\caption{\small (Color online) Group ({\color{black} ---}) and phase ({\color{red} - -}) velocities of the plane wave for $\beta=0$.}
\caption{\small (Color online) PDF of the KEE ({\color{blue} ---}) vs. PDF of the cubic NLSE ({\color{red} - . -}) for $k=0.1$.}
\caption{\small (Color online) PDF of the KEE ({\color{blue} ---}) vs. PDF of the cubic NLSE ({\color{red} - . -}) for $k=0.8$.}
\caption{\small (Color online) PDF of the KEE ({\color{blue} ---}) vs. PDF of the cubic NLSE ({\color{red} - . -}) for $k=3.2$ .}
\caption{\small (Color online) PDF of the KEE for $\beta=0.67$ ({\color{blue} ---}) vs. PDF of the KEE for $\beta=-0.67$ ({\color{red} - . -}).}
\caption{\small (Color online) PDF of the KEE ({\color{blue} ---}) vs. PDF of the quintic NLSE ({\color{red} - . -}).}
\caption{Effect of an AC field on a supported bilayer. Filled symbols are data from this work. Black squares (\textcolor{black}{$\blacksquare$}), green circles (\textcolor{green}{$\bullet$}) and red triangles (\textcolor{red}{$\blacktriangle$}) correspond to different experiments. Solid lines correspond to the full electrostatic contribution (Poisson-Boltzmann) to $\Gamma_{el}$ and $K_{el}$ which can be decomposed in a membrane contribution (dotted line) and an electric double layer contribution (dashed line). Linear Debye-H\"uckel theory is shown as dashed-dotted lines.$\kappa_D^{-1}=800$ nm (red curves), $\kappa_D^{-1}=300$ nm (green curves) and $\kappa_D^{-1}=150$ nm (black curves). Mechanical pressure as a function of distance (A). Blue stars are from Ref. \cite{Hemmerle(PNAS2012)} where pressure was applied by osmotic stress on similar double bilayers. Empty symbols are from Ref. \cite{petrache(pre1998)} (osmotic stress on multilayer stacks). The solid line is calculated after Ref. \cite{Hemmerle(PNAS2012)} using dispersive, electrostatic and entropic contributions to the inter-bilayer potential. Number of water molecules per lipid $n_w$ as a function of electrostatic pressure $\Pi$ (B). Blue circles (\textcolor{blue}{$\circ$}) were obtained by NMR spectroscopy for osmotically stressed DMPC multilayer stacks\cite{mallikarjunaiah2011solid}. Electrostatic contribution to the membrane tension $\Gamma_{el}$ as a function of frequency for a fixed voltage $V_0=5$ V (C) and as a function of the local voltage $V_{\rm loc}$ at the membrane (D). Electrostatic contribution to the membrane rigidity $K_{el}$ as a function of frequency for a fixed voltage $V_0=5$ V (E) and $K_{el} \kappa_D /k_BT$ as a function of the local potential difference $V_{\rm loc}$ (F). }
\caption{Computation numbers for varying perturbation levels $\bar w\in\{0,.1,.2\}$ and number of obstacles $p=0$ ({\protect \tikz[baseline=-0.5ex]{\protect\node{\protect\includegraphics[width=.5cm]{legend1}};}}), $p=5$ ({\protect \tikz[baseline=-0.5ex]{\protect\node{\protect\includegraphics[width=.5cm]{legend2}};}}), $p=10$ ({\protect \tikz[baseline=-0.5ex]{\protect\node{\protect\includegraphics[width=.5cm]{legend3}};}}). The upper row shows the run-times for the computation of the sets $R_i$. The lower row shows the number of half-spaces used to represent the sets $R_i$.}
\caption{Dimensionless pressure drop as a function of $R_b$. Lines indicate a constant $KC_D$ for the values: $KC_D = 1.3$ ({\Large{\color[rgb]{0.75,0.75,0.75}{$\bullet$}}}), $KC_D = 2.8$ ({\color[rgb]{0.75,0.75,0.75}{$\blacktriangle$}}), $KC_D = 3.6$ ({\color[rgb]{0.55,0.55,0.55}{$\blacktriangledown$}}), and $KC_D = 5.4$ ({\color[rgb]{0.0,0.0,0.0}{$\blacklozenge$}}).}
\caption{Hot-tail runaway density obtained using \CODE{} -- with \blue{(blue, dashed)} and without \blue{(yellow, dash-dotted; red, dotted)} an electric field included during the temperature drop -- and the analytical estimates Eqs. (\ref{eq:Smith18}) and (\ref{eq:Smith19}) \blue{(black, solid)}, for the temperature and E-field evolution in Fig.~\ref{fig:hot-tail_dist}a. \blue{An a) $\xi$-dependent and b) isotropic lower boundary of the runaway region was used}. The collision operator in Eq.~(\ref{eq:C}) was used for the \blue{blue and yellow lines}, whereas its non-relativistic limit was used for the \blue{red and black lines}.}
\caption{The conservation of Hamiltonian when the friction and noise vanish. The time evolution of the cell kinetic energy (green), \redc{$\chi\kt\ln[\det(h)]$ (pink)}, the instantaneous enthalpy (blue) and the Hamiltonian (red) are presented. The kinetic energy and \redc{$\chi\kt\ln[\det(h)]$} use the left y-axis, while the instantaneous enthalpy and the Hamiltonian use the right y-axis, as the arrows in the Figure indicate. The unit of the energy is kJ/mol. }
\caption{(a) The missing mass distribution, with the proton as the missing particle, (b) the polar angle difference of measured and missing proton, (c) the azimuthal angle difference of meson and proton, and (d) the $\gamma\gamma$ invariant mass distribution. The distributions are shown---after all other cuts discussed in the text are applied---for butanol ({\scriptsize$\square$}), carbon({\color{red}$\circ$}), and their difference ({\color{blue}\scriptsize$\triangle$}).}
\caption{Comparison of average mean squared error between reaching trials for all five subjects and predictions made on those trials by four models. \textcolor{red}{Red} bars indicate MSE for the $L_{2}$ norm based \textcolor{red}{minimum jerk model} (\emph{continuous}). MSEs shown in \textcolor{green}{green}, \textcolor{cyan}{cyan}, and \textcolor{purple}{purple} are all based on the $L_{\infty}$ norm, and thus result from \emph{sparse} control signals. They represent \textcolor{green}{minimum jerk} (\emph{sparse}), \textcolor{cyan}{minimum snap}, and \textcolor{purple}{minimum crackle}, respectively. The sparse control signals on average have smaller error than the continuous control signal (resulting from use of the $L_{2}$ norm). In all comparisons but one, the sparse control signals result in lower MSE than the continuous control signal. The one exception is in the case of subject C where we see the the minimum crackle sparse control signal produce a greater error that the $L_{2}$ norm minimum jerk (continuous) control signal, however, the difference between the errors is not statistically significant (as determined by a Wilcoxon rank-sum test).}
\caption{The \textcolor{red}{red} plot shows a minimized jerk control signal as measured by the $L_{2}$ norm. Note that it is parabolic and continuous (non-sparse). The \textcolor{blue}{blue} plot shows a minimized jerk control signal as measured by the $L_{\infty}$ norm. Note the distinct qualitative difference between the two. The $L_{\infty}$ based signal is encoded by a finite number of abrupt changes, which can be encoded . Such a signal can be characterized by Dirac delta functions as show in Figures \ref{fig:pulse} and \ref{fig:spikedUp}. These spikes resemble neuronal spikes. In contrast, the $L_{2}$ norm based signal and has no obvious mapping to the physiology of neurons. (figure originally used in \cite{Yazdani2012})}
\caption{Evaluation using HoG features on HumanEva-I. Positive {\em \% Gain} for each subject is shown in {\bf bold}, and in \textcolor{red}{red} otherwise. In the table, / shows that the values are not available (no training samples); Average gives the averaged {\em \% Gain} for the different motions of the same subject; C1 means image feature are computed only from the first camera; C1+C2+C3 means image features from three cameras are combined in a single descriptor. Columns TGP and HOTGP indicate the mean absolute error while the {\em \% Gain} column indicates the percentage reduction on error.}
\caption{Evaluation using HoG features on HumanEva-I. Positive {\em \% Gain} for each subject is shown in {\bf bold}, and in \textcolor{red}{red} otherwise. In the table, / shows that the values are not available (no training samples); Average gives the averaged {\em \% Gain} for the different motions of the same subject; C1 means image feature are computed only from the first camera; C1+C2+C3 means image features from three cameras are combined in a single descriptor. Columns HSIC and HOHSIC (Gegen.) indicate the mean absolute error while the {\em \% Gain} column indicates the percentage reduction on error.}
\caption{Evaluation using HoG features on HumanEva-I. Positive {\em \% Gain} for each subject is shown in {\bf bold}, and in \textcolor{red}{red} otherwise. In the table, / shows that the values are not available (no training samples); Average gives the averaged {\em \% Gain} for the different motions of the same subject; C1 means image feature are computed only from the first camera; C1+C2+C3 means image features from three cameras are combined in a single descriptor. Columns HSIC and HOHSIC indicate the mean absolute error while the {\em \% Gain} column indicates the percentage reduction on error.}
\caption{Evaluation using HoG features on HumanEva-I. Positive {\em \% Gain} for each subject is shown in {\bf bold}, and in \textcolor{red}{red} otherwise. In the table, / shows that the values are not available (no training samples); Average gives the averaged {\em \% Gain} for the different motions of the same subject; C1 means image feature are computed only from the first camera; C1+C2+C3 means image features from three cameras are combined in a single descriptor. Columns TGP and HOTGP {\bf (Gegen) (1,5)} indicate the mean absolute error while the {\em \% Gain} column indicates the percentage reduction on error.}
\caption{% Two-point velocity correlation functions: (a,d,g) $R_{uu}$; (b,e,h) $R_{vv}$; (c,f,i) $R_{ww}$, in (a-c) streamwise, (d-f) vertical, and (g-i) spanwise directions, for $Re_z=2000$ and $A_{yz}=1$. \dashedtridown, $A_{xz}=1.5$; \solidcirc, 3; \chndottri, $8$. The arrows are in the sense of increasing $A_{xz}$. }
\caption{% Two-point velocity correlation functions: (a) $R_{uu}$, (b) $R_{vv}$, and (c) $R_{ww}$, for $Re_z=2000$ and $A_{xz} = 3$. \solidtridown, $A_{yz}=0.5$; \dashedsquare, 1; \solidcirc, 2; \dashedcirc, 4; \solidtri, 8. }
\caption{ (a-c) The time history of box-averaged squared velocity intensities, normalised by $(SL_z)^2$: \dashed, $u^2(t)$; \solid, $v^2(t)$; \chndot, $w^2(t)$. (a) $A_{xz} = 1.5$; (b) 4; (c) 8. In all cases, $Re_z=2000$, $A_{yz}=1$. The vertical dashed lines mark the top of the box cycle. % (d) Premultiplied frequency spectra of $v^2(t)$ in (a-c), $f E_{v2}(f)$, as functions of the period $T =2\pi/ f$. \solid, (a); \dashed, (b); \chndot, (c). The inset is a zoom of the region $1.3 \le ST \le 1.7$, and defines the amplification in Fig. \ref{fig:pre_fre_tb_g}(c). % }
\caption{ (a) Total spectral energy in the harmonics of the box period in the frequency spectrum of $\bra v^2\ket_V$, normalized with the total energy, and versus the box period. Symbols as in Table~\ref{tab:zones}. % (b) Spectral energy in each of the harmonics of the box period, $T_{sn}= T_s/n_x$ ($n_x=1,2$ and 4), as in (a). Open symbols are for $n_x=1$; grey, $n_x=2$; black, $n_x=4$. %. (c) The amplification $G$ of the individual harmonics of the box period, as defined in the inset in Fig.~\ref{fig:traces}(d). Symbols as in (b). The dashed line is Eq.~(\ref{eq:G}). % (d) Lines without symbols are the segment of the history of integrated intensities in Fig.~\ref{fig:traces}(b), around $St= 35$. For comparison, the lines with symbols are the inviscid linearized solution (see Eq.~\ref{eq:rdt_sol_1}) of the fundamental mode $k_x=2\pi/L_x, k_z=0$, scaled to the same amplitude of the peak of the corresponding mode, $\widehat{v}_{10}$, of DNS. \dashed, $\bra u^2\ket_V$; \solid, $\bra v^2\ket_V$; \chndot, $\bra w^2\ket_V$. }
