le théorème suivant se déduit du théorème \[bouqibouc\] de l’appendice, lui même démontré dans [@Periodes]. \[decHTO\] Soit $G$ un groupe $p$-divisible sur $\O_K$. Il y a une décomposition de Hodge-Tate : une suite de $\Gal (\overline{K}|K)$-modules continus $$0\ldrt \omega_G^* \otimes \mathcal{O}_{\widehat{\overline{K}}} (1) \ldrt T_p (G)\otimes_{\Zp} \mathcal{O}_{\widehat{\overline{K}}} \ldrt \omega_{G^D} \otimes \mathcal{O}_{\widehat{\overline{K}}} \ldrt 0$$ dont la cohomologie comme suite de $\mathcal{O}_{\widehat{\overline{K}}}$-modules est annulée par $p^{\frac{1}{p-1}}$. Si $\O_{K_0} \subset \O_K$ est un anneau de Cohen il existe des morphismes $\O_{K_0}$-linéaires Galois équivariantes des deux cotés de la suite exacte tels que composés avec les applications de droite et de gauche on obtienne la multiplication par un élément de valuation $p$-adique $\frac{1}{p-1}$. Bien sûr la suite est définie sans recours à la théorie de Hodge p-adique, celle-ci ne sert que pour démontrer que la cohomologie de la suite est annulée par tout élément de valuation supérieure ou égale à $\frac{1}{p-1}$. Rappelons en effet que $$\a_G: T_p (G) \ldrt \omega_{G^D}\otimes \O_{\widehat{\overline{K}}}$$ est défini de la façon suivante : pour $x\in T_p (G) = \Hom (\Qp/\Zp, G_{/\O_{\overline{K}}})$ celui ci donne par dualité de Cartier un morphisme $$x^D : G_{/\O_{\overline{K}}}^D\ldrt \mu_{p^\infty}$$ sur $\spec (\O_{\overline{K}})$, donc sur $\spf (\O_{\widehat{\overline{K}}})$. Il induit un morphisme $$(x^D)^*: \O_{\widehat{\overline{K}}} \frac{dT}{T}\ldrt \omega_{G^D}\otimes \O_{\widehat{\overline{K}}}$$ et alors $$\a_G ( x) = (x^D)^* \frac{dT}{T}$$ L’application $ \omega_G^* \otimes \mathcal{O}_{\widehat{\overline{K}}} (1) \ldrt T_p (G)\otimes_{\Zp} \mathcal{O}_{\widehat{\overline{K}}}$ est définie par dualisation de $\a_{G^D}$, c’est $\,^t\a_{G^D}(1)$ après identification de $T_p (G^D)$ avec $T_p (G)^*(1)$. Lorsque $K|K_0$ est de degré fini il existe de telles quasi-sections $\O_{K}$-linéaires telles que les composées donnent la multiplication par un élément de valuation $\frac{1}{p-1} + v_p (\mathcal{D}_{K/K_0})$ où $\mathcal{D}$ désigne la différente. cf. [@Fontaine1]. Cela peut s’obtenir avec les méthodes précédentes en calculant l’annulateur de $t$ dans un gradué de $A_{cris}\otimes_{\O_{K_0}}\O_{K}$. Dans notre cas infiniment ramifié cette suite exacte n’est en général pas scindée après inversion de $p$, il n’y a pas de décomposition de Hodge-Tate; l’obstruction provient de la Connexion de Gauss Manin et du fait que $\Omega^1_{K/K_0} \neq 0$ alors que ce module est nul dans l’article [@Fontaine1]. Un tel théorème ne peut s’obtenir par les méthodes de [@Fontaine1] qui sont purement “de Rham” : il faut utiliser les propriétés cristallines de certains objets associés aux groupe $p$-divisibles. \[HTet\] Si pour une extension $K'$ de $K$ contenue dans $\overline{K}$ l’action de $\Gal( \overline{K} | K)$ sur $T_p (G)$ se factorise via $\Gal (K' |K)$ il y a alors une décomposition de Hodge-Tate sur $K'$ c’est à dire une suite $$0\ldrt \omega_G^* \otimes \mathcal{O}_{\widehat{K'}} (1) \ldrt T_p (G)\otimes_{\Zp} \mathcal{O}_{\widehat{K'}} \ldrt \omega_{G^D} \otimes \mathcal{O}_{\widehat{K'}} \ldrt 0$$ dont la cohomologie est annulée par tout élément de valuation $\geq \frac{1}{p-1}$. Cela résulte de ce que la suite peut être définie sans recours à la théorie de Fontaine comme ci-dessus et que $\O_{\widehat{\overline{K}}}$ est fidèlement plat sur $\mathcal{O}_{\widehat{K'}}$. Soit $\Lie \, E(G)$ l’algèbre de Lie de l’extension vectorielle universelle de $G$. Elle possède une filtration donnée par la partie vectorielle $$0 \ldrt \omega_{G^D} \ldrt \Lie \, E(G) \ldrt \omega_G^* \ldrt 0$$ Alors, l’application de Hodge-Tate $\a_G$ se décrit également de la façon suivante $$\xymatrix@R=2mm@C=5mm{ \Hom (\Qp/\Zp, G) \ar[r] & \Hom_{\text{\it Filtré}} \left ( \Lie \, E(\Qp/\Zp), \Lie \, E(G)\right ) \ar@{=}[r] & \Hom_{\O_K} (\O_K,\omega_{G^D}) \ar[r] & \omega_{G^D} \\ && f \ar@{|->}[r] & f(1) }$$ puisque $\Lie \, E(\Qp/\Zp) =\Fil\, \Lie \, E(\Qp/\Zp) = \O_K$. Décomposition de Hodge-Tate d’un $\O$-module $\pi$-divisible ------------------------------------------------------------ On reprend les notations de l’appendice B de [@Cellulaire]. On y note $F$ une extension de degré fini de $\Qp$ et $\O=\O_F$ son anneau des entiers. Soit $K$ comme dans les sections précédentes et supposons que $K|F$. Soit $G$ un $\O$-module $\pi$-divisible sur $\O_K$, c’est à dire un groupe $p$-divisible muni d’une action de $\O$ induisant l’action naturelle sur son algèbre de Lie. Soit la suite exacte de Hodge-Tate définie dans la section précédente $$0\ldrt \omega_G^*\otimes\widehat{\overline{K}} \ldrt V_p (G)\otimes_{\Qp} \widehat{\overline{K}} \ldrt \omega_{G^D} \otimes\widehat{\overline{K}} \ldrt 0$$ Il s’agit d’une suite de $\O\otimes_{\Zp} \O_K$-modules.\ Rappelons ([@Cellulaire]) que l’on pose $$\widetilde{\omega}_{G^D}= \omega_{G^D}/I.\Lie E(G) \text{ où } I=\ker (\O\otimes_{\Zp} \O_K \twoheadrightarrow \O_K )$$ qui est la partie vectorielle de la $\O$-extension vectorielle universelle de $G$. Rappelons en effet que si $$0\ldrt V(G) \ldrt E(G) \ldrt G \ldrt 0$$ est l’extension vectorielle universelle de $G$ alors la $\O$-extension vectorielle universelle est le poussé en avant par l’application $V(G)\twoheadrightarrow V(G)/I.\Lie \, E(G)$ de cette extension.\ Considérons le diagramme $$\xymatrix@R=6mm{ 0 \ar[r] & \omega_G^*\otimes\widehat{\overline{K}} \ar[r] & V_p (G) \otimes_{\Qp} \widehat{\overline{K}} \ar[r] \ar@{->>}[d] & \omega_{G^D} \otimes\widehat{\overline{K}} \ar[r] \ar@{->>}[d] & 0 \\ && V_p (G)\otimes_F \widehat{\overline{K}} \ar[r] & \widetilde{\omega}_{G^D} \otimes \widehat{\overline{K}} }$$ \[relhtO\] La suite $$0 \ldrt \omega_G^*\otimes\widehat{\overline{K}} \ldrt V_p (G)\otimes_F \widehat{\overline{K}} \ldrt \widetilde{\omega}_{G^D} \otimes \widehat{\overline{K}} \ldrt 0$$ extraite du diagramme précédent est exacte. La surjectivité de l’application de droite est claire. De plus la composée des deux applications est nulle. D’après la remarque B.9 de l’appendice B de [@Cellulaire] $$\dim \omega_G^*\otimes\widehat{\overline{K}} + \dim \widetilde{\omega}_{G^D} \otimes \widehat{\overline{K}} = \dim V_p (G)\otimes_F \widehat{\overline{K}}$$ Il suffit donc de montrer que l’application de gauche est injective c’est à dire $$\omega_G^*\otimes\widehat{\overline{K}} \cap I.V_p(G)\otimes_{\Qp} \widehat{\overline{K}} = 0$$ Mais $I[\frac{1}{p}]$ est un produit de corps, il existe donc $e\in I[\frac{1}{p}]$ tel que $e.I[\frac{1}{p}]= I[\frac{1}{p}]$. Mais $I[\frac{1}{p}]. \omega_G^*\otimes\widehat{\overline{K}}=0$, d’où le résultat. Comme dans la section précédente l’application $\Hom_\O (F/\O_F ,G) \ldrt \widetilde{\omega}_{G^D}$ se définit de la façon suivante en termes de la $\O$-extension vectorielle universelle $E_\O (G)$ : si $x:F/\O_F\ldrt G_{O_K}$ alors il induit un morphisme $x_*:\O_{K}\ldrt \text{Lie} (E_\O (G))$ puisque $\O_{S/\Sigma}$ le faisceau structural du site cristallin s’identifie au cristal algèbre de Lie de la $\O$-extension vectorielle universelle de $F/\O_F$. Propriétés particulières de l’application de Hodge-Tate pour les groupes $p$-divisibles formels de dimension un =============================================================================================================== On reprend les notations de la section précédente. Les périodes de Hodge-Tate vivent dans l’espace de Drinfeld ----------------------------------------------------------- Nous utiliserons la proposition suivante afin de définir l’isomorphisme au niveau des points. Soit $G$ un groupe $p$-divisible sur $\O_K$. Considérons la suite de Hodge-Tate de $G^D$ $$0 \ldrt \omega_{G^D}^*\otimes \mathcal{O}_{\widehat{\overline{K}}} (1) \ldrt T_p (G)^* (1)\otimes_{\Zp} \mathcal{O}_{\widehat{\overline{K}}} \ldrt \omega_G \otimes \mathcal{O}_{\widehat{\overline{K}}} \ldrt 0$$ et plus particulièrement l’application $ \a_{G^D} : T_p (G)^* (1) \ldrt \omega_G \otimes \mathcal{O}_{\widehat{\overline{K}}} $. Si $G$ est un groupe $p$-divisible formel c’est à dire si sa fibre spéciale $G_k$ ne possède pas de partie étale alors l’application $\a_{G^D}$ est injective. Si $x:\Qp/\Zp\ldrt G_{/\O_{\overline{K}}}^D$ et $x^D : G_{/\O_{\overline{K}}}\ldrt \mu_{p^\infty}$ alors, $x=0 \ssi x^D=0$. De plus, $G$ étant formel $x^D=0$ ssi l’application tangente associée est nulle, d’où le résultat. \[DansOmega\] Soit $G$ un $\O$-module $\pi$-divisible formel de dimension $1$ sur $\O_K$. Notons $\Omega \subset \mathbb{P} ( V_p (G)^*(1))$ l’espace de Drinfeld au sens de Berkovich obtenu en enlevant les hyperplans $F$-rationnels. Alors, avec les notations précédentes $$(V_p (G)^*(1)\otimes_{F}\widehat{\overline{K}} \twoheadrightarrow \omega_G \otimes \widehat{\overline{K}}) \in \Omega ( \widehat{\overline{K}})$$ Raffinement, d’après Faltings ----------------------------- Soit $G$ un groupe p-divisible sur $\spf (\O_{K})$ et $$0 \ldrt \omega_{G^D} \ldrt E (G) \ldrt G \ldrt 0$$ son extension vectorielle universelle sur $\spf ((\O_{K})$. Alors, si $$\begin{aligned} E(G) (\O_{\widehat{\overline{K}}}) &=& \underset{n}{\limp} E(G) (\O_{\overline{K}}/ p^n \O_{\overline{K}}) \\ G(\O_{\widehat{\overline{K}}}) &=& \underset{n}{\limp} G(\O_{\overline{K}}/ p^n \O_{\overline{K}}) \end{aligned}$$ il y a une suite exacte $$0 \ldrt \omega_{G^D}\otimes \O_{\widehat{\overline{K}}} \ldrt E(G) (\O_{\widehat{\overline{K}}}) \ldrt G (\O_{\widehat{\overline{K}}}) \ldrt 0$$ Cela résulte aisément de la lissité des morphismes $E(G) \ldrt G$ réduits mod $\O_{\overline{K},\geq \l}$ pour tout $\l\in \Q_{>0}$ et de la surjectivité sur $\overline{k}$. Nous n’auront besoin de cela que pour $\Qp/\Zp$ pour lequel les points à valeurs dans une $\Zp$-algèbre de $E(\Qp/\Zp)$ sont $(R\oplus\Qp)/\Zp$, et cela est donc évident. \[fonda\] Soit $G$ un groupe p-divisible sur $\O_{K}$. Notons $$\a_G : T_p (G) \ldrt \omega_{G^D}\otimes \O_{\widehat{\overline{K}}}$$ l’application de Hodge-Tate. Soit $x=(x_n)_{n\geq 1} \in T_p (G)$ où $x_n\in G[p^n](\O_{\overline{K}})$. Soit $k$ un entier positif. Alors, $\a_G (x)\in p^k \omega_{G^D}\otimes \O_{\widehat{\overline{K}}}$ ssi $x_k$ se relève en un point de $p^k$-torsion dans l’extension vectorielle universelle de $G$ i.e. un point de $E(G)[p^n](\O_{\widehat{\overline{K}}})$. Considérons le diagramme $$\xymatrix{ 0 \ar[r] & \O_{\widehat{\overline{K}}} \ar[r] \ar[d]^{\times \a_G (x)} & (\O_{\widehat{\overline{K}}} \oplus \Qp)/\Zp \ar[r] \ar[d]^{E(x)} & \Qp/\Zp \ar[r]\ar[d]^x & 0 \\ 0 \ar[r] & \omega_{G^D} \otimes \O_{\widehat{\overline{K}}} \ar[r] & E(G) (\O_{\widehat{\overline{K}}}) \ar[r] & G( \O_{\widehat{\overline{K}}}) \ar[r] & 0 }$$ L’élément $x_k\in G[p^k]$ se relève en un élément $\widetilde{x_k}= E(x) ([(0,p^{-k})]) \in E(G) (\O_{\widehat{\overline{K}}})$ et de plus $$p^k \widetilde{x_k}=\a_G (x)$$ Les relèvements de $x_k$ à $E(G)(\O_{\widehat{\overline{K}}})$ formant un $\omega_{G^D}\otimes \O_{\widehat{\overline{K}}}$-torseur le résultat s’en déduit.\ La proposition qui suit s’applique en particulier aux groupes $p$-divisibles formels de dimension $1$ et est une légère amélioration d’un résultat de Faltings. Soit $G$ un groupe $p$-divisible formel sur $\O_K$. Considérons la suite de Hodge-Tate de $G^D$ $$0 \ldrt \omega_{G^D}^*\otimes \mathcal{O}_{\widehat{\overline{K}}} (1) \ldrt T_p (G)^* (1)\otimes \mathcal{O}_{\widehat{\overline{K}}} \xrig{\; \a_{G^D}\otimes Id\;} \omega_G \otimes \mathcal{O}_{\widehat{\overline{K}}} \ldrt 0$$ Supposons qu’il existe un groupe p-divisible isocline $H$ sur $k$ de hauteur $h$ tel que $\Z_{p^h}\hookrightarrow \End (H_{\bar{k}})$ et une isogénie $\rho : H\otimes_k \O_{K}/ p \O_{K} \ldrt G\otimes \O_{K}/p \O_{K}$ telle que $p^n \rho^{-1}$ soit une isogénie. Alors, si on note pour $M$ un “réseau” et $m\in M$ $\delta (m)= \sup \{k\;|\; p^{-k} m \in M\}$ on a $$\forall x\in T_p(G)^*(1) \;\;\; \delta (x) \leq \delta (\a_{G^D} (x)) \leq \delta (x) + n +1$$ Tout repose sur le lemme précédent. On peut supposer $k$ algébriquement clos. Soit $H_0$ un relèvement C.M. par $\Z_{p^h}$ de $H$ à $W(k)$. Soit $x\in T_p (G^D) \setminus p T_p (G^D)$. Supposons que $\a_{G^D} (x)\in p^{n+1} \omega_G\otimes \O_{\widehat{\overline{K}}}$. Le lemme précédent implique que $x_{n+1}\in G^D [p^{n+1}](\O_{\overline{K}})$ se relève en un élément $$\widetilde{x_{n+1}} \in E(G^D)[p^{n+1}]( \O_{\widehat{\overline{K}}})$$ Il résulte de la nature cristalline de l’extension vectorielle universelle que $\rho^D$ et $(p^n \rho^{-1})^D= p^n (\rho^D)^{-1}$ se relèvent en des morphismes $$\xymatrix@C=2cm{ E(H_0^D)_{/\O_{K}} \ar@<1ex>[r]^{E(p^n (\rho^D)^{-1})} & \ar@<1ex>[l]^{E ( \rho^D)} E (G^D) }$$ tels que $E(\rho^D) E( p^n (\rho^D )^{-1})=p^n$ et $ E( p^n (\rho^D )^{-1}) E(\rho^D ) = p^n$. Considérons l’élément $y=E(\rho^D ) (\widetilde{x_{n+1}})\in E ( H_0^D) [p^{n+1}] ( \O_{\widehat{\overline{K}}})$. Alors, $$y\neq 0$$ car sinon on aurait $$p^n \widetilde{x_{n+1}}= E(p^n (\rho^D)^{-1}) (y) = 0$$ mais via $E(G^D)\drt G^D$, $p^{n} \widetilde{x_{n+1}}\mapsto x_1\in G[p]^D$ ce qui est en contradiction avec $x_1\neq 1$ puisque $x\notin pT_p (G^D)$. Donc, $\exists z\in E(H_0^D)[p] ( \O_{\widehat{\overline{K}}}),z \neq 0$. Soit $w\in H_0[p]^D (\O_{\overline{K}})$ son image. On a $w\neq 0$ car $$\ker ( E(H_0^D)[p] ( \O_{\widehat{\overline{K}}}) \ldrt H_0 [p]^D ( \O_{\widehat{\overline{K}}})) = (\omega_{H_0}\otimes \O_{\widehat{\overline{K}}}) [p] =0$$ Soit $\iota : \Z_{p^h} \drt \End (H_0^D)$ l’action C.M. obtenue par dualisation de Cartier de l’action C.M. sur $H_0$. Le module de Tate $T_p (H_0^D)$ est $\Z_{p^h}$-libre de rang $1$. Donc, $\iota (\Z_{p^h}).w=H_0[p]^D(\O_{\overline{K}})$ qui est dans l’image de $E(H_0^D)[p]\drt H_0^D[p]$ puisque le morphisme $E(H_0^D)\drt H_0^D$ est $\Z_{p^h}$-équivariant. Par application du lemme \[fonda\] on en déduit que $$\a_{H_0^D} (T_p (H_0^D)) \subset p \omega_{H_0} \otimes \O_{\widehat{\overline{K}}}$$ ce qui est en contradiction avec le fait que le conoyau du morphisme de Hodge-Tate pour $H_0^D$ est annulé par un élément de valuation $\frac{1}{p-1}$.\ [*Interprétation :* ]{} Dans les cas des groupes $p$-divisibles formels de dimension $1$ la proposition précédent dit plus précisément que si le point $(G,\rho)$ est “à distance $\leq n$” du centre de l’espace de Lubin-Tate alors ses périodes de Hodge-Tate dans l’espace de Drinfeld sont “dans une boule de rayon $n+1$” dans l’immeuble associé à l’espace de Drinfeld. Formule exacte -------------- Dans l’article [@Rami] nous donnons une formule exacte plus précise que la proposition précédente dans le cas des groupes formels de dimension $1$ pour le point associé dans l’espace de Drinfeld (cf. corollaire \[DansOmega\]). Soit $|\mathcal{I}|$ la réalisation géométrique de l’immeuble de Bruhat-Tits du groupe $\text{PGL}_n$. Il y a une application $s: \Omega (\widehat{\overline{K}})\ldrt |\mathcal{I}|$. Cette formule exprime l’image par $s$ du point de Hodge-Tate en fonction de la filtration de ramification sur le module de Tate du groupe $p$-divisible. Notations concernant les espaces de Lubin-Tate et de Drinfeld {#kuipom} ============================================================= Nous reprenons les notations du premier chapitre de [@Cellulaire]. Rappelons que l’on note $\O=\O_F$. On note $L=W_\O(\Fqb)_\Q$ et $\s$ son Frobenius . Soit $\mathbb{H}$ un $\O$-module $\pi$-divisible formel de dimension $1$ et hauteur $n$ sur $\overline{\mathbb{F}}_p$ qui est défini sur $\mathbb{F}_q$. On a donc $\mathbb{H}= \Hb^{(q)}$. On note alors $\Pi=\text{Frob}_q\in \End (\Hb)$. Alors, $$\O_D= \O_{F_n} [\Pi] \iso \End (\Hb )$$ où l’on pose $$\O_{F_n}= W_\O (\mathbb{F}_{q^n})$$ Posons $\Gb=\Hb^n$ qui est un $\O$-module $\pi$-divisible formel de dimension $n$ et $\O$-hauteur $n^2$. Munissons le d’une structure de $\O_D$-module formel spécial au sens de Drinfeld en posant $$\iota: \O_D \ldrt \End ( \Gb) = M_n (\O_D)$$ tel que $$\forall x\in \O_{F_n} \;\; \iota ( x) = \text{diag} (x,x^\s,\dots, x^{\s^{n-1}} )$$ et $$\iota (\Pi) = \text{diag} ( \Pi,\dots,\Pi )$$ et donc $$\forall x\in \O_D\;\; \iota (x)= \text{diag} ( x, \Pi x\Pi^{-1},\dots, \Pi^{n-1} x \Pi^{-(n-1)} )$$ Tous les indices de $\Gb$ sont critiques et comme élément de $\text{PGL}_n (F) \bc \widehat{\Omega} ( \overline{\mathbb{F}_q})$ cela définit un point dans l’intersection de $n$ composantes irréductibles de la fibre spéciale. Il y a de plus une isogénie de $\O$-modules $\pi$-divisibles munis d’une action de $\O_D$ $$\begin{aligned} \label{ioD} \text{Id} \oplus \Pi \oplus \dots \oplus \Pi^{n-1}: \Hb^n \ldrt \Gb\end{aligned}$$ de degré $q^{\frac{1}{2} n (n-1)}$ où $\O_D$ agit diagonalement sur $\Hb^n$ (qui n’est pas spécial). Il y a donc un isomorphisme d’isocristaux munis d’une action de $D$ $$\mathbb{D} (\Gb)_\Q \iso \mathbb{D} (\Hb)_\Q^n$$ où les modules de Dieudonné sont les modules de Dieudonné covariants relatifs à $\O$ (il s’agit de l’évaluation du cristal algèbre de Lie de la $\O$-extension vectorielle universelle sur $\O_L\twoheadrightarrow k$, cf. l’appendice B de [@Cellulaire]).\ [*Convention : Désormais on notera $\mathbb{D}$ le module de Dieudonné covariant relatif à $\O$ d’un $\O$-module $\pi$-divisible noté $\mathbb{D}_\O$ dans l’appendice B de [@Cellulaire] et par cristal de Messing on entendra le cristal de Messing $\O$-extension vectorielle universelle défini dans l’appendice B de [@Cellulaire].*]{} Modules de Dieudonné -------------------- [*Nous n’utiliserons pas cette sous-section plus tard.*]{}\ Il y a une identification $$\mathbb{D} (\Hb) = \O_D\otimes_{\O_{F_n}} \O_L$$ muni de $\iota : \O_D \ldrt \End ( \Hb)$ tels que si $V$ désigne le Verschiebung $$\begin{aligned} \forall d\otimes \l \in \mathbb{D} (\Hb) \;\;\, V ( d\otimes \l) & = & d \Pi \otimes \l^{\s^{-1}} \\ \forall d'\in \O_D\;\;\; \iota (d') ( d\otimes \l) & =& d' d \otimes \l \end{aligned}$$ Alors, $$\mathbb{D} (\Gb) = \O_D\otimes_{\O_F} \O_L$$ muni de son action de $\iota$ de $\O_D$ et $$\begin{aligned} \forall d\otimes \l \in \mathbb{D} (\Gb) \;\;\; V ( d\otimes \l) & = & d \Pi \otimes \l^{\s^{-1}} \\ \forall d'\in \O_D\;\;\; \iota (d') ( d\otimes \l) & =& d' d \otimes \l \end{aligned}$$ Dès lors on a un isomorphisme de $\O_L$-modules $$\begin{aligned} \O_D\otimes_{\O_F} \O_L & \xrig{\;\sim \;} & \prod_{k\in \Z/n\Z} \O_D\otimes_{\O_{F_n},\s^{k}} \O_L \\ d\otimes \l & \longmapsto & ( d \otimes \l)_{k\in \Z/n\Z}\end{aligned}$$ Et dans ces coordonnées l’identification $\mathbb{D} (\mathbb{H})^n = \mathbb{D} (\mathbb{G})$ est donnée par $$\begin{aligned} (\O_D\otimes_{\O_{F_n}} \O_L )^n &\ldrt & \prod_{k\in \Z/n\Z} \O_D \otimes_{\O_{F_n},\s^{k}} \O_L \\ (x_k\otimes\l_k)_{0\leq k \leq n-1} &\longmapsto & (\Pi^{-k} x_k\Pi^k\otimes \l_k)_{k\in\Z/n\Z} \end{aligned}$$ et l’isogénie (\[ioD\]) par $$\begin{aligned} (\O_D\otimes_{\O_{F_n}} \O_L )^n &\ldrt & \prod_{k\in \Z/n\Z} \O_D \otimes_{\O_{F_n},\s^{k}} \O_L \\ (x_k\otimes\l_k)_k &\longmapsto & (x_k\Pi^k\otimes \l_k)_k \end{aligned}$$ Notations concernant les espaces de Lubin-Tate ---------------------------------------------- Rappelons que l’on note $\breve{F} =\widehat{F^{nr}}$ le complété de l’extension maximale non-ramifiée de $F$ dans une clôture algébrique de celui-ci. On fixe un isomorphisme entre $\overline{\mathbb{F}}_q$ et le corps résiduel de $\breve{F}$. Celui-ci induit un isomorphisme $L\simeq \breve{F}$. Néanmoins on ne confondra pas toujours ces deux corps, l’un, $L$, étant un corps “abstrait”, l’autre, $\breve{F}$, un corps plongé (un corps reflex). Soit $K$ un corps valué complet (pour une valuation de rang $1$ c’est à dire à valeurs dans $\R$) extension de $\breve{F}$. On appelle point à valeurs dans $K$ de la tour de Lubin-Tate un triplet $(H,\rho,\eta)$ à isomorphisme près où - $H$ est un $\O$-module $\pi$-divisible formel de dimension $1$ sur $\mathcal{O}_{K}$ - $\rho$ est une quasi-isogénie $\mathbb{H}\otimes_{\overline{\mathbb{F}}_q} \O_{K}/ p \O_{K} \drt H\otimes \O_{K}/ p \O_{K}$ - $\eta : \O^n \iso T_p (H)$ est un isomorphisme de $\Gal(\overline{K}|K)$-modules, le membre de gauche étant muni de l’action triviale de Galois. On note $\M^{\LT}_\infty (K)$ cet ensemble qui est muni d’une action de $D^\times \times \GL_n (\O_F)$, où $D^\times $ agit à gauche par $d.\rho=\rho \circ d^{-1}$ et $\GL_n (\O_F)$ à droite par $\eta.g=\eta\circ g$, et d’une donnée de descente $\a$ vers $F$. On renvoie à la section 1.4 de [@Cellulaire] pour la définition de la donnée de descente $\a$. On verra bientôt que cet ensemble est en fait muni d’une action de $\GL_n (F)$. Dans la définition précédente, la définition de $\eta$ peut être remplacée par : $\eta$ est un isomorphisme $$\eta:\O^n\iso \Hom_\O ( F/\O, H)$$ \[ruygop\] On note $\M^{\LT} (K)/\sim$ l’ensemble des couples $(H,\rho)$ comme précédemment à isogénie déformant un élément de $F^\times$ près i.e. $(H,\rho)\sim (H',\rho') \ssi \exists f : H\ldrt H'$ une isogénie et un $x\in F^\times$ tels que le diagramme suivant commute $$\xymatrix{ H \text{ mod } \pi \ar[r]^{f \text{ mod } \pi} & H'\text{ mod } \pi \\ \mathbb{H}\otimes \O_K/\pi\O_K \ar[r]^{x}\ar[u]^\rho & \mathbb{H}\otimes \O_K/\pi\O_K \ar[u]^{\rho'} }$$ Cet ensemble est muni d’une action de $D^\times$. Une définition équivalente consiste à remplacer à isogénie déformant un élément de $F^\times$ près par à isogénie déformant une puissance de $\pi$ près. Une troisième définition équivalente consiste à dire que $(H,\rho)$ est équivalent à $(H',\rho')$ ssi il existe une quasi-isogénie $f:H\ldrt H'$ sur $\spec (\O_K)$ telle que $f\rho=\rho'$ i.e. ssi la quasi-isogénie $\rho'\rho^{-1}$ sur $\spec (\O_K/p\O_K)$ se relève en une quasi-isogénie sur $\spec(\O_K)$. Notations concernant les espaces de Drinfeld -------------------------------------------- Soit $K$ un corps valué complet extension de $\breve{F}$. On appelle point à valeur dans $K$ de la tour de Drinfeld un triplet $(G,\rho,\eta)$ à isomorphisme près où - $G$ est un $\O_D$-module $\pi$-divisible formel spécial sur $\spf (\O_{K})$ - $\rho$ est une quasi-isogénie $\O_D$-équivariante $\mathbb{G} \otimes_{\overline{\mathbb{F}}_q} \O_{K}/p\O_{K}\ldrt G\otimes_{\O_{K}}\O_{K}/p\O_{K} $ - $\eta: \O_D\iso T_p (G)$ est un isomorphisme de $\O_D$-modules galoisiens. On note $\M^\D_\infty (K)$ cet ensemble muni de son action de $\GL_n (F)\times D^\times$ où $\GL_n (F)= \End_{D-\text{éq.}}(\mathbb{G})_\Q^\times$ agit à gauche via $g.\rho=\rho\circ g^{-1}$ et $\O_D^\times$ agit à droite via $\eta.d = \eta \circ d^{-1}$ où $ d^{-1}: \O_D\iso \O_D$ est $d'\mapsto d' d^{-1}$. L’action de $\Pi\in D^\times$ est définie par $\Pi. (G,\rho,\eta) = (G/G[\Pi], \ph \circ \rho, \eta.\Pi)$ où $\ph : G\twoheadrightarrow G/G[\Pi]$ et $$\xymatrix@C=15mm{ \O_D \ar[d]^{.\Pi} \ar[r]^\eta_\sim & T_p (G) \ar@{^(->}[d]^{\ph_*} \\ \O_D \ar[r]^{\eta.\Pi}_\sim & T_p ( G/G[\Pi]) }$$ On renvoi à la section 1 de [@iso4] pour la définition générale de l’action d’un élément de $D^\times$. On notera que $\{(x,x)\;|\; x\in F^\times \}\subset \GL_n (F)\times D^\times$ agit trivialement. Dans la définition précédente, la définition de $\eta$ peut être remplacée par : on se donne un élément $$\eta (1)\in \Hom_\O (F/\O, G)\setminus \Pi. \Hom_\O (F/\O, G)$$ Le groupe associé à l’espace de Rapoport-Zink précédent et noté $G$ dans [@RZ] est dans notre cas $(D^{opp})^\times$. Dans la définition précédente on l’a identifié à $D^\times$ via $d\mapsto d^{-1}$ ce qui explique que l’on pose $d.\eta = \eta \circ d^{-1}$ alors que la définition usuelle des correspondances de Hecke est pour $g\in G(\Qp)$, $\eta \mapsto \eta \circ g$. L’identification $\GL_n (F)= \End_{D-\text{éq}} (\mathbb{G})_\Q^\times$ se déduit, avec les notations des sections qui suivent, de l’identification $$\End_{D-\text{éq}} (\mathbb{G})_\Q = \End_{D-\text{éq}}\left (\mathbb{D}(\mathbb{G})_\Q, V \right ) = \End_\O \left ( \mathbb{D}(\mathbb{G})_{\Q,0}^{V^{-1}\Pi}\right )$$ et d’un choix d’une base de $\mathbb{D}(\mathbb{G})_{\Q,0}^{V^{-1}\Pi}$ (cf. les section suivantes). Comme dans la définition \[ruygop\] on note $\M^{\D} (K)/\sim$ l’ensemble des couples $(G,\rho)$ à isogénie déformant un élément de $F^\times$ près. Cet ensemble est muni d’une action de $\GL_n (F)$. Quelques rappels sur Drinfeld classique --------------------------------------- Notons $$N^{\D}= \mathbb{D} (\Gb)_\Q = D\otimes_F L$$ un isocristal relativement à l’extension $L|F$. On a $$\begin{aligned} N^\D & =& \bigoplus_{i\in \Z/n \Z} N^\D_i \;\text{ où } N^\D_i =\{n\in N\; |\; \forall x\in F_n \; \iota (x).n= x^{\s^{-i}} n \; \} \end{aligned}$$ où relativement à cette graduation $\deg V= \deg \Pi =+1$. On a de plus $$\begin{aligned} D\otimes_F L &=& \bigoplus_{a\in \Z/n\Z} D\otimes_{F_n,\s^{-a}} L \\ &=& \bigoplus_{a,b\in \Z/n\Z} L. \underbrace{\iota(\Pi^b)(e_a)}_{e_{a,b}} \end{aligned}$$ où $e_a =1\otimes 1\in D\otimes_{F_n,\s^{-1}}L$ et $e_{a,b} =\Pi^b\otimes 1$. De plus, avec ces coordonnées $$V(e_{a,b})= \iota(\Pi)(e_{a,b}) = \pi^{\delta_{b,n-1}} e_{a,b+1}$$ et $$N^\D_i = \bigoplus_{a+b =i} L.e_{a,b}$$ L’opérateur $$V^{-1} \Pi : N_0^\D\ldrt N^\D_0$$ est de pente $0$ et fait de $(N_0^\D,V^{-1} \Pi)$ un isocristal unité : $$\begin{aligned} (N^\D_0, V^{-1} \Pi)\simeq ((N^\D_0)^{V^{-1} \Pi}\otimes_F L, Id\otimes \s) \\ (N^\D)^{V^{-1} \Pi}_0 = \bigoplus_{a+b = 0} F.e_{a,b} \simeq F^n\end{aligned}$$ Rappelons maintenant qu’il y a des bijections $$\begin{aligned} \left \{ {\text{Sous isocristaux D-stables de} \atop (N^\D,\ph)} \right \} &\iso & \left \{ \text{Sous } F\text{-ev. de} \atop (N^\D)^{V^{-1} \Pi} \right \} \\ N & \longmapsto & N_0^{V^{-1} \Pi}\end{aligned}$$ où $\ph$ désigne le Frobenius, et pour $K|L$ une extension valuée comme dans le premier chapitre $$\begin{aligned} \left \{ \text{ Filtrations D-stables de codimension } n \atop \text{dans } N^\D\otimes_L K \right \} &\iso & \left \{ \text{Filtrations de codimension } 1 \text{ dans}\atop (N_0^\D)^{V^{-1} \Pi} \otimes_F K \right \} \\ \Fil=\oplus_{i\in \Z/n \Z} \Fil_i &\longmapsto & \Fil_0\end{aligned}$$ Alors, $(N^\D,\ph,\Fil)$ est faiblement admissible ssi pour tout sous-isocristal $N\subset N^\D$ on a $$t_H ( N, N\otimes_L K\cap \Fil) \leq t_N (N,\ph)$$ où $t_H$, resp. $t_N$, désigne le point terminal du polygone de Hodge, resp. Newton. L’existence de la filtration de Harder-Narashiman et sa canonicité sur $(N^\D,\ph,\Fil)$ impliquent que celle-ci est $D$-stable si $\Fil$ l’est et qu’il suffit donc dans ce cas là de tester la condition d’admissibilité faible sur les sous-isocristaux $D$-stables ([@RZ] chapitre 1). Restreignons-nous aux filtrations $\Fil$ $D$-stables de codimension $n$. Via les deux bijections ci dessus on trouve aisément : $$\left \{ \text{Fil} \subset N^\D\otimes_L K \text{ telles que } (N^\D,\ph,\Fil ) \atop \text{ est faiblement admissible } \right \} \iso \left \{ \Fil'\subset (N^\D_0)^{V^{-1} \Pi}\otimes K \; \text{ telles que } \atop \forall N'\subset (N^\D_0)^{V^{-1} \Pi} \text{ de dim.} 1 \; \Fil'\cap (N'\otimes K) = (0) \right \}$$ la condition portant donc sur les sous-isocristaux $D$-stables de dimension $n$ dans $(N^\D,\ph)$ qui sont paramétrés par $\check{\mathbb{P}}\left ( ( N^\D_0)^{V^{-1} \Pi} \right )(F)$ où pour $E$ un $F$-e.v. on note $\mathbb{P} (E)$ l’espace des filtrations de codimension $1$ de $E$ et $\check{\mathbb{P}} (E) = \mathbb{P} \left ( \check{E} \right )$ l’espace des droites de $E$. La relation d’incidence entre $\mathbb{P} (E)$ et $\check{\mathbb{P}} (E)$ est donnée par : pour $H\in \check{\mathbb{P}} (E)$, $H$ définit une droite $D$ dans $E$ et donc un hyperplan encore noté $H$ dans $\mathbb{P} (E)$ formé des filtrations contenant $D$. Ainsi $ \check{\mathbb{P}} (E)$ paramètre les hyperplans dans $\mathbb{P} (E)$. On a donc $$\left \{ (N^\D,\ph,\Fil) \text{ faiblement admissible} \atop \Fil \; D\text{-stable} \right \} \iso {\mathbb{P}}\left ( N^\D_0\right )(K)\setminus \bigcup_{H\in \check{ \mathbb{P}}\left ((N^\D_0)^{V^{-1} \Pi}\right )(F)} H(K) = \Omega (K)$$ où $\Omega$ désigne l’espace de Drinfeld vu comme espace de Berkovich. Dit d’une autre façon, $\Omega (K)$ est l’ensemble des filtrations $\Fil$ de codimension $1$ dans $N^\D_0\otimes_L K$ telles que l’application suivante soit injective $$(N^\D_0)^{V^{-1} \Pi} \hookrightarrow N^\D_0\otimes_L K/\Fil$$ Description de $\M^{\LT} (K)/\sim$ en termes de modules filtrés =============================================================== Il s’agit ici d’étendre la théorie de Fontaine des modules faiblement admissibles pour les groupes de dimension $1$ à des corps $K$ non forcément de valuation discrète. Soit un couple $(H,\rho)$ où $ H$ est un $\O$-module $\pi$-divisible sur $\spf (\O_K)$ et $\rho$ une rigidification avec $\mathbb{H}$ modulo $p$. L’algèbre de Lie de la $\O$-extension vectorielle universelle de $H$, $M$, est munie d’une filtration $\Fil \subset M$ telle que $M/\Fil \simeq \omega_H^*$. La nature cristalline de la $\O$-extension vectorielle universelle induit un isomorphisme $$\rho_* : \mathbb{D} (\Hb)_\Q \otimes_{L} K \iso M\unp$$ et $\Fil_H= \rho_*^{-1} (\Fil\unp) \subset\mathbb{D} (\Hb)_\Q \otimes K$ est une filtration de codimension $1$. \[four\_D\] L’application précédente $(H,\rho_H)\mapsto \Fil_H$ induit une bijection $D^\times$-équivariante $$\M^{\LT} (K)/\sim \iso \P \left ( \mathbb{D} (\Hb)_\Q \right ) (K)$$ C’est une conséquence de l’existence du domaine fondamental de Gross Hopkins et du fait que la restriction du morphisme des périodes à ce domaine est un isomorphisme d’espaces rigides (cf. [@HopkinsGross] et [@Cellulaire]). Plus précisément, étant donné une filtration $\Fil \in \mathbb{P}(K)$ il existe un entier $i$ tel que $\Pi^i.\Fil$ soit dans l’image du domaine fondamental de Gross-Hopkins. Il existe donc un couple $(H,\rho)\in \M^{\LT} (K)$ induisant $\Pi^i.\Fil$. Alors, $\Pi^{-i}.(H,\rho)$ induit $\Fil$, d’où la surjectivité de l’application. Pour l’injectivité, si $(H,\rho)$ et $(H',\rho')$ induisent la même filtration alors la quasi-isogénie $$\rho'\circ \rho^{-1} : H\otimes {\O_K/p\O_K} \ldrt H'\otimes \O_K/p\O_K$$ est telle que l’application induite $$(\rho'\circ \rho^{-1})_* : \text{Lie} (E(H))\ldrt \text{Lie} ( E ( H'))$$ vérifie $$\exists a\in \N\; (\rho'\circ \rho^{-1})_* (\omega_{H^D}) \subset p^{-a} \omega_{{H'}^D}$$ et donc d’après la théorie de Messing l’isogénie $$p^a \rho'\circ \rho^{-1} : (H,\rho) \ldrt (H',\rho') \text{ mod } p$$ sur $\spec(\O_K/p\O_K)$ se relève en une isogénie : $(H,\rho)\sim (H',\rho')$. Il se peut qu’en général pour des espaces de périodes plus généraux l’application des périodes ne soit pas surjectivité sur les $K$-points ! i.e. qu’un module faiblement admissible ne provienne pas forcément d’un groupe p-divisible sur $K$ quelconque. Par contre le calcul des fibres de l’application des périodes (la partie injectivité dans la démonstration précédente) reste valable en général. Description de $\M^\D (K)/\sim$ en termes de module filtré ========================================================== Soit $(G,\rho)$ où $G$ est un $\O_D$-module formel spécial sur $\spf (\O_K)$ et $\rho$ une rigidification avec $\mathbb{G}$. Soit $M$ l’extension vectorielle universelle de $G$ qui est filtrée via $\Fil\subset M$ où $M/\Fil \simeq \omega_G^*$. Il y a une décomposition sous l’action de $F_n$ $$M=\bigoplus_{i\in \Z/n \N} M_i \;\text{ et } \; \Fil= \bigoplus_{i\in\Z/ n\Z} \Fil_i$$ d’où un élément $\Fil_{G,0} = \rho_*^{-1} (\Fil _0) \subset \mathbb{D} (\Gb)_{\Q,0} \otimes K, \;\Fil_{G,0}\in \Omega (K)$. \[DrP\] L’application $(G,\rho)\mapsto \Fil_{G,0}$ induit une bijection $\GL_n (F)$-équivariante entre $\M^\D (K)/\sim$ et $\Omega (K)$ avec $$\Omega (K) = {\mathbb{P}} \left ( \mathbb{D} (\Gb)_{\Q,0} \right )(K)\setminus \bigcup_{H\in \check{ \mathbb{P}} ((\mathbb{D}(\Gb)_{\Q,0})^{V=\Pi})(F)} H(K)$$ Cela résulte de ce que l’application des périodes est un isomorphisme rigide pour les espaces de Drinfeld, de l’égalité $\Omega (K)=\widehat{\Omega} (\O_K)$ où $\widehat{\Omega}$ est le schéma formel de Drinfeld, du théorème de Drinfeld et du fait que les deux applications de périodes, celle définie par Drinfeld et celle définie dans [@RZ], coïncident. Prolongement des isogénies ========================== Prolongement ------------ Soit $H$ un groupe $p$-divisible sur $\spec (\O_K )$. Soit $\mathcal{C}$ l’ensemble des classes d’isomorphismes de couples $(H',f)$ où $H'$ est un groupe $p$-divisibles sur $\O_K$ et $f:H'\drt H$ une quasi-isogénie sur $\spec (\O_K)$. Pour un tel couple $T_p (f): T_p (H')\hookrightarrow V_p (H)$. \[adsch\] L’application $$\begin{aligned} \mathcal{C} & \drt & \{ \text{ réseaux } \Gal (\overline{K}|K)-\text{stables dans } V_p (H)\; \} \\ (H',f) & \mapsto & Im \;T_p (f)\end{aligned}$$ est une bijection. C’est une conséquence de ce que pour $\mathcal{G}$ un groupe fini localement libre sur $\O_K$ et $I\subset \mathcal{G}_\eta$ un sous-groupe fini localement libre il existe un unique prolongement $\mathcal{I}\hookrightarrow \mathcal{G}$ de l’inclusion $I\subset \mathcal{G}_\eta$ où $\mathcal{I}$ est un sous-groupe fini localement libre. Le groupe $\mathcal{I}$ est obtenu par adhérence schématique (cf. le chapitre 2 de [@Ray1] pour le cas de valuation discrète et le lemme qui suit en général). Soit $E\subset K^n$ un sous-$K$-espace vectoriel. Alors, $E\cap \O_K^n$ est un $\O_K$-module libre de rang fini facteur direct. Définition de l’action de $\GL_n (F)$ sur $\M^{\LT}_\infty (K)$ --------------------------------------------------------------- On a définit une action de $\GL_n (\O_F)$ sur $\M^{\LT}_\infty (K)$. Utilisons le lemme précédent pour étendre cette action à $\GL_n (F)$. Soient $[(H,\rho,\eta)]\in \M_\infty^{\LT} (K)$ et $g\in \GL_n (F)$ (les crochets signifient que l’on prend une classe d’isomorphisme de triplets). D’après le lemme précédent au réseau Galois-stable $\eta (g. \O_F^n)$ dans $V_p (H)$ correspond une quasi-isogénie $\ph :H\ldrt H'$ sur $\spec (\O_K)$ telle que $\ph_*^{-1} ( T_p (H')) = \eta ( g.\O_F^n)$ où $\ph_* : V_p (H)\iso V_p (H')$. On pose alors $$g.[(H,\rho,\eta)] = [(H',(\ph \text{ mod } p)\circ \rho,\ph_*\circ \eta \circ g)$$ Description de $\M^{\LT}_\infty (K)$ en termes de modules filtrés rigidifiés ============================================================================ Soit $[(H,\rho)]\in \M^{\LT}(K)/\sim$. Une rigidification du module de Tate de la classe d’isogénie $[(H,\rho)]$ est un isomorphisme de modules galoisiens $$\eta : F^n\iso V_p (H)$$ qui induit donc naturellement $\forall (H',\rho')\sim (H,\rho)$ un isomorphisme $$F^n\iso V_p (H')$$ via l’isomorphisme canonique $V_p (H)\iso V_p (H')$ induit par le relèvement de la quasi-isogénie $\rho'\rho^{-1}$ sur $\spec (\O_K)$. \[peftu\] L’application naturelle $$\M_\infty^{\LT} (K)\ldrt \{ \left ([(H,\rho)],\eta \right )\; |\; [(H,\rho)]\in \M^{\LT} (K)/\sim\text{ et } \eta \text{ une rigidification} \;\}$$ est une bijection $\GL_n (F)\times D^\times$-équivariante. Décrivons l’inverse de cette application. Soit $([(H,\rho)],\eta)$ dans le membre de droite. D’après le lemme \[adsch\] il existe un unique $(H',\rho')\sim (H,\rho)$ tel que $$\eta : \O_F^n\iso T_p (H')$$ On associe alors à $([(H,\rho)],\eta)$ le triplet $(H',\rho',\eta')$. On vérifie aisément que ces deux applications sont inverses l’une de l’autre. Ainsi les fibres de l’application $\M_\infty^{\LT} (K) \ldrt \M^{\LT} (K)/\sim$ sont des $\GL_n (F)$-torseurs. Soit $U\subset \GL_n (\O_F)$ un sous-groupe compact-ouvert et $\M_U^{\LT} (K)$ l’ensemble des $K$-points de l’espace de Lubin-Tate en niveau $U$ (que nous n’avons pas défini dans cet article). En fait on a la décomposition plus précise d’extensions “Galoisiennes” $$\xymatrix{ \ar@{-}@/_4pc/[dd]_{GL_n (F)} \M_\infty^{\LT} (K) \ar[d] \ar@{-}@/^2pc/[d]^{U} \\ \M_U^{\LT} (K) \ar[d] \\ \M^{\LT} (K)/\sim \ar[r]^{\sim} & \mathbb{P} ( \mathbb{D} (\Hb)) (K) }$$ et on peut donc ainsi dire que l’espace des périodes de Gross-Hopkins est le quotient de l’espace de Lubin-Tate en niveau infini par $\GL_n (F)$. \[derhj\] Il y a une bijection $\GL_n (F)\times D^\times$-équivariante entre l’ensemble $\M_\infty^{\LT} (K)$ et l’ensemble des couples $(\Fil,\zeta)$ où $$\Fil\in\mathbb{P} \left (\mathbb{D} (\mathbb{H} \right )_\Q) (K)$$ $\zeta= (\zeta_1,\dots,\zeta_n)$ où $$\forall i\; \zeta_i\in\Hom_\O (F/\O_F,\mathbb{H}\otimes_{\overline{\mathbb{F}}_p} \O_K/p\O_K)\unp$$ sont linéairement indépendants sur $F$ i.e. induisent une injection $F^n\hookrightarrow \Hom_\O (F/\O_F,\mathbb{H}\otimes \O_K/p\O_K)\unp$, et $\forall i\;$ le morphisme induit sur l’évaluation du cristal de Messing sur l’épaississement $\O_K\twoheadrightarrow \O_K/p\O_K$, $\zeta_{i*} : K\drt \mathbb{D} (\mathbb{H})\otimes K$, vérifie $\zeta_{i*}(1)\in \Fil$. Il s’agit d’une conséquence de la proposition précédente couplée à la proposition \[four\_D\] et au critère de relèvement de Messing vis à vis de l’idéal à puissances divisées engendré par $p$. Dans cet énoncé, si $(e_i)_{1\leq i \leq n}$ est la base canonique de $F^n$ on a posé $$\zeta_i = \rho^{-1} \circ \left ( \eta (e_i) \text{ mod } p \right)$$ Description de $\M_\infty^{\LT} (K)$ en termes d’algèbre linéaire {#bujiko} ----------------------------------------------------------------- Cette sous-section ne sera pas nécessaire dans la suite mais donne une description purement en termes d’algèbre linéaire de $\M_\infty^{\LT} (K)$. \[deall\] Supposons $F=\Qp$. Il y a une bijection $\GL_n (F)\times D^\times$-équivariante entre l’ensemble $\M_\infty^{\LT} (K)$ et l’ensemble des couples $(\Fil,\xi)$ tels que $$\Fil\in\mathbb{P} (\mathbb{D} (\mathbb{H})_\Q) (K)$$ et $$\xi : F^n\hookrightarrow \Fil\left ( \mathbb{D}(\mathbb{H})_\Q\otimes_L B^+_{cris} (\O_K)\right )^{\ph = p}$$ (qui est alors automatiquement un isomorphisme) Il s’agit d’une conséquence de la proposition \[peftu\] couplée à la proposition \[four\_D\] et au théorème \[crop\]. Description de $\M_\infty^{\D} (K)$ en termes de modules filtrés rigidifiés =========================================================================== Soit $[(G,\rho)]\in\M^{\D} (K)/\sim$. Une rigidification de la classe d’isogénie $[(G,\rho)]$ est un isomorphisme de $D$-modules galoisiens $$\eta : D\iso V_p(G)$$ qui induit donc naturellement $\forall (G',\rho')\sim (G,\rho)$ un isomorphisme $$D\iso V_p (G')$$ via l’identification $V_p (G)\iso V_p (G')$ induite par le relèvement de $\rho'\rho^{-1}$. \[huyto\] Dans la proposition précédente la donnée de $\eta$ est équivalente à celle de $\zeta = \eta (1) \in \Hom_\O (F/\O_F, G) \unp\setminus \{ 0 \}$. L’application naturelle $$\M_\infty^{\D} (K) \ldrt \{\; \left ([(G,\rho)],\eta \right )\; |\; [(G,\rho)]\in\M^{\D}(K)/\sim \text{ et } \eta \text{ une rigidification}\;\}$$ est une bijection $\GL_n (F)\times D^\times$-équivariante. Elle est identique à celle de la proposition \[peftu\] en utilisant le lemme \[adsch\].