\caption{(a) Non--destructively probed Rabi oscillations. Each probe pulse contains about $10^5$ photons, is $0.2\,\mu$s long and the repetition period of the pulses is $10.3\,\mu$s. The data corresponds to an average of 10 runs of the experiment, each sampling the atoms $\sim 500$ times. The solid line corresponds to a fit, using a cosine function with exponential amplitude damping. (b) Changing the probe strength from $1.5 \times 10^5$ photons per pulse (data indicated with {\tiny $\blacksquare$}) by a moderate factor of 4, a drastic change in the Rabi oscillation's envelope occurs (data indicated by {\color{red}$ \bullet$}). The solid lines are to guide the eye, only. }
\caption{(a) Ramsey fringes, discrete points represent experimental data, the solid lines are cosine fits to the data. The data indicated with {\color{Gray} $\bullet$} represents the reference trace where no light shifting pulse has been applied to the superposition state during the Ramsey sequence. From (a1)--(a4), the probe pulse photon number has been doubled in each step, starting with $2.6 \times 10^6$. (b) Normalized phase shift and fringe amplitude of Ramsey fringes, extracted from data similar to (a1)--(a4). The solid lines represent the theory curves and follow the experimental data remarkably well. The dashed line in the phase shift data corresponds to the linear dependence expected for a homogeneous system.}
\caption{(Color online) The iteration diagram for map~(\ref{eq:QuadraticMap}) (\textit{a}) $\epsilon>0$ and (\textit{b}) $\epsilon<0$. The stable and unstable fixed points of~(\ref{eq:QuadraticMap}) are shown by \textcolor{red}{$\bullet$} and \textcolor{red}{$\circ$}, respectively}
\caption{(Color online) (\textit{a}) The dependence of the mean length $T$ of the laminar phases on the criticality parameter $\varepsilon=|\epsilon|$ ($\epsilon<0$) for~(\ref{eq:TestQuadraticMap}). The points obtained by the iteration of~(\ref{eq:TestQuadraticMap}) are shown by symbols (\textcolor{blue}{$\bullet$}). The theoretical law ${\ln T\sim \varepsilon^{3/2}}$ is shown by the solid line. (\textit{b}) The distribution of the laminar phase lengths for map~(\ref{eq:TestQuadraticMap}), the criticality parameter value has been selected as $\varepsilon=10^{-11}$. The theoretical exponential law~(\ref{eq:LamPahseLengthDistribution}) is shown by the solid line}
\caption{(Color online) (\textit{a}) The dependencies of the mean length $T$ of the laminar phases on the criticality parameter ${\epsilon=(A_c-A)}$ for the driven van der Pol oscillator with the stochastic force~(\ref{eq:DrivenVdPOscillatorAndNoise}). The points obtained by the numerical integration of~(\ref{eq:DrivenVdPOscillatorAndNoise}) are shown by symbols \textcolor{blue}{$\bullet$}. The theoretical law ${\ln T\sim \varepsilon^{3/2}}$ is shown by the solid line. (\textit{b}) The distribution of the laminar phase lengths for the driven van der Pol oscillator, with the amplitude of the external signal $A=0.0245$ being taken above the critical point $A_c=0.0238$ ($\epsilon=-7\times10^{-4}$). The ordinate axis is presented in the logarithmic scale. The theoretical exponential law~(\ref{eq:LamPahseLengthDistribution}) is shown by the solid line}
\caption{(Color online) The distributions of the laminar phase lengths for the driven periodic oscillator in the presence of noise obtained experimentally. The amplitude $V_m$ of the external signal and the noise dispersion $\sigma$ have been selected the following: (\textcolor{red}{$\blacklozenge$})~${V_m=170}$~mV, ${\sigma=475.17}$~mV; (\textcolor{green}{$\blacksquare$})~${V_m=170}$~mV, ${\sigma=298.11}$~mV; (\textcolor{blue}{$\bullet$})~${V_m=150}$~mV, ${\sigma=141.72}$~mV. The ordinate axis is presented in the logarithmic scale. The approximations corresponding to the theoretical exponential law~(\ref{eq:LamPahseLengthDistribution}) are shown by the dashed lines }
\caption{ (1) {\color{magenta} $\circ$}, $-\langle U \rangle r^3$, averaged interaction between dielectric particle and rough dielectric surface. (2) $\diamond$, $-U_s r^3$, interaction between particle and flat surface. (3) {\color{red} $\triangle$}, $\sigma_u r^3$, variance of interaction for rough surfaces. (4) {\color{green} $\square$}, $\delta U r^3$, difference in mean interaction energy between a flat and a rough surface. Solid lines: $r^{-3.5}$ and $r^{-4}$. $L=1000$. Two weeks of simulation time. Cholesky factor $2.5 {\rm GB}$. $N_g=20$.}
\caption{(color online) Phase diagrams of irradiated BSCCO crystals showing FO (\textcolor[rgb]{0.94,0.00,0.00}{$\circ$}) and SO (\textcolor[rgb]{0.00,0.00,1.00}{$\bullet$}) transitions. (a) In $B_{\phi}=$5G sample the transition is FO at all temperatures. (b) For $B_{\phi}=$10G a FO-SO-FO behavior is found. (c) The SO segment expands in $B_{\phi}=$20G sample. (d) A different $B_{\phi}=$20G sample with lower doping that exhibits a FO-SO-FO-SO sequence. The values of $B_{\phi}$ are shown by dashed lines. \label{phase_diagrams}}
\caption{% \Ca{40} photon-count histograms with $t_b=420\us$, for bright \lev{S}{1/2} state ($\circ$) and dark \lev{D}{5/2} state ($\bullet$) preparations. %%@ Above-threshold events in the dark histogram give $\epsilon_D$; below-threshold events in the bright histogram give $\epsilon_B$. Insets show %%@ time-resolved counts for two individual 17-photon events, one from each histogram, with likelihood ratios $p_B/p_D$. A PMT dark count histogram is %%@ also shown (\smalltriangle), normalized to the same area. Dashed curves are Poisson distributions with the same means as the histograms. The solid %%@ curve is the expected \lev{D}{5/2} distribution taking into account spontaneous decay. }
\caption[agile lc]{ (a): AGILE--GRID \gray light-curve at $\approx 1$-day resolution for E$>$100~MeV in units of $10^{-8}$\,\phcmsec. The downward arrow represents a 2-$\sigma$ upper-limit. The dashed line represents the weighted mean flux. (b): EGRET (triangles) and AGILE--GRID (circle) gamma-ray lightcurve in units of $10^{-8}$\,\phcmsec. EGRET data are from \citet{Hartman1999:3eg}. \label{3c454:fig:lcagile}}
\caption{Heat capacity of TmSe$_{0.45}$Te$_{0.55}$ at high pressures a) Phase diagram constructed from a number of different measurements: {\textcolor[rgb]{0.50,0.50,0.50}{$\bullet$}} (gray) onset of resistivity increase; $\blacktriangle$ maximum of resistivity; {\textcolor[rgb]{1.00,0.00,0.00}{$\blacksquare$}} maximum of ac heat capacity; $\bigstar$ first order transition of heat conductivity; $\blacktriangledown$ first order transition of resistivity and {\color{green}$\blacklozenge$} the onset of deviation of ac heat capacity of pressure run L. The inset shows the pressure-dependent Hall number $n$ at 5 K \cite{bucher1}, where $n$ decreases in phase A and electrons start to delocalize again in phase B. b) ac heat capacity data for different pressure runs: black (zero pressure), red (run M), blue (run N, metallic), green (run L) and magenta (run K). A ferromagnetic phase transition at 6 K is discernible in the metallic phase. The inset to Fig. 2 b) represents the experimental results for seven pressure runs that cross the semimetal to phase B boundary. Peaks in $C_{ac}$ ({\color{red}$ \blacksquare$}) are denoted in Fig. 2 a). The green arrow indicates the phase transition from phase A to phase B. }
\caption{\label{structure} Streamlines inside the drop (without rotation) and their frequencies $\Om\left(\psi\right)$ as given by Eq.~(\ref{Per}). } \end{figure} On every streamline $\Gm_{\psi,\phi}$, we introduce a uniform phase $\chi~\mod (2\pi)$ such that $\chi=0$ on the $x-y$ plane (with $\rho \le 1/\sqrt{2}$) and $\dot{\chi}=\Om\left(\psi\right)$. The unperturbed system, which can be rewritten in terms of $(\psi,\phi,\chi)$ as \[ \dot{\psi} =0, \quad \dot{\phi} =0, \quad \dot{\chi} = \Om(\psi), \] belongs to the class of action-action-angle flows. \subsection{\label{subsec:modelII} Perturbed case} In the perturbed case $0<\eps\ll 1$, the time evolution of the two invariants of the unperturbed system is given by \begin{eqnarray} \nonumber \dot{\psi}&=&-2 a(t)\om_x\psi\sin\phi G\left(\psi,\chi\right),\\ \label{psiphi} \dot{\phi}&=&a(t)\om_z - a(t)\om_x\cos\phi G\left(\psi,\chi\right), \end{eqnarray} where $G(\psi,\chi)=z/\rho$ is $2\pi$ periodic in $\chi$ and has zero average in $\chi$. The time evolution equation for $\chi$ is \be \nonumber \dot{\chi}=\Om(\psi)+a(t) H(\psi,\phi, \chi), \ee where $H$ is $2\pi$ periodic in $\chi$. The dynamics possesses two time scales, a fast one (of order one) associated with $\chi$, and a slow one (of order $1/\eps$) associated with $\psi$ and $\phi$.\\ If $\Om$ and $\om$ are incommensurate, then the averaging over $\Om$ and over $\om$ can be performed independently. In this case, the time-periodic terms in Eq.~(\ref{psiphi}) average out, and the {\em averaged system} reduces to $ \dot{\psi} = 0 , \quad \dot{\phi} = - \eps/2$. Thus in the averaged system the value of $\psi$ is conserved as it was in the unperturbed system; in other words, $\psi$ is an invariant of the averaged system. Each trajectory of the averaged system evolves on two-dimensional nested tori ${\mathcal T}_\psi$. In the perturbed system, $\psi$ is an adiabatic invariant and the motion follows adiabatically the tori ${\mathcal T}_\psi$. \section{\label{sec:resul}Methods and results} \subsection{\label{sec:resul}Mixing generation via resonance phenomena} We now turn to the generation of a 3D chaotic mixing region inside the drop, for which we seek to control both the location and the size. The strategy used for this purpose is to bring a chosen family of unperturbed tori ${\mathcal T}_{\psi}$ into resonance with the perturbation $a(t)$ by adjusting the frequency $\om$ to satisfy the resonance condition \be \label{resonance} n\Om(\psi) - \om =0, \ee for some $n \in \mathbb{N}$ (see Fig.~\ref{structure}). For any fixed $\om$ we denote by $\left\{{\mathcal T}^{(n)}(\om)~|n \in \mathbb{N}\right\}$ the set of resonant tori ${\mathcal T}_{\psi}$ satisfying (\ref{resonance}). Hereafter, we denote the chaotic mixing region generated around ${\mathcal T}^{(1)}(\om)$ by CMR. \subsection{\label{sec:resul}Control of the mixing} Figures~\ref{f2} and ~\ref{sizeII} present {\it Liouvillian sections} of the perturbed system, which consist of 2D projections of time-periodic 3D flows by a combination of a stroboscopic map and a Poincar\'e section (here, the $y=0$ plane). Figure~\ref{f2} shows that a perturbation $a(t)$ creates a 3D CMR around ${\mathcal T}^{(1)}(\om)$ and its location is controlled by varying $\om$ according to Eq.~(\ref{resonance}). In what follows, we analyze the location and the size of the CMR as $\om$ and $\eps$ vary. \begin{figure}[t] \center\epsfig{file=figure2, width=3in} \caption{\label{f2} Liouvillian sections for the amplitude $\eps=0.03$ and the frequencies $\om=0.55, 0.93,1.28,1.41$ (a-d). The (red) dashed line inside the CMR is the torus ${\mathcal T}^{(1)}$.} \end{figure} For small values of $\om$, all resonances are located near the pole-to-pole heteroclinic connections (at $\psi=0$, near the $z$ axis and near the boundaries of the drop, see Fig.~\ref{f2}a). As $\om$ is increased, the CMR penetrates deeper into the drop (Fig.~\ref{f2}b). In the interval $0<\om< \sqrt{2}$, the CMR is the largest chaotic region (compared to chaotic regions corresponding to higher order resonances), with all the other chaotic regions localized close to the $z$ axis and near the drop boundaries (around the heteroclinic orbits); this is due to the shape of $\Om\left(\psi\right)$. As $\om$ is increased further, the CMR moves toward the location of the elliptic fixed points of the unperturbed system, closely following the location of the resonant torus ${\mathcal T}^{(1)}(\om)$ (Fig.~\ref{f2}c). As the value of $\om$ approaches $\sqrt{2}$, the CMR shrinks to the circle of elliptic fixed points (Fig.~\ref{size}). \begin{figure}[t] \center\epsfig{file=figure3, width=3in} \caption{\label{sizeII} Liouvillian sections for the frequency $\om=1.376$ and the amplitudes $\eps=0.01,0.05,0.10,0.20$ (a-d).} \end{figure} Whereas the frequency $\om$ of the rigid body rotation is mostly responsible for the location of the CMR, it is its amplitude $\eps$ which mostly determines its size. Figure~\ref{sizeII} shows that the size of the chaotic mixing regions created by the $n=1$ resonance and by higher order resonances (mostly the $n=2$ resonance) increases as the amplitude of the perturbation increases. Around $\eps\apx 0.20$, the chaotic regions around the heteroclinic orbits and the CMR join together to cover the entire drop volume. \begin{figure}[htbp] \center\epsfig{file=figure4, width=3in} \caption{\label{order} Projection of three characteristic trajectories on the slow phase plane, with $\phi_0=0$ and $\psi_0=0.010,0.073,0.125$.} \end{figure} \begin{figure}[htbp] \center\epsfig{file=figure5, width=3in} \caption{\label{size} Size of the chaotic mixing region; Upper panel: Normalized $\Dt\psi$ {\it vs.} $\om$ for the amplitudes $\eps=0.01,0.05,0.10,0.20$ (a-d); Lower panel: Normalized $\Dt\psi$ {\it vs.} $\eps$ for $\om=0.55,0.93,1.28,1.41$ (a-d).} \end{figure} Recall that in the averaged system the adiabatic invariant $\psi$ is constant. In the exact system, however, along a given trajectory starting at $\psi=\psi_0$ it varies between $\psi^-\left(\psi_0; \om,\eps \right)$ and $\psi^+\left(\psi_0; \om,\eps \right)$. The width $\Dt \psi= \psi^+\left(\psi_0; \om,\eps\right)- \psi^-\left(\psi_0; \om,\eps \right)$ is small away from the resonance, and increases significantly closer to the resonance. The projection of three characteristic trajectories onto the $(\psi,\phi)$-plane (called the {\it slow plane} in dynamical systems) is presented in Fig.~\ref{order}. The narrow regions on the sides are off-resonance trajectories that stay quite close to the corresponding tori ${\mathcal T}_{\psi}$. In between, the middle trajectory deviates much further from its ${\mathcal T}_{\psi}={\mathcal T}^{(1)}(\om)$ and fills the entire CMR. The quantity $\Dt \psi$ is probably the most convenient quantity to estimate the size of the CMR (around ${\mathcal T}^{(1)}(\om)$ for $\om < \sqrt{2}$). The volume between the tori ${\mathcal T}_{\psi^-}(\om)$ and ${\mathcal T}_{\psi^+}(\om)$ gives the CMR size in 3D.\\ The dependence of the size of the CMR (in terms of $\Dt \psi$) on $\eps$ and $\om$ is illustrated in Fig.~\ref{size}. The curves (a)-(d) in the upper and lower panels correspond to the Figs.~\ref{f2}a-d and ~\ref{sizeII}a-d, respectively. For a given $\om$ value (i.e. for a given ${\mathcal T}^{(1)}(\om)$), the size can be controlled by adjusting the value of $\eps$; for example in the range of frequencies $1.181\leq \om \leq 1.357$, the entire droplet exhibits chaotic mixing for $\eps\geq 0.175$. For each smaller value of $\eps$ the size reaches a maximum for a certain value $\om^m\left(\eps\right)$ of the frequency. On the one hand, this property can be used as an optimization technique to obtain the maximal CMR size one can reach for a given amplitude $\eps$ of the rotation. On the other hand, $\Dt \psi$ versus $\eps$ increases quite monotonically for all values of $\om$. The derivation of the maxima locations and estimates of $\Dt \psi$ as functions of the parameters and the order of resonance, will be addressed elsewhere. The structure of the CMR in our case is rather different from that obtained in other problems that possess resonance-induced chaotic advection. Namely, here the size of the CMR vanishes as $\eps$ goes to $0$ and the CMR is localized near the resonance. In contrast, in the flow considered in, e.g., \cite{VWG:2007}, the mixing is caused by resonances, but the CMR occupies a volume on the scale of the whole system. The difference comes from the fact that the averaged change of the frequency of the fast system vanishes in the current system, thus preventing the trajectories starting away from the resonance from approaching it. This property makes the kind of flows investigated here useful as it may be advantageous to localize the mixing in certain parts of the system only. \section{\label{sec:conclu}Conclusion} In summary, we have shown that by applying a judicious oscillatory rotation to a translating drop (an integrable system), one can create a chaotic mixing zone with a prescribed location and size. The appropriate values of the parameters of the perturbation (here, a rotation of a given frequency and amplitude) are determined by quantitative features of the integrable case. For any amplitude of the rotation, the frequency optimizing the CMR size has been obtained. Such an optimization could be useful in guiding the design of practical mixing devices aiming at the best possible mixing rate within individual drops. \begin{acknowledgments} This article is based upon work partially supported by the NSF (grants CTS-0626070 (N.A.), CTS-0626123 (P.S.) and 0400370 (D.V.)). D.V. is grateful to the RBRF (grant 06-01-00117) and to the Donors of the ACS Petroleum Research Fund. C.C. acknowledges support from Euratom-CEA (contract EUR~344-88-1~FUA~F). \end{acknowledgments} \begin{thebibliography}{99} \bibitem{Ismagilov:2003} H. Song, J.D. Tice and R. F. Ismagilov, Angew. Chem. int. Ed. {\bf 42}, 768 (2003). \bibitem{Oddy:2001} M.H. Oddy, J.G Santiago, J.C. Mikkelsen, Anal. Chem. {\bf 73}, 5822 (2001). \bibitem{Bau:2001} H.H. Bau, J. Zhong and M. Yi, Sensors and Actuators B {\bf 79}, 207 (2001). \bibitem{Ouldelmoctar:2003} A. Ould El Moctar, N. Aubry and J. Batton, Lab Chip {\bf 3}, 273 (2003). \bibitem{GlasgowAubry:2003} I.K. Glasgow and N. Aubry, Lab Chip {\bf3}, 114 (2003). \bibitem{Glasgow:2004} I.K. Glasgow, J. Batton and N. Aubry, Lab Chip {\bf 4}, 558 (2004). \bibitem{Goullet:2006} A. Goullet, I.K. Glasgow and N. Aubry, Mech. Res. Commun. {\bf 33}, 739 (2006). \bibitem{Niu:2003} X. Niu and Y-K. Lee, J. Micromech. Microeng. {\bf 13}, 454 (2003). \bibitem{Bottausci:2004} F. Bottausci {\it et~al.}, Phil. Trans. Royal Soc. A {\bf 362}, 1001 (2004). \bibitem{Stremler:2004} M.A. Stremler, F.R. Haselton and H. Aref, Phil. Trans. Royal Soc. A {\bf 362}, 1019 (2004). \bibitem{Baj} K. Bajer and H.K. Moffatt, J. Fluid Mech. {\bf 212}, 337 (1990). \bibitem{KroujilineandStone:1999} D. Kroujiline and H.A. Stone, Physica D {\bf 130}, 105 (1999). \bibitem{Lee:2000} S.M. Lee, D.J. Kim and I.S. Kang, Phys. Fluids {\bf 12}, 1899 (2000). \bibitem{WardandHomsy:2001} T. Ward and G.M. Homsy, Phys. Fluids {\bf 13}, 3521 (2001). \bibitem{Grigoriev:2005} R.O. Grigoriev, Phys. Fluids {\bf 17}, 033601 (2005). \bibitem{Homsy:2007} X.M. Xu and G.M. Homsy, Phys. Fluids {\bf 19}, 013102 (2007). \bibitem{VWG:2007} D. Vainchtein, J. Widloski and R. Grigoriev, Phys. Rev. Lett. {\bf 99}, 094501 (2007). \bibitem{WardandHomsy:2003} T. Ward and G.M. Homsy, Phys. Fluids {\bf 15}, 2987 (2003). \bibitem{GSS:2006} R.O. Grigoriev, M.F. Schatz and V. Sharma, Lab Chip, {\bf 6}, 1369 (2006). \bibitem{FKP:1988} M. Feingold, L. Kadanoff and O. Piro, J. Stat. Phys. {\bf 50}, 529 (1988). \bibitem{VVN:1996a} D. Vainchtein, A. Vasiliev and A. Neishtadt, Chaos {\bf 6}, 67 (1996). \bibitem{VNM:2006} D. Vainchtein, A. Neishtadt and I. Mezi\'c, Chaos{\bf 16}, 043123 (2006). \bibitem{Lima:1990} R. Lima and M. Pettini, Phys. Rev. A {\bf 41}, 726 (1990). \bibitem{CFP2:1996} J.H.E. Cartwright, M. Feingold and O. Piro, J. of Fluid Mech. {\bf 316}, 259 (1996). \bibitem{AubrySingh:2006} N. Aubry and P. Singh, Electrophoresis {\bf 27}, 703 (2006). \end{thebibliography} \end{document} % % ****** End of file apssamp.tex ****** section{Figures and Tables} Figures and tables are typically ``floats'' which means that their final position is determined by \LaTeX\while the document is being typeset.\LaTeX\isn't always successful in placing floats optimally. Figures may be inserted by using either the\texttt{graphics} or \texttt{graphix} packages. These packages both define the \verb+\includegraphics{#1}+ command, but they differ in how optional arguments for specifying the orientation, scaling, and translation of the figure. Fig.~\ref{fig:epsart} shows a figure that is small enough to fit in a single column. It is embedded using the \texttt{figure} environment which provides both the caption and the imports the figure file. \begin{figure} \includegraphics{fig_1}% Here is how to import EPS art \caption{\label{fig:epsart} A figure caption. The figure captions are automatically numbered.} \end{figure} Fig.~\ref{fig:wide} is a figure that is too wide for a single column, so instead the \texttt{figure*} environment has been used. \begin{figure*} \includegraphics{fig_2}% Here is how to import EPS art \caption{\label{fig:wide}Use the figure* environment to get a wide figure that spans the page in \texttt{twocolumn} formatting.} \end{figure*} }
\caption{\red{Spinodal-like behavior of 6400 interacting particles within the region where $d^2f/d\rho^2 < 0$. In the numerics, $\tilde\rho$ is defined by convolution of $\rho$ with a smooth function of finite range $\pm 1/2$; we set $v_0 = 2.5, \lambda = 0.01, \varphi = 250$. The box length is $16$, with periodic boundary conditions, and the mean density is $400$. Dashed lines show the common-tangent densities ($\rho = 60; 800$). }}
\caption[]{Amplitude of the waves versus the driving velocity for varying magnetic induction for the ferrofluids APG\,512a (a) and EMG\,909 (b). The triangles mark the highest induction, the circles an interim value and the boxes the lowest field. The data have been captured at following magnetic inductions:%\opencircle \opentriangle \opentriangledown \fullsquare \opensquare \fullcircle \opendiamond $\bigtriangleup$:~$15.0\,\mathrm{mT}$, $\odot$:~$13.4\,\mathrm{mT}$, $\boxdot$:~$10.4\,\mathrm{mT}$, $\blacktriangle$:~$20.9\,\mathrm{mT}$, \fullcircle:~$14.9\,\mathrm{mT}$, $\blacksquare$:~$10.4\,\mathrm{mT}$. The solid lines display a fit by (\ref{amplitudeeq}).}
\caption{Average depth of the shower maximum \Xmax as function of energy as measured by the \PAO \cite{augerxmaxicrc07}, as well as the Fly's Eye \cite{flyseye}, Haverah Park \cite{haverahpark00}, HiRes/MIA \cite{hires00}, HiRes \cite{hiresxmax}, and Yakutsk \cite{yakutsk} experiments. \TTop: measured values are compared to predictions for primary protons and iron nuclei for different hadronic interaction models QGSJET 01 \cite{qgsjet} (\line), QGSJET II-3 \cite{qgsjet2} (\dashed), SYBILL 2.1 \cite{sibyll21} (\dotted), and DPMJET 2.55 \cite{dpmjet} (\dashdot). \BBottom: comparison to astrophysical models according to \cite{allardxmx} (\line) as well as \cite{berezinskyxmx}, for the latter a dip (\dashed) and an ankle (\dotted) scenario are distinguished. \label{xmax}}
\caption{%----- REVISED ----- % %\small % Near-infrared $JHK$ three-color image of Cloud 2 obtained with QUIRC near-infrared camera. North is up, and east is to the left. $^{12}$CO (1-0) data obtained with Nobeyama 45 m telescope is overplotted as the white contour. The locations of bright NIR sources (IRS 1, 4, 5; Kobayashi \& Tokunaga 2000) and a visible% \textcolor{black}{early-type} % star MR-1 are also indicated. A loosely packed embedded cluster (Cloud 2-N cluster) is seen as an aggregation of red sources in the northern CO peak of the two CO peaks in the Cloud 2-N molecular cloud. A dense embedded cluster (Cloud 2-S cluster) is seen near the CO peak of the Cloud 2-S molecular cloud. This cluster was originally identified as IRS 2 in Kobayashi \& Tokunaga (2000). IRS 3 near Cloud 2-S was also found to be a very small cluster (see also Fig. 2). Most of other faint red sources in this field are faint background galaxies. The NIR image was Gaussian-convolved for viewing purpose. The resultant FWHM of the image is about 0\farcs7. % \label{fig: f1}}
\caption{%----- REVISED ----- % The locations of Cloud 2 and forming stars with respect to the SNR shell GSH 138-01-94 (Stil \& Irwin 2001). The center of the shell is shown with a white cross, while the edge of the shell is shown with a dashed line. White contours show the\ion{H}{1} surface brightness integrated over the velocity range $v_\mathrm{LSR} = -94$ to $-105$ km s$^{-1}$ (from CGPS data), while the blue contours show the $^{12}$CO (1-0) surface brightness (from our Nobeyama data). The background image is the IRAS 60\,$\mu$m image from IRAS IGA data. The open yellow star mark, filled green star marks, and red circles show the locations of the visible % \textcolor{black}{early-type} % star MR-1, the bright NIR sources, and the two embedded clusters, respectively. % The \ion{H}{1} and IRAS images are 1- and 2-pixel Gaussian-convolved, respectively, for better comparison to the CO map. % The bright IRAS source in the center of the image is an unrelated foreground object (a combination of two bright NIR sources, IRS 6 and 7 in Kobayashi \& Tokunaga 2000), and shows the average PSF of this IRAS image.% See the discussion in the main text for detail. % % \label{fig: f5}}
\caption{QFI vs. input configuration parameter $\bet$ for $\gam\in\seq{0,0.25}$. \red{$\triangle$} are the $\ninput=10$ available.}
\caption[Figure1]{ \textbf{Figure 1 | Principle of dark soliton generation.} \textbf{a} Optical setup: A spatial light modulator (SLM) is used to imprint a phase step by exposing part of the condensate to a far-detuned laser beam. \textbf{b} Theoretical curve of a dark soliton's density $|\Psi|^2$ ({\color{blue}\rule[2pt]{10pt}{2pt}}) and phase $\phi$ ({\color{lightblue}\rule[2pt]{10pt}{2pt}}), as described by Eqn. 2. \textbf{c} A typical absorption image of the condensate, taken directly after preparation of the soliton and a subsequent free expansion of $11\,\mathrm{ms}$. Optical density is color- and height-coded for better visibility. \textbf{d} Integrated column density ({\color{red}$\bullet$}) of the data in c together with a fit to the data ({\color{blue}\rule[2pt]{10pt}{2pt}}). \textbf{e} Image of a soliton after an evolution time of $2.8\,\mathrm{s}$.}
\caption[Figure2]{ \textbf{Figure 2 | Dark soliton oscillations in a trapped BEC.} \textbf{a} A set of absorption images showing the soliton position at various times after phase imprinting. The soliton propagates to the right and is reflected off the edge of the condensate after $t\approx$ 80 ms. The corresponding evolution time for each image is given in units of the oscillation period $T$. \textbf{b} Results of a numerical calculation solving the 1D Gross-Pitaevski equation corresponding to our parameters in units of $T$ are presented. Experimentally observed features like density modulations caused by a density wave on the left side of the condensate as well as the development of a tiny second soliton are reproduced. \textbf{c} Axial positions of the soliton ({\color{blue}$\bullet$}) with respect to the center of mass and normalized to the width of the condensate. The oscillation frequency is $\Omega=2\pi \times (3.8\pm 0.1)\,\mathrm{Hz}$. The positition of a second tiny soliton ({\color{lightblue}$\bullet$}) as well as a sinusoidal fit ({\color{blue}\rule[2pt]{10pt}{2pt}}) to the position of the soliton are shown. Each data point was obtained from a different experimental run. The scatter is due to small fluctuations in the preparation process. Errors in extracting the solitons position from the individual images are typically less than 0.02 and therefore not plotted.}