\caption{ (a) Normalised premultiplied two-dimensional spectrum $k_xk_zE_{vv}/v'^2$. Contours are [0.045:0.02:0.085]. % (b) Two-point correlation function of the vertical velocity, $R_{vv}(r_x,r_y)$. Contours are [0.1:0.2:0.9]. % Shaded contours are a minimal channel~\cite{FloresJimenez2010}: $Re_\tau=1840,\, L_x=\pi h/2,\, L_z=\pi h/4$; \solid, full channel~\cite{HoyasJimenez2006}: $Re_\tau=2003,\, L_x=8\pi h,\, L_z=3\pi h$; \dashed, SS-HST (M32). The spectra for both channels are at $y/h \approx 0.15$, where $Re_{\lambda}\approx 100$ as in the SS-HST case. This is also the reference height for the correlations. Lengths are normalized with the Corrsin scale $L_c=L_\varepsilon {(3/S^*)}^{3/2}$. }
\caption{% (a) P.d.f. of the instantaneous growth rates of the kinetic energy. Lines without symbols are minimal channels~\cite{FloresJimenez2010} in the band $y/L_z\approx 0.13$--0.25: \solid, $(Re_\tau, Re_\lambda)=(1830, 125)$; \chndot, $(1700, 80)$; \dashed, $(950, 90)$. % Lines with symbols are SS-HST from Table~\ref{table:HST_Axz3}: $\circ$, L32; $\vartriangle$, M32; $\triangledown$, H32. % The horizontal bar is the range of growth rates for weak initial shearing of isotropic turbulence. % (b) Temporal autocorrelation function of $\bra v^2\ket_V$. The channels are as in (a), but the SS-HST are now: $\circ$, L32; $\vartriangle$, M32; $\triangledown$, M34. }
\caption{Relative error $u_{e}$ for the streamwise velocity, compared with the corresponding linear RDT solution. % (a) For different grids in $y$, as a function of the CFL. \solidcirc, $N_y=16$; \solidtri 24; \solidsquare, 32; \dashedcirc, 48; \dashedtri, 64. In all cases, $N_x = N_z = 18$. The chaindotted line has slope 3. For other parameters, see text. % (b) As in (a), as a function of $N_y$. \solidcirc, CFL$=0.02$; \solidtri, 0.05; \solidsquare, 0.1; \solidtridown, 0.2; \dashedcirc, 0.4; \dashedtri, 0.6; \dashedsquare, 0.8. The chaindotted line has slope $-6$. }
\caption{% Effect of the grid resolution. Case HOM23U.~\cite{RogersMoin1985} (a) The time evolution of the effective resolutions, $\itbold{k}_\mathrm{max} \eta(t)$. Lines with symbols are $k_x \eta$; without symbols are $k_y \eta$. % (b) Evolution of the energy dissipation rate. $\circ$, HOM23U.~\cite{RogersMoin1985} In both figures, \dashed, fine grid, ($510, 384, 254$); \solid, coarse grid ($126, 192, 126$). }
\caption{ Auto-correlation functions for streamwise (a), vertical (b), and spanwise (c) velocities along $x$- (black), $y$- (red), $z$- (blue). \solid, present DNS ($192\times192\times192$ with CFL=0.6); \dashedcirc, HOM23U in Ref.~\onlinecite{AGARD}, from Rogers {\it et al.}~\cite{RogersMoin1985}. $ St = 10.0 $. }
\caption{(a) Parameter space where a large bubble is entrapped (\textcolor{green}{$\blacktriangle$}) or not (\textcolor{red}{$\CIRCLE$}). All simulations are for a prolate water drop with horizontal axis $a= D_h/D = 0.9$. No large bubble entrapment was observed for other drop aspect ratios when $a$ in 0.95, 1, 1.05 and 1.1. (b) Equivalent spherical diameter of the large air bubble $D_b$, based on its total volume and normalized by the drop diameter ($D_b^* = D_b/D$), for $D =$~5~mm and $a =$~0.9.}
\caption{Time evolution of the vorticity maximum in the main vortex ring shed behind the neck (a), and its location (b), starting near the center and then moving outwards radially. (c) \& (d) show the time evolutions of the its radial and vertical location. The red curve represents the case where the large-bubble is entrapped. The\textbf{+} indicates the time of maximal radial location, corresponding to $t^{*} = $~1.80 (fifth panel in Fig.~\ref{fig:largeBubble}).}
\caption{Time series of the droplet interface, and non-dimensional interface velocity, $\hat{v}_{i}$, showing the evolution of the contact line at atmospheric pressure. This time series is magnified $15$ times compared to Figure~\ref{figure:timeSeriesContact}. The unit vector is shown as (% %\protect\tikzsetnextfilename{fig6e} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend1}};}% %%{\protect\tikz[baseline,yshift=0.2em,trim left=-6pt,trim right=14pt] \protect\node at (0,0) {\ref{legend1}};}% \protect\includegraphics{fig6e} ), and the non-dimensional interface velocity as (% %\protect\tikzsetnextfilename{fig6f} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] at (0,0) {\ref{legend2}};}% %%{\protect\tikz[baseline,yshift=0.2em,trim left=-6pt,trim right=14pt] \protect\node at (0,0) {\ref{legend2}};}% \protect\includegraphics{fig6f} ). Light blue represents the liquid phase, and the axes are in \si{\mu\meter}. The contour lines are iso-velocity lines, $\abs{v}$, and the difference between each line is $2.5 \si{\m.\s^{-1}}$. \label{figure:timeSeriesSlip} }
\caption{a) Interface evolution of droplet impact at $2.5 \si{\m.\s^{-1}}$ with a liquid viscosity 10 times that of ethanol. A maximum in the horizontal velocity ( %\protect\tikzsetnextfilename{fig8a} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend3}};}% %{\protect\tikz[baseline,yshift=0.2em,trim left=0pt,trim right=10pt] \protect\node at (0,0) {\ref{legend3}};}% \protect\includegraphics{fig8a} ) can be observed on the interface early after impact and can be traced to the time of lamella formation. At the moment of lamella formation there is a bifurcation for points following the cusp ( %\protect\tikzsetnextfilename{fig8b} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend4}};}% %{\protect\tikz[baseline,yshift=0.2em,trim left=0pt,trim right=10pt] \protect\node at (0,0) {\ref{legend4}};}% \protect\includegraphics{fig8b} ) and the lamella ( %\protect\tikzsetnextfilename{fig8c} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend5}};}% %{\protect\tikz[baseline,yshift=0.2em,trim left=0pt,trim right=10pt] \protect\node at (0,0) {\ref{legend5}};}% \protect\includegraphics{fig8c} ). b) The radial position plotted as function of time for the same data points as in Figure (a). The $t^{1/2}$ scaling suggests that the flow is dominated by inertia. c) The vertical position of the velocity maximum, lamella and cusp as function of time. \label{uEvolution}}
\caption{a) Non-dimensional vertical position of the velocity maximum as function of non-dimensional time for $2.5 \si{\m.\s^{-1}}$ ( %\protect\tikzsetnextfilename{fig9a} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend6}};}% %%{\protect\tikz[baseline,yshift=0.2em,trim left=0pt,trim right=10pt] \protect\node at (0,0) {\ref{legend6}};}% \protect\includegraphics{fig9a} ), $5.0 \si{\m.\s^{-1}}$ ( %\protect\tikzsetnextfilename{fig9b} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend7}};}% %%{\protect\tikz[baseline,yshift=0.2em,trim left=0pt,trim right=10pt] \protect\node at (0,0) {\ref{legend7}};}% \protect\includegraphics{fig9b} ), and $10.0 \si{\m.\s^{-1}}$ ( %\protect\tikzsetnextfilename{fig9c} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend8}};}% %%{\protect\tikz[baseline,yshift=0.2em,trim left=0pt,trim right=10pt] \protect\node at (0,0) {\ref{legend8}};}% \protect\includegraphics{fig9c} ). The surface tension is that of ethanol and the viscosity is $10$ times that of ethanol. The scaling suggests that at early times, lamella formation is dominated by viscosity. b) Non-dimensional vertical position of the velocity maximum as function of non-dimensional time for $2.5 \si{\m.\s^{-1}}$ ( %\protect\tikzsetnextfilename{fig9d} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend9}};}% %%{\protect\tikz[baseline,yshift=0.2em,trim left=0pt,trim right=10pt] \protect\node at (0,0) {\ref{legend9}};}% \protect\includegraphics{fig9d} ), $5.0 \si{\m.\s^{-1}}$ ( %\protect\tikzsetnextfilename{fig9e} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend10}};}% %%{\protect\tikz[baseline,yshift=0.2em,trim left=0pt,trim right=10pt] \protect\node at (0,0) {\ref{legend10}};}% \protect\includegraphics{fig9e} ), and $10.0 \si{\m.\s^{-1}}$ ( %\protect\tikzsetnextfilename{fig9f} %{\protect\tikz[baseline,yshift=-0.5em,trim left=3pt,trim right=14pt] \protect\node[inner ysep=1pt] (0,0) {\ref{legend11}};}% %%{\protect\tikz[baseline,yshift=0.2em,trim left=0pt,trim right=10pt] \protect\node at (0,0) {\ref{legend11}};}% \protect\includegraphics{fig9f} ). The viscosity is the viscosity of ethanol and the surface tension is $20$ times that of ethanol. The scaling suggests that at late times, the scaling for these droplets is dominated by surface tension. \label{scaling}}
\caption{Subspaces of a schematic two-dimensional phase space containing a fixed point $\mathbf{x}_\star$: \textcolor{new}{perturbations are sampled from the domain of the perturbation probability density $\rho$ (area within dashed line). Their transients return to the fixed point when sampled from the basin of attraction $\mathcal{B}(\textbf{x}_\star)$ (gray area). The set $C$ (white within solid line) is the set of initial conditions which lead to transients that satisfy a given constraint, e.g. on the $x_1$-component. While BS is computed as the fraction of perturbations within the basin of attraction, CBS is the fraction of perturbations within the intersection $C \cap \mathcal{B}(\textbf{x}_\star)$ (striped area). Thus, CBS reflects the stability with respect to perturbations whose transients fulfil a given constraint.}}
\caption{(Color online) \textcolor{new}{Illustrations of the phase space structure of the L63 system defined by \Cref{eq:L63_i,eq:L63_m,eq:L63_f}. Both panels show cross sections, obtained by cutting along the plane containing the origin with normal $(1,1,0)$, of the basins of attraction $\mathcal{B}(\mathbf{x}_*^\pm)$ and their intersections $C^\pm \cap \mathcal{B}$ with the sets $C^\pm$ for $r=15$ (upper panel) and $r=23$ (lower panel). In both panels, the blob-shaped region around each fixed point (black dots) is $C^\pm \cap \mathcal B$. Top panel: the total basin of a fixed point is composed of successive layers: it is given by $C^\pm \cap \mathcal B$ combined both with the region in the respective other half-space ($x>0$ or $x<0$) directly surrounding $C^\pm \cap \mathcal B$ and with the next layer of the same color in the fixed point's half plane. In the lower panel (coloring identical), the fractal structure of the basins, visible as intermingled green and blue sets, is apparent; it is associated with transient chaos. One observes that the fraction of the window covered by the sets $C^\pm$ shrinks as $r$ increases. This illustrates the general behavior observed in the L63 system and quantified by CBS, namely that the volume fraction of the three-dimensional sampling region occupied by $C^\pm$ decreases continuously as $r$ approaches $r_3$. }}