\ Comme pour l’espace de Lubin-Tate il y a une description pour $U\subset \O_D^\times$ un sous-groupe ouvert $$\xymatrix{ \ar@{-}@/_4pc/[dd]_{D^\times} \M_\infty^{\D} (K) \ar[d] \ar@{-}@/^2pc/[d]^{U} \\ \M_U^{\D} (K) \ar[d] \\ \M^{\D} (K)/\sim \ar[r]^(.6){\sim} & \Omega (K) }$$ \[descDri\] Il y a une bijection $\GL_n (F)\times D^\times$-équivariante entre l’ensemble $\M^{\D}_\infty (K)$ et l’ensemble des couples $(\Fil,\zeta)$ où $$\Fil\in \Omega (K)\subset \mathbb{P}\left (\mathbb{D}(\mathbb{G})_{\Q,0}\right )(K)$$ et $$\zeta \in \Hom_\O (F/\O_F, \mathbb{G}\otimes \O_K/p\O_K)\unp \setminus \{ 0\}$$ est tel que si $\zeta_* : K\drt \mathbb{D} (\mathbb{G})_{\Q}\otimes K$ est l’application induite sur l’évaluation du cristal de Messing alors si $$\zeta_* (1) = \oplus_i \zeta_*(1)_i\in \bigoplus_{i\in \Z/n\Z} \mathbb{D}(\mathbb{G})_{\Q,i}\otimes K$$ on a $$\forall i\;\; \Pi^{-i} \zeta_* (1)_i \in \Fil$$ C’est une conséquence de la proposition précédente couplée à la remarque \[huyto\] et au théorème de relèvement de Messing. Dans l’énoncé on a posé $$\zeta = \rho^{-1} \circ \left ( \eta (1) \text{ mod } p \right )$$ Description de $\M_\infty^{\D} (K)$ en termes d’algèbre linéaire ---------------------------------------------------------------- Comme dans la section \[bujiko\] cette section ne servira pas dans la suite. Supposons $F=\Qp$. Il y a une bijection $\GL_n (F)\times D^\times$-équivariante entre l’ensemble $\M^{\D}_\infty (K)$ et l’ensemble des couples $(\Fil,\xi)$ où $$\Fil\in \Omega (K)\subset \mathbb{P}(\mathbb{D}(\mathbb{G})_{\Q,0})(K)$$ et $$\xi \in \Fil(\mathbb{D}(\mathbb{G})_{\Q,0}\otimes B^+_{cris}(\O_K))^{(V^{-1}\Pi\otimes \ph)^n=p}\setminus \{ 0 \}$$ La bijection au niveau des points {#coeulop} ================================= On fixe $$\Delta : \Hb^n \ldrt\Gb$$ une quasi-isogénie compatible à l’action de $\O_D$ comme par exemple celle définie dans la section \[kuipom\]. On fixe un isomorphisme $$\begin{aligned} \label{out1} \mathbb{D} (\mathbb{H})_{\Q,0}^{V=\Pi} \simeq F\end{aligned}$$ qui induit via $\Delta$ un isomorphisme $$\begin{aligned} \label{out2} \mathbb{D} (\mathbb{G})_{\Q,0}^{V=\Pi} \simeq F^n\end{aligned}$$ et induit donc $$\begin{aligned} \label{out3} \mathbb{D} (\mathbb{G})_{\Q,0} \simeq L^n\end{aligned}$$ De plus, l’isomorphisme (\[out1\]) induit également un isomorphisme $\mathbb{D}(\mathbb{H})_{\Q,0}\simeq L$ et $$\begin{aligned} \label{out4} \mathbb{D}(\mathbb{H})_\Q =\bigoplus_{j\in\Z/n\Z} \mathbb{D}(\mathbb{H})_{\Q,j} \underset{\sum_j \Pi^{-j}}{\xrig{\;\;\sim\;\;}} \mathbb{D} (\mathbb{H})_{\Q,0}^n \simeq L^n\end{aligned}$$ Le choix de l’isomorphisme (\[out1\]) et de ceux qui s’en suivent n’est pas vraiment nécessaire à la démonstration mais permet d’identifier les espaces de périodes du coté Lubin-Tate et Drinfeld à des sous-espaces de $\mathbb{P}^n$. Rappelons que l’on a des extensions de corps valués $$K| \breve{F} | F |\Qp$$ et un isomorphisme $L\simeq \breve{F}$. L’application $\M^\D_\infty (K)\ldrt \M^{\LT}_\infty (K)$ {#dltisog} --------------------------------------------------------- Soit $(G,\rho_G,\eta_G)\in \M_\infty^\D (K)$. Rappelons (cf. théorème \[descDri\]) qu’on lui associe un couple $(\Fil_G,\zeta_G)$ où $$\Fil_G\in \Omega (K)$$ et $$\zeta_G\in \Hom_\O (F/\O_F,\mathbb{G}\otimes \O_K/p\O_K)\unp \setminus \{ 0 \}$$ Soit $$\zeta_{G*} : K\ldrt \mathbb{D} (\mathbb{G})_{\Q}\otimes K$$ l’application induite au niveau de l’évaluation des cristaux de Messing sur l’épaississement $\O_K\twoheadrightarrow \O_K/p\O_K$. Considérons le composé $$\xymatrix@R=5mm{ K\ar[r]^(.3){\zeta_{G*}} & \mathbb{D}(\mathbb{G})_\Q\otimes K \ar[r]^{\Delta_*^{-1}} & \mathbb{D} (\mathbb{H})_\Q^n\otimes K \ar@{=}[r] & \bigoplus_{j\in \Z/n\Z} \mathbb{D}(\mathbb{H})_{\Q,j}^n\otimes K \\ 1 \ar@{|->}[rrr] &&& (a_{ij})_{1\leq i\leq n,j\in \Z/n\Z} }$$ et posons $x_{ij}=\Pi^{-j}a_{ij}\in \mathbb{D}(\mathbb{H})_{\Q,0}\otimes K\simeq K$. Soit $$X=(x_{ij})_{i,j}\in\text{M}_n (K)$$ \[enkgf\] L’application $K$-linéaire de $K^n$ dans lui même induite par $X$ a pour image $\Fil_G$. D’après le théorème \[decHTO\] et la remarque \[HTet\] le sous-module engendré par l’image de $\zeta_{G*}$ dans $\mathbb{D}(\mathbb{G})_\Q\otimes K$ est $\bigoplus_{j\in \Z/n\Z} \Pi^j \Fil_G$. Il en résulte aussitôt que $$K(x_{i 0})_i+\dots + K (x_{in})_i =Fil_G \subset K^n=\mathbb{D}(\mathbb{G})_{\Q,0}\otimes K$$ On note $\Fil_H\in \mathbb{P}^n (K)$ l’image de $\,^t X$ dans $K^n$ i.e. le sous-espace engendré par les lignes de $X$. Via l’identification $\mathbb{D}(\mathbb{H})_\Q\otimes K \simeq K^n$ si $$\xymatrix@R=5mm{ K \ar[r]^(.3){\zeta_{G*}} & \mathbb{D}(\mathbb{G})_\Q\otimes K \ar[r]^{\sim} & \mathbb{D}(\mathbb{H})_\Q^n\otimes K \\ 1 \ar@{|->}[rr] && (\a_i)_i }$$ alors $\Fil_H \in \mathbb{P} ( \mathbb{D} (\Hb))(K)$ est l’image de l’application $K^n\ldrt \mathbb{D}(\mathbb{H})_\Q\otimes K$ définie par $(\a_i)_i$.\ La filtration $\Fil_H$ définit donc d’après la proposition \[four\_D\] un élément $\M^{\LT} (K)/\sim$ dont il reste à définir la rigidification du module de Tate (cf. théorème \[derhj\]). Considérons le composé $$F/\O_F \xrig{\;\; \zeta_G \;\;} \mathbb{G}\otimes \O_K/p\O_K \xrig{\;\;\Delta^{-1}\;\;} \mathbb{H}^n\otimes \O_K/p\O_K$$ qui fournit des éléments $$(\zeta_{H,i})_{1\leq i \leq n} \in \Hom_\O (F/\O_F, \mathbb{H}\otimes \O_K/p\O_K)\unp$$ Pour un entier $i\in \{1,\dots,n\}$ le morphisme induit entre cristaux de Messing évalués sur $\O_K\twoheadrightarrow \O_K/p\O_K$ est $$\xymatrix@R=5mm{ (\zeta_{H,i})_* : & K \ar[r]^(.3){\zeta_{G*}} & \mathbb{D} (\mathbb{G})_\Q\otimes K \ar[r]^{\Delta_*^{-1}} & \mathbb{D} (\mathbb{H})_\Q^n\otimes K \ar[r]^{\;\text{proj}_i\;} & \mathbb{D}(\mathbb{H})_\Q\otimes K \\ & 1 \ar@{|->}[rrr] &&& \a_i }$$ et donc $\forall i\; \text{Im}( (\zeta_{H,i})_*)\subset \Fil_H$.\ Il reste à voir que les $(\zeta_{H,i})$ sont $F$-linéairement indépendants. Mais si $(\l_i)_i\in F^n$ est tel que $\sum_i \l_i \zeta_{H,i}=0$ alors $(\l_i)_i$ définit une forme linéaire sur $F^n$. Et l’égalité $\sum_i \l_i \zeta_{H,i}$ implique sur l’évaluation des cristaux l’égalité $\sum_i \l_i (a_{ij})_j = 0$ et donc $\sum_i \l_i (x_{ij})_j = 0$ ce qui implique d’après le lemme \[enkgf\] que la forme linéaire associée à $(\l_i)_i$ s’annule sur $\Fil_G$. Si $(\l_i)_i$ est non-nul alors $\Fil_G$ est égal au noyau de la forme linéaire associée, est donc défini sur $F$ et contient donc à fortiori une droite $F$-rationnelle. Donc, puisque $\Fil_G\in \Omega (K)$, $\forall i\; \l_i= 0$.\ D’après le théorème \[derhj\] on en déduit un triplet $(H,\rho_H,\eta_H)\in \M_\infty^{\LT} (K)$. L’application $\M_\infty^{\LT}(K)\ldrt \M_\infty^{\D} (K)$ ---------------------------------------------------------- Soit $(H,\rho_H,\eta_H)\in\M_\infty^{\LT} (K)$. Rappelons (cf. théorème \[derhj\]) qu’on lui associe un couple $(\Fil_H,\zeta_H)$ où $$\Fil_H\in \mathbb{P}^n (K)$$ et $$\zeta_H=(\zeta_{H,i})_i : F^n\hookrightarrow \Hom_\O (F/\O_F,\mathbb{H}\otimes \O_K/p\O_K)\unp$$ Soit $$\xymatrix@R=4mm@C=12mm{ K \ar[rr]^(.37){(\zeta_{H,1*},\dots,\zeta_{H,n*})} && \mathbb{D}(\mathbb{H})_\Q^n\otimes K \ar@{=}[r] & \bigoplus_{j\in \Z/n\Z} \mathbb{D} (\mathbb{H})_{\Q,j}^n\otimes K \\ 1 \ar@{|->}[rrr] &&& (a_{ij})_{i,j} }$$ l’application induite sur l´évaluation des cristaux sur l’épaississement $\O_K\twoheadrightarrow \O_K/p\O_K$. Notons $x_{ij}=\Pi^{-j} a_{ij}\in\mathbb{D}(\mathbb{H})_{\Q,0}\otimes K \simeq K$. Alors, $$X=(x_{ij})_{i,j}\in \text{M}_n (K)$$ L’endomorphisme $K$-linéaire de $K^n$ induit par $\,^tX$ a pour image $\Fil_H$. C’est une conséquence du théorème \[decHTO\]. On note $\Fil_G=\text{Im} (X)\in \mathbb{P}^n (K) = \mathbb{P}(\mathbb{D}(\mathbb{G})_{\Q,0})(K)$ via $\mathbb{D}(\mathbb{H})_{\Q,0}^n\simeq \mathbb{D}(\mathbb{G})_{\Q,0}$ induit par $\Delta$. $\Fil_G\in \Omega (K) \subset \mathbb{P}\left ( \mathbb{D}(\mathbb{G})_{\Q,0}^{V=\Pi}\right )(K)$ On a $$\Fil_G = K(x_{i1})_i+\dots + K ( x_{in})_i$$ Soit $(\mu_j)_j\in K^n$ tel que $$\sum_j \mu_j (x_{ij})_i\in F^n$$ Soit alors la forme $F$-linéaire à valeurs dans $F$ définie sur $V_p(H)$ par $$\ph : \zeta_{H,i}\longmapsto \sum_j \mu_j x_{ij}\in F$$ Elle définit un élément de $V_p (H)^*$. D’après le corollaire \[DansOmega\] dans la suite de Hodge-Tate de $H^D$ $$0\drt \omega_{H^D}^*\otimes K \drt V_p (H)^*\otimes_{F} K \drt \omega_H\otimes K\drt 0$$ l’application $V_p(H)^*\ldrt \omega_H\otimes K$ est injective. Or $$\left [ V_p (H)^*\otimes_F K\twoheadrightarrow \omega_H\otimes K\right ] =\left [ \ker \a \hookrightarrow V_p(H)\otimes_F K \right ]^*$$ où $\a : V_p (H) \otimes_F K \ldrt \mathbb{D}(\mathbb{H})_\Q\otimes K$ est l’application de matrice $\,^t(a_{ij})_{i,j}$. Mais $\ph_{|\ker \a}=0$ par définition. Donc $\ph=0$ et $\sum_j \mu_j (x_{ij})_i=0$. On obtient donc d’après la proposition \[DrP\] un couple $(G,\rho_G)\in \M^{\D}(K)/\sim$. Reste à définir une rigidification de $G$. Considérons le morphisme composé $$\zeta_G : F/\O_F \xrig{\;(\zeta_{H,1},\dots,\zeta_{H,n})\;} \mathbb{H}^n\otimes \O_K/p\O_K \xrig{\;\Delta\;}\mathbb{G}\otimes\O_K/p\O_K$$ On vérifie aussitôt que par définition de $\Fil_G$ le morphisme induit au niveau de l’évaluation des cristaux a son image contenue dans $\Fil_G$. D’après le théorème \[descDri\] on obtient donc un élément $(G,\rho_G,\eta_G)$. Les deux applications sont inverses l’une de l’autre ---------------------------------------------------- Cela ne pose pas de problème. Retraçage des actions --------------------- Dans la bijection entre $\M^{\LT}_\infty (K)$ et $\M_\infty^\D(K)$ si $$(H,\rho_H,\eta_H) \longmapsto (G,\rho_G,\eta_G)$$ et $(g,d)\in \GL_n (F)\times D^\times$ alors $$(g,d).(H,\rho_H,\eta_H) \longmapsto (\,^t g,d^{-1}) .(G,\rho_G,\eta_G)$$ Cela ne pose aucun problème. Bijection entre les points de l’espace de Berkovich associé ----------------------------------------------------------- Soit $K\hookrightarrow K'$ une extension isométrique de corps valués complets pour une valuation de rang $1$. On vérifie alors aussitôt que le diagramme suivant est commutatif $$\xymatrix{ \M_\infty^{\LT} (K) \ar[d] \ar[r]^\sim & \M_\infty^{\D} (K) \ar[d] \\ \M_\infty^{\LT} (K') \ar[r]^\sim & \M_\infty^{\D} (K') }$$ Pour $*\in \{ \LT,\D\}$ posons $$\left |\M^*_\infty \right |= \coprod_K \M_\infty^* (K) /\sim$$ où pour $x\in \M^*_\infty (K_1)$ et $y\in \M^*_\infty (K_2)$ $\; x\sim y$ ssi il existe une extension valuée comme précédemment $$\xymatrix@R=2mm{ K_1 \ar@{^(->}[rd] \\ & K_3 \\ K_2 \ar@{^(->}[ru] }$$ telle que $x$ et $y$ aient même image dans $\M^*_\infty (K_3)$. Cette classe d’équivalence est bien définie au sens où c’est un ensemble. Cela résulte de l’existence de modèles entiers de nos espaces qui implique que l’on peut se limiter à des corps $K$ de cardinalité bornée. On a donc une bijection $\GL_n (F)\times D^\times$-équivariante $$\left | \M^{\LT}_\infty \right | \iso \left | \M^{\D}_\infty \right |$$ Le choix de modèles entiers de nos espaces permet de munir ces ensembles d’une structure d’espace topologique localement compact et l’existence de l’isomorphisme au niveau de ces modèles entiers ([@iso4]) impliquera que la bijection précédente est un homéomorphisme. La matrice $X$ {#kfyygpzr5rrppn} ============== Les triplets $(H,\rho_H,\eta_H)$ et $(G,\rho_G,\eta_G)$ qui se correspondent ont en commun la matrice de rang $n-1$ $\;X\in\text{M}_n (K)$ où comme précédemment on identifie $$\DH_\Q^n\otimes K \simeq \bigoplus_{j\in \Z/n\Z} \DH_{\Q,j}^n\otimes K \underset{\sim}{\xrig{\oplus_j \Pi^{-j}}} \bigoplus_{j\in \Z/n\Z} \DH_{\Q,0}^n\otimes K \simeq M_n (K)$$ Celle-ci vérifie - Les lignes de $X$ engendrent $\Fil_H\in \mathbb{P}^n (K)$ - Les colonnes de $X$ engendrent $\Fil_G\in\Omega (K)$ Supposons maintenant pour simplifier que $F=\Qp$. Alors, $X$ possède un relèvement $$X_{cris}\in \text{M}_n ( B^+_{cris}(\O_K))$$ i.e. via $\theta : B^+_{cris} (\O_K)\twoheadrightarrow \O_{\widehat{K}}$ on a $\theta (X_{cris})=X$. Ce relèvement est défini de manière similaire à $X$ en évaluant les cristaux sur $A_{cris}(\O_K)\twoheadrightarrow \O_{\widehat{\overline{K}}}/p \O_{\widehat{\overline{K}}}$ au lieu de $\O_K\twoheadrightarrow \O_K/p\O_K$ (cf. les propositions \[docyum1\] et \[docyum2\] de la section \[ruhgp\] pour plus de détails). De plus, $$\det (X_{cris}) \in \Qp^\times .t$$ où la valuation de l’élément de $\Qp^\times$ est liée aux hauteurs de $\rho_H,\rho_G$ et $\Delta$ (cf. la section suivante). Si $$\Phi=\left ( \begin{matrix} 0 & p & & \\ 0 & 0& \ddots & \\ \vdots & &\ddots & p \\ 1 &0 & & 0 \end{matrix}\right )$$ est la matrice du Frobenius de $\mathbb{D}(\mathbb{H})_\Q$ relativement à la base définie par l’isomorphisme $\mathbb{D}(\mathbb{H})_{\Q,0}\simeq L$ et $\DH_\Q\simeq \oplus_j \DH_{\Q,j} \underset{\sim}{\xrig{\sum_j \Pi^{-j}}} \oplus_j \DH_{\Q,0}$ $$\begin{aligned} \label{eqslf} \boxed{ \ph (X_{cris}) \,^{t}\Phi = p X_{cris} }\end{aligned}$$ où $\ph (X_{cris})$ est obtenue en appliquant le Frobenius cristallin à tous les éléments de la matrice $X_{cris}$. On peut montrer la proposition suivante : Les espaces $\M_\infty^{\LT} (K)$ et $\M_\infty^{\D}(K)$ s’envoient surjectivement sur l’ensemble des matrices $X\in\text{M}_n (K)$ de rang $n-1$ possédant un relèvement $X_{cris}$ à $B^+_{cris}$ de déterminant non-nul et vérifiant l’équation (\[eqslf\]). De plus les fibres de cette application en $X$ sont en bijection avec l’ensemble des relèvements de $X$ de déterminant non-nul satisfaisant à l’équation (\[eqslf\]). Cette ensemble de matrices est muni d’une action de $\GL_n (F)\times D^\times$ de la façon suivante : l’isomorphisme $\mathbb{D}(\mathbb{H})_\Q\simeq L^n$ induit un plongement $D^\times \hookrightarrow \GL_n (L)$. Alors $$\forall (g,d)\in \GL_n (F)\times D^\times \;\;\forall X \;\;\;\;\;\; (g,d).X= \,^tg X d^{-1}$$ Les applications précédentes sont compatibles à cette action. $$\xymatrix@C=6mm@R=22mm{ \M_\infty^{\D} (K) \ar[rrd]\ar[dd]_{Periodes} \ar[rrrr]^{\sim} &&&& \M_\infty^{\LT}(K) \ar[lld] \ar[dd]^{Periodes} \\ && \{X\in \text{M}_n (K)\; |\; \text{rg} X =n-1\; \} \ar[lld]^{X\mapsto \text{Im} \,^t X} \ar[rd]^{X\mapsto \text{Im} X} \\ \mathbb{P}^n (K) && & \mathbb{P}^n (K) & \Omega(K) \ar@^{_(->}[l] }$$ Il s’agit d’une application du théorème \[deall\] énoncé dans l’appendice. Dans la proposition précédente on peut enlever “de déterminant non-nul” si l’on se restreint aux $X$ telles que $\text{Im}\, X \in \Omega (K)$. Pour les corps locaux de caractéristique positive A.Genestier et V.Lafforgue donnent une interprétation géométrique comme points d’un certain schéma formel de l’analogue en caractéristique positive de l’ensemble des matrices $X_{cris}$ solutions de l’équation précédente. Il n’existe pas de telle interprétation en caractéristique zéro. L’isomorphisme conserve le degré {#ruhgp} ================================= Il s’agit de démontrer que pour un choix convenable de $\Delta$ le diagramme suivant commute $$\xymatrix{ \M_\infty^\D (K) \ar[rd] \ar[rr]^\sim && \M_\infty^{\LT} (K) \ar[ld] \\ & \Z }$$ où les applications vers $\Z$ sont les applications hauteur renormalisées des quasi-isogénies universelles. Cela implique en particulier que les tours de Lubin-Tate et Drinfeld classiques, les fibres en hauteur zéro, sont isomorphes. Plus précisément, si $f$ est une quasi-isogénie entre $\O$-modules formels on note $\text{ht}_\O (h) = \text{ht} (f) /[F:\Qp]$. Alors l’application $\M_\infty^\D (K) \ldrt \Z$ est $$(G,\rho_G,\eta_G)\longmapsto \text{ht}_\O (\rho_G) /n$$ et l’application $\M_\infty^{\LT} (K)\ldrt \Z$ est $$(H,\rho_H,\eta_H)\longmapsto \text{ht}_\O (\rho_H)$$ On va appliquer le théorème \[det\_periodes\_unite\] de l’appendice. Afin de simplifier on se limite au cas $F=\Qp$, le cas général étant laissé au lecteur.\ Supposons que l’isomorphisme $\DH_{\Q,0} \simeq L$ fixé au début de la section \[coeulop\] provienne d’un isomorphisme $\DH_0\simeq \O_L$. \[docyum1\] Soit $(H,\rho_H,\eta_H)\in \M_\infty^{\LT} (K)$. Soit $\xi : \Qp/\Zp \ldrt \Hb^n\otimes \O_K/p\O_K$ associé à $\eta_H$, $\rho_H$ et la base canonique de $\Zp^n$. Soit $$\xi_* : B^+_{cris} \ldrt \DH_\Q^n\otimes_L B^+_{cris}$$ le morphisme associé sur l’évaluation des cristaux. Soit $X_{cris} \in \M_n \left ( B^+_{cris} (\O_K)\right ) $ la matrice associée via $$\DH_\Q^n\otimes_L B^+_{cris} = \bigoplus_{j\in\Z/n\Z} \DH_{\Q,j}^n \otimes_L B^+_{cris} \underset{\sim}{\xrig{\; \oplus_j \Pi^{-j}\; }} \bigoplus_{j\in \Z/n\Z} \DH_{\Q,0}^n\otimes_L B^+_{cris}$$ Alors $\det \left ( X_{cris}\right ) = \l .t \in \Qp^\times .t$ et de plus $$v_p (\l) = -\text{ht} (\rho_H) - \frac{1}{2} n (n-1)$$ Soit $M_0 \subset \DH_\Q$ le module de Dieudonné de $H\otimes_{\O_K} k$ $$M_0 = \mathbb{D} (\rho_H^{-1}) \left ( \mathbb{D} (H_k)\right )$$ D’après le théorème \[det\_periodes\_unite\] de l’appendice, dans l’isomorphisme $$T_p (H) \otimes_{\Zp} B^+_{cris} \simeq M_0 \otimes_{\O_L} B^+_{cris}$$ on a $$\bigwedge^n T_p (H)\otimes \O_L = t. \bigwedge^n M_0$$ De plus, $[M_0 : \DH ] = \text{ht} (\rho_H)$. Via l’isomorphisme $\DH_\Q \simeq \oplus_{j\in \Z/n\Z} \DH_{\Q,0}$ le réseau $\DH$ s’envoit sur $$\bigoplus_{j\in\Z/n\Z} \Pi^{-j} \DH_0$$ Soit $W$ l’image de $T_p (H)$ dans $\oplus_j \DH_0 \otimes_{\O_L} B^+_{cris}$. Alors $$\bigwedge^n W\otimes \O_L = p^{-\text{ht} (\rho_H)-\frac{1}{2} n (n-1)} t. \DH_0$$ D’où le résultat. \[docyum2\] Soit $(G,\rho_G,\eta_G)\in\M_\infty^\D (K)$. Soit $\zeta= \rho_G^{-1}\circ \eta_G (1) : \Qp/\Zp \ldrt \Gb\otimes \O_K/p\O_K$ et $\xi =\Delta^{-1} \circ \zeta : \Qp/\Zp \ldrt \Hb^n\otimes \O_K/p\O_K$. Soit $X_{cris}$ la matrice associée comme dans la proposition précédente. Alors $\det ( X_{cris})= \l . t \in \Qp^\times t$ et $$v_p (\l) = -\text{ht} (\rho_G)/n - \text{ht} (\Delta)$$ La démonstration est identique à celle de la proposition précédente. On applique le théorème \[det\_periodes\_unite\] de l’appendice. La matrice des périodes de $G$ est $n$-copies de la matrice $X_{cris}$. On en déduit que $\det (X_{cris})^n$ est donné par la valeur annoncée. Si $\text{ht}_\O (\Delta) = \frac{1}{2} n (n-1)$ alors la bijection entre $\M_\infty^{\LT} (K)$ et $\M_\infty^\D (K)$ respecte les hauteurs normalisées. Soit $(H,\rho_H,\eta_H)\in \M_\infty (K)$ tel que $\text{ht} (\rho_H)=0$. Alors, le couple $(G,\rho_G)$ associé dans l’espace de Rapoport-Zink sas niveau est l’élément de $\Omega (K) \subset \M^{\D} (K)$ donné par le dual de la décomposition de Hodge-Tate de $H$ couplé à $\eta_H$. Plus généralement, soit $(G,\rho_G)$ le point de $\Omega (K)$ associé à la décomposition de Hodge-Tate de $H$ et à $\eta_H$. Le point de $\M^{\D} (K)$ associé par l’isomorphisme de Faltings est alors $$(G,\rho_G\circ \Pi^{\text{ht} (\rho_H)})$$ Un point de vue différent sur la bijection {#ptovue} ========================================== Nous allons redémontrer la bijection précédente d’un point de vue “dual”. En particulier, au lieu de considérer des filtrations $\Fil_H\in \mathbb{P}^n, \Fil_G\in \Omega$ nous considérerons plutôt les quotients $K^n\twoheadrightarrow K^n/\Fil_H, K^n\twoheadrightarrow K^n/\Fil_G$. Ce point de vue se prête mieux lorsque l’on travaillera sur une base quelconque (c’est à dire plus sur un point comme dans cet article) puisque que pour un fibré vectoriel $\mathcal{E}$ il est plus commode de définir $\mathbb{P} (\mathcal{E})$ comme classifiant les quotients localement libres de rang $1$ $\mathcal{E}\twoheadrightarrow \mathcal{L}$ (en tous cas cela est plus facile à platifier par éclatements). En effet, dans le cas d’une base quelconque l’approche des sections précédentes nous conduirait à définir $\Fil_H$ comme sous-module engendré par certaines sections ce qui est moins commode. De plus la matrice définissant ces sous-modules serait une section de $\DH_\Q^n\otimes \O_\X \unp\simeq M_n ( \O_\X \unp )$ où $\X$ est l’espace de Lubin-Tate ou de Drinfeld en niveau infini. Il faudrait donc procéder à de nouveaux éclatements afin de rendre l’espace engendré par les lignes et celui par les colonnes localement facteur direct entier (cette dernière justification est quelque peu hypocrite puisque de toutes façons dans la démonstration finale on devra à un endroit rendre entier l’application des périodes). Le point de vue qui suit permet de construire directement une application entière du sommet de la tour de Lubin-Tate (resp. la tour de Drinfeld) vers le schéma formel de Drinfeld (resp. $\widehat{\P}^{n-1}$). Commençons par faire le lien entre le point de vue précédent et celui qui va suivre. Identification de $K^n\twoheadrightarrow K^n/\Fil_H$ avec l’application de Hodge-Tate de $G^D$ ---------------------------------------------------------------------------------------------- Soit $(G,\rho_G,\eta_G)\in \M^{\D}_\infty (K)$. Rappelons qu’on a défini une matrice $X$ telle que $\text{Im} X =\Fil_G$ et $\text{Im}\,^t X =\Fil_H$. Tout repose sur une bidualisation, l’exactitude de la suite de Hodge-Tate ainsi que sur un analogue de l’énoncé d’algèbre linéaire suivant : soit $u:E_1\drt E_2$ une application linéaire entre deux $K$-espaces vectoriels de dimension finie, il y a alors une identification canonique $$\left [ E_1^*\twoheadrightarrow E_1^*/\text{Im} \,^t u\right ] \simeq \left [ \ker u \hookrightarrow E_1 \right ]^*$$ L’analogue est le suivant : pour $E$ un $D\otimes_F K$-module de type fini (à gauche ou à droite) posons $$\Gamma (E) =\Hom_{D\otimes_F L}(E, \mathbb{D}(\mathbb{H})_{\Q})$$ un $K$-espace vectoriel de dimension finie. Alors, $$\begin{aligned} \label{ergjk} \forall u : E_1\drt E_2\;\;\; \left [ \Gamma (E_1)\twoheadrightarrow \Gamma (E_1)/ \text{Im} \,^\Gamma u \right ] \simeq \Gamma \left [\ker u \hookrightarrow E_1\right ] \end{aligned}$$ Notons maintenant pour $W$ un $K$-e.v. $\Phi (W) =W^*$ le dual usuel. Si $E$ est un $D\otimes_L K$-module de type fini à gauche, resp. à droite, alors $\Phi (E)$ est naturellement un $D\otimes_L K$-module de type fini à droite, resp. à gauche. Il y a alors un isomorphisme naturel de bidualité pour $E$ comme ci-dessus $$\GG\circ \Phi (E) \simeq \mathbb{D} (\mathbb{H})_{\Q,0} \otimes_L E_0$$ où comme d’habitude $E_0 = \{x\in E\; |\; \forall a\in F_n\; a\otimes 1. x = 1\otimes a.x \;\}$. Soit l’application $D\otimes_F K$-linéaire composée $$\xymatrix@C=14mm{ u : D\otimes_F K \ar[r]^{\eta_G}_{\sim} & V_p (G)\otimes_F K \ar[r]^{\a_G} & \mathbb{D}(\mathbb{G})_\Q\otimes_L K }$$ Procédons aux identifications suivantes : $$\begin{aligned} \Gamma (D\otimes_F K) & \iso & \mathbb{D}(\mathbb{H})_\Q\otimes_L K \\ f &\longmapsto & f (1\otimes 1)\end{aligned}$$ et $$\begin{aligned} \Gamma (\mathbb{D} (\mathbb{G})_\Q\otimes_L K) &\iso & \Hom_L ( \mathbb{D}(\mathbb{G})_{\Q,0},\mathbb{D}(\mathbb{H})_{\Q,0})\otimes_L K \simeq K^n \\ f &\longmapsto & f_{|\mathbb{D}(\mathbb{G})_{\Q,0}\otimes K}\end{aligned}$$ via l’isomorphisme (\[out1\]) et $\Delta$. Alors $$\,^\Gamma u : K^n\ldrt \mathbb{D} ( \mathbb{H})_\Q\otimes K \;\;\;\;(\simeq K^n)$$ s’identifie à $\,^t X$ et donc $\text{Im} (\,^\Gamma u) =\Fil_H$. Il suffit de retracer les différentes identifications. Étant donné que le module de Tate de $G$ est trivialisé et que $\det V_p (G)\simeq \Qp (1)$ on en déduit que $\Qp (1)$ est trivialisé et que donc l’application de Hodge-Tate de $G^D$ est définie sur $K$ : $\a_{G^D} : V_p (G^D) \ldrt \omega_G\otimes K$. En d’autres termes $\Qp (\mu_{p^{\infty}})\subset K$. Il y a des identifications canoniques $$\begin{aligned} \left [ \mathbb{D}(\mathbb{H})_\Q\otimes K \twoheadrightarrow \mathbb{D}(\mathbb{H})_\Q\otimes K/\Fil_H \right ] &\simeq & \left [ (V_p (G^D)\otimes_F K)_0 \overset{\a_{G^D}}{\twoheadrightarrow} \omega_{G,0}\otimes K \right ] (-1)\otimes_L \mathbb{D} (\mathbb{H})_{\Q,0} \\ &\simeq & \left [\mathbb{D}(\mathbb{H})_\Q\otimes K\twoheadrightarrow \omega_{G,0}(-1)\otimes_L \mathbb{D}(\mathbb{H})_{\Q,0}\right ]\end{aligned}$$ où $\a_{G^D} $ est l’application de Hodge-Tate de $G^D$. Soit $$\left [ E_1 \xrig{\;\a_G\;} E_2 \right ] = \left [ V_p (G)\otimes_F K \xrig{\;\a_G\;} \mathbb{D} (\mathbb{G})_\Q\otimes_L K \right ]$$ D’après le lemme précédent $$\left [ \mathbb{D} (\mathbb{H})_\Q\otimes K \twoheadrightarrow \mathbb{D} (\mathbb{H})_\Q\otimes K/\Fil_H \right ] \simeq \left [ \GG (E_1) \overset{\,^\GG \a_G}{\twoheadrightarrow} \GG (E_1)/\text{Im} (\,^\GG \a_G ) \right ]$$ qui d’après l’identification (\[ergjk\]) est isomorphe à $$\GG \left [ \ker \a_G \hookrightarrow E_1 \right ]$$ Mais d’après la proposition \[relhtO\] (décomposition de Hodge-Tate) la suite $$0 \ldrt \omega_G^*\otimes K(-1) \xrig{\;\,^\Phi\a_{G^D} (-1)\;} V_p (G)\otimes_F K \xrig{\; \a_G\; } \mathbb{D} (\mathbb{G} )_\Q\otimes_L K$$ est exacte. Donc, $$\begin{aligned} \simeq \GG \left [ \ker \a_G \hookrightarrow E_1 \right ]& \simeq& \GG\circ \Phi \left [ V_p (G^D)\otimes_F K \twoheadrightarrow \omega_G\otimes K \right ](-1) \\ &\simeq & \left[ \left ( V_p (G^D)\otimes_F K \right )_0 (-1) \twoheadrightarrow \omega_{G,0}\otimes K(-1) \right ]\otimes_L\mathbb{D} (\mathbb{H})_{\Q,0}\end{aligned}$$ Le dernier isomorphisme dans l’énoncé du corollaire provient de la rigidification $\eta : D\iso V_p (G)$ qui induit $$\GG\circ \Phi \left ( V_p (G)\otimes_F K\right ) \iso \mathbb{D} (\mathbb{H})_{\Q}\otimes K$$ Identification de $K^n\twoheadrightarrow K^n/\Fil_G$ avec l’application de Hodge-Tate de $H^D$ ---------------------------------------------------------------------------------------------- Soit $(H,\rho_H,\eta_H)\in \M_\infty^{\LT} (K)$. On procède comme dans la section précédente. Pour $E$ un $K$-espace vectoriel posons $$\Psi (E) = \Hom_L (E,\DH_{\Q,0} )$$ Soit l’application composée $$\xymatrix{ v:K^n \ar[r]^(.4){\eta_H}_(.4){\sim} & V_p (H)\otimes_F K \ar[r] & \DH_\Q\otimes_L K }$$ Procédons aux identifications suivantes : $$\begin{aligned} \Psi (K^n)&\simeq& \DH_{\Q,0}^n\otimes K \simeq \DG_{\Q,0}\otimes K\text{ via } \Delta \\ \Psi ( \DH_\Q\otimes_L K) &\simeq & \Hom ( \DH_\Q,\DH_{\Q,0})\otimes K \simeq K^n\end{aligned}$$ où la second identification utilise $$\DH_\Q = \bigoplus_{j\in\Z/n\Z} \DH_{\Q,j} \underset{\sim}{\xrig{\oplus_i \Pi^{-i}}} \DH_{\Q,0}^n\simeq L^n$$ Alors $\,^\Psi v : K^n \ldrt \DG_{\Q,0} \otimes K$ s’identifie à la matrice $X$ et donc $\text{Im} (\,^\Psi v)=\Fil_G$. Il y a des identifications $$\begin{aligned} \left [ \DG_{\Q,0}\otimes K \twoheadrightarrow \DG_{\Q,0}\otimes K / \Fil_G \right ] & \simeq & \left [ V_p (H^D) \xrig{\;\a_{H^D}\;}\omega_H\otimes K\right ](-1) \otimes_L \DH_{\Q,0} \\ & \simeq & \left [ \DG_{\Q,0}\otimes K \twoheadrightarrow \omega_H\otimes K(-1) \otimes_L \DH_{\Q,0} \right ]\end{aligned}$$ Soit $$\left [ E_1\xrig{\;\a_{H}\;} E_2\right ] = \left [ V_p (H)\otimes_F K \xrig{\;\a_H\;} \DH\otimes_L K \right ]$$ Alors $$\begin{aligned} \left [ \DG_{\Q,0}\otimes K \twoheadrightarrow \DG_{\Q,0}\otimes K / \Fil_G\right ] & \simeq & \left [ \Psi (E_1) \twoheadrightarrow \Psi ( E_2) \right ] \\ & \simeq & \Psi \left [ \ker \a_H \hookrightarrow E_1 \right ] \\ & \simeq & \Psi\circ \Phi \left [ V_p (H^D)\otimes_F K(-1) \xrig{\;\a_{H^D} (-1)\;} \omega_H\otimes K(-1) \right ] \\ & \simeq & \left [ V_p (H^D)\otimes_F K \overset{\a_{H^D}}{\twoheadrightarrow} \omega_H\otimes K \right ](-1) \otimes_L \DH_{\Q,0} \end{aligned}$$ où on a utilisé l’exactitude de la suite de Hodge-Tate pour $H$ (proposition \[relhtO\]). L’application $\M^\D_\infty (K)\ldrt \M^{\LT}_\infty (K)$ {#dltisog2} --------------------------------------------------------- Soit $(G,\rho_G,\eta_G)\in \M_\infty^\D (K)$. ### Première étape : on met une structure d’isocristal sur le module de Tate Posons $$N=\Hom_D ( V_p (G), \mathbb{D} (\Hb)_\Q )$$ un isocristal relativement à l’extension $L|F$ si on le munit de $\ph : N\iso N$ défini par $\ph.f=\ph \circ f$. ### Deuxième étape : la rigidification du module de Tate induit une rigidification de l’isocristal L’isomorphisme de $D$-modules $\eta_G : D\iso V_p (G)$ induit un isomorphisme d’isocristaux $$\mathbb{D} (\Hb)_\Q\iso ( N,\ph)$$ ### Troisième étape : la filtration de Hodge-Tate induit une filtration du module de Dieudonné Considérons l’application de Hodge-Tate de $G^D$ tordue par $K(-1)$ $$(\mathcal{S}) = \left [ V_p (G)^*\otimes_F K \overset{\a_{G^D}(-1)}{\twoheadrightarrow} \omega_G\otimes K(-1) \right ] = \bigoplus_{j\in \Z/n\Z} \left [ \left ( V_p (G)^*\otimes_F L\right )_j\otimes_L K \twoheadrightarrow \omega_{G,j}\otimes K(-1) \right ]$$ Tensorisons cette suite $$(\mathcal{S})\otimes_{D\otimes_F L} \DH_\Q \simeq \left [ N\otimes_F L\twoheadrightarrow \omega_{G,0}(-1)\otimes_L \DH_{\Q,0} \right ]$$ qui fournit donc via la deuxième étape un élément $$\Fil_H\in \mathbb{P} \left ( \DH_\Q \right ) (K)$$ On obtient ainsi un couple $[(H,\rho_H)]\in \M^{\LT} (K)/\sim$. ### Quatrième étape : construction d’éléments dans le module de Tate de $H$ Construisons un morphisme $$\Psi : \Hom_ {\O_D} (\Gb, \Hb)\unp \ldrt V_p (H)$$ Soit $$\zeta_G \in \Hom (F/\O_F, \Gb\otimes \O_K/p\O_K)\unp$$ comme dans le théorème \[descDri\]. Soit $f\in \Hom_{\O_D} (\Gb,\Hb)\unp$. Considérons $$f\circ \zeta_G \in \Hom (F/\O_F, \Hb \otimes \O_K/p\O_K)\unp$$ Pour le morphisme induit sur les cristaux évalués sur l’épaississement $\O_K\twoheadrightarrow \O_K/p\O_K$ $$(f\circ \zeta_G)_* : K \ldrt \DH_\Q\otimes K$$ montrons que $(f\circ \zeta_G)_* (1) \in \Fil_H$.\ Via $\eta_G : D\iso V_p (G)$ le morphisme $f_* : \DG_\Q\otimes K \ldrt \DH_\Q\otimes K$ s’identifie à $$\left ( V_p (G)^*\otimes_F K \right )\otimes_{D\otimes L} \left [ \DG_\Q \xrig{\; \mathbb{D} (f)\;} \DH_\Q \right ]$$ On doit montrer que $\zeta_{G*} (1) \in \left ( V_p (G)^* \otimes K\right )\otimes_{D\otimes L} \DG_\Q$ s’envoit sur zéro par la composée donnée par la ligne pointillées diagonale dans le diagramme suivant $$\xymatrix@C=4mm@R=6mm{\left ( V_p(G)^*\otimes K \right )\ar[dd]^{\a_{G^D}(-1)} & \otimes_{D\otimes L} [ \DG_\Q \ar[r]^(0.6){\mathbb{D}(f)} \ar@<3ex>[dd]^{Id} \ar@{-->}@<1.5ex>[rdd] & \DH_\Q\ar[dd]^{Id} ] \\ & & \\ \omega_G\otimes K(-1) & \otimes_{D\otimes L} [ \DG_\Q \ar[r]^{\mathbb{D}(f)} & \DH_\Q ] }$$ Il résulte de ce diagramme qu’il suffit de vérifier que l’image de $\zeta_{G*} (1)$ par l’application $$\a_{G^D} (-1) \otimes Id : \left ( V_p(G)^*\otimes K \right ) \otimes_{D\otimes L} \DG_\Q \ldrt \omega_G\otimes K(-1) \otimes_{D\otimes L} \DG_\Q$$ est nulle. Mais si $\iota : \widetilde{\omega}_{G^D} \otimes K \hookrightarrow \DG_\Q\otimes K$ et $\,^t \a_{G} : \widetilde{\omega}^*_{G^D}\otimes K \ldrt V_p (G)^*\otimes K$ est l’application de gauche dans la suite de Hodge Tate de $G^D$ tordue par $K(-1)$, $\,^t \a_{G} \in V_p (G)^*\otimes K \otimes \omega_{G^D} $, alors $$\zeta_{G*}(1) = (Id \otimes \iota) \left (\,^t\a_G \right )$$ Le résultat se déduit donc du fait que dans la suite de Hodge-Tate de $G^D$ la composée des deux applications est nulle : $\,^t \a_{G} \circ \a_{G^D} (-1) = 0$. On a donc bien défini l’application $\Psi$. ### Sixième étape : l’application $\psi$ est un isomorphisme Il suffit de démontrer qu’elle est injective. Soit donc $f$ tel que $\psi (f)=0$. On vérifie alors sur l’évaluation des cristaux que cela implique que pour $\mathbb{D} (f) : \DG_\Q\ldrt \DH_\Q$ $$\mathbb{D} (f)_{|\widetilde{\omega}_{G^D}\otimes K} =0$$ Mais $$\xymatrix@R=2mm{ \mathbb{D} (f) \in \Hom_{D,\ph} \left ( \mathbb{D} (\Gb)_\Q,\mathbb{D} (\Hb)_\Q\right ) \ar[r]^\sim & \Hom\left (\mathbb{D} (\Gb)_{\Q,0}^{V=\Pi},\DH_{\Q,0}^{V=\Pi}\right ) \\ h \ar@{|->}[r] & h_{|\DG_{\Q,0}^{V= \Pi}} }$$ or $$\left [ \mathbb{D} (\Gb)_{\Q,0}^{V=\Pi}\otimes K \twoheadrightarrow \DG_{\Q,0} \otimes K/ \widetilde{\omega}_{G^D,0}\otimes K \right ]\in \Omega (K) \subset \mathbb{P} \left ( \mathbb{D} (\Gb)_{\Q,0}^{V=\Pi}\right ) (K)$$ Donc, $\mathbb{D} (f)_{|\omega_{G^D}\otimes K}=0 \impl \mathbb{D} (f)=0$ et donc $f=0$.\ La quasi-isogénie $\Delta$ induit un isomorphisme $$F^n \iso \Hom_{\O_D} (\Gb,\Hb)\unp$$ et on obtient donc une rigidification $\eta_H$ de $[(H,\rho_H)]$ ce qui détermine le triplet $(H,\rho_H,\eta_H)$ d’après le théorème \[derhj\]. L’application $\M_\infty^{\LT} (K) \ldrt \M_\infty^\D (K)$ ---------------------------------------------------------- Soit maintenant $(H,\rho_H,\eta_H)\in \M_\infty^{\LT} (K)$. ### Première étape : On met une structure d’isocristal sur le module de Tate Soit $$N=\Hom_F ( V_p (H), \mathbb{D} (\Hb)_\Q)$$ qui est muni d’une structure d’isocristal en posant $\ph.f= \ph\circ f$. Notons que cet isocristal est muni d’une action de $D$ puisque c’est le cas de $\mathbb{D} (\Hb)_\Q$. #### Deuxième étape : $\eta_H$ rigidifie l’isocristal L’isomorphisme $\eta_H:F^n \iso V_p (H)$ couplé à la quasi-isogénie $\Delta$ induit un isomorphisme d’isocristaux munis d’une action de $D$ $$\mathbb{D} ( \Gb)_\Q \simeq \mathbb{D} (\Hb)_{\Q}^n \iso N$$ ### Troisième étape : la filtration de Hodge-Tate induit une du module de Dieudonné Considérons l’application de Hodge-Tate de $H^D$ tordue par $K(-1)$ $$(\mathcal{T})=\left [ V_p (H)^*\otimes K \twoheadrightarrow \omega_H \otimes K(-1) \right ]$$ Après application de $-\otimes_F \DH_{\Q}$ on obtient via la deuxième partie une filtration $D$-invariante $$(\mathcal{T})\otimes_L \DH_\Q \simeq \left [ \DG_{\Q}\otimes K \twoheadrightarrow \omega_H\otimes K(-1) \otimes \DH_\Q \right ]$$ La partie “indice zéro” de cette application est obtenue par $$(\mathcal{T})\otimes_L \DH_{\Q,0} \simeq \left [ \DG_{\Q,0}\otimes K \twoheadrightarrow \omega_H\otimes K(-1) \otimes \DH_{\Q,0} \right ]$$ qui définit $\Fil_G\in \mathbb{P} (\DG_{\Q,0}) (K)$. La quasi-isogénie $\Delta$ couplée à $\eta_H$ induit un isomorphisme $$V_p (H)^*\otimes_F \DH_{\Q,0}^{V^{-1}\Pi} \iso \DG_{\Q,0}^{V^{-1} \Pi}$$ D’après le corollaire \[DansOmega\] $(\mathcal{T}) \in \Omega (K) \subset \mathbb{P} (V_p (H)^*)(K)$. Donc $$\Fil_G \in \Omega (K)\subset \mathbb{P} \left ( \DG_{\Q,0}^{V^{-1}\Pi} \right ) (K)$$ On obtient donc ainsi un couple $[(G,\rho_G)]\in \M_\infty^{\D} (K)/\sim$. ### Quatrième étape : Construction d’éléments dans le module de Tate de $G$ A la rigidification $\eta_H: F^n \iso V_p (H)$ est associée un élément $$\xi \in \Hom (F/\O_F,\Hb^n\otimes \O_K/p\O_K)\unp$$ qui composé avec $\Delta$ fournit $\zeta_G\in \Hom (F/\O_F,\Gb\otimes \O_K/p\O_K)\unp$. Il s’agit de montrer que sur l’évaluation des cristaux $$\zeta_{G*}(1) \in \bigoplus_{j\in \Z/n\Z} \Pi^j \Fil_G\subset \DG_\Q\otimes K$$ Pour cela il suffit de vérifier que $$\xi_* (1) \in \ker \left [ \DH_\Q^n\otimes K \overset{\eta_H}{\simeq} V_p (H)^*\otimes K \otimes \DH_\Q \twoheadrightarrow \omega_H\otimes K(-1) \otimes \DH_\Q \right ]$$ Or cela résulte de ce que $\xi_* (1) \in V_p (H)^*\otimes K\otimes \DH_\Q$ est donné par $$V_p (H) \xrig{\; \a_H \;} \DH_\Q\otimes K$$ et de ce que $\a_{H^D}(-1)\circ \a_H=0$ dans la suite de Hodge-Tate de $H^D$. Donc, $\zeta_G$ définit un élément de $V_p (G)$. #### Cinquième étape : Rigidification Il suffit de montrer que $\zeta_G$ est non-nul mais cela est clair. On obtient donc d’après le théorème \[descDri\] un triplet $(G,\rho_G,\eta_G)\in \M_\infty^\D (K)$. Les deux applications sont inverses l’une de l’autre ---------------------------------------------------- Cela est moins clair que dans la première description de l’isomorphisme.\ Partons de $(G,\rho_G,\eta_G)\in \M_\infty^\D (K)$ et soit $(H,\rho_H,\eta_H)\in \M_\infty (K)$ le triplet associé. Soit $(G',\rho_{G'},\eta_{G'})\in\M_\infty^\D (K)$ le triplet associé à $(H,\rho_H,\eta_H)$. Rappelons que la filtration de $\DG_{\Q}\otimes K$ définissant $(G',\rho_{G'})\in \M^\D (K)/\sim$ s’identifie à $$\left [ V_p (H)^*\otimes K \twoheadrightarrow \omega_H\otimes K(-1) \right ] \otimes_L \DH_\Q$$ via $$V_p (H)^* \otimes \DH_\Q \underset{\text{via }\eta_H}{\xrig{\;\;\sim\;\;}} \DH_\Q^n \xrig{\;\mathbb{D} (\Delta)\;} \DG_\Q$$ Rappelons que l’on a un isomorphisme $$\Hom_{\O_D} ( \Gb,\Hb) \iso V_p (H)$$ et qu’alors l’identification ci-dessus se résume à $$\xymatrix@R=2mm{ \DG_\Q \ar[r]^(.24)\sim & \Hom_F \left ( \Hom_{\O_D} (\Gb,\Hb)\unp, \DH_\Q\right ) & \ar[l]_(.34)\sim V_p (H)^*\otimes \DH_\Q \\ x \ar@{|->}[r] & \left [ f \longmapsto \mathbb{D} (f) (x) \right ] }$$ D’après l’exactitude de la suite de Hodge-Tate de $H^D$ $$\ker \left ( V_p (H)^*\otimes K\twoheadrightarrow \omega_G\otimes K(-1) \right)\otimes_L \DH_\Q = \{ \;h :V_p (H)\otimes K\ldrt \DH_\Q\otimes K \; |\; \a_H (x) =0 \impl h(x) =0 \;\}$$ Mais via l’identification $\Hom_{\O_D} (\Gb,\Hb)\unp \simeq V_p (H)$ l’application $\a_H$ est $$f\longmapsto \left ( \mathbb{D} (f) \otimes Id \right ) ( \zeta_{G*} (1) )$$ où $\mathbb{D} (f)\otimes Id : \DG\otimes K\ldrt \DH\otimes K$, $\zeta_G= \rho_G \circ \eta_G (1)$ et $\zeta_{G*}$ est l’application induite sur les cristaux. On en déduit que la filtration définissant $(G',\rho_{G'})$ sur $\DG_\Q\otimes K$ est $$\{\, x\in \DG_\Q\otimes K \; |\; \forall f \in \Hom_{D,\ph} \left ( \DG_\Q,\DH_\Q \right ) \; (f\otimes Id) (\zeta_{G*} (1))=0 \impl (f\otimes Id) (x) =0 \;\} =\left ( Im (\a_G)^\perp \right )^\perp$$ qui est donc égal à $Im \a_G $ qui par surjectivité de l’application de Hodge-Tate de $G$ est la filtration définissant $G$. Donc $(G',\rho_{G'}) = (G,\rho_G)$. Il est maintenant aisé de vérifier que les rigidifications des modules de Tate de $G'$ et $G$ coïncident puisqu’il suffit de vérifier qu’elles coïncident modulo $p$, dans $\Hom (F/\O_F, \Gb\otimes \O_K/p\O_K)$ ce qui est immédiat. Donc $(G',\rho_{G'},\eta_{G'}) = (G,\rho_G, \eta_G)$.