\caption[Figure3]{ \textbf{Figure 3 | Creation and oscillation of a dark-bright soliton.} \textbf{a} A local population transfer in the center of the trapped BEC is achieved by a coherent two-photon Raman process between the two hyperfine states $F=1$ and $F=2$ leading to the generation of a dark-bright soliton (\textbf{b}). \textbf{c} A set of double exposure absorption images showing the density distributions of the two components which undergo slow oscillations in the axial direction. \textbf{d} Time series of the axial positions of the dark ({\color{blue}$\bullet$}) and bright ({\color{red}$\blacktriangle$}) component of the soliton in addition to corresponding sinusoidal fits to the position. Note that the time scale is different by almost an order of magnitude as compared to Fig.\,2c. For details of the first$175\,\mathrm{ms}$ see Fig.\,4.}
\caption[Figure4]{ \textbf{Figure 4 | Collision of a dark and a dark-bright soliton.} A detailed plot of the first $175\,\mathrm{ms}$ of Figure 3 reveals the reflection of an extra dark soliton off the dark-bright soliton. The axial position of the extra dark soliton ({\rule[0pt]{6pt}{6pt}}) is plotted together with a fit to the data (\rule[2pt]{10pt}{2pt}). The mean $e^{-2}$ width of the bright soliton is indicated ({\color{lightred}\rule[0pt]{10pt}{4pt}}). The reflection of the extra dark soliton is very close to that expected from a hard-wall reflection. The fit corresponds to a sine-function mirrored at the reflection time $t_{r}=117\,\mathrm{ms}$.}
\caption{\label{fig1:kinet} (Color online). Dynamics of the transient reflectivity of YVO$_3$ following optical excitation with a power density of ($a$) 25~mJ/cm$^2$, ($b$) 45~mJ/cm$^2$, ($c$) 85~mJ/cm$^2$, ($d$) 100~mJ/cm$^2$, and ($e$) 170~mJ/cm$^2$. Solid lines show the experimental data measured at 25~K, dotted lines are a fit of a double exponential function to the experimental data. The inset shows the amplitude of the fast (\textcolor{red}{$\bigtriangleup$}) and slow ($\blacksquare$) decay components of the transient reflectivity extracted from the fits as a function of the excitation power density. }
\caption{\label{fig2:power} (Color online). Dependence of the transient reflectivity on the pump power density probed at $4$~ps (\textcolor{blue}{$\circ$}) and at $100$~ps (\textcolor{red}{$\blacksquare$}) after the optical excitation at (a) $T=50$~K, (b) $T=100$~K, and (c) $T=140$~K.}
\caption{\label{fig3:temper} (Color online). Temperature dependence of the transient reflectivity probed at $4$~ps (\textcolor{blue}{$\circ$}) and $100$~ps (\textcolor{red}{$\blacksquare$}) for (a) $85$~mJ/cm$^2$ and (b) $190$~mJ/cm$^2$ excitation power density. Lines are guided for eye. Insert shows the stationary reflectance calculated using the optical constants taken from Ref.~\cite{Tsvetkov04}.}
\caption{The spectrum of the diffuse emission for $330^\circ<l<30^\circ,|b|<15^\circ$ as calculated in the optimised GALPROP model for the ISRF with maximal metallicity gradient, {\it Left:} with primary electrons only, {\it Right:} with secondary electrons and positrons only. Line-styles: red solid -- $\pi^0$-decay, green broken -- IC (optical [long dash], IR [short dash], CMB [dot]), green solid -- total IC, cyan solid -- bremsstrahlung, black solid -- extragalactic \gray{} background \citep{SMR04a}, blue solid -- total. Data points: red -- EGRET and green -- COMPTEL, as in \citet{strong05}; magenta -- INTEGRAL/SPI (broken lines: components in fit to positronium + positron annihilation line + unresolved point sources; shaded region: power-law continuum) \citet{bouchet07}. For the SPI power-law continuum the uncertainty is estimated as described in the text. In this and subsequent figures, the identifier (e.g., 53\_6102029RH) corresponds to the GALPROP version and run used; all parameters of the model are contained in the ``GALDEF'' parameter file for future reference and are available from the GALPROP website at\tt http://galprop.stanford.edu. }
\caption{One realization of the random locations of the boxes among the 16 node locations (\textcolor{red}{$\times$}). Each box (grey rectangles) occupies two pixels of this graph and can be placed either parallel or perpendicular to x-axis.}
\caption{Link pairs with identical link geometry in a grid deployment. Each link pair is shown with one link as a dotted (\textcolor{red}{-~-}) line and another link as a solid lines (--). All link pairs with identical link geometry are shown.}
\caption{\textcolor[named]{Black}{ (A) A log-log plot of velocity during growth steps. The line shows $v \sim l_1^{-1}$. (B) Mean value and fluctuation of the fraction of the folded state with $l_p = 12$. } \label{fig.simu.growth} }
\caption{ Time evolution of a long chain ($N=2048$) during the unfolding transition. (A) Time evolution shows slow relaxation after initial expansion (the lower black line). The size is normalized with the equilibrium size, $R^{\rm eq}_{\rm L}$. \textcolor[named]{Black}{The both axes are shown with logarithmic scale.} The kinetics of an ideal polymer is also shown (the upper red line). (B) Interaction between monomers quickly decreases at an early stage of the transition. The two solid lines show $t^{1/4}$ and $t^{1/8}$. The size at the equilibrium state is estimated from the additional simulations at a fixed value of epsilon. \label{fig.long.polymer.unfolg} }
\caption{ Time evolution of the long-axis length of a polymer at an early stage for various persistence length. Bare plot for $l_p=4$ is shown in (A). We determine the initial values for the disentanglement process from $R_1$ in the figure. \textcolor[named]{Black}{ (B) A log-log plot of the time evolution.} Both the long-axis length and time steps are normalized with characteristic space and time scales (see Eqs. (\ref{eq.unfolding.disentanglement.scaling})). \label{fig:simu_unfold} }
\caption{ \textcolor[named]{Black}{ A log-log plot of time evolution of the long-axis length of a polymer at a later stage for various persistence length.} Both the long-axis length and time steps are normalized with characteristic space and time scales according to Eq. (\ref{eq.unfolding.swelling.scaling}) in the text. The dashed and dotted lines show the equilibrium size for $l_p=4$ and $l_p=14$, respectively. \label{fig:simu_unfold:relaxation} }
\caption{Plots relevant to the decay $\pp\to\dipi\jpsi$ (top) and $\pp\to\dipiz\jpsi$ (bottom). The left plots show the dipion recoil mass spectrum and the right plots the dipion mass spectrum. The $J/\psi$ candidates in the continuum sample arise from the tail of the $\psi(2S)$. \symbols \label{fig:pipij} }
\caption{ $\pp\to\eta\jpsi$, $\eta\to\gamma\gamma$ (top) and $\eta\to\dipi\piz$ (middle), and $\pp \to \piz\jpsi$ (bottom): The $\jpsi$ momentum (left) and invariant mass of the decay products (right). \symbols\\label{fig:etaj_and_pi0j} }
\caption{For the decays $\pp\to\gamma\gamma\jpsi$, $\egammalow$ (top) before applying the full-event kinematic fit to the area above $\egammalow = 200\mev$, the di-photon recoil mass for $\chi_{c2,1}$ (middle [lower] left for $J=2$ [$J=1$]), and $\egammalow$ after the full-event kinematic fit (lower right). \symbols\\label{fig:chicJ} }
\caption{System and particle type dependence at \sNNtwohundred for identified trigger particles. (a) Dependence of \jet on \pttrig for \stdassoc (b) Dependence of \jet and \ridge on \ptassoc for \stdtrig (c) Dependence of \jet on \npart (d) Dependence of \ridge on \npart. Data from the fits in (b) are shown in \tref{Table}. The systematic errors due to \vtwo are comparable to those shown for unidentified triggers for \kaon triggers and roughly 3/2 times larger for \lam and \xibary\cite{Jana,Me}. Colour online.}
\caption{System and particle type dependence of the \jet at \sNNtwohundred for identified associated particles. (a) Dependence on \pttrig for \stdassoc (b) Dependence of \ptassoc for \stdtrig (c) Particle ratios in the \ridge and \jet as compared to the inclusive particle ratios. Data from the fits in (b) are shown in \tref{Table}. Colour online.}
\caption{Inverse slope parameter k (\MeV) of \ptassoc for fits of data in \Fref{Figure1}(b), \Fref{Figure2}(b), and \Fref{Figure3}(b) to $A e^{-p_T/k}$. The inverse slope parameter from a fit to $\pi^-$ in \Au from \cite{Pion} above 1.0 \GeV is $k$ = 280.9 $\pm$ 0.4 \MeV for \sNNsixtytwo and is $k$ = 330.9 $\pm$ 0.3 \MeV for \sNNtwohundred. Statistical errors only.}
\caption{Degree distributions $P(k)$ of the \email data (solid line) and in selected models fitted to it. The box plots display medians and first and third quartiles in $100$ network realizations. Whiskers extend from each end of the box to the most extreme values in the data within $1.5$ times the interquartile range from the ends of the box. Outliers are denoted by \textcolor{red}{$+$}. }
\caption{ Left: Purity and nonG of random states as points in the plane $(\mu,\delta)$. Different colors correspond to different dimension of the Hilbert space. The black points correspond to the average (typical) purity and nonG at each dimension. \textcolor{blue}{blue}: $N=2$. \textcolor{green}{green}: $N=5$. \textcolor{yellow}{yellow}: $N=10$. orange $N=15$. \textcolor{red}{red}: $N=20$. Right: Typical purity and nonG in the plane $(\mu,\delta)$ varying the dimension $N=2,\dots,20$. The blue line correspond to the approximate formula reported in the text.\label{f:pur_nong}}
\caption{$\Delta^\mathcal{C}_{b_i}$ represents a basis vector $\left(1,0,\dots,1,0\right)$ of $\mathcal{C}$.} \label{fig:trianrep cycle} \end{figure} For every basis vector $b_i$ we construct a triangular configuration $\Delta^C_{b_i}$. \begin{figure}[hp] \begin{center} \includegraphics[width=150pt]{minors3.eps} % minors3.eps: 1179666x1179666 pixel, 0dpi, infxinf cm, bb= \end{center} \caption{An example of triangular representation $\Delta^\mathcal{C}_B$ of $\mathcal{C}$.} \label{fig:trianrep} \end{figure} Triangular configuration $\Delta^\mathcal{C}_{b_i}$ is obtained from $B^n\cup \mathcal{S}^m$, where $m$ is even and $m\geq n$, $m\geq 4$. Let $J^i$ be the set of indices of nonzero entries of $b_i$. For each $j\in J^i$ we join triangle $\mathcal{S}^m_j$ of $\mathcal{S}^m$ with triangle $B^n_j$. Then, we remove triangle $\mathcal{S}^m_j$ from $\mathcal{S}^m$. Finally, we remove triangles of $B^n$ that are not joined with the sphere. The example of $\Delta^\mathcal{C}_{b_i}$ for $b_i=\left(1,0,\dots,1,0\right)$ is depicted in Figure~\ref{fig:trianrep cycle}. Thus triangular configuration $\Delta^\mathcal{C}_{b_i}$ contains $B^n_j$ if and only if $j\in J^i$. We note that \begin{prop} \label{o:eve} $|\Delta^\mathcal{C}_{b_i}|-w(b_i)$ is always even. \end{prop} Triangular configurations $\Delta^\mathcal{C}_{b_i}$, $i=1,\dots,d$, share the triangles of $B^n$ and do not share spheres $\mathcal{S}^m$. Hence, $\mathcal{A}(\Delta^\mathcal{C}_{b_i})\cap\mathcal{A}(\Delta^\mathcal{C}_{b_j})\subseteq\mathcal{A}(B_n)$ holds for $i<j$, $i,j\in\left\{1,\dots,d\right\}$. Finally, triangular representation $\Delta^\mathcal{C}_B$ of $\mathcal{C}$ is the union of $\Delta^\mathcal{C}_{b_i}$, $i=1,\dots,d$. An example of triangular representation $\Delta^\mathcal{C}_B$ of $\mathcal{C}$ is depicted in Figure~\ref{fig:trianrep}. %with respect to basis $(1,0,\dots,1,0), (1,0,\dots,1,0),\dots,(0,0,\dots,0,1)$. %Let $\mathcal{C}$ be a binary linear code of dimension $d$. Triangular representation $\Delta^\mathcal{C}_B$ of $\mathcal{C}$ is \dfn{balanced} if there is an integer $e$ such that $\left\lvert\Delta^\mathcal{C}_{b_i}\right\rvert-w(b_i)=e$ for all $i=1,\dots,d$. This $e$ is denoted by $\ex{\Delta^\mathcal{C}_B}$. %As $S_m$ exists for every even $k\geq4$, balanced representation always exists. Let $c$ be a codeword of $\mathcal{C}$. Let $c=\sideset{}{^2}\sum_{i\in I}b_i$ be the unique expression of $c$, where $b_i\in B$. The \dfn{degree} of $c$ with respect to basis $B$ is defined to be the cardinality $\left\rvert I\right\lvert$ of the index set. The degree is denoted by $\vd{c}$. We denote by $\ker\Delta^\mathcal{C}_B$ the cycle space of the triangular configuration $\Delta^\mathcal{C}_B$ %Let $C$ be a binary code. Let $ $ be its triangular repre We define linear mapping $f\colon\mathcal{C}\mapsto \ker\Delta^\mathcal{C}_B$ in the following way: let $c$ be a codeword of $\mathcal{C}$; let $c=\sideset{}{^2}\sum_{i\in I}b_i$ be the unique expression of $c$, where $b_i\in B$. %We give a numbering ot We define $f(c):=\chi(\bigtriangleup_{i\in I} \Delta^\mathcal{C}_{b_i})$. %Let us remind of the incidence vector definition. Hence, the entries of $f(c)$ are indexed by the triangles of $\Delta^\mathcal{C}_B$. We have $f(c)^{B^n_j}=1$ if and only if $\bigtriangleup_{i\in I} \Delta^\mathcal{C}_{b_i}$ contains triangle $B^n_j$. \begin{prop} \label{prop:mappingf} Denote $\left\lvert \bigtriangleup_{i\in I} \Delta^\mathcal{C}_{b_i} \right\rvert$ by $m$. Let $c=(c^1,\dots,c^n)$ and \begin{equation*} f(c)=\left(f(c)^{B^n_1},\dots,f(c)^{B^n_n},f(c)^{n+1},\dots,f(c)^{m}\right). \end{equation*} Then, $f(c)^{B^n_j}=c^j$ for all $j=1,\dots,n$ and all $c\in\mathcal{C}$. \end{prop} \begin{pf} We show the proposition by induction on the degree $\vd{c}$ of $c$. Codeword $c$ is equal to $\sideset{}{^2}\sum_{i\in I}b_i$. If $\vd{c}=0$, then $c=0$ and $f(c)=0$ is the incidence vector of the empty triangular configuration. So, the proposition holds for vectors of degree $0$. If $\vd{c}$ is greater than $0$, then $\lvert I\rvert\geq 1$. We choose some $k$ from $I$. Codeword $c+^2b_k$ has degree less than $c$. By induction assumption, the proposition holds for $c+^2b_k$. Let $b_k=(b_k^1,\dots,b_k^n)$. From the definition of $\Delta^\mathcal{C}_{b_k}$, $b_k^j=\chi(\Delta^\mathcal{C}_{b_k})^{B^n_j}$ for all $j=1,\dots,n$. Therefore, \begin{equation*} c^j=(c^j+^2b_k^j)+^2b_k^j=\chi(\bigtriangleup_{i\in I\setminus\{k\}}\Delta^\mathcal{C}_{b_i})^{B^n_j}+^2\chi(\Delta^\mathcal{C}_{b_k})^{B^n_j}=f(c)^{B^n_j} \end{equation*} for all $j=1,\dots,n$. \qed \end{pf} \begin{cor} \label{cor:inj} Mapping $f$ is injective. \end{cor} \begin{lem} \label{lem:cycletriangles} Every non-empty cycle of $\Delta^\mathcal{C}_B$ contains $\Delta^\mathcal{C}_{b_i}-B^n$ as a subconfiguration for some $i\in \left\{1,\dots,d\right\}$. \end{lem} \begin{pf} Every cycle of $\Delta^\mathcal{C}_B$ contains either all triangles or no triangle of $\Delta^\mathcal{C}_{b_i}-B^n$, since $\Delta^\mathcal{C}_{b_i}\cap \Delta^\mathcal{C}_{b_j}\subseteq B^n$ for all distinct $i,j\in \left\{1,\dots,d\right\}$. $B^n$ does not contain non-empty cycles, since the triangles of $B^n$ are disjoint. Therefore, every non-empty cycle contains a triangle of $\Delta^\mathcal{C}_{b_i}- B^n$ for some $i\in \left\{1,\dots,d\right\}$. Hence, every non-empty cycle contains $\Delta^\mathcal{C}_{b_i}- B^n$. \qed \end{pf} %Before the next theorem, let us recall that $(\ker \Delta^\mathcal{C}_B)/S$ means puncturing code $\ker\Delta^\mathcal{C}_B$ along $S$ as it is defined in the end of Section~\ref{sec:prelim}. \begin{thm} \label{thm:trianrep} Let $\mathcal{C}$ be a binary code. Let $\Delta^\mathcal{C}_B$ be its triangular representation with respect to a basis $B$. Mapping $f$ defined above is a bijection of the binary linear codes $\mathcal{C}$ and $\ker\Delta^\mathcal{C}_B$ which maps minimal codewords to minimal codewords. %where $S$ is the set triangles of %$\Delta^\mathcal{C}_B$ out of $B^n$. %Moreover, a bijection between $\mathcal{C}$ and $\ker\Delta^\mathcal{C}_B$ exists, that %maps minimal codewords to minimal codewords, and %$\dim\mathcal{C}=\dim\ker\Delta^\mathcal{C}_B$. \end{thm} \begin{pf} By Corollary~\ref{cor:inj}, $f$ is injective. %Codes $\mathcal{C}$ and $(\ker\Delta^\mathcal{C}_B)/S$ have the same length. So we %identify index $i$ with index $B^n_i$. By Proposition~\ref{prop:mappingf}, $f(c)^i=c^i$ %for all $i=1,\dots,n$. Therefore, $c=f(c)/S$, where $f(c)/S$ means deleting the entries %indexed by the elements of $S$ from $f(c)$. %Thus, $\mathcal{C}\subseteq(\ker\Delta^\mathcal{C}_B)/S$. It remains to be proven that $\dim\mathcal{C}=\dim\ker\Delta^\mathcal{C}_B$. %$\ker\Delta^\mathcal{C}_B$ contains $\left\{f(b_1),\dots,f(b_d)\right\}$, %$\dim\mathcal{C}\leq \dim\ker\Delta^\mathcal{C}_B$. Suppose on the contrary that some codeword of $\ker\Delta^\mathcal{C}_B$ is not in the span of $\left\{f(b_1),\dots,f(b_d)\right\}$. Let $c$ be such a codeword with the minimal possible weight $w(c)$. Let $K$ be a cycle of $\Delta^\mathcal{C}_B$ such that $\chi(K)=c$. By Lemma~\ref{lem:cycletriangles}, cycle $K$ contains $\Delta^\mathcal{C}_{b_i}- B^n$ for some $i\in \left\{1,\dots,d\right\}$. Since $\left\lvert\Delta^\mathcal{C}_{b_i}- B^n\right\rvert>\left\lvert B^n\right\rvert$, $\left\lvert K\bigtriangleup \Delta^\mathcal{C}_{b_i}\right\rvert < \left\lvert K\right\rvert$. Therefore, $w(c)>w(\chi(K\bigtriangleup \Delta^\mathcal{C}_{b_i}))$. This is a contradiction. %Since $\left\lvert %\mathcal{C}\right\rvert=\left\lvert\ker\Delta^\mathcal{C}_B\right\rvert$, mapping $f$ is a %bijection, so $\mathcal{C}=(\ker\Delta^\mathcal{C}_B)/S$. Finally we show that $f$ maps minimal codewords to minimal codewords. Let $d$ be a minimal codeword. Suppose on the contrary that $f(d)$ is not a minimal codeword of $\ker\Delta^\mathcal{C}_B$. Then for some codeword $c$, $f(c)\prec f(d)$. However, $c^i=f(c)^i=1$ implies that $d^i=f(d)^i=1$, so $c\prec d$. This contradicts the minimality of $d$. \qed %So, codeword $f(c)$ distinct of $f(d)$ exists such that $f(c)\preceq f(d)$. %$c^i=f(c)^i=1$ implies $d^i=f(d)^i=1$. Hence, $c\preceq d$ and $c$ is distinct of $d$. This is a contradiction with minimality of $d$. \qed \end{pf} Let $t$ be a triangle of a triangular configuration $\Delta$. A \dfn{subdivision} of triangle $t$ is a triangular configuration obtained from $\Delta$ by exchanging triangle $t$ is replaced by triangles $t_1,t_2,t_3$ in the way depicted in Figure~\ref{fig:trian div}. \begin{figure}[hp] \centering \includegraphics[width=150pt]{subdivision.eps} % subdivision.eps: 1179666x1179666 pixel, 0dpi, infxinf cm, bb= \caption{Triangle subdivision} \label{fig:trian div} \end{figure} \begin{prop} \label{prop:balanced} Every binary code $\mathcal{C}$ of length $n$ and dimension $d$ has a balanced triangular representation $\Delta^\mathcal{C}_B$ such that $\ex{\Delta^\mathcal{C}_B}>n$, where $B$ is an arbitrary basis for $\mathcal{C}$. \end{prop} \begin{pf} Let $\Delta^\mathcal{C}_B$ be an arbitrary triangular representation of $\mathcal{C}$ with respect to a basis $B=\left\{b_1,\dots,b_d\right\}$. We denote by $k_i$ the number $\left\lvert\Delta^\mathcal{C}_{b_i}\right\rvert-w(b_i)$. Every $k_i$ is even by Proposition~\ref{o:eve}. Let $n'$ be the smallest even number greater than $n$. Let $k:=\max\left\{n',k_i \vert i=1,\dots,d\right\}$. For each $i\in\left\{1,\dots,d\right\}$ such that $k_i\neq k$, the following step is applied. We choose a triangle $t$ from $\Delta^\mathcal{C}_{b_i}- B^n$ and subdivide $t$. The number $k_i$ is increased by $2$. If $k_i$ still does not equal to $k$, then we repeat this step. After this procedure, $\Delta^\mathcal{C}_B$ is balanced and $\ex{\Delta^\mathcal{C}_B}>n$. \qed \end{pf} %{\bf nerozumim proc dale musite uvazovat jen sude} \begin{prop} \label{prop:weightf} Let $\mathcal{C}$ be an even binary linear code. Let $\Delta^\mathcal{C}_B$ be its balanced triangular representation with respect to a basis $B$. Then $w(f(c))=w(c)+\vd{c}\ex{\Delta^\mathcal{C}_B}$ for every codeword $c\in \mathcal{C}$. \end{prop} \begin{pf} Write $c$ as $\sideset{}{^2}\sum_{i\in I}b_i$, where $b_i\in B$. Then $f(c)=\chi(\bigtriangleup_{i\in I} \Delta^\mathcal{C}_{b_i})$. Now, $\bigtriangleup_{i\in I} \Delta^\mathcal{C}_{b_i}$ contains all triangles of $\Delta^\mathcal{C}_{b_i}- B^n$ for all $i\in I$. The number of these triangles is $\vd{c}\ex{\Delta^\mathcal{C}_B}$, since $\left\lvert\Delta^\mathcal{C}_{b_i}- B^n\right\rvert=\ex{\Delta^\mathcal{C}_B}$ and $\left\lvert I\right\rvert=\vd{c}$. By Proposition~\ref{prop:mappingf}, $\bigtriangleup_{i\in I} \Delta^\mathcal{C}_{b_i}$ contains triangle $B^n_k$ if and only if $c_k=1$. The number of these triangles is $w(c)$. Therefore, $w(f(c))=w(c)+\vd{c}\ex{\Delta^\mathcal{C}_B}$. \qed \end{pf} \section{Weight enumerator} In this section, we state the connection between the weight enumerator of a code and the weight enumerator of its triangular representation. This will provide a proof of Theorem~\ref{thm:main}. % The \dfn{double code}, denoted by $\mathcal{C}^2$, of a binary linear code $\mathcal{C}$ of length $n$ is the code $\left\{\left(c_1,\dots,c_n,c_1,\dots,c_n\right) : c\in C\right\}$. % \begin{prop} % \label{prop:doubled} % Let $\mathcal{C}$ be a binary linear code. Let $\mathcal{C}^2$ be its double code. Every codeword of $\mathcal{C}^2$ has even weight, and % %. The following equation holds for the weight enumerators: % \begin{equation*} % W_\mathcal{C}(x^2)=W_{\mathcal{C}^2}(x). % \end{equation*} % \end{prop} % \begin{pf} % Every codeword $cc$ of $\mathcal{C}^2$ has even weight, since $w(cc)=2w(c)$ for some codeword $c$ of $\mathcal{C}$. % \begin{equation*} % W_\mathcal{C}(x^2)=\sum_{c\in \mathcal{C}} (x^2)^{w(c)}=\sum_{cc\in \mathcal{C}^2} x^{w(cc)}=W_{\mathcal{C}^2}(x). % \qed % \end{equation*} % \end{pf} % \begin{cor} % The double code of a binary linear code has a balanced triangular representation. % \end{cor} We define the \dfn{extended weight enumerator} (with respect to a fixed basis) by \begin{equation*} W^k_\mathcal{C}(x):=\sum_{\substack{c\in \mathcal{C}\\\vd{c}=k}} x^{w(c)}. \end{equation*} If code $\mathcal{C}$ has dimension $d$, then \begin{equation*} W_\mathcal{C}(x)=\sum_{k=0}^d W^k_\mathcal{C}(x). \end{equation*} \begin{prop} \label{prop:extended polynomial} Let $\mathcal{C}$ be a binary code and let $\Delta^\mathcal{C}_B$ be its balanced triangular representation $\Delta^\mathcal{C}_B$ with respect to the fixed basis $B$. Then \begin{equation*} W^k_{\ker\Delta^\mathcal{C}_B}(x)=W^k_\mathcal{C}(x)x^{k\ex{\Delta^\mathcal{C}_B}}. \end{equation*} \end{prop} \begin{pf} Let $f$ be the mapping defined in Section~\ref{sec:representation}. For every codeword $c$ of degree $k$ of $\mathcal{C}$ there is codeword $f(c)$ of degree $k$ of $\ker\Delta^\mathcal{C}_B$. By Proposition~\ref{prop:weightf}, $w(f(c))=w(c)+k\ex{\Delta^\mathcal{C}_B}$. Therefore, \begin{equation*} \begin{split} W^k_{\ker\Delta^\mathcal{C}_B}(x)& =\sum_{\substack{f(c)\in \ker\Delta^\mathcal{C}_B\\\vd{f(c)}=k}} x^{w(f(c))} =\sum_{\substack{c\in \mathcal{C}\\\vd{c}=k}} x^{w(c)+k\ex{\Delta^\mathcal{C}_B}} =W^k_{\mathcal{C}}(x)x^{k\ex{\Delta^\mathcal{C}_B}}. \qed \end{split} \end{equation*} \end{pf} %Now we show how to compute weight polynomial $W_\mathcal{C}(x)$ of a code $\mathcal{C}$ from the weight polynomial $W_{\Delta^\mathcal{C}_B}(x)$ of its triangular representation $\Delta^\mathcal{C}_B$. \begin{prop} \label{prop:weights} Let $\mathcal{C}$ be a binary code of length $n$. Let $\Delta^\mathcal{C}_B$ be a balanced triangular representation of $\mathcal{C}$. The inequality $k\ex{\Delta^\mathcal{C}_B}\leq w(c)\leq k\ex{\Delta^\mathcal{C}_B}+n$ holds for every codeword $c$ of degree $k$ of $\ker\Delta^\mathcal{C}_B$. \end{prop} \begin{pf} By Proposition~\ref{prop:weightf}, $w(c)=w(f^{-1}(c))+k\ex{\Delta^\mathcal{C}_B}$. Since $0\leq w(f^{-1}(c))\leq n$ holds for every $c\in \ker\Delta^\mathcal{C}_B$, $k\ex{\Delta^\mathcal{C}_B}\leq w(c)\leq k\ex{\Delta^\mathcal{C}_B}+n$. \qed \end{pf} \begin{cor} Let $\mathcal{C}$ be a binary code of dimension $d$ and length $n$. Let $\Delta^\mathcal{C}_B$ be a balanced triangular representation of $\mathcal{C}$ such that $n<\ex{\Delta^\mathcal{C}_B}$. Denote $\ex{\Delta^\mathcal{C}_B}$ by $e$. If $\sum_{i=0}^{de+n}a_ix^i$ is the weight enumerator of $\ker\Delta^\mathcal{C}_B$, then \begin{equation*} W^k_{\ker\Delta^\mathcal{C}_B}(x)=\sum_{i=ke}^{ke+n}a_ix^i. \end{equation*} \end{cor} \begin{pf} By Proposition~\ref{prop:weights}, $w(c)\leq(k-1)e+n$ holds for all codewords $c\in \ker\Delta^\mathcal{C}_B$ of degree less than $k$. Since $n<e$, $w(c)\leq ke-e+n<ke$. By Proposition~\ref{prop:weights}, $(j+1)e\leq w(c)$ holds for all codewords $c\in \ker\Delta^\mathcal{C}_B$ of degree greater than $k$. Since $n<e$, $ke+e<ke+n\leq w(c)$. Hence, enumerator $W^k_{\ker\Delta^\mathcal{C}_B}(x)$ is the sum over all codewords of weight between $ke$ and $ke+n$. \qed \end{pf} \begin{thm} \label{thm:weight polynomial} Let $\mathcal{C}$ be a binary code of dimension $d$ and length $n$. Let $\Delta^\mathcal{C}_B$ be a balanced triangular representation of $\mathcal{C}$ such that $n<\ex{\Delta^\mathcal{C}_B}$. Denote $\ex{\Delta^\mathcal{C}_B}$ by $e$. If $\sum_{i=0}^{de+n}a_ix^i$ is the weight polynomial of $\ker\Delta^\mathcal{C}_B$, then \begin{equation*} W_\mathcal{C}(x)=\sum_{i=0}^{de+n}a_ix^{i\bmod e}. \end{equation*} \end{thm} \begin{pf} $w(c)\leq n$ for every codeword $c\in \mathcal{C}$. Let $f$ be the mapping defined in Section~\ref{sec:representation}. By Proposition~\ref{prop:weightf}, $w(f(c))=w(c)+\vd{c}e$ for every codeword $c$ of $\mathcal{C}$. Since $n<e$, \begin{equation*} w(f(c))\bmod e = (w(c)+\vd{c}e)\bmod e = w(c). \end{equation*} Hence, \begin{equation*} W_\mathcal{C}(x)=\sum_{i=0}^{de+n}a_ix^{i\bmod e}. \qed \end{equation*} \end{pf} Finally, we prove Theorem~\ref{thm:main}. \begin{pf*}{PROOF of Theorem~\ref{thm:main}.} Let $\mathcal{C}$ be a linear binary code of length $n$. %For the double code $\mathcal{C}^2$, $n$ denotes the length of $\mathcal{C}^2$ and $d$ denotes the dimension of $\mathcal{C}^2$. By Proposition~\ref{prop:balanced}, we can construct a balanced triangular representation $\Delta$ of $\mathcal{C}$ such that $\ex{\Delta}>n$. Denote $\ex{\Delta}$ by $e$. Let $W_{\Delta}(x)=\sum_{i=0}^{de+n}a_ix^i$ be the weight enumerator of $\Delta$. % By Theorem~\ref{thm:weight polynomial}, the weight enumerator of $\mathcal{C}^2$ is equal to $\sum_{i=0}^{de+n}a_ix^{i\bmod e}$. %By proposition~\ref{prop:doubled}, the weight enumerator of $\mathcal{C}$ is given by a simple formula of the weight enumerator of $\mathcal{C}^2$. Hence, \begin{equation*} W_\mathcal{C}(x)=\sum_{i=0}^{de+n}a_ix^{i\bmod e}. \qed \end{equation*} \end{pf*} \section{Matching} In this section we reduce computation of the weight enumerator of the even subconfigurations to computation of the weight enumerator of the perfect matchings. Let $\Delta$ be a triangular configuration. A \dfn{matching} of $\Delta$ is a subconfiguration $M$ of $\Delta$ such that $t_1\cap t_2$ does not contain an edge for every distinct $t_1,t_2\in T(M)$. Let $\Delta$ be a triangular configuration. Let $M$ be a matching of $\Delta$. Then the \dfn{defect} of $M$ is the set $E(T)\setminus E(M)$. We denote matching with this defect by $M_{E(T)\setminus E(M)}$. The \dfn{perfect matching} of $\Delta$ is a matching with empty defect. We denote the set of all perfect matchings of $\Delta$ by $\mathcal{P}(\Delta)$. %The \dfn{empty triangle} is the set of edges $t=\{e_1,e_2,e_3\}$ forming boundary of triangle. The weight enumerator of perfect matchings in $\Delta$ is defined to be $P_\Delta(x)=\sum_{P\in\mathcal{P}(\Delta)} x^{w(P)}$, where $w(P):=\sum_{t\in P} w_t$. % \subsection{Reduction for graphs} % The part of reduction for triangular configuration is inspired by reduction for graphs. % So we start with showing reduction for graphs which is taken from Loebl~\cite{}. % % Let $G=(V,E)$ a graph and $v$ a vertex. % % % % The reduction consists from some basic blocks. First is triangular configuration $P$. \subsection{Triangular configuration $P$} \begin{figure}[ht] \centering \includegraphics{pyramid.eps} \caption{Triangular configuration $P$} \label{fig:piramid} \end{figure} The triangular configuration $P$ is depicted in Figure~\ref{fig:piramid}. \begin{prop} Triangular configuration $P$ has exactly two perfect matchings $\{t_1,t_3,t_5,t_7\}$, $\{t_2,t_4,t_6,t_8\}$. \end{prop} %We denote these matchings by $M_\emptyset$ and $M_{t_1t_2t_3}$. \subsection{Closed triangular tunnel $T$} \begin{figure}[h] \begin{center} \includegraphics[width=320pt]{closedtrianedge.eps} \end{center} \caption{Closed triangular tunnel $T$.