\caption{(Color online) \textcolor{new}{BS (green (upper) line) and CBS (red (lower) line)} of the positive fixed points in the L63 system as the parameter $r$ is varied. From left to right, the vertical lines indicate the appearance of the chaotic set ($r_1$), the boundary crisis ($r_2$) and the stability loss of the fixed point ($r_3$). The difference between the two curves represents the fraction of the basin from where chaotic transients evolve. Note that BS exhibits a jump at $r_2$ which is blurred by the finite simulation time: the simulation stopped before all (increasingly long) transients had reached the fixed point. Because of the system's x-y symmetry the negative fixed point has the same BS and CBS. The grey envelopes represent $\pm 3\sigma$ according to \Cref{eq:sigma_BS}. }
\caption{(Color online) \textcolor{new}{Illustrations of the phase space structure of the Anderies system, \Cref{eq:Andi_i,eq:Andi_f}. Both panels show the (globally attracting) desirable fixed point (black dot) and the set $C$ (red, upper part of triangle) and its complement $\mathcal{B}/C$ (green, lower part of triangle) for $\alpha=0.15$ (upper panel) and $\alpha=0.35$ (lower panel). The white dot at $(0,0.5)$ is the desert state fixed point. One observes that the fraction of the two-dimensional finite phase space covered by the set $C$ shrinks as $\alpha$ increases, in agreement with \Cref{BS_CBS_Anderies}.}}
\caption{(Color online) \textcolor{new}{ BS (straight lines) and CBS (curved red line) of $\mathbf{x}_\star^d$ and $\mathbf{x}_\star^{ud}$ versus the human carbon offtake rate $\alpha\in[0,0.6]$ for $t_{crit} = 90$. BS of $\mathbf{x}_\star^d$ is represented by the green straight line ($\alpha\leq \alpha_{crit}\approx 0.4$), BS of $\mathbf{x}_\star^{ud}$ by the blue straight line ($\alpha\leq \alpha_{crit}$).} At low values of $\alpha$, the desirable state is stable against any strength of perturbation while the desert state is unstable. For $\alpha > 0.03$, an increasing fraction of perturbation-induced trajectories take longer than $t_{crit}$ to return to the desirable fixed point until at $\alpha \approx 0.32$, CBS vanishes. By contrast, BS of both fixed points exhibits a jump at $\alpha_{crit}$ and thus no precursory phenomena can be observed there. The grey envelope represents $\pm 3 \sigma$ according to \Cref{eq:sigma_BS}.}
\caption{\label{zigzag_opg} (Color online) Lowest optical and two-photon gaps in 3-ZGNR versus the inverse of number of unit cells. As the ground state spin value changes on increasing the number of monomer units, the lowest optical gaps and lowest two-photon gaps in both $S=0$ and $S=1$ sectors are plotted. Symbols represent the following: lowest optical gap in singlet space ({\color{Red}$\mathbf{\vartriangle}$}); lowest two-photon gap in singlet space ($\square$); lowest optical gap in triplet space ({\color{Blue}$\blacktriangle$}); lowest two-photon gap in triplet space ({\color{ForestGreen}$\blacksquare$}). Inset: Magnified plot of the lowest optical gaps and its extrapolation in 3-ZGNR systems, 11-16 monomer units.}
\caption{\label{tpg_opg} (Color online) Lowest one-photon (optical) ({\color{Red}$\vartriangle$}) and two-photon ($\square$) gaps in (a) 6-AGNR and (b) 5-AGNR, versus the inverse of the number of unit cells. The ground state remains in both the cases in $S=0$ space. The extrapolated one-photon and two-photon gaps are also indicated in the figures.}
\caption{ (a) Measured ($\bullet$) and simulated (\textcolor{red}{$+$}) diameters of ablation marks produced by 10\,ps pulses with various energies on a glass surface placed at the position of peak air ionization. The dashed curves are fits proportional to the square-root of the pulse energy. (b) Photograph of the ablation mark produced by a 10\,J, 10\,ps pulse. (c) The simulated fluence profile for a 9\,J, 10\,ps pulse. The 5\,J/cm$^{2}$ level corresponding to the ablation threshold fluence for 10\,ps pulse duration is marked with a dashed line. The inset in (a) shows the simulated beam diameter at the 5\,J/cm$^{2}$ level, vs. the distance from the geometrical focus, for a 9\,J, 10\,ps pulse.}
\caption{ (a) Measured ($\bullet$) and simulated (\textcolor{red}{$+$}) average laser fluence at the position of peak air ionization, vs. pulse duration. At any given pulse duration, the average fluence is clamped, i.e., approximately independent of the pulse energy. (b)-(d) Computed fluence profiles at the positions of peak ionization, for various pulse configurations, showing fluence clamping.}
\caption{\label{electrical_scheme} In black the basic electrical circuit for making a magnetic field with a coil L\subs{exp} and a current source. Added are a \textcolor{figorange}{superconducting shortcut} that is brought to its resistive state by a \textcolor{figred}{switch coil L\subs{sw}}. The response is measured by a \textcolor{figgreen}{voltmeter}. The \textcolor{figblue}{blue dashed box} contains the parts that are at low temperature inside a cryostat.}
\caption{In H-KPFM (\textbf{\textcolor{light-blue}{$\blacksquare$}} a,b) an alternating voltage is applied at a frequency $f_{2}-f_{1}$ (\textbf{$\textcolor{red} \downarrow$}). The cantilever's response is {mixed with oscillation at the carrier frequency} in order to be detected at one resonance (${\textcolor{bright-blue} \uparrow}$). {The carrier oscillation occurs at} another resonance and is { also} used to maintain time-averaged distance to the surface ($\mid$). Likewise, in the sideband implementation of FM-KPFM ($\square$, c) a voltage is applied and the response detected at different frequencies: the alternating voltage is applied at $ f_{A} \ll f_{1}$ and detected at $f_{1}+f_{A}$. In AM-KPFM ({\textcolor{light-red}{$\blacksquare$}}), the alternating voltage is applied at the same frequency at which the cantilever response is detected ($\textcolor{purple}{\updownarrow}$). The magnitude of the cantilever transfer function $G(f)$ with each eigenmode modeled as a point-mass, is shown in (e). }
\caption{ Few layer graphene on silicon shows (a) height contrast and (b) significant voltage contrast. (c) A histogram of the KPFM dat shows that the potential distribution is bimodal. (d,e) A watershed algorithm is applied to gradient magnitude of the potential image (d) in order to calculate the boundary (e, {\color{red}$\mathbf{- - -}$}). (f) Voltages are summed as a function of the distance from the boundary ({\color{blue}$\bullet$}) and fit to a tanh function (black solid line), from which the 10-90 resolution is deduced. }
\caption{Partial sums of $^{16}\text{O}$ in HO basis (a) and HF basis (b) for the NN+3N-full interaction with $\alpha=0.08\text{fm}^4$ and truncation parameters $N_{\text{max}}=2$ (\redcircle),\,$4$ (\bluetriangleup) and $6$ (\orangestar). The corresponding energy corrections for each order are displayed in (c) and (d), respectively. All calculations are performed at oscillator frequency $\hbar \Omega =24\,\text{MeV}$.}
\caption{Partial sums for varying flow parameters in HF-MBPT for \elem{He}{4} (a), \elem{O}{16} (b) and \elem{O}{24} (c). The corresponding energy corrections are given in (d), (e) and (f), respectively. The model space for the first and second panel are truncated at $N_{\text{max}}=6$. The truncation for the third panel is given by $N_{\text{max}}=4$. The flow parameters for the different data sets are $\alpha=0.02 \,\text{fm}^4$ (\redcircle), $0.04 \,\text{fm}^4$ (\bluetriangleup) and $0.08 \,\text{fm}^4$(\orangestar). All calculations use a NN+3N-full interaction with oscillator frequency $\hbar \Omega =24\,\text{MeV}$.}
\caption{Panel (a) shows the ground-state energies per nucleon from third-order HF-MBPT (\redcircle) in comparison to $\text{CR-CC}(2,3)$ (\bluetriangleup) results for selected closed-shell nuclei. Panel (b) shows the correlation energy per nucleon, $E_0^{(2)}$ (\redcircleopen) as well as $E_0^{(2)} + E_0^{(3)}$ (\redcircle) for HF-MBPT. Additionally, the correlation energy per nuclei for CCSD (\bluetriangleupopen) and CR-CC(2,3) (\bluetriangleup) are shown. All calculations were performed with the NN+3N-full interaction with $\alpha=0.08\,\mathrm{\text{fm}^4}$, $\hbar \Omega = 24\,\text{MeV}$ in an $e_{\text{max}}=12$ truncated model space. Experimental values are indicated by black bars.}
\caption{Individual contributions of the diagrams appearing at third-order perturbation theory. Show are the contributions per nucleon from the pp-diagram(\reddiamond) , the hh-diagram (\bluestar) and the ph-diagram (\orangetriangleup). The overall contribution of the third-order correction is depicted in (\greencircle). The first panel corresponds to the NN+3N-full interaction and the second panel to a NN+3N-induced interaction with $\alpha=0.08\,\mathrm{\text{fm}^4}$, $\hbar \Omega = 24\,\text{MeV}$, and $e_{\text{max}}=12$.}
\caption{Scheme of the Methodology employed in this study, including the two computational tools FLUKA and ANSYS AUTODYN\textregistered}
\caption{\label{fig:fra-db}Sample annotations from FRA-db, showing \textcolor{magenta}{hands}, \textcolor{green}{faces}, \textcolor{blue}{facial-landmarks} and \textcolor{red}{action objects}. Best viewed online.}
\caption{(Top panel:) The projection of the best-fit function in Figure \ref{fig:func_best} onto the time direction (red histogram). The same projection of the event distribution of Figure \ref{fig:selevt} is given in crosses. The vertical axis is the event rate in the unit of counts s$^{-1}$. (Bottom panel:) The residual of the best\textcolor{red}{-}fit. }
\caption{(Top panel:) The projection of the best-fit function in Figure \ref{fig:func_best} onto the anode direction (red histogram). The same projection of the event distribution of Figure \ref{fig:selevt} is given in crosses. The vertical axis is the event rate in the unit of counts mm$^{-1}$. (Bottom panel:) The residual of the best\textcolor{red}{-}fit. }
\caption{Luminance in $cd/m^{2}$ of Reference Table Patches measured by our PR-655 SpectraScan\textregistered Spectroradiometer. We can observe that their spectral reflectance descreased monotonically from the 1st patch to the 65th.}
\caption{\textcolor{blue}{\(\Delta_p V\leq0\)}, \textcolor{green}{\(\Delta_p V=0\)}, \textcolor{red}{\(\Delta_p V\geq0\)}}
\caption{Average global 4-fold bond-orientational order parameter $\langle |Q_4| \rangle$ versus $\Gamma$ for C1 (a) and C2 (b). Open (solid) symbols represent data obtained for increasing (decreasing) $\Gamma$ ramps, with the following rates: $\Delta\Gamma/\Delta t \approx 0.005$ min$^{-1}$ ({\color{blue} $\triangle$, $\blacktriangledown$}) and $\Delta\Gamma/\Delta t \approx 0.02$ min$^{-1}$ ({\color{red} $\circ$, $\bullet$}). Continuous lines in (b) correspond to fits of the linear trend $Q_4^L = a\Gamma +b$ for $2.5<\Gamma<5$, with $a=0.011\pm0.001$ and $b=0.380\pm0.002$, and a supercritical-like behavior $ \langle |Q_4| \rangle=Q_4^L + c(\Gamma-\Gamma_c)^\beta$, with $\beta=1/2$, observed for $\Gamma \gtrsim 5$. Figure from \citet{Castillo}. \copyright 2012, American Physical Society.}
\caption{(color online) The total spin $z$-component ($T_i^z$) along a finite chain with $L=60$ sites and for different values of $U$. The symbols \textopenbullet, \textcolor{red}{\Square} and \textcolor{blue}{$\triangle$} denote $U/W=0$, $1$ and $3$, respectively. The other parameters are $J_K=W$, $\Delta=2$, $J_H=0.5W$ in all cases.}