\ On vérifie de la même façon que l’application composée $$\M_\infty^{\LT} (K) \ldrt \M_\infty^\D (K) \ldrt \M_\infty^{\LT} (K)$$ est l’identité. Théorèmes de comparaison entiers relatifs pour les groupes $p$-divisibles d’après Faltings ========================================================================================== Dans cet appendice on explique les résultats auxquels on peut parvenir à partir des méthodes de [@Faltings1] pour les périodes cristallines et de Hodge-Tate des groupes $p$-divisibles sur les anneaux d’entiers de corps non-archimédiens. Les démonstrations seront données dans [@Periodes]. Groupes p-divisibles sur les anneaux d’entiers de corps non-archimédiens ------------------------------------------------------------------------ Soit $K|\Qp$ un corps valué complet pour une valuation de rang $1$, c’est à dire à valeurs dans $\R$, étendant celle de $\Qp$. On note $\O_{K}$ son anneau des entiers et $k$ son corps résiduel. Fixons $\O_{K_0}\subset \O_{K}$ un anneau de Cohen. On a des extensions valuées $K|K_0 |\Qp$ . Le choix de l’anneau de Cohen fixe en particulier une section $\e$ du morphisme $\O_{K}/p\O_{K} \twoheadrightarrow k$. Si $k$ est parfait $\O_{K_0}\simeq W(k)$. Par définition un groupe $p$-divisible sur $\spf (\O_K)$ est un système compatible de groupes $p$-divisibles sur les $(\spec (\O_{K}/p^n \O_{K}))_{n\geq 1}$ (cf. le début de la section 1 de [@Cellulaire]). Soit $G$ un groupe $p$-divisible sur $\spf (\O_K)$. Son équivalents : - $G\otimes_{\O_{K}} \O_{K}/p \O_{K}$ est isogéne à un groupe constant $H$ sur $k$ : $$G\otimes_{\O_{K}} \O_{K}/p \O_{K}\sim H\otimes_{k,\e} \O_{K}/p \O_{K}$$ - Il existe un nombre réel $\l \geq 1$ et $H'$ un groupe p-divisible sur $k$ tels que si $\mathfrak{m}_{K,\l}=\{ x \in K \;|\; v_p (x)\geq \l \;\}$ alors $$G\otimes \O_{K}/ \mathfrak{m}_{K,\l} \simeq H'\otimes_{k,\e} \O_{K}/ \mathfrak{m}_{K,\l}$$ - Soit $R\simeq \O_{K_0}[[x_i]]_{i\in I}$ l’anneau universel des déformations du groupe $G\otimes k$ sur des $\O_{K_0}$-algèbres locales complètes et $G^{univ}$ la déformation universelle. Il existe un morphisme $$x : \spf (\O_{K}) \ldrt \spf ( R)$$ i.e. $x\in \spf ( R)^{an} (K)$ où $\,^{an}$ désigne la fibre générique au sens des espaces de Berkovich, tel que $G\simeq x^* G^{univ}$ Elle ne pose pas de problème. On remarquera que si les conditions du lemme précédent sont vérifiées alors nécessairement $H'\simeq G_k$ et on peut choisir $H=G_k$. Un groupe p-divisible satisfaisant les conditions équivalentes du lemme précédent sera dit isotrivial mod $p$. En utilisant des morphismes non-continus $\O_{K_0}[[x_i]]_{i\in I} \ldrt \O_{\Cp}$ on peut construire des groupes $p$-divisibles non-isotriviaux mod $p$ sur $\O_{\Cp}$. Théorèmes de comparaison ------------------------ On reprend les notations de la section précédente. On fixe un relèvement de Frobenius $\s: \O_{K_0}\drt \O_{K_0}$. Soit $G$ un groupe p-divisible sur $\spf (\O_{K})$. On note $M$ l’algèbre de Lie de l’extension vectorielle universelle de $G$. Celle-ci est filtrée : $$0 \ldrt \omega_{G^D} \ldrt M \ldrt \omega_G^* \ldrt 0$$ On note $\text{Fil}\, M = \omega_{G^D}$. On note $\mathcal{E}$ le cristal de Messing (covariant) de $G\otimes \O_{K}/ p \O_{K}$ sur le gros site cristallin nilpotent de [@BBM] $$NCRIS(\spec (\O_{K}/p \O_{K})/ \spec (\O_{K_0}))$$ où $\spec (\O_{K_0}) \ldrt \spec (\O_{K}/p \O_{K})$ est défini via la section $\e$, et $$M_0 = \mathcal{E}_{\O_{K_0}\twoheadrightarrow k}$$ le module de Dieudonné “classique” de $G\otimes k$ si $k$ est parfait. Il est muni d’une application $\s$-linéaire $\varphi:M_0\ldrt M_0$ associée au relèvement de Frobenius $\s$. Rappelons ([@Periodes]) que l’on dispose d’une $\O_{K_0}$-algèbre $A_{cris} (\O_{K})$ augmentée via $\theta : A_{cris} (\O_{K}) \twoheadrightarrow \mathcal{O}_{\widehat{\overline{K}}}$. On note $\varphi $ le Frobenius cristallin sur $A_{cris}$. L’évaluation $$E = \mathcal{E}_{ A_{cris} \twoheadrightarrow \mathcal{O}_{\widehat{\overline{K}}}}$$ est un $A_{cris}$-module libre muni d’un morphisme $\varphi$-linéaire $\varphi : E \ldrt E$ tel que $$E\otimes_{\theta} \mathcal{O}_{\widehat{\overline{K}}} \simeq M\otimes_{\O_{K}} \otimes \mathcal{O}_{\widehat{\overline{K}}}$$ ce qui permet de filtrer $E$ via $\text{Fil}\, E = \theta^{-1} (\text{Fil}\, M \otimes \mathcal{O}_{\widehat{\overline{K}}})$. On a donc un $\varphi$-module filtré $(E,\varphi, \text{Fil}\, E)$. \[crop\] Il y a un isomorphisme naturel $\Gal (\overline{K}|K)$-équivariant $$T_p (G) \iso \left ( \text{Fil}\; E \right )^{\varphi =p}$$ \[bouqibouc\] Il y a deux inclusions naturelles strictement compatibles aux filtrations et à l’action de Galois $$t E \subset T_p (G)\otimes_{\Zp} A_{cris} \subset E$$ Elles sont compatibles aux Frobenius cristallins lorsque l’on munit $T_p (G) \otimes A_{cris}$ de $p\otimes \ph$. Dans le théorème précédent la filtration est indexée de la façon suivante $$\begin{aligned} \Fil^{\;i} \left ( T_p (G)\otimes A_{cris}\right ) &=& T_p (G)\otimes \Fil^{\; i} A_{cris} \\ \forall i\leq -1 \;\Fil^{\;i}\, E &=& E \\ \Fil^{\; 0}\, E &= & \theta^{-1} (\text{Fil} \,M \otimes \mathcal{O}_{\widehat{\overline{K}}}) \\ \forall i\geq 1\; \Fil^{\; i}\, E &=& \Fil^{\; i} A_{cris}. \Fil^{\; 0} \end{aligned}$$ Lorsque le groupe $p$-divisible $G$ est isotrivial mod $p$ son isocristal est engendré par ses section horizontales. On en déduit le théorème suivant. \[comp\_modulaire\] Supposons de plus que $G$ est isotrivial mod $p$. Il y a alors des isomorphismes $$M_0 \unp \otimes_{K_0} K \simeq M \unp$$ et $$E \unp \simeq M_0 \unp \otimes_{K_0} B^+_{cris}$$ comme $\varphi$-modules où $B^+_{cris} = A_{cris} \unp$. Il y a donc des inclusions strictement compatibles aux filtrations $$t M_0 \unp \otimes_{K_0} B^+_{cris} \subset V_p (G) \otimes_{\Qp} B^{+}_{cris} \subset M_0 \unp \otimes_{K_0} B^+_{cris}$$ un isomorphisme $$M_0\unp\otimes_{K_0} B_{cris} \simeq V_p (G) \otimes_{\Qp} B_{cris}$$ avec $B_{cris} =B^+_{cris} \unp$ et un autre isomorphisme $$V_p (G) \iso \Fil \left ( M_0\unp \otimes_{K_0} B^+_{cris}\right )^{\varphi = p }$$ où sur $M_0\otimes B^+_{cris}$, $\varphi= \varphi \otimes \varphi$ et $\Fil$ est la filtration associée à celle de $M\unp$ via l’isomorphisme $M_0\unp \otimes_{K_0} K \simeq M\unp$ et $\theta$. Le déterminant des périodes divisé par $2i\pi$ est une unité $p$-adique ----------------------------------------------------------------------- ### Énoncé et dévissage au cas C.M. Soit $K$ comme précédemment et supposons de plus que son corps résiduel est algébriquement clos (ce que l’on peut toujours réaliser quitte à étendre les scalaires). Soit $G$ un groupe $p$-divisible modulaire sur $\spf (\O_{K})$ de dimension $d$ et hauteur $h$. L’isomorphisme de $B_{cris}$-modules $$V_p (G) \otimes_{\Qp} B_{cris} \iso M_0 \unp \otimes_{K_0} B_{cris}$$ du théorème \[comp\_modulaire\] induit un isomorphisme $\ph$-équivariant $$\a: \det V_p (G) \otimes_{\Qp} B_{cris} \iso ( \det M_0 )\unp \otimes_{K_0} B_{cris}$$ Plus précisément, il y a une inclusion de $B^+_{cris}$-modules compatible aux filtrations et Frobenius $$u : T_p (G) \otimes B^+_{cris} \hookrightarrow M_0\otimes_{\O_{K_0}} B^+_{cris}$$ de conoyau annulé par $t$, d’où une inclusion $$\det u : \det T_p (G) \otimes_{\Zp} B^+_{cris} \hookrightarrow \left ( \det M_0 \right ) \otimes_{\O_{K_0}} B^+_{cris}$$ Via l’application de réduction modulo $\Fil^1 B^+_{cris}$ et l’identification $M_0 \otimes_{\O_{K_0}} \widehat{\overline{K}} \simeq M \otimes_{\O_K} \widehat{\overline{K}}$ (où $M=\Lie\, E (G)$) $$\left ( M_0 \otimes_{\O_{K_0}} B^+_{cris} \right ) / \Fil^1 B^+_{cris}.\left ( M_0 \otimes_{\O_{K_0}} B^+_{cris} \right ) \simeq M \otimes_{\O_K} \widehat{\overline{K}}$$ et $\Fil\left ( M_0\otimes B^+_{cris}\right ) $ est l’image réciproque de $\Fil \, M \otimes \widehat{\overline{K}}$ (où $\Fil\, M = \omega_{G^D}$). Il existe donc d’après le lemme de Nakayama une base $(e_1,\dots,e_n)$ du $B^+_{cris}$-module $M_0\otimes B^+_{cris}$ telle que $$\Fil \left ( M_0\otimes_{\O_{K_0}} B^+_{cris}\right ) = B^+_{cris} e_1\oplus \dots \oplus B^+_{cris} e_{n-d-1} \oplus \Fil^1 B^+_{cris} e_{n-d} \oplus \dots \Fil^1 B^+_{cris} e_n$$ L’image de $u$ est incluse dans $\Fil\left (M_0\otimes B^+_{cris}\right )$. On en déduit que $$\det u : \det T_p (G) \otimes B^+_{cris} \hookrightarrow \det M_0 \otimes_{\O_{K_0}} \Fil^d B^+_{cris}$$ qui définit donc un élément $$\beta \in \left ( \det T_p (G)^{-1} \otimes_{\Zp} \det M_0 \otimes_{\O_{K_0}} \Fil^d\, B^+_{cris} \right )^{\ph = Id}$$ où rappelons que $\ph$ agit par la multiplication par $p$ sur $T_p (G)$ (cf. théorème \[bouqibouc\]). Mais le corps résiduel de $K$ étant algébriquement clos $$(\det M_0, \det \ph) \simeq (\O_{K_0}, p^{h-d} \s)$$ Donc, $$\beta \in \det (T_p (G))^{-1} \otimes_{\Zp} \left ( \det ( M_0)\right )^{\ph= p^{h-d}} \otimes_{\Zp} \left (\Fil^d\, B^+_{cris} \right )^{\ph = p^d}$$ Mais puisque (cf. [@Periodes]) $$\left (\Fil^d\, B^+_{cris} \right )^{\ph = p^d} = \Qp.t^d$$ Donc $$\beta \in \left ( \det T_p (G)\right )^{-1}\otimes \left ( \det M_0\right )^{\ph= p^{h-d}} \otimes \Qp.t^d$$ Les structures entières $\det T_p (G) \simeq \Zp$, $\det M_0 \simeq \O_{K_0}$ induisent une $\Zp$-structure sur $\Qp t^d$. L’élément $\beta$ fournit donc un élément de $ \Qp^\times/\Zp^\times. t^d $ \[Faltings\] \[det\_periodes\_unite\] L’élément $\beta$ est entier : $\beta \in \Zp^\times.t^d$. Le groupe $G$ étant modulaire on peut le mettre en famille : pour un morphisme $x:\spf (\O_K)\drt \spf (\O_{K_0} [[x_i]]_{1\leq i\leq d(h-d)})$, $G=x^*G^{univ}$ où $G^{univ}$ désigne la déformation universelle de la fibre spéciale de $G$. Il existe de plus un rationnel $\a\in\Q_{>0}$ tel que $\forall i \; v_p ( x^* (x_i))\geq \a$. Le morphisme $x$ se factorise donc en $$x: \spf (\O_{K}) \ldrt \mathcal{C} \ldrt \spf (\O_{K_0} [[x_i]]_i)$$ où $\mathcal{C}=\spf (R)$ est un modèle formel $p$-adique normal sans $p$-torsion sur $\spf (O_{K_0})$ de la boule rigide formée des éléments de valuation supérieure ou égale à $\a$. Il y a donc un morphisme de spécialisation $A_{cris} (R)\twoheadrightarrow A_{cris} (\O_{K})$. Sur $\spec (R/pR)$ le groupes $H\otimes R/pR$ est isogéne au groupe définissant l’espace des déformations. Donc, si $H$ désigne le groupe $p$-divisible restriction de $G^{univ}$ à $\mathcal{C}$ il y a un isomorphisme de comparaison $$V_p (H) \otimes_{\Qp} B_{cris} (R) \iso M_0 \unp \otimes_{K_0} B_{cris} (R)$$ (cf. [@Periodes]). On obtient comme précédemment un élément $$\beta'\in \left ( ( \det T_p (H))^{-1} \otimes_{\Zp} \det M_0 \otimes_{\O_{K_0}} \Fil^d\, B_{cris}^+(R)\right )^{\ph = Id} \simeq \Qp t^d$$ où la dernière égalité résulte de [@Periodes]. Et bien sûr, via $A_{cris} (R) \twoheadrightarrow A_{cris} (\O_{K})$, $\beta'\mapsto \beta$. L’énoncé du théorème est donc invariant par transport parallèle : on peut transporter $\beta$ parallèlement en n’importe quel point du disque unité $\spf (\O_{K_0}[[x_i]]_i)^{an}$ (quitte à agrandir le rayon de la boule). Le mieux est de choisir un point pour lequel la matrice des périodes est diagonalisable. On peut par exemple prendre un point C.M. ayant multiplication complexe par $\Z_{p^h}$ où $h$ est la hauteur de $G$. En effet, l’action de $\Z_{p^h}$ permet de diagonaliser la matrice des périodes si l’on prend des bases de vecteurs propres pour cette action. Le résultat est démontré dans la section suivante dans ce cas particulier. ### Étude des périodes entières des groupes p-divisibles ayant multiplication complexe par un ordre maximal non-ramifié {#CM} Nous étudions ici les périodes des groupes p-divisible C.M. les plus simples, ceux ayant multiplication complexe par l’anneau des entiers d’une extension non-ramifiée de $\Qp$. Soit donc $G$ un groupe p-divisible de hauteur $h$ et dimension $d$ sur $\Z_{p^h}$ muni d’une action $\iota:\Z_{p^h}\drt \End (G)$. On pourra par exemple prendre lorsque $d=1$ le groupe formel d’exponentielle $$f(T)=\sum_{n\geq 0} \frac{T^{p^{nh}}}{p^n}$$ qui est bien muni d’une action de $\Z_{p^h}$ puisque $\forall \zeta \in \mu_{p^h-1}\;\; f(\zeta T) =\zeta f(T)$ ou n’importe quel groupe de Lubin-Tate de hauteur $1$ pour l’extension $\Q_{p^h}$. Notons $$\chi: \Gal(\Qpb |{\Q_{p^h}}) \ldrt \Z_{p^h}^\times$$ le caractère de Lubin-Tate. Notons $\s \in \Gal (\Q_{p^h}|\Qp)$ le Frobenius et $\forall i\; \chi^{\s^i}= \s^i \circ \chi$. Il résulte de la classification des représentations cristallines abeliennes (Fontaine, cf. [@Wintemberger1]) qu’il existe des entiers $a_i$, $0\leq a_i \leq d$ tels que $$V_p (G)= \prod_{i=0}^{h-1} \chi^{a_i \s^i}$$ où $\sum_i a_i = d$ et $a_i=\text{rg}_{\Z_{p^h}} \text{Lie} (G)_i$ avec $$\Lie G = \bigoplus_{i\in \Z/h\Z} \Lie (G)_i$$ $\iota (\Z_{p^h})$ agissant sur $ \text{Lie} (G)_i$ via $\s^i$. Considérons l’application de comparaison $$T_p (G) \otimes_{\Zp} A_{cris} (\Z_{p^h}) \hookrightarrow M_0\otimes_{\Z_{p^h}} A_{cris} (\Z_{p^h})$$ où $M_0$ est le module de Dieudonné de la fibre spéciale de $G$. L’action $\iota$ permet de décomposer $$\begin{aligned} T_p (G) \otimes_{\Zp} A_{cris} (\Z_{p^h}) &=& \bigoplus_{i\in \Z/ h\Z} \left (T_p (G) \otimes_{\Zp} A_{cris} (\Z_{p^h})\right )_i \\ M_0 &=& \bigoplus_{i\in \Z/h \Z} M_{0,i}\end{aligned}$$ où $\iota (\Z_{p^h})$ agit sur la composante indexée par $i$ via $\s^i$. La naturalité du morphisme de comparaison implique que celui-ci est somme directe de morphismes $$\left ( T_p (G) \otimes_{\Zp} A_{cris} (\Z_{p^h}) \right )_i \hookrightarrow M_{0,i} \otimes_{\Z_{p^h}} A_{cris} (\Z_{p^h})$$ Après choix d’une base de $T_p (G)$ et $M_{0,i}$ comme $\Z_{p^h}$-modules ce morphisme est donnée par la multiplication par un élément $x_i \in A_{cris}$ bien défini modulo $\Z_{p^h}^\times$. D’après la compatibilité stricte aux filtrations du morphisme de comparaison si $a_i=0, x_i\notin \Fil^1 A_{cris}$ et $x_i \in \Fil^1 A_{cris}$ sinon. Notons alors $y_i\in \O_{\C_p}$ l’image de $x_i$ dans $\text{Gr}^0 A_{cris}$, resp. $\text{Gr}^1 A_{cris}$. Supposons $d=1$ et soit $i_0$ l’unique indice tel que $a_{i_0}\neq 0$. Alors, $$\begin{aligned} \forall i<i_0 \; v(y_i) &=& \frac{p^{h+i- i_0}}{p^h-1} \\ \forall i\geq i_0 \; v(y_i) &=& \frac{p^{i-i_0}}{p^h -1}\end{aligned}$$ La compatibilité du morphisme de comparaison à l’action de $\Gal(\Qpb | {\Q_{p^h}})$ est équivalente à ce que $$\forall \tau \in G_{\Q_{p^h}}\;\;\; x_i^\tau =\chi_G (\tau)^{\s^i} x_i$$ où $\chi_G$ désigne le caractère galoisien associé à $V_p (G)$. Puisque $d=1$, $\chi_G=\chi^{\s^{i_0}}$. Considérons maintenant le lemme suivant : Soit $z\in \O_{\C_p}$ tel que $\forall \tau \in \Gal (\Qpb|\Q_{p^h}) \;\; z^\tau = \chi (\tau)^{\s^i} z$. Alors, $$\exists j\in \N\;\; v(z)\in \frac{p^i}{p^{jh} (p^h -1)} + \Z$$ Pour $\l \in \Q$ notons $\O_{\C_p,\geq \l}$, resp. $\O_{\C_p,> \l}$, les éléments de valuations supérieure à $\l$, resp. strictement supérieure à $\l$. Notons $q=p^h$. Décrivons l’action de $\Gal(\Qpb|\Q_{p^h})$ sur $\O_{\C_p,\geq \l}/\O_{\C_p,> \l} \simeq \overline{\Fp}$. Fixons $p^\l\in \Qpb$ un élément tel que si $\l=\frac{r}{s}$ avec $r\wedge s=1$ on ait $ (p^\l)^s = p^r$. Soit $\tau_0\in \Gal (\Qpb | \Q_{p^h}(p^\l))$ un relèvement du Frobenius $x\mapsto x^{p^h}$. L’inertie $I_{\Q_{p^h}}$ agit sur $\O_{\C_p,\geq \l}/\O_{\C_p,> \l}$ via le caractère modéré $$\tau \mapsto \frac{\tau (p^\l)}{p^\l} \text{ mod } \O_{\C_p,> \l} \in \overline{\Fp}^\times$$ Écrivons $z= u p^\l$ où $u\in \O_{\C_p}^\times$. Alors, modulo $\O_{\C_p,> \l}$ $$\begin{aligned} \forall \tau \in I_{\Q_{p^h}}\; \forall k\in \Z\; \; \frac{\tau_0^k\tau (z)}{z} &\equiv &\frac{\tau_0^k (u)}{u} . \frac{\tau_0^k \tau (p^\l)}{p^\l} \;\text{ car } \frac{\tau (u)}{u} \equiv 1 \\ &\equiv & u^{q^k-1} \left ( \frac{\tau (p^\l)}{p^\l} \right )^{q^k} \;\text{ car } \tau_0 (p^\l)=p^\l\end{aligned}$$ Quant au caractère $\chi$ il vérifie la congruence $$\forall \tau'\in \Gal (\Qpb| \Q_{p^h})\;\; \chi ( \tau') \equiv \frac{\tau' (p^{\frac{1}{q-1}})}{p^{\frac{1}{q-1}}} \in \mu_{q-1}$$ L’hypothèse du lemme implique que $$\forall k\in \Z\; \forall \tau\in I_{\Q_{p^h}} \;\; u^{q^k-1} \frac{\tau ( p^{q^k \l-\frac{p^i}{q-1}})}{ p^{q^k \l-\frac{p^i}{q-1}}} \equiv 1$$ ce qui implique d’abord avec $\tau= Id$ que $\bar{u}\in {\mathbb{F}_q}^\times$ et que l’on peut donc supposer que $u=1$ dans la congruence ci-dessus. De plus, pour $\mu \in\Q$, le caractère modéré $$\begin{aligned} I_{\Q_{p^h}}&\ldrt & \overline{\Fp}^\times \\ \tau & \longmapsto & \frac{\tau(p^\mu)}{p^\mu}\end{aligned}$$ est trivial ssi $\mu\in \Z [ \frac{1}{p}]$. Donc, $\forall k\in \Z\; q^k \l -\frac{p^i}{q-1} \in \Z[\frac{1}{q}]$ ce qui implique facilement le lemme. Il résulte du lemme précédent que $$\begin{aligned} \forall i>i_0\; \exists j\in \N\; v_p (y_i)\in \frac{p^{i-i_0}}{q^j ( q-1)} + \Z \\ \text{ et } \forall i<i_0\; \exists j\in \N\; v_p (y_i)\in \frac{p^{h+i - i_0}}{q^j (q-1)}+\Z \end{aligned}$$ Quant à $y_{i_0}$, étant donné que dans $\text{Gr}^1 A_{cris}$ $t$ se transforme via $N_{\Q_{p^h}/\Qp} \circ \chi$ et que $v(t)=\frac{1}{p-1}$, $$\exists j\in \N\; v_p (y_{i_0})\in \frac{1}{q^j (q-1)} + \Z$$ Le morphisme de comparaison possède un quasi-inverse tel qu’avec composition avec celui-ci on obtienne la multiplication par $t$ qui est de valuation $1/(p-1)$. On en déduit que $$\begin{aligned} \forall i \leq i_0 \;\; 0 \leq v(y_i) \leq \frac{p^{i-i_0}}{q-1} \\ \forall i <i_0 \; \; 0 \leq v(y_i) \leq \frac{p^{h+i-i_0}}{q-1}\end{aligned}$$ Remarquons maintenant que $$\prod_i x_i = \beta t$$ où $\beta \in \Zp$, et $$v(\beta) + \frac{1}{p-1} = \sum_i v_p (y_i) \leq \frac{1}{p-1}$$ et donc $v(\beta)=0$ i.e. $\beta \in \Zp^\times$ et les valuations des $y_i$ sont celle annoncées. Il résulte donc de la démonstration de la proposition précédente que L’énoncé du théorème \[det\_periodes\_unite\] est vrai dans le cas de dimension $1$ et de multiplication complexe par $\Z_{p^h}$· Attaquons maintenant le cas de dimension $d$ quelconque. Commençons par remarquer qu’il résulte de l’étude du cas de dimension $1$ et du théorème de Tate, $H^0 (G_{\Q_{p^h}},\C_p)=\Q_{p^h}$, que Pour tout entier $i$ compris entre $0$ et $h-1$, $$H^0 (G_{\Q_{p^h}}, \C_p (\chi^{-\s^i}))=\Q_{p^h}.z_i$$ où $z_i\in \C_p$ est un élément de valuation $\dpt{\frac{p^i}{p^h-1}}$. Il en résulte que l’on a une majoration des valuations des éléments $y_i$ (les périodes partielles). Sachant que $\prod_i y_i = \beta $ on en déduit le résultat facilement. On renvoie également à l’article [@Colmez1].
{ "pile_set_name": "ArXiv" }
--- abstract: 'In last passage percolation models, the energy of a path is maximized over all directed paths with given endpoints in a random environment, and the maximizing paths are called [*geodesics*]{}. The geodesics and their energy can be scaled so that transformed geodesics cross unit distance and have fluctuations and scaled energy of unit order. Here we consider Poissonian last passage percolation, a model lying in the KPZ universality class, and refer to scaled geodesics as [*polymers*]{} and their scaled energies as [*weights*]{}. Polymers may be viewed as random functions of the vertical coordinate and, when they are, we show that they have modulus of continuity whose order is at most $t^{2/3}\big(\log t^{-1}\big)^{1/3}$. The power of one-third in the logarithm may be expected to be sharp and in a related problem we show that it is: among polymers in the unit box whose endpoints have vertical separation $t$ (and a horizontal separation of the same order), the maximum transversal fluctuation has order $t^{2/3}\big(\log t^{-1}\big)^{1/3}$. Regarding the orthogonal direction, in which growth occurs, we show that, when one endpoint of the polymer is fixed at $(0,0)$ and the other is varied vertically over $(0,z)$, $z\in [1,2]$, the resulting random weight profile has sharp modulus of continuity of order $t^{1/3}\big(\log t^{-1}\big)^{2/3}$. In this way, we identify exponent pairs of $(2/3,1/3)$ and $(1/3,2/3)$ in power law and polylogarithmic correction, respectively for polymer fluctuation, and polymer weight under vertical endpoint perturbation. The two exponent pairs describe [@H121; @H122; @H123] the fluctuation of the boundary separating two phases in subcritical planar random cluster models.' address: - 'Departments of Mathematics and Statistics, University of California at Berkeley, Berkeley, CA, USA.' - 'Department of Statistics, University of California at Berkeley, Berkeley, CA, USA.' author: - Alan Hammond - Sourav Sarkar bibliography: - 'HolderLPP.bib' date: - - title: Modulus of continuity for polymer fluctuations and weight profiles in Poissonian last passage percolation --- [^1] Introduction ============ In 1986, Kardar, Parisi, and Zhang [@KPZ86] predicted universal scaling behaviour for many planar random growth processes, including first and last passage percolation as well as corner growth processes, though rigorous validation has been subsequently provided for only a handful of them. In such models, fluctuation in the direction of growth is governed by an exponent of one-third, with this fluctuation enduring on a scale governed by an exponent of two-thirds in the orthogonal, or transversal, direction. Poissonian last passage percolation illustrates these effects. We will define it shortly, since it is our object of study; briefly, the model specifies a growth process whose height at a given moment is the maximum number of points (or the [*energy*]{}) obtainable in a directed path through a planar Poisson point process. Baik, Deift and Johansson [@BDJ99] established the $n^{1/3}$-order fluctuation of the maximum number of Poisson points on an increasing path from $(0,0)$ to $(n,n)$, deriving the GUE Tracy-Widom distributional limit of the scaled energy. Later Johansson [@J00] proved the transversal fluctuation exponent of two-thirds in this model. These are exactly solvable models, for which certain exact distributional formulas are available, and the derivations of these formulas typically employ deep machinery from algebraic combinatorics or random matrix theory. It is interesting to study geometric properties of universal KPZ objects by approaches that, while they are reliant on certain integrable inputs, are probabilistic in flavour: for example, [@BSS14],[@BSS17++] and [@BSS17+] are recent results and applications concerning geometric properties of last passage percolation paths. It is rigorously understood, then, that last passage percolation paths experience fluctuation in their energy and transversal fluctuation governed by scaling exponents of one-third and two-thirds. It is very natural to view such paths via the lens of scaled coordinates, in which transversal fluctuation and path energy each has unit order. We will be more precise very shortly, when suitable notation has been introduced, but for now we mention that our aim in this article is to refine rigorous understanding of the magnitude and geometry of fluctuation in last passage percolation paths. We shall call the scaled geodesics [*polymers*]{}, and refer to the scaled energy as [*weight*]{}. We will see that polylogarithmic corrections to the scaled laws implied by the exponents of one-third and two-thirds arise when we consider natural geometric problems concerning the weights and the maximum fluctuation among polymers in a unit order region. The techniques for verifying our claims will employ geometric and probabilistic tools rather than principally integrable ones, since problems involving maxima as both endpoints of a last passage percolation path are varied are not usually amenable to integrable techniques. Model definition and main results {#ss:defres} --------------------------------- Let $\Pi$ be a homogeneous rate one Poisson point process (PPP) on $\R^2$. We introduce a partial order on $\R^2$: $(x_1,y_1)\preceq (x_2,y_2)$ if and only if $x_1\leq x_2$ and $y_1\leq y_2$. For $u\preceq v$, $u,v\in \R^2$, an increasing path $\gamma$ from $u$ to $v$ is a piecewise affine path, viewed as a subset of $\R^2$, that joins points $u=u_0\preceq u_1\preceq u_2 \preceq \ldots \preceq u_k = v$ such that $u_i\in \Pi$ for $i\in \llbracket 1,k-1\rrbracket$. Here and later, $\llbracket a,b\rrbracket$ for $a,b\in \Z$ with $a\leq b$ denotes the integer interval $\{a,\cdots,b\}$. Also let $|\gamma|$ denote the [*energy*]{} of $\gamma$, namely the number of points in $\Pi\setminus\{v\}$ that lie on $\gamma$; (the last vertex is excluded from the definition of energy so that the sum of the energies of two paths equals the energy of the concatenated path, as we will see in Section \[ss:pocat\]). Then we define the last passage time from $u$ to $v$, denoted by $X_u^v$, to be the maximum of $|\gamma|$ as $\gamma$ varies over all increasing paths from $u$ to $v$. Any such maximizing path is called a geodesic. There may be several such, but if $\Gamma_u^v$ denotes any one of them, we have $$\label{e:defE} X_u^v=|\Gamma_u^v| \, .$$ Note that, in this notation, the starting and ending points of the geodesic, $u$ and $v$, are assigned subscript and superscript placements. We will often use this convention, including in the case of the scaled coordinates that we will introduce momentarily. When $u\preceq v$, any geodesic from $u$ to $v$ may be viewed as a function of its horizontal coordinate, since it contains a vertical line segment with probability zero. The operations of maximum and minimum may be applied to any pair of such geodesics, and the results are also geodesics. For this reason, we may speak unambiguously of $\Gamma_{u}^{\leftarrow;v}$, the uppermost geodesic between $u$ and $v$, and of $\Gamma^{\rightarrow;v}_{u}$, the lowermost geodesic between $u$ and $v$. (The notation $\leftarrow$ and $\rightarrow$ is compatible with these two paths being equally well described as the leftmost and rightmost geodesics. This choice of notation also anticipates the form of these paths when viewed in the scaled coordinates that we are about to introduce.) When the endpoints are $(0,0)$ and $(n,n)$, we will call these geodesics $\Gamma_n^{\leftarrow}$ and $\Gamma_n^{\rightarrow}$. ### Introducing scaled coordinates We rotate the plane about the origin counterclockwise by $45$ degrees, squeeze the vertical coordinate by a factor $2^{1/2}n$ and the horizontal one by $2^{1/2}n^{2/3}$, thus setting $$\label{e:deftrans} T_n:(x,y)\mapsto\left(2^{-1}n^{-2/3}(x-y),2^{-1}n^{-1}(x+y)\right) .$$ The horizontal line at vertical coordinate $t$ is the image under $T_n$ of the anti-diagonal line through $(nt,nt)$. It is easy to see that, for $(x,t)\in \R^2$, $T_n^{-1}(x,t)=(nt+xn^{2/3},nt-xn^{2/3})$. Paths that are the image of geodesics under $T_n$ will be called [*polymers*]{}; we might say $n$-polymers, but the suppressed parameter will always be $n$. Geodesics from $(0,0)$ to $(n,n)$ transform to polymers $(0,0)$ to $(0,1)$. Figure \[f:scaling\] depicts a geodesic $\Gamma$ and its image polymer $\rho$. The polymer between planar points $u$ and $v$ that is the image of the uppermost geodesic given the preimage endpoints will be denoted by $\rho^{\leftarrow;v}_{n;u}$, and, naturally enough, called the leftmost polymer from $u$ to $v$. The rightmost polymer from $u$ to $v$ is the image of the corresponding lowermost geodesic and will be denoted by $\rho^{\rightarrow;v}_{n;u}$. The simpler notation $\rho^{\leftarrow}_n$ and $\rho^{\rightarrow}_n$ will be adopted when $u = (0,0)$ and $v = (0,1)$. When $u=(x_1,t_1), v=(x_2,t_2)$, with $x_1,x_2,t_1,t_2\in \R$, $t_1<t_2$, such that $T_n^{-1}(x_1,t_1)\preceq T_n^{-1}(x_2,t_2)$, we will, when it is convenient, regard any polymer $\rho$ from $u$ to $v$ as a function of its vertical coordinate: that is, for $t \in [t_1,t_2]$, $\rho(t)$ will denote the unique point such that $(\rho(t),t)\in \rho$. (This definition makes sense since an increasing path can intersect any anti-diagonal at most once.) We regard the vertical coordinate as time, as the $t$-notation suggests, and will sometimes refer to the interval $[t_1,t_2]$ as the [*lifetime*]{} of the polymer. In particular, when $t_1=0,t_2=1$, writing $C[0,1]$ for the space of continuous real-valued functions on $[0,1]$ (equipped for later purposes with the topology of uniform convergence), we may thus view $\rho=\{\rho(t)\}_{t \in [0,1]}$ as an element of $C[0,1]$. ![The scaling map $T_n$ applied to the left figure produces the figure on the right. The point $e$ in the geodesic $\Gamma$ is the preimage of the point $(\rho(t),t)$ in the polymer $\rho$. []{data-label="f:scaling"}](scaling.pdf) ### Condition for existence of polymers {#sss:comp} For $u=(x_1,t_1),v=(x_2,t_2)$ with $x_1,x_2,t_1,t_2\in \R$, $t_1<t_2$, we have that $T_n^{-1}(u)=(nt_1+x_1n^{2/3},nt_1-x_1n^{2/3})$ and $T_n^{-1}(v)=(nt_2+x_2n^{2/3},nt_2-x_2n^{2/3})$. Thus $T_n^{-1}(u)\preceq T_n^{-1}(v)$ is and only if $|x_1-x_2|<n^{1/3}(t_2-t_1)$. Indeed, we will write $u\overset{n}{\preceq}v$ to mean that $|x_1-x_2|<n^{1/3}(t_2-t_1)$; this condition ensures that polymers exist between the endpoints $u$ and $v$. The first of our three main results shows that polymers, viewed as functions of the vertical coordinate, enjoy modulus of continuity of order $t^{2/3} \big( \log t^{-1} \big)^{1/3}$. \[t:main1\] - The sequence $\{\rho_n^{\leftarrow}\}_{n\in \N}$ is tight in $(C[0,1],\|\cdot\|_\infty)$. - There exists a constant $C > 0$ such that, for the weak limit $\rho_*^{\leftarrow}$ of any weakly converging subsequence of $\{\rho_n^{\leftarrow}\}_{n\in \N}$, almost surely, $$\label{e:modcon} \limsup_{t\searrow 0}\sup_{0\leq z\leq 1-t}{t^{-2/3}\big(\log t^{-1}\big)^{-1/3}}{|\rho_*^{\leftarrow}(z+t)-\rho_*^{\leftarrow}(z)|}\leq C \, .$$ The same result holds for the rightmost polymer. Note that the constant $C$ does not depend on the choice of the weakly converging subsequence. The exponent pair $(2/3,1/3)$ for power law and polylogarithmic correction is thus demonstrated to hold in an upper bound on polymer fluctuation. We believe that a lower bound holds as well, in the sense that the limit infimum counterpart to is positive. A polymer is an object specified by a global constraint, and it by no means clearly enjoys independence properties as it traverses disjoint regions, even though the underlying Poisson randomness does. In order to demonstrate the polymer fluctuation lower bound, this subtlety would have to be addressed. We choose instead to demonstrate that the exponent pair $(2/3,1/3)$ describes polymer fluctuation by proving a lower bound of this form for the maximum fluctuation witnessed among a natural class of short polymers in a unit region. This alternative formulation offers a greater supply of independent randomness. Indeed, we now specify a notion of [*maximum transversal fluctuation*]{} over a collection of short polymers. Fix any two points $u=(x_1,t_1),v=(x_2,t_2)$ such that $t_2>t_1$. Let $\Phi_{n;u}^v$ denote the set of all polymers $\rho$ from $u$ to $v$. Let $\l_{u}^v$ denote the planar line segment that joins $u$ and $v$; extending an abuse of notation that we have already made, we write $\l_u^v(t)$ for the unique point such that $(\l_u^v(t),t)\in \l_u^v$, where $t\in [t_1,t_2]$. Then, for any polymer $\rho$, the transversal fluctuation $\mathrm{TF}(\rho)$ of $\rho$ is specified to be $$\label{e:TFfixedpol} \mathrm{TF}(\rho):=\sup_{t\in [t_1,t_2]}|\rho(t)-\l_u^v(t)|,$$ and the transversal fluctuation between the points $u$ and $v$ to be $$\label{e:TFbetwnpoints} \mathrm{TF}_{n;u}^v:=\max_{\rho\in \Phi_{n;u}^v}\mathrm{TF}(\rho)=\max\left\{\mathrm{TF}(\rho_{n;u}^{\leftarrow;v}),\mathrm{TF}(\rho_{n;u}^{\rightarrow;v})\right\}.$$ Also, let $$\mathsf{InvSlope}_{(x_1,t_1)}^{(x_2,t_2)}=\frac{x_2-x_1}{t_2-t_1}$$ denote the reciprocal of the slope of the interpolating line. Since $t_2>t_1$, $\mathsf{InvSlope}_{(x_1,t_1)}^{(x_2,t_2)} \in \R$. Now fix some large constant $\psi>0$. Then, for any fixed parameter $t\in (0,1]$ and any $n\in \N, n>\psi^3$, we define the set of [*admissible endpoint pairs*]{} $$\begin{aligned} \label{e:defComp} \mathrm{AdEndPair}_{n,\psi}(t):=\Big\{((x_1,t_1),(x_2,t_2)):t_2-t_1\in (0,t] \, , \, \left|\mathsf{InvSlope}_{(x_1,t_1)}^{(x_2,t_2)}\right|\leq \psi,\nonumber \\ x_1,x_2\in [-1,1] \, , \, t_1,t_2\in [0,1]\Big\}.\end{aligned}$$ Since $n>\psi^3$, $$|x_2-x_1|n^{2/3}\leq \psi(t_2-t_1)n^{2/3}<(t_2-t_1)n\,.$$ Recalling the notation at the start of Subsection \[sss:comp\], we thus have $(x_1,t_1)\overset{n}\preceq (x_2,t_2)$, so that polymers do exist between such endpoint pairs. We then define $$\label{e:defMTF} \mathrm{MTF}_n(t) = \mathrm{MTF}_{n,\psi}(t):=\sup\left\{\mathrm{TF}_{n;u}^{v}:(u,v)\in \mathrm{AdEndPair}_{n,\psi}(t) \right\},$$ so that $\mathrm{MTF}_n(t)$ is the [*maximum transversal fluctuation*]{} over polymers between all endpoint pairs at vertical distance at most $t$ such that the slope of the interpolating line segment is bounded away from being horizontal; (we suppress the parameter $\psi$ in the notation). Our second theorem demonstrates that the exponent pair $(2/3,1/3)$ governs this maximum traversal fluctuation. \[t:flucshortpol\] There exist $\psi$-determined constants $0<c<C<\infty$ such that $$\liminf_n \, \P\left( \, t^{-2/3}\big(\log t^{-1}\big)^{-1/3} \mathrm{MTF}_n(t) \in [c,C] \, \right) \to 1 \quad \mbox{as } \, \, {t\searrow 0}\,.$$ ### Scaled energies are called weights It is natural to scale the energy of a geodesic when we view the geodesic as a polymer after scaling. Scaled energy will be called [*weight*]{} and specified so that it is of unit order for polymers that cross unit-order distances. For $t_1<t_2$, let ${t_{1,2}}$ denote $t_2-t_1$; (this is a notation that we will often use). Let $(x,t_1),(y,t_2)\in \R^2$ be such that $|x-y|<{t_{1,2}}n^{1/3}$. (This condition ensures that $(x,t_1)\overset{n}\preceq (y,t_2)$, so that polymers exist between this pair of points.) Since $T_n^{-1}((x,t_1))=( nt_1 + xn^{2/3},nt_1 - xn^{2/3})$ and $T_n^{-1}((y,t_2))=( nt_2 + yn^{2/3},nt_2 - yn^{2/3})$, it is natural to define the scaled energies, which we call *weights*, in the following way. Define $$\label{e:defweight} {\mathsf{W}}_{n;(x,t_1)}^{(y,t_1)} = n^{-1/3} \bigg( X_{( nt_1 + n^{2/3}x, nt_1 - n^{2/3}x)}^{(nt_2 + n^{2/3}y, nt_2 - n^{2/3}y)} \Big) \, - \, 2 n {t_{1,2}}\bigg) \, .$$ Because of translation invariance of the underlying Poisson point process, ${t_{1,2}}$ is a far more relevant parameter than $t_1$ or $t_2$. The notation on the left-hand side of (\[e:defweight\]) is characteristic of our presentation in this article: a scaled object is being denoted, with planar points $(\cdot,\cdot)$ in the subscript and superscript indicating starting and ending points. ### A continuous modification of the weight function For the statement of our third theorem, we prefer to make an adjustment to the polymer weight to cope with a minor problem concerning discontinuity of geodesic energy under endpoint perturbation. For $n\in \N$, define $X_n:[1,2]\mapsto [0,\infty)$, $$X_n(t):=X_{(0,0)}^{(nt,nt)} \, .$$ Observe that $X_n(t)$ is integer-valued, non-decreasing, right continuous and has almost surely a finite number of jump discontinuities. Let $d_0 = 1$ and $d_m = 2$. Record in increasing order the points of discontinuity of $X_n$ as a list $\big( d_1,d_2,\cdots,d_{m-1} \big)$. We specify a [*modified*]{} and continuous form of the function $X_n$ by linearly interpolating it between these points of discontinuity, setting $$X_n^{\mbox{\tiny{mod}}}(t):=X_n(d_i)+(t-d_i)(d_{i+1}-d_i)^{-1}\big(X_n(d_{i+1})-X_n(d_i)\big), \mbox{ for } t \in [d_i,d_{i+1}],$$ for $i=1,2,\cdots,m-1$. Because almost surely no two points in a planar Poisson point process share either their horizontal or vertical coordinate, $X_n(d_{i+1})-X_n(d_i)=1$ for all $i$. Thus, for all $t\in [1,2]$, $$\label{e:Xmod} X_n(t)\leq X_n^{\mbox{\tiny{mod}}}(t)\leq X_n(t)+1\,.$$ Now define the *modified weight* function $\mathsf{Wgt}_n:[1,2]\mapsto \R$ for polymers from $(0,1)$ to $(\cdot,1)$: $$\label{e:defWgt} \mathsf{Wgt}_n(t):=n^{-1/3}\left(X_n^{\mbox{\tiny{mod}}}(t)-2nt\right)$$ Because of , $$\label{e:Wmod} \left| \, \mathsf{Wgt}_n(t)-{\mathsf{W}}_{n;(0,0)}^{(0,t)} \, \right| \leq n^{-1/3} \, .$$ By construction, $\mathsf{Wgt}_n$ sending $t \in [1,2]$ to $\mathsf{Wgt}_n(t)$ is an element of $C[1,2]$, the space of continuous functions on $[1,2]$; (similarly to before, this space will be equipped with the topology of uniform convergence). Our third main result demonstrates that the exponent pair $(1/3,2/3)$ offers a description of the modulus of continuity of polymer weight when one endpoint is varied vertically. \[t:holdweight\] The sequence $\{\mathsf{Wgt}_n\}_{n\in \N}$ is tight in $(C[1,2],\|\cdot\|_\infty)$. There exist constants $0<c<C<\infty$ such that, for the weak limit $\mathsf{Wgt}_*$ of any weakly converging subsequence of $\{\mathsf{Wgt}_n\}_{n\in \N}$, almost surely $$\begin{aligned} \label{e:wgtholdbounds} c&\leq &\liminf_{t\searrow 0}\sup_{1\leq z\leq 2-t} \, t^{-1/3}\big(\log t^{-1}\big)^{-2/3} \, \Big\vert \mathsf{Wgt}_*(z+t)-\mathsf{Wgt}_*(z) \Big\vert \\ &\leq & \limsup_{t\searrow 0}\sup_{1\leq z\leq 2-t} \, t^{-1/3}\big(\log t^{-1}\big)^{-2/3} \, \Big\vert \mathsf{Wgt}_*(z+t)-\mathsf{Wgt}_*(z) \Big\vert \, \leq \, C \, .\nonumber\end{aligned}$$ Note that, as in Theorem \[t:main1\], the constants $c$ and $C$ do not depend on the choice of weak limit point or converging subsequence. Beyond these three theorems, we present a proposition, which is needed for the proof of Theorem \[t:flucshortpol\] but which also has independent interest. The maximum fluctuation of any geodesic joining $(0,0)$ and $(n,n)$ around the interpolating line has probability at most $e^{-c k^3}$ of exceeding $k n^{2/3}$. This upper bound has essentially been obtained in [@BSS14 Theorem $11.1$ and Corollary $11.7$], though we will state and prove this result, with the power of three in the exponent inside the exponential, as Theorem \[t:transversal\]. Our next proposition is the matching lower bound, stated using scaled coordinates. Observe from that, for any polymer $\rho$ between $(0,0)$ and $(0,1)$, $\mathrm{TF}(\rho)=\sup_{y\in [0,1]}|\rho(y)|$. Also recall that $\Phi_{n;(0,0)}^{(0,1)}$ is the set of all polymers from $(0,0)$ to $(0,1)$. \[p:lowbndTF\] There exist positive constants $c^*$, $n_0$, $s_0$ and $\alpha_0$ such that, for all $t_1,t_2$ with ${t_{1,2}}=t_2-t_1>0$ and all $n{t_{1,2}}\geq n_0$ and $s\in [s_0,\alpha_0(n{t_{1,2}})^{1/3}]$, $$\P\left( \min \left\{ \mathrm{TF}(\rho) : \rho\in \Phi_{n;(0,t_1)}^{(0,t_2)} \right\} \geq s {t_{1,2}}^{2/3} \right) \, \geq \, \exp \big\{-c^*s^3 \big\} \, .$$ A few words about the proofs ---------------------------- The main ingredients in the proofs of Theorem \[t:main1\] and Theorem \[t:flucshortpol\] are the estimates from integrable probability assembled in Section  \[s:estinteg\] and a [*polymer ordering*]{} property elaborated in Lemma \[l:porder\] that propagates control on polymer fluctuation among polymers whose endpoints lie in a discrete mesh to all polymers in the region of this mesh. The basic tools in the proof of the upper bound in Theorem \[t:holdweight\] and that of Proposition \[p:lowbndTF\] are surgical techniques and comparisons of the weights of polymers, and are reminiscent of the techniques developed and extensively used in [@BSS14] and [@BSS17++]. Phase separation and KPZ ------------------------ Certain random models manifest the scaling exponents of KPZ universality and some of its qualitative features, without exhibiting the richness of behaviour of models in this class. For example, the least convex majorant of the stochastic process $\R \to \R: x \to B(x) - t^{-1} x^2$ is comprised of planar line segments, or facets, the largest of which in a compact region has length of order $t^{2/3 + o(1)}$ when $t > 0$ is high; and the typical deviation of the process from its majorant scales as $t^{1/3 + o(1)}$. Some such models form a testing ground for KPZ conjectures. Phase separation concerns the form of the boundary of a droplet of one substance suspended in another. When supercritical bond percolation on $\Z^2$ is conditioned on the cluster (or droplet) containing the origin being finite and large, namely of finite size at least $n^2$, with $n$ high, the interface at the boundary of this cluster is expected to exhibit KPZ scaling characteristics, with the scaling parameter $n$ playing a comparable role to $t$ in the preceding example. Indeed, in [@H121; @H122; @H123], a surrogate of this interface, expressed in terms of the random cluster model, was investigated. The maximum length of the facets that comprise the boundary of the interface’s convex hull was proved to typically have the order $n^{2/3} \big( \log n \big)^{1/3}$, while the maximum local roughness, namely the maximum distance from a point on the interface to the convex hull boundary, was shown to be of the order of $n^{1/3} \big( \log n \big)^{2/3}$. Viewed in this light, the present article validates for the KPZ universality class the implied predictions: that exponent pairs of $(1/3,2,3)$ and $(2/3,1/3)$ for power-law and logarthmic-power govern maximal polymer weight change under vertical endpoint displacement and maximal transversal polymer fluctuation. In a natural sense, these two exponent pairs are accompanied by a third, namely $(1/2,1/2)$, for interface regularity. In the example of parabolically curved Brownian motion, $x \to B(x) - x^2 t^{-1}$, the modulus of continuity of the process on $[-1,1]$ is easily seen to have the form $s^{1/2} \big( \log s^{-1} \big)^{1/2}$, up to a random constant, and uniformly in $t \geq 1$. In KPZ, this assertion finds a counterpart when it is made for the Airy$_2$ process, which offers a limiting description in scaled coordinates of the weight of polymers of given lifetime with first endpoint fixed. This assertion has been proved in [@H16 Theorem $1.11(1)$]. Recently, for a very broad class of initial data, the polymer weight profile was shown in [@H17 Theorem  $1.2$] to have a modulus of continuity of the order of $s^{1/2}\big( \log s^{-1}\big)^{2/3}$, uniformly in the scaling parameter and the initial condition. Organization ------------ We continue with two sections that offer basic general tools. The first, Section \[s:estinteg\], provides useful estimates available from the integrable probability literature. Then, in Section \[s:polorder\], we state and prove the polymer ordering lemmas and some other basic results, which are essential tools in the proofs of the main theorems. The remaining four sections, \[s:2/3,1/3\] – \[s:last\], contain the main proofs. Consecutively, these sections are devoted to proving: - the polymer [Hölder ]{}continuity upper bound Theorem \[t:main1\]; - the modulus of continuity for maximum transversal fluctuation over short polymers, Theorem \[t:flucshortpol\], subject to assuming Proposition \[p:lowbndTF\]; - [Hölder ]{}continuity for the polymer weight profile, Theorem \[t:holdweight\]; - and the lower bound on transversal polymer fluctuation, Proposition \[p:lowbndTF\]. We will stick to scaled coordinates in the results’ statements and, except in Section \[s:estinteg\], in their proofs. A bridge between scaled coordinates and the original ones is offered in this next section, in whose proofs we use the scaling map $T_n$ from and weight function ${\mathsf{W}}$ from to transfer unscaled results to their scaled counterparts. Scalings and estimates from integrable probability {#s:estinteg} ================================================== In this section, we assemble some results from integrable probability. Most of these results were derived in terms of unscaled coordinates in [@BSS14] and [@BSS17++]. Point-to-point estimates of last passage percolation geodesics were used crucially in [@BSS14] to resolve the “slow-bond" conjecture, and in [@BSS17++] to show the coalescence of nearby geodesics, and those estimates will be crucially employed in this paper as well. We state the results in scaled coordinates, and the proofs detail how to obtain these statements from their unscaled versions available in the literature. In going from the unscaled to scaled coordinates, we shall use the definitions of the scaling map in and the weight in . First we observe some simple relations between the different scaled versions of these quantities that will be used in the proofs of the theorems in this section.