} \label{fig:tunnel2} \end{figure} Closed triangular tunnel $T$ is depicted in Figure~\ref{fig:tunnel2}. We call triangles $\{a,b,c\}=t_2$ and $\{1,2,3\}=t_1$ ending triangles. \begin{prop} Closed triangular tunnel $T$ has two perfect matchings $M^T_{t_1}=\{t_1,s_4,s_5,s_6\}$, $M^T_{t_2}=\{t_2,s_1,s_2,s_3\}$. % exactly two matchings of cardinality $3$. One with defect $\{1,2,3\}$, second with defect $\{a,b,c\}$. \end{prop} % \begin{lem} % \label{lem:dif} % Let $\Delta_1$, $\Delta_2$ be triangular configurations. Let $\mathcal{P}(\Delta_1)$, $\mathcal{P}(\Delta_2)$ be sets of perfect matchings of $\Delta_1$, $\Delta_2$, respectively. Denote by $I$ the intersection $\Delta_1\cap\Delta_2$. Then % \begin{equation} % \mathcal{P}(\Delta_1\bigtriangleup\Delta_2)=\{(P_1\setminus I)\cup (P_2\setminus I)\vert P_1\cap P_2=\emptyset, P_1\cup P_2\supseteq I, P_1\in \mathcal{P}(\Delta_1), P_2\in\mathcal{P}(\Delta_2)\} % \end{equation} % \end{lem} \subsection{Triangular configuration $E_{pq}$} The \dfn{matching triangular edge} is triangular configuration which is obtained from triangular configuration $P$ and two closed triangular tunnels $T$ in the following way: Let $T_1$ and $T_2$ be closed triangular tunnels and let $t_1^{T_1},p^{T_1}$ and $t_1^{T_2},q^{T_2}$ be ending triangle of $T_1$ and $T_2$, respectively. We identify $t_1^{T_1}$ with $t_1^P$ and $t_1^{T_2}$ with $t_3^P$. Then $E_{pq}$ is defined to be $T_1\bigtriangleup P\bigtriangleup T_2$. Triangular configuration $E_{pq}$ is depicted in Figure~\ref{fig:matchedge}. \begin{prop} \label{prop:trianedgematch} Matching triangular edge has two perfect matchings denoted by $N^1_{pq}$ and $N^0_{pq}$. \end{prop} \begin{pf} There are two matchings. First matching is $N^0_{pq}:=M^{T_1}_{t_1}\cup M^{T_2}_{t_1}\cup \{t_5^P,t_7^P\}$. Second matching is $N^1_{pq}:=M^{T_1}_{p}\cup M^{T_2}_{q}\cup \{t_2^P,t_4^P,t_6^P,t_8^P\}$. Any perfect matching of $E_{pq}$ has to contain $\{t_5^P,t_7^P\}$ or $\{t_2^P,t_4^P,t_6^P,t_8^P\}$. This determines remaining triangles in a perfect matching. Hence, there are just two perfect matchings. \qed \end{pf} \begin{figure}[hp] \centering \includegraphics{matchedge.eps} % matchedge.eps: 7x7 pixel, 0dpi, infxinf cm, bb= \caption{Matching triangular edge} \label{fig:matchedge} \end{figure} We denote the matching $N^1_{pq}$ by $M^1_{pq}$ and the matching $N^0_{pq}\setminus{p,q}$ by $M^0_{pq}$. \subsection{Triangular configuration $T_{pqr}$} The \dfn{matching triangular triangle} is triangular configuration which is obtained from triangular configuration $P$ and three closed triangular tunnels $T$ in the following way: Let $T_1$, $T_2$ and $T_3$ be closed triangular tunnels and let $t_1^{T_1},p^{T_1}$, $t_1^{T_2},q^{T_2}$ and $t_1^{T_3},r^{T_3}$ be ending triangle of $T_1$, $T_2$ and $T_3$, respectively. We identify $t_1^{T_1}$ with $t_1^P$, $t_1^{T_2}$ with $t_3^P$ and $t_1^{T_3}$ with $t_5^P$. Then $T_{pqr}$ is defined to be $T_1\bigtriangleup P\bigtriangleup T_2\bigtriangleup T_3$. Triangular configuration $T_{pqr}$ is depicted in Figure~\ref{fig:matchtrian}. \begin{prop} Matching triangular triangle has two perfect matchings denoted by $N^1_{pqr}$ and $N^0_{pqr}$. \end{prop} \begin{pf} There are two matchings. First matching is $N^0_{pqr}:=M^{T_1}_{t_1}\cup M^{T_2}_{t_1}\cup M^{T_3}_{t_1}\cup \{t_7^P\}$. Second matching is $N^1_{pqr}:=M^{T_1}_{p}\cup M^{T_2}_{q}\cup M^{T_3}_{r}\cup \{t_2^P,t_4^P,t_6^P,t_8^P\}$. Any perfect matching of $T_{pqr}$ has to contain $\{t_5^P,t_7^P\}$ or $\{t_2^P,t_4^P,t_6^P,t_8^P\}$. This determines remaining triangles in a perfect matching. Hence, there are just two perfect matchings. \qed \end{pf} We denote the matching $N^1_{pqr}$ by $M^1_{pqr}$ and the matching $N^0_{pqr}\setminus{p,q,r}$ by $M^0_{pqr}$. \begin{figure}[hp] \centering \includegraphics{matchtriangle.eps} % matchtriangle.eps: 1179666x1179666 pixel, 0dpi, infxinf cm, bb= \caption{Matching triangle} \label{fig:matchtrian} \end{figure} \subsection{Triangular configuration $C_{t_1t_2\dots t_n}$} This part of reduction is analogous to the reduction for graphs described in Galluccio et al~\cite{galluccio01optimization}. Let $t^C_1,t'_1$ be empty disjoint triangles. Let $t_2^C,\dots,t^C_n,t'_2,\dots,t'_n$ be disjoint triangles. Then $C_{t_1t_2\dots t_n}$ is defined to be $\left(\bigtriangleup_{i=1}^nt_i\right)\bigtriangleup\left(\bigtriangleup_{i=1}^nt'_i\right)\bigtriangleup\left(\bigtriangleup_{i=1}^nE_{t_it'_i}\right)\bigtriangleup\left(\bigtriangleup_{i=2}^nE_{t_it'_{i-1}}\right)\bigtriangleup\left(\bigtriangleup_{i=1}^{n-1}E_{t'_it'_{i+1}}\right)$. The configuration is depicted in Figure~\ref{fig:C}. \begin{figure}[h] \centering \includegraphics{C.eps} % C.eps: 0x0 pixel, 300dpi, 0.00x0.00 cm, bb= \caption{Triangular configuration $C_{t_1t_2\dots t_n}$} \label{fig:C} \end{figure} \begin{prop} Let $M^I_C$ denote perfect matching containing triangles $t_i,i\in I$. Then there exists exactly one perfect matching $M^I_C$ of $C_{t_1t_2\dots t_n}$ if and only if $\lvert I\rvert$ is even. \end{prop} \begin{pf} We construct matching by the following algorithm. In the first step. If $t_1\in I$ then $M:=M^1_{t_1t'_1}$ else $M:=M^0_{t_1t'_1}$. We say that first step is even if $t'_1$ is covered otherwise odd. (first step is even if and only if $t_1\notin I$). In the $i$-th step. \begin{itemize} \item If $t'_i$ is covered by $M$ and $t_{i+1}\notin I$ then $M:=M\cup M^0_{t'_it_{i+1}}$. \item If $t'_i$ is covered by $M$ and $t_{i+1}\in I$ then $M:=M\cup M^0_{t'_it_{i+1}}\cup M^0_{t'_it'_{i+1}}\cup M^1_{t_{i+1}t'_{i+1}}$. \item If $t_i$ is not covered by $M$ and $t_{i+1}\notin I$ then $M:=M\cup M^0_{t'_it_{i+1}}\cup M^1_{t'_it'_{i+1}}$. \item If $t'_i$ is not covered by $M$ and $t_{i+1}\in I$ then $M:=M\cup M^1_{t'_it_{i+1}}$. \end{itemize} The $i$-th step is even if $t_i$ is covered. From definition of $i$-th step follows that parity of step changes if $t_{i+1}\in I$. Hence, the algorithm succed in construction of perfect matching if and only if $\lvert I\rvert$ is even. \qed \end{pf} \subsection{Reduction} Let $\Delta$ be a triangular configuration. We construct triangular configuration $\Delta'$ such that every even subconfiguration of $\Delta$ uniqually coresponds to one perfect matching of $\Delta'$ and a natural weight-preserving bijection between the set of the even subconfiguration of $\Delta$ and the set of the perfect matchings of $\Delta'$. We put into $\Delta'$ empty disjoint triangles $t_e$ for every tuple $(t,e)$ where $e\in E(\Delta)$ and $t\in T(\Delta)$. We add to $\Delta'$ matching triangle $T_{t_at_bt_c}$ for every triangle $t\in T(\Delta)$, where $a,b,c$ are edges of $t$. %We denote a matching of $\Delta'$ containing only triangles of $T_{t_at_bt_c}$ by $M^t$. We assign weight $w(t):=1$ to one triangle in matching $M_t^1$ and weight $0$ otherwise. We add to $\Delta'$ triangular configuration $C_{t^1_e\dots t^n_e}$ for every edge $e\in E(\Delta)$ where $t^1_e,\dots,t^n_e$ are triangles incident with $e$ in $\Delta$. %We denote a matching of $\Delta'$ containing only triangles of $C_{t^1_e\dots t^n_e}$ by $M^e$. We assign weight $w_t:=0$ to all triangles of $C_{t^1_e\dots t^n_e}$. \begin{thm} Let $\Delta$ be a triangular configuration and $\Delta'$ be matching reduction of $\Delta$. Let $C$ be an even subconfiguration of $\Delta$. Then there exists exactly one perfect matching $M_C$ in $\Delta'$. In $\Delta'$ are not any others perfect matchings. \end{thm} \begin{pf} Let $C$ be an even subconfiguration of $\Delta$. We construct perfect matchings $M_C$ in $\Delta'$. We denote matchings $M^1_{t_at_bt_c}$ and $M^0_{t_at_bt_c}$ of $T_{t_at_bt_c}$ by $M_t^1$ and $M_t^0$, respectively. We denote the set $\{i\vert e\in T(t_i),t_i\in C\}$ by $I_e$. $M_C:=\{M_t^1\vert t\in C\}\cup\{M_t^0\vert t\notin C, t\in T(\Delta)\}\cup\{M_e^{I_e}\vert e\in E(\Delta)\}$. %$M':=\{M_\emptyset^t\vert t\notin C, t\in T(\Delta)\}\cup\{M_{t_at_bt_c}^t\vert t\in C\}\cup\{M^e_{t_i,e\in T(t_i),t_i\notin C,t_i\in T(\Delta)}\vert e\in E(\Delta)\}$. Matching $M_C$ is perfect. We show that there is no other matching. Every matching triangle $T_t$ has to be covered by $M^1_t$ or $M^0_t$. Thus $C_e$ has to be covered by $M^I_e$ for some even $I$. So every perfect matching in $\Delta'$ defines even subset in $\Delta$. \qed \end{pf} \begin{prop} Let $\Delta$ be a triangular configuration. Let $\Delta'$ be its matching representation. Let $C$ be an even subconfiguration and $M_C$ be coresponding perfect matching then $\lvert C\rvert=w(M_C)$. \end{prop} \begin{pf} \begin{align*} w(M_C)&=\sum_{t\in C}w(M^1_t) + \sum_{t\notin C, t\in T(\Delta)}w(M^0_t) + \sum_{e\in E(\Delta)}w(M_e^{\{i\vert e\in T(t_i),t_i\in C\}})\\ &=\sum_{t\in C}1 + \sum_{t\notin C, t\in T(\Delta)}0 + \sum_{e\in E(\Delta)}0\\ &=\lvert C\rvert \end{align*} \qed \end{pf} The following theorem is a consequence. \begin{thm} Let $\Delta$ be a triangular configuration. Let $\Delta'$ be its matching representation. Then $W_\Delta(x)=P_{\Delta'}(x)$. \end{thm} \begin{ack} This article extends a result of my master thesis, written under the direction of Martin Loebl. I would like to thank Martin Loebl for helpful discussions and continuous support. \end{ack} \bibliography{references}{} \bibliographystyle{elsart-num} \end{pf}\end{document} }
\caption{(color online) Minimum conductivity normalized to its value at $T=5$~K as a function of $T$ for three devices before (S1:\textcolor{red}{$\square$},S2:\textcolor{blue}{$\circ$}) and after (S1:\textcolor{red}{$\blacksquare$},S2:\textcolor{blue}{$\bullet$},S3:\textcolor{green}{$\blacktriangle$}) annealing. Data for S3 before current annealing are not available. The $T$-dependence increases considerably after annealing. Inset: Maximum resistivity for S2 before (\textcolor{blue}{$\blacktriangle$}) and after (\textcolor{red}{$\blacksquare$}) current annealing. }
\caption{Azimuthal angular distribution of elastic scattered projectiles for 3 keV $^{3}$He atoms impinging on LiF(001) along the direction \TEXTsymbol{<}% 110\TEXTsymbol{>}, with the incidence angle $\protect\theta _{i}=1.1\deg $. Solid line, differential probability derived from the surface eikonal approach, including polarization effects; dashed line, surface eikonal results without including the projectile polarization. Full stars, experimental spots of Fig. 5 of Ref. [1]. Empty circles, classical distribution, as explained in the text, in \textit{absolute} scale.}
\caption{Similar to Fig. 3 for 8.6 keV $^{4}$He atoms impinging on LiF(001) along the direction \TEXTsymbol{<}100\TEXTsymbol{>}, with the incidence angle $\protect\theta _{i}=0.71\deg $. Dash-dotted line, first Born approximation [Eq. (7)]. (a) Full stars, experimental spots, and (b) thick solid line, experimental intensity, both drawn from Fig. 2 of Ref. [3].}
\caption{Similar to Fig. 3 for 0.2 keV $^{4}$He atoms impinging on LiF(001) along the direction \TEXTsymbol{<}110\TEXTsymbol{>}, with the incidence angle $\protect\theta _{i}=1.5\deg $. Full stars, experimental data from Fig. 1 of Ref. [2]}
\caption{\label{fig-square_out} Transmission coefficient (-.-.) of the NLSE for the rectangular potential well (\ref{potwell}) (parameters $m=\hbar=1$, depth $V_0=50$, width $2a=40$ and attractive nonlinearity $g=-1$). Also shown are the transmission probabilities for the left (width $10$, \textcolor{rot}{$---$}) and right potential (width $30$, \textcolor{blau}{$-$}) obtained from dividing $V(x)$ into two parts. These two probabilities are equal at the resonance at $\mu_{\rm res}$ where the potential $V(x)$ is transparent.}
\caption{The $L_1$ norm of the error between the CFB equations' solution and the entropy solution. The error is displayed for four different values of $t$. $t=0$ \solid, $t=1$ \dashed, $t=2$ \dashdot, and $t=3$ \dashdotdot. The error approaches zero roughly linearly as $\alpha \to 0$.}
\caption{The canonical phase diagram of the extended FKM with Hund coupling for the 1D lattice and $U=2$, $J=0.5$. The red crosses \color{red}$\times$\\color{black} and horizontal straight line segments mark stability points or intervals of periodic phases. Their unit cells are drawn as sequences of small circles and plus and minus signs that correspond to sites non-occupied, occupied by the spin up and occupied by the spin down \emph{f-electrons}, respectively.}
\caption{The canonical phase diagram of the extended FKM with Hund coupling for the 2D square lattice and $U=4$, $J=0.5$. The lines $\rho_f=1-\rho_d$, $\rho_f=2-\rho_d$ and $\rho_f=1-\rho_d/2$ are merely visual guides. The red crosses \color{red}$\times$\\color{black} and horizontal straight line segments mark stability points and intervals of the periodic phases, respectively. Their unit cells are drawn as sequences of small circles and plus and minus signs that correspond to sites non-occupied, occupied by the spin up and occupied by the spin down \emph{f-electrons}, respectively. A number of pairs of phases have the same unit cells but different translation vectors. Unit cells of the phases are displayed along the horizontal lines in the middle between the lines $\rho_f=1-\rho_d$ and $\rho_f=1-\rho_d/2$, and in the middle between the lines $\rho_f=1-\rho_d/2$ and $\rho_f=2-\rho_d$. The configurations located along the line $\rho_f=1$ are presented in Fig. 3 and a set of characteristic configurations D1, D2, D3, D4, D5 is shown in Fig. 6.}
\caption{ (Color online) The dependence on $X$ of $\Lambda^{(1)}(X,M)$ in the geometry of Fig.~\ref{fig:3Q1D geo}(a) with transverse RBC is shown for $M=4$ ($\textcolor{blue}{\odot}$), $8$ ($\textcolor[rgb]{0,0.7,0.25}{\blacksquare}$), and $16$ (\textcolor{red}{$\bullet$}). The dependence on $X$ of $\Lambda^{(2)}(X,M)$ with transverse RBC is shown for $M=4$ ($\textcolor{blue}{\times}$), $8$ ($\textcolor[rgb]{0,0.7,0.25}{+}$), and $16$ ($\textcolor{red}{*}$). The dependence on $X$ of $\Lambda^{(1)}(X,M)$ with transverse PBC is shown for $M=4$ (blue dashed curve), $8$ (green dotted curve), and $16$ (red solid curve). The vertical dashed lines identify the critical points $X^{\ }_{s}$ and $X^{\ }_{l}$ deduced in Ref.~\onlinecite{Obuse07b} when transverse PBC are imposed. The inset displays the same data points with a logarithmic vertical scale. }
\caption{ (Color online) (a) Dependence on $\alpha^{(\kappa)}$ of the multifractal spectra $f^{(\kappa)}$ in the bulk $(\kappa)=(2)$ (\textcolor[rgb]{0,0.7,0.25}{$\blacktriangle$}), at the $\textsf{S}$ boundaries $(\kappa)=(1,\mathrm{O})$ ($\textcolor{blue}{\odot}$), and at the $\textsf{S}'$ boundaries $(\kappa)=(1,\mathbb{Z}^{\ }_{2})$ ($\textcolor{red}{\bullet}$), at $X=X^{\ }_{l}$. Dependencies on $q$ of $\alpha^{\ }_{q}$ and $f^{\ }_{q}$ are shown in (b) and (c), respectively, with the same symbols as in (a). The bulk and boundary multifractal spectra for the SU(2) model defined in Ref.% ~\onlinecite{Asada02} are shown by solid and dashed curves, respectively. }
\caption{(color online). Comparison of jamming coverage $\theta_J$ for $\epsilon = -5.0, \ldots, 0.24$ from the series expansion up to order $13$ using Pad\'e approximant$[6,6]$, $[6,7]$ and $[7,6]$ for the optimization described in the text ($+$) and from MC simulations for a $200\times200$ - lattice and $100$ iterations (red $\sline$). Errors in simulation data points are of order $0.001$. Inset: Sticking probability $S$ as function of coverage $\theta$ up to order $N = 9$ for three different $ \epsilon = 0.1 (\sline)$, $\epsilon = 0.0 (\dashline)$, and $\epsilon = -0.5$ (\ddashline). }
\caption{The thickness of the shocks formed in CFB vary depending upon the value of $\alpha$. (a) As $\alpha$ decreases the thickness of the shock decreases. $\alpha=0.08$=\dashed, $\alpha=0.05$=\full, $\alpha=0.02$=\dotted (b) The thickness of the traveling shock decreases linearly with $\alpha$. Helmholtz filter, $g(x)=\frac{1}{2}exp(-|x|)$=\full, Gaussian filter, $g(x)=\pi^{-1/2}exp(-x^2)$=\dashed, Hat filter, $g(x)=\{x-1 \text{ for } x\in(-1,0),1-x \text{ for } x\in(0,1), 0 \text{ otherwise}\}$=\chain, Tophat filter, $g(x)=\{1 \text{ for } x\in [-\frac{1}{2},\frac{1}{2}], 0 \text{ otherwise} \}$ =\dashddot.}
\caption{Spectral energy decompositions for 1D and 2D CFB using Helmholtz and Gaussian filters . All simulations were performed with $\alpha = 0.08$. (a) and (b) are one dimensional simulations at $t=3$ with initial conditions $u(x)=\sin(x)$ for Helmholtz and Gaussian filters respectively. In figure (a) one can see that the energy cascade slope drops dramatically after wavelength $\frac{1}{\alpha}$. This occurs similarly in figure (b), with the exception that the spectral energy decreases exponentially due to the Gaussian filter. (c) and (d) are two dimensional simulations at $t=1$ with Gaussian pulses as the initial conditions, again with Helmholtz and Gaussian filters respectively. The spectral energy decompositions show similar characteristics as the one dimensional simulations. The spectral energy of $\ub$ \full, $\ubarb$ \dashed, and a reference slope of -2 \chain \, are shown.}
\caption{Energy decay for 1D and 2D CFB. $\int \ub \cdot \ub$ \chain, $\int \ub \cdot \ubarb$ \dashed, $\int \ubarb \cdot \ubarb$ \full. Figures (a) and (b) show the energy decay for 1D CFB with the Helmholtz and Gaussian filter respectively. Initial conditions $u(x,0)=exp(-3x^2)$, with $\alpha=0.05$ and 1024 gridpoints. Figures (c) and (d) show the energy decay for 2D CFB with the Helmholtz and Gaussian filter. Initial conditions $u_1(x,y,0)=exp(-30x^2-30y^2)$ and $u_2(x,y,0)=exp( -30x^2-30y^2)$, with $\alpha=0.05$ and 128 $\times$ 128 gridpoints. Decay rates are similar to those seen in figure \ref{viscousenergy}. }
\caption{Here the energies for CFB are shown. $\int \ub \cdot \ub$ \chain, $\int \ub \cdot \ubarb$ \dashed, $\int \ubarb \cdot \ubarb$ \full. In (a) the initial conditions are $u_0=C(x -\pi)/(1+100(x-\pi)^4) $, with $C$ chosen such that $\max(u_0) =1$, and (b) with a random initial condition. In both cases the energies increase initially, but then behave with normal energy decaying behavior.}
\caption{ The energy $\int \ub \cdot \ub$ is compared for three different values of $\alpha$. $\alpha=0.08$ \chain, $\alpha=0.05$ \dashed, and $\alpha=0.02$ \full. The initial conditions are $u_0=C(x -\pi)/(1+100(x-\pi)^4) $, with $C$ chosen such that $\max(u_0) =1$. For all three values of $\alpha$ there is a brief increase of energy, but as $\alpha$ becomes smaller, this increase in energy becomes less substantial. }
\caption{\label{work} Irreversible work calculations associated with the process of transforming the interacting, defect-free solid into an Einstein crystal (Panel $a$) and the process of introducing a monovacancy into the defect-free solid (Panel $b$) as a function of the process duration $t_{\rm sim}$ measured in MC sweeps. Results are displayed for the forward processes ({\color{blue} $\triangle$}), backward processes ({\color{blue}$\bigcirc$}), as well as the corresponding unbiased estimator of Eq.~(\ref{w}) ($\ast$). Each point was obtained as the average over thirty independent simulations carried out $N=500$ particles. The solid line is an average of the last four values of the unbiased work estimator, for which the processes have reached the linear-response regime.}
\caption{\label{f_QHA_sw} Vacancy-formation free energy as a function of temperature. Results were obtained for $b = 1.72\sigma$ and $N=500$. HA ({\color{blue} $\times$}) and RW ($\bigcirc$) results. The lines are guides to the eyes.}
\caption{\label{free_size} Vacancy-formation free energy as a function of system size at $k_{B}T=1$. The straight was obtained through a least-square linear fit. Results obtained for $b=1.72\sigma$ and $N=108$ ({\color{magenta} $\square$}), $N=256$ ({\color{blue} $\blacksquare$}), $N=500$ ({\color{black} $\bigcirc$}), $N=864$ ({\color{Orange} $\bigtriangleup$}) and $N=1372$ ($\bigtriangledown$). Inset shows the formation free energy as a function of $k_{B}T$ for the mentioned system sizes.}
\caption{\label{free} Extrapolated vacancy-formation free energy as a function of $b$ at $k_{B}T=1.0$. Results obtained using the RW method ($\blacktriangle$) and the HA ({\color{blue} $\times$}). Lines are plotted to guide the eyes.}
\caption{\label{c_dens} Concentration of vacancies as a function of density for $b=1.72\sigma$. Results obtained using RW method ($\blacktriangle$) and HA({\color{blue} $\times$}). Lines are plotted to guide the eyes.}
\caption{\label{wigseitz} Arrhenius plot of Wigner-Seitz rejection probability as a function of temperature. Results were obtained for $b=1.72\sigma$ and $N=108$ ({\color{magenta} $\square$}), $N=256$ ({\color{blue} $\blacksquare$}), $N=500$ ({\color{black} $\bigcirc$}), $N=864$ ({\color{Orange} $\bigtriangleup$}) and $N=1372$ ($\bigtriangledown$). The line represents the linear regression to the results obtained with $N=1372$.}
\caption{Reduction from \planarcolor to immersibility or embeddability of a curve.}
\caption{(Color online) Temporal evolution, $\tau_{eddy} \approx 4.5$, for $1536^3$ DNS (solid, black), $256^3$ $k_\alpha = 33$ \lamhda (dashed, green online), $256^3$ under-resolved ``DNS'' (dotted, red online), \neu{and $384^3$ $k_\alpha = 33$ nonhyperdiffusive-\lamhda (dash-dotted, blue online).}{\bf (a)} Time evolution of the energies: kinetic (lower curves), magnetic (middle curves) and total (upper curves). {\bf (b)} Time evolution of total enstrophy, $\left<j^2+\omega^2\right>$ ($\left<j^2+\boldsymbol{\omega}\cdot\bar{\boldsymbol{\omega}}\right>$ for \lamhda \neu{and $\left<\vec{j}\cdot\bar{\vec{j}}+\boldsymbol{\omega}\cdot\bar{\boldsymbol{\omega}}\right>$ for the nonhyperdiffusive case).} Note that \lamhda gives a better agreement to the total dissipation rate up to the maximum time that the high resolution DNS is performed. Also note that the DNS equivalent to the \lamhda run presented here is not feasible on present-day computers at a reasonable cost.} \label{FIG:EVST} \end{center} \end{figure} Having now shown that \lamhda does not suffer the same drawbacks with regards to energy spectra as \lans, we may turn our attention to \resp{}{a practical application.} The purpose of a SGS model or LES is to make predictions about large Reynolds number flows at a reduced computational expense. From the scaling arguments in Refs. \cite{FHT01,PGHM+07a}, using simulations conducted at $Re \approx 2200$, and assuming a $k^{-1}$ scaling, we can estimate $\alpha=1/33$ for a $256^3$ \lamhd-LES ``prediction'' of our $1536^3$ MHD-DNS. Time evolution of the energies and the total enstrophy are shown in Fig. \ref{FIG:EVST} for much later times than reasonably attainable with the MHD DNS with present-day computers. Also shown are results for solving the MHD equations, Eqs. (\ref{eq:mhd}) with $\nu =2 \cdot 10^{-4}$ and a resolution of $256^3$: a so-called ``unresolved DNS'' \resp{}{and the non-hyperdiffusive modified-\lamhda from the previous section.} Before the peak of dissipation, $t\approx4$, the unresolved DNS gives a poorer prediction of the total dissipation and total energy which is then followed by a significantly larger and somewhat later peak of dissipation, at $t\approx5$ than the resolved DNS and the \lamhda LES. \resp{}{The non-hyperdiffusive \lamhda is not expected to perform well as a SGS model and it is seen to be clearly under-dissipative. The ratio of magnetic to kinetic dissipation is $\approx1.5$ for the DNS, $\approx2.9$ for \lamhd, $\approx1.1$ for the under-resolved DNS, and $1.4$ for the non-hyperdiffusive model.} \resp{}{Together with Fig. \ref{FIG:EVST} (b) these ratios show that \lamhda achieves accurate total dissipation by an excess of magnetic dissipation and a reduction of kinetic dissipation (both at the small scales). This feature has already been depicted in Fig. 15 of Ref. \cite{MMP05a}.} Compensated energy spectra for \resp{}{the peak of dissipation} ($t\in[2.7,3.7]$) are shown in Fig. \ref{FIG:COMP2}. For the under-resolved DNS, \resp{}{we observe} the appearance of a tail at large wavenumbers with a $k^2$ spectrum as predicted using statistical mechanics arguments for truncated systems in the ideal ($\nu=0$, $\eta=0$) case \cite{FrPoLe+1975}. The under-resolved spectra are not significantly different from the resolved DNS, but note that a reliable and convincing determination of spectral indices, beyond visual inspection, does require high resolutions. Comparing now the resolved DNS and the \lamhda run, the quality of the spectra are similar for scales larger than $\alpha$. Recall that differences at the largest scales, stem from the differences in initial conditions as stated in Section \ref{SEC:LAMHDGOOD}, and from time evolution of the flow. Finally, noting that the computer saving here is $6^3$ in memory and $6^4$ in running time, we conclude that the \lamhda continues to behave satisfactorily, as already shown both in two space dimensions \cite{MMP05a,PGMP05,PGHM+06} and in 3D \cite{MMP05b}, in particular in the context of the dynamo problem of generation of magnetic fields by velocity gradients; thus, \lamhda may prove to be a useful tool in many astrophysical contexts where magnetic fields are dynamically important, such as in the solar and terrestrial environments, or in the interstellar and intergalactic media. %Given this demonstration of the spectral properties of \lamhda as a LES with a decrease in linear resolution of a factor of $6$ and \note{ considering previous tests in Pablo's and our papers, dynamo run, etc.