\caption{The phase diagram of the anisotropic model as functions of the anisotropy and the Hubbard interaction strength for $J_K=W$ and $J_H=0.5W$. The filled and open circles correspond to Haldane and N\'eel ground state, respectively. The dotted line denotes the approximate phase boundary.}
\caption{\textbf{Reported FRBs to date}: This list contains a total of {\color{blue}sixteen} events among which {\color{blue}\underline{FRB $121102$} and \underline{FRB $110523$} were detected at Arecibo and Green Bank Telescope respectively} while the remaining {\color{blue}fourteen} were all detected at Parkes. The FRB parameters, including the redshifts are taken from the published references listed in the Table. {\color{blue}\protect\cite{ch15} have not mentioned the redshift but provide $DM_{MW}$ from which we have estimated $z$ using eq. (\ref{eq:c3}).} Note that none of the FRBs have a direct redshift measurement, and the $z$ values in this Table have all been inferred from the observed dispersion measure. \protect\cite{BM14} have proposed that FRB $010621^{\star}$ probably has a galactic origin in which case the estimated redshift is not meaningful. However, in our work we assume an extragalactic origin for this FRB and take the estimated redshift to be correct. }
\caption{View of the data along the circumference of the bullet (circular segment of about 30 degrees).\label{fig:sidex3p}}{% \includegraphics[width=.65\textwidth]{sidex3p.png} \hspace{1cm} \includegraphics[width=.25\textwidth]{side-sketch.png} }
\caption{Frontal view of a bullet land impression (lower end of the view is the bottom of the bullet). \label{fig:topx3p}}{% \includegraphics[width=.65\textwidth]{topx3p.png} \hspace{1cm} \includegraphics[width=.15\textwidth]{top-sketch.png} }
\caption{(a) Hexagonal structure of buckled two dimensional lattice where ${\boldsymbol\delta}_{1}=a\left( 1,0,{1}/{2\sqrt{2}} \right)$, ${\boldsymbol\delta}_{2}=\frac{a}{2}\left(-1,{\sqrt{3}},{1}/{\sqrt{2}} \right)$, ${\boldsymbol\delta}_{3}=\frac{a}{2}\left( -1,{-\sqrt{3}},{1}/{\sqrt{2}} \right)$ are the nearest neighbors position vectors, the lattice vectors are ${{\boldsymbol a}_{1}}=\frac{a}{2}\left(3,\sqrt{3} \right)$, ${{\boldsymbol a}_{2}}=\frac{a}{2}\left( 3,-\sqrt{3} \right)$, and the next nearest position vectors are ${{{\boldsymbol \delta }'}_{1}}=\pm{{\boldsymbol a}_{1}}$, ${{{\boldsymbol \delta }'}_{2}}=\pm{{\boldsymbol a}_{2}}$ and ${{{\boldsymbol \delta }'}_{3}}=\pm ({{\boldsymbol a}_{2}}-{{\boldsymbol a}_{1}})$. (b) Side view of buckled structure for silicene. \label{fig1}} \end{figure} Graphene-like materials could be considered as honeycomb lattice structures, similar to graphene meanwhile the SOC of buckled honeycomb structures contains parallel and perpendicular terms. The Hamiltonian of the buckled honeycomb lattice in tight-binding approximation in the presence of SOCs can be written as \begin{eqnarray} H=-t\sum\limits_{\left\langle ij \right\rangle \alpha }{c_{i\alpha }^{\dagger }}{{c}_{j\alpha }}+ i{{t}_{SO}}\sum\limits_{\left\langle \left\langle ij \right\rangle \right\rangle \alpha \beta }{{{u}_{ij}}c_{i\alpha }^{\dagger }\sigma _{\alpha \beta }^{z}{{c}_{j\beta }}} -i{{t}_{intR}}\sum\limits_{\left\langle \left\langle ij \right\rangle \right\rangle \alpha \beta }{{{\mu }_{ij}}\hat{c}_{i\alpha }^{\dagger }\left( \vec{\sigma }\times {{{\vec{d}}}_{ij}} \right)_{\alpha \beta }^{z}{{{\hat{c}}}_{j\beta }}}+ i{{t}_{extR}}\sum\limits_{\left\langle ij \right\rangle \alpha \beta }{\hat{c}_{i\alpha }^{\dagger }\left( \vec{\sigma }\times {{{\vec{d}}}_{ij}} \right)_{\alpha \beta }^{z}{{{\hat{c}}}_{j\beta}}} \end{eqnarray} where the operator ${\mathop{c^\dagger_{j\alpha}}} (\mathop{c}_{j\alpha})$ creates (annihilates) an electron with spin $\alpha$ at site $j$ and $t$ is the nearest neighbor hopping amplitude. The values of these parameters for different materials are given in table~ \ref{table-structure}. The ${t}_{SO}$ is the next-nearest neighbor hopping, ${{u}_{ij}}={{{{\vec{d}}}_{i}}\times{{{\vec{d}}}_{j}}} /{\left|{{{\vec{d}}}_{i}}\times{{{\vec{d}}}_{j}} \right|}$ where ${{\vec{d}}}_{i}$ and ${{\vec{d}}}_{j}$ are the two nearest bonds that connect the next-nearest neighbors, Where ${{u}_{ij}}=1$ if the next-nearest neighbor hopping is counterclockwise and ${{u}_{ij}}=-1$ when it is clockwise with respect to the positive $z$ axis \cite{ezawa2013spin}. The $ {\left\langle\left\langleij \right\rangle \right\rangle} $ run over all the next-nearest neighbor hopping sites and ${\sigma}_{z}$ is the Pauli matrix. ${t}_{\operatorname{int}R}$ and $t_{ext R}$ are the strength of intrinsic and extrinsic Rashba SOCs respectively and $\mu_{ij}=+1(-1)$ stands for the A (B) site. The strength of the external Rashba coupling can be manipulated by an external gate voltage or by the selected substrate. The extrinsic Rashba coupling arises as a result of the inversion symmetry breaking due to an applied perpendicular electric field or interaction with substrate \cite{bychkov1984oscillatory}. Dielectric function and the screening of the charged impurity and also the dynamical polarization which gives collective excitations could be captured by the polarization function $\Pi(\omega,q)$. Dielectric function and collective density oscillations of an electron liquid (plasmons), have been observed in different metals and superconductors \cite{ando1982electronic,giuliani2005quantum}. At the static limit ($\hbar \omega=0$) polarization function gives the screening behavior of the coulomb potential. The dielectric function is relevant to plasmonic studies meanwhile the transport and phonon spectra are also another relevant fields \cite{kaasbjerg2013acoustic}. \begin{figure}[h] \includegraphics[width=0.35\linewidth]{polarization.png} \centering \caption{The bare polarization bubble diagram corresponding to Eq. (\ref{aa}). \label{bubble}} \end{figure} The electron-electron interaction has been considered within the random phase approximation characterizes by the density-density correlation function or polarization function (Fig. \ref{bubble}) \cite{pyatkovski2008polarization,pyatkovskiy2009dynamical, scholz2012dielectric,gorbar2002magnetic,hwang2007dielectric,wunsch2006dynamical, sensarma2010dynamic,scholz2013plasmons,giuliani2005quantum,mahan2000many}. In this approach dielectric function is given by \begin{equation}\label{bb} \epsilon \left( \omega ,\vec{q} \right)=1-V\left( q \right)\Pi \left( \omega ,\vec{q} \right) \end{equation} Where $V\left(q \right)$ is the 2D Coulomb potential, here $V\left( q \right)={2\pi e^2 }/{q}$. Within the Dirac point approximation an effective Coulomb potential could be employed in which $V\left( q \right)={2\pi \alpha }/{q}$ and $\alpha$ is the ratio of coulomb to kinetic energy and named effective fine structure constant where equal to $\alpha ={{{e}^{2}}}/{\left( \hbar {{\varepsilon }_{0}}\upsilon \right)}$ and ${{\varepsilon}_{0}}$ is the bare dielectric constant. Unlike to the graphene where the value of fine structure constant could be determined experimentally in different substrates \cite{hwang2012fermi}, for other buckled honeycomb lattices one can set $\alpha=0.8$ \cite{tabert2014dynamical}. The polarization function in one loop approximation is calculated directly from the bubble diagram that shown in Fig (\ref{bubble}). \begin{eqnarray}\label{aa} \Pi \left(\omega, \vec{q} \right)&=&\sum\limits_{s {s}' k k'}{\frac{f_{k}^{s}-f_{k'}^{{{s}'}}}{\omega +E_{k}^{s}-E_{k'}^{{{s}'}}} |<k' \lambda^{s'}_{k'}\mid e^{iq.r}\mid k \lambda^s_{k}>|^2}\nonumber\\ &=&\sum\limits_{s{s}'k}{\frac{f_{k}^{s}-f_{k+q}^{{{s}'}}}{\omega +E_{k}^{s}-E_{k+q}^{{{s}'}}}{{F}_{s'{s}}}\left(\vec{k}+\vec{q},\vec{k} \right)}, \end{eqnarray} here, the summation performed over the full Brillouin zone and all of the spin and pseudo-spin dependent eigenstates in which $f_{k}^{s}=\frac{1}{\exp \beta \left( E_{k}^{s}-E_F \right)+1}$ is the Fermi distribution function, and $ E_F$ is the Fermi energy. The form factor is given by $F_{s',s}( \vec{k}',\vec{k} )=|<k' \lambda^{s'}_{k'}\mid e^{iq.r}\mid k \lambda^s_{k}>|^2=|<k' \lambda^{s'}_{k'}\mid k \lambda^{s}_{k}>|^2\delta_{\vec{k}',\vec{k}+\vec{q}}$ in which $\mid k \lambda^{s}_{k}>=\mid k>\otimes \mid \lambda^{s}_{k}>$ are the eigenstates of the Hamiltonian where $\mid \lambda^{s}_{k}>$ is the eigenstate in the spin and pseudo-spin subspaces. $\delta_{\vec{k}',\vec{k}+\vec{q}}$ represents the momentum conservation rule as a general condition for contributing transitions. Within the Dirac point approximation the above expression of the polarization function has been assumed to be \cite{wunsch2006dynamical} \begin{eqnarray} \Pi \left(\omega ,\vec{q} \right)=g\sum\limits_{s{s}'k}{\frac{f_{k}^{s}-f_{k+q}^{{{s}'}}}{\omega +E_{k}^{s}-E_{k+q}^{{{s}'}}}{{F}_{{s}'s}}\left(\vec{k}+\vec{q},\vec{k} \right)}\label{Pd} \end{eqnarray} where $g$ is valley degeneracy factor and the summation runs around a single dirac point. Moreover in the absence of the spin-orbit interactions the form factor is reduced to: ${{F}_{s'{s}}}\left(\vec{k}+\vec{q},\vec{k} \right)=\frac{1}{2}[1+s{s}'cos\left( 2\theta \right)]$, with $ \theta$ being the angle between $k$ and $k+q$. It should be noted that in this relation the degeneracy factor, $g$, implies identical contribution of the different Dirac points in the polarization function at given $\vec{q}$. As discussed before the valley degeneracy could result in identical contribution of the Dirac points in scalar quantities such as total energy. Nevertheless this degeneracy cannot indicate the identical contribution of the Dirac points in the wave number dependent quantities such as polarization function. %Therefore in the current %case application of the Dirac point approximation in the Eq. (\ref{aa}) cannot result in the %mentioned expression of the polarization function which has been employed by authors in this %field {\bf{do you think is it necessary to talk about another authors? i think it is better to let the readr's judgment}}. Consequently, this type of calculations should be performed beyond the Dirac point approximation even when the Fermi energy level lies close to the cones intersecting points. This is due to the fact that the contribution of the different Dirac points, $\vec{K}_D$ and $\vec{K}_D'$, is not the same and depends on the location of the Dirac point in the $k$-space that has been identified by $\vec{K}_D$. When the integration has been reduced to the Fermi circle of a single Dirac point this assumption automatically ignores the contribution of the inter-valley transitions in which the initial and final states belong to different Fermi circles. This could take place when the transferred momentum, $q$, satisfies $q\sim\mid\vec{K}_D-\vec{K}_D'\mid$. This means that the general form of the Dirac point approximation ignores the inter-valley transition. The anisotropy of the dielectric function results from this type of transitions when the Fermi energy is exactly zero. Within a second-order perturbation