\ [**The scaling principle.**]{} Because of translation invariance and the definition , it is easy to see that for any $x,y,t_1,t_2\in \R$ with ${t_{1,2}}=t_2-t_1>0$ and $(x,t_1)\overset{n}{\preceq}(y,t_2)$ (see Subsection \[sss:comp\]), for any $\theta\in [0,1]$, $$\label{e:screl} \rho_{n;(x,t_1)}^{\leftarrow;(y,t_2)}(t_1+\theta {t_{1,2}})\overset{d}{=}{t_{1,2}}^{2/3}\rho_{n;(x{t_{1,2}}^{-2/3},0)}^{\leftarrow;(y{t_{1,2}}^{-2/3},1)}(\theta)\overset{d}{=}{t_{1,2}}^{2/3}\rho_{n;(0,0)}^{\leftarrow;((y-x){t_{1,2}}^{-2/3},1)}(\theta)\,.$$ The same statement holds for the rightmost polymers as well. Here and throughout $\overset{d}{=}$ denotes that the two random variables on either side have the same distribution. We will sometimes call the displayed assertion the [*scaling principle*]{}. Also by translation invariance and the definition of weight in , it follows that $$\label{e:eqdist} {t_{1,2}}^{-1/3}{\mathsf{W}}_{n;(x,t_1)}^{(y,t_2)}\overset{d}{=}{\mathsf{W}}_{n{t_{1,2}};(xt_{1,2}^{-2/3},0)}^{(yt_{1,2}^{-2/3},1)}\overset{d}{=}{\mathsf{W}}_{n{t_{1,2}};(0,0)}^{((y-x)t_{1,2}^{-2/3},1)}\,.$$\ [**Boldface notation for applying results.**]{} In our proofs, we will naturally often be applying tools such as those stated in this section. Sometimes the notation of the tool and of the context of the application will be in conflict. To alleviate this conflict, we will use boldface notation when we specify the values of the parameters of a given tool in terms of quantities in the context of the application. We will first use this notational device shortly, in one of the upcoming proofs. The next theorem was proved in [@BDJ99]. \[t:BDJ\] As $n\rightarrow \infty$, $${\mathsf{W}}_{n;(0,0)}^{(0,1)}\Rightarrow F_{TW},$$ where the convergence is in distribution and $F_{TW}$ denotes the GUE Tracy-Widom distribution. For a definition of the GUE Tracy-Widom distribution, also called the $F_2$ distribution, see [@BDJ99]. Moderate deviation inequalities for this centred and scaled polymer weight will be important. Such inequalities follow immediately from [@LM01 Theorem  $1.3$], [@LMS02 Theorem $1.2$] and . These are essential inequalities, used repeatedly in this paper. In fact, it should be possible to recover the theorems of this paper for other integrable models for which such moderate deviation estimates are known. \[t:moddev\] There exist positive constants $c,s_0$ and $n_0$ such that, for all $t_1<t_2$ with $n{t_{1,2}}>n_0$ and $s>s_0$, $$\P\left({t_{1,2}}^{-1/3}{\mathsf{W}}_{n;(0,t_1)}^{(0,t_2)}\geq s\right)\leq e^{-cs^{3/2}},$$ and $$\P\left({t_{1,2}}^{-1/3}{\mathsf{W}}_{n;(0,t_1)}^{(0,t_2)}\leq -s\right)\leq e^{-cs^{3/2}}.$$ Also, we shall need not just tail bounds for weights of point to point polymers, but uniform tail bounds on polymer weights whose endpoints vary over fixed unit order intervals. The unscaled version of this theorem follows from [@BSS14 Propositions $10.1$ and $10.5$]. \[t:unifmoddev\] There exist $C,c \in (0,\infty), C_0\in (1,\infty)$ and $n_0 \in \N$ such that, for all $t_1<t_2$ with $n{t_{1,2}}\geq n_0$, $s\in [0,10(n{t_{1,2}})^{2/3}]$, $A=C_0^{-1}s^{1/4}n^{1/6}{t_{1,2}}^{5/6}$ and $I$ and $J$ intervals of length at most ${t_{1,2}}^{2/3}$ that are contained in $\left[-A,A\right]$, $$\P \bigg( \, \sup_{x \in I, y \in J} \Big\vert {t_{1,2}}^{-1/3} {\mathsf{W}}_{n;(x,t_1)}^{(y,t_2)} + {t_{1,2}}^{-4/3} (x-y)^2 \Big\vert > s \, \bigg) \, \leq \, C \exp \big\{ -c s^{3/2} \big\} \, .$$ [**Proof.**]{} First we prove the theorem when $t_1=0$ and $t_2=1$ by invoking the unscaled version of this theorem from [@BSS14]. At the end we prove Theorem \[t:unifmoddev\] for general $t_1<t_2$. Observe that $|x-y|<2C_0^{-1}s^{1/4}n^{1/6}{t_{1,2}}^{5/6}<2^{-1}n{t_{1,2}}^{1/3}$ for $C_0>2\cdot10^{1/4}$ since $s\leq 10(n{t_{1,2}})^{2/3}$. This ensures that ${\mathsf{W}}_{n;(x,t_1)}^{(y,t_2)}$ is well defined. Let $u=T_n^{-1}(x,0)=(xn^{2/3},-xn^{2/3})$ and $v=T_n^{-1}(y,1)=(n+yn^{2/3},n-yn^{2/3})$. If $S_{u,v}$ denotes the slope of the line segment joining $u$ and $v$, then $|x-y|<2^{-1}n$ ensures that $3^{-1}<S_{u,v}<3$. Then, using the first order estimates (see [@BSS14 Corollary $9.1$]) and a simple binomial expansion giving $|(1-x)^{1/2}-(1-2^{-1}x)|\leq C_1x^2$ for $x\in (-1,1)$, we get that $$\left|\E [X_u^v]-(2n-(x-y)^2n^{1/3})\right|\leq C_2n^{-1/3}(x-y)^4+C_2n^{1/3}\,,$$ for some constants $C_1,C_2>0$, where $X_u^v$ is defined in . Since $|x-y|\leq 2C_0^{-1}s^{1/4}n^{1/6}$, $$C_2n^{-2/3}(x-y)^4\leq 2^4C_0^{-4}C_2s<2^{-1}s$$ for $C_0>2^{5/4}C_2^{1/4}$. Hence, using the definition of the weight function in , for all $s\geq 6C_2$, $$\begin{aligned} \Bigg\{\left|{\mathsf{W}}_{n;(x,0)}^{(y,1)}+(x-y)^2\right|>s\Bigg\}&\subseteq & \Bigg\{n^{-1/3}\left|X_u^v-\E X_u^v\right|>s-C_2n^{-2/3}(x-y)^4-C_2\Bigg\}\\ &\subseteq & \Bigg\{n^{-1/3}\left|X_u^v-\E X_u^v\right|>3^{-1}s\Bigg\}\,.\end{aligned}$$ Let $U=T_n^{-1}(I\times \{0\})$ and $V=T_n^{-1}(J\times \{1\})$. For $u\in U,v\in V$, since $3^{-1}<S_{u,v}<3$, we can invoke the proofs of [@BSS14 Propositions $10.1$ and $10.5$]. Observe that, for Poissonian last passage percolation, [@BSS14 Corollary $9.1$] strengthens to $$\label{e:locc1} \P(|X_u^{u'}-\E X_u^{u'}|>\theta r^{1/3})\leq e^{-C_1\theta^{3/2}}\,.$$ Following the proofs of Proposition $10.1$ and $10.5$ of [@BSS14] verbatim, and using the above bound in  in place of Corollary $9.1$ of [@BSS14], one thus has for all $n,s$ large enough, $$\P\Bigg(\sup_{u\in U,v\in V}n^{-1/3}\left|X_u^v-\E X_u^v\right|>2^{-1}s\Bigg)\leq e^{-cs^{3/2}}\,.$$ Thus, for $n$ large enough, and $I$ and $J$ intervals of at most unit length contained in the interval of length $2C_0^{-1}s^{1/4}n^{1/6}$ centred at the origin, $$\label{e:01case} \P \bigg( \, \sup_{x \in I, y \in J} \Big\vert {\mathsf{W}}_{n;(x,0)}^{(y,1)} + (x-y)^2 \Big\vert > s \, \bigg) \, \leq \, C \exp \big\{ -c s^{3/2} \big\} \, .$$ We now make a first use of the boldface notation for applying results specified at the beginning of Section \[s:estinteg\]. For general $t_1<t_2$, set $\bm{n}=n{t_{1,2}},\bm{x}=x{t_{1,2}}^{-2/3},\bm{y}=y{t_{1,2}}^{-2/3}, \bm{I}={t_{1,2}}^{-2/3}I, \bm{J}={t_{1,2}}^{-2/3}J$ and $\bm{s}=s$ in . Recall that the boldface variables are those of Theorem \[t:unifmoddev\] and that these are written in terms of non-boldface parameters specified by the present context. From the hypothesis of Theorem \[t:unifmoddev\], $\bm{I}$ and $\bm{J}$ are intervals of at most unit length contained in $[-\bm{n}^{1/6},\bm{n}^{1/6}]$. Thus, applying and using the scaling principle , we get Theorem \[t:unifmoddev\]. The following lower bound on the tail of the polymer weight distribution follows from [@LM01 Theorem $1.3$] and . \[t:moddevlow\] There exist constants $c_2,s_0,n_0>0$ such that, for all $t_1<t_2$ with $n{t_{1,2}}>n_0$ and $s>s_0$, $$\P\left({t_{1,2}}^{-1/3}{\mathsf{W}}_{n;(0,t_1)}^{(0,t_2)}\geq s\right)\geq e^{-c_2s^{3/2}}.$$ Moving to unscaled coordinates, the transversal fluctuations for paths between $(0, 0)$ and $(n, n)$ around the interpolating line joining the two points were shown to be $n^{ 2/3+o(1)}$ with high probability in [@J00]. More precise estimates were established in [@BSS14]. However, the fluctuation of the geodesic at the point $(r,r)$ for any $r\leq n$ is only of the order $r^{2/3}$. This is the content of the next theorem which in essence is the scaled version of [@BSS17++ Theorem $2$] adapted for Poissonian LPP. Recall that, for $u,v\in \R^2$, $\Phi_{n;u}^v$ is the set of all polymers from $u$ to $v$, and $\l_u^v$ is the straight line joining $u$ and $v$. \[l:at0\]\[t:carsestimate2\] There exist positive constants $n_0, s_1, c$ such that for all $x,y,t_1,t_2\in \R$ with ${t_{1,2}}=t_2-t_1>0$ and $|x-y|\leq 2^{-1}n^{1/3}{t_{1,2}}$ and for all $n{t_{1,2}}\geq n_0, s\geq s_1$ and $t\in [t_1,t_2]$, $$\label{e:cars} \P\left(\max \left\{ \left|\rho(t)-\l_{(x,t_1)}^{(y,t_2)}(t)\right| : \rho\in \Phi_{n;(x,t_1)}^{(y,t_2)} \right\} \geq s\Big((t-t_1)\wedge (t_2-t)\Big)^{2/3} \right)\leq 2e^{-cs^3}.$$ Here $a\wedge b$ denotes $\min\{a,b\}$. [**Proof of Theorem \[t:carsestimate2\].**]{} First we prove the theorem when $t_1=0,t_2=1,x=0$. Observe that in this case it is enough to bound the probabilities of the events $$\left\{\left|\rho_{n;(0,0)}^{\leftarrow;(y,1)}(t)-\l_{(0,0)}^{(y,1)}(t)\right|\geq s\Big(t\wedge (1-t)\Big)^{2/3}\right\} \, \, \, \textrm{and} \, \, \, \left\{\left|\rho_{n;(0,0)}^{\rightarrow;(y,1)}(t)-\l_{(0,0)}^{(y,1)}(t)\right|\geq s\Big(t\wedge (1-t)\Big)^{2/3}\right\} \, ,$$ and use a union bound to obtain . We first prove an upper bound for the probability of the first of these two events. Also, first assume that $t\in [0,2^{-1}]$. To prove the bound in this case, we move to unscaled coordinates, and use [@BSS17++ Theorem $2$]. To this end, let $\Gamma:=\Gamma_{(0,0)}^{\leftarrow;(n+yn^{2/3},n-yn^{2/3})}$ be the leftmost geodesic, and $\mathcal{S}$ the straight line from $(0,0)$ to $(n+yn^{2/3},n-yn^{2/3})$. For $r\in [0,n+yn^{2/3}]$, let $\Gamma(r)$ and $\mathcal{S}(r)$ be such that $(r,\Gamma(r))\in \Gamma$ and $(r,\mathcal{S}(r))\in \mathcal{S}$. Now, for $r=nt$, $$\begin{aligned} \label{e:scalingfluct} &&\left\{\left|\rho_{n;(0,0)}^{\leftarrow;(y,{t_{1,2}})}(t)-\l_{(0,0)}^{(y,{t_{1,2}})}(t)\right|\geq st^{2/3}\right\}\\ &=&\left\{\left|n^{2/3}\rho_{n;(0,0)}^{\leftarrow;(y,1)}(rn^{-1})-n^{2/3}\l_{(0,0)}^{(y,1)}(rn^{-1})\right|\geq sr^{2/3}\right\}\nonumber\\ &\subseteq &\left\{\left|\Gamma(r')-\mathcal{S}(r')\right|\geq sr^{2/3}\right\}=:\mathsf{B} \, , \nonumber\end{aligned}$$ where $r'$ is such that the anti-diagonal line passing through $(r,r)$ intersects $\mathcal{S}$ at $(r',\mathcal{S}(r'))$. The last inclusion follows from the definition of the scaling map $T_n$ in . Since $|y|\leq 2^{-1}n^{1/3}$, $2^{-1}r\leq r'\leq 2r$. Thus, $$\mathsf{B}\subseteq \left\{ \left|\Gamma(r')-\mathcal{S}(r')\right|\geq 2^{-1}s(r')^{2/3}\right\}=:\mathsf{C}\,.$$ Thus it is enough to bound the probability of the event $\mathsf{C}$. This local fluctuation estimate for the leftmost geodesic in was proved for exponential directed last passage percolation in [@BSS17++ Theorem $2$ and Corollary $2.4$]. The proof goes through verbatim for the leftmost (and also the rightmost) geodesic in Poissonian last passage percolation. Moreover, the refined bounds of Theorem \[t:unifmoddev\] give corresponding improvements for Poissonian LPP: see [@BSS17++ Remark $1.5$]. This gives that, for some positive constants $n_0,r_0,s_0$, and for $n\geq n_0, r'\geq r'_0$ and $s\geq s_0$, $$\label{e:local} \P(\mathsf{C})\leq e^{-cs^3}\,.$$ However, observe that holds only when $r'\geq r'_0$. Now assume $r'\leq r'_0$, so that $r\leq r_0$, where $r_0=2r_0'$. Let the anti-diagonal passing through $(r,r)$ intersect the geodesic $\Gamma$ at $v$ and the line $\mathcal{S}$ at $w$. Clearly $\|v-(r,r)\|_2 \leq 2^{1/2} r$. Also, since $|y|\leq 2^{-1}n^{1/3}$, $$\|w-(r,r)\|_2= 2^{1/2} |y|rn^{-1/3}\leq r\,.$$ Thus, with $r=nt\leq r_0$, $$\left|\rho_{n;(0,0)}^{\leftarrow;(y,1)}(t)-\l_{(0,0)}^{(y,1)}(t)\right|=2^{-1/2}n^{-2/3}\|v-w\|_2 \leq 2^{-1}(2^{1/2}+1)n^{-2/3}r\leq 2r_0^{1/3}t^{2/3}\,.$$ Define $s_1=\max\{s_0,2r_0^{1/3}\}$. Then for $n\geq n_0,s\geq s_1$ and $t\in [0,2^{-1}]$, $$\P\left(\left|\rho_{n;(0,0)}^{\leftarrow;(y,1)}(t)-\l_{(0,0)}^{(y,1)}(t)\right|\geq st^{2/3}\right)\leq e^{-cs^3}\,.$$ For $t\in [2^{-1},1]$, we consider the reversed polymer and translate it by $-y$ so that its starting point is $(0,0)$, that is, $\rho'(v)=\rho_{n;(0,0)}^{\leftarrow;(y,1)}(1-v)-y$ for $v\in [0,1]$. Now we follow the same arguments as above to get the bound for the probability of the event $$\left\{\left|\rho_{n;(0,0)}^{\leftarrow;(y,1)}(t)-\l_{(0,0)}^{(y,1)}(t)\right|\geq s\Big(t\wedge (1-t)\Big)^{2/3}\right\}\,.$$ Since the same arguments work for the rightmost polymer $\rho_{n;(0,0)}^{\rightarrow;(y,1)}$, we get for $n\geq n_0,s\geq s_1$ and all $t\in [0,1]$, $$\label{e:fix} \P\left(\max \left\{ \left|\rho(t)-\l_{(0,0)}^{(y,1)}(t)\right| : \rho\in \Phi_{n;(0,0)}^{(y,1)} \right\} \geq s\Big(t\wedge (1-t)\Big)^{2/3} \right)\leq 2e^{-cs^3}.$$ Now for general $t_1<t_2$, set $\bm{n}=n{t_{1,2}}, \bm{y}=(y-x){t_{1,2}}^{-2/3},\bm{s}=s$ and $\bm{t}={t_{1,2}}^{-1}(t-t_1)$. Then from the hypothesis of Theorem \[t:carsestimate2\], $|\bm{y}|\leq 2^{-1}\bm{n}^{1/3}$ since $|y-x|\leq 2^{-1}n^{1/3}{t_{1,2}}$. Thus applying and using the scaling principle , we get the theorem. The following theorem bounds the transversal fluctuation of polymers; (recall the definitions in and ). The theorem essentially follows from [@BSS14 Theorem  $11.1$]; however, we replace the exponent in the upper bound with its optimal value. \[l:flucpoly\]\[t:transversal\]There exist positive constants $c$, $n_0$ and $k_0$ such that, for $t\in (0,1]$, $k\geq k_0$ and $n\geq n_0t^{-1}$, $$\P\left(\mathrm{TF}_{n;(0,0)}^{(0,t)}\geq kt^{2/3}\right)\leq 2e^{-ck^3}.$$ [**Proof.**]{} Because of , it is enough to bound the probabilities of the events $\left\{\mathrm{TF}\big(\rho_{n;(0,0)}^{\leftarrow;(0,t)}\big)\geq kt^{2/3}\right\}$ and $\left\{\mathrm{TF}\big( \rho_{n;(0,0)}^{\rightarrow;(0,t)} \big) \geq kt^{2/3}\right\}$ and use a union bound. We bound only the first event, the arguments for the second event being the same. Then, as in the proof of Theorem \[l:at0\], going to the unscaled coordinates, and defining $\Gamma=\Gamma_{(0,0)}^{\leftarrow;(nt,nt)}$, it is enough to show that $$\label{e:maxTF1} \P\left(\sup_{r\in [0,nt]}\left|\Gamma(r)-r\right|\geq k(nt)^{2/3}\right)\leq e^{-ck^3}\,.$$ From Theorem \[t:carsestimate2\], it is easy to see that there exist constants $c>0$ and $n_0,k_0>0$ such that, for all $k>k_0$ and $nt\geq n_0$, $$\P\left(\left|\Gamma\left(2^{-1}nt\right)-2^{-1}nt\right|>k(nt)^{2/3}\right)\leq e^{-ck^3}\,.$$ Using the above bound in place of [@BSS14 Lemma $11.3$], and following the rest of the proof of [@BSS14 Theorem $11.1$] verbatim, we get . Basic tools {#s:polorder} =========== Fundamental facts about ordering and concatenation of polymers will be used repeatedly in the proofs of the main theorems. Polymer concatenation and superadditivity of weights {#ss:pocat} ---------------------------------------------------- Let $n\in \N$ and $(x,t_1),(y,t_2)\in \R^2$ with $t_1<t_2$ and $|x-y|<n^{1/3}(t_2-t_1)$. (This condition ensures that $(x,t_1)\overset{n}\preceq (y,t_2)$, see Subsection  \[sss:comp\].) Let $u=T_n^{-1}(x,t_1)$ and $v=T_n^{-1}(y,t_2)$ and let $\zeta$ be an increasing path from $u$ to $v$. Let $\gamma=T_n(\zeta)$. We call $\gamma$ an *$n$-path*. We shall often consider $\gamma$ as a subset of $\R^2$, and call $(x,t_1)$ its *starting point* and $(y,t_2)$ its *ending point*. Moreover, similarly to the definition of the weight of a polymer in , we define the weight of an $n$-path as $$\label{e:defweightpath} n^{-1/3}\left(|\zeta|-2n{t_{1,2}}\right)\,,$$ where $|\zeta|$ denotes the energy of $\zeta$, that is, the number of points in $\Pi\setminus\{v\}$ that lie on $\zeta$. Now, let $(x,t_1),(y,t_2),(z,t_3)\in \R^2$ be such that $t_1<t_2<t_3$, $|x-y|< n^{1/3}(t_2-t_1)$ and $|y-z|<n^{1/3}(t_3-t_2)$, so that there exist polymers from $(x,t_1)$ to $(y,t_2)$; and from $(y,t_2)$ to $(z,t_3)$. Let $\rho_1$ be any polymer from $(x,t_1)$ to $(y,t_2)$, and $\rho_2$ any polymer from $(y,t_2)$ to $(z,t_3)$. The union of these two subsets of $\R^2$ is an $n$-path from $(x,t_1)$ to $(z,t_3)$. We call this $n$-path the concatenation of $\rho_1$ and $\rho_2$ and denote it by $\rho_1\circ \rho_2$. The weight of $\rho_1\circ \rho_2$ is ${\mathsf{W}}_{n;(x,t_1)}^{(y,t_2)}+{\mathsf{W}}_{n;(y,t_2)}^{(z,t_3)}$. This additivity is the reason that the endpoint $v$ was excluded from the definition of path energy in Section \[ss:defres\]. Again, let $n\in \N$ and $(x,t_1),(y,t_2),(z,t_3)\in \R^2$ be such that $t_1<t_2<t_3$ and $|x-y|< n^{1/3}(t_2-t_1)$ and $|y-z|<n^{1/3}(t_3-t_2)$. Then $$\label{e:superadd} {\mathsf{W}}_{n;(x,t_1)}^{(z,t_3)}\geq {\mathsf{W}}_{n;(x,t_1)}^{(y,t_2)}+{\mathsf{W}}_{n;(y,t_2)}^{(z,t_3)}\,.$$ Indeed, taking a polymer $\rho_1$ from $(x,t_1)$ to $(y,t_2)$ and a polymer $\rho_2$ from $(y,t_2)$ to $(z,t_3)$, the weight of $\rho_1\circ\rho_2$ is a lower bound on ${\mathsf{W}}_{n;(x,t_1)}^{(z,t_3)}$. Polymer ordering lemmas ----------------------- The first lemma roughly says that if two polymers intersect at two points during their lifetimes, then they are identical between these points. \[l:basic\]Let $n\in \N$ and $(x_1,t_1),(x_2,t_2),(y_1,s_1),(y_2,s_2)\in \R^2$ and $t,s\in \R$ be such that $t_1<t<s<s_1$, $t_2<t<s<s_2$, $|x_1-y_1|<n^{1/3}(s_1-t_1)$ and $|x_2-y_2|<n^{1/3}(s_2-t_2)$. Suppose that $\rho_{n;(x_1,t_1)}^{\leftarrow;(y_1,s_1)}$ and $\rho_{n;(x_2,t_2)}^{\leftarrow,(y_2,s_2)}$ intersect at two points $z_1=(x,t)$ and $z_2=(y,s)$. Then $\rho_{n;(x_1,t_1)}^{\leftarrow;(y_1,s_1)}$ and $\rho_{n;(x_2,t_2)}^{\leftarrow,(y_2,s_2)}$ are identical between $t$ and $s$. The same statement holds for the rightmost polymers. To simplify notation in the proof, we write $\rho_1=\rho_{n;(x_1,t_1)}^{\leftarrow;(y_1,s_1)}$ and $\rho_2=\rho_{n;(x_2,t_2)}^{\leftarrow,(y_2,s_2)}$. ![This illustrates Lemma \[l:porder\]. The points of the underlying Poisson process lying on a polymer are marked by dots, and the polymer is obtained by linearly interpolating between the points. The figure shows that both the paths cannot be leftmost polymers between their respective endpoints, since by joining the dashed lines, one obtains an alternative increasing path where the Poisson points between the intersecting points $z_1$ and $z_2$ in the two polymers are interchanged. []{data-label="f:PO"}](cropped_PO.pdf) [**Proof of Lemma \[l:basic\].**]{} First, for any polymer $\rho$, call a point $u\in \rho$ a *Poisson* point of $\rho$ if $T_n^{-1}(u)\in \Pi\cap\Gamma$, where $\Gamma$ is the geodesic $T_n^{-1}(\rho)$ and $\Pi$ is the underlying unit rate Poisson point process. Also, for $r_1,r_2\in \rho$, let $\rho[r_1,r_2]$ denote the part of the polymer between the points $r_1$ and $r_2$, and let $\#\rho[r_1,r_2]$ denote the number of Poisson points that lie in $\rho[r_1,r_2]$. We first claim that $\#\rho_1[z_1,z_2]=\#\rho_2[z_1,z_2|$ where $z_1$ and $z_2$ appear in the lemma’s statement. For, if not, without loss of generality assume that $\#\rho_1[z_1,z_2]<\#\rho_2[z_1,z_2|$ and let $u_1$ and $v_1$ be the Poisson points of $\rho_1$ immediately before $z_1$ and immediately after $z_2$; and let $u_2$ and $v_2$ be the Poisson points of $\rho_2$ immediately after $z_1$ and immediately before $z_2$: see Figure \[f:PO\]. Then joining $u_1$ to $u_2$ and $v_1$ to $v_2$ (shown in the figure by dashed lines), one gets an alternative path $\rho'$ between $(x_1,t_1)$ and $(y_1,s_1)$ that has more Poisson points than $\rho_1$, thereby contradicting that $\rho_1$ is a polymer between $(x_1,t_1)$ and $(y_1,s_1)$. Thus, $\#\rho_1[z_1,z_2]=\#\rho_2[z_1,z_2|$. Since both $\rho_1$ and $\rho_2$ are leftmost polymers between their respective endpoints, we see that $\rho_1[z_1,z_2]=\rho_2[z_1,z_2]$. This proves the lemma. The next result roughly says that two polymers that begin and end at the same heights, with the endpoints of one to the right of the other’s, cannot cross during their shared lifetime. \[l:porder\] Fix $n\in \N$. Consider points $(x_1,t_1),(x_2,t_1),(y_1,t_2),(y_2,t_2)\in \R^2$ such that $t_1 < t_2$, $x_1\leq x_2$, $y_1\leq y_2$, $|x_1-y_1|<n^{1/3}(t_2-t_1)$ and $|x_2-y_2|<n^{1/3}(t_2-t_1)$. Then $\rho_{n;(x_1,t_1)}^{\leftarrow;(y_1,t_2)}(t)\leq \rho_{n;(x_2,t_1)}^{\leftarrow;(y_2,t_2)}(t)$ and $\rho_{n;(x_1,t_1)}^{\rightarrow;(y_1,t_2)}(t)\leq \rho_{n;(x_2,t_1)}^{\rightarrow;(y_2,t_2)}(t)$ for all $t\in [t_1,t_2]$. Let $\rho_1=\rho_{n;(x_1,t_1)}^{\leftarrow;(y_1,t_2)}$ and $\rho_2=\rho_{n;(x_1,t_1)}^{\rightarrow;(y_1,t_2)}$. [**Proof of Lemma \[l:porder\].**]{} Supposing otherwise, there exists $z=(x,y)\in \rho_2$ such that $x<\rho_1(y)$. But then there exist $z_1,z_2\in \rho_1 \cap\rho_2$ straddling the point $z$. By Lemma \[l:basic\], $\rho_1[z_1,z_2]=\rho_2[z_1,z_2]$, and hence $z\in \rho_1\cap \rho_2$, a contradiction. By ordering, a polymer whose endpoints are straddled between those of a pair of polymers becomes sandwiched between those polymers. \[c:po\]Fix $n\in \N$. Consider points $(x_1,t_1),(x_2,t_1),(x_3,t_1),(y_1,t_2),(y_2,t_2),(y_3,t_2)\in \R^2$ such that $t_1 < t_2$, $x_1\leq x_2\leq x_3$, $y_1\leq y_2\leq y_3$ and $|x_i-y_i|<n^{1/3}(t_2-t_1)$ for $i=1,2,3$. Let $t\in (t_1,t_2)$. Let $\rho_i=\rho_{n;(x_i,t_1)}^{\leftarrow;(y_i,t_2)}$ for $i=1,2,3$. Then $$|\rho_2(t)-\rho_2(t_1)|\leq \max_{i\in\{1,3\}}|\rho_i(t)-\rho_i(t_1)|+\max_{i\in\{1,3\}}|x_i-x_2|\,.$$ The same result holds for rightmost polymers. [**Proof.**]{} By Lemma \[l:porder\], $$\rho_1(t)\leq \rho_2(t)\leq \rho_3(t)\,.$$ The result now follows immediately. Exponent pair $(2/3,1/3)$ for a single polymer: Proof of Theorem \[t:main1\] {#s:2/3,1/3} ============================================================================ In this section, we show that the sequence $\big\{\rho^{\leftarrow}_n : n \in \N \big\}$ of leftmost $n$-polymers from $(0,0)$ to $(0,1)$ is tight, and any weak limit is [Hölder ]{}$2/3-$-continuous with a polylogarithmic correction of order $1/3$. The main two ingredients in this proof are the local regularity estimate Theorem \[t:carsestimate2\] and the polymer ordering Lemma \[l:porder\]. First, we bound the fluctuation of the polymer near any given point $z\in [0,1]$. \[l:boundatz\] There exist positive constants $n_0,s_1$ and $c$ such that, for all $n\geq n_0,s\geq s_1$, $z\in [0,1]$ and $0\leq t\leq 1-z$, $$\label{e:boundatz} \P \left( |\rho_n^{\leftarrow}(z+t)-\rho_n^{\leftarrow}(z)|\geq st^{2/3} \right) \leq 10t^{-2/3}e^{-cs^3}.$$ The same statement holds for $\rho_n^{\rightarrow}$. As we now explain, the proposition will be proved by reducing to the case that $z=0$, when the result follows from Theorem \[l:at0\]. For any fixed $z\in (0,1)$, Theorem \[l:at0\] again guarantees that the polymer $\rho_n^{\leftarrow}$ is at distance at most $s$ from the point $(0,z)$ with probability at least $1-e^{-cs^3}$. We break the horizontal line segment of length $2s$ centred at $(0,z)$ into a sequence of consecutive intervals of length $2^{-1}st^{2/3}$, and consider the leftmost polymers starting from each of these endpoints and ending at $(0,1)$, as in Figure \[f:Fig1\]. Due to the Corollary \[c:po\] of the polymer ordering Lemma \[l:porder\], a big fluctuation of $\rho_n^{\leftarrow}$ between times $z$ and $z+t$ creates a big fluctuation for one of the polymers starting from these deterministic endpoints. The probability of the latter fluctuations is controlled via Theorem \[l:at0\] and since the number of these polymers is of the order of $t^{-2/3}$, a union bound gives .  \ [**Proof of Proposition \[l:boundatz\].**]{} First observe that for $s> (nt)^{1/3}$, the probability in is zero by the definition of the scaling map $T_n$ in and the geodesics being increasing paths. Hence we assume that $s\leq (nt)^{1/3}$. Fix $s\leq (nt)^{1/3}$ and $z\in [0,1]$. For $t\geq 8^{-3}$, $$\left\{|\rho_n^{\leftarrow}(z+t)-\rho_n^{\leftarrow}(z)|\geq st^{2/3}\right\}\subseteq \left\{|\rho_n^{\leftarrow}(z+t)-\rho_n^{\leftarrow}(z)|\geq 8^{-2}s\right\}\subseteq\left\{\mathrm{TF}_{n;(0,0)}^{(0,1)}\geq 2^{-1}8^{-2}s\right\}\,,$$ where $\mathrm{TF}_{n;(0,0)}^{(0,1)}$ is defined in . Hence, applying Theorem \[l:flucpoly\] with the parameter specifications $\bm{t}=1$ and $\bm{k}=2^{-1}8^{-2}s$, we get that holds for all $n,s$ large enough. Hence we assume that $t\leq 8^{-3}$. Also, let us assume for now that $z\in [0,2^{-1}]$. Let $L$ be the line segment $[-s,s]\times\{z\}$. Let $\mathsf{E}$ be the event that $\rho_n^{\leftarrow}$ passes through $L$. By Theorem \[l:at0\] with $\bm{n}=n, \bm{t}=z, \bm{x}=0,\bm{y}=0, \bm{s}=s,\bm{t_1}=0$ and $\bm{t_2}=1$, we have that, for $n\geq n_1$ and $s\geq s_1$, $$\P(\mathsf E)\geq 1-2e^{-cs^3}.$$ Now, we divide $L$ into $\lceil{4t^{-2/3}}\rceil$-many adjacent intervals of length at most ${2^{-1}st^{2/3}}$, and let $(x_i,z), i=0,1,2,\cdots,\lceil4t^{-2/3}\rceil$ be the endpoints of these intervals, i.e., $$x_i=-s+2^{-1}ist^{2/3} \quad \mbox{for } i=0,1,2,\cdots, \lceil4t^{-2/3}\rceil.$$ Let $\rho_n^{(i)}:=\rho_{n;(x_i,z)}^{\leftarrow;(0,1)}$ be the leftmost polymer from $(x_i,z)$ to $(0,1)$. ![The proof of Proposition \[l:boundatz\] is illustrated here. We mark the line segment $L$ with a number of equally spaced points. As the leftmost polymer from $(0,0)$ to $(0,1)$ passes between two such points on the line $L$, it is, in view of polymer ordering, sandwiched between the two leftmost polymers, shown as dotted lines, originating from those points and ending at $(0,1)$. Hence it is sufficient to bound the fluctuations of the polymers originating from these equally spaced points on $L$.[]{data-label="f:Fig1"}](Thm1Figv2.pdf){width="80.00000%"} By Corollary \[c:po\], on $\mathsf{E}$, $$\begin{aligned} \label{e:polord} \left|\rho_n^{\leftarrow}(z+t)-\rho_n^{\leftarrow}(z)\right|\leq \max_{i\in \llbracket 0, \lceil4t^{-2/3}\rceil\rrbracket}\left|\rho_n^{(i)}(z+t)-\rho_n^{(i)}(z)\right| +2^{-1}st^{2/3}.\end{aligned}$$ Also, for any fixed $i \in \llbracket 0, \lceil4t^{-2/3}\rceil\rrbracket$, let $\l^{(i)}=\l_{(x_i,z)}^{(0,1)}$ be the straight line segment joining $(x_i,z)$ and $(0,1)$. Then, since $z\in [0,2^{-1}]$ and $t\leq 8^{-3}$, for any $i\in 0,1,2,\cdots,\lceil4t^{-2/3}\rceil$, $$\left|\l^{(i)}(z)-\l^{(i)}(z+t)\right| \leq \frac{st}{1-z} \leq 2st \leq 4^{-1}st^{2/3} \, .$$ Since $\rho_n^{(i)}(z)=\l^{(i)}(z)=x_i$, $$\begin{aligned} \left|\rho_n^{(i)}(z+t)-\rho_n^{(i)}(z)\right| &\leq& \left|\rho_n^{(i)}(z+t)-\l^{(i)}(z+t)\right|+|\l^{(i)}(z+t)-\l^{(i)}(z)|\\ &\leq& \left|\rho_n^{(i)}(z+t)-\l^{(i)}(z+t)\right|+{4^{-1}st^{2/3}}\,.\end{aligned}$$ Thus, on the event $\mathsf{E}$, by , $$\left|\rho_n^{\leftarrow}(z+t)-\rho_n^{\leftarrow}(z)\right|\leq \max_{i\in \llbracket 0, \lceil4t^{-2/3}\rceil\rrbracket}\left|\rho_n^{(i)}(z+t)-\l^{(i)}(z+t)\right|+ \tfrac{3}{4} st^{2/3} \, .$$ From here, it follows by taking a union bound that $$\begin{aligned} &&\P\left(\left|\rho_n^{\leftarrow}(z+t)-\rho_n^{\leftarrow}(z)\right|\geq st^{2/3}\right)\\ &\leq& \P(\mathsf{E}^c)+\sum_{i=0}^{\lceil 4t^{-2/3}\rceil}\P\left(\left|\rho_n^{(i)}(z+t)-\l^{(i)}(z+t)\right|\geq 4^{-1}st^{2/3} \right)\\ &\leq & 10t^{-2/3}e^{-cs^3},\end{aligned}$$ for some absolute positive constant $c$ and all $n\geq 2n_0$. Here the last inequality follows by applying Theorem \[l:at0\] to each of the polymers $\rho^{(i)}$. For given $i$, set the parameters $\bm{n}=n,\bm{t_1}=z,\bm{t_2}=1,\bm{t}=t+z,\bm{x}= -s+2^{-1}ist^{2/3},\bm{y}=0$ and $\bm{s}=4^{-1}s$. Since $z\in [0, 2^{-1}]$ and $s\leq (nt)^{1/3}$, we have that $|\bm{x}-\bm{y}|\leq s\leq n^{1/3}t^{1/3}\leq 8^{-1}n^{1/3}\leq 4^{-1}\bf{n}^{1/3}\bf{{t_{1,2}}}$. Thus one can apply Theorem \[l:at0\] to get the above inequality for all $\bm{n{t_{1,2}}}\geq 2^{-1}n\geq n_0$. For $z\in [2^{-1},1]$, define the reversed polymer $\widehat{\rho}_n^{\leftarrow}$ by $\widehat{\rho}_n^{\leftarrow}(a)={\rho}_n^{\leftarrow}(1-a)$ for $a\in [0,1]$, and follow the above argument. Next we show the tightness of the members of the sequence $\{\rho_n^{\leftarrow}\}_{n\in \N}$ as elements in the space $(C[0,1],\|\cdot\|_\infty)$. We prove that Proposition \[l:boundatz\] guarantees that Kolmogorov-Chentsov’s tightness criterion is satisfied. [**Proof of Theorem \[t:main1\]$(a)$.**]{} Fix $n\geq n_0$ and any $\lambda>0$. Fix $t\in (0,1]$ small enough that $\lambda t^{-2/3}\geq s_1$, where $n_0$ and $s_1$ are as in Proposition \[l:boundatz\]. Also fix some $M\in \N$ large enough that $2M-2/3>1$. Then it follows from Proposition \[l:boundatz\] that for any $z,z'\in [0,1]$ with $|z-z'|=t$, $$\label{e:KolCh} \P\left(|\rho_n^{\leftarrow}(z')-\rho_n^{\leftarrow}(z)|\geq \lambda\right)\leq 10t^{-2/3}e^{-c(\lambda^3 t^{-2})}\leq K_M\lambda^{-3M} t^{2M-2/3}=K_M\lambda^{-3M} |z'-z|^{2M-2/3},$$ where $K_M:=\sup_{x\geq 0}x^Me^{-cx}<\infty$. Since $2M-2/3>1$, by Kolmogorov-Chentsov’s tightness criterion (see for example [@Dur10 Theorem $8.1.3$]), it follows that the sequence $\{\rho_n^{\leftarrow}\}_{n\in \N}$ is tight in $(C[0,1],\|\cdot\|_\infty)$. Modulus of continuity --------------------- Here we prove Theorem \[t:main1\](b), thus finding the modulus of continuity for any weak limit of a weakly converging subsequence of $\{\rho_n^{\leftarrow}\}_{n\in \N}$. We will follow the arguments used to derive the Kolmogorov continuity criterion, where one infers [Hölder ]{}continuity of a stochastic process from moment bounds on the difference of the process between pairs of times. Thus we introduce the set of dyadic rationals $$D=\bigcup_{i=0}^\infty 2^{-i}\Z \, .$$ Next is the first step towards proving the modulus of continuity. \[l:dyad\] Let $\rho_*^{\leftarrow}$ be the weak limit of a weakly converging subsequence of $\{\rho_n^{\leftarrow}\}_{n\in \N}$. Then there exists a universal positive constant $C$ (not depending on the particular weak limit $\rho_*^{\leftarrow}$) such that, almost surely, for some random $m_0(\omega)\in \N$ and for all $s,t\in D\cap [0,1]$ with $|t-s|\leq 2^{-m_0(\omega)}$, $$|\rho_*^{\leftarrow}(t)-\rho_*^{\leftarrow}(s)|\leq C(t-s)^{2/3}\left(\log({t-s})^{-1}\right)^{1/3} .$$ [**Proof.**]{} For $m\in \N$, let $S_m$ be the set of all intervals of the form $[j2^{-m},(j+1)2^{-m}]$, for $j\in\{0,1,2,\cdots,2^m-1\}$. Fix $c_0>(\frac{5}{3c})^{1/3}$, where $c$ is the constant in Proposition \[l:boundatz\]. Writing $\Rightarrow$ for convergence in distribution, let $\{\rho_{n_k}^{\leftarrow}\}_{k\in \N}$ be a subsequence of $\{\rho^{\leftarrow}_n\}_{n\in \N}$ such that $\rho_{n_k}^{\leftarrow}\Rightarrow \rho_*^{\leftarrow}$ as random variables in $(C[0,1],\|\cdot\|_\infty)$. Since for $a,b\in [0,1]$, the map $\tau_{a,b}$ defined by $(C[0,1],\|\cdot\|_\infty)\mapsto (\R,|\cdot|):f\mapsto |f(a)-f(b)|$ is continuous, $$\mathsf{U}:=\bigcup\left\{\tau^{-1}_{(j+1)2^{-m},j2^{-m}}\left(c_02^{-\frac{2m}{3}}\left(\log 2^m\right)^{1/3},\infty\right):j=0,1,\cdots,2^m-1\right\}$$ is an open set. Thus, by the Portmanteau theorem, $$\begin{aligned} \label{e:main1} &&\P\Bigg(\sup_{j\in \{0,1,\cdots,2^m-1\}}|\rho_*^{\leftarrow}((j+1)2^{-m})-\rho_*^{\leftarrow}(j2^{-m})|>c_02^{-\frac{2m}{3}}\left(\log 2^m\right)^{1/3}\Bigg)\\ &\leq & \liminf_k \, \P\Bigg(\sup_{j\in \{0,1,\cdots,2^m-1\}}|\rho_{n_k}^{\leftarrow}((j+1)2^{-m})-\rho_{n_k}^{\leftarrow}(j2^{-m})|>c_02^{-\frac{2m}{3}}\left(\log 2^m\right)^{1/3}\Bigg)\nonumber\\ &\leq & \limsup _n \, \P\Bigg(\sup_{j\in \{0,1,\cdots,2^m-1\}}|\rho_{n}^{\leftarrow}((j+1)2^{-m})-\rho_{n}^{\leftarrow}(j2^{-m})|>c_02^{-\frac{2m}{3}}\left(\log 2^m\right)^{1/3}\Bigg)\nonumber\,.\end{aligned}$$ Now, for all $m$ large enough that $\left(\log 2^m\right)^{1/3}\geq s_1$, where $s_1$ is as in Proposition \[l:boundatz\], and all $n\geq n_0$, applying Proposition \[l:boundatz\] and a union bound, $$\begin{aligned} &&\P \, \Bigg(\sup_{j\in \{0,1,\cdots,2^m-1\}}|\rho_n^{\leftarrow}((j+1)2^{-m})-\rho_n^{\leftarrow}(j2^{-m})|>c_02^{-\frac{2m}{3}}\left(\log 2^m\right)^{1/3}\Bigg)\nonumber\\ &\leq & 10\cdot 2^m\left(\frac{1}{2^{m}}\right)^{c_0^3c-2/3}\leq 10\left(\frac{1}{2^{m}}\right)^{c_0^3c-5/3}.\nonumber\end{aligned}$$ Hence, from , $$\P \, \Bigg(\sup_{j\in \{0,1,\cdots,2^m-1\}}|\rho_*^{\leftarrow}((j+1)2^{-m})-\rho_*^{\leftarrow}(j2^{-m})|>c_02^{-\frac{2m}{3}}\left(\log 2^m\right)^{1/3}\Bigg)\leq 10\cdot 2^{-m(c_0^3c-5/3)}\,.$$ As the right hand side is summable in $m$ (by the choice of $c_0$ made at the beginning of the proof), the Borel-Cantelli lemma implies that there exists a null set $N_0$, such that, for each $\omega \notin N_0$, there is some $m_0(\omega)$ for which $m\geq m_0(\omega)$ entails that $$\label{e:dyad} |\rho_*^{\leftarrow}(t)-\rho_*^{\leftarrow}(s)|\leq c_0(t-s)^{2/3}\left(\log({t-s})^{-1}\right)^{1/3} \quad \mbox{ for all } [s,t]\in S_m \, .$$ Now, let $\omega \notin N_0$ and $s,t\in D\cap [0,1]$ be such that $|s-t|\leq 2^{-m_0(\omega)}$. Let $m=m(s,t)$ be the greatest integer such that $|s-t|\leq 2^{-m}$; then clearly, $m\geq m_0(\omega)$. Also, consider the binary expansions of $s$ and $t$: $$s=s_0+\sum_{j>m} \sigma_j2^{-j}, \quad t=t_0+\sum_{j>m} \tau_j2^{-j},$$ where $\sigma_j,\tau_j\in \{0,1\}$, and each of the sequences is eventually zero. Either $s_0=t_0$ or $[s_0,t_0]\in S_m$. Moreover, for $n\geq 1$, let $$s_n=s_0+\sum_{m<j\leq m+n} \sigma_j2^{-j}.$$ Then, for $n\geq 1$, either $s_n=s_{n-1}$ or $[s_{n-1},s_n]\in S_{m+n}$. Since $m\geq m_0(\omega)$, by , $$|\rho_*^{\leftarrow}(t_0)(\omega)-\rho_*^{\leftarrow}(s_0)(\omega)|\leq c_02^{-\frac{2m}{3}}\left(\log 2^{m}\right)^{1/3}.$$ Also, $$\begin{aligned} |\rho_*^{\leftarrow}(s)(\omega)-\rho_*^{\leftarrow}(s_0)(\omega)|\leq \sum_{n=1}^\infty|\rho_*^{\leftarrow}(s_n)(\omega)-\rho_*^{\leftarrow}(s_{n-1})(\omega)|&\leq & \sum_{n=1}^\infty c_02^{-\frac{2(m+n)}{3}}\left(\log 2^{m+n}\right)^{1/3}\\ &\leq & C_12^{-\frac{2(m+1)}{3}}\left(\log 2^{m+1}\right)^{1/3},\end{aligned}$$ and similarly $$|\rho_*^{\leftarrow}(t)(\omega)-\rho_*^{\leftarrow}(t_0)(\omega)|\leq C_2 2^{-\frac{2(m+1)}{3}}\left(\log 2^{m+1}\right)^{1/3},$$ for some absolute constants $C_1$ and $C_2$. Hence, $$|\rho_*^{\leftarrow}(t)-\rho_*^{\leftarrow}(s)|\leq |\rho_*^{\leftarrow}(t)-\rho_*^{\leftarrow}(t_0)|+|\rho_*^{\leftarrow}(t_0)-\rho_*^{\leftarrow}(s_0)|+|\rho_*^{\leftarrow}(s)-\rho_*^{\leftarrow}(s_0)|\leq C2^{-\frac{2m}{3}}\left(\log 2^{m}\right)^{1/3}.$$ Since by definition $2^{-m-1}\leq |s-t|\leq 2^{-m}$, the result follows. [**Proof of Theorem \[t:main1\]$(b)$.**]{} For any $s,t\in [0,1]$ satisfying $s<t$ and $|s-t|\leq 2^{-m_0(\omega)}$, choose $s_k,t_k\in D\cap [s,t]$ such that $s_k\searrow s$ and $t_k\nearrow t$. Then, since $|s_k-t_k|\leq |s-t|\leq 2^{-m_0(\omega)}$, by Lemma \[l:dyad\], $$|\rho_*^{\leftarrow}(t_k)-\rho_*^{\leftarrow}(s_k)|\leq C(t_k-s_k)^{2/3}\left(\log({t_k-s_k})^{-1}\right)^{1/3} .$$ Since $\rho_*^{\leftarrow}(t_k)(\omega)\rightarrow \rho_*^{\leftarrow}(\omega)$ and $\rho_*^{\leftarrow}(s_k)(\omega)\rightarrow \rho_*^{\leftarrow}(s)(\omega)$, the theorem follows by taking the limit as $k\rightarrow \infty$. The same argument applies without any change for the rightmost polymers as well. Exponent pair $(2/3,1/3)$ for maximum fluctuation over short polymers:\ Proof of Theorem \[t:flucshortpol\] {#s:shortpol} ======================================================================= In this section, we shall prove Theorem \[t:flucshortpol\]. It is the upper bound that is the more subtle. Recall the notation of transversal fluctuations from and , $\mathrm{AdEndPair}_n(t)$ from and $\mathrm{MTF}_n(t)$ from . Here is the idea behind the proof. Proposition \[p:lowbndTF\] offers a lower bound on the transversal fluctuation of a polymer between two given points. By considering order-$t^{-1}$ endpoint pairs with disjoint intervening lifetimes of length $t$, we obtain a collection of independent opportunities for the fluctuation lower bound to occur. By tuning the probability of the individual event to have order $t$, at least one among the constituent events typically does occur, and the lower bound in Theorem \[t:flucshortpol\] follows. On the other hand, suppose that a big swing in the unit order region happens between a certain endpoint pair, with an intervening duration, or height difference, of order $t$. Members of the endpoint pair may be exceptional locations when viewed as functions of the underlying Poisson point field, both in horizontal and vertical coordinate. Thus, the upper bound in Theorem \[t:flucshortpol\] does not follow directly from a union bound of a given endpoint estimate over elements in a discrete mesh, since such a mesh may not capture the exceptional endpoints. However, polymer ordering forces exceptional behaviour to become typical and to occur between an endpoint pair in a discrete mesh. To see this, assume that the original polymer between exceptional endpoints makes a big left swing. (Figure \[f:shortpoly\] illustrates the argument.) We take a discrete mesh endpoint pair whose lifetime includes that of the original polymer but has the same order $t$, and whose lower and upper points lie to the left of the original endpoint locations, about halfway between these and the leftmost coordinate visited by the original polymer. Then we consider the leftmost mesh polymer at the beginning and ending times of the original polymer. If the mesh polymer is to the right of the original polymer at any of these endpoints, then the mesh polymer has already made a big rightward swing at one of these endpoints. If, on the other hand, the mesh polymer is to the left of the original polymer at both the endpoints of the original polymer, then by polymer ordering Lemma \[l:porder\], the mesh polymer cannot cross the original polymer during the latter’s lifetime. Hence the big left swing of the original polymer forces a significant left swing for the mesh polymer as well. ![The figure illustrates the proof of the upper bound in Theorem \[t:flucshortpol\]. If the leftmost polymer between $(u,t_1)$ and $(v,t_2)$ (shown in red) makes a huge leftward fluctuation and the leftmost polymer between points $e_{i,j}^{(1)}$ and $f_{i,j}^{(1)}$ (shown in blue) is to the left of $u$ and $v$ at $t_1$ and $t_2$ respectively, then the blue polymer stays to the left of the red polymer between times $t_1$ and $t_2$ by polymer ordering. Thus the big left fluctuation transmits from the red to the blue polymer. If, however, the blue polymer reaches to the right of either $u$ or $v$, then it creates a big right fluctuation for the blue polymer. Thus by bounding the fluctuations of a small number of polymers between deterministic endpoints, one can bound the fluctuation between all admissible endpoint pairs.[]{data-label="f:shortpoly"}](shortpoly.pdf)  \ [**Proof of Theorem \[t:flucshortpol\].**]{} The lower bound follows in a straightforward way from Proposition \[p:lowbndTF\]. For any $t\in (0,1)$ and $i\in \left\{0,1,2,\cdots,\left[t^{-1}\right]-1\right\}$, define $$\mathsf{F}_{i,t,n}=\left\{\mathrm{TF}_{n;(0,it)}^{(0,(i+1)t}\geq ct^{2/3}\big(\log t^{-1}\big)^{1/3}\right\}.$$ For given such $(t,i)$, we apply Proposition \[p:lowbndTF\] with parameter settings $\bm{n}=n,\bm{t_1}=it,\bm{t_2}=(i+1)t$ and $\bm{s}=c(\log t^{-1}\big)^{1/3}$, to find that, when $c(\log t^{-1})^{1/3}\geq s_0$ and $n\geq \max\{\alpha_0^{-3}c^3t^{-1}\log t^{-1},n_0t^{-1}\}$, $$\P(\mathsf{F}_{i,t,n})\geq e^{-c^*c^3\log t^{-1}}=t^{c^*c^3 }\,,$$ where the proposition specifies the quantities $\alpha_0,n_0$ and $s_0$. Thus, for all $t\leq e^{-(c^{-1}s_0)^3}$ and $i\in \left\{0,1,2,\cdots,\left[t^{-1}\right]-1\right\}$, $$\begin{aligned} \liminf_n t^{-1}\P(\mathsf{F}_{i,t,n})=\liminf_n t^{-1}\P(\mathsf{F}_{0,t,n}) \geq t^{c^*c^3 - 1}\,.