}, we can confidently recommend continued use of this model. \begin{figure}[htbp]\begin{center}\leavevmode \centerline{% \begin{tabular}{c@{\hspace{.15in}}c} \includegraphics[width=8.95cm]{fig_comp2a} & \includegraphics[width=8.95cm]{fig_comp2b} \end{tabular}} \caption{ Spectra compensated by $k^{3/2}$ for the kinetic {\bf (a)} and magnetic {\bf (b)} energies \neu{averaged over $t \in[2.7,3.7]$;} labels are as in Fig. \ref{FIG:EVST} and the dashed vertical line indicates $k_\alpha=33$. \neu{Note} the $k^2$ tail at high wavenumber that is known to develop for under-resolved runs, a prediction stemming from statistical mechanics. } \label{FIG:COMP2} \end{center} \end{figure} \vier{We also computed a $512^3$ \lamhd-LES ($\alpha=1/85$) which retains more of the small scales than the $256^3$ \lamhd-LES while still yielding significant computational savings over the $1536^3$ DNS. We compare this with the result for $\alpha=1/18$ (chosen not as a LES but to stress the model) in Fig. \ref{FIG:SHEETS}. The structure of sheets observed in MHD dissipative structures is preserved in the \lamhda simulations, although current and vortex sheets become thicker in \lamhda as a result of the filter as $\alpha$ is increased. This is necessary to achieve reduced resolution computations. Note that these sheets are different in nature from the fat 'rigid bodies' observed in \lans, as the turbulent energy transfer to small scales is not quenched and there is no super-filter-scale bottleneck.} \begin{figure}[htbp]\begin{center}\leavevmode \centerline{% \begin{tabular}{c@{\hspace{.15in}}c} \includegraphics[width=8.95cm]{fig_current85} & \includegraphics[width=8.95cm]{fig_current18} \\ \includegraphics[width=8.95cm]{fig_current33} & \includegraphics[width=8.95cm]{fig_currentNS} \\ \end{tabular}} \caption{\vier{2D cross sections of square current, $j^2$, for $512^3$ \lamhd-LES ($\alpha=1/85$) {\bf (Upper Left)} and model-stress-case ($\alpha=1/18$) {\bf (Upper Right)}. MHD dissipative structures, sheets, are retained which become thicker as $\alpha$ is increased. {\bf (Lower Left)} $256^3$ \lamhd-LES ($\alpha=1/33$) and {\bf (Lower Right)} $256^3$ unresolved DNS. For the unresolved run, current sheets are somewhat smeared out by numerical noise.}} \label{FIG:SHEETS} \end{center} \end{figure} \section{Discussion} In this paper, we have tested the \lamhda model against high Reynolds number direct numerical simulations (up to \resp{}{Reynolds numbers of $\approx 9200$)} and in particular we have focused our attention on the dynamics of small scales near the $\alpha$ cut-off. We find that the small-scale spectrum presents no particular defect; specifically, we find that, unlike in the hydrodynamical case, the Lagrangian-averaged modeling for MHD exhibits, even at large Reynolds numbers, neither a positive-power-law spectrum nor any contamination of the super-filter-scale spectral properties. \resp{}{This difference between \lansa and \lamhda} is not due to the inclusion of a hyper-diffusive term in \lamhda that stems from the derivation of the model; rather, it stems from fundamental differences between hydrodynamics and MHD. Indeed, neither the (non-consistent) removal of hyperdiffusion from \lamhda nor the examination of scales much smaller than $\alpha$ gave any indication of problems similar to those caused by the zero-flux regions found in computations using \lans. These regions limited the computational gains of using \lansa as a LES in hydrodynamics to a factor of only $10$ in computational degrees of freedom or $30$ in computation time. \lamhda is not subject to the same limitations and, as we demonstrated, a gain of a factor of $200$ in the number of degrees of freedom, or a factor of $1300$ in computation time, obtains when comparing to the highest Reynolds number in turbulent MHD available today in a DNS. There are two obvious candidates to explain the lack of a (super-filter-scale) bottleneck effect in \lamhd: the enhanced (hyper-)diffusion in \lamhda compared with \lans, and \resp{}{physical differences between fluids and magneto-fluids, specifically, spectrally nonlocal transfer via Alfv\'en waves and its associated} breaking of the circulation conservation. The first candidate would eliminate the super-filter-scale bottleneck by removing energy from the system and precluding the formation of a secondary range below the filtering scale $\alpha$ (note that this term becomes of the same order as the ordinary diffusion when $l\sim\alpha$). Simulations of \lamhda performed without the hyper-diffusion term \resp{}{ruled out} this scenario, as no super-filter bottleneck was found. The second candidate is the \resp{}{presence of the Lorentz force in MHD (and \lamhd) which} breaks down the circulation conservation \resp{}{and provides the restoring force for Alfv\'en waves. Both properties were shown to be preserved by \lamhd. In Navier-Stokes, the development of helical filaments could quench local interactions \cite{MT92,T01} depleting the energy transfer and leading to the viscous bottleneck. However, in MHD, the conservation of the circulation ($d\Gamma/dt=0$ in the absence of dissipation) is broken by the Lorentz force, which modifies Kelvin's theorem (see Eq. (\ref{eq:kelvin})). The forcing term is associated with the Alfv\'en waves, and represents the removal of circulation (and of kinetic energy) that is transfered to the magnetic field. Note that in Fourier space, the term scales as $k E_M(k)$ and is dominant compared to the dissipation in the inertial range. This term precludes the formation of rigid bodies, giving as a result a larger net flux towards smaller scales and a resulting larger dissipation in MHD/\lamhd. This is illustrated in Fig. \ref{FIG:PDFS}. This sink of circulation may also be the cause of the lack of a viscous-scale bottleneck in MHD. In LANS it was shown \cite{PGHM+07a,PiGrHoMi+2008} that conservation of the circulation (except for viscosity) leads to the formation of rigid bodies that fill a substantial volume of the fluid, and that in turn substantially decrease the energy flux to small scales, reduce dissipation, and create the super-filter scale bottleneck.} In \lamhd, the destruction of sub-filter-scale rigid bodies by large scale magnetic field and shear \resp{}{results} as the presence of a magnetic field permits the development of long-range interactions in spectral space \cite{MAP05,AMP05a,Al2007a}. This can also explain why $\alpha-$models for other non-local equations, or for problems that do not preserve the circulation provide good SGS models. As an example, the use of \lansa in primitive equations ocean modeling gives satisfactory results, e.g. in its reproducing the Antarctic circumpolar current baroclinic instability that can be seen only at substantially higher resolutions when using direct numerical simulations \cite{HeHoPe+2008}. \resp{}{Energy is dissipated in MHD flows through two different processes. Viscosity is responsible for the dissipation of mechanical energy, while Ohmic losses are responsible for dissipation of magnetic energy. Mechanical and magnetic energy are not conserved separately, but rather coupled as illustrated by the existence of Alfv\'en waves, which correspond to oscillations of the magnetofluid with the velocity field parallel or anti-parallel to the magnetic field, and associated to the interchange of magnetic and kinetic energy. In MHD, it is believed that most of the total energy in the flow is finally dissipated (mediated by this interchange) through Ohmic losses, in a process that involves reconnection of magnetic field lines. This is supported by several simulations of MHD turbulence \cite{HaBrDo2003,Mi2007a} and is consistent with phenomenology. While in hydrodynamics small scales are permeated by a myriad of vortex filaments, in MHD the dominant dissipative structures are current sheets, where strong gradients of the magnetic field and their associated strong currents lead to rapid Ohmic dissipation. Sub-grid models attempt to replace the physical processes of small-scale dissipation by processes that mimic the non-linear transfer of energy to smaller scales (where energy is in reality dissipated, but now in scales that are not resolved by the model). In traditional LES, this is done with enhanced turbulent viscosities. Note that the eddy viscosity is not obtained from the linear dissipative term (the term that describes the actual physical process responsible for the dissipation) but from the non-linear terms in the equations (the terms that describe the coupling between fields at different scales). The final goal is not to capture the dissipation processes, but to be able to preserve (with computational gains) the large scale dynamics.} \resp{}{Lagrangian averaged models take a different (although related, see e.g., \cite{PGHM+06}) approach. Besides adding (in some cases, as in the case of MHD) an enhanced viscosity, the non-linear terms are modified at small scales. This modification changes the time-scale of the energy cascade, and as a result changes the scaling law of the energy spectrum $E(k)$ at sub-filter scales. This change leads to changes in the dissipation, as the dissipation is in the original equations proportional to $k^2E(k)$. The end result (an enhanced dissipation that is intended to mimic the transfer of energy to smaller scales in the unresolved scales) should be the same as in a traditional LES: gains in computing costs preserving as much information of the large scale flow as possible. As in the case of LES, the actual dissipation process is not as important as the fact that large-scale dynamics should be reproduced with minimal contamination \vier{by} the sub-grid model. We believe the results presented here (and in earlier work \cite{MMP05a,MMP05b,PGMP05,PGHM+06,Mi2006a}) show this is the case, and allow the use of the LAMHD equations as a subgrid model of MHD turbulence. However, considering the differences observed between LANS and LAMHD, we discuss the dissipation processes in LAMHD. Two mechanisms for dissipation can be identified in LAMHD: dissipation of mechanical energy through the viscosity, and dissipation of magnetic energy through (enhanced) Ohmic losses. From the equations, the total variation of energy goes as \cite{MMP05a}: $dE/dt = -\nu\left<\boldsymbol{\omega}\cdot\bar{\boldsymbol{\omega}}\right> -\eta\left<j^2\right>$ and as a result the mechanical energy dissipation scales as $k^2E_V(k)$ while the magnetic energy dissipation scales as $(1+ \alpha^2k^2)k^2E_M(k)$. The extra $k^2$ factor in the latter gives more dissipation than in the LANS case. This excess of magnetic dissipation in LAMHD mimics, as previously mentioned, the dominant contribution to dissipation by Ohmic losses in MHD. This hyperdiffusion is required in the sub-filter scales to accurately model the total energy dissipated at the unresolved scales. This was demonstrated by our experiments with a modified \lamhd, where we (non-consistently) removed the hyperdiffusive term and found the resulting model to fail as a LES.} \resp{}{Yet} another way to understand the differences between \lansa (for incompressible isotropic and homogeneous flows) and \lamhda is to consider the derivation of these models \cite{H02a} using the generalized Lagrangian-mean (GLM) formalism \cite{AM78}. This form of Lagrangian averaging describes wave, mean-flow interactions. For the case of weak turbulence, where the nonlinear transfer is dominated by waves, GLM requires in principle no closure. As a result, GLM gives an exact closed theory for the evolution of the wave activity. On the other hand, when there are no waves (as in incompressible Navier-Stokes) or when eddies dominate the transfer, a closure is required. One possible closure assumes that fast fluctuations are just advected by the mean flow (basically, Taylor's frozen-in hypothesis for the small scale turbulent fluctuations) and leads to the several "$\alpha$-models" that include \lansa and \lamhd. In this context, it is not surprising for subgrid models based on GLM to perform better in the presence of Alfv\'en waves (for\lamhd) or Rossby and gravity waves (for the Lagrangian-averaged primitive equations \cite{HeHoPe+2008}). The more relevant the waves are to the dynamics, and to the non-linear coupling of modes in the system, the less relevant is the hypothesis behind the closure. Furthermore, the $\alpha$-model equations can then be expected to be a better approximation to the problem at hand, that is, to be closer to an exact closure of the original system of equations. \vier{In the fluid case, the application of the ``Taylor'' closure that smaller-than-$\alpha$ scale fluctuations are swept along by the large-scale flow results in the fluctuations having greatly reduced interactions. This allows for a reduction in computational expense and leads to the super-filter-scale bottleneck by quenching spectrally non-local interactions. In the \lamhda case, the small-scale $\vec{z}^+$ ($\vec{z}^-$) fluctuations are swept along by the large-scale $\bar{\vec{z}}^-$ ($\bar{\vec{z}}^+$) flow. Small-scale fluctuations advected by two different fields may now collide and nonlinearly interact. The second part of the model is the preferential hyperdiffusion of Alfv\'en waves with wavelengths shorter than$\alpha$. This damps rather than quenches nonlinear interactions among the small scales. This more gentle suppression of the transfer of energy to smaller scales reduces the numerical resolution requirements without forming a bottleneck.} It was noted in \cite{MMP05b} when assessing the properties of \lamhda in the dynamo context that the overall temporal evolution was satisfactory, e.g. with a correct growth rate, although the growth of the magnetic seed field started slightly earlier in the \lamhda run than in the DNS. One can speculate as to whether this delay is linked to the super-bottleneck effect of \lansa (which prevails when the magnetic field is negligible compared to the velocity, the two modeling approaches, \lamhda and \lans, being dynamically consistent). This point is left for future work; one could determine as well at what ratio of magnetic to kinetic energy the overshooting of spectra in \lansa disappears for \lamhd. Also deserving of a separate study is to investigate the behavior of \lamhda when anisotropies that appear at small scales \cite{MiPo2007a} are present; this would be essential when a uniform magnetic field is imposed to the overall flow. The evaluation of the behavior of the model when computing spectra in the perpendicular and parallel directions (with respect to a quasi-uniform magnetic field, computed by locally averaging the field in a sphere of radius comparable to the integral scale) remains to be done but is somewhat time consuming. An analysis of the structures that develop in the highly turbulent \lamhda flow studied in the preceding section is also left for future work; of particular interest is the occurrence of Kelvin-Helmholtz like roll-up of current sheets as observed at high resolution \cite{MiPo2007a}; however, the choice of the parameter $\alpha$ in the present paper was made on the basis of questioning the existence or lack thereof of a rigid-body high-wavenumber $k^{+1}$ spectrum and, thus, was not optimized for the study of the inertial range properties of the flow for which a much smaller value of the length $\alpha$ could be used. Finally, how far resolution can be reduced when using \lamhda as a LES for various statistics of interest will also require further detailed study. The present study shows that, to reproduce the super-filter-scale energy spectrum in three dimensions, gains by a factor of 1300 in computing time can be achieved. The need to reproduce higher order statistics can decrease these gains. As an example, in two-dimensional MHD, it was shown that gains when using \lamhda as a subgrid model depend for high order moments on the order that one wants to see to be accurately reproduced \cite{PGHM+06}. %The lack of strong contamination of super-filter-scales in \lamhda when compared with LANS can be associated with the interaction between the magnetic and the velocity field, which breaks the formation of rigid bodies in the flow. Associated with this process, the Lorentz force in \lamhda breaks the conservation of the circulation which is preserved in LANS. We can expect similar behavior of the Lagrangian-averaged models for other flows with mechanisms that can change the circulation, such as rotating or baroclinic flows. \begin{acknowledgments} Computer time was provided by GWDG, NCAR, and the National Science Foundation Terascale Computing System at the Pittsburgh Supercomputing Center. The NSF Grant No. CMG-0327888 at NCAR supported this work in part and is gratefully acknowledged. \add{}{PDM is a member of the Carrera del Investigador Cient\'{\i}fico of CONICET.} \resp{}{The anonymous referees are gratefully acknowledged for improving the clarity of the discussion of our results.} \end{acknowledgments} %\bibliographystyle{plain} %\bibliography{phd} \begin{thebibliography}{10} \bibitem{MK00} C.~{Meneveau} and J.~{Katz}. \newblock{Scale-Invariance and Turbulence Models for Large-Eddy Simulation}. \newblock{\em Annual Review of Fluid Mechanics}, 32:1--32, 2000. \bibitem{PFL76} A.~{Pouquet}, U.~{Frisch}, and J.~{Leorat}. \newblock{Strong MHD helical turbulence and the nonlinear dynamo effect}. \newblock{\em Journal of Fluid Mechanics}, 77:321--354, 1976. \bibitem{Y87} A.~{Yoshizawa}. \newblock{Subgrid modeling for magnetohydrodynamic turbulent shear flows}. \newblock{\em Physics of Fluids}, 30:1089--1095, 1987. \bibitem{ChLe1981} J.-P. {Chollet} and M.~{Lesieur}. \newblock{Parameterization of Small Scales of Three-Dimensional Isotropic Turbulence Utilizing Spectral Closures.} \newblock{\em Journal of Atmospheric Sciences}, 38:2747--2757, 1981. \bibitem{BaPoPo+2008} J.~{Baerenzung}, H.~{Politano}, Y.~{Ponty}, and A.~{Pouquet}. \newblock{Spectral modeling of magnetohydrodynamic turbulent flows}. \newblock{\em \pre}, 78(2):026310--+, August 2008. \bibitem{AMP05a} A.~{Alexakis}, P.~D. {Mininni}, and A.~{Pouquet}. \newblock{Shell-to-shell energy transfer in magnetohydrodynamics. I. Steady state turbulence}. \newblock{\em \pre}, 72(4):046301--+, 2005. \bibitem{MAP05} P.~{Mininni}, A.~{Alexakis}, and A.~{Pouquet}. \newblock{Shell-to-shell energy transfer in magnetohydrodynamics. II. Kinematic dynamo}. \newblock{\em \pre}, 72(4):046302--+, 2005. \bibitem{ZSG02} Y.~{Zhou}, O.~{Schilling}, and S.~{Ghosh}. \newblock{Subgrid scale and backscatter model for magnetohydrodynamic turbulence based on closure theory: Theoretical formulation}. \newblock{\em \pre}, 66(2):026309--+, 2002. \bibitem{GaNaNe+2000} S.~{Galtier}, S.~V. {Nazarenko}, A.~C. {Newell}, and A.~{Pouquet}. \newblock{A weak turbulence theory for incompressible magnetohydrodynamics}. \newblock{\em Journal of Plasma Physics}, 63:447--488, 2000. \bibitem{GoSr1995} P.~{Goldreich} and S.~{Sridhar}. \newblock{Toward a theory of interstellar turbulence. 2: Strong alfvenic turbulence}. \newblock{\em \apj}, 438:763--775, 1995. \bibitem{I64} P.~S. {Iroshnikov}. \newblock{Turbulence of a Conducting Fluid in a Strong Magnetic Field}. \newblock{\em Soviet Astronomy}, 7:566--+, 1964. \bibitem{K65} R.~H. {Kraichnan}. \newblock{Inertial-range spectrum of hydromagnetic turbulence}. \newblock{\em Physics of Fluids}, 8:1385--1387, 1965. \bibitem{MaCaBo2008} J.~{Mason}, F.~{Cattaneo}, and S.~{Boldyrev}. \newblock{Numerical measurements of the spectrum in magnetohydrodynamic turbulence}. \newblock{\em \pre}, 77(3):036403--+, 2008. \bibitem{MiPo2007a} P.~D. {Mininni} and A.~{Pouquet}. \newblock{Energy Spectra Stemming from Interactions of Alfv{\'e}n Waves and Turbulent Eddies}. \newblock{\em Physical Review Letters}, 99(25):254502--+, 2007. \bibitem{PoMiMo+2008} Annick {Pouquet}, Pablo {Mininni}, David {Montgomery}, and Alexandros {Alexakis}. \newblock{Dynamics of the Small Scales in Magnetohydrodynamic Turbulence }. \newblock In Yukio {Kaneda}, editor, {\em Proceedings of the IUTAM Symposium on Computational Physics and New Perspectives in Turbulence, Nagoya University, Nagoya, Japan, September, 11-14, 2006}, volume~4, pages 305--312. Springer-Verlag, 2008. \bibitem{AMK+01} O.~{Agullo}, W.-C. {M{\"u}ller}, B.~{Knaepen}, and D.~{Carati}. \newblock{Large eddy simulation of decaying magnetohydrodynamic turbulence with dynamic subgrid-modeling}. \newblock{\em Physics of Plasmas}, 8:3502--3505, 2001. \bibitem{TFS94} M.~L. {Theobald}, P.~A. {Fox}, and S.~{Sofia}. \newblock{A subgrid-scale resistivity for magnetohydrodynamics}. \newblock{\em Physics of Plasmas}, 1:3016--3032, 1994. \bibitem{HB06} N.~E.~L. {Haugen} and A.~{Brandenburg}. \newblock{Hydrodynamic and hydromagnetic energy spectra from large eddy simulations}. \newblock{\em Physics of Fluids}, 18:5106--+, 2006. \bibitem{MC02} W.-C. {M{\"u}ller} and D.~{Carati}. \newblock{Dynamic gradient-diffusion subgrid models for incompressible magnetohydrodynamic turbulence}. \newblock{\em Physics of Plasmas}, 9:824--834, 2002. \bibitem{LS91} D.~W. {Longcope} and R.~N. {Sudan}. \newblock{Renormalization group analysis of reduced magnetohydrodynamics with application to subgrid modeling}. \newblock{\em Physics of Fluids B}, 3:1945--1962, 1991. \bibitem{KM04} B.~{Knaepen} and P.~{Moin}. \newblock{Large-eddy simulation of conductive flows at low magnetic Reynolds number}. \newblock{\em Physics of Fluids}, 16:1255--+, 2004. \bibitem{PMM+05} Y.~{Ponty}, P.~D. {Mininni}, D.~C. {Montgomery}, J.-F. {Pinton}, H.~{Politano}, and A.~{Pouquet}. \newblock{Numerical Study of Dynamo Action at Low Magnetic Prandtl Numbers}. \newblock{\em Physical Review Letters}, 94(16):164502--+, 2005. \bibitem{PPP04} Y.~{Ponty}, H.~{Politano}, and J.-F. {Pinton}. \newblock{Simulation of Induction at Low Magnetic Prandtl Number}. \newblock{\em Physical Review Letters}, 92(14):144503--+, 2004. \bibitem{H02b} D.~D. {Holm}. \newblock{Averaged Lagrangians and the mean effects of fluctuations in ideal fluid dynamics}. \newblock{\em Physica D Nonlinear Phenomena}, 170:253--286, 2002. \bibitem{H02a} D.~D. {Holm}. \newblock{Lagrangian averages, averaged Lagrangians, and the mean effects of fluctuations in fluid dynamics}. \newblock{\em Chaos}, 12:518--530, 2002. \bibitem{MP02} D.~C. {Montgomery} and A.~{Pouquet}. \newblock{An alternative interpretation for the Holm ``alpha model''}. \newblock{\em Physics of Fluids}, 14(9):3365--3366, 2002. \bibitem{MMP05a} P.~D. {Mininni}, D.~C. {Montgomery}, and A.~G. {Pouquet}. \newblock{A numerical study of the alpha model for two-dimensional magnetohydrodynamic turbulent flows}. \newblock{\em Physics of Fluids}, 17:5112--+, 2005. \bibitem{PGMP05} J.~Pietarila Graham, P.~D. {Mininni}, and A.~{Pouquet}. \newblock{Cancellation exponent and multifractal structure in two-dimensional magnetohydrodynamics: Direct numerical simulations and Lagrangian averaged modeling}. \newblock{\em \pre}, 72(4):045301(R)--+, 2005. \bibitem{PGHM+06} J.~Pietarila Graham, D.~D. {Holm}, P.~{Mininni}, and A.~{Pouquet}. \newblock{Inertial range scaling, K\'arm\'an-Howarth theorem, and intermittency for forced and decaying Lagrangian averaged magnetohydrodynamic equations in two dimensions}. \newblock{\em Physics of Fluids}, 18:045106, 2006. \bibitem{MMP05b} P.~D. {Mininni}, D.~C. {Montgomery}, and A.~{Pouquet}. \newblock{Numerical solutions of the three-dimensional magnetohydrodynamic {$\alpha$} model}. \newblock{\em \pre}, 71(4):046304--+, 2005. \bibitem{Mi2006a} P.~D. {Mininni}. \newblock{Turbulent magnetic dynamo excitation at low magnetic Prandtl number}. \newblock{\em Physics of Plasmas}, 13:056502--+, 2006. \bibitem{CHO+05} Alexey Cheskidov, Darryl~D. Holm, Eric Olson, and Edriss~S. Titi. \newblock{On a Leray$-\alpha$ model of turbulence}. \newblock{\em Proceedings of the Royal Society of London}, A461:629--649, 2005. \bibitem{PGHM+07a} J.~Pietarila Graham, D.~{Holm}, P.~{Mininni}, and A.~{Pouquet}. \newblock{Highly turbulent solutions of the Lagrangian-averaged Navier-Stokes alpha model and their large-eddy-simulation potential}. \newblock{\em \pre}, 76:056310--+, 2007. \bibitem{MiPoSu2008} P.D. {Mininni}, A.~{Pouquet}, and P.~{Sullivan}. \newblock{Two examples from geophysical and astrophysical turbulence on modeling disparate scale interactions}. \newblock In Roger {Temam} and Joe {Tribbia}, editors, {\em Summer school on mathematics in geophysics}. Springer-Verlag, 2008. \newblock to appear. \bibitem{GMD05b} D.~O. {G{\'o}mez}, P.~D. {Mininni}, and P.~{Dmitruk}. \newblock{MHD simulations and astrophysical applications}. \newblock{\em Advances in Space Research}, 35:899--907, 2005. \bibitem{GMD05} D.~O. {G{\'o}mez}, P.~D. {Mininni}, and P.~{Dmitruk}. \newblock{Parallel Simulations in Turbulent MHD}. \newblock{\em Physica Scripta Volume T}, 116:123--127, 2005. \bibitem{MaPoMi+2008} W.~H. {Matthaeus}, A.~{Pouquet}, P.~D. {Mininni}, P.~{Dmitruk}, and B.~{Breech}. \newblock{Rapid Alignment of Velocity and Magnetic Field in Magnetohydrodynamic Turbulence}. \newblock{\em Physical Review Letters}, 100(8):085003--+, 2008. \bibitem{MiPoMo2006} P.~D. {Mininni}, A.~G. {Pouquet}, and D.~C. {Montgomery}. \newblock{Small-Scale Structures in Three-Dimensional Magnetohydrodynamic Turbulence}. \newblock{\em Physical Review Letters}, 97(24):244503--+, 2006. \bibitem{FHT01} C.~{Foias}, D.~D. {Holm}, and E.~S. {Titi}. \newblock{The Navier-Stokes-alpha model of fluid turbulence}. \newblock{\em Physica D Nonlinear Phenomena}, 152-153:505--519, May 2001. \bibitem{LKT+07} E.~{Lunasin}, S.~{Kurien}, M.~{Taylor}, and E.~{Titi}. \newblock{A study of the Navier-Stokes-alpha model for two-dimensional turbulence}. \newblock{\em ArXiv Physics e-prints}, 2007. \bibitem{PP98b} H\'el\`ene{Politano} and Annick {Pouquet}. \newblock{Dynamical length scales for turbulent magnetized flows}. \newblock{\em Geophysical Research Letters}, 25(3):273--276, 1998. \bibitem{K67} R.~H. {Kraichnan}. \newblock{Inertial Ranges in Two-Dimensional Turbulence}. \newblock{\em Physics of Fluids}, 10:1417--1423, 1967. \bibitem{CHT05} C.~{Cao}, D.~D. {Holm}, and E.~S. {Titi}. \newblock{On the Clark {$\alpha$} model of turbulence: global regularity and long-time dynamics}. \newblock{\em Journal of Turbulence}, 6:19--+, 2005. \bibitem{DmGoMa2003} P.~{Dmitruk}, D.~O. {G{\'o}mez}, and W.~H. {Matthaeus}. \newblock{Energy spectrum of turbulent fluctuations in boundary driven reduced magnetohydrodynamics}. \newblock{\em Physics of Plasmas}, 10:3584--3591, 2003. \bibitem{Al2007a} A.~{Alexakis}. \newblock{Nonlocal Phenomenology for Anisotropic Magnetohydrodynamic Turbulence}. \newblock{\em \apjl}, 667:L93--L96, 2007. \bibitem{HSL+82} J.~R. {Herring}, D.~{Schertzer}, M.~{Lesieur}, G.~R. {Newman}, J.~P. {Chollet}, and M.~{Larcheveque}. \newblock{A comparative assessment of spectral closures as applied to passive scalar diffusion}. \newblock{\em Journal of Fluid Mechanics}, 124:411--437, 1982. \bibitem{LMG95} D.~{Lohse} and A.~{M{\"u}ller-Groeling}. \newblock{Bottleneck Effects in Turbulence: Scaling Phenomena in r versus p Space}. \newblock{\em Physical Review Letters}, 74:1747--1750, 1995. \bibitem{MCD+97} D.~O. {Mart{\'{\i}}nez}, S.~{Chen}, G.~D. {Doolen}, R.~H. {Kraichnan}, L.-P. {Wang}, and Y.~{Zhou}. \newblock{Energy spectrum in the dissipation range of fluid turbulence}. \newblock{\em Journal of Plasma Physics}, 57:195--201, 1997. \bibitem{MAP06} P.~D. {Mininni}, A.~{Alexakis}, and A.~{Pouquet}. \newblock{Large-scale flow effects, energy transfer, and self-similarity on turbulence}. \newblock{\em \pre}, 74(1):016303--+, 2006. \bibitem{MT92} H.~K. {Moffatt} and A.~{Tsinober}. \newblock{Helicity in laminar and turbulent flow}. \newblock{\em Annual Review of Fluid Mechanics}, 24:281--312, 1992. \bibitem{T01} Arkady {Tsinober}. \newblock{\em{An Informal Introduction to Turbulence}}. \newblock Kluwer Academic Publishers, Dordrecht, 2001. \bibitem{CHM+99} S.~{Chen}, D.~D. {Holm}, L.~G. {Margolin}, and R.~{Zhang}. \newblock{Direct numerical simulations of the Navier-Stokes alpha model}. \newblock{\em Physica D Nonlinear Phenomena}, 133:66--83, 1999. \bibitem{PiGrHoMi+2008} J.~Pietarila Graham, D.~D. {Holm}, P.~D. {Mininni}, and A.~{Pouquet}. \newblock{Three regularization models of the Navier-Stokes equations}. \newblock{\em Physics of Fluids}, 20(3):035107--+, 2008. \bibitem{FrPoLe+1975} U.~{Frisch}, A.~{Pouquet}, J.~{L\'eorat}, and A.~{Mazure}. \newblock{Possibility of an inverse cascade of magnetic helicity in magnetohydrodynamic turbulence}. \newblock{\em Journal of Fluid Mechanics}, 68:769--778, 1975. \bibitem{HeHoPe+2008} M.~W. {Hecht}, D.~D. {Holm}, M.~R. {Petersen}, and B.~A. {Wingate}. \newblock{Implementation of the LANS-{$\alpha$} turbulence model in a primitive equation ocean model}. \newblock{\em Journal of Computational Physics}, 227:5691--5716, 2008. \bibitem{HaBrDo2003} N.~E.~L. {Haugen}, A.~{Brandenburg}, and W.~{Dobler}. \newblock{Is Nonhelical Hydromagnetic Turbulence Peaked at Small Scales?} \newblock{\em \apjl}, 597:L141--L144, 2003. \bibitem{Mi2007a} P.~D. {Mininni}. \newblock{Inverse cascades and {$\alpha$} effect at a low magnetic Prandtl number}. \newblock{\em \pre}, 76(2):026316--+, 2007. \bibitem{AM78} D.~G. {Andrews} and M.~E. {McIntyre}. \newblock{An exact theory of nonlinear waves on a Lagrangian-mean flow}. \newblock{\em J. Fluid Mech.}, 89:609--646, 1978. \end{thebibliography} \end{document} % LocalWords: Leray LANS DNS subgrid pseudospectral streamwise eps pdfs Yukawa % LocalWords: discretization Howarth Kraichnan advected Subdominance Eq IK % LocalWords: ABC Eqs Lyapunov MHD SGS NCAR Terascale Supercomputing CMG LAMHD % LocalWords: Sonnensystemforschung Katlenburg Lindau Departamento Facultad % LocalWords: Ciencias Exactas Naturales Universidad Buenos Ciudad diffusivity % LocalWords: Universitaria hyperdiffusivity hyperviscosity hyperdiffusion emf % LocalWords: Beltrami enstrophy BaPoPo Prandtl electromotive barotropic PDFs % LocalWords: Iroshnikov hydrodynamical baroclinic GLM Rossby }