approach it has been realized that the exchange interaction of the localized spins $S_1$ and $S_2$ with the conducting electrons results in an effective magnetic interaction between these localized magnetic moments known as RKKY interaction given by\cite{hwang2008screening} $H_{RKKY}(r)=JS_1.S_2\Pi(r)$ in which $J$ is the exchange coupling constant between the conducting electrons and localized magnetic moments and $\Pi(r)$ is the Fourier transform of the $k$-space polarization function $\Pi(q)$ . Therefore the characteristic properties of this interaction could be captured by the polarization function of the mediating electrons. Accordingly it is expected that the anisotropy of the polarization function could manifest itself in the anisotropy of the RKKY interaction as it appeared in the dielectric function. The polarization function could be separated into the inter-band (if $s\ne {s}'$ ) and the intra-band (if $s=s'$) contributions\cite{sensarma2010dynamic}. In addition each of these contributions could be classified as intra-valley and inter-valley contributions that correspond to different transitions in which the transferred momentum, $\vec{q}$, is $q\leq k_F<\mid \textbf{K}_D-\textbf{K}_D'\mid$ or $q\sim \mid \textbf{K}_D-\textbf{K}_D'\mid$ respectively (where $k_F$ is the radius of the Fermi circle and $\textbf{K}_D$, $\textbf{K}_D'$ are different Dirac points). To illustrate the oscillations of the induced charge density a charged impurity has been considered to be inserted in the honeycomb structure. The static polarization function is of particular importance as it determines the screened potential of a charge impurity. The screening particle density ${\delta n}(r)$ due to the central impurity $Ze$ is, \begin{equation}{{ \delta n}}\left(\vec{r} \right)=Ze\frac{1}{{{\left(2\pi \right)}^{2}}}\int{\left[\frac{V\left(\vec{q} \right)\Pi \left(0 ,\vec{q} \right)}{\epsilon \left(0, \vec{q} \right)} \right]\exp \left(i\vec{q}.\vec{r} \right)d^2q}. \end{equation} It is really important to note that the above integration goes beyond the limit of intra-valley transitions (the radius of Fermi circle) and therefore the contribution of the inter-valley transitions should be included. This means that the pattern of Friedel oscillations requires the whole information of the static response function in the first Brillouin zone. Therefore inter-valley transitions between the different Dirac cones should be included. Therefore this type of transitions cannot be captured within the single Dirac cone approximation. We want to remark that the Friedel oscillation curves in the RPA, Hubbard vertex correction and Singwer-Sj{\"o}lander are very similar\cite{hubbard1958description,singwi1970electron,mahan2000many}. The reason, for example in the Hubbard Model, is that the Hubbard local field factor ${{G}_{H}}(q)$ is appeared in correlations that are deal with two-body or more, however the screening is essentially a one-body property and the correlations are not appear in one-body amplitudes \cite{mahan2000many}. \section*{Non-identical contribution of different Dirac cones} \label{app} %____________________ As depicted in Fig.~(\ref{Dirac1}) at low Fermi energies the Fermi curves have been appeared as separated islands around each Dirac point. Therefore the amount of the dielectric and polarization functions in a given $\vec{q}$ wave number have significantly been determined by the orientation of the wave number with respect to the Dirac points position vectors. This anisotropy of the $q$-space is reflected in the real space quantities such as Friedel oscillations. When the Fermi level is appeared as distinct circular curves in k-space. One might conclude that the circular shape of the Fermi contour around of the Dirac points implies that all of the extractable physical properties of the system should be isotropic as long as the Dirac point approximation is applied. However it should be considered that the isotropic and circular shape of the Fermi contours around each of the Dirac points cannot results in isotropic properties at least in when the inter-valley transitions are taken into account. This is due to the anisotropy of the band structure in honeycomb systems which indicates that the Dirac points themselves are located in k-space in an anisotropic manner. Accordingly it is important to note that the results of present study cannot be compared with the results that have been obtained within the single Dirac point approximation\cite{Deng2015}. Meanwhile, for calculation of vector and tensor dependent quantities we should consider that the contribution of inter-valley transitions cannot be identical in all of the $q$-space directions even when $E_F=0$. Moreover when $E_F>0$ both of the intra-valley and inter-valley transitions result in anisotropic dependence of the dielectric function in $q$-space. Therefore different directions in of a given momentum transfer, $q$, have not identical contribution even when the Dirac points are degenerate. Accordingly the conventional Dirac point approximation could not describe all of the physics of the vector or vector dependent parameters at low wave-length limit. This could also result in anisotropic electric and thermal conductivity in graphene-like materials for short range scatterers (for example in the case of the delta-shaped scatterers) in which all of the intra-valley and inter-valley scatterings are possible. \begin{figure}[t] \includegraphics[width=0.35\linewidth]{trigonal_varping2.png} \centering \caption{Dirac points of monolayer graphene and Fermi curves at different Fermi energies. Fermi contours have been depicted for $E_F=0.6eV,1.0eV~1.5eV,~2.0eV,~3.0eV,~4.0eV$ and $5.0eV$. For a given wave vector $\vec{q}$ the contribution of the different Dirac points on $\epsilon(\vec{q})$ strictly depends on the orientation and position of the $\vec{q}$ with respect to the six Dirac vectors. Trigonal warping of the Fermi curves at different Fermi energies has also indicated in this figure. Single Dirac cone approximation could take into account the anisotropic effects comes from the trigonal warping of a single Fermi curve, however, since the orientation of the deformed Fermi curves are not the same, the anisotropic contribution of the other cones are not identical.\label{Dirac1}} \end{figure} When the Dirac points are not located exactly on the Fermi level the intra-valley transitions could take place within the range of $q\leq k_F$. Meanwhile we have assumed that the Fermi energy is still low enough to employ the linear dispersion relation of the Dirac cone. It can be shown that both of the intra-valley ($q\leq k_F$ for this case) and inter-valley transitions should be considered as anisotropic contributions in the dielectric function. In this case since the inter-band transitions ( $\mid k \lambda^s_k>\rightarrow\mid k' \lambda^{s'}_{k'}> ~~ s\neq s' $ ) are absent in the static limit ($\hbar\omega=0$) therefore all of the contributing terms (both intra-valley and inter-valley transitions) are intra-band. Consequently, the contribution of $q\sim 0$ transitions in the static dielectric function decreases by increasing the Fermi energy. Since it can be shown that for $E_F\neq 0$ we have $F_{s s'}\left( \vec{k}+\vec{q},\vec{k}\right)=0~~$ when $q=0$ and $ s\neq s'$. %__________________________ Different types of transitions could be contributed in the dielectric function of the honeycomb structures. In this case transitions could be either intra-valley or inter-valley. Where in the intra-valley transitions initial and final states $\vec{k}$ and $\vec{k'}$ belong to the same Dirac valley cone while in the inter-valley transitions $\vec{k}$ and $\vec{k'}$ belong to different Dirac cones (Fig. \ref{qval} (a) and (b)). The momentum conservation rule for each transition between the states $\vec{k}$ and $\vec{k'}$ with transferred momentum $\vec{q}$ could be satisfied when $\vec{k}$ and $\vec{k'}$ sweep the Fermi circles as shown in Fig. \ref{qval}. It can be inferred that this condition could be satisfied for intra-valley transitions when $0 \leq q \leq k_F$ where $k_F$ is the radius of the Fermi circle Fig.~\ref{qval} (a) and inter-valley transitions could take place $q\sim \mid \textbf{K}_D-\textbf{K}_D'\mid$ where $\textbf{K}_D$ and $\textbf{K}_D'$ are different Dirac points. Fig. \ref{qval} shows the Fermi circles of a planar honeycomb lattice has been depicted in the absence of the SOCs. It can be shown that due to this six-fold band rotational symmetry of the system if the transition rule is satisfied for a given transferred momentum ($\vec{q}$) it will also be satisfied for the sixfold rotated wave number $\mathcal{R}^n_{{2\pi}/{6}} \vec{q}$ (Fig. \ref{qval} (a) and (b)). In which $\mathcal{R}_{{2\pi}/{6}}$ is the sixfold rotation operator. This means that \begin{eqnarray}\label{rule} \delta_{\vec{k}+\vec{q},\vec{k'}}&=&\delta_{\mathcal{R}^n_{{2\pi}/{6}}\vec{k}+\mathcal{R}^n_{{2\pi}/{6}}\vec{q},~\mathcal{R}^n_{{2\pi}/{6}}\vec{k'}}\nonumber\\ &=&\delta_{\vec{k}_n+\vec{q}_n,\vec{k'}_n} \end{eqnarray} Meanwhile the form factor of the graphene is also invariant under the sixfold rotations in the absence of the spin-orbit couplings. \begin{eqnarray}\label{form2}{{F}_{s'{s}}}\left(\vec{k}+\vec{q}, \vec{k}\right)={{F}_{{s}'s}}\left(\mathcal{R}^n_{{2\pi}/{6}}\vec{k}+\mathcal{R}^n_{{2\pi}/{6}}\vec{q}, \mathcal{R}^n_{{2\pi}/{6}}\vec{k} \right). \end{eqnarray} In both cases i.e. for inter-valley and intra-valley transitions band symmetry of the honeycomb structures manifests itself as $E_{\vec{k}}^{s}=E_{\mathcal{R}^n_{{2\pi}/{6}}\vec{k}}^{s}$. Therefore Eq. \ref{aa} reveals that for a given transferred momentum, $\vec{q}$, satisfying the transition rule $\vec{k}=\vec{k}'-\vec{q}$. The contribution of the $\vec{k}$-state in the dielectric function is identical with the contributions of the rotated states: $\mathcal{R}_{{2\pi}/{6}}\vec{k}$, $\mathcal{R}^2_{{2\pi}/{6}}\vec{k}$,... $\mathcal{R}^6_{{2\pi}/{6}}\vec{k}$ regardless of the inter-valley or intra-valley nature of the transitions. In the other words it could be inferred that $\epsilon(\vec{ q})$ $=\epsilon( \mathcal{R}_{{2\pi}/{6}} \vec{ q})$ $=\epsilon( \mathcal{R}^2_{{2\pi}/{6}} \vec{ q})$...$=\epsilon( \mathcal{R}^5_{{2\pi}/{6}} \vec{ q})$. \begin{figure}[t] \includegraphics[width=1.0\linewidth]{valley_new.png} \centering \caption{(Color online) Intra-valley (a) and inter-valley (b) transitions for a given transferred momentum $\vec{q}$. Dashed circles indicate the Fermi circles of the honeycomb system in the absence of the spin-orbit couplings. When the momentum conservation rule is satisfied for $\vec{q}$ the initial and final states should be placed on the Fermi circles. In this case the contribution of the given states (black vectors) is identical with the contribution of the sixfold-rotated states (cyan vectors). One can imagine about another type of possible transitions (c) with constant value of the transferred momentum $q=q_1$ between the equi-energy states $E^s_k=E^s_{k'}=E^s_{k'_1}=E^s_{k''}$ where the corresponding pair vectors ($\vec{q}$ $\vec{q}_1$), ($\vec{k}$ $\vec{k'}_1$) and ($\vec{k'}$ $\vec{k''}$) are not related by sixfold symmetry operators e.g. $\mathcal{R}^n_{{2\pi}/{6}}\vec{q}\neq\vec{q}_1$. It can be shown that form factor of these transitions are different i.e. $F(k,k')\neq F(k'_1,k'')$. At zero Fermi energy (d) i.e. when $k_F=0$ intra-valley transitions occur at $q=0$ which result in central peak of the dielectric function. However inter-valley transitions (light green vectors) are still the source of anisotropy of the dielectric function.