\end{aligned}$$ By choosing $c>0$ small enough that $c^*c^3<1$, one has $\liminf_n t^{-1}\P(\mathsf{F}_{0,t,n})\rightarrow \infty$ as $t \searrow 0$. For such $c>0$, using the definition (\[e:defMTF\]) of $\mathrm{MTF}_n(t)$ and independence of the events $\mathsf{F}_{i,t,n}$ for $i\in \left\{0,1,2,\cdots,\left[t^{-1}\right]-1\right\}$, $$\P\left({\mathrm{MTF}_n(t)}{t^{-2/3}\big(\log t^{-1}\big)^{-1/3}}<c\right) \, \leq \, \P\left(\bigcap_{i=0}^{[t^{-1}]-1}\mathsf{F}_{i,t,n}^c\right)=\prod_{i=0}^{[t^{-1}]-1}\P\left(\mathsf{F}_{i,t,n}^c\right)\,.$$ Thus, $$\begin{aligned} &&\limsup_n \, \P\left({\mathrm{MTF}_n(t)}{t^{-2/3}\big(\log t^{-1}\big)^{-1/3}}<c\right)\\ &\leq&\limsup_n \, \Big(1-\P\big(\mathsf{F}_{0,t,n}\big) \Big)^{[t^{-1}]}\leq \limsup_n \, \exp\left\{ -\left[t^{-1}\right]\P(\mathsf{F}_{0,n})\right\} \, \rightarrow \, 0\,,\end{aligned}$$ the latter convergence as $t\searrow 0$. Now we show the upper bound. Fix $t\in (0,1]$ small enough that $\psi t\leq t^{2/3}$, where the parameter $\psi$ appears in the definition  of $\mathrm{AdEndPair}_n(t)$. For any $i=0,1,2,\ldots, \lceil t^{-1}\rceil$ and $j\in \left\llbracket-\left\lceil t^{-2/3}\big(\log t^{-1}\big)^{-1/3}\right\rceil, \left\lceil t^{-2/3}\big(\log t^{-1}\big)^{-1/3}\right\rceil \right\rrbracket$, define the rectangle $\mathcal{A}_{i,j}$ with lower-left corner $\left((j-1)t^{2/3}\big(\log t^{-1}\big)^{1/3},it\right)$, width $2 t^{2/3}\big(\log t^{-1}\big)^{1/3}$ and height $2t$. Figure \[f:shortpoly\] illustrates this rectangle and the arguments that follow. Let $C > 0$ be an even integer whose value will later be specified. For such $i,j$ as above, define planar points $$\begin{aligned} e_{i,j}^{(1)}&:=&\left((j-2^{-1}C)t^{2/3}\big(\log t^{-1}\big)^{1/3},it\right),\quad f_{i,j}^{(1)}:=\left((j-2^{-1}C)t^{2/3}\big(\log t^{-1}\big)^{1/3},(i+2)t\right),\\ e_{i,j}^{(2)}&:=&\left((j+2^{-1}C)t^{2/3}\big(\log t^{-1}\big)^{1/3},it\right),\quad f_{i,j}^{(2)}:=\left((j+2^{-1}C)t^{2/3}\big(\log t^{-1}\big)^{1/3},(i+2)t\right).\end{aligned}$$ Then we claim that, whatever the value of $C > 0$, $$\label{e:imp} \mathsf{B}_{i,j}:=\left\{\sup\left\{\mathrm{TF}_{n;(x_1,y_1)}^{(x_2,y_2)}:(x_1,y_1),(x_2,y_2)\in \mathcal{A}_{i,j},y_2>y_1\right\}>Ct^{2/3}\big(\log t^{-1}\big)^{1/3}\right\}\subseteq \mathsf{D}^{(1)}_{i,j}\cup \mathsf{D}^{(2)}_{i,j},$$ where $$\mathsf{D}^{(1)}_{i,j}:=\left\{\mathrm{TF}_{n;e_{i,j}^{(1)}}^{f_{i,j}^{(1)}}\ge(2^{-1}C-1)t^{2/3}\big(\log t^{-1}\big)^{1/3}\right\}$$ and $$\mathsf{D}^{(2)}_{i,j}:=\left\{\mathrm{TF}_{n;e_{i,j}^{(2)}}^{f_{i,j}^{(2)}}\ge(2^{-1}C-1)t^{2/3}\big(\log t^{-1}\big)^{1/3}\right\}.$$ To see , define the vertical lines: $$L_2 = \big\{ x = (j-C+1)t^{2/3}\big(\log t^{-1}\big)^{1/3} \big\} \, \, \, \, \textrm{and} \, \, \, \, L_2' = \big\{ x = (j+C-1)t^{2/3}\big(\log t^{-1}\big)^{1/3} \big\} \, .$$ Then, on the event $\mathsf{B}_{i,j}$, there exists a pair of points $(u,t_1),(v,t_2)\in \mathcal{A}_{i,j}$ such that either $\rho_{n;(u,t_1)}^{\leftarrow;(v,t_2)}$ intersects $L_2$ or $\rho_{n;(u,t_1)}^{\rightarrow;(v,t_2)}$ intersects $L_2'$. We now show that, when $\rho_{n;(u,t_1)}^{\leftarrow;(v,t_2)}$ intersects $L_2$, the event $\mathsf{D}_{i,j}^{(1)}$ occurs. Let $$\rho:=\rho_{n;e_{i,j}^{(1)}}^{\leftarrow;f_{i,j}^{(1)}}\,.$$ Let $\l_{i,j}^{(1)}$ be the line segment joining $e_{i,j}^{(1)}$ and $f^{(1)}_{i,j}$. If $\rho(t_1)>u$, then $$\rho(t_1)-\l_{i,j}^{(1)}(t_1)\geq (j-1)t^{2/3}\big(\log t^{-1}\big)^{1/3}-(j-2^{-1}C)t^{2/3}\big(\log t^{-1}\big)^{1/3}\geq (2^{-1}C-1)t^{2/3}\big(\log t^{-1}\big)^{1/3}\,,$$ and thus $\mathsf{D}_{i,j}^{(1)}$ holds. Similarly, if $\rho(t_2)>v$, then $\mathsf{D}_{i,j}^{(1)}$ holds. Now assume that $\rho(t_1)<u$ and $\rho(t_2)<v$. Polymer ordering Lemma \[l:porder\] then implies that $\rho(t)\leq \rho_{n;(u,t_1)}^{\leftarrow;(v,t_2)}(t)$ for all $t\in [t_1,t_2]$. Thus $\rho$ intersects $L_2$ as well, and hence $\mathsf{D}_{i,j}^{(1)}$ occurs. By similar reasoning, we see that, when $\rho_{n;(u,t_1)}^{\rightarrow;(v,t_2)}$ intersects $L_2'$, the event $\mathsf{D}_{i,j}^{(2)}$ occurs. We have proved . For any compatible pair of points $(u,v)\in \mathrm{AdEndPair}_n(t)$, there exists a pair $(i,j)$ for which $u,v\in \mathcal{A}_{i,j}$; here we use $\psi t\leq t^{2/3}$. Hence, $$\begin{aligned} & & \left\{{ t^{-2/3}\big(\log t^{-1}\big)^{-1/3} \mathrm{MTF}_n(t)} > C\right\}\\ &\subseteq & \bigcup \left\{\mathsf{B}_{i,j}: i \in \llbracket 0,\lceil {t^{-1}}\rceil \rrbracket \, , \, j \in \left\llbracket-\left\lceil t^{-2/3} \big(\log t^{-1}\big)^{-1/3} \right\rceil, \left\lceil t^{-2/3} \big(\log t^{-1}\big)^{-1/3} \right\rceil\right\rrbracket\right\}\\ &\subseteq &\bigcup\left\{\mathsf{D}_{i,j}^{(1)}\cup \mathsf{D}_{i,j}^{(2)}:i \in \llbracket 0,\lceil {t^{-1}}\rceil \rrbracket \, , \, j \in \left\llbracket-\left\lceil t^{-2/3} \big(\log t^{-1}\big)^{-1/3} \right\rceil, \left\lceil t^{-2/3} \big(\log t^{-1}\big)^{-1/3} \right\rceil\right\rrbracket\right\} \, ,\end{aligned}$$ where (\[e:imp\]) was used in the latter inclusion. Thus, with $c,k_0,n_0$ as in the statement of Theorem  \[l:flucpoly\], for any fixed $t$ small enough that $\log t^{-1}\geq 2^2k_0^3$, and all $n\geq n_0(2t)^{-1}$, we have by a union bound and the translation invariance of the environment, $$\begin{aligned} && \P\left(\, t^{-2/3} \big(\log t^{-1}\big)^{-1/3} \mathrm{MTF}_n(t) > C \, \right)\\ &\leq & \big( 2 t^{-2/3}\big(\log t^{-1}\big)^{-1/3}+2 \big)(t^{-1}+2)\P\left(\mathrm{TF}_{n;(0,0)}^{(0,2t)}>(2^{-1}C-1)t^{2/3}\big(\log t^{-1}\big)^{1/3}\right)\\ &\leq & 2(t^{-2/3}+1)(t^{-1}+2)\exp\left\{-c(2^{-1}C-1)^3\log t^{-1}\right\}\leq 8\cdot t^{c(C/2-1)^3-5/3}.\end{aligned}$$ Here the second inequality follows from Theorem \[l:flucpoly\] with $\bm{t}=2t,\bm{k}=2^{-2/3}(2^{-1}C-1)\big(\log t^{-1}\big)^{1/3}$ and $\bm{n}=n$ being the parameter settings. The assumptions $\log t^{-1}\geq 2^2k_0^3$, and $n\geq n_0(2t)^{-1}$ ensure that $\bm{n}\geq n_0\bm{t}^{-1}$ and $\bm{k}\geq k_0$ for any $C\geq 2$. Finally, choosing $C$ large enough that $c\left(C/2-1\right)^3>5/3$, we learn that $$\P\left(\, t^{-2/3} \big(\log t^{-1}\big)^{-1/3} \mathrm{MTF}_n(t) > C \, \right)\rightarrow 0\quad \mbox{as } \, \, {t\searrow 0}\, ,$$ whenever $n = n(t)$ verifies $n \geq n_0(2t)^{-1}$. This completes the proof of Theorem \[t:flucshortpol\]. Exponent pair $(1/3,2/3)$ for polymer weight: Proof of Theorem \[t:holdweight\] {#s:1/3,2/3} =============================================================================== A lemma and two propositions will lead to the proof of Theorem \[t:holdweight\] on the [Hölder ]{}continuity of $[1,2]\mapsto\R:t\mapsto\mathsf{Wgt}_n(t)$, the polymer weight profile under vertical displacement. \[l:upwtineq\] There exist positive constants $n_0,r_0,s_0,c_0$ such that, for all $n \geq n_0$, $z\in [1,2]$, $t\in [r_0n^{-1},2-z]$ and $s\in [s_0,10(nt)^{2/3}]$, $$\P\left(|\mathsf{Wgt}_n(z+t)-\mathsf{Wgt}_n(z)|\geq st^{1/3}\right)\leq 5e^{-c_0s^{3/2}}\,.$$ We postpone the proof to Section \[ss:upbdwt\] and first see how the lemma implies the upper bound in Theorem \[t:holdweight\]. This bound follows from Lemma \[l:upwtineq\] similarly to how Theorem \[t:main1\] is derived from Proposition \[l:boundatz\]. \[t:upbdwt\] The sequence $\{\mathsf{Wgt}_n\}_{n\in \N}$ is tight in $(C[1,2],\|\cdot\|_\infty)$. Moreover, if $\mathsf{Wgt}_*$ is the weak limit of a weakly converging subsequence of $\{\mathsf{Wgt}_n\}_{n\in \N}$, then there exists a positive constant $C$ not depending on the particular weak limit $\mathsf{Wgt}_*$ such that, almost surely, $$\label{e:wtup} \limsup_{t\searrow 0}\sup_{1\leq z\leq 2-t}{\left|\mathsf{Wgt}_*(z+t)-\mathsf{Wgt}_*(z)\right|} \, {t^{-1/3}\big(\log t^{-1}\big)^{-2/3}}\leq C\,.$$ Lemma \[l:upwtineq\] holds only for $t\in [\max\{r_0n^{-1},10^{-3/2}s^{3/2}n^{-1}\},2-z]$ for some fixed constant $r_0>0$, and not for all $t\in [0,1-z]$, as was the case in Proposition \[l:boundatz\]. Hence, we directly show tightness in the following proof instead of applying Kolmogorov-Chentsov’s tightness criterion. [**Proof of Proposition \[t:upbdwt\].**]{} To show the first statement, concerning tightness, we follow the proof of the tightness criterion used to derive [@Bil68 Theorem $12.3$]. To this end, it is enough to show that, for given $\epsilon,\eta>0$, there exist $\delta \in [0,1]$, which we may harmlessly suppose to verify $\delta^{-1} \in \N$, and $N_0 \in \N$ such that, for all $n\geq N_0$, $$\label{e:toshowtight} \sum_{j<\delta^{-1}} \P\left(\sup_{j\delta\leq u\leq (j+1)\delta}|\mathsf{Wgt}_n(1+u)-\mathsf{Wgt}_n(1+j\delta)|\geq \epsilon\right)<\eta\,.$$ Assume then that $\epsilon,\eta>0$ are given small constants. For the time being, fix some $\delta>0$ small to be chosen later (depending on $\epsilon$ and $\eta$). Now fix any $M>1$. For any $z_1,z_2\in [1,2]$ such that $|z_1-z_2|=10^{-1}\epsilon n^{-2/3}$, set $t = \vert z_1 - z_2 \vert$. For all $\lambda\in [0,\epsilon]$, clearly $\lambda t^{-1/3}\leq 10(nt)^{2/3}$. Hence, choosing $s=\lambda t^{-1/3}$ in Lemma \[l:upwtineq\], one gets, for all $n$ large enough, $$\label{e:onepointwt} \P\Big(|\mathsf{Wgt}_n(z_1)-\mathsf{Wgt}_n(z_2)|\geq \lambda\Big)\leq K_M\lambda^{-3M}|z_1-z_2|^M,$$ for some constant $K_M$ depending only on $M$. To establish tightness, the general strategy is to bound the distribution of the maximum of certain fluctuations. To achieve this, we crucially use the bound in together with the inequality in [@Bil68 Theorem $12.2$] that bounds the maximum of partial sums. To this end, fix $j<\delta^{-1}$, and break the interval $[j\delta,(j+1)\delta]$ into $\lceil \delta\beta^{-1}\rceil$-many subintervals of length $\beta:=10^{-1}\epsilon n^{-2/3}$ each, and follow the proof of the inequality in [@Bil68 Theorem $12.2$] to obtain $$\label{e:wgt1} \P\left(\max_{0\leq i\leq \lceil \delta \beta^{-1}\rceil}|\mathsf{Wgt}_n(1+j\delta+i\beta)-\mathsf{Wgt}_n(1+j\delta)|\geq \frac{\epsilon}{2}\right)\leq K'_M\epsilon^{-3M}\delta^M,$$ for some appropriate constant $K'_M$ depending only on $M$. Note that by [@Bil68 Theorem $12.2$] it directly follows that if holds for all $\lambda>0$, then holds for all $\epsilon>0$. However, in our case holds for all $\lambda\in [0,\epsilon]$, instead of all $\lambda>0$. Hence, we resort to the proof of [@Bil68 Theorem $12.2$] which shows that if for some fixed $\epsilon>0$, holds for all $\lambda\in [0,\epsilon]$, then holds for that particular $\epsilon$. Now, fix any $i\in \llbracket 0,\lceil \delta \beta^{-1}\rceil-1\rrbracket$. For any $u\in [j\delta+i\beta,j\delta+(i+1)\beta]$, it clearly follows from the definition , $${\mathsf{W}}_{n;(0,1+u)}^{(0,1+j\delta+(i+1)\beta)}\geq -2n^{2/3}(1+j\delta+(i+1)\beta-(1+u))\geq -2n^{2/3}\beta\,,\,\,\mbox{and}$$ $${\mathsf{W}}_{n;(0,1+j\delta+i\beta)}^{(0,1+u)}\geq -2n^{2/3}(1+u-(1+j\delta+i\beta))\geq -2n^{2/3}\beta\,.$$ Thus, for any $u\in [j\delta+i\beta,j\delta+(i+1)\beta]$, by superaddivity of polymer weights described in , $${\mathsf{W}}_{n;(0,0)}^{(0,1+j\delta+i\beta)}-2n^{2/3}\beta\leq {\mathsf{W}}_{n;(0,0)}^{(0,1+j\delta+i\beta)}+{\mathsf{W}}_{n;(0,1+j\delta+i\beta)}^{(0,1+u)}\leq {\mathsf{W}}_{n;(0,0)}^{(0,1+u)}\,\,\, \mbox{and}$$ $${\mathsf{W}}_{n;(0,0)}^{(0,1+u)}\leq {\mathsf{W}}_{n;(0,0)}^{(0,1+j\delta+(i+1)\beta)}-{\mathsf{W}}_{n;(0,1+u)}^{(0,1+j\delta+(i+1)\beta)}\leq {\mathsf{W}}_{n;(0,0)}^{(0,1+j\delta+(i+1)\beta)}+2n^{2/3}\beta\,.$$ This, together with , imply that for any $i\in \llbracket 0,\lceil \delta \beta^{-1}\rceil-1\rrbracket$ and $u\in [j\delta+i\beta,j\delta+(i+1)\beta]$, $$\begin{aligned} \label{e:wgt2} n^{1/3}\left|\mathsf{Wgt}_n(1+u)-\mathsf{Wgt}_n(1+j\delta)\right|\leq 2n\beta+2+ n^{1/3}\max\Big\{\left|\mathsf{Wgt}_n(1+j\delta+i\beta)-\mathsf{Wgt}_n(1+j\delta)\right|,\nonumber\\ \left|\mathsf{Wgt}_n(1+j\delta+(i+1)\beta)-\mathsf{Wgt}_n(1+j\delta)\right|\Big\}.\end{aligned}$$ Since $2n\beta= 5^{-1}\epsilon n^{1/3}$, for all $n$ large enough that $2n^{-1/3}\leq \epsilon/5$, and imply $$\begin{aligned} &&\P\left(\sup_{j\delta\leq u\leq (j+1)\delta}\left|\mathsf{Wgt}_n(1+u)-\mathsf{Wgt}_n(1+j\delta)\right|\geq \epsilon\right)\\ &\leq& \P\left(\max_{0\leq i\leq \lceil \delta \beta^{-1}\rceil}|\mathsf{Wgt}_n(1+j\delta+i\beta)-\mathsf{Wgt}_n(1+j\delta)|\geq \frac{\epsilon}{2}\right)\leq K'_M\epsilon^{-3M}\delta^M.\end{aligned}$$ Thus, by choosing $\delta$ small enough that $K'_M\epsilon^{-3M}\delta^{M-1}<\eta$, we obtain , and hence tightness. To show , we follow the proof of Theorem \[t:main1\](b). Let $n_0,r_0,s_0$ and $c_0$ be as in Lemma \[l:upwtineq\]. For any fixed $m\in \N$ such that $c_1\left(\log 2^m\right)^{2/3}\geq s_0$, and any $j\in \{0,1,2,\cdots,2^m-1\}$, and all $n\geq \max\{r_02^m,10^{-3/2}c_1^{3/2}2^m\log 2^m\}$, by applying Lemma \[l:upwtineq\] with the parameters $\bm{n}=n,\bm{t}=2^{-m}$ and $\bm{s}=c_1\left(\log 2^m\right)^{2/3}$, it follows that $$\P\Bigg( \big\vert \mathsf{Wgt}_n(1+(j+1)2^{-m})-\mathsf{Wgt}_n(1+j2^{-m}) \big\vert > c_12^{-\frac{m}{3}}\left(\log 2^m\right)^{2/3}\Bigg) \, \leq \, 5\cdot 2^{-m(c_0c_1^{3/2})}\,.$$ Now, observe that in the proof of Lemma \[l:dyad\] carries over verbatim to the present case. By choosing $c_1$ high enough that $c_0c_1^{3/2}>1$, and exactly imitating the rest of the proof of Lemma \[l:dyad\] followed by the proof of Theorem \[t:main1\](b), we complete the proof of Proposition \[t:upbdwt\]. Turning to prove the lower bound in , we restate it now. \[p:wgtlwb\] There exists a constant $c>0$ such that, almost surely, $$\liminf_{t\searrow 0}\sup_{1\leq z\leq 2-t} \, t^{-1/3}\big(\log t^{-1}\big)^{-2/3} \big\vert \mathsf{Wgt}_*(z+t)-\mathsf{Wgt}_*(z) \big\vert \, \geq \, c\,.$$ This result will follow directly from weight superadditivity, i.e. ${\mathsf{W}}_{n;(0,0)}^{(0,1+z+t)}-{\mathsf{W}}_{n;(0,0)}^{(0,1+z)}\geq {\mathsf{W}}_{n;(0,1+z)}^{(0,1+z+t)}$ for $z,t>0$, control on weight with given endpoints via Theorem \[t:moddevlow\], independence in disjoint strips, and the weight ${\mathsf{W}}_{n;(0,1+z)}^{(0,1+z+t)}$ depending on the configuration in the strip delimited by the lines $y=1+z$ and $y=1+z+t$. The proof is reminiscent of an argument for a similar statement made for Brownian motion: see the proof on page $362$ of Exercise $1.7$ in the book [@MP]. [**Proof of Proposition \[p:wgtlwb\].**]{} We need to show that, for some constant $c>0$, almost surely, there exists $\epsilon>0$ such that, for all $0<t<\epsilon$ and some $z\in[1,2-t]$, $$|\mathsf{Wgt}_*(z+t)-\mathsf{Wgt}_*(z)|\geq ct^{1/3}\big(\log t^{-1}\big)^{2/3}.$$ Let $c>0$ satisfy $2^{3/2}c_2c^{3/2}<1$, where $c_2$ arises from Theorem \[t:moddevlow\]. For integers $n,m\geq 1$ and $k\in \{0,1,2,\cdots,m-1\}$, we define the events $$\mathsf{A}_{k,m,n}=\left\{\mathsf{Wgt}_n\left(1+{(k+1)}{m}^{-1}\right)-\mathsf{Wgt}_n\left(1+{k}{m}^{-1}\right)\geq c{m^{-1/3}}\left(\log m\right)^{2/3}\right\}$$ and $$\mathsf{A}_{k,m}=\left\{\mathsf{Wgt}_*\left(1+{(k+1)}{m}^{-1}\right)-\mathsf{Wgt}_*\left(1+{k}{m}^{-1}\right)\geq c{m^{-1/3}}\left(\log m\right)^{2/3}\right\}\,.$$ Also let $$\mathsf{B}_{k,m,n}=\left\{{\mathsf{W}}_{n;(0,1+km^{-1})}^{(0,1+(k+1)m^{-1})}\geq c{m^{-1/3}}\left(\log m\right)^{2/3}+2n^{-1/3}\right\}\,.$$ Let $n_0,s_0$ and $c_2$ be as in Theorem \[t:moddevlow\], and let $m_0$ be large enough that $2c(\log m_0)^{2/3}\geq \max\{s_0,4n_0^{-1/3}\}$. Then from Theorem \[t:moddevlow\] with parameter settings $\bm{t_1}= 1+km^{-1},\bm{t_2}=1+(k+1)m^{-1},\bm{{t_{1,2}}}=m^{-1},\bm{n}=n$ and $\bm{s}=2c(\log m)^{2/3}$, for all $m\geq m_0$ and $n\geq n_0m$, $$\label{e:Bomn} \P(\mathsf{B}_{0,m,n})\geq \P\left({\mathsf{W}}_{n;(0,1+km^{-1})}^{(0,1+(k+1)m^{-1})}\geq 2c{m^{-1/3}}\left(\log m\right)^{2/3}\right)\geq e^{-2^{3/2}c_2c^{3/2}\log m}=m^{-2^{3/2}c_2c^{3/2}}\,.$$ Here the first inequality follows because $$cm^{-1/3}\left(\log m\right)^{2/3}\geq cm^{-1/3}\left(\log m_0\right)^{2/3}\geq 2n_0^{-1/3}m^{-1/3}\geq 2n^{-1/3}\,$$ for $m\geq m_0$, $2c(\log m_0)^{2/3}\geq 4n_0^{-1/3}$ and $n\geq n_0m$. Now $\mathsf{B}_{k,m,n}$ are i.i.d. random variables for $k\in\{0,1,2,\cdots,m-1\}$ as the weights of polymers over disjoint regions are independent. Also using ${\mathsf{W}}_{n;(0,0)}^{(0,1+(k+1)m^{-1})}-{\mathsf{W}}_{n;(0,0)}^{(0,1+km^{-1})}\geq {\mathsf{W}}_{n;(0,1+km^{-1})}^{(0,1+(k+1)m^{-1})}$ by superadditivity of polymer weights, together with , we get that $\mathsf{B}_{k,m,m}\subseteq \mathsf{A}_{k,m,n}$. Thus, using , for all $m\geq m_0$ and $n\geq n_0m$, $$\begin{aligned} \label{e:Acom} \P\left(\bigcap_{k=0}^{m-1} \mathsf{A}^c_{k,m,n}\right)&\leq& \P\left(\bigcap_{k=0}^{m-1} \mathsf{B}^c_{k,m,n}\right) =(1-\P(\mathsf{B}_{0,m,n}))^m\nonumber\\ &\leq & \exp\left\{-m\P(\mathsf{B}_{0,m,n})\right\}\leq \exp\Big\{-m^{1-2^{3/2}c_2c^{3/2}}\Big\}\,,\end{aligned}$$ where we use that $1-x\leq e^{-x}$ for all $x\geq 0$. Next, similarly to the first part of the proof of Lemma \[l:dyad\], let $\{\mathsf{Wgt}_{n_r}\}_r$ be a subsequence of $\{\mathsf{Wgt}_{n}\}_n$ such that $\mathsf{Wgt}_{n_r}\Rightarrow \mathsf{Wgt}_*$ as random variables in $(C[1,2],\|\cdot\|_\infty)$ (where $\Rightarrow$ denotes convergence in distribution). Since for $a,b\in [1,2]$, the map $T_{a,b}$ defined by $(C[1,2],\|\cdot\|_\infty)\mapsto (\R,|\cdot|):f\mapsto f(a)-f(b)$ is continuous, the set $$\mathsf{U}:=\bigcap\left\{T^{-1}_{1+(k+1)m^{-1},1+km^{-1}}\left(-\infty,cm^{-1/3}\big(\log m\big)^{2/3}\right):k=0,1,\cdots,m-1\right\}$$ is open. Thus, by the Portmanteau theorem, $$\P\left(\bigcap_{k=0}^{m-1} \mathsf{A}^c_{k,m}\right)\leq \liminf_r\P\left(\bigcap_{k=0}^{m-1} \mathsf{A}^c_{k,m,n_r}\right)\leq \limsup_n\P\left(\bigcap_{k=0}^{m-1} \mathsf{A}^c_{k,m,n}\right)\,.$$ From here, using and that our given choice of the constant $c$ ensures $2^{3/2}c_2c^{3/2}<1$, we get $$\begin{aligned} \sum_{m=m_0}^\infty \P\left(\bigcap_{k=0}^{m-1} \mathsf{A}^c_{k,m}\right)\leq \sum_{m=m_0}^\infty \limsup_n\P\left(\bigcap_{k=0}^{m-1} \mathsf{A}^c_{k,m,n}\right)\leq \sum_{m=m_0}^\infty \exp\Big\{-m^{1-2^{3/2}c_2c^{3/2}}\Big\} <\infty\,.\end{aligned}$$ Hence, using the Borel-Cantelli lemma, almost surely there exists $M_0\in \N$ such that for all $m\geq M_0$, one has some $k_m\leq m-1$ with $z=1+{k_m}{m}^{-1}$ satisfying $$\left|\mathsf{Wgt}_*(z+{m}^{-1})-\mathsf{Wgt}_*(z)\right|\geq c{m^{-1/3}}\left(\log m\right)^{2/3}.$$ Let $\epsilon= M_0^{-1}$. Also let $M_0^{-1}$ be small enough in the sense of Proposition \[t:upbdwt\]: namely, almost surely for all $t\in [0,M_0^{-1}]$, $\sup_{1\leq z\leq 2-t}{\left|\mathsf{Wgt}_*(z+t)-\mathsf{Wgt}_*(z)\right|}{t^{-1/3}\big(\log t^{-1}\big)^{-2/3}}\leq 2C$. Then, for any given $t\in [0,\epsilon]$, let $m$ be such that ${(m+1)^{-1}}<t\leq {m}^{-1}$. Then for $z=1+{k_m}{m}^{-1}$, $$\begin{aligned} &&|\mathsf{Wgt}_*(z+t)-\mathsf{Wgt}_*(z)|\\ &\geq &\left|\mathsf{Wgt}_*\left(z+{m}^{-1}\right)-\mathsf{Wgt}_*(z)\right|-\left|\mathsf{Wgt}_*(z+t)-\mathsf{Wgt}_*\left(z+{m}^{-1}\right)\right|\\ &\geq&c{m^{-1/3}}\left(\log m\right)^{2/3}-2C\left({m}^{-1}-{(m+1)}^{-1}\right)^{1/3}\Big(\log\left({m}^{-1}-{(m+1)}^{-1}\right)^{-1}\Big)^{2/3}.\end{aligned}$$ As the second term decays much faster than the first, choosing $M_0$ large enough so that the second term is smaller that $2^{-1} c {m^{-1/3}}(\log m)^{2/3}$ gives the result. [**Proof of Theorem \[t:holdweight\].**]{} This result follows from Proposition \[t:upbdwt\] and Proposition \[p:wgtlwb\]. Upper bound on polymer weight fluctuation: Proof of Lemma \[l:upwtineq\] {#ss:upbdwt} ------------------------------------------------------------------------ In this subsection, we complete the proof of Theorem \[t:holdweight\]. The remaining element, Lemma \[l:upwtineq\], will be derived from Lemmas \[l:upineq1\] and \[l:upwtineq2\]. \[l:upineq1\] There exist positive constants $s_0,r_0$ and $c_0$ such that for $s\geq s_0$, $z\in [1,2]$ and $t\in [r_0n^{-1},2-z]$, $$\P\left({\mathsf{W}}_{n;(0,0)}^{(0,z)}\geq {\mathsf{W}}_{n;(0,0)}^{(0,z+t)}+st^{1/3}\right)\leq e^{-c_0s^{3/2}}.$$ [**Proof.**]{} Using ${\mathsf{W}}_{n;(0,0)}^{(0,z+t)}\geq {\mathsf{W}}_{n;(0,0)}^{(0,z)}+{\mathsf{W}}_{n;(0,z)}^{(0,z+t)}$, we see that, for $nt\geq r_0$ and $s\geq s_0$, $$\P\left({\mathsf{W}}_{n;(0,0)}^{(0,z)}\geq {\mathsf{W}}_{n;(0,0)}^{(0,z+t)}+st^{1/3}\right)\leq \P\big({\mathsf{W}}_{n;(0,z)}^{(0,z+t)}\leq -st^{1/3}\big)\leq e^{-cs^{3/2}},$$ where the latter inequality follows from the moderate deviation estimate Theorem \[t:moddev\], with $\bm{t_1}=z,\bm{t_2}=z+t,\bm{n}=n$ and $\bm{s}=s$, and setting $r_0$ and $s_0$ to equal $n_0$ and $s_0$ respectively from the statement of Theorem \[t:moddev\]. Next is the more subtle of the two constituents of Lemma \[l:upwtineq\]. \[l:upwtineq2\]There exist positive constants $n_0,s_2,r_1$ and $c_0$ such that, for $n\geq n_0$, $t\in [r_1n^{-1},2-z]$, $s\in [s_2,10(nt)^{2/3}]$ and $z\in [1,2]$, $$\label{e:wgtuplem} \P\left({\mathsf{W}}_{n;(0,0)}^{(0,z+t)}\geq {\mathsf{W}}_{n;(0,0)}^{(0,z)}+st^{1/3}\right)\leq 4 \, e^{-c_0s^{3/2}}.$$ This proof is reminiscent of arguments used in [@BSS14] and [@BSS17++]. We first explain the basic idea, which is illustrated in Figure \[f:fig2weight\]. A path may be formed from $(0,0)$ to $(0,z)$ by following the route of a polymer from $(0,0)$ to $(0,z+t)$ until its location, $(U,z-t)$ say, at height $z-t$; and then following a polymer from $(U,z-t)$ to $(0,z)$. The discrepancy in weight between the original polymer, from $(0,0)$ to $(0,z+t)$, and the newly formed path, from $(0,0)$ to $(0,z)$, is equal to the difference in weights between the polymer from $(U,z-t)$ to $(0,z+t)$ and that from $(U,z-t)$ to $(0,z)$. The latter two polymers have duration of order $t$; Theorem \[t:unifmoddev\] may then show that their weights have order $t^{1/3}$. Thus, the weight difference ${\mathsf{W}}_{n;(0,0)}^{(0,z+t)} - {\mathsf{W}}_{n;(0,0)}^{(0,z)}$, which is at most the discrepancy we are considering, is seen to be unlikely to exceed order $t^{1/3}$. [**Proof of Lemma \[l:upwtineq2\].**]{} To implement this idea, we will consider, for definiteness, the leftmost polymer from $(0,0)$ to $(z+t,0)$, namely $\rho_{n;(0,0)}^{\leftarrow;(z+t,0)}$. In accordance with the notation in the plan, we will set $U = \rho_{n;(0,0)}^{\leftarrow;(z+t,0)}(z-t)$. The height-$(z-t)$ polymer location $U$ typically has order $t^{2/3}$. The plan will run into trouble if $U$ is atypically high, because then the two short polymers running to $(0,z+t)$ and $(0,z)$ from $(U,z-t)$ will have large negative weights dictated by parabolic curvature. To cope with this difficulty, we introduce a [*good*]{} event $\mathsf{G}$, $$\mathsf{G}=\left\{\left|U\right|\leq \phi\right\}\,,$$ specified in terms of a parameter $\phi$ that is set equal to $D^{-1}s^{1/2}(2t)^{2/3}$. Here, the constant $D$ is chosen to be $2^{2/3}10^{1/2}{C_0}$, with $C_0$ given by Theorem \[t:unifmoddev\]. In view of Theorem \[t:carsestimate2\], this choice of $\phi$ ensures that the event $\mathsf{G}$ fails to occur with probability of order $\exp \big\{ - \Theta(1) s^{3/2} \big\}$. (The appearance of the factor of $D^{-1}$ in $\phi$ is a detail concerning values of $s$ in Lemma \[l:upwtineq2\] close to the maximum value $10 (nt)^{2/3}$. ) Indeed, applying Theorem \[t:carsestimate2\] with $\bm{n}=n,\bm{t_1}=0,\bm{t_2}=z+t,\bm{t}=z-t,\bm{x}=0,\bm{y}=0$ and $\bm{s}=D^{-1}s^{1/2}$, we find that, when $n\geq n_0$ (a bound which ensures that the hypothesis that $\bm{n{t_{1,2}}} \geq n_0$ is met) and $s \geq s_1$, $$\label{e:boundGcom} \P(\mathsf{G}^c) \, \leq \, 2 \, \exp \big\{ -c D^{-3} s^{3/2} \big\} \, ,$$ where the positive constants $c$ and $s_1$ are provided by the theorem being applied. ![When the thick blue polymer $\rho_{n;(0,0)}^{\leftarrow;(z+t,0)}$ crosses height $z-t$ without immoderately high fluctuation, it may be diverted via the red polymer to form a path of comparable weight from $(0,0)$ to $(z,0)$.[]{data-label="f:fig2weight"}](Fig2weight.pdf){width="90.00000%"} When $\mathsf{G}$ occurs, $$|U|\leq D^{-1}s^{1/2}(2t)^{2/3}\leq D^{-1}2^{2/3}10^{1/2}tn^{1/3}<tn^{1/3}\,,$$ because $s\leq 10(nt)^{2/3}$, $D = 2^{2/3} 10^{1/2} C_0$ and $C_0>1$. As we saw in Subsection \[sss:comp\], it is this bound on $\vert U \vert$ that ensures the existence of polymers between $(U,z-t)$ and $(0,z)$. By superadditivity of polymer weights, we thus have $${\mathsf{W}}_{n;(0,0)}^{(0,z)}\geq {\mathsf{W}}_{n;(0,0)}^{(U,z-t)}+{\mathsf{W}}_{n;(U,z-t)}^{(0,z)}\,.$$ Thus, when $\mathsf{G}$ occurs, $$\begin{aligned} {\mathsf{W}}_{n;(0,0)}^{(0,z+t)}- {\mathsf{W}}_{n;(0,0)}^{(0,z)}&\leq& {\mathsf{W}}_{n;(0,0)}^{(0,z+t)}-{\mathsf{W}}_{n;(0,0)}^{(U,z-t)}-{\mathsf{W}}_{n;(U,z-t)}^{(0,z)}\\ &=&{\mathsf{W}}_{n;(U,z-t)}^{(0,z+t)}-{\mathsf{W}}_{n;(U,z-t)}^{(0,z)} \, \leq \, \sup_{x\in [-\phi,\phi]}\left({\mathsf{W}}_{n;(x,z-t)}^{(0,z+t)}-{\mathsf{W}}_{n;(x,z-t)}^{(0,z)}\right)\,, \end{aligned}$$ where the equality is dependent on the definition of $U$ and the final inequality on the occurrence of $\mathsf{G}$. We see then that $$\begin{aligned} &&\P\left(\mathsf{G}\cap \left\{{\mathsf{W}}_{n;(0,0)}^{(0,z+t)}\geq {\mathsf{W}}_{n;(0,0)}^{(0,z)}+st^{1/3}\right\}\right) \nonumber \\ &\leq& \P\left(\sup_{x\in [-\phi,\phi]}\left({\mathsf{W}}_{n;(x,z-t)}^{(0,z+t)}-{\mathsf{W}}_{n;(x,z-t)}^{(0,z)}\right)\geq st^{1/3}\right)\nonumber\\ & \leq & \P\left(\sup_{x\in [-\phi,\phi]}\left|{\mathsf{W}}_{n;(x,z-t)}^{(0,z+t)}\right|>2^{-1}st^{1/3}\right)+ \P\left(\sup_{x\in [-\phi,\phi]}\left|{\mathsf{W}}_{n;(x,z-t)}^{(0,z)}\right|>2^{-1}st^{1/3}\right) \label{e:ineqgcap} \,.\end{aligned}$$ The latter two probabilities will each be bounded above by a union bound over several applications of Theorem \[t:unifmoddev\]. Addressing the first of these probabilities to begin with, we set parameters for a given application of the theorem, taking $\bm{I}$ to be a given interval of length at most $t^{2/3}$ contained in $[-\phi,\phi]$ and $\bm{J} = \{ 0 \}$, and also setting $\bm{n}=n, \bm{t_1}=z-t,\bm{t_2}=z$ and $\bm{s}=4^{-1}s$. The theorem’s hypothesis concerning inclusion for the interval ${\bf I}$ (and ${\bf J}$) is ensured because $$|x|\leq D^{-1}s^{1/2}(2t)^{2/3}\leq 2^{2/3}10^{1/4}D^{-1}s^{1/4}n^{1/6}t^{5/6}< C_0^{-1}s^{1/4}n^{1/6}t^{5/6}\,,$$ for $x\in [-\phi,\phi]$, where here we use $s\leq 10(nt)^{2/3}$ and $D= 2^{2/3}10^{1/2}C_0>2^{2/3}10^{1/4}C_0$. In these applications of Theorem \[t:unifmoddev\], the parabolic curvature term inside the supremum, $t^{-4/3} x^2$, is at most $t^{-4/3} \phi^2$. It is thus also at most $s/4$, because $\phi = D^{-1} s^{1/2}(2t)^{2/3}$ and $D \geq 2^{5/3}$. Thus, dividing $[-\phi,\phi]$ into $\lceil 2^{5/3}D^{-1}s^{1/2}\rceil$-many consecutive intervals of length at most $t^{2/3}$, we are indeed able to apply Theorem \[t:unifmoddev\] and a union bound, finding that, for $n_0 \in \N$ and $C,c > 0$ the constants furnished by the theorem, and for $nt \geq n_0$, $$\begin{aligned} &&\P\left(\sup_{x\in [-\phi,\phi]}\left|{\mathsf{W}}_{n;(x,z-t)}^{(0,z)}\right|>2^{-1}st^{1/3}\right)\\ &\leq &\P\left(\sup_{x\in [-\phi,\phi]}\left|t^{-1/3}{\mathsf{W}}_{n;(x,z-t)}^{(0,z)}+t^{-4/3}x^2\right|>4^{-1}s\right)\leq \lceil 2^{5/3}D^{-1}s^{1/2}\rceil C e^{-cs^{3/2}}\leq e^{-c's^{3/2}}\,,\end{aligned}$$ for $c'=2^{-1}c$ and $s\geq s_0$ where $s_0$ is chosen in such a way that $e^{2^{-1}cs_0^{3/2}}\geq C \lceil 2^{5/3}D^{-1}s_0^{1/2}\rceil$. The second probability in is bounded above by similar means. Several applications of Theorem \[t:unifmoddev\] will be made. In a given application, the parameters $\bm{I}, \bm{J}, \bm{n}$ and $\bm{s}$ are chosen as before, but we now set $\bm{t_1}=z-t$ and $\bm{t_2}=z+t$, so that $\bm{{t_{1,2}}}$ equals $2t$, rather than $t$. The curvature term $(2t)^{-4/3} x^2$ is bounded above by $(2t)^{-4/3} \phi^2$, a smaller bound than before, so that the preceding bound of $s/4$ remains valid. The condition for inclusion for the intervals $\bm{I}$ (and $\bm{J}$), namely $\phi \leq C_0^{-1}s^{1/4}n^{1/6}(2t)^{5/6}$, is weaker than it was previously and is thus satisfied. Hence, using Theorem \[t:unifmoddev\] and a union bound, we find that, for all $n\geq 2^{-1}n_0 t^{-1}$, $$\P\left(\sup_{x\in [-\phi,\phi]}\left|{\mathsf{W}}_{n;(x,z-t)}^{(0,z+t)}\right|>2^{-1}st^{1/3}\right)\leq e^{-c's^{3/2}}\,,$$ for $s\geq s_0$. Combining and with the two bounds just derived, we obtain Lemma \[l:upwtineq2\] by taking $c_0 > 0$ to be less than $\min\{c D^{-3},c'\}$, $s_2$ to be suitably greater than $\max\{s_0, s_1\}$, and $r_1=2^{-1}n_0$. [**Proof of Lemma \[l:upwtineq\].**]{} This follows immediately using and from Lemmas \[l:upineq1\] and \[l:upwtineq2\] and a union bound. Lower bound on transversal fluctuation: Proof of Proposition \[p:lowbndTF\] {#s:last} =========================================================================== In this last section we shall prove the lower bound on the transversal fluctuation of the polymer, the corresponding upper bound of which was proved in [@BSS14 Theorem $11.1$] (and is stated here, with the optimal exponent in the bound, as Theorem \[t:transversal\]). In fact, Proposition \[p:lowbndTF\] does slightly more than just providing a corresponding lower bound on the quantity whose upper bound is proved in Theorem \[t:transversal\]. Indeed, in Proposition \[p:lowbndTF\], one takes the minimum over the transversal fluctuations of all the polymers between two fixed points, and not just the transversal fluctuation of the leftmost one. The proof of Proposition \[p:lowbndTF\] crucially uses the polymer weight lower tail Theorem \[t:moddevlow\]. We also fix the constant $\alpha_0$ in this Proposition \[p:lowbndTF\] as $\alpha_0=C_0^{-2}3^{-5/3}10^{-1/2}$, where $C_0$ is as in Theorem \[t:unifmoddev\]. This choice of $\alpha_0$ ensures that the condition in the hypothesis of Theorem \[t:unifmoddev\] is met whenever it is applied. [**Proof of Proposition \[p:lowbndTF\].**]{} We prove the proposition for $t_1=0$ and $t_2=1$. The case for general $t_1<t_2$ follows readily using the scaling principle . A box is a subset of $\R^2$ of the form $[a,b] \times [r_1,r_2]$, where $a \leq b$ and $r_1 \leq r_2$. Any box has a lower and an upper side, namely $[a,b] \times \{ r_1 \}$ and $[a,b] \times \{ r_2 \}$. The key box for the proof is ${{\rm Strip}}$, now specified to be $[-s,s] \times [0,1]$. Proposition \[p:lowbndTF\] is, after all, a lower bound on the probability that there exists a polymer between $(0,0)$ and $(0,1)$ that escapes ${{\rm Strip}}$. We divide ${{\rm Strip}}$ into three further boxes, writing ${{\rm Mid}}$ for the box $[-s,s] \times [1/3,2/3]$, and ${{\rm South}}$ and ${{\rm North}}$ for the boxes obtained from ${{\rm Mid}}$ by vertical translations of $-1/3$ and $1/3$. We further set ${{\rm West}}$ to be the box obtained from ${{\rm Mid}}$ by a horizontal translation of $-2s$. See Figure \[f.casehigh\]. ![In Case High, the high weight path $\rho$ is extended to form a path from $(0,0)$ to $(0,1)$ whose weight exceeds that of any path between these points that remains in ${{\rm Strip}}= {{\rm North}}\cup {{\rm Mid}}\cup {{\rm South}}$.[]{data-label="f.casehigh"}](casehigh.pdf) Recall that, when $(x,t_1)$ and $(y,t_2)$ verify $n^{1/3} {t_{1,2}}\geq |y - x|$, we denote the polymer weight with this pair of endpoints by ${\mathsf{W}}_{n;(x,t_1)}^{(y,t_2)}$. We now use a set theoretic notational convention to refer in similar terms to the set of weights of polymers between two collections of endpoint locations. Indeed, let $I$ and $J$ be compact real intervals. We will write $${\mathsf{W}}_{n;(I,t_1)}^{(J,t_2)} = \Big\{ {\mathsf{W}}_{n;(x,t_1)}^{(y,t_2)} : x \in I, y \in J \Big\} \, ;$$ we will ensure that whenever this notation is used, $(x,t_1) \overset{n}{\preceq} (y,t_2)$ for all $x \in I$ and $y \in J$ in the sense of Subsection \[sss:comp\]. When an interval is a singleton, $I = \{ x \}$ say, we write $(x,t_1)$ instead of $(\{x\},t_1)$ when using this notation. To any box $B$ and $s \in \R$, we define the event ${{\rm High}}(B,s)$ that the weight of some path that is contained in $B$ with starting point in the lower side of $B$ and ending point in the upper side of $B$ is at least $s$. Our approach to proving Proposition \[p:lowbndTF\] gives a central role to the event ${{\rm High}}({{\rm Mid}},300s^2)$. It may be expected that the order of probability of this event is $\exp \big\{ - \Theta(1) s^3 \big\}$, but we do not attempt to prove this. Rather, we analyse two cases, called [*High*]{} and [*Low*]{}, according to the value of the event’s probability. We will quantify the notion of high or low probability for ${{\rm High}}({{\rm Mid}},300s^2)$ in terms of the decay rate for a very high weight polymer between $(0,0)$ and $(0,1)$. Indeed, noting from Theorem \[t:moddevlow\] that there exists $C > 0$ such that, for $s\geq s_0$, $$\label{e.wthou} \P \Big( {\mathsf{W}}_{n;(0,0)}^{(0,1)} \geq 1000s^2 \Big) \geq \exp \big\{ - C s^3 \big\} \, ,$$ we declare that Case High occurs if $$\P \big( {{\rm High}}({{\rm Mid}},300s^2) \big) \geq \exp \big\{ -2C s^3 \big\} \, ;$$ Case Low occurs when Case High does not. In order to analyse Case High, we introduce a [*favourable*]{} event $\mathsf{F}$. The event is specified as the intersection of the following events: - $\mathsf{G}_1=\left\{\inf{\mathsf{W}}_{n;(0,0)}^{([-3s,-s],1/3)} \geq - 50s^2\right\}$; - $\mathsf{G}_2=\left\{\inf{\mathsf{W}}_{n;([-3s,-s],2/3)}^{(0,1)} \geq - 50s^2\right\}$; - $\mathsf{G}_3=\left\{\sup {\mathsf{W}}_{n;(0,0)}^{([-s,s],1/3)} \leq 50s^2\right\}$; - $\mathsf{G}_4=\left\{\sup {\mathsf{W}}_{n;([-s,s],2/3)}^{(0,1)} \leq 50s^2\right\}$; - and $\mathsf{G}_5$ is the event that ${{\rm High}}({{\rm Mid}},50s^2)$ does not occur. Thus, the occurrence of $\mathsf{F}$ forces the absence of any high weight path inside ${{\rm Mid}}$ that crosses this box from its lower to its upper side, while also ensuring that any polymer connecting $(0,0)$ (or $(0,1)$) to the lower (or upper) sides of ${{\rm Mid}}$ and ${{\rm West}}$ is not of very low weight. We claim that $\mathsf{F}$ is a high probability event, proving this by applying Theorem \[t:unifmoddev\]. Indeed, for the events $\mathsf{G}_1$ and $\mathsf{G}_3$ entailed by $\mathsf{F}$, we make several applications of Theorem  \[t:unifmoddev\]. For a given application, we consider the parameter settings $\bm{n}=n, \bm{t_1}=0,\bm{t_2}=1/3, \bm{s}=10s^2, \bm{I}=\{0\}$ and $$\bm{J}= \big[-3s+(i-1)3^{-2/3},\max\{-3s+i3^{-2/3},s\} \big]$$ for some $i\in \{1,2,\cdots,\lceil4\cdot 3^{2/3}s\rceil\}$. The condition on inclusion for the intervals $\bm{I}$ and $\bm{J}$ is satisfied since for $y\in [-3s,s]$, $$|y|\leq 3s\leq \bm{s}^{1/2}\leq 10^{1/4}\alpha_0^{1/2}\bm{n}^{1/6}\bm{s}^{1/4}\leq 3^{5/6}10^{1/4}\alpha_0^{1/2}\bm{n}^{1/6}\bm{s}^{1/4}\bm{{t_{1,2}}}^{5/6}=C_0^{-1}\bm{n}^{1/6}\bm{s}^{1/4}\bm{{t_{1,2}}}^{5/6}\,,$$ where we use that $s\leq \alpha_0n^{1/3}$ and our given choice of $\alpha_0$ has been made so that $\alpha_0=C_0^{-2}3^{-5/3}10^{-1/2}$. Also the parabolic curvature inside the supremum is $$\sup_{y\in [-3s,s]}3^{4/3}y^2\leq 3^{4/3}\cdot 3^2s^2<40s^2\,.$$ Thus, dividing $[-3s,s]$ into $\lceil4\cdot 3^{2/3}s\rceil$-many intervals of length at most $3^{-2/3}$ and using Theorem \[t:unifmoddev\] and a union bound, it follows that, for $s$ large enough and $n\geq 3n_0$, $$\P(\mathsf{G}_1^c\cup \mathsf{G}_3^c)\leq \P\left(\sup_{y\in [-3s,s]}\left|3^{1/3}{\mathsf{W}}_{n;(0,0)}^{(y,1/3)}+3^{4/3}y^2\right|>10s^2\right)\leq \lceil 4\cdot 3^{2/3}s\rceil Ce^{-cs^3}\leq 6^{-1}\,.$$ Similarly for the events $\mathsf{G}_2$ and $\mathsf{G}_4$, in a given application of Theorem \[t:unifmoddev\], we set the parameters $\bm{n}=n,\bm{t_1}=2/3,\bm{t_2}=1, \bm{s}=10s^2,\bm{I}=[-3s+(i-1)3^{-2/3},\max\{-3s+i3^{-2/3},s\}]$ and $\bm{J}=\{0\}$, for some $i\in \{1,2,\cdots,\lceil4\cdot 3^{2/3}s\rceil\}$. The condition on the inclusion for the intervals $\bm{I}$ and $\bm{J}$ is ensured exactly in the same way as before, and the parabolic curvature is bounded above by $40s^2$. Hence, using Theorem \[t:unifmoddev\] and a union bound, it follows that, for $s$ large enough and $n\geq 3n_0$, $$\P(\mathsf{G}_2^c\cup \mathsf{G}_4^c)\leq \lceil 4\cdot 3^{2/3}s\rceil Ce^{-cs^3}\leq 6^{-1}\,.$$ Finally, for $\mathsf{G}_5$, observe that, since paths between two fixed endpoints constrained to stay in a box have smaller weight than does the polymer between these endpoints, we can again use Theorem \[t:unifmoddev\]. For a given application of Theorem \[t:unifmoddev\], take $\bm{n}=n, \bm{t_1}=1/3,\bm{t_2}=2/3, \bm{s}=40s^2,\bm{I}=[-s+(i-1)3^{-2/3},\max\{-s+i3^{-2/3},s\}]$ and $\bm{J}=[-s+(j-1)3^{-2/3},\max\{-s+j3^{-2/3},s\}]$ for $i\in \{1,2,\cdots,\lceil 2\cdot 3^{2/3}s\rceil\}$ and $j\in \{1,2,\cdots,\lceil 2\cdot 3^{2/3}s\rceil\}$. As before, the condition on inclusion for $\bm{I}$ and $\bm{J}$ is satisfied, and the parabolic curvature is at most $3^{4/3}s^2$, which is less than $10s^2$. Thus, applying Theorem \[t:unifmoddev\] and a union bound, we find that, for $n \geq 3n_0$ and $s$ large, $$\P(\mathsf{G}_5^c)\leq \P\left(\sup {\mathsf{W}}_{n;([-s,s],1/3)}^{([-s,s],2/3)}>50s^2\right)\leq \lceil 2\cdot 3^{2/3}s\rceil^2Ce^{-cs^3}\leq 6^{-1}\,.$$ Thus we have $\P(\mathsf{F}) \geq 1/2$ by a union bound. In Case High, we also have $$\P \big( {{\rm High}}({{\rm West}},300s^2) \big) \geq \exp \big\{ -2C s^3 \big\} \, ,$$ because ${{\rm West}}$ is a translate of ${{\rm Mid}}$. Since the interior of ${{\rm West}}$ is disjoint from the regions that dictate the occurrence of $\mathsf{F}$, we see that $$\label{e.fbound} \P \Big( {{\rm High}}({{\rm West}},300s^2) \cap \mathsf{F} \Big) \geq 2^{-1} \exp \big\{ -2C s^3 \big\} \, .$$ When ${{\rm High}}({{\rm West}},300s^2) \cap \mathsf{F}$ occurs, a high weight path connecting $(0,0)$ to $(0,1)$ may be formed by running it through ${{\rm West}}$. Indeed, and as Figure \[f.casehigh\] depicts, let $\rho$ denote a polymer running across, and contained in, ${{\rm West}}$, whose weight is at least $300s^2$. If $x,y \in [-3s,-s]$ are such that $(x,1/3)$ and $(y,2/3)$ are $\rho$’s endpoints, then the path $\rho_{n;(0,0)}^{\leftarrow;(x,1/3)} \circ \rho \circ \rho_{n;(y,2/3)}^{\leftarrow,(0,1)}$ connects $(0,0)$ to $(0,1)$ and has weight at least $-50s^2 + 300s^2 - 50s^2$, in view of the first two conditions that specify $\mathsf{F}$. On the other hand, the final three conditions specifying $\mathsf{F}$ ensure that, when this event occurs, any path from $(0,0)$ to $(0,1)$ whose $x$-coordinate never exceeds $s$ in absolute value has weight at most $50s^2 + 50s^2 + 50s^2$; indeed, the weight of any such path may be represented as a sum of the weights of the three subpaths formed by cutting the path at heights one-third and two-thirds. We thus find that, on ${{\rm High}}({{\rm West}},300s^2) \cap \mathsf{F}$, any path from $(0,0)$ to $(0,1)$ that remains in ${{\rm Strip}}$ has weight at most $150s^2$; at the same time, a path of weight at least $200s^2$ connects these two points. Thus, we see that any polymer from $(0,0)$ to $(0,1)$ has maximum transversal fluctuation at least $s$ in this event. By (\[e.fbound\]), we find that $$\label{e.conchigh} \P\left( \min \left\{ \mathrm{TF}(\rho) : \rho\in \Phi_{n;(0,0)}^{(0,1)} \right\} \geq s \right) \, \geq \, 2^{-1} \exp \big\{ -2C s^3 \big\} \, .$$ Suppose now instead that Case Low holds. We will argue that $$\label{e.tbargued} \P \Big( {\mathsf{W}}_{n;(0,0)}^{(0,1)} \geq 1000 s^2 \, , \, \neg \, {{\rm High}}\big( [-s,s] \times [0,1] , 900 s^2 \big) \Big) \geq 2^{-1} \exp \big\{ - C s^3 \big\} \, ,$$ where $\neg\, A$ denotes the complement of the event $A$. Before we do so, we show that the event on this left-hand side entails that any polymer from $(0,0)$ to $(0,1)$ must leave the strip $[-s,s] \times [0,1]$; thus, (\[e.conchigh\]) holds in Case Low, even when the factor of $2$ is omitted from the right-hand exponential. When the last left-hand event occurs, any path from $(0,0)$ to $(0,1)$ that remains in the strip has weight at most $900s^2$. At the same time, the weight of any polymer from $(0,0)$ to $(0,1)$ is at least $1000s^2$. It is thus impossible for any polymer to remain in the strip. To derive (\[e.tbargued\]), note that, because ${{\rm North}}$ and ${{\rm South}}$ are translates of ${{\rm Mid}}$, Case Low entails that $$\P \Big( {{\rm High}}({{\rm South}},300s^2) \cup {{\rm High}}({{\rm Mid}},300s^2) \cup {{\rm High}}({{\rm North}},300s^2) \Big) < 3 \exp \big\{ -2C s^3 \big\} \, .$$ The bound (\[e.wthou\]) then yields (\[e.tbargued\]), since $3 \exp \big\{ -2 C s^3 \big\} \leq 2^{-1} \exp \big\{ - C s^3 \big\}$ for all $s$ large enough. The bound (\[e.conchigh\]) has been derived in both of the cases, so that proof of Proposition \[p:lowbndTF\] is complete. [^1]: The first author is supported by NSF grant DMS-$1512908$.