\caption{\label{fig:figure2} Kinetic (red curves) and magnetic (black curves) energies (see text for detail): $E_{\mbox{\tiny{kin}}}^{\mbox{\tiny{Sh}}}$ (\textcolor{red}{\full}), $E_{\mbox{\tiny{kin}}}^{\theta}$ (\textcolor{red}{\chain}), $E_{\mbox{\tiny{kin}}}^{\mbox{\tiny{N}}}$ (\textcolor{red}{\dashed}), and $E_{\mbox{\tiny{kin}}}^{\mbox{\tiny{S}}}$ (\textcolor{red}{\dotted}). $E_{\mbox{\tiny{mag}}}^{\mbox{\tiny{Sh}}}$ (\full), $E_{\mbox{\tiny{mag}}}^{\mbox{\tiny{N}}}$ (\dashed), and $E_{\mbox{\tiny{mag}}}^{\mbox{\tiny{S}}}$ (\dotted). We also plot $E_{\mbox{\tiny{mag}}}^{\mbox{\tiny{Sh}}}$ for an unperturbed model (\chain). ($1$~B=$10^{51}$~erg.)}
\caption{\label{fig:figure3} Magnetic energy inside the shocked cavity, $E_{\mbox{\tiny{mag}}}^{\mbox{\tiny{Sh}}}$, for models with various values of initial magnetic field strength $B_{0}$ at the PNS surface: $10^{8}$~G (\dashed), $10^{9}$~G (\dotted), $10^{10}$~G (\full), and $10^{11}$~G (\chain), respectively. }
\caption{\SiII\in\synow\for increasing excitation temperatures. The trough of the blue feature due to\SiII\\SiIIblue\increases relative to the trough of the\SiII\\SiIIred\feature as temperature increases, as predicted by Eq.~\ref{eq:SiTauRedBlue}. The \SiII\reference line is\SiII\\SiIIred\, and only the excitation temperature is varied, which explains the$\sim 6100$~\AA\trough stability. (Note that the unrealistic excitation temperature of 40000K leads to a very strong feature around 5400~\AA\that is not observed in SNe~Ia spectra.)}
\caption{Temperature for day 20 models in the $8000$-$16000$~\kms\region where\SiII\ions are found, as a function of the model parameter$M_{bol}$. The horizontal blue line marks the 7000 K limit at which \TiII\lines have been suggested to dominate over\SiII\\SiIIblue\\citep{garn99by04}.}
\caption{The na\"ive blackbody fit would peak where the full synthetic spectrum does, that is around 5000~\ang, far to the red of the physical blackbody. Top Panel: Full \phx\day 20 spectrum in black, continuum-only spectrum forming below 2000~\kms\in dotted blue,$13685K$ blackbody in plain blue. Bottom Panel: Continuum-only day 20 spectrum in dotted blue, and $13685K$ blackbody in plain blue.}
\caption{\FeII\and\FeIII\single ion spectra for our faintest simulated spectra, 20 days after explosion. The top panel is\FeIII\in red and\FeII\in pink. The bottom panel is\FeII\+\FeIII\two-ion spectra in red. In both cases the black spectrum is the full simulated spectrum and the blue spectrum is the continuum-only spectrum.}
\caption{\FeII\(top panel) and \FeIII\(bottom panel) day 20 single ion spectra as a function of $M_{bol}$. Each spectrum is rescaled to the bolometric flux of the corresponding full synthetic spectrum. Since \FeIII\forms deeper it is less sensitive to a change in the bolometric luminosity.\FeII\lines are more efficient at transferring flux from blue to red.}
\caption{Single ion spectra for our faintest luminosity 20 days after explosion. The top panel shows spectra without \SII, the bottom panel includes \SII. They show that \SiII\lines alone can not account for the\RSi\wavelength region shape.\FeII\and\FeIII\contribute with their weak lines to the ``pseudo-continuum''. The blue edge ($<$5600~\AA) of the \RSi\region is dominated by\SII\lines.}
\caption{The full synthetic day 20 spectrum (solid line) is compared to the day 20 four-ion spectra including \FeIII, \FeII, \SiII, and \SII\(dotted line). The top panel shows $M_{bol}= -18.0$, the bottom panel $M_{bol}= -19.0$. The pink vertical line indicates \FeIII\$5128$~\AA\feature, while the green vertical region shows\FeII\strong feature$5170-5291$~\AA\that dominates at lower luminosities. The blue vertical zones are zones where\RSiS\is defined.}
\caption{The three order parameters used to study the square to rectangle phase transformation in crystals of cubic symmetry; {\color{red}red}=positive, {\color{blue}blue}=negative value.}
\caption{Stress field around one edge dislocation with its Burgers vector parallel to the $x_1$ axis ({\color{blue}blue}=negative, {\color{red}red}=positive values). Because periodic boundary conditions are used in this calculation, the stress field shown here corresponds to one dislocation from a periodic array of dislocations of the same kind.}
\caption{The dependence of the wavelength $\lambda$ on the rate $\gamma$. Analytical predictions from the CGLE~(\ref{CGLE}) (divided by a factor $1.55$), red line, are compared to numerical results (\textcolor{blue}{\scriptsize $\Box$}). For $\gamma<\gamma_E\approx 2$ spirals are convectively instable. When approaching the bifurcation point, $\gamma\to 0$, computation of the correlation length ({\scriptsize $\odot$}) shows that the spatial structures are no longer determined by the (diverging) wavelength, but reach a constant size $l_c$, see text. \label{gamma_lambda}}
\caption{Regimes of stable, unstable, and neutrally stable biodiversity. They are revealed by computing the extinction probability for $\gamma=0$ and $\rho=1$, depending on $D$, after a waiting time $t=N$, for increasing system size $N$. We show curves for $N=20\times 20$ (\textcolor{green}{\scriptsize $\Box$}), $N=25\times 25$ (\textcolor{red}{\large $\circ$}), $N=40\times 40$ (\textcolor{blue}{\scriptsize $\bigtriangleup$}), and $N=100\times 100$ ({\scriptsize $\diamondsuit$}). The transitions occur at $D_c^{(1)}\approx 3.5\times 10^{-3}$ and $D_c^{(2)}\approx 0.2$. \label{Pext}}
\caption{\label{36}Finite kagom\'e lattice of$N=36$ sites. The solid lines represent the NN bonds $J_1$ and the dashed lines the diagonal bonds $J_d$. The coloured circles indicate the spin orientations of the twelve-sublattice classical GS relevant for $J_d>0$. Along the chains formed by diagonal bonds there is an antiparallel (N\'{e}el) spin alignment (e.g. the spins at the light and dark green circles). On each triangle formed by NN bonds $J_1$ (e.g. spins on sites 21, 22, 23) there is a $120^{\circ}$ spin arrangement. In addition, two antiparallel (N\'{e}el) spin-sublattices are perpendicular to one other group of two N\'{e}el-like sublattices, i.e. \textcolor{blue}{ \huge $ \bullet$ }{\large $\perp \;$}{\color{brown!70!black}\huge $\bullet$ } \quad ; \quad{\color{green!70!blue} \huge $ \bullet$ }{\large $\perp \;$}{\color{magenta!70!blue} \huge $\bullet$ } \quad ; \quad \textcolor{cyan!70!blue}{ \huge $ \bullet$ }{\large $\perp \;$} \textcolor{red!90!black}{ \huge $\bullet$ }.}
\caption{(Color online) (a) Average level $\langle l \rangle$ of the random walker versus the time $n$. The heptagonal lattice of the size $L$ is first built, and we perform random walks as in Fig.~\ref{fig:walker}(b). We see $\langle l \rangle \propto n$ for small $n$, while $\langle l \rangle$ eventually saturates as the walker approaches the external boundary. (b) Average distance $\left< d_h \right>$ from the origin versus the time $n$, obtained from the method in Fig.~\ref{fig:walker}(c). The deviation from the linear behavior is not due to the finiteness of the lattice as in (a), but due to accumulated numerical errors. All the averages were taken from $10^3$ independent realizations.} \label{fig:dist} \end{figure} We first study numerically the random walk problem in our heptagonal lattice structure, and the results are compared with numerical solutions of the diffusion equation for the hyperbolic geometry. A possible way of simulating a random walker on the heptagonal lattice is to identify the connection structure first and then let a particle move along the edges. In Fig.~\ref{fig:dist}(a), the position of the random walker measured by the level $l$ is shown as a function of the time step $n$ for the heptagonal lattices of sizes $L=7$, 10, and 14. It is clearly shown that the random walker drifts away from the starting position $l=1$ linearly in time. As the lattice becomes larger (as $L$ is increased), the linear diffusion regime becomes extended, indicating $\langle l \rangle \propto n$ in the thermodynamic limit. One drawback of this approach is that the number of points increases exponentially as we add concentric layers. The memory constraint restricts the distance from the origin, and therefore the results severely suffer from finiteness of the model system under consideration. %Indeed, recalling that the Brownian motion is transient %there~\cite{kendall}, this algorithm is absolutely inefficient in that most %of the lattice points are unnecessary. A better alternative is obtained from the hyperbolic tessellation. Since the heptagonal lattice is dual to $\{3,7\}$, the transition between the neighboring points coincides with reflecting a hyperbolic triangle [see Fig.~\ref{fig:walker}(c)]. Since the Poincar{\'e} disk can be identified with a unit disk on the complex plane, we start from a regular triangle $(z_1, z_2, z_3)$ where $z_i$'s are complex numbers. Let one of the vertices, say $z_1$, be located at the origin. Setting the interior angle at that point to be $2\pi/7$, the triangle is regular only when $|z_1-z_2| = |z_1-z_3| \approx 0.496 97$, as this surface has its own intrinsic length scale, i.e., the curvature~\cite{green}. If a triangle $(z_i, z_j, z_k)$ is reflected around the edge along $z_i$ and $z_j$, we get a new triangle $(z_i, z_j, z_k')$ with $z_k' = w + \xi^2/(\overline{z_k}-\overline{w})$, where $w=(|z_i|^2z_j - z_i|z_j|^2-z_i + z_j)/(\overline{z_i}z_j - z_i\overline{z_j})$, $\xi=|w-z_i|=|w-z_j|$, and $\overline{z}$ is the complex conjugate of $z$. The center of a triangle, $z$, has the hyperbolic distance from the origin by $d_h(z) = \ln[(1+|z|)/(1-|z|)]$. Note that $d_h$ diverges to infinity as $|z| \rightarrow 1$, which is consistent with the definition of the Poincar{\'e} disk. %The random walk on the heptagonal lattice is therefore equivalent to %choosing one vertex randomly at each time step. The result is depicted in Fig.~\ref{fig:dist}(b), which clearly shows that the distance is linearly proportional to time. The numerical inaccuracy becomes larger as we repeat the reflection of triangles. In order to confirm the origin of the deviation from the linear behavior $\langle d_h \rangle \sim n$ at large $n$, we intentionally assign numerical precision and observe how $\langle d_h \rangle$ depends on it. In Fig.~\ref{fig:dist}(b), it is seen that as we use better numerical precision, the linear regime becomes more extended, indicating that the deviation from the linear behavior originates from the simple artifact of the numerical accuracy. Beyond the deviation point of each curve in Fig.~\ref{fig:dist}, the numerical values are not trustworthy and thus the decrease of $\langle d_h \rangle$ should not be taken as real. For the random walk on a 2D flat surface, the walker has zero probability of escape by P{\'o}lya's theorem which states that the walk becomes transient for dimensions larger than two~\cite{rudnick}. If we denote the probability density of finding a particle at a position ${\bf r}$ after time $t$ as $\phi({\bf r},t)$, the diffusion process is described by ${\partial \phi}/{\partial t} = \nabla^2 \phi$. %with $i = 1,2$ for two dimensions and the repeated index implying %the sum over the index (Einstein summation convention). Solving this equation by Fourier transformation, one can see that the expected displacement from the origin is proportional to the square root of time in the 2D flat surface. The Laplace operator $\nabla^2$ is changed to the Laplace-Beltrami operator for the hyperbolic metric~\cite{young}, %For the hyperbolic metric tensor, %\[ g = \{g_{ij}\} = \left[ \begin{array}{cc} 1 & 0 \\ 0 & \sinh^2 r %\end{array} \right],\] %the Laplace-Beltrami operator is given by~\cite{young} %\begin{eqnarray*} %\triangle &=& \frac{1}{\sqrt{|g|}} \partial_i \left( \sqrt{|g|}g^{ij} %\partial_j \right)\\ %%&=& \frac{1}{\sinh r} \frac{\partial}{\partial r} \left( \sinh r %\frac{\partial }{\partial r} \right) + \frac{1}{\sinh^2 r} \frac{\partial^2 %}{\partial \theta^2}. %\end{eqnarray*} \begin{equation} \triangle = \frac{1}{\sinh~r} \frac{\partial}{\partial r} \left( \sinh~r \frac{\partial }{\partial r} \right) + \frac{1}{\sinh^2~r} \frac{\partial^2 }{\partial \theta^2}, \end{equation} and we get the solution of the hyperbolic diffusion equation, $\partial \phi/\partial t = \triangle \phi$~\cite{monthus,banica}: \begin{equation} \phi(r,\theta;t) \propto \frac{e^{-t/4}}{t^{3/2}} \int\int dr' d\theta' \phi(r', \theta';0) I(t,\rho) \sinh~r' \label{eq:sol} \end{equation} with the distance $\rho$ between two positions $(r,\theta)$ and $(r', \theta')$, and $I(t,\rho) \equiv \int_{\rho}^{\infty} (s e^{-s^2/4t}/\sqrt{\cosh~s - \cosh~\rho})ds$. %The solution is not written in a closed form, since the plain wave on the %hyperbolic space becomes a lot different from that on the flat %plane~\cite{helgason}. The numerical integration is then performed to get the result presented in Fig.~\ref{fig:dif}, which again shows the linear diffusion $\langle r \rangle \propto t$. We thus conclude that our different approaches, i.e., discrete random walks by triangle reflections on the Poincar{\'e} disk and the solution of the hyperbolic diffusion equation, unanimously confirm that the distance from the origin increases linearly in time. In comparison to $\langle d \rangle \sim \sqrt{t}$ for the 2D flat surface, the diffusion occurs much faster in the negatively curved surface. \begin{figure} \includegraphics[width=0.46\textwidth]{fig3.eps} \caption{(Color online) The solution $\phi(r,t)$ of the hyperbolic diffusion equation. Inset: Average distance from the origin $\left<r \right>= \int r \phi(r) \sinh~r ~dr$ increases linearly in time.} \label{fig:dif} \end{figure} %\subsection{Quantum case}\label{subsec:quantum} We next study the diffusion of a tight-binding quantum particle in the heptagonal lattice. The quantum diffusion phenomena, and the existence and nature of the localization transition have been studied on regular lattice structures for a long time. Motivated by the intensive research interest of complex networks, recent years have observed the beginning of research on quantum mechanical systems put on the structure of complex networks~\cite{bjkim,mulken,cpzhu-giraud}. We assume that the wave function is localized on the first layer containing seven points [see Fig.~\ref{fig:walker}(a)]. The time evolution of the quantum particle is governed by the Schr{\"o}dinger equation $ i (\partial / \partial t) |\psi\rangle = {\cal H} |\psi\rangle $ ($\hbar \equiv 1$). By using the perfectly localized states as basis kets, we apply the tight-binding approximation that ${\cal H}_{ij} = 1 (0)$ if two points $i$ and $j$ are connected (not connected). The Schr{\"o}dinger equation is then numerically integrated by the fourth order Runge-Kutta method. The spreading of the wave packet is measured by the average layer, $\left<l \right>$, depicted in Fig.~\ref{fig:level}. Again, one can see clearly that the particle diffuses linearly in time as in the classical diffusion. The saturation comes from the finiteness of our lattice. It is to be noted that in the negatively curved heptagonal lattice, both the quantum and the classical diffusion exhibit the linear diffusion. \begin{figure} \includegraphics[width=0.46\textwidth]{fig4.eps} \caption{(Color online) Diffusion of the tight-binding quantum particle on the heptagonal lattice. As the time $t$ evolves, the quantum particle diffuses toward the upper layers. Note that the linear regime extends as the size becomes larger.} \label{fig:level} \end{figure} We then take an alternative approach to investigate the quantum diffusion problem, through the use of the mapping of the above tight-binding Hamiltonian to the one for a free particle. We note that the shift of energy by a constant amount does not change any measurable quantity and makes the transformation ${\cal H} \rightarrow {\cal H} - k I$ with a degree $k$ (the number of neighbors, e.g., $k = 3$ for the heptagonal lattice) and the identity operator $I$. This simple transformation makes the Hamiltonian proportional to the lattice Laplacian, except for the outermost points where $k \neq 3$. In short, the tight-binding Hamiltonian can be phenomenologically treated as that of the free particle simulated on the discrete lattice, described by $i \partial \Psi / \partial t = \triangle \Psi$ with the wave function $\Psi = \langle {\bf r}| \psi \rangle$. Accordingly, if we substitute the time $t$ in the Schr{\"o}dinger equation by the imaginary time $i t$, it takes exactly the same form as the diffusion equation. It is then straightforward to apply the previous result on classical diffusion in hyperbolic lattice (see Fig.~\ref{fig:dif}) to obtain the propagating solution of the quantum particle on the hyperbolic plane~\cite{banica} (see Fig.~\ref{fig:sch}). It is revealed that the average distance again exhibits the linear diffusion property as shown in the inset of Fig.~\ref{fig:sch} in accordance with Fig.~\ref{fig:level}. \begin{figure} \includegraphics[width=0.46\textwidth]{fig5.eps} \caption{(Color online) Solution of the quantum free particle Schr{\"o}dinger equation in the negatively curved surface. Inset: Average distance from the origin $\left< r \right> = \int r |\Psi(r)|^2 \sinh~r ~dr$ is shown to increase linearly in time.} \label{fig:sch} \end{figure} %\section{Discussion}\label{sec:discussion} In the Euclidean or disordered structures, the quantum diffusion has been known to be faster than the classical one~\cite{bjkim,mulken}. In contrast, the classical motion now becomes comparable to the quantum one in speed on the negatively curved surface, with the aid of the geometrical drift. Some differences, however, seem to remain: The geometrical drift pushes the particle only in the outward direction from the starting point. Based on the report for tree structures, a quantum particle is believed to readily propagate back to the origin~\cite{childs}. In fact, a classically diffusing particle acts somewhat like a flow, even if the initial direction may be chosen randomly, in the sense that it appears to have some nonzero velocity. If one introduces a real flow here, however, the flux would soon become negligible because of the exponentially increasing boundary. %Therefore we believe that %the transport should be mainly driven by a diffusive way in such a geometry. Our observation is qualitatively similar to the report for the Bethe lattice~\cite{monthus}, since the angular movement is effectively suppressed in the long run. Yet it is still notable that the heptagonal lattice provides a better representation for the negatively curved surface, and that the numerical simulation could be taken with more ease by the hyperbolic tessellation technique explained above. One may expect that the transport can be enhanced by introducing a negative Gaussian curvature in biological or engineering applications. In summary, we have investigated the classical and quantum diffusions in the heptagonal lattice, which is a discrete representation of the surface with a constant negative curvature. Even a classical particle has been shown to diffuse so fast that the average displacement is linearly proportional to time. The quantum diffusion has also been shown to exhibit the same linear diffusion behavior. Those results on discrete lattices were also confirmed in a continuum by solving a hyperbolic diffusion equation in real and imaginary times for classical and quantum diffusion, respectively. \acknowledgments This work was supported by the Korea Research Foundation Grant funded by the Korean Government (MOEHRD) with the Grant No. KRF-2005-005-J11903 (S.K.B.) and KRF-2006-312-C00548 (B.J.K.). \bibliographystyle{revtex} \begin{thebibliography}{10} \providecommand*{\bibinfo}[2]{#2} \providecommand*{\eprint}[1]{#1} \providecommand*{\url}[1]{#1} \bibitem{rudnick} \bibinfo{author}{J.~Rudnick} and \bibinfo{author}{G.~Gaspari}, \bibinfo{title}{\emph{Elements of the Random Walk: An Introduction for Advanced Students and Researchers}} (\bibinfo{publisher}{Cambridge University Press}, New York, \bibinfo{year}{2004}). \bibitem{bjkim} \bibinfo{author}{B.~J. Kim}, \bibinfo{author}{H.~Hong}, and \bibinfo{author}{M.~Y. Choi}, \bibinfo{journal}{Phys. Rev. B} \bibinfo{volume}{\textbf{68}}, \bibinfo{pages}{014304} (\bibinfo{date}{2003}). \bibitem{mulken} O. M{\"u}lken and A. Blumen, Phys. Rev. E {\bf 73}, 066117 (2006). %title: Efficiency of quantum and classical transport on graphs %quantum mechanical transport is faster in general \bibitem{jdnoh} \bibinfo{author}{J.~D. Noh} and \bibinfo{author}{H.~Rieger}, \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{92}}, \bibinfo{pages}{118701} (\bibinfo{date}{2004}); \bibinfo{author}{J.~D. Noh} and \bibinfo{author}{S.-W. Kim}, \bibinfo{journal}{J. Korean Phys. Soc.} \bibinfo{volume}{\textbf{48}}, \bibinfo{pages}{S202} (\bibinfo{date}{2006}). \bibitem{shima-belo-sausset} \bibinfo{author}{H.~Shima} and \bibinfo{author}{Y.~Sakaniwa}, \bibinfo{journal}{J. Stat. Mech. (2006) P08017}; \bibinfo{journal}{J. Phys. A: Math. Theor.} \bibinfo{volume}{\textbf{39}}, \bibinfo{pages}{4921} (\bibinfo{date}{2006}); \bibinfo{author}{L.~R.~A. Belo}, \bibinfo{author}{N.~M. Oliveira-Neto}, \bibinfo{author}{W.~A. Moura-Melo}, \bibinfo{author}{A.~R. Pereira}, and \bibinfo{author}{E.~Ercolessi}, \bibinfo{journal}{Phys. Lett. A} \bibinfo{volume}{\textbf{365}}, \bibinfo{pages}{463} (\bibinfo{date}{2007}); \bibinfo{author}{F.~Sausset} and \bibinfo{author}{G.~Tarjus}, %\bibinfo{title}{\emph{Periodic boundary conditions on the pseudosphere}} %(\bibinfo{date}{2007}), %\eprint{cont-mat/0703326v1}. \bibinfo{journal}{J. Phys. A: Math. Theor.} \bibinfo{volume}{\textbf{40}}, \bibinfo{pages}{12873} (\bibinfo{date}{2007}). \bibitem{baek} \bibinfo{author}{S.~K. Baek} and \bibinfo{author}{B.~J. Kim}, \bibinfo{journal}{Europhys. Lett.} \bibinfo{volume}{\textbf{79}}, \bibinfo{pages}{26002} (\bibinfo{date}{2007}). \bibitem{thurston} \bibinfo{author}{W.~P. Thurston} and \bibinfo{author}{J.~R. Weeks}, \bibinfo{journal}{Sci. Am.} \bibinfo{volume}{\textbf{251}}, \bibinfo{pages}{108--120} (\bibinfo{date}{1984}). \bibitem{anderson} \bibinfo{author}{J.~W. Anderson}, \bibinfo{title}{\emph{Hyperbolic Geometry}} (\bibinfo{publisher}{Springer-Verlag}, London, \bibinfo{year}{1999}). \bibitem{nano} D. Vanderbilt and J. Tersoff, Phys. Rev. Lett. {\bf 68}, 511 (1992); N. Park, M. Yoon, S. Berber, J. Ihm, E. Osawa, and D. Tom{\'a}nek, {\it ibid.} {\bf 91}, 237204 (2003). \bibitem{kendall} \bibinfo{author}{W.~S. Kendall}, \bibinfo{journal}{S{\'e}m. Probab. (Strasbourg)} \bibinfo{volume}{\textbf{18}}, \bibinfo{pages}{70} (\bibinfo{date}{1984}). \bibitem{monthus} \bibinfo{author}{C.~Monthus} and \bibinfo{author}{C.~Texier}, \bibinfo{journal}{J. Phys. A: Math. Theor.} \bibinfo{volume}{\textbf{29}}, \bibinfo{pages}{2399} (\bibinfo{date}{1996}). \bibitem{karlsson} \bibinfo{author}{A.~Karlsson}, in \emph{Proceedings of a Workshop at the Schr{\"o}dinger Institute, Vienna, 2001} (\bibinfo{publisher}{de Gruyter}, Berlin, \bibinfo{year}{2004}), \bibinfo{pages}{p. 459}. %\bibitem{aste} T. Aste, Phys. Rev. E {\bf 55}, 6233 (1997). \bibitem{coxeter} \bibinfo{author}{H.~S.~M. Coxeter}, \bibinfo{journal}{Bull. Can. Math.} \bibinfo{volume}{\textbf{40}}, \bibinfo{pages}{158} (\bibinfo{date}{1997}). \bibitem{green} \bibinfo{author}{M.~J. Greenberg}, \bibinfo{title}{\emph{Euclidean and Non-Euclidean Geometries: Development and History}} (\bibinfo{publisher}{W. H. Freeman}, New York, \bibinfo{year}{1993}), 3rd ed. %\bibitem{bender} %\bibinfo{author}{E.~A. Bender} and \bibinfo{author}{L.~B. Richmond}, % \bibinfo{journal}{Ann. Prob.} \bibinfo{volume}{\textbf{12}}, % \bibinfo{pages}{274} (\bibinfo{date}{1984}). \bibitem{young} \bibinfo{author}{E.~C. Young}, \bibinfo{title}{\emph{Vector and Tensor Analysis}}, 2nd ed. (\bibinfo{publisher}{Marcel Dekker}, New York, \bibinfo{year}{1993}). \bibitem{banica} \bibinfo{author}{V.~Banica}, \bibinfo{journal}{Commun. Partial Differ. Equ.} \bibinfo{volume}{\textbf{32}}, \bibinfo{pages}{1643} (\bibinfo{date}{2007}). \bibitem{cpzhu-giraud} C.-P. Zhu and S.-J. Xiong, Phys. Rev. B {\bf 62}, 14780 (2000); O. Giraud, B. Georgeot, and D. L. Shepelyansky, Phys. Rev. E {\bf 72}, 036203 (2005). % localization-delocalization transition in quantum small-world network \bibitem{childs} \bibinfo{author}{A.~M. Childs}, \bibinfo{author}{E.~Farhi}, and \bibinfo{author}{S.~Gutmann}, \bibinfo{journal}{Quantum Inf. Process.} \bibinfo{volume}{\textbf{1}}, \bibinfo{pages}{35} (\bibinfo{date}{2002}). \end{thebibliography} \end{document} }
\caption{\label{fig:rat_comp} The ratio of $K^0$($K^+$) yields produced by pions (protons) on heavy and light targets plotted as a function of the momentum, $p$, in the lab. system. The full squares depict the yield\,-\,ratio of$K^0_S$ produced on Pb and C targets in this experiment. A similar ratio of $K^+$ yields measured in proton-induced reactions on Au and C targets is represented by full circles~\cite{Nekipelov02}. The results of the HSD model with different strength of the KN\,-\,potential are depicted by solid (black, dashed (red) and blue (dotted) lines.}