\label{qval}} \end{figure} For example in the case of intra-valley transitions the different Dirac point or Fermi curves which have been related by sixfold rotation operators have the same contribution in the dielectric function for those transferred momentums which have the same symmetry relation. Consequently, the contribution of the Fermi circle located around the $\vec{K_D}$ Dirac point on the dielectric function of $\vec{q}$ is identical with the contribution of $\vec{K}_{D}^{(i)}$ Fermi circle on the dielectric function of of $ \mathcal{R}^i_{{2\pi}/{6}} \vec{ q}$ where these two Dirac points are related by $\vec{K}_{D}^{(i)}=\mathcal{R}^i_{{2\pi}/{6}} \vec{K_D} $. If we continue the same procedure for parallel wave numbers, which satisfying the transition rule, one can realize the anisotropy of the dielectric function which manifests itself by sixfold symmetric curve at low transferred momentums Fig.~\ref{eps-mg-ef-05}. These relations identify different class of the states which have identical contribution in the dielectric function i.e. the for different states with transferred momentum $q$ in this case the class of the identical contributions for both inter and intra valley transitions is specified by \begin{eqnarray}\label{class} [q]=\{\vec{q},~ \mathcal{R}_{{2\pi}/{6}}\vec{q},~ \mathcal{R}^2_{{2\pi}/{6}}\vec{q},~ ...\mathcal{R}^5_{{2\pi}/{6}}\vec{q}\}. \end{eqnarray} Which corresponds to the transitions: $\vec{k}\rightarrow\vec{k}+\vec{q}$, $\mathcal{R}^5_{{2\pi}/{6}}\vec{k}\rightarrow\mathcal{R}_{{2\pi}/{6}}\vec{k}+\mathcal{R}_{{2\pi}/{6}}\vec{q}$, ...$\mathcal{R}^5_{{2\pi}/{6}}\vec{k}\rightarrow\mathcal{R}_{{2\pi}/{6}}\vec{k}+\mathcal{R}_{{2\pi}/{6}}\vec{q}$. The first consequence of the above argument is that the different Fermi curves of each Dirac point have not identical contribution on the dielectric function at a given $\vec{q}$ wave number. In the other words the contribution of each Fermi circle (corresponding to $K_D$ dirac point) with a given transferred momentum, $\vec{q}$, has been determined by the orientation of the $\vec{q}$ with respect to the $K_D$. When $\vec{q}$ satisfies the transition rule for a specific Dirac cone (for example in an intra-valley process) this rule will be satisfied for $\mathcal{R}^n_{{2\pi}/{6}}\vec{q}~~(n=1..5)$ at other Dirac cones and $\vec{q}$ itself could not satisfy the momentum conservation rule or could not give the same contribution at these Dirac cones. Therefor the contribution of different Fermi circles on the dielectric function of $\vec{q}$ is not identical. Accordingly the dielectric function should be anisotropic in the $q-$space with sixfold symmetry which was originated from the symmetry of the band structure. All of the other possible transitions with a given fixed value of the transferred momentum could take place between the isoenergy states as shown in Fig. \ref{qval} (c). In this group of the transitions, transferred momentum is the same $q=q_1$ and both of the initial final states are located at Fermi circle $E^s_k=E^s_{k'}=E^s_{k'_1}=E^s_{k''}=E_F$. However, corresponding pair vectors are not related by sixfold symmetry operators i.e. $\mathcal{R}^n_{{2\pi}/{6}}\vec{q}\neq\vec{q}_1$, $\mathcal{R}^n_{{2\pi}/{6}}\vec{k}\neq\vec{k}'_1$ and $\mathcal{R}^n_{{2\pi}/{6}}\vec{k'}\neq\vec{k}''$. The transition rule has been satisfied for these transition where we have $\vec{k}+\vec{q}=\vec{k}'$ and $\vec{k}'_1+\vec{q}_1=\vec{k}''$. Meanwhile, the form factor of the transitions are not the same $F(k,k')\neq F(k'_1,k'')$ which implies that the corresponding contributions are not identical (Fig. \ref{qval} (c)). When the Fermi energy located at Dirac points i.e. $E_F=0$ then the intra-valley transitions occur in $\vec{q}=0$ between different bands (Fig \ref{qval} (d)). Unlike the previous case the intra-valley contributions are identical here. These contributions result in the central peak of the dielectric function. Since the $V(q)=2\pi e^2/q$ diverges at $q=0$ where the intra-valley transitions could contribute at this point. However the inter-valley transitions occur away from the $\Gamma$-point ($q=0$) result in anisotropic contributions as the previous case. Therefore one can expect that the anisotropy of the dielectric function would be appeared at $q \geq \mid \textbf{K}_D-\textbf{K}_D'\mid$ Fig. (\ref{eps-mg-0-2d}). \section*{Discussions and numerical results} In conclusion, we have analyzed the band induced anisotropic effects in graphene-like structures. We have also investigated the influence of the spin-orbit interactions on dielectric function and Friedel oscillations in this type of materials. Due to the foregoing discussions for wave number-dependent or non-scalar quantities such as dielectric function, electric and thermal conductivities we have to concern about the position of the Dirac points relative to the direction of the characteristic vector of the physical quantity (such as transfered momentum) even when the Dirac point approximation is valid. For sharp scattering potentials we have to consider the inter-valley transitions in this type of the quantities. Within the Dirac point approximation the integration over the state-resolved contributions is generally performed over a single Fermi circle. This could be a correct approach when we assume the identical contribution of each Dirac point and ignore the inter-valley transitions. In this case if we put aside the Dirac point approximation and perform the integration over the whole Brillouin zone the correct contribution of each Dirac point could be obtained. This means that beyond the Dirac point approximation dielectric function and therefore Friedel oscillations for graphene-like honeycomb structures should be anisotropic in k-space and the real space respectively. % % % % % % % % % % % % % % % % % % % % % % % % % % %^^^^^^^^^^^ Based on RPA formalism, which accounts for electron-electron scattering, we have shown that the Dirac points have not identical contribution for wave number dependent quantities such as dielectric function even when the Fermi energy is close to these points. The main limitations for the use of the Dirac point approximation has been discussed within the current work. Non-identical contribution of the Dirac points results in anisotropic dielectric function in $k$-space. Moreover, the anisotropy of the dielectric function leads to anisotropic Friedel oscillation in graphene-like materials. We have shown that the isotropic Friedel oscillations cannot be considered as a consequence of the Dirac point approximation itself. Since in this case the isotropic Friedel oscillations could be obtained just when the inter-valley transitions are ignored. The influence of the Rashba SOC on the dielectric function and Friedel oscillations has also been discussed in the present study where we have shown that increasing the Rashba coupling strength cannot results in significant change in the dielectric function and Friedel oscillation. Meanwhile in the presence of the Rashba interaction the inversion symmetry of the dielectric function has been lifted in $k$-space. The extrinsic Rashba interaction could be manipulated by an external gate voltage. By changing the extrinsic Rashba coupling strength, polarization function and consequently the dielectric function will be altered. Dielectric function has been determined by density-density correlations in the system. In the present work this correlation function has been obtained within the random phase approximation. Results of the numerical calculations depicted in the following figures of this section. As mentioned before the band induced anisotropic effects could be obtained even when the Fermi energy is close to the Dirac points. This is due to the fact that the optical transitions which could be available for a given transferred momentum $\vec{q}$ are not identical for different Dirac points in $k$-space. At the level of the single Dirac point approximation anisotropic position of the Dirac points and its relevant effects have been ignored. Meanwhile, the anisotropy of single Fermi circle reshaping, which arises by increasing the Fermi energy, could be obtained within the single Dirac cone approximation. In the present work we have obtained the anisotropic properties which could be originated from the anisotropic location of the Dirac points at low Fermi energies. According to the numerical results, two-dimensional graphene like materials show anisotropic Friedel oscillations beyond the Dirac cone approximation even when the Dirac points located at the Fermi level. In addition results of the present study show that the extrinsic Rashba coupling has not any considerable effect on the Friedel oscillations. Meanwhile the influence of the spin-orbit couplings is relatively evident in the Friedel oscillations of the monolayer germanene. It is important to consider that the general integral expression for the polarization function (Eq. \ref{aa}) goes beyond the states in which the Dirac point approximation is not valid. Nevertheless, far from the Dirac points most of the states in each band are empty with no contribution in the dielectric function in the static limit ($\hbar \omega =0$) even at room temperature. For gap-less structures, at room temperature, thermal transitions could be taken place within the range of thermal energy $K_BT=0.025$eV. Therefore, when the Fermi level is located at the Dirac point, the contribution of the states which were not in the permitted range of the Dirac point approximation automatically have been eliminated by the distribution function implemented in Eq. (\ref{aa}). The linear dispersion relation (and therefore circle like Fermi curves) around the Dirac points valid even up to $E_F \sim 1$ eV in graphene and the thermal transitions at room temperature with $K_BT=0.025$ eV could not induce any considerable contribution from those states which have been located far from the Dirac points. So it seems that the Dirac point approximation could still describe the physics of the honeycomb lattice and the linear dispersion relation around the Dirac points could be employed for the calculation of the dielectric function. This means that the non-linear part of the band structure and the anisotropy that might be induced by this part could be ignored. This is due to the fact that this part of the band structure (which could be considered the energy states with $E^s_{k}>1 eV$) has not been occupied even as a result of the thermal transitions or could not contribute in the isoenergy transitions of the static limit. However, in the current work we have shown that the