{ "pile_set_name": "ArXiv" }
--- abstract: 'A low-energy enhancement of the $\gamma$-ray strength function in several light and medium-mass nuclei has been observed recently in $^3$He-induced reactions. The effect of this enhancement on $(n,\gamma$) cross-sections is investigated for stable and unstable neutron-rich Fe, Mo and Cd isotopes. Our results indicate that the radiative neutron capture cross sections may increase considerably due to the low-energy enhancement when approaching the neutron drip line. This could have non-negligible consequences on r-process nucleosynthesis calculations.' author: - 'A. C. Larsen' - 'S. Goriely' title: 'Impact of a low-energy enhancement in the $\gamma$-ray strength function on the radiative neutron-capture' --- Introduction ============ The $\gamma$-ray strength function or radiative strength function (RSF) characterizes average electromagnetic decay properties of excited nuclei. This quantity is important for describing the $\gamma$-emission channel in nuclear reactions. It is also indispensable for calculating nuclear reaction cross sections and reaction rates relevant for astrophysical applications. Recent studies [@ar07; @go98] clearly show the importance of a precise description of the $\gamma$-ray strength function at low energies, especially for a proper understanding of the nucleosynthesis of the elements heavier than iron by the rapid neutron-capture process (r-process). The r-process nucleosynthesis is called for to explain the origin of about half of the stable nuclides heavier than iron observed in nature and is believed to result from an extremely large neutron irradiation on timescales of the order of about one second. So far, however, the astrophysical site hosting such an r-process remains unknown. Although an $(n,\gamma$)–($\gamma,n$) equilibrium might take place in an environment with high neutron density and high temperature (in which case the r-abundance distribution remains rather insensitive to the reaction rates), more general r-process simulations require a reliable determination of the radiative neutron capture rates for all nuclei involved [@ar07]. Indeed, it should be kept in mind that even if the thermodynamic conditions of the r-process site remain unknown, the assumption of an $(n,\gamma$)–($\gamma,n$) equilibrium can only be tested if the neutron capture, $\beta$-decay and photodisintegration rates for all neutron-rich nuclei synthesized during such a process can be estimated reliably [@go96]. Furthermore, in specific sites such as the decompression of neutron star matter or the so-called cold neutrino-driven wind, neutron captures are in competition with $\beta$-decays, not with photodisintegrations, so that no $(n,\gamma$)–($\gamma,n$) equilibrium can be achieved and the final r-abundance distribution may well be sensitive to reaction rates. Finally, in any site, the final abundance distribution is likely to depend more or less on the freeze-out conditions for which the $(n,\gamma$)–($\gamma,n$) competition comes out of equilibrium. Under such conditions, an accurate and reliable determination of all the ingredients of relevance in the calculation of the neutron capture rates, including in particular the $\gamma$-strength function, is required. The nuclear physics group in Oslo has performed measurements on the $\gamma$-ray strength functions below neutron threshold of various light and medium-mass nuclei [@Fe_Alex; @Fe_Emel; @Mo_RSF; @V; @Sc]. These data have revealed an unexpected increase in the $\gamma$-decay probability at low $\gamma$-ray energies. This enhancement is seen to be present typically for $E_\gamma \leq 3 $ MeV. In contrast to other soft resonances observed previously such as the $M1$ scissors mode [@PragueM1; @SchillerM1], the physical origin of the enhancement remains unknown. There is, for the time being, no established theory that is able to explain this *upbend* phenomenon. For nuclei close to the valley of stability, one might expect that the low-energy enhancement would have little influence on the neutron-capture cross section. Naturally, the most important energy region for the neutron-capture reaction is in the vicinity of the neutron separation energy $S_n$, and low-lying structures in the $\gamma$-ray strength function would probably have a relatively small effect. However, for neutron-rich nuclei approaching the neutron drip line, the neutron separation energy rapidly decreases and enters the energy region where the enhanced strength is observed. This work aims at investigating how such a very low-energy strength enhancement may influence the neutron-capture cross section of exotic neutron-rich nuclei. Since this upbend phenomenon has been clearly seen in Mo isotopes, the present study will focus on Mo. The Mo case is also of particular interest due to the many observational constraints that can help us to determine the full RSF; these include the resonance spacing and the average total radiative width at the neutron separation energies, photoneutron cross-section data, and the measurements by the Oslo group. In Sect. \[sect\_exp\], we will describe the experimental results obtained by the Oslo method and the parameterizations used to model the low-energy $E1$ strength, in particular in the vicinity of the upbend structure. In Sect.  \[sect\_res\], the neutron-capture cross sections and the corresponding astrophysical rates are estimated for the Mo nuclei, as well as Fe and Cd isotopic chains. Finally, implications of these results and conclusions are discussed in Sect. \[sect\_conc\]. Experimental and theoretical description of the upbend structure {#sect_exp} ================================================================ Experimental results -------------------- During the last decade, the Oslo group has developed an experimental method capable of extracting information on the nuclear level densities (NLD) and $\gamma$-ray strength functions by analyzing particle-$\gamma$ coincidence data from neutron pickup ($^3$He,$\alpha\gamma$) and inelastic scattering ($^3$He,$^3$He$^{\prime}$$\gamma$) reactions. Details about the method can be found in [@Schiller00]. One major result obtained through the Oslo method concerns an increase of the RSF at decreasing photon energy. This upbend structure has been observed in $^{44,45}$Sc, $^{50,51}$V, $^{56,57}$Fe, and $^{93-98}$Mo at energies typically smaller than 3 MeV. Heavier isotopes of Sn, Sm, Dy, Er and Yb for which similar experiments have been conducted do not show such a low-energy behavior. At the moment, only the Fe results have been confirmed by another experimental technique (the two-step cascade method, see [@Fe_Alex]), and theoretically there is no model that can provide possible explanations. Even the multipolarity of the strength remains unknown. Guttormsen et al. [@Mo_RSF] showed that in the case of an $E1$ strength, the reduced strength in the $1-3$ MeV range would correspond to an average $B(E1)$ value of 0.02 $e^2$fm$^2$ (i.e. about 0.07% of the $E1$ sum rule), in case of an $M1$, $B(M1)\simeq 2~\mu_N^2$ which is 3 to 4 times larger than the observed strength to mixed-symmetry $1^+$ states around 3 MeV [@Fransen1; @Fransen2], and in case of an $E2$ strength, $B(E2)\simeq 15000~e^2$fm$^4$ which is 5 to 15 times the strength of the (de)excitation of the first $2^+$ states in the even Mo isotopes. Recent results on $^{60}$Ni investigated with the two-step cascade method applied on $(p,2\gamma)$ data [@Alex_Ni], indicate that there is an upbend in the $M1$ component of the RSF and possibly also in the $E1$ component. However, one should note that this specific nucleus has only positive-parity states below $E_x \approx 4$ MeV, and one of the conclusions in Ref. [@Alex_Ni] is in fact that the low-energy enhancement is probably due to secondary $M1$ transitions in this excitation-energy region, while the Oslo results reveal the RSF in the quasi-continuum region above $E_x \approx 4$ MeV. Thus, the multipolarity and the electromagnetic character of the upbend in other cases are still unknown. Although we cannot exclude any of the above-mentioned posibilities, we will assume that it can be associated with $E1$ $\gamma$-ray transitions for the nuclei studied in this work. For the Mo case, additional information exists for the strength function below the neutron threshold. It concerns the measured $E1$ strength for $^{93,95}$Mo at $\approx 7$ MeV ([@RIPL] and references therein), and for $^{92,94,96,98,100}$Mo from $(\gamma,\gamma')$ experiments [@Rusev]. The latter, however, shows an RSF with a shape quite different (convex rather than concave) from the one extracted from the Oslo data. In addition, the absolute value of the data presented in Ref. [@Rusev] appears to overestimate the experimental average radiative width $\left< \Gamma_{\gamma}(S_n) \right>$. The reason for this disagreement is not yet understood, but the explanation might be connected to the different reaction parameters and selectivity of the populated states (such as restrictions on the spin range and/or parity) compared to the Oslo data. Also, these data reach energies down to $E_{\gamma} \approx 5-6$ MeV only, depending on the isotope studied. We will therefore in the following use the results from the Oslo method to constrain the $E_\gamma \rightarrow 0$ limit. The existence of the upbend structure could also be questioned on the basis of the various assumptions related to the Oslo method. In particular, one fundamental assumption behind the extraction procedure relies on the Brink hypothesis [@br55], which states that collective excitation modes built on excited states have the same properties as those built on the ground state. This hypothesis allows to express the probability of the $\gamma$ decay in the statistical regime as being proportional to a separable product of the final-state level density and the RSF. Although the Brink assumption can be questionable, in particular from the point of view of some models such as the Fermi liquid theory [@ka83], it should be emphasized that the presence of the upbend structure has been tested against such an assumption. Specifically, the RSFs for $^{56,57}$Fe [@Fe_Alex], $^{96,98}$Mo [@Mo_RSF], and $^{45}$Sc [@Sc] have been determined for various initial energies and shown to present the upbend structure for all the excitation bins studied. From such a test, it can be inferred that the eventual temperature-dependence of the RSF at low energy is small with respect to the strength of the upbend structure. In this low-energy region, the validity of the Brink hypothesis remains a fundamental open question. Uncertainties in the subtraction procedure of the Oslo method, including the estimate of the statistical multiplicity have also been studied in [@Mo_reanalyzed] and shown not to affect the RSF significantly and definitely not the conclusion regarding the existence of the upbend pattern. Finally, it should also be stressed that quantitatively the experimental determination of the NLD and RSF is model-dependent, as the method enables a unique determination only of the *functional form* of the NLD and RSF. In order to obtain the absolute value of the total level density (and hence the RSF) from the measured data, the experimental NLD needs to be normalized to the total level density at the neutron separation energy $S_n$, which in turn is derived from the neutron resonance spacing. The parity and spin distributions need to be known to estimate the total level density from the resonance spacing. As shown in Ref. [@go08], uncertainties within a factor of two can still affect this NLD normalization procedure and consequently could affect the low-energy RSF. This uncertainty has been considered here. In Fig. \[fig00\], the NLDs of $^{93-98}$Mo have been renormalized to the calculated total level densities at $S_n$ taken from Ref. [@go08], which were themselves normalized to available experimental s-wave spacings. One exception is $^{98}$Mo, for which the newly recommended value of the s-wave spacing $D_0 = 60(10)$ eV [@RIPL3] has been adopted. This new value leads to $\rho(S_n) = 1.38(69) \times 10^5$ MeV$^{-1}$ if calculated with the same prescription as in [@Mo_RSF], and to $\rho(S_n) = 1.90(95) \times 10^5$ MeV$^{-1}$ following [@go08]. It is seen that both normalizations reproduce well the known, discrete levels taken from [@ENSDF]. In Fig. \[fig01\] the corresponding RSFs for both normalizations are displayed, and it is seen how the slope of the RSFs is changed. Note that for both cases the total RSF is normalized in absolute value to the average total radiative width $\left< \Gamma_{\gamma}(S_n) \right>$. For the cases of $^{93,95,97,98}$Mo the upbend structure is clearly present and relatively strong, for both normalizations of the NLD, while for $^{94,96}$Mo the enhancement is reduced. However, this new normalization does not question the presence of the upbend in the RSF. ![image](MoNLD_testnorm.pdf){width="1.5\columnwidth"} ![image](testnorm_MoRSF.pdf){width="1.5\columnwidth"} In summary, the Oslo method has now been widely tested to confirm its capacity to determine the NLD and RSF. It has proven to be an excellent procedure and there is, at present, no reason not to trust the low-energy RSF data showing an upbend pattern. For that reason, we will now assume that this structure is present in all elements lighter than typically Cd (no upbend has been seen in Sn isotopes [@Sn]), and discuss how it can be modeled as an $E1$ strength. Model parameterizations ----------------------- To describe the $\gamma$-ray strength function with the enhancement observed at low energy, we consider different ways to model its contribution to the total $E1$ strength. More specifically, we have applied the widely used Generalized Lorentzian (GLO) model [@ko87; @ko90] for the $E1$ strength, and modified it in order to describe the observed upbend structure. The GLO model makes use of a temperature dependence that increases the $E1$ de-excitation strength at low energies and gives an $E_\gamma \rightarrow 0$ non-zero limit, but no upbend pattern. This non-zero limit was first introduced by Kadmenski[ĭ]{}, Markushev and Furman in 1983 on the basis of theoretical calculations within the Fermi liquid model [@ka83]. This was done in order to take into account quasi-particle fragmentation of the $E1$ strength, and to explain the observed low-energy data on the $\gamma$-ray strength function of several medium-mass and heavy nuclei studied with $(n,\gamma\alpha)$ and $(n,\gamma f)$ reactions (see [@ka83] and references therein). Later, Kopecky and Chrien [@ko87], and Kopecky and Uhl [@ko90] found it necessary to introduce a temperature dependence in order to describe the strength of primary $\gamma$-ray data from average resonance capture (ARC) reactions, and introduced the GLO model [@ko87; @ko90]. This model relies also on the theory of Fermi liquids and accounts for microscopic properties of the GEDR. The advantage of the GLO model is its capability to reproduce both photoabsorption cross-section data as well as the above-mentioned sub-threshold data reasonably well. The GLO strength function is given by [@ko90] $$\begin{aligned} &&f_{\rm GLO}(E_{\gamma},T_f) = \frac{1}{3\pi^2\hbar^2c^2}\sigma_{E1}\Gamma_{E1} \times \\ \nonumber && \left[\frac{ E_{\gamma} \Gamma(E_{\gamma},T_f)}{(E_\gamma^2-E_{E1}^2)^2 + E_{\gamma}^2 \left[\Gamma (E_{\gamma},T_f)\right]^2} + \;0.7\frac{\Gamma(E_{\gamma}=0,T_f)}{E_{E1}^3} \right], \label{eq:GLo}\end{aligned}$$ where $\sigma_{E1}$, $\Gamma_{E1}$, and $E_{E1}$ are the Giant Electric Dipole Resonance (GEDR) peak cross section, width, and centroid energy, respectively. The energy- and temperature-dependent width reads $$\Gamma(E_{\gamma},T_f) = \frac{\Gamma_{E1}}{E_{E1}^2} (E_{\gamma}^2 + 4\pi^2T_f^2), \label{eq:edepwidth}$$ identical to the prediction of [@ka83]. Here, the first term reflects the spreading of particle-hole states into more complex configurations, and the second term accounts for collisions between quasiparticles. The spreading width thus depends on the nuclear temperature of the final states $T_f$. The Oslo method is based on the Brink hypothesis, which means no dependence on the excitation energy and thus on the nuclear temperature $T_f$ of final states in the RSF. Introducing a constant temperature is, however, not in contradiction with the Brink hypothesis or the extraction procedure to get the NLD and RSF from the coincidence data. We have therefore considered $T_f$ to be constant in the GLO model. We have chosen $T_f = 0.30$ MeV, to give a reasonable global agreement with experimental data of $^{93-98}$Mo for $E_{\gamma} \gtrsim 3$ MeV, as illustrated in Fig. \[fig02\]; this first adaptation of the GLO model is in the following referred to as GLO-lo. For the GEDR parameters we take experimental values from [@RIPL; @iaea00] when available (for $^{98}$Mo a new improved determination has been performed), and interpolated values from the even-even $^{92,94,96,98}$Mo for the missing odd isotopes $^{93,95,97}$Mo. For the heavier Mo nuclei ($A>98$), we used the systematics recommended in Ref. [@RIPL]. The final GEDR parameters for $^{93-98}$Mo are summarized in Table \[tab:parameters\]. ----------- ---------- --------------- --------------- Nucleus $E_{E1}$ $\sigma_{E1}$ $\Gamma_{E1}$ (MeV) (mb) (MeV) $^{93}$Mo 16.59 173.5 4.82 $^{94}$Mo 16.36 $185.0$ 5.50 $^{95}$Mo 16.28 $185.0$ 5.76 $^{96}$Mo 16.20 $185.0$ 6.01 $^{97}$Mo 16.00 $187.0$ 5.98 $^{98}$Mo 16.50 $ 220.0$ 8.00 ----------- ---------- --------------- --------------- : Parameters used for the GEDR strength of $^{93-98}$Mo. \ \[tab:parameters\] We have chosen two ways to model the upbend: - [introducing a low-lying resonance of the form of a standard Lorentzian (SLO).]{} - [modifying the energy-dependent width of the GLO model.]{} It should be stressed that both approaches are completely phenomenological, since there is at present no proper theoretical description of the upbend structure. For the first approach, the upbend is given by a low-lying resonance described by an SLO shape: $$f_{\mathrm{up1}}(E_\gamma)=\frac{1}{3\pi^2\hbar^2c^2} \frac{\sigma_{\mathrm{up1}}E_\gamma\Gamma_{\mathrm{up1}}^2} {(E_\gamma^2-E_{\mathrm{up1}}^2)^2+E_\gamma^2\Gamma_{\mathrm{up1}}^2}, \label{eq:lorup}$$ located at a resonance energy of $E_{\mathrm{up1}}=1.5$ MeV with a full width at half maximum $\Gamma_{\mathrm{up1}}=1.5$ MeV and a peak cross section $\sigma_{\mathrm{up1}}=0.05$ mb. This resonance is added to the GLO-lo strength in order to make the total RSF fit with the low-energy data as well; this model is referred to as GLO-up1 and its Mo description is shown in Fig. \[fig02\]. For the second approach, we have modified the temperature-dependent width of the GLO model in the following way: $$\Gamma_{\rm up2}(E_{\gamma},T_f) = \frac{\Gamma_{E1}}{E_{E1}^2} \left[E_{\gamma}^2 + \frac{4\pi^2T_f^2E_{E1}}{(E_{\gamma}+\delta)}\right], \label{eq:glowidthmod}$$ where the introduction of an $E_{\gamma}^{-1}$ dependence in the second term enables an increasing RSF for decreasing $E_{\gamma}$. The constant parameter $\delta = 0.05$ MeV is applied to ensure a finite value of $f_{\rm GLO}$ for $E_\gamma \rightarrow 0$. In this approach, the temperature $T_f $ has also been assumed to remain constant and the value of $T_f= 0.16$ MeV has been adjusted to reproduce at best the experimental Mo RSF, as seen in Fig. \[fig02\]. Note that the $T_f$ value is obviously different from the value extracted wihtin the GLO-lo model because of the new functional (\[eq:glowidthmod\]) considered for the energy-dependent width. This modified width allows us to use for the RSF a closed form identical to Eq. (\[eq:GLo\]) that makes it possible to describe the RSF at all energies and also to reproduce the upbend structure phenomenologically without calling for the presence of an extra low-lying resonance. We call this model GLO-up2. Since most of the reaction calculations are performed with the original version of the GLO model [@ko90], we also consider the corresponding model with a variable temperature $T_f$, estimated from the well-known expression [@RIPL] $$T_f = \sqrt{(E^* -\Delta - E_\gamma)/a}, \label{eq:temp}$$ where $E^*$ is the initial excited state in the compound nucleus (for neutron capture at incoming neutron energy $E_n$, $E^*=E_n+S_n$), $\Delta$ is a pairing correction, and $a$ is the level density parameter at $S_n$. For all $E1$ strength functions considered here, we have also added an $M1$ strength described by a Lorentzian shape: $$f_{M1}(E_\gamma)=\frac{1}{3\pi^2\hbar^2c^2} \frac{\sigma_{M1}E_\gamma\Gamma_{M1}^2} {(E_\gamma^2-E_{M1}^2)^2+E_\gamma^2\Gamma_{M1}^2}. \label{eq:M1}$$ Here, $\sigma_{M1}$, $\Gamma_{M1}$, and $E_{M1}$ are the peak cross section, width, and centroid energy, respectively, of the Giant Magnetic Dipole Resonance (GMDR) related to spin-flip transitions between major shells . The peak cross section is normalized to the $E1$ strength function at $E_{\gamma} = 7$ MeV as described in Ref. [@RIPL], while the width and the peak position is determined from systematics [@RIPL]. With such a parameterization the $M1$ contribution to the total RSF remains low compared to the $E1$ contribution. The GLO-lo, GLO-up1 and GLO-up2 models are shown together with experimental data on $^{93-98}$Mo in Fig. \[fig02\]. For $^{98}$Mo, also the GLO model with $E_n = 1$ MeV is displayed. It is seen that both the GLO-up1 and the GLO-up2 models give reasonable agreement with data at all energies, while GLO-lo fails to describe the low-energy region. ![image](MoRSF_GLO.pdf) In the following section, the impact of the upbend structure is estimated by comparing the above-mentioned models (GLO-up1, GLO-up2, and GLO-lo) and the original and widely used $T$-dependent GLO model. Results {#sect_res} ======= We now perform the calculation of $(n,\gamma)$ cross-sections and astrophysical rates with the code TALYS [@TALYS; @go08b] in order to study the impact the upbend structure might have on the radiative neutron capture for stable as well as neutron-rich nuclei. By default, all calculations are performed using the nuclear structure properties determined within the HFB-17 mass model [@go09], the neutron-nucleus optical potential of Ref. [@ko03] and the NLD obtained within the combinatorial method [@go08]. The latter model not only reproduces quite accurately resonance spacing data, but also the energy-dependence of the total level density extracted consistently for the Mo isotopes through the Oslo method. The Oslo data and the theoretical NLD are compared in Fig. 9 of Ref. [@go08]. To remain coherent, it is of prime importance to use the same NLD prescription and normalizing values at the neutron separation energy for the cross section calculation as those used in the extraction of the RSF with the Oslo method. As shown in Fig. \[fig01\], different NLD models lead to different experimental RSFs. ![\[fig04\] (Color online) Comparison of experimental [@ka64; @mu76] and calculated $^{97}$Mo$(n,\gamma)^{98}$Mo cross section obtained with the GLO (solid line), GLO-lo (blue dashed line), GLO-up1 (red dotted line) and GLO-up2 (green, dash-dot line) models. ](fig_97Mong_glo.pdf) We compare in Fig. \[fig04\] experimental $^{97}$Mo$(n,\gamma)^{98}$Mo data with the calculated cross section obtained with the original GLO model, the GLO-lo model with $T_f=0.3$ MeV, and the two parameterizations of the upbend, GLO-up1 and GLO-up2. It is seen from Fig. \[fig04\] that the increase in the predicted cross section using our description of the upbend structure for this specific stable Mo isotope can reach about 20% if we adopt the GLO-up1 parameterization with respect to the GLO-lo one, and roughly 50% if we adopt the GLO-up2 model. Similar results are obtained for the other Mo isotopes. In general, the upbend structure improves the agreement with experimental data, although there is clearly some strength missing, partially due to a possible extra contribution lying in the 8-10 MeV region of $^{98}$Mo that is not properly described by the present parameterizations (see Fig. \[fig02\]). Note, however, that no effort is done here to reproduce experimental data perfectly, for example by modifying the NLD model or the RSF parameterization. When considering the original GLO model, the estimated cross section is slightly higher due to the high temperatures $T_f$ encountered in the compound system in comparison with the constant values adopted. On the basis of the input models described above, we now perform calculations on the Maxwellian-averaged neutron capture rates of astrophysical interest for the full isotopic chains of Mo, as well as Fe and Cd up to the neutron drip line. The GLO-up1, GLO-up2, GLO-lo predictions are compared with the widely used GLO estimates in Figs. \[fig05\] – \[fig07\] for a temperature of $T=10^9$ K typical of the r-process nucleosynthesis [@ar07]. As already demonstrated in Fig. \[fig04\], close to the stability line the upbend structure has a relatively small influence. However, for exotic neutron-rich nuclei the impact may become large, essentially due to the low neutron separation energies allowing only for $\gamma$-decays with energies lower than typically 2 MeV. In this case, the strength in the low-energy region dominates the decay. ![\[fig05\] (Color online) Ratios of Maxwellian-averaged $(n,\gamma)$ reaction rates at $T=10^9$ K for the Fe, Mo and Cd isotopic chains up to the neutron drip line, using the GLO-lo and GLO model.](fig_sigv_glo_lo.pdf) ![\[fig06\] (Color online) Same as Fig. \[fig05\] for the GLO-up1 and the GLO model.](fig_sigv_glo_up1.pdf) ![\[fig07\] (Color online) Same as Fig. \[fig05\] for the GLO-up2 and the GLO model.](fig_sigv_glo_up2.pdf) In Fig. \[fig05\], the rates obtained using the GLO-lo with constant temperature of $T_f=0.3$ MeV are compared to the frequently used standard GLO model with variable temperature as defined in Eq. (\[eq:temp\]). For nuclei close to the valley of stability, it is seen that the constant-temperature approach gives lower rates than the original GLO model, which is easy to understand from the higher absolute value of the GEDR tail when using a variable temperature which is found to be higher than the 0.3 MeV considered in the GLO-lo model (see Fig. \[fig02\] for $^{98}$Mo). However, when approaching the neutron drip line, the rates of the GLO-lo model become comparable and even larger than the ones using the GLO model. This is due to the fact that the neutron separation energy drops rapidly and so does the temperature $T_f$, at least for neutron incident energies of about 100 keV (corresponding to the $T=10^9$ K temperature considered here). The original GLO model when applied to the neutron capture by exotic neutron-rich nuclei can therefore be approximated by the $T_f=0$ case. We see from Fig. \[fig05\] that assuming a constant temperature $T_f=0.3$ MeV (i.e the GLO-lo case) can give an order of magnitude increase in the reaction rates for such exotic nuclei. Including the upbend structure through the GLO-up1 model may give another significant increase of the rates as shown in Fig. \[fig06\]. In particular, the rates for neutron-rich Cd isotopes gain an additional order of magnitude due to the low-energy RSF contribution that become effective as soon as $S_n$ drops after crossing the closed neutron shell at $N=82$. As demonstrated in Fig. \[fig07\], the GLO-up2 parameterization gives a similar large increase of the rates with respect to the traditional calculation based on the GLO model. The predictions are even larger than when considering the GLO-up1 model. Similar conclusions can be drawn for the Fe and Mo isotopes. In general, we see that the influence of the upbend structure on the $(n,\gamma)$ cross sections and thus the reaction rates becomes more and more important as the number of neutrons increases. In particular, as soon as a major neutron shell is crossed, the neutron separation energy $S_n$ drops and the RSF in the vicinity of the upbend structure starts to play a major role in the radiative decay. It can be seen that the combination of the upbend structure and applying a constant temperature may lead to an increase of the reaction rates by up to a factor of 300. This increase is observed in all the isotopic chains studied here when applying the GLO-up2 model. These calculations show that a proper understanding of the $E_\gamma \rightarrow 0$ limit of the RSF can be of crucial importance in the determination of radiative neutron-capture cross sections for exotic neutron-rich nuclei. This effect could have a non-negligible impact on the neutron captures that can potentially take place in astrophysical environments characterized by high neutron densities, in particular during the r-process nucleosynthesis. Conclusions {#sect_conc} =========== As shown experimentally by the Oslo group, the RSF at very low $\gamma$-ray energy might be characterized by a significant enhancement with respect to the usual rapidly decreasing $\gamma$-decay strength. This upbend structure has been observed systematically in V, Sc, Fe and Mo isotopes at energies typically smaller than 3 MeV, but is absent in elements heavier than Sn. The very origin of this extra strength remains unexplained theoretically. The impact of this upbend structure is found to be relatively small on the neutron capture cross section of stable nuclei, since it modifies the RSF in an energy region which hardly takes part in the reaction mechanism. However, for exotic neutron-rich nuclei, this effect becomes significant and could potentially increase the reaction rates of astrophysical relevance by one or even two orders of magnitude. This effect is particularly pronounced for nuclei with a low neutron separation energy, crossing a major neutron shell. This effect could have a non-negligible impact on the neutron capture rates essential for the r-process nucleosynthesis. [99]{} M. Arnould, S. Goriely, and K. Takahashi, Phys. Rep. [**450**]{}, 97 (2007). S. Goriely, Phys. Lett. [**B436**]{}, 10 (1998). S. Goriely, M. Arnould, Astron. Astrophys. [**312**]{} 327 (1996). A. Voinov, E. Algin, U. Agvaanluvsan, T. Belgya, R. Chankova, M. Guttormsen, G.E. Mitchell, J. Rekstad, A. Schiller and S. Siem, Phys. Rev. Lett [**93**]{}, 142504 (2004). E. Algin, U. Agvaanluvsan, M. Guttormsen, A. C. Larsen, G. E. Mitchell, J. Rekstad, A. Schiller, S. Siem, and A. Voinov, Phys. Rev. C [**78**]{}, 054321 (2008). M. Guttormsen, R. Chankova, U. Agvaanluvsan, E. Algin, L.A. Bernstein, F. Ingebretsen, T. L[ö]{}nnroth, S. Messelt, G.E. Mitchell, J. Rekstad, A. Schiller, S. Siem, A.C. Sunde, A. Voinov and S. [Ø]{}deg[å]{}rd, Phys. Rev. C [**71**]{}, 044307 (2005). A. C. Larsen, R. Chankova, M. Guttormsen, F. Ingebretsen, T. L[ö]{}nnroth, S. Messelt, J. Rekstad, A. Schiller, S. Siem, N. U. H. Syed, A. Voinov, and S. W. [Ø]{}deg[å]{}rd, Phys. Rev. C **73, 064301 (2006). A. C. Larsen, M. Guttormsen, R. Chankova, F. Ingebretsen, T. L[ö]{}nnroth, S. Messelt, J. Rekstad, A. Schiller, S. Siem, N. U. H. Syed, and A. Voinov, Phys. Rev. C **76, 044303 (2007). M. Krticka, F. Becvár, J. Honzátko, I. Tomandl, M. Heil, F. Käppeler, R. Reifarth, F. Voss, K. Wisshak, Phys. Rev. Lett. [**92**]{}, 172501 (2004). A. Schiller, A. Voinov, E. Algin, J. A. Becker, L. A. Bernstein, P. E. Garrett, M. Guttormsen, R. O. Nelson, J. Rekstad, S. Siem, Phys. Lett. B [**633**]{}, 225 (2006). A. Schiller, L. Bergholt, M. Guttormsen, E. Melby, J. Rekstad, S. Siem, Instrum. Methods Phys. Res. A [**447**]{} 494 (2000). C. Fransen [*et al*]{}., Phys. Rev. C **67, 024307 (2003). C. Fransen [*et al*]{}., Phys. Rev. C **70, 044317 (2004). A. Voinov, S. M. Grimes, C. R. Brune, M. Guttormsen, A. C. Larsen, T. N. Massey, A. Schiller, and S. Siem, Phys. Rev. C **81, 024319 (2010). T. Belgya, O. Bersillon, R. Capote, T. Fukahori, G. Zhigang, S. Goriely, M. Herman, A. V. Ignatyuk, S. Kailas, A. Koning, P. Oblozinsky, V. Plujko and P. Young, *Handbook for calculations of nuclear reaction data*, RIPL-2. **IAEA-TECDOC-1506** (IAEA, Vienna, 2006); also available online at http://www-nds.iaea.org/RIPL-2/ G. Rusev et al., Phys. Rev. C [**79**]{} 061302 (2009). D. M. Brink, PhD thesis, Oxford University (1955). S. G. Kadmenski[ĭ]{}, V.P. Markushev, and V.I. Furman, Sov. J. Nucl. Phys. [**37**]{}, 165 (1983). Ê M. Guttormsen, R. Chankova, U. Agvaanluvsan, E. Algin, L. A. Bernstein, F. Ingebretsen, T. L[ö]{}nnroth, S. Messelt, G.E. Mitchell, J. Rekstad, A. Schiller, S. Siem, A.C. Larsen, A. Voinov and S. [Ø]{}deg[å]{}rd, Los Alamos preprint server:\ <http://xxx.lanl.gov/abs/0801.4667>. S. Goriely, S. Hilaire, and A.J. Koning, Phys. Rev. C **78**, 064307 (2008). R. Capote, et al., Nuclear Data Sheets, [**110**]{}, 3107 (2009). See also *http://www-nds.iaea.org/RIPL-3/.*********** Data extracted using the NNDC On-Line Data Service from the ENSDF database. U. Agvaanluvsan, A. C. Larsen, R. Chankova, M. Guttormsen, G. E. Mitchell, A. Schiller, S. Siem, and A. Voinov, Phys. Rev. Lett. **102 , 162504 (2009). J. Kopecky and R. E. Chrien, Nucl. Phys. **A468, 285 (1987). J. Kopecky and M. Uhl, Phys. Rev. C [**41**]{}, 1941 (1990). Photonuclear data for applications; cross sections and spectra 2000, IAEA-Tecdoc-1178. A. Bohr and B. Mottelson, [*Nuclear Structure*]{}, (Benjamin, New York, 1975), Vol. II, p. 636. A. J. Koning, S. Hilaire, and M. C. Duijvestijn, *“TALYS-1.0”*, in Proceedings of the International Conference on Nuclear Data for Science and Technology, April 22-27, 2007, Nice, France. Editors: O. Bersillon, F. Gunsing, E. Bauge, R. Jacqmin, and S. Leray, EDP Sciences, 211 (2008). S. Goriely, S. Hilaire, A.J. Koning, *Astron. Astrophys.*, **487**, 767 (2008). S. Goriely, N. Chamel, and J.M. Pearson, Phys. Rev. Lett. **102**, 152503 (2009). A. J. Koning and J.-P. Delaroche, Nucl. Phys. [**A713**]{}, 231 (2003). H. Beil, R. Bergère, P. Carlos, A. Leprêtre, A. De Miniac, A. Veyssière, Nucl. Phys. **A227**, 427 (1974). S.¬P. Kapchigashev, Yu. P. Popov, Neutron Interactions Conf., Dubna, p.104 (1964). A. R. Del. Musgrove, B. J. Allen, J. W. Boldeman, R. L. Macklin, Nucl. Phys. [**A270**]{}, 108 (1976).****
{ "pile_set_name": "ArXiv" }