\caption{Results for the observations which showed the highest values of \redc. Each observation is given a case number (column one) and the observation identification number is shown in column two. We give the date, the exposure time (Exp.), the degrees of freedom ($\nu$) after any flare-filtering, the \redc\value and then state the ratio between the$\chi^{2}$ value of the line lightcurve and that of the continuum lightcurve (\cratio). We also state the average \fin\(FF) ratio (and error on this value) between the MOS1 and MOS2. The final column gives a comment about the observation as described in the text.}
\caption{Histogram of \redc\values.}
\caption{Scatter plot of \redc\versus\cratio\values. Filled circle are used for observations with a suspected or previously published SWCX enhancement.}
\caption{\label{fig4} (color online). Transverse scans in the $ab$ plane through the $(100)$ magnetic Bragg peak at (a) $T$=5~K and (b) $T$=21~K. The data were collected on the 4-circle TriCS diffractometer at PSI. Each peak is fitted (full line) by the sum of a broad and a narrow Gaussian function, together with a flat background. (c) FWHM and (d) integrated intensity of the broad (\textcolor{blue}{$\bigtriangleup$}) and narrow (\textcolor{magenta}{$\blacksquare$}) Gaussian components of the $(100)$ peak as a function of $T$. The total integrated intensity {\large($\bullet$)} is also shown. }
\caption{(color online). Dependence of $H_{sat}$ on $d_{Cr}$ in Nb(20nm)/Fe/$\{$Cr/Fe$\}_{\times 7}$ films for: $d_{Fe}$=0.6 nm (\textcolor[rgb]{1.00,0.00,0.00}{$\square$}); and $d_{Fe}$=1.0 nm ($\blacksquare$). Solid lines are to guide the eye. Inset: illustration of S/FM/$\{$NM/FM$\}_{N-1}$ structure with $N$ FM layers and $N-1$ NM layers. Vertical arrows indicate FM polarization direction.\label{Hsatvsdcr}}
\caption[]{Schematic metabolic network as a bipartite graph. Edges go from reaction nodes (\opensquare, blue) to metabolite nodes (\opendiamond, red). Reaction fluxes, $\va$, metabolite shadow prices, $\mi$, and the stoichiometry matrix, $\Sia$, combine to make the edge-associated yield flux network, $\Jia$. It can be shown (see text) that the $\Jia$ are conserved at all nodes except a limited number of exchange reactions which serve as sources, and the biomass demand reaction which serves as a sink. The biomass shadow price is unity, and the flux through the biomass demand reaction is the specific growth rate $\mu$.\label{fig:schem}}
\caption{Evolution of the granulation velocity and characteristic time versus the height (left) and the MPSI index (right). The linear trend with MPSI has been computed after removing the trend with height. \textcolor{blue}{$+$} and \textcolor{blue}{$\square$}: respectively BW1 and BW2; \textcolor{red}{$\triangle$}: RW. The solid lines in each panel are the best fitting gradient of linear regressions for independent series and for the separate blue- and red-wing data sets. Note the jump between red and blue operating mode. All panels show overlapping data for illustrative purposes, with errors plotted only on the independant data.}
\caption{\label{mmaxntot} Mass of the most massive star versus the number of stars in the cluster (for better visibility, a small random scatter was applied to the (discrete) masses). The data are collected from the literature, with the main sources Testi et al. ({\color{red} $\blacklozenge$}) and Weidner \& Kroupa ({\color{green} $\blacksquare$}). The references for the other points are given in Appendix \ref{datatable}. The solid line is the mean value of $m_\mathrm{max}$ depending on $n$. The dotted lines follow the $1/6$ and $5/6$ quantiles, and should confine $2/3$rd of the observed data. }
\caption{\label{quantiles} A plot of the ordered cumulative probabilities versus the observed cumulative probabilities derived from the distribution of $m_\mathrm{max}$ for a particular $n$. The shown data fullfill the selection criterion discussed in Section \ref{analysisdataset}, i.e. they lie in the dashed wedge in Fig. \ref{mmaxntot}. The filled symbols follow with the observed $n$, whereas for the open symbols $2n$ was used, corresponding to the right end of the error bars in Fig. \ref{mmaxntot} ({\color{green}$\blacksquare$}: Weidner \& Kroupa;{\color{red}$\blacklozenge$} Testi et al.). If the data were uniformly distributed they should follow the diagonal. }
\caption{Distribution of the conductance at transition points in the unitary class with a PCC at $L_x=L_y=129$ (\full) %(solid line) and without a PCC at $L_x=L_y=128$ (\dashed). %(dashed line). The size of the system is twice as large as in the case of Ref. \cite{Hirose}. }
\caption{Distribution of the conductance at transition points in the symplectic class with a PCC at $L_x=L_y=129$ (\full) %(solid line) and without a PCC at $L_x=L_y=128$ (\dashed). %(dashed line). }
\caption{Casimir energy per unit area as a function of the plate separation $\lb = L/(2\pi)$, normalized to the healing length $\zeta$. Recall that $\zeta = \frac {\hbar} 2 \sqrt{1/(g n m)}$ with $g$ the effective interaction constant and $n$ the BEC density. The Casimir energy has been normalized to its value $\esc ( L ) = - (\pi^2/90) \hbar c / L^3$ for a massless scalar field propagating with a velocity $c = \hbar/ (2 m \zb)$; this limit is approached at large distance (\dashed). The broken curve (\longbroken) shows the asymptotic expansion for $L/\zeta \gg 1$ \pr{eqn:cas_ex}, while the dotted curve (\chain) shows the opposite limit $L/\zeta \to 0$ (non-interacting Bose gas), of eqn.~\pr{eqn:E_nonint}. The full black curve is obtained by numerically evaluating the integral in \pr{eqn:DOS1} with the mode density $\rho( x )$ of \pr{eqn:rho_an}. It smoothly describes the dependence of $\ec$ on the interaction strength in the regime where both of the asymptotic expansions diverge.}
\caption{Linear-log plot of $P(T)$ versus $T$: the fitted $\alpha$'s for the four curves with $\Wi=0, 2.0, 6.1$, and $20.3$ are $1.20, 0.57, 0.375$, and $0.326$ respectively. The two datasets with symbols {\color{red} $\boxdot$} and {\color{blue}$\blacktriangle$} in (c) are obtained by switching off the thermal noise along $x$ direction ($\eta_1=0$) in \eref{evolve} and $\dot{\gamma} = 0.2$ and $0.6$ respectively ---both fit well with the analytical $\alpha = 0.324$ line. All data are for $N = 10$. Inset: Typical $R_x(t)$ versus $t$ corresponding to the cases (a) both $\dot{\gamma}\not=0$ and $\eta_1\not=0$, (b) $\dot{\gamma}=0$ and $\eta_1\not=0$, and (c) $\dot{\gamma}\not=0$ and $\eta_1=0$.}
\caption{% Predicted age evolution in the observed ACS colors at redshift $z{=}0.034$ for Bruzual \& Charlot (2003) single-burst stellar population models with five different metallicities, labeled by their[Fe/H] values. We also show the expected colors at this redshift for six different empirical galaxy templates (see text) with arbitrary placement along the horizontal axis. % The shaded areas delineate the color selection criteria for the UCD candidates. The broader baseline \gIcolor\color is used for the more stringent selection cut, based on the expected range of stellar populations in UCDs. The less-sensitive\rIcolor\cut is simply to ensure the objects have reasonable colors for galaxies at this redshift.}
\caption{\sersic\index$n$ is plotted against the circularized half light radius $\rc$ for the Galfit \sersic\model fits. The dashed lines show the biweight mean values of$1.47\pm0.15$ and $1.07\pm0.07$ for the objects with $10<\rc<100$ pc and $100<\rc<400$ pc, respectively.}
\caption{Locations of the 15 UCD candidates (blue diamonds), bright globular cluster candidates with $\Iacs<25$ and $\rc<10$\,pc (red squares), larger compact galaxies from Fig.\,\ref{fig:rest} (open circles), and all other objects in the field with $17<\Iacs<25$ and meeting our color cuts (small dots). The orientation is the same as in Fig.\,\ref{fig:pic}, although here we represent the full ${\sim\,}3\farcm4{\times}3\farcm4$ field. The contours show elliptical isophotes of \esog\with major axes of 0\farcm5, 1\farcm0, and 1\farcm5. The GCs preferentially align along the galaxy's major axis. \hbox{Two-thirds} of the UCD candidates also fall along this direction. \vspace{0.2cm} }
\caption{F814W magnitude versus size for UCD candidates (filled diamonds), larger compact galaxies in the 100-300~pc range (circles) and all other objects (open squares) in the \esog\field that meet our color selection criteria and are within the plotted magnitude and size limits. Objects with$\rc{\,<\,}10$~pc are designated globular cluster candidates, while the UCD candidates are chosen as having $\rc = 10$ to 100~pc and ellipticity $<\,$0.5. However, there may be a separation between the most compact UCD candidates with $\rc<20$~pc, similar to large globular clusters, and those with $\rc\gta40$~pc, which may be true compact dwarfs. Completely unresolved objects with $\rc\approx0$ fall off the edge of this logarithmic plot. We show the expected location for M32 at this distance; no similar galaxies are found in our sample. \vspace{0.1cm}}
\caption{\pttrig dependence of the \njet for \Cu and \Au at \sNNsixtytwo and \dAu, \Cu, and \Au at \sNNtwohundred compared to the yield from PYTHIA scaled by 2/3. Color online.}
\caption{\ptassoc dependence of \njet for \Cu and \Au at \sNNsixtytwo and \dAu, \Cu, and \Au at \sNNtwohundred compared to the yield from PYTHIA scaled by 2/3. The inverse slope parameters from fits of an exponential to the data and to PYTHIA are given in \tref{Table}. Color online.}
\caption{Inverse slope parameter k (\MeV) of \ptassoc for fits of data in \Fref{AssocPt}. The inverse slope parameter from a fit to $\pi^-$ in \Au from \cite{Pion} above 1.0 \GeV is $k$ = 280.9 $\pm$ 0.4 \MeV for \sNNsixtytwo and is $k$ = 330.9 $\pm$ 0.3 \MeV for \sNNtwohundred. Statistical errors only.}
\caption{\npart dependence of the \njet for \Cu and \Au at \sNNsixtytwo and \dAu, \Cu, and \Au at \sNNtwohundred compared to the yield from PYTHIA. Color online.}
\caption{(a) distribution of trigger particles \zT and (b) \pthat distribution in PYTHIA at \sNNsixtytwo from PYTHIA 8.1 for \stdassoc and \stdtrig at \sNNsixtytwo and \sNNtwohundred. Color online.}
\caption{\nridge dependence on \npart for \sNNsixtytwo and\sNNtwohundred. Color online.}
\caption{\nridge/\njet dependence on \npart for \sNNsixtytwo and\sNNtwohundred. Color online.}
\caption{$\eta_{\mathrm{v}}\left(k\right)$ for [\textcolor{blue}{$\times$}]$\Psi_{\mathrm{t}}^{(1,\mathrm{A})}$, [\textcolor{myGreen}{$+$}]$\Psi_{\mathrm{t}}^{(2,\mathrm{A})}$ and [\textcolor{myRed}{\opencircle}], reported by Armour and Baker \cite{Armour1987}.} \label{fig:PSvK1} \end{figure} \begin{figure} \centering \includegraphics{figure2.eps} \caption{$Z_{\mathrm{eff}}\left(k\right)$ for [\textcolor{blue}{$\times$}]$\Psi_{\mathrm{t}}^{(1,\mathrm{A})}$, [\textcolor{myGreen}{$+$}]$\Psi_{\mathrm{t}}^{(2,\mathrm{A})}$ and [\textcolor{myRed}{\opencircle}], reported by Armour and Baker \cite{Armour1987}.}
\caption{The dependence of \ps on \pc at $k=0.04$, for [\textcolor{blue}{$\times$}]$\Omega^{(1)}$ and [\textcolor{myGreen}{$+$}]$\Omega^{(2)}$. Basis functions have been removed successively from \targA.}
\caption{The dependence of \zeff on \pc at $k=0.04$, for [\textcolor{blue}{$\times$}]$\Omega^{(1)}$ and [\textcolor{myGreen}{$+$}]$\Omega^{(2)}$. Basis functions have been removed successively from \targA.}
\caption{$\eta_{\mathrm{v}}\left(k\right)$ for [\textcolor{blue}{$\times$}]$\Psi_{\mathrm{t}}^{(1,\mathrm{A})}$, [\textcolor{myGreen}{$+$}]$\Psi_{\mathrm{t}}^{(2,\mathrm{A})}$, [\textcolor{myRed}{--$\cdot$--}]$\Psi_{\mathrm{t}}^{(1,\mathrm{B})}$ and [---]$\Psi_{\mathrm{t}}^{(2,\mathrm{B})}$.} \label{fig:PSvK2} \end{figure} \begin{figure} \centering \includegraphics{figure8.eps} \caption{$Z_{\mathrm{eff}}\left(k\right)$ for [\textcolor{blue}{$\times$}]$\Psi_{\mathrm{t}}^{(1,\mathrm{A})}$, [\textcolor{myGreen}{$+$}]$\Psi_{\mathrm{t}}^{(2,\mathrm{A})}$, [\textcolor{myRed}{{\textbf{ --$\cdot$--}}}]$\Psi_{\mathrm{t}}^{(1,\mathrm{B})}$ and [---]$\Psi_{\mathrm{t}}^{(2,\mathrm{B})}$.}
\caption{The dependence of \ps on \pc at $k=0.04$, for [\textcolor{blue}{$\times$}]$\Omega^{(1)}$ and [\textcolor{myGreen}{$+$}]$\Omega^{(2)}$. Basis functions have been removed successively from \targB.}
\caption{The dependence of \zeff on \pc at $k=0.04$, for [\textcolor{blue}{$\times$}]$\Omega^{(1)}$ and [\textcolor{myGreen}{$+$}]$\Omega^{(2)}$. Basis functions have been removed successively from \targB.}
\caption{(color online). Measured \Leff~values as a function of Xe nuclear recoil energy. Symbols correspond to (\textcolor{red}{$\circ$})--this work; (\textcolor[rgb]{0,.5,0}{\bf$\square$})--Chepel et al.~\cite{Chepel:06}; (\textcolor{blue}{$\bigtriangleup$})--Aprile et al.~\cite{Aprile:05}; (\textcolor[rgb]{.6,0,0}{$\lozenge$})--Akimov et al.~\cite{Akimov:02}; ($\times$)--Bernabei et al.~\cite{Bernabei:01}; (\textcolor[rgb]{1.,0,1.}{$\bigtriangledown$})--Arneodo et al.~\cite{Arneodo:00}. The solid gray curve is the result from a recent best-fit analysis of XENON10 AmBe source data and MC \cite{Sorensen:09}. Also shown is the theoretical prediction of Hitachi (dashed line) \cite{Hitachi:BiEx}.}
\caption{\label{fig2}Number (\ding{110}) and temperature (\textcolor{blue}{\ding{108}}) of cooled atoms as a function of $h_p$ (height of the pump beam above the trough vertex). The positive slope of $T_O^{(z)}$ reflects energy gained by atoms in free fall. For $h_p > 100\:\mu\text{m}$, the additional energy increases the loss rate from the optical trough. For $h_p < 100\:\mu\text{m}$ spatial overlap of the pump beam and optical trough beams reduces the excitation probability and hence the capture rate. The highest phase-space density is achieved at $h_p = 41\:\mu\text{m}$.}
\caption{\label{fig3}Atom capture efficiency as a function of the initial magnetic trap temperature. The solid line (\textcolor{red}{-}) represents the upper bound capture efficiency given by (\ref{eq:2}). Circles are experimental data. Above $40\:\mu\text{K}$, we measure efficiencies which clearly surpass the predicted limit. We attribute this divergence to an increasing collision rate (inset).}
\caption{The fraction of ``red'' galaxies (\fred; left) and passive galaxies (\fpass; right) for central ({\it top set of plots}) and satellite ({\it bottom set of plots}) galaxies in the (\mhalo, \mgal) plane. The fraction of red/passive galaxies in a given pixel in (\mhalo, \mgal) is indicated by the colour, where red colours indicate a higher red/passive fraction, as shown in the scale. To guide the eye, we draw a solid (dotted) line showing the approximate upper envelope of the central (satellite) galaxy mass distribution for the observational group catalog, and repeat this same line on every panel. Central and satellite galaxies in the observational group catalogs show noticably different joint dependencies on stellar mass and halo mass. The models with AGN feedback qualitatively reproduce the trends for central galaxies, but do not reproduce the empirical trends for satellites. }
\caption{The fraction of "red" (\fred) (left panels) and "passive" galaxies ($f_{\rm passive}$) (right panels) in models without dust corrections. The colour scale is as in Fig.~\ref{fig:fred2d}. We present the results for central galaxies (left) and satellite (right) galaxies seperately. We see that the details of the \fred\distribution are quite sensitive to the dust correction, whereas$f_{\rm passive}$ is not noticably affected by the dust correction. We also note that the dust-free \fred\results appear more similar to the\fpass\results, which presumably probe the physical properties of galaxies more directly.}
\caption{ The fraction of ``red'' galaxies (\fred) (left panels) and of ``passive'' galaxies ($f_{\rm passive}$) (right panels) without selection criteria. The colour scale is as in Fig.~\ref{fig:fred2d}. We present the results for central galaxies (left) and satellite (right) galaxies seperately. It can be inferred that our main results do not depend on our selection criteria. }
\caption{The fraction of ``red'' galaxies (\fred) as a function of halo mass and stellar mass, where halo masses have been assigned in the semi-analytic models using an approach similar to that used in the SDSS group catalogs. The colour scale is as in Fig.~\ref{fig:fred2d}. We present the results for central galaxies (left) and satellite (right) galaxies seperately. We see that the procedure used to assign halo masses in the SDSS group catalogs reduces the scatter in \mgal\at fixed halo mass, and washes out many of the detailed features of the 2d distribution that are visible in the raw model predictions.}
\caption{Close-up of the egress of \vela from the first eclipse using a pulse averaged light curve of \vela (time resolution of 283.5\,s). The flux increases exponentially with time until a normal flux level is reached as indicated by the solid red line. The egress time is determined as the time when the count rate was no longer consistent with zero and is marked by a thick vertical line.}
\caption{Fit results for the overall spectrum excluding the eclipse. When fixing $\Gamma$ to one of the values found in the literature \citep[fixed~1 and fixed~2;][]{kreykenbohm99a,labarbera03a}, \ecut and \efold are better constrained, but the resulting \redchi-values are higher. All uncertainties here and elsewhere in the paper are on a 90\% confidence level. The DOF are the degrees of freedom.}
\caption{The trajectories of the poles of $S_{+}$ for the symmetric states of the one-dimensional symmetric rectangular potential. The full circles (\fullcircle) indicate the attractive poles and the open circles (\opencircle) the repulsive poles. The arrows given along the curves indicate the directions of the movement of poles as the phase parameter $\alpha$ increases. \label{fig: RECT1}}
\caption{The trajectories of the poles of $S_{-}$ for the antisymmetric states of the one-dimensional symmetric rectangular potential. The full circles (\fullcircle) indicate the attractive poles and the open circles (\opencircle) the repulsive poles. The arrows indicate the directions of the movement of poles as the phase parameter $\alpha$ increases. \label{fig: RECT2}}
\caption{Convergence of [\textcolor{blue}{---}]$\eta_{\mathrm{v}}\left(\tau\right)$ to [\textcolor{myGreen}{+}]$\hat{\eta}_{\mathrm{v}}$ as $\tau\rightarrow\hat{\tau}_{\mathrm{s}}$ at $k=0.2$.} \label{fig:convergence} \end{figure} We can reasonably conclude that we have encountered anomaly-free singularities. Before we can develop an optimization for $\tau$ based on these singularities, however, there are two outstanding issues to be addressed. Firstly, we have already noted that no real-valued solutions of (\ref{eq:qeq}) were found at $k=0.65$ and $k=0.66$. This is potentially problematic as, for a sufficiently accurate trial function, we should expect at least one real root of (\ref{eq:qeq}) at each $k$, corresponding to an anomaly-free singularity. However, inspection of figure \ref{fig:sMap} shows that the solutions of (\ref{eq:qeq}) are close together in the regions either side of $k=0.65$ and $k=0.66$. Were the two roots to coincide at some $k$, then $\mathcal{B}^2=4\mathcal{A}\mathcal{C}$. Near a point of coincidence, $\mathcal{B}^2\sim4\mathcal{A}\mathcal{C}$ and small errors in the values of $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ could erroneously give rise to $\left(\mathcal{B}^2-4\mathcal{A}\mathcal{C}\right)<0$. The values of $\tau$ solving (\ref{eq:qeq}) at $k=0.65$ and $k=0.66$ were found to be, respectively, $\tau_{\mathrm{s}}\sim2.02\pm0.01\rmi$ and $\tau_{\mathrm{s}}\sim1.98\pm0.03\rmi$. In both cases, the fact that $\Im\left[\tau_{\mathrm{s}}\right]\ll\Re\left[\tau_{\mathrm{s}}\right]$ suggests that singularities do genuinely exist for $\tau\in\mathbb{R}$ at these values of $k$, but small errors in our calculations of $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ due to inexact numerical integration have prevented us from finding them. Having investigated this problem in more detail, in figure \ref{fig:imroot} we show the calculated values of $\Im\left[\tau_{\mathrm{s}}\right]$ for $31$ values of $k$ equidistant in the range $0.64 \leq k \leq 0.67$. There is a clearly defined region of $k$ where no real roots of (\ref{eq:qeq}) have been found. The smoothness of $\Im\left[\tau_{\mathrm{s}}\right]$ over $k$ in this region does not necessarily preclude the notion that the failure to find real-valued solutions is due to small numerical errors in our calculations. It is conceivable that inaccuracies in the calculated values of $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ could also arise from systematic errors in the algorithm \cite{NAG1} used to calculate the determinants. Nevertheless, the results illustrated in figure \ref{fig:imroot} are interesting; their exact origin may be speculated upon and will remain a subject of our ongoing investigations. \begin{figure} \centering \includegraphics{figure7.eps} \caption{Values of $\Im\left[\tau_{\mathrm{s}}\right]$ for $0.64 \leq k \leq 0.67$.} \label{fig:imroot} \end{figure} The second difficulty concerns the choice of $\hat{\tau_{\mathrm{s}}}$ from the two available solutions of (\ref{eq:qeq}). A method is needed for identifying at each $k$ the root of (\ref{eq:qeq}) corresponding to a legitimate singularity. This can easily be achieved by inspecting values of \ps at values of $\tau$ either side of each singularity, although this approach is not ideal as it requires solutions of (\ref{eq:KohnEq}) to be found. In practice, at each $k$ it should be possible to determine by inspection which of the two phase shifts is anomaly-free by examining corresponding results at singularities for nearby values of $k$. For example, figure \ref{fig:sMap} clearly shows that only one curve in the $\left(\tau,k\right)$ plane corresponds to a physically acceptable variation of phase shift over $k$. With these considerations in mind, we claim that choosing $\tau=\hat{\tau}_\mathrm{s}$ at each $k$ defines a consistent optimization that can be used to avoid anomalies due to Schwartz singularities appearing at other values of $\tau$. To evaluate the success of this approach, we have found it helpful to consider an alternative optimization of $\tau$. For the calculations of \ps carried out with $p=1001$, choosing the value of $\tau$ at each $k$ giving rise to the median phase shift, $\langle\eta_{\mathrm{v}}\rangle$, should also mitigate anomalous behaviour. In figure \ref{fig:comparison} we have compared results for $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$ and $\hat{\eta}_{\mathrm{v}}\left(k\right)$ for momenta in the range $0.01\leq k \leq1$. Here, of the two candidates for $\hat{\eta}_{\mathrm{v}}$ corresponding to the two singularities, at each $k$ we have chosen the one whose absolute value is closest to $\vert\langle{\eta}_{\mathrm{v}}\rangle\vert$. For clarity, we have included in the figure values of $\hat{\eta}_{\mathrm{v}}\left(k\right)$ for only $50$ values of $k$ equidistant in the range $0.01 \leq k \leq 0.99$. \begin{figure} \centering \includegraphics{figure8.eps} \caption{A comparison of optimization schemes for $\tau$, [\textcolor{blue}{\full}] $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$ and [\textcolor{myGreen}{$\times$}] $\hat{\eta}_{\mathrm{v}}\left(k\right)$.} \label{fig:comparison} \end{figure} Both optimization schemes successfully avoid anomalous behaviour at most values of $k$ and there is good agreement between the two sets of results at all momenta. However, the intriguing feature of the figure is the anomaly appearing in both sets of results at $k\sim0.71$. For $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$, we have shown this anomaly in greater detail by including in the figure results of a further $100$ Kohn calculations for momenta equidistant in the range $0.7\leq k\leq0.72$, although it is practical to consider henceforth only the anomalous behaviour occurring precisely at $k=0.71$. We believe the anomaly shown in figure \ref{fig:comparison} is of a different type to those shown in figure \ref{fig:psMap}, which are due to singularities found by varying only $\tau$ at a given $k$. In the following section we will examine the circumstances under which persistent anomalies of the kind shown in figure \ref{fig:comparison} could appear, before going on to discuss methods designed to avoid them. \subsection{Persistent anomalous behaviour}\label{ss:persistent} It is often claimed that anomalous results observed in the region of singularities arise from $A$ having a determinant close to zero. Statements of this kind can be misleading, as the determinant of any nonsingular $A$ can be made arbitrarily close to zero by an appropriate scalar multiplication, without altering the sensitivity of the solution, $x$, to small errors in the elements of $A$ or $b$. A better measure for identifying regions where anomalies may occur is the condition number, \cond, defined for nonsingular $A$. The condition number is independent of the normalization constant, $N$, and for the linear system (\ref{eq:KohnEq}) is defined as the maximum ratio of the relative error in $x$ and the relative error in $b$. Formally, it can be shown that \begin{equation}\label{eq:condition} \kappa\left(A\right)=\parallel A\parallel\parallel A^{-1}\parallel, \end{equation} with respect to some matrix norm, $\parallel A \parallel$ \cite{Higham2002}. The value of \cond is dependent upon the choice of norm. In our calculations, we have considered the matrix 1-norm \cite{Higham2002} of $A$, \begin{equation}\label{eq:onenorm} \parallel A \parallel_{1}=\max_{1\leq j\leq\left(M+2\right)}\sum_{i=1}^{M+2}\vert a_{ij}\vert, \end{equation} where $a_{ij}$ is the element in the $i^{\mathrm{th}}$ row and $j^{\mathrm{th}}$ column of $A$. In what follows, the particular choice of the 1-norm in our calculations will implicitly be assumed. A matrix with a large condition number is said to be ill-conditioned, and the solution of the corresponding linear system may not be reliable if the elements of $A$ and $b$ are not known exactly. For any invertible $A$, the condition number can be used to formalize the definition of closeness to singularity in the following way. If $\Delta A$ is defined to be any matrix such that $A+\Delta A$ is singular, then the relative distance to singularity, \dist, for $A$, is defined \cite{Higham2002} to be \begin{equation}\label{eq:distance} \Lambda\left(A\right)=\min\left(\frac{\parallel \Delta A\parallel}{\parallel A\parallel}:\quad \det\left(A+\Delta A\right)=0\right). \end{equation} This definition holds for any consistent norm. Further, if \dist and \cond are evaluated using the same choice of norm, it can be shown \cite{Higham2002} that \begin{equation}\label{eq:distcond} \Lambda\left(A\right)= \left[\kappa\left(A\right)\right]^{-1}. \end{equation} In section \ref{ss:pscalcs} we noted that (\ref{eq:qeq}) has, in general, no more than two zeros if only variations of $\tau\in\left[0,\pi\right)$ are considered. However, if $\mathcal{A}=\mathcal{B}=\mathcal{C}=0$ then \deter is identically zero independently of $\tau$ and no consistent value of phase shift can be calculated, either by solving (\ref{eq:KohnEq}) or directly from (\ref{eq:afm}). There is no obvious physical reason why this circumstance should arise at any $k\in\mathbb{R}$. However, it is conceivable that $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ could coincidentally be close to zero, in some sense, over a narrow range of $k$. Small errors in the evaluation of $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ could then give rise to both unreliable solutions of (\ref{eq:qeq}) and persistent anomalies in the calculation of \ps due to ill-conditioning in the Kohn equations (\ref{eq:KohnEq}). To see how the latter case arises, using (\ref{eq:detan}) we note that \begin{eqnarray} \mathcal{C}&=&\det\left[A\left(\tau=0\right)\right],\\ \mathcal{A}&=&\det\left[A\left(\tau=\frac{\pi}{2}\right)\right],\\ \mathcal{B}&=&2\det\left[A\left(\tau=\frac{\pi}{4}\right)\right]-\mathcal{A}-\mathcal{C}, \end{eqnarray} so that the notion of $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ being close to zero is immediately formalized in terms of \dist at $\tau=0$, $\tau=\pi/4$ and $\tau=\pi/2$. In our calculations, we have used a numerical algorithm \cite{NAG2} to calculate \diste, an estimate of \dist. Values of \diste at $\tau=0$, $\tau=\pi/4$ and $\tau=\pi/2$ are shown in figure \ref{fig:ABC} for $100$ values of $k$ in the range $0.7 \leq k \leq 0.72$. Values of \diste for $\tau=0$ and $\tau=\pi/4$ are anomalously small at $k\sim0.71$, with the values for \diste at $\tau=\pi/2$ also passing through a clear minimum at $k\sim0.711$. Ordinarily, we would not expect ill-conditioning to occur over a very broad range of $\tau$ at a given $k$. At $k=0.71$, the small values of \diste at $\tau=0$, $\tau=\pi/4$ and $\tau=\pi/2$ therefore point to a manifestation of ill-conditioning which is unusually widespread in $\tau$. In fact, in our calculations we have failed to find any value of $\tau\in\left[0,\pi\right)$ such that $A$ is sufficiently well-conditioned to avoid anomalous results at $k=0.71$. We have also confirmed that $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ all pass through zero at least once between $k=0.71$ and $k=0.7104$. \begin{figure} \centering \includegraphics{figure9.eps} \caption{Values of \diste over $k$ for [\textcolor{blue}{\dashed}] $\tau=0$, [\textcolor{myGreen}{\chain}] $\tau=\frac{\pi}{4}$ and [{\dotted}] $\tau=\frac{\pi}{2}$.