contribution of the Dirac points in wave-number dependent quantities should be calculated with some care even when the Dirac point approximation is valid. It seems that the degeneracy of the K-points results in identical contribution of each Dirac point in all of the physical quantities when the Fermi energy is close to the Dirac points. However, it should be noted that, this is not the case for some of the quantities which directly depending on the direction of the transfered wave number. In this case due to the anisotropic position of the Dirac points in the Brillouin zone, each of the Dirac points has not identical contribution for this type of the quantities, even when the Dirac point approximation is valid. Meanwhile, since the inter-valley transitions could take place between the different Dirac cones, these type of transitions cannot be captured within the single Dirac cone approximation. Anisotropic Friedel oscillations in two-Dimensional structures have been observed before \cite{hofmann1997anisotropic}. However, in the present case the anisotropic effects are direct manifestation of non-identical contribution of Fermi circles and Dirac points in wave number dependent quantities. Some of the physical quantities, such as dielectric function, are given by integration over the Brillouin zone as expressed in Eq. (\ref{aa}). This integration goes beyond the states in which the Dirac point approximation and the linear dispersion relation no longer valid. However distribution function at low temperatures and Fermi energies picks up the contribution of the states which were located near to the Dirac points. Nevertheless for evaluating non-scalar quantities that depending on the wave number and its direction we should consider that the orientation of the wave number, $q$, (relative to the position vector of the Dirac points in k-space) determines the contribution of each Dirac point. It can be clearly demonstrated that the dielectric function shows anisotropic directional dependence in $q$-space. It should be noted that increasing of the Fermi energy results in deformation of circle-like Fermi curves (Fermi-circles) of low Fermi energies around the Dirac points \cite{Shih2017}. In this case the isotropic form of the Fermi circles change into the trigonal-shaped contours (known as trigonal warping effect) and the isotropic form of the Fermi curve around of the Dirac points has totally been removed at high Fermi energies. This type of deformation could results in a new source of anisotropy at high Fermi energies which could be captured within the single cone approximation. Meanwhile trigonal warping induced anisotropy has not been considered in the current work which was limited to the low Fermi energies. At low Fermi energies optical transitions around each Dirac points, where take place within a single Fermi circle, have the main contribution in the dielectric function of the honeycomb systems. This manifests itself as a central peak of the dielectric function in the middle of the Brillouin zone. It could be shown that beyond the Dirac point approximation the central peak contains the contribution of all of the Dirac points via the inter-band transitions. When the Fermi energy is close to Dirac points ($E_F=0$) and at the high wave length limit ($q\ll \mid K_D-K'_D $) the occupation factor $f^s_k-f^{s'}_{k'}$ could have a significant value only when the $k$ and ${k'}$ states are close to a single Dirac point of different bands ($s\neq s'$). Meanwhile the Kronecker delta, $\delta_{\vec{k'},\vec{k}+\vec{q}}$, in the expression of the polarization indicates that the contribution of the Dirac points should be selected by $\Gamma$-point $(\vec{q}=0)$. Since at this limit the main contribution is due to the intra-valley transitions which take place near the Dirac points in which $k\approx K_D$ and $k'\approx K_D$ therefore the mentioned Kronecker delta, which reflects the momentum conservation, imposes that the contribution of the Dirac points should be manifest themselves at the $\Gamma$-point $(\vec{q} \approx 0)$ of the $q$-space. The mentioned argument reveals the fact that the central peak (Figs. \ref{eps-mg-0} and \ref{eps-SG}) of the dielectric function is exactly sum of all of the intra-valley contributions from each Dirac point. In this case since the main contribution belongs to the case of zero transferred momentum ($q\sim 0$) the question of the anisotropy in $q$-space seems to be irrelevant at $E_F=0$. Although the contribution of the intra-valley transitions are dominant at $q\sim 0$, however, this is not the case for the wave numbers within the range of $ \mid K_D -K'_D \mid-2k_F \leq q< \mid K_D -K'_D \mid+2k_F$. It should be noted that the contribution of inter-valley transitions between different Dirac points ($q\sim \mid K_D -K'_D \mid$) are not identical. In this case the contribution of each pair of contributing Dirac point (${\bf{K}}_D$ and ${\bf{K}'}_D$) has been determined by the direction of the $\vec{q}$ vector. In order to obtain the anisotropic effects which have been induced merely by band energy we have to switch off both types of the spin orbit couplings. It was reported that the Rashba coupling strength in graphene is greater than the intrinsic spin orbit interaction \cite{dedkov2008rashba}. Therefore one can ignore intrinsic spin orbit coupling in monolayer graphene. At zero Rashba interaction both intrinsic and extrinsic spin-orbit couplings are absent. This enables us to obtain the anisotropic effects which could be induced merely by band energy. Dielectric function at zero Rashba coupling and zero Fermi energy ($E_F=0$) has been obtained as depicted in Fig. (\ref{eps-mg-0}). At the first look it seems that there is no anisotropy in the dielectric function of the honeycomb structures (Figs. \ref{eps-mg-0} and \ref{eps-SG}), however, it should be noted that the anisotropy of the dielectric function has been hidden behind the large central peak at $q\sim 0$. The amount of the dielectric anisotropy is very small in comparison with the value of the dielectric function at the $\Gamma$-point. Accordingly this fact prevents the identification of the directional dependence of the dielectric function. At low wave numbers i.e. in the range of the intra-valley transitions dielectric function seems to be quite isotropic in $q$-space Fig. (\ref{eps-mg-0-2d} (a) and (b)). However, far from the central region if we select different symmetric slices of the dielectric surface, the anisotropy of the dielectric function will be evident Fig. (\ref{eps-mg-0-2d} (c) and (d)). A similar discussion holds for the case of silicene and germanene where as shown in Figs. \ref{eps-s} and \ref{eps-g} dielectric function has completely different behavior on symmetrically chosen slices, especially far enough the $\Gamma$ point. These figures evidently show that for different directions in the $k$-space behavior of the dielectric function are quite different. This reveals that the contributing inter-valley transitions in finite wave length limit ($q\sim \mid K_D -K'_D \mid$) introduce the anisotropic behaviors. \begin{figure}[t] \includegraphics[width=0.5\linewidth]{mono_diel_ef_0.png} \centering \caption{$k$-space dielectric function of monolayer graphene at $t_{extR}=0$. Additionally it was assumed that the intrinsic spin-orbit coupling has been also negligible. This enables us to compute the net band induced anisotropic effects. \label{eps-mg-0}} \end{figure} \begin{figure}[t] \includegraphics[width=0.75\linewidth]{S_G_Diel.png} \centering \caption{$k$-space dielectric function of monolayer silicene and germanene $t_{extR}=0.01eV$. \label{eps-SG}} \end{figure} \begin{figure}[h] \includegraphics[width=0.7\linewidth]{2d_diel.png} \centering \caption{Dielectric function of monolayer graphene at different symmetrically chosen slices for $t_{extR}=0$ in $k$-space. (a) At $q_x=0.0015{\AA}^{-1}$ plane. (b) $q_y=0.0015{\AA}^{-1}$ plane. (c) At $q_x=0.15{\AA}^{-1}$ and $q_y=0.15{\AA}^{-1}$ slices. (d) At $q_x=0.26{\AA}^{-1}$ and $q_y=0.26{\AA}^{-1}$ planes. Anisotropic effects appear far away from the origin where the contribution of the inter-valley transitions should be taken into account.\label{eps-mg-0-2d}} \end{figure} \begin{figure}[t] \includegraphics[width=0.8\linewidth]{slice_silicene.png} \centering \caption{Dielectric function of monolayer silicene at different symmetrically chosen slices for $t_{extR}=0.01 $eV in $k$-space. (a) At $q_\alpha=0.06{\AA}^{-1}$ planes ($\alpha = x,y$). (b) $q_\alpha=0.005{\AA}^{-1}$ planes. (c) At $q_\alpha=0.09{\AA}^{-1}$ slices. (d) At $q_\alpha=0.01{\AA}^{-1}$ planes. Anisotropic effects appear far away from the origin where the contribution of the inter-valley transitions is dominant.\label{eps-s}} \end{figure} \begin{figure}[t] \includegraphics[width=0.8\linewidth]{germanene_slice.png} \centering \caption{Dielectric function of monolayer silicene at different symmetrically chosen slices for $t_{extR}=0.01 $eV in $k$-space. (a) At $q_\alpha=0.02{\AA}^{-1}$ planes ($\alpha = x,y$). (b) $q_\alpha=0.01{\AA}^{-1}$ planes. (c) At $q_\alpha=0.04{\AA}^{-1}$ slices. (d) At $q_\alpha=0.001{\AA}^{-1}$ planes. Anisotropic effects appear far away from the origin where the contribution of the inter-valley transitions is dominant.\label{eps-g}} \end{figure} \begin{figure}[h] \includegraphics[width=0.5\linewidth]{mono_friedel.png} \centering \caption{Real space anisotropic Friedel oscillations in monolayer graphene at different Rashba couplings. (a)-(c) As shown in these figures the Rashba interaction has not a significant influence on the Friedel oscillations at intermediate Rashba coupling strength. (d) Unlike to the previous case for $t_{extR}=0.2$~eV Friedel oscillations are not similar even for $x$ and $y$ directions. \label{Fr-mg}} \end{figure} \begin{figure}[h] \includegraphics[width=0.5\linewidth]{S_G_friedel.png} \centering \caption{Real space anisotropic Friedel oscillations in monolayer silicene ((a) and (b)) and germanene ((c) and (d)) at different Rashba couplings. The Rashba interaction has not a significant influence on the Friedel oscillations. \label{Fr-s}} \end{figure} \begin{figure}[t] \includegraphics[width=1.0\linewidth]{polar2.png} \centering \caption{Angular dependence of the normalized graphene Friedel oscillations, $\delta n(r,\theta)/n_0$, for $t_{extR}=0$, $E_F=0.1 t$ and $K_BT=0$ at (a) $r=4\AA$, (b) $r=5\AA$, (c) $r=14\AA$ and (d) $r=25\AA$ in which we have defined $n_0=\delta n(\textbf{r}=0)$. \label{polar}} \end{figure} As discussed before band induced anisotropic effects ( see Fig.