--- abstract: 'Using a hydrodynamical model we study how the order of phase transition in the equation of state of strongly interacting matter affects single particle spectra, elliptic flow and higher order anisotropies in Au+Au collisions at RHIC ($\sqrt{s_{\mathrm{NN}}}=200$ GeV energy). We find that the single particle spectra are independent of the order of phase transition and that the fourth harmonic $v_4(p_T)$ shows only a weak dependence in the $p_T$ region where hydrodynamics is expected to work. The differential elliptic flow, $v_2(p_T)$, of baryons shows the strongest dependence on equation of state. Surprisingly the closest fit to data was obtained when the equation of state had a strong first order phase transition and a lattice inspired equation of state fits the data as badly as a purely hadronic equation of state.' address: | Helsinki Institute of Physics, P.O. Box 64, FIN-00014 University of Helsinki, Finland\ and\ Department of Physics, P.O. Box 35, FIN-40014 University of Jyväskylä, Finland author: - Pasi Huovinen title: Anisotropy of flow and the order of phase transition in relativistic heavy ion collisions --- HIP-2005-14/TH relativistic heavy ion collisions, elliptic flow, order of phase transition, hydrodynamic model 25.75.-q, 25.75.Ld, 25.75.Nq Introduction ============ In non-central heavy ion collisions at the Relativistic Heavy Ion Collider (RHIC) of BNL the particle distributions exhibit quite large anisotropies [@Phenix-v2; @Phobos; @STAR-compiled]. The second Fourier coefficient of the azimuthal distribution of particles, so called elliptic flow, has been extensively studied [@flowreview] since it is sensitive to the early dense stage of the evolution [@Sorge]. Recently also higher harmonics have been measured [@STAR-compiled; @STAR-v4]. It has been claimed that they should be even more sensitive to the initial configuration of the system [@Kolbv4]. Ideal fluid hydrodynamics has been particularly successful in describing the observed anisotropy of particles at low $p_T$ in minimum bias collisions [@review; @Peter-review]. This success has been interpreted as a sign of formation of thermalized matter rapidly after the primary collision [@Heinz]. Studies of both single particle spectra and anisotropies have also shown that a reasonable reproduction of data favours an Equation of State (EoS) of strongly interacting matter with a phase transition [@Derek-RQMD; @Heinz2]. The lattice QCD calculations of the EoS of strongly interacting matter support such a scenario by predicting a phase transition at $T_c\approx 170$ MeV temperature. For a physical scenario of two light and one heavier quark, the phase transition is predicted to be a smooth crossover at small values of baryochemical potential. Contrary to naive expectations, lattice QCD predicts that pressure and energy density do not reach their ideal Stefan-Boltzmann values immediately above the critical temperature, but approach them slowly [@Karsch]. At mid-rapidity at collisions at RHIC, the net baryon density is small and the relevant EoS should exhibit a crossover transition. However, so far all hydrodynamical calculations of elliptic flow [@Heinz; @Derek-RQMD; @Heinz2; @letter1; @letter2; @alkutilat; @Hirano1; @Hirano; @Peter-Ralf] have used an EoS with a strong first order phase transition and ideal parton gas to describe the plasma phase. The usual point of view has been that it is unlikely that the details of phase transition would lead to significant dynamical effects [@Peter-review]. This standpoint has been supported by the early calculations [@Blaizot; @Dirk] where it was found that the width of the phase transition region, $\Delta T$, had only little effect on the final flow pattern in one dimensional flow. Thus it was considered safe to claim that the final particle distributions would not be sensitive to $\Delta T$ either. However, full three dimensional expansion is more complicated than one dimensional. It is known that in three dimensional expansion the differential elliptic anisotropy, $v_2(p_T)$, of (anti)protons is sensitive to the existence of phase transition and its latent heat [@Derek-RQMD; @Heinz2; @letter2]. The anisotropy of flow might thus be sensitive to other details of phase transition as well. In this paper we address this possible sensitivity. We use a hydrodynamical model to calculate single particle spectra, elliptic flow and higher order anisotropies in $\sqrt{s_{\mathrm{NN}}} = 200$ GeV Au+Au collisions using four different EoSs with different phase transitions and plasma properties. As a representative of lattice QCD results, we use an EoS based on the thermal quasiparticle model of Schneider and Weise [@Schneider] (EoSqp). This model is tuned to reproduce the lattice QCD EoS and provides a method to extrapolate the results to physical quark masses. To facilitate comparison with earlier calculations we use as reference points the EoSs Q and H used in Refs. [@Heinz; @Heinz2; @letter1; @letter2; @alkutilat]. EoSQ has a first order phase transition between hadron gas and an ideal parton gas whereas EoSH is a hadron gas EoS without any phase transition. To study the effects of the order of phase transition and slow approach to the Stefan-Boltzmann limits separately we also use a simple parametrisation for an EoS (EoST) where the hadron gas and ideal parton gas phases are connected using a hyperbolic tangent function. Such an EoS has a smooth crossover transition but the plasma properties approach their ideal values much faster than in EoSqp. We find that the main sensitivity to the details of an EoS lies in the differential elliptic flow of heavy particles ($m \gtrsim 1$ GeV) where EoSQ with a first order phase transition leads to an anisotropy closest to the data. Surprisingly, the lattice inspired EoSqp reproduces the data as badly as purely hadronic EoS. EoST with a crossover transition leads to almost as good results as EoSQ. Thus hydrodynamical description of elliptic flow does not require a strong first order phase transition, but it does require sufficiently large increase in entropy and energy densities within sufficiently small temperature interval. Equation of State ================= Until recently the lattice QCD calculations were restricted to vanishing net baryon densities, $\mu_B = 0$. Even if there are some recent results for $\mu_B \neq 0$ [@Karsch], we limit our discussion to zero net baryon density for the sake of simplicity. Since we are interested in the behaviour of the collision system at midrapidity at RHIC where net baryon density is small, this approximation is unlikely to cause a large effect. Thermal models suggest that around phase transition temperature, the baryon chemical potential is below 50 MeV [@PBM] corresponding to a quark chemical potential of about 15 MeV. At these small values of $\mu$ the critical temperature is expected to change by less than a percent from that at $\mu = 0$ [@Karsch]. One of our EoSs (EoS Q, see below) also includes extension to non-zero baryon densities. We have checked that for this EoS, the results obtained when the finite baryon density is included in the EoS or approximated by zero, differ by less than two percents. Even if we do not include finite baryon density to the EoS, we still have finite baryon current in our hydrodynamical calculation. This allows us to have different baryon and anti-baryon yields at freeze-out and thus finite net proton yields. So far the lattice QCD calculations with quarks must be done using unphysically large quark masses. The calculated equation of state must therefore be extrapolated to physical quark mass values. For this purpose we use the thermal quasiparticle model of Schneider and Weise [@Schneider]. In this model the lattice QCD results are described in terms of quasiparticles with temperature dependent effective masses and effective number of degrees of freedom. In this approach the EoSs obtained in lattice calculations for pure glue [@lat1] and different number of quark flavours [@lat2] are well reproduced. Since the mass of quarks is an explicit parameter in this model, it is easy to extrapolate the results to physical quark masses. Here we use the quasiparticle EoS for two light quark flavours ($m_{u,d} = 0$) and a heavier strange quark ($m_s \simeq 170$ MeV) to describe the plasma phase of an EoSqp. The quasiparticle model is compared to the lattice results in Fig. \[hilakuva\], where pressure and energy density are shown as a function of temperature. The lattice result for pressure [@lat2] is extrapolated to the continuum limit by assuming a 10% correction, i.e., $p_{\mathrm{cont}}\approx 1.1 p_{\mathrm{lat}}$ [@Schneider], whereas the result for energy density is shown without such an extrapolation. When the quark masses in the quasiparticle model are set temperature dependent as in the lattice calculations, $m_q = 0.4T$ (light quarks) and $m_s = 1.0T$ (heavy quark), the lattice pressure is nicely reproduced (PQP, dashed line). When physical quark masses are chosen (EoSqp, thin solid line), the pressure is larger than with temperature dependent masses. There is no quasiparticle result with temperature dependent masses available for energy density, but comparison of quasiparticle model with physical quark masses (EoSqp) to the lattice shows nice reproduction of the lattice energy density just above $T_c$ but much larger density above $1.5T_c$. This can be partly explained by the missing extrapolation to continuum limit of the lattice result. If one assumes similar 10% correction than for pressure, the difference between lattice and EoSqp is quite similar for both pressure and energy density at high temperature. The parametrized EoST (see later in the text) is included for comparison’s sake and is shown to lead to much larger pressure and energy density than lattice calculations. The large difference between EoSqp and lattice below $T_c$ is intentional and not related to the quasiparticle model. In the present lattice simulations pions turn out too heavy and therefore their contribution to pressure and entropy is strongly suppressed. Thus one may expect the lattice calculations to give too small pressure and entropy density in the hadronic phase below $T_c$. The quasiparticle model reproduces also this feature of the lattice EoS and one has to describe the hadronic phase using another model. We adopt the usual approach of using an EoS of noninteracting hadron resonance gas to describe the hadronic phase. It has been shown that such an EoS describes interacting hadron gas reasonably well at temperatures around pion mass [@Prakash] and that the hadron resonance gas approach reproduces the lattice results below $T_c$ if the same approximations are used in both [@Redlich]. The properties of hadron resonance gas depend on the number of particles included in the model. Here we include all the strange and non-strange particles and resonances listed in the Particle data Book up to 2 GeV mass. The details of constructing this EoS can be found in Ref. [@Josef]. To circumvent our ignorance of the behaviour of the EoS around $T_c$, we use the approach outlined in Ref. [@Renk]: We use the hadron resonance gas EoS up to a temperature $T_c-\Delta T$, the quasiparticle EoS above $T_c$ and interpolate smoothly between these two regimes. In practice we choose the values $T_c=170$ MeV and $\Delta T=5$ MeV and connect the entropy densities of both models using a polynomial function. We require that the first, second and third temperature derivatives of entropy density are continuous to approximate a smooth crossover from hadronic to plasma phase. Below $T_c-\Delta T$ we use the hadron resonance gas values for pressure and energy density. Above this limit we obtain $P(T)$ and $\varepsilon(T)$ by using the thermodynamical relations ${\mathrm{d}}P = s\,{\mathrm{d}}T$ and $\varepsilon = Ts-P$. This EoS is called EoSqp in the following. For comparison’s sake we also carry out the calculations using the EoSQ and H used in Refs. [@Heinz; @Heinz2; @letter1; @letter2; @alkutilat]. EoS H is a purely hadronic EoS without any phase transition. It is constructed by extending the previously described hadron resonance gas EoS to arbitrarily high temperatures. EoSQ, on the other hand, is inspired by a bag model and contains a first order phase transition from hadron gas to ideal parton gas. The hadron phase is again described by an hadron resonance gas and the plasma phase by a gas of ideal massless quarks and gluons with a bag constant. To approximate the effect of the finite strange quark mass we use the number of quark flavours $N_f=2.5$. The phase boundary is determined using the Gibbs criterion $p_{\mathrm{HG}}(T_c) = p_{\mathrm{QGP}}(T_c)$ and the two phases are connected using the Maxwell construction at $T_c=165$ MeV. Details of constructing these two EoSs can be found in Ref. [@Josef]. As can be seen in Fig. \[eoskuva\] the quasiparticle and bag model inspired EoSs (qp and Q, respectively) lead to quite different behaviour around critical temperature. Since we want to study the effects of the order of phase transition and slow approach to ideal Stefan-Boltzmann values separately, we construct yet another EoS. We follow the idea presented in Ref. [@Blaizot] and connect the hadron and parton phases of the EoS by a hyperbolic tangent function. We refine this approach by using hadron resonance gas EoS instead of ideal pion gas to describe the hadron phase. In Ref. [@Blaizot] the entropy densities of hadron and parton phases are connected in this way. This leads to similar behaviour of energy density than the Maxwell construction of EoSQ — above $T_c$ energy density rises above the ideal Stefan-Boltzmann limit and approaches the ideal values from above. There is no sign of this kind of behaviour in the lattice results. Therefore we use hyperbolic tangent to connect the energy density of the different phases instead of entropy density. Energy density is given by $$\varepsilon(T) = \frac{1}{2}\left[\varepsilon_{\mathrm{HRG}}(T) \left(1-\tanh\frac{T-T_c}{\Delta T}\right) +\frac{169}{120}\pi^2T^4 \left(1+\tanh\frac{T-T_c}{\Delta T}\right)\right]$$ where the latter term is the energy density of ideal parton gas with 3 colours and 2.5 quark flavours. We use again $T_c=170$ MeV and make the crossover rapid by choosing $\Delta T = 5$ MeV. After obtaining $\varepsilon(T)$ we again use standard thermodynamical relations, $(\partial S/\partial E)_{N,V} = 1/T$ and $P = Ts-\varepsilon$, to obtain entropy density and pressure as a function of temperature. This EoS is called EoST in the following. All four EoSs are compared in Fig. \[eoskuva\] where entropy and energy density are shown as functions of temperature and pressure and the square of the speed of sound are shown as functions of energy density for each EoS. As can be seen the behaviour of the lattice inspired EoSqp is quite different from the previously used EoSQ with a first order phase transition. The latter has a relatively large latent heat of 1.15 GeV/fm$^3$ whereas in the former the region where the speed of sound is small and the EoS soft is much smaller. The parametrised EoST is a compromise between these two. It can also be seen that above the phase transition region the EoSQ has the largest speed of sound and is therefore hardest whereas the EoSH without phase transition is softest. It is worth noticing that EoSqp depicts a smaller rise in both energy and entropy densities around $T_c$ than what could be expected from lattice calculations. This is not a property of the quasiparticle model used here, but due to the use of hadron resonance gas EoS below $T_c$. As mentioned before, lattice calculations lead to too high pion mass and correspondingly too small densities below $T_c$. If realistic pion masses are used in hadron resonance gas, its pressure and densities are well above lattice results below $T_c$. Initialization {#init} ============== We use the same boost-invariant hydrodynamic code than in Refs. [@letter1; @letter2; @alkutilat] and described in detail in Ref. [@Peterinpitka]. To fix the parameters of the model, we require that the model reproduces the $p_T$ spectra of pions and net protons ($p - \bar{p}$) in most central collisions and the centrality dependence of pion multiplicity at midrapidity. We use net protons instead of protons and/or antiprotons because our model assumes chemical equilibrium to hold down to kinetic freeze-out temperature and is unable to reproduce proton and antiproton yields simultaneously. Some parametrisations to fix the initial density distributions were explored in Ref. [@alkutilat]. None of them reproduces the observed centrality dependence of multiplicity, but a linear combination of them does. Here we use the same combination than in Refs. [@Heinz; @Heinz2]. The local entropy density is taken to scale with a linear combination of the density of participants and binary collisions in the transverse plane with weights of 0.75 and 0.25, respectively. This kind of scaling can be interpreted as particle production from “soft” and “hard” processes. For the sake of simplicity, the initial baryon number density is taken to scale with the number of participants. The initial time of the calculation, $\tau_0 = 0.6$ fm/$c$, is taken from earlier calculations for $\sqrt{s_{\mathrm{NN}}}=130$ GeV energy [@letter1; @letter2]. The freeze-out energy density is chosen to reproduce the slopes of pion and net proton spectra in most central collisions (see upper left panel of Fig. \[spektrit\]). The stiffer the EoS, the sooner, i.e. at higher density, the necessary flow velocity to fit the spectra has been built up. We find that effectively the EoSqp is stiffest since it requires the highest decoupling energy density $\varepsilon_{fo} = 0.14$ GeV/fm$^3$ ($\langle T_{fo}\rangle = 141$ MeV) to fit the data. Even if the ideal parton gas EoS is stiff, the mixed phase of the EoSQ makes it effectively the softest EoS here. The stiffening of the phase transition region and softening of the plasma phase in the EoST cancel each other. It is almost as soft as EoSQ and can use the same freeze-out energy density $\varepsilon_{fo} = 0.08$ GeV/fm$^3$ ($\langle T_{fo}\rangle = 130$ MeV). The purely hadronic EoS H is in between these extremes and requires $\varepsilon_{fo} = 0.10$ GeV/fm$^3$ ($\langle T_{fo}\rangle = 135$ MeV) to fit the data. As can be seen in the upper left panel of Fig. \[spektrit\], these choices of $\varepsilon_{fo}$ allow all EoSs to fit the data equally well. Results ======= $p_T$-spectra ------------- The transverse momentum spectra for pions, kaons and net protons for various centralities are shown in Fig. \[spektrit\]. Pion and net proton spectra in most central collisions were used to fix the parameters of the model, but the data are well reproduced at other centralities too. Only at the most peripheral collisions the data tend to favour flatter spectra than calculated. The kaon spectra was not taken into account when choosing the parameters and the calculated spectra is a prediction in most central collisions too. The fit to data is surprisingly good when one takes into account that the freeze-out temperature is well below the $T_{\mathrm{chem}}\approx 174$ MeV chemical freeze-out temperature where particle yields are fixed [@PBM]. The net proton spectra is well reproduced up to $p_T = 2.5$ - 3.0 GeV except in the most peripheral collisions, where the data begins to deviate form the calculation at lower $p_T$. It is worth noticing that we are able to fit the $p_T$ spectra of net protons without any initial transverse velocity field whereas the fit of protons in Ref. [@Peter-Ralf] required a non-zero initial transverse velocity. One reason for this is that due to larger errors, it is easier to fit the net-proton than proton spectra. The main cause is, however, the different EoS in the hadronic phase. In [@Peter-Ralf], the authors assumed a separate thermal and kinetic freeze-outs and only a partial chemical equilibrium in the hadronic phase whereas in this work a full chemical equilibrium is assumed. Although the relation between pressure and energy density is almost independent of these assumptions, the relation between temperature and energy density depends strongly on them [@Hirano; @Derek-eos]. Thus the relation between collective and thermal motion in a hydrodynamical model depends on the assumption of chemical equilibrium or non-equilibrium and very different initial states can be required to fit the data. Elliptic anisotropy ------------------- The second Fourier coefficient, $v_2$, of the azimuthal distribution of charged particles as function of centrality is shown as a histogram in Fig. \[v2cent\]. Note that the data measured by the STAR [@STAR-v4] and PHENIX [@Phenix-v2] collaborations have different pseudorapidity and $p_T$ cuts. After these cuts have been applied to the calculations, the results differ slightly. Therefore the comparison with the data is done in two separate panels. The agreement with data is similar to that seen in $\sqrt{s_{\mathrm{NN}}} = 130$ GeV collisions [@Derek-RQMD; @letter1; @QM02]: at most central collisions ($< 10$% of cross section, $b \lesssim 4.6$ fm) the observed anisotropy is above the hydrodynamical result. At semicentral collisions the calculations fit the data (10 - 30% of cross section, $4.6 \lesssim b \lesssim 8$ fm, depending on the EoS) and at peripheral collisions the calculated anisotropy is well above the observed. One possible explanation for larger observed than calculated anisotropy in most central collisions is fluctuations in the initial state geometry [@Miller]. The present experimental procedure cannot distinguish between the enhancing and suppressing effects of fluctuations on anisotropy in most central collisions and consequently leads to too large value of $v_2$. The sensitivity of the anisotropy to the EoS depends on centrality. In the most central and semi-peripheral collisions EoSQ leads to the lowest anisotropy and EoSH to the largest, but in most peripheral collisions the lowest anisotropy is achieved using EoSqp. The stiffest EoS does not always lead to the largest anisotropy and the softest to smallest because of the interplay of collective and thermal motion. Stiff EoS may necessitate decoupling at higher temperature when larger thermal motion dilutes the flow anisotropy. The momentum dependence of elliptic flow, $v_2(p_T)$, in minimum bias collisions is shown in Fig. \[v2piap\] for positive pions and antiprotons and in Fig. \[v2KL\] for neutral kaons and a sum of lambdas and antilambdas. For pions the behaviour is similar to the charged particle $v_2(p_T)$ at $\sqrt{s_{\mathrm{NN}}} = 130$ GeV collisions [@Heinz2; @letter1]. Regardless of the EoS the calculated anisotropy reproduces the data up to $p_T \approx 1.5$ GeV where the data begins to saturate but the hydrodynamical curve keeps increasing. Major differences between different EoSs are at the high $p_T$ region where no EoS fits the data. Closer look at low $p_T$ region ($p_T < 1$ GeV) reveals that EoSH leads to slightly larger $v_2$ than the other EoSs, but the difference is equal to the difference between the STAR and PHENIX data. The antiprotons show much stronger sensitivity to the EoS than pions. Below $p_T = 2$ GeV the results form two groups. EoSsqp and H lead to almost identical $v_2(p_T)$ which is clearly above the data whereas EoSsQ and T lead to anisotropy very close to the data. The phase transition crossover in EoST is very rapid with $\Delta T = 5$ MeV. We have tested that increase in $\Delta T$ leads to larger antiproton $v_2(p_T)$ at low $p_T$ and worse fit with the data. For example $\Delta T = 17$ MeV moves the $v_2(p_T)$ curve roughly halfway between results for EoSQ and qp. At high values of $p_T$ the order of results is changed with EoSQ leading to highest and EoSH to the lowest anisotropy. The antiproton data follows the hydrodynamical calculation to much higher values of $p_T$ than the pion data. Even the highest data point at $p_T=3.2$ GeV is fitted while using EoSqp or H. Even when EoSQ is used, we can not reproduce the antiproton $v_2(p_T)$ as well as in earlier studies [@letter2]. The main reason is that in Ref. [@letter2] freeze-out temperature was lower $T_f \approx 120$ MeV, but after constraining the freeze-out to fit the $p_T$ spectra we are forced to use higher freeze-out temperature which does not allow as good description of the $v_2$ data. The general behaviour of antiproton $v_2(p_T)$ suggests that the larger the latent heat, the smaller the $v_2(p_T)$ at low $p_T$. However, this is not the case. To test this hypothesis we used also an EoS with a first order phase transition and larger latent heat than EoSQ (2 GeV/fm$^3$ instead of 1.15 GeV/fm$^3$). The fit to antiproton anisotropy was no better than for EoSQ (similar behaviour was already seen in Ref. [@Derek-RQMD] for EoSs with latent heats 0.8 and 1.6 GeV/fm$^3$). Comparison with the strange particle data ($K_s^0$, $\Lambda + \bar{\Lambda}$) in Fig. \[v2KL\] shows similar trends. The larger the particle mass the larger the differences between EoSs at low $p_T$. The data deviates from the overall behaviour of hydrodynamical calculation at lower $p_T$ for mesons than for baryons – the kaon data deviates already around $p_T \approx 1.2$ GeV whereas hydrodynamical calculation is close to lambda data up to $p_T \approx 3.5$ GeV. On the other hand the overall fit to data is worse for strange than non-strange particles. Even EoSQ leads to calculated anisotropy which is above the data also at low $p_T$. Smaller $v_2$ at low $p_T$ can not be interpreted as a sign of strange particles freezing out earlier at higher temperature. For kaons and lambdas that would mean *larger* $v_2$ at small $p_T$. The good fit to kaon $p_T$ spectra is also against a higher kinetic freeze-out temperature for strange particles. As shown in Ref. [@alkutilat], the different parametrisations of the initial state can lead to similar pion $v_2(p_T)$ but different proton $v_2(p_T)$ in minimum bias collisions. We have checked if it would be possible to bring the antiproton $v_2(p_T)$ down to fit the data using EoSH but different initial state as speculated in Ref. [@review]. To do this we assumed that at each value of impact parameter $b$, the initial energy density was proportional to the density of binary collisions in the transverse plane (parametrisation eBC of Ref. [@alkutilat]), but the proportionality constant depended on impact parameter to reproduce the observed centrality dependence of multiplicity. Because this parametrisation led to steeper initial gradients than our usual parametrisation, we had to use freeze-out energy density $\varepsilon_f = 0.12$ GeV/fm$^3$ instead of $\varepsilon_f = 0.1$ GeV/fm$^3$ ($\langle T_f\rangle = 138$ MeV and $\langle T_f\rangle = 135$ MeV, respectively) to reproduce the $p_T$ distributions of pions and net-protons. As a result, the earlier decoupling negated the change due to the different initial shape and the final proton $v_2(p_T)$ was almost similar to that shown in Fig. \[v2piap\] and well above the data. We conclude that the anisotropies shown in Figs. \[v2piap\] and \[v2KL\] are typical for each EoS and robust against small variations in the initial parametrisation of the system. Higher harmonics ---------------- 55 mm Recently there has been interest in measuring the higher harmonics of the azimuthal distribution of particles [@STAR-compiled; @STAR-v4]. It has been proposed that these higher coefficients should be even more sensitive to the initial configuration of the system than the elliptic flow coefficient $v_2$ [@Kolbv4]. A detailed study of these coefficients would require checking how different initial configurations would affect these coefficients. Instead we calculate the fourth and sixth harmonics of distribution, $v_4$ and $v_6$, using the initial state defined above as a first attempt to see how an EoS affects higher harmonics. The fourth and sixth harmonics of the charged particle distribution in minimum bias collisions, $v_4(p_T)$ and $v_6(p_T)$ as function of transverse momenta are shown in the left panel of Fig. \[v46charged\]. The EoS has significant effect only above $p_T\approx 2$ GeV, i.e. in the region where $v_2(p_T)$ is no longer reproduced by hydrodynamics. EoSQ leads to $v_4$ peaking around $p_T\approx 3$ GeV whereas all the other EoSs lead to monotonous increase of $v_4(p_T)$ with increasing $p_T$. The data, on the other hand, increases up to $p_T\approx 3$ GeV and saturates. Except for the high $p_T$ region all the EoSs lead to $v_4$ which is smaller than the experimentally measured values. The measured values of the sixth harmonics of the distribution, $v_6$, are consistent with zero, although the errors are large enough not to exclude any of the calculations here. The calculated values of $v_6$ are also small but show a qualitative dependence on the EoS: EoSQ leads to negative $v_6$ whereas all the other EoSs lead to positive values of $v_6$. The centrality dependence of the $p_T$ averaged fourth harmonic $v_4$ is shown in the right panel of Fig. \[v46charged\]. It shows qualitatively similar behaviour to $v_4(p_T)$. The EoS has only a weak effect on $v_4$ except in peripheral collisions. In central and semicentral collisions the calculated values are below the observed ones. In Fig. \[v4particles\] the $p_T$ dependence of fourth harmonic $v_4$ in minimum bias collisions is shown for identified pions, kaons ($K_s^0$), antiprotons and lambdas. As was the case for charged hadrons, the EoS has only a weak effect on results below $p_T\approx 2$ GeV. The pion data is above the hydrodynamical calculations. The errors for other particles are large and the calculations fit the data except at the highest $p_T$ where kaon data seems to favour EoSQ and lambda data all the other EoSs. Flow on decoupling surface ========================== To understand how different EoSs lead to different anisotropies, we study the properties of the freeze-out surface in Au+Au collision with impact parameter $b=6$ fm. We try to find a set of parameters to describe the surface similar to those presented in Ref. [@STAR130] for a blast wave model. Freeze-out temperature, average transverse flow velocity and two anisotropy coefficients are shown in Table \[omituisuudet\]. To characterise the spatial anisotropy of the surface we generalise the usual spatial anisotropy $\epsilon_x$ [@Peterinpitka] for hypersurfaces: $$\epsilon_x = \frac{\int \partial\sigma_\mu s^\mu\, (y^2 - x^2)} {\int \partial\sigma_\mu s^\mu\, (y^2 + x^2)},$$ where the usual integral over ${\mathrm{d}}x\,{\mathrm{d}}y$ is replaced by an integral over space-time hypersurface and instead of energy density, entropy density is used as a weight. To characterise the azimuthal modulation of the flow field, we first calculate average flow velocity as function of flow angle, $\langle v_r(\phi_v)\rangle$, where $\phi_v = \arctan(v_y/v_x)$. We use the second Fourier coefficient of this distribution as a measure of anisotropy of the flow field: $$a_2 = \frac{\int {\mathrm{d}}\phi\, \langle v_r (\phi)\rangle\, \cos(2\phi)} {\int {\mathrm{d}}\phi\, \langle v_r (\phi)\rangle}.$$ This allows us to separate the spatial anisotropy from the flow anisotropy. EoSqp EoSQ EoSH EoST ------------------------------- ------- ------- ------- ------- $\langle T_{fo}\rangle$ (MeV) 141 130 134 130 $\langle v_r \rangle$ 0.47 0.47 0.49 0.49 $\epsilon_x$ 0.058 0.033 0.056 0.034 $a_2$ 0.027 0.027 0.025 0.026 : Freeze-out temperature, average transverse flow velocity, spatial eccentricity and flow anisotropy on the decoupling surface in Au+Au collision with impact parameter $b=6$ fm using four different EoSs.[]{data-label="omituisuudet"} The average flow velocity and anisotropy of the velocity field are surprisingly similar in all four cases. The main differences at freeze-out are freeze-out temperature and the shape of the surface. As seen in Ref. [@Lisa] where the anisotropies are studied using a parametrisation of the freeze-out surface, at this temperature and velocity range the lower temperature should lead to larger anisotropies for both pions and protons. As can be expected, in parametrisation smaller spatial anisotropy is seen to lead to smaller $v_2$ of particles. This behaviour is different from what we see here where EoSQ leads to lowest $v_2(p_T)$ at low $p_T$. Smaller spatial anisotropy can not explain this alone, since its effect should be cancelled by lower temperature. Also the differences between EoSQ and T are such that one would expect EoST to lead to lower $v_2$ for both pions and protons, but that is not the case. Clearly the average values do not characterise the flow well enough. The reason for different anisotropies must lie in the details of the flow profiles. To have a closer look at the properties of flow on the decoupling surface, we have plotted the flow velocity on decoupling surface as a function of radial coordinate in Fig. \[profiles\]. In the left panel the flow velocity is shown as function of $y$ when $x=0$ and in the right panel as function of $x$ when $y=0$. As expected from very similar spectra and differential anisotropies, the velocity distribution for EoSs qp and H is also close to each other. EoSQ, on the other hand, leads to different flow profile with slower increase of velocity with increasing radius, a distinctive “shoulder” at $r\approx 5$ fm where the velocity can even slightly decrease with increasing $r$ (in $x$-direction) and very rapid rise of flow velocity close to maximum radius of the system. EoST on the other hand is somewhere in between these two with the slow rise at low $r$ and very rapid rise at large $r$ but with much weaker structure around $r\approx 5$ fm. Even if the flow velocity distributions shown in Fig. \[profiles\] do not look too different from each other, the amount of particles emitted from fluid elements at different velocities is very different. To characterise this, the velocity curves in Fig. \[profiles\] are divided into segments so that each segment corresponds to 20% of entropy flowing through surface and thus $\sim 20$% of particles emitted. Also the entropy flow as function of flow velocity on the decoupling surface is shown in Fig. \[entropy\]. As can be seen EoSQ leads to very different distribution with much more particles being emitted at small flow velocities. Especially the “shoulder” in flow profile around $r = 5$ fm leads to a peak in entropy distribution at $v_r \approx 0.38$ whereas EoSs qp and H lead to distributions peaking at $v_r\approx 0.6$, close to maximum values of flow velocity. EoST is again a compromise between these two extremes. The entropy flow has a peak both at the “shoulder” at $v_r\approx 0.42$ and close to maximum velocity at $v_r\approx 0.68$. The largest flow velocity is also larger than for EoSs qp and H and close to the maximum for EoSQ. The flow on decoupling surface is thus weighted very differently for each EoS and the average values of flow velocity and anisotropy do not completely describe $p_T$ differential anisotropies of particles. In Ref. [@Derek-RQMD] similar velocity profiles were considered linear and corroborating the general use of linear velocity profiles in hydrodynamically inspired fits to particle spectra. As seen here the deviations from linear behaviour are important at least in non-central collisions. Thus the parameter values from fits can deviate from values obtained in full-fledged hydrodynamical calculations. Conclusions =========== In this paper we have examined how the order of deconfinement phase transition affects the anisotropy in a hydrodynamical description of relativistic nuclear collision. We used four different Equations of State – one lattice inspired EoS with a crossover transition from hadronic to partonic phase (EoSqp), one where a simple Maxwell construction between different phases creates a first order phase transition (EoSQ), a purely hadronic EoS with no phase transition at all (EoSH) and an EoS where different phases were smoothly connected with a hyperbolic tangent function (EoST). The $p_T$ distributions of various particles could be reproduced equally well using each of these EoSs when the freeze-out density was chosen accordingly. Our result is thus different from Ref. [@Derek-RQMD] where $p_T$ distributions were sensitive to the amount of latent heat of the EoS. This difference is due to different treatment of freeze-out. In Ref. [@Derek-RQMD], the hadronic stage was described using RQMD cascade model which does not have freeze-out temperature or density as a free parameter. The main sensitivity to the EoS was seen in the differential anisotropy of heavy particles at low $p_T$, i.e. antiprotons and lambdas. None of the EoSs was able to reproduce the data, but the EoS with the first order phase transition, EoSQ was closest. Surprisingly the lattice based EoSqp was as far from the proton $v_2$ data as the EoSH without any phase transition. The basic rule was that the sharper the rapid rise in entropy and energy density at phase transition and the larger the latent heat, the lower the differential anisotropy of antiprotons at low $p_T$ was. This, however, is valid only among the EoSs discussed here. The results here favour EoSQ and first order phase transition over lattice inspired EoSqp. One should not interpret this to mean that hydrodynamical description of elliptic flow requires a first order phase transition since EoST with a crossover transition lead to only marginally worse results than EoSQ. The main difference between EoSs qp and T is in the size of the increase in energy and entropy densities around the critical temperature and consequently how wide is the region where the speed of sound is small. Thus the acceptable description of elliptic flow seems to require very fast and sufficiently large increase in entropy and energy densities around $T_c$. However, these results must be taken as only preliminary. For simplicity hadron gas was assumed to maintain chemical equilibrium until kinetic freeze-out in these calculations. As mentioned in section \[init\], this assumption does not allow the reproduction of observed particle yields but only the slopes of their spectra and approximatively their anisotropies [@Heinz; @letter1]. The recent calculations where this assumption is relaxed and a separate chemical and kinetic freeze-outs included in the model [@Hirano; @Peter-Ralf], have lead to much worse description of the data [@phenixwhite]. It looks like it is very difficult to describe the data using ideal fluid hydrodynamics while the hadron gas is not in chemical equilibrium [@miklos-tetsu]. On the other hand, if the hadronic phase is described using RQMD cascade which allows chemical non-equilibrium, the data is again reproduced [@Derek-RQMD]. Thus the correct treatment of the hadronic phase in a hydrodynamical model is an open question and it is not yet possible to draw final conclusions about the details of the EoS based on the observed anisotropies. Nevertheless our results point to that a large and rapid increase in densities around critical temperature is necessary in hydrodynamical description to describe the observed anisotropies. The failure of lattice inspired EoS to do this raises the questions whether the lattice result used here is sufficiently accurate around $T_c$, whether the hadron resonance gas description of the EoS below $T_c$ is inaccurate or whether some finite size effects make the EoS relevant for heavy ion collisions differ from lattice QCD results. Acknowledgements {#acknowledgements .unnumbered} ================ I am grateful for Thorsten Renk and Roland Schneider for allowing us to use a parametrisation of their quasiparticle EoS. Fruitful discussions with P. V. Ruuskanen, R. Snellings and S. S. Räsänen are also thankfully acknowledged. This work was partially supported by the Academy of Finland under project no. 77744. [99]{} S. S. Adler [*et al.*]{} \[PHENIX Collaboration\], Phys. Rev. Lett.  [**91**]{} (2003) 182301 \[arXiv:nucl-ex/0305013\]. B. B. Back [*et al.*]{} \[PHOBOS Collaboration\], arXiv:nucl-ex/0407012. J. Adams [*et al.*]{} \[STAR Collaboration\], arXiv:nucl-ex/0409033. For reviews see, e.g., J. Ollitrault, Nucl. Phys. A [**638**]{}, 195 (1998) \[arXiv:nucl-ex/9802005\]; A. M. Poskanzer, arXiv:nucl-ex/0110013; or S. A. Voloshin, Nucl. Phys. A [**715**]{}, 379 (2003) \[arXiv:nucl-ex/0210014\]. H. Sorge, Phys. Rev. Lett.  [**78**]{} (1997) 2309 \[arXiv:nucl-th/9610026\]. J. Adams [*et al.*]{} \[STAR Collaboration\], Phys. Rev. Lett.  [**92**]{} (2004) 062301 \[arXiv:nucl-ex/0310029\]; A. M. Poskanzer \[STAR Collaboration\], J. Phys. G [**30**]{} (2004) S1225 \[arXiv:nucl-ex/0403019\]. P. F. Kolb, Phys. Rev. C [**68**]{} (2003) 031902 \[arXiv:nucl-th/0306081\]. P. Huovinen, in Quark-Gluon Plasma 3, eds. R. C. Hwa and X. N. Wang (World Scientific, Singapore, 2004) \[arXiv:nucl-th/0305064\]. P. F. Kolb and U. Heinz, in Quark-Gluon Plasma 3, eds. R. C. Hwa and X. N. Wang (World Scientific, Singapore, 2004) \[arXiv:nucl-th/0305084\]. U. W. Heinz and P. F. Kolb, Nucl. Phys. A [**702**]{} (2002) 269 \[arXiv:hep-ph/0111075\]. D. Teaney, J. Lauret and E. V. Shuryak, arXiv:nucl-th/0110037. U. W. Heinz and P. F. Kolb, arXiv:hep-ph/0204061. F. Karsch and E. Laermann, in Quark-Gluon Plasma 3, eds. R. C. Hwa and X. N. Wang (World Scientific, Singapore, 2004) \[arXiv:hep-lat/0305025\]. P. F. Kolb, P. Huovinen, U. W. Heinz and H. Heiselberg, Phys. Lett. B [**500**]{} (2001) 232 \[arXiv:hep-ph/0012137\]. P. Huovinen, P. F. Kolb, U. W. Heinz, P. V. Ruuskanen and S. A. Voloshin, Phys. Lett. B [**503**]{} (2001) 58 \[arXiv:hep-ph/0101136\]. P. F. Kolb, U. W. Heinz, P. Huovinen, K. J. Eskola and K. Tuominen, Nucl. Phys. A [**696**]{} (2001) 197 \[arXiv:hep-ph/0103234\]. T. Hirano, Phys. Rev. C [**65**]{} (2002) 011901 \[arXiv:nucl-th/0108004\]. T. Hirano and K. Tsuda, Phys. Rev. C [**66**]{} (2002) 054905 \[arXiv:nucl-th/0205043\]; P. F. Kolb and R. Rapp, Phys. Rev. C [**67**]{} (2003) 044903 \[arXiv:hep-ph/0210222\]. J. P. Blaizot and J. Y. Ollitrault, Phys. Rev. D [**36**]{} (1987) 916. D. H. Rischke and M. Gyulassy, Nucl. Phys. A [**597**]{} (1996) 701 \[arXiv:nucl-th/9509040\]. R. A. Schneider and W. Weise, Phys. Rev. C [**64**]{} (2001) 055201 \[arXiv:hep-ph/0105242\]. P. Braun-Munzinger, D. Magestro, K. Redlich and J. Stachel, Phys. Lett. B [**518**]{} (2001) 41 \[arXiv:hep-ph/0105229\]. G. Boyd, J. Engels, F. Karsch, E. Laermann, C. Legeland, M. Lutgemeier and B. Petersson, Nucl. Phys. B [**469**]{} (1996) 419 \[arXiv:hep-lat/9602007\]. F. Karsch, E. Laermann and A. Peikert, Phys. Lett. B [**478**]{} (2000) 447 \[arXiv:hep-lat/0002003\]; F. Karsch, Nucl. Phys. A [**698**]{} (2002) 199 \[arXiv:hep-ph/0103314\]. R. Venugopalan and M. Prakash, Nucl. Phys. A [**546**]{} (1992) 718. F. Karsch, K. Redlich and A. Tawfik, Eur. Phys. J. C [**29**]{} (2003) 549 \[arXiv:hep-ph/0303108\]. J. Sollfrank, P. Huovinen, M. Kataja, P. V. Ruuskanen, M. Prakash and R. Venugopalan, Phys. Rev. C [**55**]{} (1997) 392 \[arXiv:nucl-th/9607029\]. T. Renk, R. A. Schneider and W. Weise, Phys. Rev. C [**66**]{} (2002) 014902 \[arXiv:hep-ph/0201048\]. P. F. Kolb, J. Sollfrank and U. W. Heinz, Phys. Rev. C [**62**]{} (2000) 054909 \[arXiv:hep-ph/0006129\]. S. S. Adler [*et al.*]{} \[PHENIX Collaboration\], Phys. Rev. C [**69**]{} (2004) 034909 \[arXiv:nucl-ex/0307022\]. D. Teaney, arXiv:nucl-th/0204023. P. Huovinen, Nucl. Phys. A [**715**]{} (2003) 299 \[arXiv:nucl-th/0210024\]. M. Miller and R. Snellings, arXiv:nucl-ex/0312008. C. Adler [*et al.*]{} \[STAR Collaboration\], Phys. Rev. Lett.  [**87**]{} (2001) 182301 \[arXiv:nucl-ex/0107003\]. F. Retiere and M. A. Lisa, Phys. Rev. C [**70**]{} (2004) 044907 \[arXiv:nucl-th/0312024\]. K. Adcox [*et al.*]{} \[PHENIX Collaboration\], Nucl. Phys. A [**757**]{} (2005) 184 \[arXiv:nucl-ex/0410003\]. T. Hirano and M. Gyulassy, arXiv:nucl-th/0506049.