} \label{fig:ABC} \end{figure} We can conclude that both optimization schemes developed in section \ref{ss:pscalcs} successfully avoid anomalies due to Schwartz singularities whose existence depends only upon the choice of $\tau$. However, at certain values of $k$, $A$ can become close to singularities whose existence is independent of $\tau$, resulting in anomalies in the calculation of \ps that persist even after $\tau$ has been optimized. In attempting to address this problem, a number of methods are available. In principle, anomalous behaviour can be reduced dramatically by allowing the Kohn trial wavefunction to be complex-valued \cite{Schneider1988,McCurdy1987}. An alternative approach for avoiding anomalies persistent in $\tau$ is to make a small change in some other parameter of the trial wavefunction. We will explore both of these techniques in the following section. \subsection{The complex Kohn method} The complex Kohn method is an extension of the original variational approach in which the boundary conditions of the trial wavefunction are complex. It was originally believed \cite{Schneider1988,McCurdy1987} that this method was anomaly-free, although anomalies were subsequently reported by Lucchese \cite{Lucchese1989}. For our complex Kohn calculations on \ehmol scattering, we have used a trial wavefunction, $\Psi_{\mathrm{t}}'$, of the form \begin{equation}\label{eq:complextrialwave} \Psi_{\mathrm{t}}' = \left(\bar{S} + a_{\mathrm{t}}'\bar{T} + p_{0}'\chi_{0}\right)\psi_{\mathrm{G}} + \sum_{i=1}^{M} p_{i}'\chi_{i}, \end{equation} where \begin{equation}\label{eq:hankel} \bar{T}=\bar{S}+\rmi\bar{C}, \end{equation} the functions \targetwave and $\Omega=\{\chi_{1},\dots,\chi_{M}\}$ being the same as in (\ref{eq:trialwave}). The unknowns $a_{\mathrm{t}}'$ and $\{p_{0}',\dots,p_{M}'\}$ will not, in general, be real. Application of the variational principle to (\ref{eq:complextrialwave}) leads to a matrix equation analogous to (\ref{eq:KohnEq}), \begin{equation}\label{eq:CKohnEq} A'x'=-b' \end{equation} where $A'$ and $b'$ are identical to $A$ and $b$, but for the function, $\bar{T}$, replacing $\bar{C}$ in (\ref{eq:matA}) and (\ref{eq:vecB}). The determinant, $\det\left(A'\right)$, conveniently reduces to \begin{equation} \det\left(A'\right) = \mathcal{D}\left(k\right)e^{-2\rmi\tau}, \end{equation} where \begin{equation}\label{eq:DABC} \mathcal{D}=\left(\mathcal{A}-\mathcal{C}\right)-\rmi\mathcal{B} \end{equation} is a complex constant with respect to variations in $\tau$. The values of $\det\left(A'\right)$ then describes a circle of radius $\vert \mathcal{D}\vert$ in the complex plane for variations of $\tau\in\left[0,\pi\right)$. Hence, singularities are obtained only if both the real and imaginary parts of $\mathcal{D}$ are zero; they can neither be located nor avoided by varying only $\tau$. We would therefore expect anomalous results due to singularities arising from the choice of $\tau$ at a given $k$ to be eliminated in the complex Kohn method. However, from (\ref{eq:DABC}) we would also expect $\mathcal{D}$ to be close to zero at $k=0.71$, in the same sense that $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ have already been seen to be close to zero at this value of $k$. It is therefore likely that the anomalies already seen to occur due to relationships between $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ rather than the choice of $\tau$, will persist even in the complex Kohn method. We have obtained values of \ps using the trial function, \ctrialwave. We have found that the differences in the calculated values of \ps at different values of $\tau\in\left[0,\pi\right)$ are negligible, for all positron momenta considered here. Without loss of generality, we can regard the complex Kohn calculation as effectively independent of $\tau$ and choose $\tau=0$ for simplicity. In figure \ref{fig:methComp} we have compared results for $\eta_{\mathrm{v}}\left(k,\tau=0\right)$ obtained with the trial function, \ctrialwave, with the results for $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$ obtained in section \ref{ss:pscalcs} with \trialwave. \begin{figure} \centering \includegraphics{figure10.eps} \caption{A comparison of [\textcolor{blue}{\full}] $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$, determined using the trial function, \trialwave, with [\textcolor{myGreen}{$\times$}] $\eta_{\mathrm{v}}\left(k,\tau=0\right)$, evaluated using the complex Kohn method with the trial function, \ctrialwave.} \label{fig:methComp} \end{figure} The results of the two methods are essentially equivalent at all positron momenta, with the optimized results for \trialwave differing from the results for \ctrialwave by no more than $0.1 \%$ at each $k$. As expected, the complex Kohn method automatically avoids anomalies at most values of $k$ without the need for an optimization of $\tau$. Nevertheless, the use of the complex trial function has predictably failed to remove the persistent anomaly at $k=0.71$. In figure \ref{fig:D} we verify that $\Lambda_{\mathrm{a}}\left(A'\right)$ is anomalously small at $k\sim0.71$ for $\tau=0$. We have found that differences between the results shown in figure \ref{fig:D} and values of $\Lambda_{\mathrm{a}}\left(A'\right)$ calculated at other values of $\tau$ are negligible. \begin{figure} \centering \includegraphics{figure11.eps} \caption{Values of $\Lambda_{\mathrm{a}}\left(A'\right)$ over $k$ for $\tau=0$.} \label{fig:D} \end{figure} Having failed to find a systematic remedy for the persistent anomalous behaviour, we consider a more {\it ad hoc} approach. It should be possible to avoid any Schwartz-type anomaly by some variation of parameters in the trial wavefunction. We have found that variations in $\tau$ are not always successful, but other candidates exist. In our complex Kohn calculations, we have varied the values of $\alpha$ and $\beta$ in (\ref{eq:srcf}), fixing $\tau=0$. Recall that the values of the these parameters have so far remained fixed at $\alpha=0.6$ and $\beta=1.0$. We now consider the results of Kohn calculations carried out for $31$ different values of $\alpha$ in the range $\alpha\in\left[0.59,0.605\right]$ and $61$ different values of $\beta$ in the range $\beta\in\left[0.65,1.25\right]$. \begin{figure} \centering \includegraphics{figure12.eps} \caption{Values of $\Delta'\left(\alpha,\beta\right)$ at $k=0.71$.} \label{fig:pssSurf} \end{figure} To illustrate persistent anomalous behaviour, it is helpful to define a function analogous to (\ref{eq:Delta1}), \begin{equation}\label{eq:Delta2} \Delta'\left(\alpha,\beta\right)=\vert\eta_{\mathrm{v}}\left(\alpha,\beta\right)-\overline{{\eta}_{\mathrm{v}}}\left(\alpha\right)\vert, \end{equation} where, for each of the values of $\alpha$ considered, $\overline{{\eta}_{\mathrm{v}}}\left(\alpha\right)$ is the median value of \ps evaluated across the range of values of $\beta$. Values of $\Delta'\left(\alpha,\beta\right)$ are shown in figure \ref{fig:pssSurf}, from which it is clear that persistent anomalies appear distributed about a curve in the $\left(\alpha,\beta\right)$ plane. For values of $\alpha$ and $\beta$ away from this curve, the calculations are free of anomalies. Hence, a small change in the values of $\alpha$ or $\beta$ can indeed be shown to successfully avoid persistent anomalous behaviour. \begin{figure} \centering \includegraphics{figure13.eps} \caption{Values of $\eta_{\mathrm{v}}\left(\gamma\right)$ at $k=0.71$.} \label{fig:varyGamma} \end{equation}\end{figure} Finally, we consider briefly that the shielding parameter, $\gamma$, in (\ref{eq:CosOpenChannel}) and (\ref{eq:Chi}) might also be varied in an effort to avoid anomalous behaviour. Values of \ps at $k=0.71$, $\alpha=0.6$ and $\beta=1.0$ for $0.5\leq\gamma\leq1.0$ are shown in figure \ref{fig:varyGamma}. It is apparent that small changes in the value of $\gamma$ have relatively little effect on the persistent anomaly at $k=0.71$. This is not unexpected, being consistent with the findings of Lucchese \cite{Lucchese1989}, who investigated the effect of varying a parameter analogous to $\gamma$ in his model potential calculations. He noted that singularities due to the choice of $\gamma$ occurred when the values of $\gamma$ and $\alpha$ were not similar, most typically when $\gamma\ll\alpha$. With this in mind, and from inspection of figures \ref{fig:pssSurf} and \ref{fig:varyGamma}, we can conclude that the anomaly observed in figure \ref{fig:methComp} is due primarily to the choices of $\alpha$ and $\beta$ rather than the choice of $\gamma$. \section{Concluding remarks} We have carried out a thorough examination of singularities and related anomalous behaviour in generalized Kohn calculations for \ehmol scattering. We have argued that singularities do not always occur spuriously and that variational calculations of the scattering phase shift can be anomaly-free at these singularities. Subsequently, we have developed an optimization scheme for choosing a free parameter of the trial wavefunction allowing anomaly-free values of the phase shift to be determined without the need to solve the linear system of equations derived from the Kohn variational principle. This approach has been seen to be largely successful, giving phase shifts in close agreement with those determined by a conventional generalization of the Kohn method, as well as those obtained with the complex Kohn method. Persistent anomalies in both sets of calculations have been identified and attributed to singularities that cannot be avoided with any choice of the parameter, $\tau$. Further, we have found that our implementation of the complex Kohn method is susceptible to the same behaviour. We have demonstrated, however, that persistent anomalies can be avoided by small changes in the nonlinear parameters of the short-range correlation functions. Hence, by studying the behaviour of $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ over $k$, we can predict the appearance of persistent anomalous behaviour quantitatively and avoid it by an appropriate change in $\alpha$ or $\beta$. \ack We wish to thank John Humberston for valuable discussions. This work is supported by EPSRC (UK) grant EP/C548019/1. \section*{References} \begin{thebibliography}{99} \bibitem{Kohn1948} Kohn W 1948 \PR{\bf 74} 1763--72 \bibitem{Nesbet1980} Nesbet R K 1980 {\it Variational Methods in Electron-Atom Scattering Theory} (New York: Plenum) \bibitem{MasseyRidley1956} Massey H S W and Ridley R O 1956 {\it Proc. Phys. Soc.} A {\bf 69} 659--67 \bibitem{Schneider1988} Schneider B I and Rescigno T N 1988 \PR A {\bf 37} 3749--54 \bibitem{VanReeth1995} Van Reeth P and Humberston J W 1995 \jpb{\bf 28} L511--7 \bibitem{VanReeth1996} Van Reeth P, Humberston J W, Iwata K, Greaves R G and Surko C M 1996 \jpb{\bf 29} L465--71 \bibitem{VanReeth1999} Van Reeth P and Humberston J W 1999 \jpb{\bf 32} 3651--67 \bibitem{Armour1990} Armour E A G, Baker D J and Plummer M 1990 \jpb{\bf 23} 3057--74 \bibitem{Cooper2007} Cooper J N and Armour E A G 2008 \NIM B {\bf 266} 452--7 \bibitem{CooperArmourPlummer2008} Cooper J N, Armour E A G and Plummer M 2008 \jpb{\bf 41} 245201 \bibitem{Schwartz1961} Schwartz C 1961 \AP{\bf 16} 36--50 \bibitem{Schwartz1961b} Schwartz C 1961 \PR{\bf 124} 1468--71 \bibitem{Nesbet1968} Nesbet R K 1968 \PR{\bf 175} 134--42 \bibitem{Brownstein1968} Brownstein K R and McKinley W A 1968 \PR{\bf 170} 1255--66 \bibitem{Shimamura1971} Shimamura I 1971 \JPSJ{\bf 31} 852--70 \bibitem{Takatsuka1979} Takatsuka K and Fueno T 1979 \PR A {\bf 19} 1011--7 \bibitem{McCurdy1987} McCurdy C W, Rescigno T N and Schneider B I 1987 \PR A {\bf 36} 2061--6 \bibitem{Lucchese1989} Lucchese R R 1989 \PR A {\bf 40} 6879--85 \bibitem{Feshbach1962} Feshbach H 1962 \APNY{\bf 19} 287--313 \bibitem{Chung1971} Chung K T and Chen J C Y 1971 \PRL{\bf 27} 1112--4 \bibitem{CharltonHumberston2005} Charlton M and Humberston J W 2005 {\it Positron Physics} ({\it Cambridge Monographs on Atomic, Molecular and Chemical Physics} vol 11) ed A Dalgarno \etal (Cambridge: Cambridge University Press) \bibitem{Temkin1967} Temkin A and Vasavada K V 1967 \PR{\bf 160} 109--17 \bibitem{Temkin1969} Temkin A, Vasavada K V, Chang E S and Silver A 1969 \PR{\bf 186} 57--66 \bibitem{Flammer1957} Flammer C 1957 {\it Spheroidal Wave Functions} (Stanford: Stanford University Press) \bibitem{Kato1950} Kato T 1950 \PR{\bf 80} 475 \bibitem{Kato1951} Kato T 1951 {\it Prog. Theor. Phys.} {\bf 6} 394--407 \bibitem{BransdenJoachain2003} Bransden B H and Joachain C J 2003 {\it Physics of Atoms and Molecules} (Harlow: Prentice Hall) \bibitem{Hylleraas1929} Hylleraas E A 1929 \ZP{\bf 54} 347--66 \bibitem{Armour1987} Armour E A G and Baker D J 1987 \JPB{\bf 20} 6105--19 \bibitem{Armour1988} Armour E A G 1988 {\it Phys. Rep.} {\bf 169} 1--98 \bibitem{Armour2008} Armour E A G, Todd A C, Jonsell S, Liu Y, Gregory M R and Plummer M 2008 \NIM B {\bf 266} 363--8 \bibitem{Higham2002} Higham N J 2002 {\it Accuracy and Stability of Numerical Algorithms} (Philadelphia: Society for Industrial and Applied Mathematics) \bibitem{ArmourHumberston1991} Armour E A G and Humberston J W 1991 {\it Phys. Rep.} {\bf 204} 165--251 \bibitem{NAG1} http://www.nag.co.uk/numeric/Fl/manual20/pdf/F03/f03aaf.pdf \bibitem{NAG2} http://www.nag.co.uk/numeric/Fl/manual20/pdf/F07/f07agf.pdf \end{thebibliography} \end{document} }\end{figure}}
\caption{Values of $\Im\left[\tau_{\mathrm{s}}\right]$ for $0.64 \leq k \leq 0.67$.} \label{fig:imroot} \end{figure} The second difficulty concerns the choice of $\hat{\tau_{\mathrm{s}}}$ from the two available solutions of (\ref{eq:qeq}). A method is needed for identifying at each $k$ the root of (\ref{eq:qeq}) corresponding to a legitimate singularity. This can easily be achieved by inspecting values of \ps at values of $\tau$ either side of each singularity, although this approach is not ideal as it requires solutions of (\ref{eq:KohnEq}) to be found. In practice, at each $k$ it should be possible to determine by inspection which of the two phase shifts is anomaly-free by examining corresponding results at singularities for nearby values of $k$. For example, figure \ref{fig:sMap} clearly shows that only one curve in the $\left(\tau,k\right)$ plane corresponds to a physically acceptable variation of phase shift over $k$. With these considerations in mind, we claim that choosing $\tau=\hat{\tau}_\mathrm{s}$ at each $k$ defines a consistent optimization that can be used to avoid anomalies due to Schwartz singularities appearing at other values of $\tau$. To evaluate the success of this approach, we have found it helpful to consider an alternative optimization of $\tau$. For the calculations of \ps carried out with $p=1001$, choosing the value of $\tau$ at each $k$ giving rise to the median phase shift, $\langle\eta_{\mathrm{v}}\rangle$, should also mitigate anomalous behaviour. In figure \ref{fig:comparison} we have compared results for $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$ and $\hat{\eta}_{\mathrm{v}}\left(k\right)$ for momenta in the range $0.01\leq k \leq1$. Here, of the two candidates for $\hat{\eta}_{\mathrm{v}}$ corresponding to the two singularities, at each $k$ we have chosen the one whose absolute value is closest to $\vert\langle{\eta}_{\mathrm{v}}\rangle\vert$. For clarity, we have included in the figure values of $\hat{\eta}_{\mathrm{v}}\left(k\right)$ for only $50$ values of $k$ equidistant in the range $0.01 \leq k \leq 0.99$. \begin{figure} \centering \includegraphics{figure8.eps} \caption{A comparison of optimization schemes for $\tau$, [\textcolor{blue}{\full}] $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$ and [\textcolor{myGreen}{$\times$}] $\hat{\eta}_{\mathrm{v}}\left(k\right)$.} \label{fig:comparison} \end{figure} Both optimization schemes successfully avoid anomalous behaviour at most values of $k$ and there is good agreement between the two sets of results at all momenta. However, the intriguing feature of the figure is the anomaly appearing in both sets of results at $k\sim0.71$. For $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$, we have shown this anomaly in greater detail by including in the figure results of a further $100$ Kohn calculations for momenta equidistant in the range $0.7\leq k\leq0.72$, although it is practical to consider henceforth only the anomalous behaviour occurring precisely at $k=0.71$. We believe the anomaly shown in figure \ref{fig:comparison} is of a different type to those shown in figure \ref{fig:psMap}, which are due to singularities found by varying only $\tau$ at a given $k$. In the following section we will examine the circumstances under which persistent anomalies of the kind shown in figure \ref{fig:comparison} could appear, before going on to discuss methods designed to avoid them. \subsection{Persistent anomalous behaviour}\label{ss:persistent} It is often claimed that anomalous results observed in the region of singularities arise from $A$ having a determinant close to zero. Statements of this kind can be misleading, as the determinant of any nonsingular $A$ can be made arbitrarily close to zero by an appropriate scalar multiplication, without altering the sensitivity of the solution, $x$, to small errors in the elements of $A$ or $b$. A better measure for identifying regions where anomalies may occur is the condition number, \cond, defined for nonsingular $A$. The condition number is independent of the normalization constant, $N$, and for the linear system (\ref{eq:KohnEq}) is defined as the maximum ratio of the relative error in $x$ and the relative error in $b$. Formally, it can be shown that \begin{equation}\label{eq:condition} \kappa\left(A\right)=\parallel A\parallel\parallel A^{-1}\parallel, \end{equation} with respect to some matrix norm, $\parallel A \parallel$ \cite{Higham2002}. The value of \cond is dependent upon the choice of norm. In our calculations, we have considered the matrix 1-norm \cite{Higham2002} of $A$, \begin{equation}\label{eq:onenorm} \parallel A \parallel_{1}=\max_{1\leq j\leq\left(M+2\right)}\sum_{i=1}^{M+2}\vert a_{ij}\vert, \end{equation} where $a_{ij}$ is the element in the $i^{\mathrm{th}}$ row and $j^{\mathrm{th}}$ column of $A$. In what follows, the particular choice of the 1-norm in our calculations will implicitly be assumed. A matrix with a large condition number is said to be ill-conditioned, and the solution of the corresponding linear system may not be reliable if the elements of $A$ and $b$ are not known exactly. For any invertible $A$, the condition number can be used to formalize the definition of closeness to singularity in the following way. If $\Delta A$ is defined to be any matrix such that $A+\Delta A$ is singular, then the relative distance to singularity, \dist, for $A$, is defined \cite{Higham2002} to be \begin{equation}\label{eq:distance} \Lambda\left(A\right)=\min\left(\frac{\parallel \Delta A\parallel}{\parallel A\parallel}:\quad \det\left(A+\Delta A\right)=0\right). \end{equation} This definition holds for any consistent norm. Further, if \dist and \cond are evaluated using the same choice of norm, it can be shown \cite{Higham2002} that \begin{equation}\label{eq:distcond} \Lambda\left(A\right)= \left[\kappa\left(A\right)\right]^{-1}. \end{equation} In section \ref{ss:pscalcs} we noted that (\ref{eq:qeq}) has, in general, no more than two zeros if only variations of $\tau\in\left[0,\pi\right)$ are considered. However, if $\mathcal{A}=\mathcal{B}=\mathcal{C}=0$ then \deter is identically zero independently of $\tau$ and no consistent value of phase shift can be calculated, either by solving (\ref{eq:KohnEq}) or directly from (\ref{eq:afm}). There is no obvious physical reason why this circumstance should arise at any $k\in\mathbb{R}$. However, it is conceivable that $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ could coincidentally be close to zero, in some sense, over a narrow range of $k$. Small errors in the evaluation of $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ could then give rise to both unreliable solutions of (\ref{eq:qeq}) and persistent anomalies in the calculation of \ps due to ill-conditioning in the Kohn equations (\ref{eq:KohnEq}). To see how the latter case arises, using (\ref{eq:detan}) we note that \begin{eqnarray} \mathcal{C}&=&\det\left[A\left(\tau=0\right)\right],\\ \mathcal{A}&=&\det\left[A\left(\tau=\frac{\pi}{2}\right)\right],\\ \mathcal{B}&=&2\det\left[A\left(\tau=\frac{\pi}{4}\right)\right]-\mathcal{A}-\mathcal{C}, \end{eqnarray} so that the notion of $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ being close to zero is immediately formalized in terms of \dist at $\tau=0$, $\tau=\pi/4$ and $\tau=\pi/2$. In our calculations, we have used a numerical algorithm \cite{NAG2} to calculate \diste, an estimate of \dist. Values of \diste at $\tau=0$, $\tau=\pi/4$ and $\tau=\pi/2$ are shown in figure \ref{fig:ABC} for $100$ values of $k$ in the range $0.7 \leq k \leq 0.72$. Values of \diste for $\tau=0$ and $\tau=\pi/4$ are anomalously small at $k\sim0.71$, with the values for \diste at $\tau=\pi/2$ also passing through a clear minimum at $k\sim0.711$. Ordinarily, we would not expect ill-conditioning to occur over a very broad range of $\tau$ at a given $k$. At $k=0.71$, the small values of \diste at $\tau=0$, $\tau=\pi/4$ and $\tau=\pi/2$ therefore point to a manifestation of ill-conditioning which is unusually widespread in $\tau$. In fact, in our calculations we have failed to find any value of $\tau\in\left[0,\pi\right)$ such that $A$ is sufficiently well-conditioned to avoid anomalous results at $k=0.71$. We have also confirmed that $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ all pass through zero at least once between $k=0.71$ and $k=0.7104$. \begin{figure} \centering \includegraphics{figure9.eps} \caption{Values of \diste over $k$ for [\textcolor{blue}{\dashed}] $\tau=0$, [\textcolor{myGreen}{\chain}] $\tau=\frac{\pi}{4}$ and [{\dotted}] $\tau=\frac{\pi}{2}$.} \label{fig:ABC} \end{figure} We can conclude that both optimization schemes developed in section \ref{ss:pscalcs} successfully avoid anomalies due to Schwartz singularities whose existence depends only upon the choice of $\tau$. However, at certain values of $k$, $A$ can become close to singularities whose existence is independent of $\tau$, resulting in anomalies in the calculation of \ps that persist even after $\tau$ has been optimized. In attempting to address this problem, a number of methods are available. In principle, anomalous behaviour can be reduced dramatically by allowing the Kohn trial wavefunction to be complex-valued \cite{Schneider1988,McCurdy1987}. An alternative approach for avoiding anomalies persistent in $\tau$ is to make a small change in some other parameter of the trial wavefunction. We will explore both of these techniques in the following section. \subsection{The complex Kohn method} The complex Kohn method is an extension of the original variational approach in which the boundary conditions of the trial wavefunction are complex. It was originally believed \cite{Schneider1988,McCurdy1987} that this method was anomaly-free, although anomalies were subsequently reported by Lucchese \cite{Lucchese1989}. For our complex Kohn calculations on \ehmol scattering, we have used a trial wavefunction, $\Psi_{\mathrm{t}}'$, of the form \begin{equation}\label{eq:complextrialwave} \Psi_{\mathrm{t}}' = \left(\bar{S} + a_{\mathrm{t}}'\bar{T} + p_{0}'\chi_{0}\right)\psi_{\mathrm{G}} + \sum_{i=1}^{M} p_{i}'\chi_{i}, \end{equation} where \begin{equation}\label{eq:hankel} \bar{T}=\bar{S}+\rmi\bar{C}, \end{equation} the functions \targetwave and $\Omega=\{\chi_{1},\dots,\chi_{M}\}$ being the same as in (\ref{eq:trialwave}). The unknowns $a_{\mathrm{t}}'$ and $\{p_{0}',\dots,p_{M}'\}$ will not, in general, be real. Application of the variational principle to (\ref{eq:complextrialwave}) leads to a matrix equation analogous to (\ref{eq:KohnEq}), \begin{equation}\label{eq:CKohnEq} A'x'=-b' \end{equation} where $A'$ and $b'$ are identical to $A$ and $b$, but for the function, $\bar{T}$, replacing $\bar{C}$ in (\ref{eq:matA}) and (\ref{eq:vecB}). The determinant, $\det\left(A'\right)$, conveniently reduces to \begin{equation} \det\left(A'\right) = \mathcal{D}\left(k\right)e^{-2\rmi\tau}, \end{equation} where \begin{equation}\label{eq:DABC} \mathcal{D}=\left(\mathcal{A}-\mathcal{C}\right)-\rmi\mathcal{B} \end{equation} is a complex constant with respect to variations in $\tau$. The values of $\det\left(A'\right)$ then describes a circle of radius $\vert \mathcal{D}\vert$ in the complex plane for variations of $\tau\in\left[0,\pi\right)$. Hence, singularities are obtained only if both the real and imaginary parts of $\mathcal{D}$ are zero; they can neither be located nor avoided by varying only $\tau$. We would therefore expect anomalous results due to singularities arising from the choice of $\tau$ at a given $k$ to be eliminated in the complex Kohn method. However, from (\ref{eq:DABC}) we would also expect $\mathcal{D}$ to be close to zero at $k=0.71$, in the same sense that $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ have already been seen to be close to zero at this value of $k$. It is therefore likely that the anomalies already seen to occur due to relationships between $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ rather than the choice of $\tau$, will persist even in the complex Kohn method. We have obtained values of \ps using the trial function, \ctrialwave. We have found that the differences in the calculated values of \ps at different values of $\tau\in\left[0,\pi\right)$ are negligible, for all positron momenta considered here. Without loss of generality, we can regard the complex Kohn calculation as effectively independent of $\tau$ and choose $\tau=0$ for simplicity. In figure \ref{fig:methComp} we have compared results for $\eta_{\mathrm{v}}\left(k,\tau=0\right)$ obtained with the trial function, \ctrialwave, with the results for $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$ obtained in section \ref{ss:pscalcs} with \trialwave. \begin{figure} \centering \includegraphics{figure10.eps} \caption{A comparison of [\textcolor{blue}{\full}] $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$, determined using the trial function, \trialwave, with [\textcolor{myGreen}{$\times$}] $\eta_{\mathrm{v}}\left(k,\tau=0\right)$, evaluated using the complex Kohn method with the trial function, \ctrialwave.