{ "pile_set_name": "ArXiv" }
--- abstract: 'Every year hundreds of people die at sea because of vessel and airplane accidents. A key challenge in reducing the number of these fatalities is to make Search and Rescue (SAR) algorithms more efficient. Here we address this challenge by uncovering hidden TRansient Attracting Profiles (*TRAPs*) in ocean-surface velocity data. Computable from a single velocity-field snapshot, TRAPs act as short-term attractors for all floating objects. In three different ocean field experiments, we show that TRAPs computed from measured as well as modelled velocities attract deployed drifters and manikins emulating people fallen in the water. TRAPs, which remain hidden to prior flow diagnostics, thus provide critical information for hazard responses, such as SAR and oil spill containment, and hence have the potential to save lives and limit environmental disasters.' author: - 'Mattia Serra[^1]' - Pratik Sathe - Irina Rypina - Anthony Kirincich - 'Shane D. Ross' - Pierre Lermusiaux - Arthur Allen - Thomas Peacock - 'George Haller[^2]' bibliography: - '/Users/serram/Dropbox/ScientificLiterature/BibFiles/ReferenceList3.bib' date: - - title: Search and rescue at sea aided by hidden flow structures --- Introduction ============ In 2016, the United Nation Migration Agency recorded over 5000 deaths among people trying to reach Europe by crossing the Mediterranean Sea [@DeathAtSeaMediterr2017; @DeathAtSeaStatsMSF2017]. This calls for an enhancement of the efficiency of SAR at sea[@UNHCR_StrenghtenSARMediterr2018], which requires improved modeling of drifting objects, as well as optimized search assets allocation (see[@Breivik2013; @stone2013search] for reviews). Flow models used in SAR operations combine sea dynamics, weather prediction and in situ observations, such as self-locating datum marker buoys[@allen1996performance] deployed from air, which enhance model precision near the last seen location. Even with the advent of high-resolution ocean models and improved weather prediction, however, SAR planning . Current SAR procedures[@Kratzke2010] approach uncertainties through Bayesian techniques, turning the modeling exercise into an ensemble integration over all unknown parameters and incorporating unsuccessful searches into locating the next target. This strategy produces probability-distribution maps for the lost object’s location, which, based on a list of assigned search assets, returns search plans, . The vast uncertain parameter space together with the continuous motion of floating objects driven by unsteady flows, however, leads to error accumulation, “making SAR planning as much art as science, where rescuers still often rely as much on their hunches as on the output of sophisticated prediction tools”[@Breivik2013]. Furthermore, the convergence of updated probability computations based on a selected prior and unsuccessful searches is usually a slow process, while timing is everything when lives are on the line. ![image](Fig_F_png/Fig1){height=".8\columnwidth"} In a SAR scenario, one would ideally have a simply interpretable tool based on key features of the ocean surface dynamics. Such a tool should narrow down the search area by promptly providing the most attracting regions in the flow toward which objects fallen in the water at uncertain locations likely converge. This raises the question: How can one rigorously assess short-term variabilities of material transport in fast-changing flows characterized by high uncertainties? Here, we address this question using the recently developed concept of Objective Eulerian Coherent Structures (OECSs)[@SerraHaller2015] from dynamical systems theory. In our context, attracting OECSs uncover hidden *TRansient Attracting Profiles* (*TRAPs*), revealing the currently strongest regions of accumulation for objects floating on the sea surface. TRAPs are quickly computable as smooth curves from a single snapshot of available modelled or remotely sensed velocity fields, providing highly specific information for optimal search-asset allocation (Fig. \[fig:DeathByRescue\]). The inset in Fig. \[fig:DeathByRescue\] shows a migrant boat that capsized on 12 April 2015 in the Mediterranean Sea, along with a schematic TRAP attracting people in the water (PIW). We confirm the predictive power of TRAPs in three field experiments emulating SAR situations south of Martha’s Vineyard in Massachusetts USA. In the first experiment, we compute TRAPs from a submesoscale ocean surface velocity field reconstructed from remotely sensed High Frequency Radar (HFR) data, and show their decisive influence on surface drifters emulating people that have fallen in water at uncertain locations. In actual SAR operations, however, HFR velocity data is generally not available in real time. We address this challenge in our second and third experiments by computing TRAPs from an ocean model velocity field that assimilates in situ experimental information. We then verify the TRAPs’ role in attracting and aligning drifters and manikins, simulating PIW, released in their vicinity through targeted deployments. Our analysis reveals a remarkable robustness under uncertainty for TRAPs: even without accounting for –typically uncertain in SAR scenarios–, the TRAPs invariably attract floating objects in water over two-to-three hours. Methods ======= Short-term variability in flow features (or coherent structures) is substantial in unsteady flows. These structures, such as fronts, jets and vortices, continue to receive significant attention in fluid mechanics due to their decisive role in organizing overall transport of material in fluids. Such transport is a fundamentally Lagrangian phenomenon, i.e., best studied by keeping track of the longer-term redistribution of individual tracers released in the flow. In that setting, Lagrangian coherent structures (LCSs) have been efficient predictors of tracer behavior in approximately two-dimensional geophysical flows, such as surface currents in the ocean [@LCSHallerAnnRev2015]. Larger-scale models and measurements of environmental flows, however, generally produce Eulerian data, i.e., instantaneous information about the time-varying velocity field governing the motion of tracers. These velocity fields can be integrated to obtain tracer trajectories, but the result of this integration will generally be sensitive to a number of factors. One such set of factors is the exact release time, release location and length of the observation period. Another major sensitivity factor is errors and uncertainties in the velocity field, which either arise from unavoidable simplifications and approximations in modeling, or from inaccuracies in remote sensing. A third source of sensitivity is the necessarily approximate nature of trajectories generated by numerical integration, due to finite spatial and temporal resolution of the velocity data, as well as to approximations in the numerical integration process. All these factors are significant in predictions for SAR purposes: in fast-changing coastal waters, uncertainties both in the available velocities and in the release location and time are high. This has prompted the use of multiple models, stochastic simulations and probabilistic predictions, all of which require substantial time to be done accurately, even though time runs out quickly in these situations. An alternative to these Lagrangian approaches is to find the short-term limits of LCSs purely from Eulerian observations, thereby avoiding all the pitfalls of trajectory integration. These limiting LCSs predict pathways and barriers to short-term material transport until the next batch of updated velocity information becomes available. While simple at first sight, this approach comes with its own challenges, given that most classic instantaneous Eulerian diagnostics (streamlines, velocity magnitude, velocity gradient, energy, vorticity, helicity, etc) are not objective[@TruesdellNoll2004], i.e., depend on the observer. As such, they cannot possibly be foolproof indicators of material transport, which is a fundamentally frame-independent concept. Indeed, different observers relying on data collected from the coast, from an airplane, from a ship or from a satellite should not come to different conclusions regarding the likely location of materials or people in the water. Yet classic Eulerian quantities would in fact give such different answers (see e.g. Fig. 3a in[@LCSHallerAnnRev2015] and Fig. 1 in [@SerraHaller2015]). In a SAR situation, this ambiguity is a serious limitation that represents high risk. These considerations led to the development of OECSs[@SerraHaller2015], which are objective (observer-independent) short-term limits of LCSs. Most relevant to our current setting are hyperbolic OECSs in two-dimensional flows, which are the strongest short-term attractors and repellers of material fluid elements. As such, OECSs are extensions of the notions of unstable (and stable) manifolds of a saddle point in a steady flow, which attract (and repel, respectively) fluid elements and hence ultimately serve as the theoretical centerpieces of deforming tracer patterns. In unsteady flows and over short times, however, such saddle-type, instantaneous stagnation points lose their connection with material transport[@SerraHaller2015]. Instead, *objective saddle points* – the cores of hyperbolic OECSs – emerge, with associated attracting and repelling OECSs (Fig. \[fig:Saddle\]), which, in turn, continuously evolve over time. In our present context, we will refer to attracting OECSs and objective saddle points as *TRAPs* and *TRAP cores*. Unlike stagnation points in steady flows, OECSs cannot be located by inspection of a (frame-dependent) streamline configuration. Instead, consider a planar velocity field $\mathbf{v}(\mathbf{x},t)$, denoting by $\mathbf{e}_2(\mathbf{x},t)$ the dominant (positive) eigenvector and by $s_1(\mathbf{x},t)$ the negative eigenvalue of the rate-of-strain tensor $\mathbf{S}(\mathbf{x},t)=\tfrac{1}{2}(\mathbf{\nabla v}(\mathbf{x},t) +[\mathbf{\nabla v}(\mathbf{x},t)]^*)$, TRAPs are short segments of curves tangent to $\mathbf{e}_2$ that emanate from local minima of $s_1$[@SerraHaller2015] (see Supplementary Information for details). As an illustration, Fig. \[fig:TRAPS\_min\_s1\] shows TRAPs in an unsteady ocean velocity data set derived from AVISO satellite altimetry (see [@SerraHaller2015] for a detailed OECSs analysis of this flow). Thus, the $s_1$ scalar field along with the TRAPs provides a skeleton of currently active attracting regions in the flow along with their relative strengths. This in turn gives specific and actionable input for SAR asset allocation, such as high-priority flight paths for discovering people in the water (Fig. \[fig:DeathByRescue\]). Remarkably, such pathways remain generally hidden in streamline plots, and can even be perpendicular to streamlines as illustrated in Fig. \[fig:TRAPS\_min\_s1\]. Figure \[fig:TRAPS\] shows that TRAPs evolve over time and attract floating objects whose uncertain initial positions are represented by an array of green dots. As Eulerian objects, TRAPs are simply computable from a single snapshot of the velocity field $\mathbf{v}(\mathbf{x},t)$. Moreover, velocity fields used in SAR are generally obtained from models that assimilate environmental data in the proximity to the last known position of a missing person[@Kratzke2010; @Breivik2013]. This represents a further challenge to Lagrangian prediction methods, as much of their trajectory forecasts tend to leave the domain of reliable velocities and hence have questionable accuracy. In the Supplementary Information, we illustrate this effect, showing that Lagrangian methods provide only partial coverage when velocities are available over a finite-size domain. A TRAP-based analysis is, therefore, not only faster but provides complete coverage by exploiting all available velocity data. Finally, owing to the structural stability of its construction[@SerraHaller2015], TRAPs necessarily persist over short times and are robust to perturbations of the underlying velocity field. In the Supplementary Information we show that the sensitivity of TRAPs to uncertainties is typically lower compared to those of trajectory-based methods. This makes TRAPs a trustworthy now-casting tool for material transport, one that is resilient under uncertainties in initial conditions and other unknown factors, such as the inertia of a drifting object or windage effects. Results ======= Here we show how TRAPs accurately predict short-term attracting regions to which objects fallen in water at uncertain nearby locations converge in ocean field experiments carried out south of Martha’s Vineyard. Figure \[fig:HFRandDriftOverview\] shows the location of the experiments and the tools we used. In our first experiment, we compute TRAPs from ocean-surface sub-mesoscale velocity derived from High-Frequency-Radar (HFR) measurements available over a uniform 800m$\times$ 800m grid spanning $[-70.7979^\circ, -70.4354^\circ]$ longitude and $[41.0864^\circ, 41.3386^\circ]$ latitude, and in time steps of $30$ minutes. The velocity field is reconstructed from HFR measurements as described in [@Kirincich2012], and is available on a uniform grid within the hatched polygon in Fig. \[fig:HFRDomain\] (Supplementary Information). To mimic objects fallen in the water, we use 68 Coastal Ocean Dynamics Experiment (CODE) drifters (Supplementary Information and Fig. \[fig:HFRandDriftOverview\]) whose GPS-tracked locations (green dots) are recorded once every 5min. Drifters of the same design are routinely used by the U.S. Coast Guard in SAR operations. The starting time of our analysis is the 4th of August 2014 at 17:00 EDT when drifters are located close to the Muskeget channel (Fig. \[fig:HFRDomain\]). Figure \[fig:HFR2014\_17:00\] shows a zoomed version of the black square inset in Fig. \[fig:HFRDomain\], along with drifter positions and the instantaneous streamlines of the HFR velocity. We then compute TRAPs every 30min with the updated velocity field. As expected, we find that the emergence of strong TRAPs at 19:00 (Fig. \[fig:HFR2014\_19:00\]) promptly organize the drifters into one-dimensional structures along TRAPs within two hours (Figs. \[fig:HFR2014\_19:00\]-\[fig:HFR2014\_20:00\]). These results show that TRAPs may be completely hidden in instantaneous streamline and divergence plots, yet predict the short-term fate of passive tracers, as well as inertial objects influenced by windage, such as drifters. Although incorporating inertial, windage and leeway effects could, in principle, provide a better prediction, in a SAR operation the inertia of the target objects is generally unknown[@maio2016evaluation] and wind information is unavailable. Although using HFR velocity would significantly enhance the success of SAR operations[@Bellomo2015], SAR planning is generally based on model velocity data. To account for this, we conducted two more experiments to identify TRAPs from the ocean surface velocity derived from the MIT Multidisciplinary Simulation, Estimation, and Assimilation Systems (MIT-MSEAS)[@Haley2010] (Supplementary Information) which assimilates local measurements, similarly to the models used in actual SAR. \ For the experiment performed on the 17th August 2017, we compute TRAPs from the 24h forecast model velocity provided on the 16th August at 7pm. We focus on a region south-east of Martha’s Vineyard and identify TRAPs from 11am on 17th August 2017 (Fig. \[ReleasePlan2017\] Left). The strongest TRAPs are located along a trench of the $s_1(\mathbf{x},t)$ field demarcating a one-dimensional structure containing several TRAPs with strong attraction rates. We note the presence of two parallel trenches from the model. Assuming that the real trench is somewhere in between these two because of modeling uncertainties, we released four drifters north of the lower trench (magenta squares). The right panels in Fig. \[ReleasePlan2017\] show later positions of fluid particles obtained by integrating the model velocity field from the target drifter release location, along with the corresponding TRAPs. The figure confirms their attracting property with respect to model data before the float deployment. Based on the release locations in Fig. \[ReleasePlan2017\], Fig. \[fig:2017HFR\] (left) shows the deployed GPS-tracked drifters position (green dots) at 11:10am within the area of interest bounded by the black rectangle in Fig. \[ReleasePlan2017\], along with the streamlines computed from the HFR velocity at the same time. The right panels show later drifter positions along with the TRAPs and the streamlines computed from the HFR velocity field. Although our deployment strategy was purely based on model velocities, the comparison with actual drifter trajectories and TRAPs computed from HFR velocity shows that the model provided reliable estimates of the actual TRAPs. While these TRAPs remained hidden in streamline plots, they nevertheless attracted drifters within three hours. In our last experiment, to mimic an even more realistic SAR scenario, we considered a larger set of initially spread-out floating objects consisting of 8 CODE drifters and 4 OSCAR Water Rescue Training manikins manufactured by Emerald Marine Products (Supplementary Information, and Fig. \[fig:DrifterManikin\]). Using a strategy similar to the 2017 experiment, we designed a deployment for the 9th of August 2018, based on the center forecast model velocity field provided on the 8th of August at 8pm. Figure \[Release2018\] shows the target model-based TRAPs at 10:15am on the 9th of August 2018, along with all released drifter (green dots) and manikin (cyan triangles) positions. We show only the strongest targeted TRAPs ranked by $s_1$. Dashed curves represent the GPS-tracked trajectories of the deployed objects from their release until 10:15am. In this experiment, we used two WHOI vessels for deployment: one for the release of drifters and manikins at the locations demarcated by A,B,C,D in Fig. \[Release2018\], and a second vessel for the remaining drifters. Figures (\[2018Mit\_11:15\]-\[2018Mit\_12:30\]) show the later positions of drifters and manikins along with their trajectories and the recomputed model-based TRAPs[^3]. Similar to the previous experiments, both drifters and manikins show a striking alignment with the strongest nearby TRAPs computed from the fluid model velocity within two hours. A closer inspection of the deployed drifter and manikin trajectories shows that these two different objects may follow different paths even after short times (less than 2h). This is clearly the case for objects released from locations A,D,C shown in Fig. \[fig:2018Experiment\]. In the inset of Fig. \[2018Mit\_11:15\], we show a zoomed version of the drifter and manikin trajectories deployed in A, together with the trajectory of a fluid particle (magenta square) obtained by integrating the model velocity from A. Even though fluid particles, drifters and manikins all follow different trajectories due to inertia, windage and other effects, they invariably converge to the same TRAP, which provides a highly robust attracting skeleton of the underlying flow. In the Supplementary Information, we compare TRAP predictions with trajectory-based ones typically used in SAR. We use nine ensemble velocity field forecasts arising from parametric uncertainty sources (Supplementary Information), and compute the corresponding trajectories using the experimental deployment locations as initial conditions. We find that even though drifter, manikins and ensemble trajectories all differ from each other, they all converge to nearby TRAPs computed from the center-forecast velocity. Using simple mathematical arguments, we also show that TRAPs are intrinsically robust under uncertainties over short times, as opposed to trajectory-based methods, whose sensitivity to uncertainties grow with the largest Lyapunov exponent of the underlying velocity field. Admittedly, TRAPs loose their predictive power over longer time scales because of their instantaneous nature. Shorter time scales, however, are precisely the relevant ones for SAR and hazard response scenarios. Conclusions =========== We have predicted and experimentally verified the existence of TRansient Attracting Profiles (TRAPs), which govern short-term trajectory behavior in chaotic ocean currents characterized by high uncertainties. We expect TRAPs to provide critical information in emergency response situations, such as SAR and oil spill containment, in which operational decisions need to be made quickly about optimal resource allocation. Existing SAR techniques handle uncertain parameters in models of floating objects by averaging several Monte Carlo Simulations and providing probability maps for the objects’ location. These maps, however, are not readily interpretable for practical use and can converge slowly due to the underlying chaotic processes. TRAPs and their attraction rates, in contrast, are easily interpretable and highly localized curves which can be computed and updated instantaneously from snapshots of the ocean surface velocity. This eliminates the need for costly trajectory calculations and yields fast input for search-asset allocation. We have emulated different SAR scenarios in three ocean field experiments carried out south of Martha’s Vineyard. We computed TRAPS both from HFR submesoscale ocean surface velocity and from model velocities similar to those available in SAR operations. Our results indicate that TRAPs have significant predictive power in assessing the most likely current locations of objects and people fallen in water at uncertain locations. We have specifically found that TRAPs invariably attract nearby floating objects within two-to-three hours, even though they remain hidden to instantaneous streamlines and divergence fields, which also rely on the same Eulerian velocity input. . We therefore envision that sea TRAPs will enhance existing SAR techniques, providing critical information to save lives and limit the fall-out from environmental disasters during hazard responses. Acknowledgements {#acknowledgements .unnumbered} ================ We are grateful to Margaux Filippi, Michael Allshouse, Javier González-Rocha, Peter Nolan, Siavash Ameli, Patrick Haley Jr. and the MSEAS team for their contribution in the field experiments. We also acknowledge the NSF Hazard funding grant no. 1520825. S.D.R acknowledges support from NSF grant no. 1821145. M.S. would like to acknowledge support from the Schmidt Science Fellowship [(https://schmidtsciencefellows.org/)](https://schmidtsciencefellows.org/). G.H. acknowledges support from the Turbulent Superstructures Program of the German National Science Foundation (DFG). Author contributions {#author-contributions .unnumbered} ==================== M.S. and G.H. designed research; M.S. and P.S. performed research; I.R. provided the drifter data; A.K. provided the HFR velocity data, S.R. provided the manikin data, P.L. provided the model velocity data, A.A. provided knowledge and expertise in SAR; T.P. and I.R. led the field experiments. All authors contributed to the field experiments. M.S., G.H., T.P. and P.S. wrote the paper. Competing interests {#competing-interests .unnumbered} =================== The authors declare no competing interests. Materials and Correspondence {#materials-and-correspondence .unnumbered} ============================ All data and codes are available upon request from the corresponding authors. Supplementary Information {#supplementary-information .unnumbered} ========================= TRAPs in two-dimensional flows {#App:CompueSEA TRAPS} ============================== **Input:** A 2-dimensional velocity field $\mathbf{v}(\mathbf{x},t)$ 1. Compute the Jacobian of the velocity field $\mathbf{\nabla} \mathbf{v}$ by numerically differentiating $\mathbf{v}$ with respect to $\mathbf{x}$, and the rate-of-strain tensor $\mathbf{S}(\mathbf{x},t)=\frac{1}{2}\left(\mathbf{\nabla} \mathbf{v}(\mathbf{x},t)+\left[\mathbf{\nabla} \mathbf{v}(\mathbf{x},t)\right]^{*}\right)$ at the current time $t$ on a grid over the $\mathbf{x} = (x_{1},x_{2}$) coordinates, where \* denotes matrix transposition. 2. Compute the smallest eigenvalue field $s_{1}(\mathbf{x},t)\leq s_{2}(\mathbf{x},t)$ and the unit eigenvector field $\mathbf{e_{2}}(\mathbf{x},t)$ of $\mathbf{S}(\mathbf{x},t)$ associated to $s_{2}(\mathbf{x},t)$. 3. Compute the set $\mathcal{S}_{m}(t)$ of negative local minima of $s_{1}(\mathbf{x},t)$. 4. Compute TRAPs as solutions of the ODE $$\begin{cases} \mathbf{r}^{\prime}(\tau)=\mathrm{sign}\left\langle \mathbf{e_{2}}(\mathbf{r}(\tau)),\mathbf{r}^{\prime}(\tau-\Delta)\right\rangle \mathbf{e_{2}}(\mathbf{r}(\tau))\\ \mathbf{r}(0)\in\mathcal{S}_{m}, \end{cases}$$ where $\tau$ denotes the arclength parameter, $^\prime$ differentiation with respect to $\tau$, and $\Delta$ the arclength increment between two nearby points on the TRAP. Stop integration when $s_{1}(\mathbf{r}(\tau))>0.3s_{1}(\mathbf{r}(0))$ or $s_{1}(\mathbf{r}(s))\geq 0$. **Output:** TRAPs at time t along with their normal attraction rate field $s_{1}(\mathbf{x},t)$. The sign term in step 4 guarantees the local smoothness of the direction field $\mathbf{e_{2}}$, and the termination conditions ensure that the attraction rate of subsets of TRAPs is at least $30\%$ of the core attraction rate, hence exerting a distinguished attraction compared to nearby structures. TRAPs and uncertainties {#App:Uncertanty} ======================= The partial differential equations generating the fluid velocities used in SAR account for a broad set of uncertainties [@lermusiaux_JCP2006; @lermusiaux_et_al_O2006b]. These affect the initial conditions of all state variables including velocity, as well as external forcing and boundary conditions such as tidal forcing, atmospheric forcing fluxes, lateral boundary conditions, etc. (see SI, Model Velocity for details). With a set of ensemble velocities at hand, typical Lagrangian methods used in SAR compute their corresponding trajectories with the last seen location and time used as initial conditions. Here we consider a total of 9 ensemble forecast velocities modelling part of the above uncertainties during the 2018 field experiment shown in Fig. \[fig:2018Experiment\], and compute trajectories of the different ensemble velocities from the drifter and manikin release locations (Fig. \[fig:Uncertanties2018Exp\]). Figure \[fig:UncertZoom\] shows the drifter and manikin trajectories, initially released at point B, along with fluid particle trajectoties of the ensemble velocities (gray) and the model-based TRAP computed from the center-forecast velocity. Even within two hours from the release, ensemble trajectories already show visible differences within each other and with the actual drifter and manikin trajectories, yet all converge towards nearby TRAPs. Figure \[fig:UncertBig\] shows the same as Fig. \[fig:UncertZoom\] for all the release location of our 2018 experiment. To gain insights about the predictive power of TRAPs and Lagrangian methods under uncertainties, one can encode the above velocity field uncertainties in the stochastic ordinary differential equation for particle motions $d\textbf{x}(t) = \textbf{v}(\textbf{x}(t),t)dt + \textbf{R}(\textbf{x}(t),t)d\textbf{W}(t)$ where $\textbf{x}(t)$ is the random position vector, $\textbf{v}(\textbf{x}(t),t)$ is the deterministic drift velocity and $\textbf{W}(t)$ is a two-dimensional Weiner process with diffusion matrix $\textbf{R}(\textbf{x}(t),t)$. Then, the uncertain rate-of-strain tensor is $\textbf{S}_u(\textbf{x},t) = \textbf{S}(\textbf{x},t)+\textbf{S}_R(\textbf{x},t)d\textbf{W}(t)$, where $\textbf{S}_R(\textbf{x},t) = \text{sym}(\partial_x\textbf{R}(\textbf{x}(t),t) + \partial_y\textbf{R}(\textbf{x}(t),t))$. In the simplest case of spatially homogeneous uncertainties $\textbf{S}_R(\textbf{x},t) = \mathbf{0}$, hence TRAP predictions remain unaffected while Lagrangian (trajectory based) predictions will have inherent errors that grow both with the largest Lyapunov exponent and $\sqrt{t}$. For TRAPs to be significantly affected by uncertainties, their spatial inhomogeneities should be comparable to $\textbf{S}(\textbf{x},t)$, which means the model is highly inaccurate. These simple considerations suggest that TRAPs are intrinsically robust predictors under uncertainties over short times. TRAPs and velocity field divergence {#App:Divergence} =================================== By Liouville’s theorem[@ArnoldODE1973], the infinitesimal phase-space volume of a dynamical system shrinks along a trajectory as long as the trajectory is contained in a domain of negative divergence. Guckenheimer and Holmes[@GuckenheimerHolmes1983] conclude that if a steady vector field points everywhere inwards along the boundary of a compact region of negative divergence, then the region contains a nonempty attractor. This criterion, however, will never hold in an unsteady flow on its extended phase space of positions and time. Consequently, there is no applicable, classical dynamical systems technique to find attractors based on the velocity-field divergence in a general, non-autonomous system. This gap has been filled by the variational theory LCS[@LCSHallerAnnRev2015] for finite-time flows, and OECS[@SerraHaller2015], as its instantaneous limit. Indeed, accumulation of particles along lines for longer time intervals is well known to happen along attracting LCS in incompressible velocity fields[@Beronetal08b; @Olascoaga2013]. Our results show that even regions of positive divergence can contain curves that collect drifters/manikins. The attraction of TRAPs arises from the combination of isotropic and anisotropic deformations (Figure \[fig:DivergeneAndSeaTraps\](a)). The isotropic component is due to the instantaneous divergence of the velocity field ($\mathbf{\nabla \cdot v}$), while the anisotropic one to shear, both of which are encoded in $\mathbf{S}$, and thus in the definition of TRAPs. ![(a) At leading order, the short-term deformation of a fluid patch arises from an isotropic contribution quantified by the divergence of the velocity field ($\mathbf{\nabla \cdot v}$), and an isotropic contribution due to shear. Because of this combination, short-term attraction and clustering can occur in regions of positive divergence (b) or zero divergence (c), invariably captured by TRAPs (red). (d) Tracers escape a region of negative divergence.[]{data-label="fig:DivergeneAndSeaTraps"}](Fig_F_png/Fig8){height=".24\columnwidth"} Figure \[fig:DivergeneAndSeaTraps\] (b-c) illustrate short-term clustering regions, correctly predicted by TRAPs, where the local divergence is positive or zero. Panel (b) is similar to Fig. 4c, panel (c) to Fig. 2b or to the findings in Ref. [@SerraHaller2015]. Figure \[fig:DivergeneAndSeaTraps\] (d) shows that tracers can escape a region of negative divergence. Datasets {#App:Dataset} ======== High Frequency Radar Velocity Field {#App:HFR Velocity} ----------------------------------- The WHOI high frequency radar system (HFR), as operated during the 2014-2017 experiments, consisted of 3 land-based sites spaced at $~$10km intervals along the south side of Martha’s Vineyard, MA. These 25MHz systems were run using a combination of 350kHz transmit bandwidth and low transmit power (10W max) which allowed all systems to achieve resolutions of 429m and ranges of 30km (see [@kirincich2016remote] for details). Doppler spectra received from each system were processed using advanced methods[@Kirincich2012] into radial velocity estimates every 15min based on a 24min averaging window. Radial velocity estimates were quality controlled before inclusion into the vector velocity estimates using standard time-series QC techniques. These data were combined into vector velocities on a uniform 800m resolution grid, given in latitude and longitude coordinates, using a unique weighted least squares technique that employed non-velocity based signal quality metrics to weight the data to increase the accuracy of the final product. Two successive estimates of the 15min radials are used to estimate the vector (east and north) velocities on a 1/2h time interval centered on the hour. The spatial extent of the vector velocities was limited by theoretical Geometrical Dilution of Precision (GDOP) values less than 1.75. An error estimate for the east, north, and the norm of the vector velocity components is given. This estimate uses the radial velocity error estimates (the weighted standard deviation of the individual HF radar radial returns found within each 5 degree azimuthal bin average) in a standard (numerical recipes) vector error calculation. Because of occasional measurement deficiency, there are grid points at which the velocity is not available within the region of interest (blue dots inside the dashed curve in Fig. \[fig:velo\_avail\_init\]). To overcome this limitation, we devise a simple interpolation scheme by which we can obtain velocity everywhere within a well-defined boundary. Specifically, we first compute the boundary of this region (the dashed curve in Fig. \[fig:interpolation\_scheme\]), using the Delaunay triangulation function in MATLAB. For each time instance at which the velocity field data is available, we obtain an estimate of the velocity at the blue points inside the boundary using a linear scattered interpolation scheme (see Fig. \[fig:velo\_avail\_fin\]). Once the velocity field is available within the black boundary, we convert it from its original units of $m s^{-1}$ to degrees per day. Finally, we note that HFR velocities are computed by averaging the raw radial velocity estimates with a 800m window radius at each grid point, spaced 800m apart from each other[@kirincich2016remote]. To yield an accurate computation of TRAPs from HFR velocities consistent with the way the data are processed, we smooth $\mathbf{\nabla} \mathbf{v}(\mathbf{x},t)$ with a spatial average filter whose width corresponds to two grid sizes (1600m), as in Ref.[@Kirincich2016]. Finite-size domain {#App:finite size} ------------------ When the velocity field is available over a finite-sized domain, as the case of HFR velocity, Lagrangian Coherent Structures (LCSs) methods[@LCSHallerAnnRev2015; @hadjighasem2017critical] invariably provide incomplete coverage because of particles leaving the domain. Even if the velocity field is derived from models, they typically assimilate in situ measurements[@allen1996performance] to enhance model predictions in a specific region of interest, making the resulting velocity field more accurate in a finite-size domain. Assuming that such a region is the domain over which HFR velocity is available, Fig. \[fig:LagrangianCoverage\] shows the coverage reduction of LCSs methods for two different integration times. We compute fluid trajectories by integrating the HFR velocity field starting on the 4th of August 2014 at 8pm, with a dense set of initial conditions covering the entire domain bounded by the dashed curve. After 2.4h (12h) any Lagrangian method can provide information only on 82% (55%) of the region where the velocity field is available. Eulerian methods, instead, provide full coverage of the domain because they do not rely on particle trajectories. Model velocity {#App:Model Velocity} -------------- We use the MIT Multidisciplinary Simulation, Estimation, and Assimilation Systems (MIT-MSEAS)[@Haley2010; @haley_et_al_OM2015] Primitive-Equation (PE) ocean modeling system to compute ocean surface velocity forecasts in the Nantucket and Martha’s Vineyard coastal region during August 2017 and 2018. The modeling system is set-up in an implicit 2-way nesting configuration, and provided forecasts of the ocean state variable fields (three-dimensional velocity, temperature, salinity, and sea surface height) every hour with a spatial resolution of 200m in the Martha’s Vineyard domain and of 600m in the larger Shelf domain. The ocean forecasts are initialized using historical and synoptic ocean CTD data from the National Marine Fisheries Service (NMFS) and the Martha’s Vineyard Coastal Observatory (MVCO), SST images from the Johns Hopkins University’s Applied Physics Lab (JHU APL), and other data from available sources. These ocean simulations are forced by atmospheric flux fields forecast by the National Centers for Environmental Prediction (NCEP) and tidal forcing from TPXO8, but adapted to the high-resolution bathymetry and coastlines[@logutov_lermusiaux_OM2008]. The deterministic 2-way nesting ocean forecast initialized from the estimated ocean state conditions at a particular time is referred to as the central forecast. The ensemble forecasts were initialized using Error Subspace Statistical Estimation procedures [@lermusiaux_JAOT2002]. The forecasts within the ensemble were commonly initialized from perturbed initial conditions of all state variables (temperature, salinity, velocity, sea surface height) and forced by perturbed tidal forcing, atmospheric forcing fluxes and lateral boundary conditions. These initial, forcing and boundary perturbations are created so as to represent the expected uncertainties in each of these quantities, respecting ocean physics and in accord with the few observed data misfits. Each ensemble members were 2-way nested in two domains and required respecting domain embedding conditions. Finally, parameter uncertainties (bottom drag, mixing coefficients, etc.) were also modeled by perturbing the values of parameters for each ensemble forecast. Ensemble forecasts were issued twice daily during the 2 weeks of the 2017 and 2018 experiment (see <http://mseas.mit.edu/Sea_exercises/NSF_ALPHA/2018/>), with the number of ensemble forecasts issued varying between 7 and 100 depending on the number of computing units available and on computational power. The nine 2018 MSEAS ensemble forecasts utilized for the present study correspond to parametric uncertainties only, representing uncertainties in the surface wind mixing, tidal mixing and tidal bottom friction. Drifters {#App:Drifters} -------- The drifters used in our experiments (Fig. 3) have technical specifications similar to the original CODE drifters designed by Dr. Russ Davis of the Scripps Institution of Oceanography. Each drifter consists of a thin (15cm in diameter) 1m long cylindrical metal body with 0.5m long metal foldable cross-shaped upper arms and lower legs that held a rectangular cloth sail. The drifter is attached by four 20cm rope segments to 4 small (15cm in diameter) round plastic surface buoys for floatation. The round buoy shape minimizes the wave effects compared to flat buoys, the ropes minimizes tilting of the sail compared to the “solid-neck” drifter design, and the large body-to-buoy size ratio insures good water-following characteristics. The drifters are equipped with GPS transmitters that provide positioning fixes every 5min. Based on land tests conducted prior to deployment, the STD of the GPS positioning error is on the order of a few meters (exact values depend on the sky view and location). Estimates of the expected wind slippage of CODE type drifters with standard sails such as ours are 1–2 cm/s in light wind conditions similar to those during our field experiment [@Ohlmann2007; @Poulain2009]. Drifters of the same design are routinely used by the U.S. Coast Guard in SAR operations, as well as in our previous field experiments south of Martha’s Vineyard, MA [@rypina2014eulerian; @rypina2016Invest]. Manikins {#App:Manikins} -------- We used OSCAR Water Rescue Training manikins (Fig. 3) manufactured by Emerald Marine Products (Edmonds, WA) for man-overboard rescue training. Each manikin consists of eight heavy-duty vinyl body parts, PVC fill/drain fittings, six stainless steel joints and two galvanized lifting shackles. The manikin filled with water replicates an 82 kg rescue victim, 1.83 m tall and 0.46 m wide (chest). For an accurate simulation of a person in water, the manikin is filled with water to float at chest level. The manikins are equipped with the same GPS transmitters used for drifters, which provide positioning fixes every 5min. [^1]: corresponding author: [email protected] [^2]: corresponding author: [email protected] [^3]: Because of a relocation of HFR towers in 2018, HFR velocity was not available in the domain shown in Fig. \[fig:2018Experiment\].
{ "pile_set_name": "ArXiv" }
--- abstract: 'Position sensitive beam monitors are indispensable for the beam diagnostics in storage rings. Apart from their applications in the measurements of beam parameters, they can be used in non-destructive in-ring decay studies of radioactive ion beams as well as enhancing precision in the isochronous mass measurement technique. In this work, we introduce a novel approach based on cavities with elliptical cross-section, in order to compensate in ion storage rings. The design is aimed primarily for future heavy ion storage rings of the FAIR project. The conceptual design is discussed together with simulation results.' address: - '$^1$GSI Helmholtzzentrum für Schwerionenforschung, 64291 Darmstadt.' - '$^2$ExtreMe Matter Institute EMMI, 64291 Darmstadt.' - '$^3$Ruprecht-Karls-Universität Heidelberg, 69117 Heidelberg.' - '$^4$AGH University of Science and Technology, 30-059 Krakow.' - $^5$Helmholtz Institut Jena author: - 'M S Sanjari$^{1,2}$, X Chen$^{1,3}$, P Hülsmann$^1$, Yu A Litvinov$^{1,3}$, F Nolden$^1$, J Piotrowski$^{1,3,4}$, M Steck$^1$, Th Stöhlker$^{1,5}$' title: Conceptual design of elliptical cavities for intensity and position sensitive beam measurements in storage rings --- [*Keywords*]{}: Non destructive ion beam detection\ Introduction ============ Cavity based BPM designs have advantages when high resolution and sensitivity are desired, albeit at the cost of narrow-band operation. First ideas date back as early as 1960s. In the literature, most successful designs were utilized in machines with small beam pipe aperture sizes. This leads to precision use of higher order modes, such as the dipole mode in the cavity, for the purpose of position determination . Some of these designs have aimed for similar constructions in ion storage rings (an overview can be found in [@chen2014-gsi]). But since the figures of merit of cavity resonant pick-ups degrade as a result of large beam pipe apertures and low beam intensities, a situation often found in heavy ion storage rings, a more suitable method is needed in order to achieve high sensitivities. Theory of operation =================== Schottky noise analysis ----------------------- The theory of Schottky noise analysis in storage rings is very well explained in the literature [@chattopadhyay1984]. Schottky noise can be used to extract several important beam parameters. Longitudinal signals provide measures of revolution frequency, momentum spread and intensity of the beam, while transversal signals can be used to extract information on tune and chromaticity. Longitudinal Schottky signals can also be used for in-ring mass and lifetime measurements of radioactive ion beams [@bosch2013]. power spectral density of the Schottky signals contains repeated bands around multiples of revolution harmonics that grow in width with increasing frequency. Since these bands essentially contain the same information on the beam, as long as these bands do not overlap, higher harmonics are preferable for higher frequency resolutions, provided the signal is mixed down to base band for analysis. Microwave Cavities as Schottky pick-ups --------------------------------------- The eigenmodes of microwave cavities can be modelled by using ideal RLC elements, where an electric current oscillates at eigenfrequency $f_\nu$ with quality factor $Q_\nu$ and shunt impedance $R_{sh,\nu}=U_\nu^2/P_{diss, \nu}$. Often the material independent *characteristic impedance* is used in order to be able to compare different cavity structures: $$\left(\frac{R_{sh}}{Q}\right)_\nu=\widehat{\left(\frac{R_{sh}}{Q}\right)}_\nu \Lambda_\nu(\beta)^2 \label{eqn:roq}$$ where $\Lambda_\nu(\beta)$ is the so called *transit time factor* that is a function of the relativistic $\beta$ of the beam for each mode. The hat notation shows the ideal or *frozen* characteristic impedance, namely for a cavity with zero length and for a beam travelling with the speed of light. A beam passing through a cavity excites oscillating fields. The resulting standing waves can be extracted out of the cavity by using proper couplers. At critical coupling the average output power of a single particle at the harmonic $m$ is [@sanjari2013] $$\langle P_{out}\rangle|_{mf_r}=\langle P_{diss}\rangle|_{mf_r}=(Ze)^2f_r^2\ \widehat{R_{sh,\nu}}\ \Lambda(\beta)^2$$ The use of RF cavities as pick-ups allows for sensitive detection of particles whenever one of the eigen-frequencies of the cavity matches with a harmonic of the beam. For intensity measurements, a longitudinally sensitive detector can be designed using a circularly cylindrical shallow pillbox, where an electromagnetic field oscillates at its fundamental oscillating eigenmode TM010. By designing proper geometrical dimensions and a tuning mechanism for the eigenfrequency, the above requirements for sensitive particle detection can easily be met. Transversal sensitivity ======================= The R/Q map ----------- The longitudinal characteristic impedance R/Q is an integrated quantity as seen by a particle beam along the axis of the cavity resonator (z axis). Nevertheless, its value depends not only on the transit time factor as seen in equation (\[eqn:roq\]). Due to the distribution (pattern) of the z component of the electric field in every specified mode, it also depends on the transversal position of the beam [@whittum1999]. Since the transit time factor can be determined numerically for a given beam species and energy, in the following we concentrate on the *frozen* characteristic impedance, which is independent from the beam velocity. A plot of the longitudinal characteristic impedance in 3D results in an *R/Q map* which can be used to indicate the sensitivity of each modes for a given transversal offset. A circularly cylindrical shallow cavity shows a relatively flat R/Q map around its center which is also the location of the beam pipe, so this construction can be used for highly sensitive beam intensity measurements (see e.g. [@nolden2011] and [@kienle2013] for single particle sensitivity). Generalization to elliptical cross-section ------------------------------------------ For the transverse sensitivity, a position dependent R/Q map is required where a linear dependency is needed only along one transversal axis and ideally no variation along the other. Application of two such detectors placed sequentially in right angles on the beam axis will provide information on the whole transversal plane. One of the geometries that allows for the above mentioned R/Q map is an elliptical shallow pillbox working in its TM010 mode. The bound electromagnetic fields differ from those of circularly cylindrical resonators [@yeh1962]. By contrast to circular shapes the fall-off of the amplitude of the $E_z$ component of the standing wave along the beam axis is more dramatic along the minor axis of the ellipse, compared to the one along the major axis. By putting the beam pipe off center one can make use of the one-sided steep fall-off while the two-sided slow fall-off along the large axis is even favourably more flattened by the presence of the large aperture beam pipe. In figure \[fig:fig\] the R/Q map is shown as a colour map showing the mode impedance versus transversal position offset using Microwave Studio^^. The results in the top row of this figure are obtained by simulating a circular pillbox cavity of radius 20 cm. The center of the beam pipe is offset to $r/2=10$ cm. The beam pipe has a radius of 9.5 cm. The bottom row shows the same configuration, but a for an elliptical cavity whose major axis is 3 times the minor axis, whereas the beam pipe is still offset halfway along the minor axis (see figure \[fig:dimensions\]). The corresponding frequencies are listed in table \[tab:tab\]. [@lllll]{} Mode& &\ &Freq \[MHz\]&Q&Freq \[MHz\]&Q\ M1& 582.787 & 27630 & 422.838 & 23381\ M2& 901.159&33698 & 509.778 & 25580\ M3& 954.758&37462 & 614.954 & 28984\ M4 & 1142.481 & 64126 & 708.236 & 31812\ \ $^{\dagger}$Calculated in perturbation mode using a perfect copper surface with $\sigma = 5.8\times 10^7$ S/m. This approach has some potential advantages over existing approaches to design circular cavity based beam position monitors that utilize a circular dipole mode as shown in figure \[fig:circ\_mode2e\]. In such designs, the dipole mode needs to be extracted and often the presence of a strong fundamental monopole (see e.g. figure \[fig:circ\_mode1e\]) mode is a problem. In the present design the strong fundamental monopole mode itself is used (figure \[fig:ellip\_mode1e\]) which results in . The fundamental mode of the elliptical design stretches out in the direction of major axis, allowing for a large area where the R/Q is decreasing almost linearly along the x direction (see figure \[fig:ellip\_mode1e\_roqx\]). Simultaneously this area of interest is nearly cleared from the unwanted fields of the dipole mode. For instance, the value of R/Q at y=-2 cm in the elliptical dipole mode as shown in figure \[fig:ellip\_mode2e\] shows a reduction of approximately one order of magnitude compared to that of the circular mode shown in figure \[fig:circ\_mode2e\]. The flatness of the R/Q map should be maximized for a given set of parameters imposed by the storage ring, such as mechanical restrictions and beam pipe aperture at the position of installation, which is preferably where the dispersion is maximum. Coupling mechanisms ------------------- The stretch in one transversal direction comes at a cost of close lying higher order modes (HOM) (see table \[tab:tab\]). This is of little concern in cavity pick-ups designed to have high sensitivities to detect a few charged particles. Also, frequencies above the beam pipe cut-off frequency (here 925 MHz for TE11) will not be trapped. Still, due to the asymmetric design, any remaining HOM fields can be selectively damped using HOM couplers. The elliptical design fortunately allows for a spatial separation (see figure \[fig:ellip\_allmode\_efieldy\]), where electrical antennas can be placed on the face along the major axis pointing into z direction. The places of interest are the maxima of the e-field of the modes in figure \[fig:ellip\_allmode\_efieldy\]. For the fundamental mode, while the signal can also be coupled electrically, a magnetic loop is much more desirable to better control the coupling constant which in turn directly affects the loaded Q value. Mode of operation ----------------- The information on position is hidden in the amplitude of the coupled signal. So by contrast to traditional beam position monitors, this design can determine position of bunched as well as coasting beams. After normalizing the signal to that of a circular reference cavity, this signal needs to be mixed down and passed through a level-detector for further use. Additionally, due to the position dependence of the R/Q, standard techniques can be used to extract information on tune and chromaticity [@caspers2009]. Finally, in order to tune the frequency of the cavity to exactly match a desired Schottky harmonic, remotely controlled plungers are needed, which can be installed symmetrically as shown in figure \[fig:ellip\_cast\]. Conclusion and outlook ====================== This work presents a novel approach to designing a highly sensitive resonant cavity pickup for non-destructive detection of intensity and position of ion beams in storage rings. By using an elliptical geometry, limitations rising from large beam pipe aperture can be circumvented. Future challenges are optimizations with regard to specific beam parameters in the CR storage ring of the FAIR project such as the isochronous ion-optical mode and the position information of cocktail beams [@walker2013]. A bench-top model of a circular pillbox cavity with offset beam pipes has been constructed as a toy model for the verification of the R/Q map. To this end, a fully automated test-bench has been constructed in order to measure this and future model cavities using bead pull perturbation method. A complete account can be found in [@chen2014]. M.S.S. acknowledges partial support by HA216/EMMI. X.C. acknowledges support of the EU contract No. PITN-GA-2011-289485. This work is partially supported by Helmholtz-CAS Joint Research Group HCJRG-108. The authors thank I. Schurig, F. Caspers, P. Kowina, S. A. Litvinov and A. Mostacci for fruitful discussions. References {#references .unnumbered} ========== [9]{} Chen X 2014 [GSI-Report-2014-3]{} Chattopadhyay S 1984 CERN84-11 Bosch F and Litvinov Yu A 2013 *Int. J. Mass. Spec.* **349-250** Sanjari M S 2013 PhD Thesis Goethe University Frankfurt Whittum D and Kolomensky Yu 1999 *Rev. Sci. Inst.* **70** 5 Nolden F et al 2011 *Nucl. Inst. Meth. A* **659** 1 Kienle P, Bosch F et al 2013 *Phys. Lett. B* **726** 4–5 Yeh C W H 1962 PhD Thesis California Institute of Technology Pasadena Caspers F 2009 CERN-2009-005.407 Walker P M et al, *Int. J. Mass Spectr.* **349-350** Chen X et al 2014 *Phys. Scr.* these proceedings
{ "pile_set_name": "ArXiv" }