} \label{fig:methComp} \end{figure} The results of the two methods are essentially equivalent at all positron momenta, with the optimized results for \trialwave differing from the results for \ctrialwave by no more than $0.1 \%$ at each $k$. As expected, the complex Kohn method automatically avoids anomalies at most values of $k$ without the need for an optimization of $\tau$. Nevertheless, the use of the complex trial function has predictably failed to remove the persistent anomaly at $k=0.71$. In figure \ref{fig:D} we verify that $\Lambda_{\mathrm{a}}\left(A'\right)$ is anomalously small at $k\sim0.71$ for $\tau=0$. We have found that differences between the results shown in figure \ref{fig:D} and values of $\Lambda_{\mathrm{a}}\left(A'\right)$ calculated at other values of $\tau$ are negligible. \begin{figure} \centering \includegraphics{figure11.eps} \caption{Values of $\Lambda_{\mathrm{a}}\left(A'\right)$ over $k$ for $\tau=0$.} \label{fig:D} \end{figure} Having failed to find a systematic remedy for the persistent anomalous behaviour, we consider a more {\it ad hoc} approach. It should be possible to avoid any Schwartz-type anomaly by some variation of parameters in the trial wavefunction. We have found that variations in $\tau$ are not always successful, but other candidates exist. In our complex Kohn calculations, we have varied the values of $\alpha$ and $\beta$ in (\ref{eq:srcf}), fixing $\tau=0$. Recall that the values of the these parameters have so far remained fixed at $\alpha=0.6$ and $\beta=1.0$. We now consider the results of Kohn calculations carried out for $31$ different values of $\alpha$ in the range $\alpha\in\left[0.59,0.605\right]$ and $61$ different values of $\beta$ in the range $\beta\in\left[0.65,1.25\right]$. \begin{figure} \centering \includegraphics{figure12.eps} \caption{Values of $\Delta'\left(\alpha,\beta\right)$ at $k=0.71$.} \label{fig:pssSurf} \end{figure} To illustrate persistent anomalous behaviour, it is helpful to define a function analogous to (\ref{eq:Delta1}), \begin{equation}\label{eq:Delta2} \Delta'\left(\alpha,\beta\right)=\vert\eta_{\mathrm{v}}\left(\alpha,\beta\right)-\overline{{\eta}_{\mathrm{v}}}\left(\alpha\right)\vert, \end{equation} where, for each of the values of $\alpha$ considered, $\overline{{\eta}_{\mathrm{v}}}\left(\alpha\right)$ is the median value of \ps evaluated across the range of values of $\beta$. Values of $\Delta'\left(\alpha,\beta\right)$ are shown in figure \ref{fig:pssSurf}, from which it is clear that persistent anomalies appear distributed about a curve in the $\left(\alpha,\beta\right)$ plane. For values of $\alpha$ and $\beta$ away from this curve, the calculations are free of anomalies. Hence, a small change in the values of $\alpha$ or $\beta$ can indeed be shown to successfully avoid persistent anomalous behaviour. \begin{figure} \centering \includegraphics{figure13.eps} \caption{Values of $\eta_{\mathrm{v}}\left(\gamma\right)$ at $k=0.71$.} \label{fig:varyGamma} \end{equation}\end{figure} Finally, we consider briefly that the shielding parameter, $\gamma$, in (\ref{eq:CosOpenChannel}) and (\ref{eq:Chi}) might also be varied in an effort to avoid anomalous behaviour. Values of \ps at $k=0.71$, $\alpha=0.6$ and $\beta=1.0$ for $0.5\leq\gamma\leq1.0$ are shown in figure \ref{fig:varyGamma}. It is apparent that small changes in the value of $\gamma$ have relatively little effect on the persistent anomaly at $k=0.71$. This is not unexpected, being consistent with the findings of Lucchese \cite{Lucchese1989}, who investigated the effect of varying a parameter analogous to $\gamma$ in his model potential calculations. He noted that singularities due to the choice of $\gamma$ occurred when the values of $\gamma$ and $\alpha$ were not similar, most typically when $\gamma\ll\alpha$. With this in mind, and from inspection of figures \ref{fig:pssSurf} and \ref{fig:varyGamma}, we can conclude that the anomaly observed in figure \ref{fig:methComp} is due primarily to the choices of $\alpha$ and $\beta$ rather than the choice of $\gamma$. \section{Concluding remarks} We have carried out a thorough examination of singularities and related anomalous behaviour in generalized Kohn calculations for \ehmol scattering. We have argued that singularities do not always occur spuriously and that variational calculations of the scattering phase shift can be anomaly-free at these singularities. Subsequently, we have developed an optimization scheme for choosing a free parameter of the trial wavefunction allowing anomaly-free values of the phase shift to be determined without the need to solve the linear system of equations derived from the Kohn variational principle. This approach has been seen to be largely successful, giving phase shifts in close agreement with those determined by a conventional generalization of the Kohn method, as well as those obtained with the complex Kohn method. Persistent anomalies in both sets of calculations have been identified and attributed to singularities that cannot be avoided with any choice of the parameter, $\tau$. Further, we have found that our implementation of the complex Kohn method is susceptible to the same behaviour. We have demonstrated, however, that persistent anomalies can be avoided by small changes in the nonlinear parameters of the short-range correlation functions. Hence, by studying the behaviour of $\mathcal{A}$, $\mathcal{B}$ and $\mathcal{C}$ over $k$, we can predict the appearance of persistent anomalous behaviour quantitatively and avoid it by an appropriate change in $\alpha$ or $\beta$. \ack We wish to thank John Humberston for valuable discussions. This work is supported by EPSRC (UK) grant EP/C548019/1. \section*{References} \begin{thebibliography}{99} \bibitem{Kohn1948} Kohn W 1948 \PR{\bf 74} 1763--72 \bibitem{Nesbet1980} Nesbet R K 1980 {\it Variational Methods in Electron-Atom Scattering Theory} (New York: Plenum) \bibitem{MasseyRidley1956} Massey H S W and Ridley R O 1956 {\it Proc. Phys. Soc.} A {\bf 69} 659--67 \bibitem{Schneider1988} Schneider B I and Rescigno T N 1988 \PR A {\bf 37} 3749--54 \bibitem{VanReeth1995} Van Reeth P and Humberston J W 1995 \jpb{\bf 28} L511--7 \bibitem{VanReeth1996} Van Reeth P, Humberston J W, Iwata K, Greaves R G and Surko C M 1996 \jpb{\bf 29} L465--71 \bibitem{VanReeth1999} Van Reeth P and Humberston J W 1999 \jpb{\bf 32} 3651--67 \bibitem{Armour1990} Armour E A G, Baker D J and Plummer M 1990 \jpb{\bf 23} 3057--74 \bibitem{Cooper2007} Cooper J N and Armour E A G 2008 \NIM B {\bf 266} 452--7 \bibitem{CooperArmourPlummer2008} Cooper J N, Armour E A G and Plummer M 2008 \jpb{\bf 41} 245201 \bibitem{Schwartz1961} Schwartz C 1961 \AP{\bf 16} 36--50 \bibitem{Schwartz1961b} Schwartz C 1961 \PR{\bf 124} 1468--71 \bibitem{Nesbet1968} Nesbet R K 1968 \PR{\bf 175} 134--42 \bibitem{Brownstein1968} Brownstein K R and McKinley W A 1968 \PR{\bf 170} 1255--66 \bibitem{Shimamura1971} Shimamura I 1971 \JPSJ{\bf 31} 852--70 \bibitem{Takatsuka1979} Takatsuka K and Fueno T 1979 \PR A {\bf 19} 1011--7 \bibitem{McCurdy1987} McCurdy C W, Rescigno T N and Schneider B I 1987 \PR A {\bf 36} 2061--6 \bibitem{Lucchese1989} Lucchese R R 1989 \PR A {\bf 40} 6879--85 \bibitem{Feshbach1962} Feshbach H 1962 \APNY{\bf 19} 287--313 \bibitem{Chung1971} Chung K T and Chen J C Y 1971 \PRL{\bf 27} 1112--4 \bibitem{CharltonHumberston2005} Charlton M and Humberston J W 2005 {\it Positron Physics} ({\it Cambridge Monographs on Atomic, Molecular and Chemical Physics} vol 11) ed A Dalgarno \etal (Cambridge: Cambridge University Press) \bibitem{Temkin1967} Temkin A and Vasavada K V 1967 \PR{\bf 160} 109--17 \bibitem{Temkin1969} Temkin A, Vasavada K V, Chang E S and Silver A 1969 \PR{\bf 186} 57--66 \bibitem{Flammer1957} Flammer C 1957 {\it Spheroidal Wave Functions} (Stanford: Stanford University Press) \bibitem{Kato1950} Kato T 1950 \PR{\bf 80} 475 \bibitem{Kato1951} Kato T 1951 {\it Prog. Theor. Phys.} {\bf 6} 394--407 \bibitem{BransdenJoachain2003} Bransden B H and Joachain C J 2003 {\it Physics of Atoms and Molecules} (Harlow: Prentice Hall) \bibitem{Hylleraas1929} Hylleraas E A 1929 \ZP{\bf 54} 347--66 \bibitem{Armour1987} Armour E A G and Baker D J 1987 \JPB{\bf 20} 6105--19 \bibitem{Armour1988} Armour E A G 1988 {\it Phys. Rep.} {\bf 169} 1--98 \bibitem{Armour2008} Armour E A G, Todd A C, Jonsell S, Liu Y, Gregory M R and Plummer M 2008 \NIM B {\bf 266} 363--8 \bibitem{Higham2002} Higham N J 2002 {\it Accuracy and Stability of Numerical Algorithms} (Philadelphia: Society for Industrial and Applied Mathematics) \bibitem{ArmourHumberston1991} Armour E A G and Humberston J W 1991 {\it Phys. Rep.} {\bf 204} 165--251 \bibitem{NAG1} http://www.nag.co.uk/numeric/Fl/manual20/pdf/F03/f03aaf.pdf \bibitem{NAG2} http://www.nag.co.uk/numeric/Fl/manual20/pdf/F07/f07agf.pdf \end{thebibliography} \end{document} }
\caption{Values of \diste over $k$ for [\textcolor{blue}{\dashed}] $\tau=0$, [\textcolor{myGreen}{\chain}] $\tau=\frac{\pi}{4}$ and [{\dotted}] $\tau=\frac{\pi}{2}$.}
\caption{A comparison of [\textcolor{blue}{\full}] $\langle{\eta}_{\mathrm{v}}\rangle\left(k\right)$, determined using the trial function, \trialwave, with [\textcolor{myGreen}{$\times$}] $\eta_{\mathrm{v}}\left(k,\tau=0\right)$, evaluated using the complex Kohn method with the trial function, \ctrialwave.}
\caption{\label{examples} The above matrices are irreducible representations in the Young-Yamanouchi basis with Young diagram \protect \includegraphics[width=0.18in]{tetris2.eps}. Here $\sigma_i$ is the permutation in $S_4$ that swaps $i$ with $i+1$.}
\caption{Phase diagram of {\laf} as a function of the doping level $x$ and temperature $T$, highlighting the unusual $\rho(T)$ as compared to that of the (approximately) linear $\rho(T)$ near 300~K. The latter ($\frac{d\tilde \rho}{dT}(T)=\frac{d\rho}{dT}(T)-\frac{d\rho}{dT}(\rm 296~K)\approx0$) gives rise to the yellow areas. The blue regions ($d\tilde \rho/dT<0$) indicate carrier localization/fluctuation and Fermi liquid-like behavior for $x\leq0.075$ and $x\geq0.1$, respectively. Across the whole phase diagram the red areas ($d\tilde \rho/dT>0$) are centered around $T_\mathrm{drop}$ and mark the signatures of the structural/magnetic transitions ($x\leq0.04$) and the corresponding remnant feature ($x\geq0.05$). The dark bars separate the non-superconducting, the underdoped and the overdoped superconducting regimes. The diagram shows also data points for \tc (\textbigcircle), $T_\mathrm{drop}$ (\textbullet), $T_\mathrm{max}$ ($\Diamond$), and, where available, \tn ($\triangle$) and \ts ($\square$) from $\mu$SR and XRD experiments \cite{Luetkens2009a}. }
\caption{Phase diagram of {\smf}. As in fig.~\ref{fig:La_phasendiagramm}, the diagram highlights where $d\tilde \rho/dT$ is larger (red), smaller (blue) or equal to the slope near 300~K (yellow). % This corresponds to the idea to highlight the unusual temperature dependence of $\rho$ as compared to that of an ordinary metal with linear $\rho(T)$. The dark bar marks the transition from the non-superconducting to the superconducting region. The diagram shows also data points for \tc (\textbigcircle), $T_\mathrm{drop}$ (\textbullet), $T_\mathrm{max}$ ($\Diamond$) extracted from the resistivity, and \ts ($\square$) from XRD experiments \cite{Hamann2008}. Note, that at $x=0.04$ $T_\mathrm{max}$ has been extracted after subtracting a linear background which matches $\rho(T)$ near 300~K.}
\caption{\label{fig:size_distribution}(Color online) Size distribution of small components obtained by numerical simulations (symbols) compared with the theoretical prediction of \eqref{eq:K} (lines). $\triangle$ and $\lozenge$ correspond to the number of type-1 nodes in small components $\big(\text{generated by }K(x_1,1;\mathbf{T})\big)$ for $\gamma=0.1$ and $\gamma=0.5$ respectively. $\bigcirc$ and $\square$ are the equivalent quantities but for type-2 nodes $\big(\text{generated by }K(1,x_2;\mathbf{T})\big)$.}
\caption{\label{fig:e2-12} $\epsilon_N/(N-1)$, where $\epsilon_N$ is the $N$-body energy problem as a function of $G N/2$, where $G$ is the coupling. Solid curve: exact 2-body energy $\epsilon_2$ which is also a lower bound to $\epsilon_N/(N-1)$ for $N>2$. Dashed curve: Gaussian variational approximation $\tilde\epsilon_N/(N-1)$, valid for any $N$. {\magenta{$\diamond$}}:variational hyperscalar approximation $\bar{\epsilon}_{3}/2$ for the 3-body case. {\red{$\bullet$}}: hyperscalar approximation $\bar{\epsilon}_{12}/11$ for the 12-body case. }
\caption{Schematic view of the experimental setup used for the rotary echo experiment in the magnetic trap (a) and the dipole trap (ODT) (b). Figure (c) shows the setup for the EIT experiment in an ODT. The ODT beam is depicted in black and propagating along the $z$-direction. The laser is linearly polarized along the $x$-axis and far red detuned (\unit[826]{nm}) with respect to the ground state. For the rotary echo sequence in the magnetic trap (dashed ellipse) both excitation lasers are collinear and propagating along the -$z$-direction. The laser for the lower transition (see \fref{fig:rotesEIT} (a)) of the two-photon excitation into the \ftS\, via the\fiveP , shown in red, has a wavelength of \unit[780]{nm} and a diameter of $~\unit[1]{mm}$. The laser for the excitation into the \ftS\, Rydberg state is shown in blue and has a wavelength of\unit[480]{nm} and a waist of $\unit[42]{\mu m}$. The lasers are $\sigma^+$ (\unit[780]{nm}) and $\sigma^-$ (\unit[480]{nm}) polarized with respect to the quantization axis (green arrow) given by a small magnetic field along the $z$-direction. In the ODT the quantization is chosen along $x$, while the \unit[780]{nm} beam is traveling along -$x$. For the EIT sequence a quantization axis along the $y$-direction is chosen. The \unit[780]{nm} beam is $\sigma^+$ polarized and imaged with a CCD camera, which is also used in the rotary echo sequence to image the ground state atoms after the Rydberg excitation. The Rydberg atoms are field ionized and detected by a multichannel plate.}
\caption{Collection of the dephasing rates $\subt{\gamma}{d}$ as a function of the measured maximal Rydberg atom number $\subt{N}{R}$. All figures are plotted double logarithmic. Figure (a) shows the results from the rotary echo measurements in the magnetic trap (\textcolor{blue}{\pmb{$\square$}}) and in the ODT (\textcolor{red}{\pmb{$\square$}}). Figure (b) shows the dephasing rates $\subt{\gamma}{d}=\subt{\gamma}{rd}$ obtained from the EIT measurements presented in figure \ref{fig:rotesEIT} (\textcolor{red}{\pmb{$\bigcirc$}}) and figure \ref{fig:blauesEIT} (\textcolor{blue}{\pmb{$\bigcirc$}}). Theoretically investigated dephasing rates for numerically calculated rotary echo experiments are shown in figure (c). The dashed line in figure (c) is a fit to the theoretical data, whilst the dashed lines in (a) and (b) have the same slope as in (c), but fitted offsets (see text).}
\caption{(A) Normalized short-time diffusion coefficient $D_S(Q)\eta_r/D_0$, (B) structure factor $S(Q)$ and (C) hydrodynamic function $H(Q)$ as a function of polymer concentration $c_{p}/c_{p}^{*}$ for different scattering vectors $QR=0.7$ (\blue{$\bullet$}), $1.35$ (\green{$\blacksquare$}), $1.9$ (\red{$\blacktriangle$}) and $2.35$ (\violet{$\blacklozenge$}).}
\caption{(A) Normalized long-time diffusion coefficient $D_L(Q)\eta_r/D_0$ and (B) stretching exponents $\beta$ as a function of polymer concentration $c_p/c_p^*$ for different scattering vectors $QR=0.7$ (\blue{$\bullet$}), $1.35$ ((\green{$\blacksquare$}), $1.9$ (\red{$\blacktriangle$}) and $2.35$ (\violet{$\blacklozenge$}).}
\caption{ (a) Phase diagram and volume ($V$) of $\rm Ba_{1-x}K_xBiO_3$ upon doping throughout the insulator-metal transition. The computed volumes are compared to neutron powder diffraction data\cite{pei}. (b) Schematic structure of the Bi-O plane for the insulating charge ordered state of pure BaBiO$_3$ ($x=0$), the bipolaronic phases at $x=0.125$ and $x=0.25$ and the metallic regime at $x=0.5$, showing the local oxygen ($\circ$) breathing environment around $\rm Bi^{3+}$ (\color{Orange}{$\bullet$}\color{black}), $\rm Bi^{5+}$ (\color{blue}{$\bullet$}\color{black}), bipolaron $\rm Bi^{3+} \to \rm Bi^{5+}$ (\color{red}{$\bullet$}\color{black}), and $\rm Bi^{4.5+}$ (\color{green}{$\bullet$}\color{black}) sites. Bi-O bond lengths between different sites are displayed. }
\caption{For members of the cluster and its outskirts, a histogram of the deviation of LDP redshifts (\zldp) from redshifts measured from higher resolution spectra (\zspec). The biweight computed standard deviation of $z_{\rm LDP}-z_{\rm spec}$ is $\sigma = 0.018$ and the offset from $z_{\rm LDP}-z_{\rm spec} = 0$ is quite small, with $\Delta z = -0.013$. The good agreement between \zspec\and\zldp\establishes the LDP in IMACS as a powerful tool in studying large quantities of galaxies at intermediate redshift.}
\caption{Spectroscopic completeness of our survey of \rxj\as a function of\Rz\and\ifil\(Note that the \Rz\color represents a Vega magnitude minus an AB magnitude. See\S~\ref{kcorrection} for a conversion between the two systems for the \Rfil\filter). The completeness fraction in a given color-magnitude bin is given by$N_z/N$, where $N_z$ is the number of galaxies with measured redshifts and $N$ is the total number of galaxies targeted from the Suprime-Cam photometry catalog with \ifil~$<23.75$~mag. Note that the first 3 slitmasks selected galaxies with \ifil~$<23$~mag while the last 2 selected galaxies down to the fainter limit of \ifil~$<23.75$~mag. The vertical dashed lines indicate these magnitude limits. The \ntot\galaxies defined to be members of the cluster superstructure are overplotted. Note that the spectroscopic completeness for galaxies on the red-sequence near\Rz~$\sim 1.6$ is above $\sim 50\%$ for \ifil~$<23$~mag. At brighter magnitudes along the red-sequence, the spectroscopic completeness quickly ramps up to $>80\%$. Overall, our spectroscopic completeness is high and the variations in this map are accounted for in our calculations in order to minimize any biases in our results.}
\caption{Color vs. stellar mass for galaxies in \rxj\and its outskirts (a). The black diagonal dashed and solid lines represent the derived limits of our spectroscopic selection criteria of\ifil~=~23~mag and \ifil~=~23.75~mag respectively. The dashed vertical line indicates our mass cut of \mcutrange\above which we are$>$50\% complete. This completeness limit can be seen on the red-sequence in Figure~\ref{fig_completeness} at \ifil~$\sim 23$~mag. The solid red line represents the color-mass relation (CMR) for red galaxies (see \S~\ref{colorfractions}). The 2$\sigma$ scatter for this relation separates red galaxies from blue ({\em dashed red line}). The 2 known AGN detected in X-ray observations are circled \citep{demarco2005}. Gray data points above the mass limit are ignored in our analysis because their colors or magnitudes are contaminated by nearby stars. The histogram of $\log M/$\msun\(b) shows the distribution of stellar masses, derived using the \citet{bell2003} relation between rest-frame $B-V$ color and $B$-band mass-to-light ratio. Our mass-limited sample consists of \nmcut\galaxies,$73\%$ of which are red. This dataset provides a large, statistically robust sample from which we can assess the role of environment in evolving massive galaxies.}
\caption{Rest-frame \ugz\color as a function of projected clustercentric radius. Red points represent galaxies with mass\hirange, orange points \midrange\and green points\lorange. Note that most of the data points for the lowest mass galaxies at $R \la 1$~Mpc are not visible but are located on the red-sequence. The colored lines correspond to a smoothed median color for galaxies in these three mass bins. The highest mass galaxies are red at all radii while lower mass galaxies span a broader range in colors at $R \ga 1$~Mpc. Given that galaxies in the outskirts will fall into the cluster core, where almost all galaxies are red, many will have to transition from bluer colors, particularly galaxies in the mass range \lorange.}
\caption{Fraction of galaxies with mass \mcutrange\on the red-sequence ({\em red circles}) or in the blue cloud ({\em blue circles}) as a function of projected clustercentric radius ($a$). Open circles represent galaxies within the two cores. Colored lines represent the color fraction at a given radius for the nearest 30 galaxies. The blue and red galaxy trends are complementary, since all galaxies are defined to be one of the two. The shaded band indicates a red galaxy fraction of $61 \pm 3\%$ for a $z \sim 0.8$ field sample from \citet{vanderwel2007b} adjusted for the rest-frame \ug\color used in this work. The red galaxy fraction is$>90\%$ in the two cores and drops to a level consistent with that of the field at $R \ga 3$~Mpc. This figure reinforces what we see in Figure~\ref{crr}, namely that the two merging cores are dominated by red galaxies, while the outer regions have more blue galaxies. As galaxies at larger radii fall into the central regions of the cluster, the blue population must be transformed into a red one. In the bottom panel ($b$), we plot local galaxy density ($\Sigma$) vs. clustercentric radius. Note that for $R \la 2$~Mpc there is a correlation between $\Sigma$ and $R$. At larger radii however, the correlation is non-existent. Consequently, at these larger radii ($R \ga 3$~Mpc) one can isolate the impact of the local environment (i.e. $\Sigma$) on galaxy properties. For example, the fluctuations in the color fractions in ($a$) at $R > 3$~Mpc may be due to the density peaks at $R \sim 4$~Mpc and $R \sim 6$~Mpc seen in ($b$).}
\caption{Rest-frame \ugz\color vs. local density for galaxies with mass\mcutrange\in\rxj\and its outskirts ($a$). The color coding is the same as in Fig.~\ref{crr}: red points represent galaxies with mass \hirange, orange points \midrange\and green points\lorange. The solid colored lines correspond to a smoothed median color for galaxies in the three mass bins. The highest mass galaxies are on the red-sequence at all densities while lower mass galaxies have a broader range in color at lower densities (cf. Figure~\ref{crr}). The fraction of galaxies with red or blue colors as a function of local density is shown in ($b$). Solid circles are observed color fractions while open circles represent the color fractions after statistically subtracting field interlopers from the sample. We use galaxies in the lowest density bin at $R>3$~Mpc for the field interloper sample. For this reason, no corrected value is shown for the lowest density bin in this sample. The number of galaxies in each density bin is indicated at the bottom of the figure. The increasing fraction of red galaxies with local density shows the expected correlation between local density and clustercentric radius at $R<2$~Mpc.}
\caption{(color online). Pulsed measurements: normalized fringe visibility for $z$ = 219($\scriptstyle{\blacksquare}$), 119(\textcolor{blue}{$\blacktriangle$}) and 9(\textcolor{red}{$\bullet$}) mm, as a function of the plane mirror translation $\Delta x$. The experimental data are fitted with a Gaussian curve.}
\caption{ Two-dimensional PDFs of the sky position for the MCMC runs as labelled. The colours show the different probability intervals (1-$\sigma$, 2-$\sigma$ and 3-$\sigma$ for red, yellow and blue respectively). The black dashed lines mark the position in the sky of the injection for each run. Left column {\bf (a)}: results for the reference runs, experiment 1 (signal injection at R.A.\= 14.3\,h, Dec = 11.5$^\circ$). The symbol \opencircle denotes the ``gap'' discussed in the text. Right column {\bf (b)}: results for experiment 3: an MCMC run with a signal injection at \opencircle (R.A.\= 13.25\,h, Dec = 23$^\circ$). For the non-spinning case, the PDFs are very similar to those in the original run, whereas they are very different for the spinning cases (notice the difference in the axis ranges). \label{fig:sky} }
\caption{(\textit{Color Online}) (a) The logarithm of the probabilities of forward and time-reversed trajectories as a function of work. It can be seen that all data of different work and different control protocols ($N=5$ (red $\bullet$), $N=10$ (blue $\bigtriangleup$), $N=15$ (green $\square$), and $N=20$ (black $\bigcirc$) ) collapse onto the same straight line. The slop of the line is equal to unity, and the line cross the horizontal axes at $W= \Delta F$. Thus the numerical result verifies the quantum Crooks FT $\ln \left[ P_{F}(W|_{a})/P_{R}(-W|_{-a})\right]=\beta(a- \Delta F)$. (b) The averaged work VS. the logarithm of averaged exponent work for different control protocols. It can be seen that the averaged work $\left\langle W \right\rangle$ (red $\bigcirc$) is always greater than the difference of free energy $\Delta F_{AB}$ and differ from one control protocol to another, while the logarithm of the exponentially averaged work $\ln \left\langle\exp[-\beta W]\right\rangle$ (blue $\square$) is always equivalent to the difference of free energy irrespective of the control protocols. Thus the numerical result verifies the JE $\ln{\left \langle \exp[-\beta W]\right \rangle} \equiv \Delta F$.}