~\ref{Dirac1}) have been reflected in the Friedel oscillations of the graphene-like structures illustrated in Figs. \ref{Fr-mg} and \ref{Fr-s}. The anisotropy of the dielectric function as discussed before is due to the non-identical contribution of the Dirac points and the nonlinear part of the spectrum cannot contribute in dielectric function when the Fermi level is close to the Dirac points. Increasing the Rashba coupling strength slightly modifies the Friedel oscillations in silicene, however the change in monolayer graphene is relatively more pronounced compared to that of other selected honeycomb structures where Figs. (\ref{Fr-mg} and \ref{Fr-s}) exactly indicates this fact. This could be explained if we consider relatively large and dominant intrinsic spin-orbit coupling in silicene. As shown in the figure (\ref{Fr-mg}) the anisotropic Friedel oscillations have been observed even when the Rashba coupling strength is very low or zero. It can be inferred from the results of the current work that the Rashba coupling is less effective in the generation of the anisotropy. Therefore one can conclude that the anisotropy of the dielectric function and Friedel oscillations mainly depends on the anisotropy of the band structure in $k$-space. There are several studies which have been performed in this field, aiming at an accurate quantitative prediction of dynamical dielectric function, screened charged impurity potential and Friedel oscillations in graphene-like materials. It was realized that the long-distance decay of Friedel oscillations in graphene depends on the symmetry of the scatterer\cite{CheianovEPJ}. In addition a faster, $\delta n \sim 1/r^3$, decay in comparison with conventional 2D electron systems has been observed in Friedel oscillations of a localized impurity inside the monolayer graphene within the Dirac point approximation \cite{CheianovEPJ,wunsch2006dynamical}. However, $1/r$ decay has been reported for bilayer graphene \cite{BenaPRL} and strong asymmetry and an inverse square-root decay has also been obtained for an anisotropic graphene-like structure when one of the nearest-neighbor hopping amplitudes is different from the others \cite{dutreix2013friedel}. Recently in rhombohedral graphene multi-layers, $1/r$ decay has been observed for impurity induced Friedel oscillations \cite{Dutreix2016}. Completely isotropic behavior has been reported for the potential of a screened charged impurity, Friedel oscillations \cite{scholz2012dielectric,wunsch2006dynamical,gomez2011measurable,Cheianov} and static dielectric function \cite{hwang2007dielectric} within the Dirac point approximation in graphene. Similarly the Dirac point approximation results in isotropic screened potential of a charged impurity in other graphene-like materials such as silicene and germanene \cite{tabert2014dynamical}. The studies based on Dirac point approximation give the correct physics of the high wave length limit ($q << k_F$) at $E_F=0$ where inter-valley transitions could not contribute in the physical processes. In the absence of the spin-orbit couplings by using the massless linear Dirac spectrum it was also shown that short wavelength spatial dependence of the local density of states leads to anisotropic Friedel oscillations which has the form\cite{Attila} \begin{equation}\label{ff}{{ \delta n}}\left(r \right)\sim c(\vec{r}) \rho_0(E_F)\frac{\sin(2k_Fr)}{r^2}. \end{equation} In which $c(\vec{r})$ is the short wavelength spatial dependence factor and $\rho_0(E)$ is the density of states. Anisotropic dependence of the Friedel oscillations has been introduced by $c(\vec{r})$ factor which was found to be invariant under threefold rotations\cite{Attila}. However, if the impurity could not produce inter-valley scatterings this factor is reduced to a constant number\cite{Attila}. Therefore the anisotropic effects have been removed in the absence of inter-valley transitions\cite{Attila}. In the current study we have observed that for finite Fermi energies $0<E_F \leq 1$eV even intra-valley transitions are the source of the anisotropic behaviors at linear energy dispersion regime. In the case of the single valley band structures where all of the transitions should be considered as intra-valley transitions. The wave length of the Friedel oscillations is modulated by Fermi wave number. However it can be easily shown that this is not the case for multi-valley band structures. In which the inter-valley transitions could contribute in the dielectric function. As indicated in Eq. (\ref{ff}) it was expected that the wavelength of the Friedel oscillations should be modulated by the Fermi wave-vector $k_F$ \cite{Attila}. Where the long range behavior of the local density of states has been obtained within the single valley approximation and linear dispersion relation\cite{Attila}. The possible transfered momentums, $q$, determine the oscillation wavelength of the induced charged and for single valley band structures in two-dimensional systems typical transfered momentum is $q\sim2k_F$. However it should be noted that for a typical graphene Fermi energy e.g. $E_F=0.1$eV one can obtain $k_F=E_F/(\hbar v_F)=0.0152\AA^{-1}$. Therefore the oscillation wavelength that corresponds to the intra-valley transitions is about $\lambda_{intra}=2\pi/2k_F\sim200\AA$. On the other hand in the present case the inter-valley transitions with momentum transfer of $q\sim|\textbf{K}-\textbf{K}'|\pm 2 k_F$ correspond to the oscillations with wavelength of $\lambda_{inter}=2\pi/(|\textbf{K}-\textbf{K}'|\pm 2 k_F)$. The wavelengths of the oscillations in the current work for monolayer graphene are in the following range $7\AA\lesssim\lambda_{inter}\lesssim13\AA$ that are at the same order of the inter-valley transition wavelengths given by $\lambda_{inter}=2\pi/(|\textbf{K}-\textbf{K}'|\pm 2 k_F)$. For example the distance between two successive Dirac points in graphene is about $\Delta K=|K-K'| \sim 1.7\AA^{-1}$. Therefore the average momentum transfer between these Dirac points is $ q \sim |K_-K'|/2=0.85\AA^{-1}$ then $\lambda_{inter}=2\pi/\Delta K=7.37\AA$. The numerical results indicate that the wavelength of the oscillations is less-sensitive to the value of the Fermi energy. This can be realized if we consider that $|\textbf{K}-\textbf{K}'|>>2k_F$ for intermediate Fermi energies. Decay rate of the Friedel oscillations are determined by fitting to the numerical results. We have examined several decay rates such as $1/r$, $1/r^2$ and $1/r^3$. Numerical fitting shows that the $1/r$ decay rate is much more close to the computational data profile. More precisely decay rate is actually $1/r^{1+\eta}$ where $0<\eta<0.2$. Another important issue about the Friedel oscillations is that how sharp the mentioned density anisotropy really is? In this way we have obtained the angular dependence of the induced density at different distances as depicted in Fig. (\ref{polar}). As indicated in this figure the anisotropy of the Friedel oscillations increases by distance. It can be realized that the angular dependence of the induced density is so sharp at intermediate distances. This provides more detectable condition for observation of the anisotropy. Interestingly it was shown that the Friedel oscillations in graphene have a strong sublattice asymmetry\cite{lawlor}. These calculations have been performed beyond the Dirac point approximation within the Born approximation which can be employed for weak scattering potentials and the stationary phase approximation (SPA) has also been applied for Brillouin zone integrations\cite{lawlor}. Anisotropic Friedel oscillations could also be inferred from the numerical results of the recent work in the absence of the spin-orbit interactions especially over short distances. Finally it is important to note, the anisotropy of the dielectric function suggests that the orientation of the bases vectors of the honeycomb lattice could be determined by full optical measurements. Since dynamical dielectric function of the graphene-like materials are possibly have the same anisotropic nature. Therefore the absorption spectra of the system should be anisotropic. Accordingly, the real space orientation of the basis vectors could be explored since the absorption spectra leads to identification of the band energy configuration in k-space. \bibliography{refrences} %\section*{Acknowledgements (not compulsory)} %Acknowledgements should be brief, and should not include thanks to anonymous referees and editors, or effusive comments. Grant or contribution numbers may be acknowledged. \section*{Author contributions statement} All authors contributed to the numerical programming. A. Phirouznia and T. Farajollahpour carried out the discussion and analysed the results. In addition final version of the manuscript has been reviewed by A. Phirouznia and T. Farajollahpour. \section*{Additional information} %To include, in this order: \textbf{Accession codes} (where applicable); \textbf{Competing financial interests} (mandatory statement). Competing financial interests: The authors declare no competing financial interests. %The corresponding author is responsible for submitting a \href{http://www.nature.com/srep/policies/index.html#competing}{competing financial interests statement} on behalf of all authors of the paper. This statement must be included in the submitted article file. \end{document} }}}\end{equation}}\end{eqnarray}}\end{eqnarray}}}
\caption{ {\it Top panels}: The surface number density profiles, $\sigmag(R)$, of SDSS photometric galaxies with different magnitude thresholds around the entire \redms cluster sample with $z\in[0.1, 0.33]$ and richness $\lambda\in[20, 100]$, are shown using symbols with errorbars. The dashed lines correspond to (sub)-halo surface density profiles in the Multidark Planck II simulation around clusters with the mass threshold similar to our sample at $z=0.248$. The threshold on subhalo $V_{\rm peak}$ values roughly correspond to the magnitude thresholds in each panel and were obtained by subhalo abundance matching (see Appendix ~\ref{app:amatch}). {\it Bottom panels}: The logarithmic slope of the surface density profiles are shown using solid and dashed lines for the observed galaxy and the subhalo surface density profiles, respectively. The observed slope of the surface density profile has a shape which is similar to that expected from simulations. Note that although the surface density profiles both in observations and simulations exhibit similar steepening, the corresponding radii of the steepest slope are at slightly different locations. }
\caption{ Left panel: The distributions of line-of-sight velocities of spectroscopic galaxies from SDSS around most probable \redms central galaxies in the two subsamples. Middle panel: The distributions of the axis ratios of the light profiles in the $i$ band for the most probable centrals in the two cluster subsamples are shown. Right panel: The distributions of the axis ratios of the satellite distributions around the most probable centrals in the two cluster subsamples are shown. We do not see a large differences between our two cluster subsamples for any of these statistics. }
\caption{High-z galaxies are ``not special''. Subsamples of local galaxies in SDSS that match the \oiiired/\halpha~ratios of high-z galaxies. The local high-z analogs are selected based on \EWha, i.e., sSFR, to be similar to $z~\sim~2$, $z~\sim~3.5$, and $z~\sim~5.5$ galaxies. This shows the potential of using local galaxies with high resolution spectra to investigate the properties of high-z galaxies. \label{fig:local}}
\caption{Some GED ($H$, $S^{\mathrm{\Rmnum{1}}}$ and $S^{\mathrm{\Rmnum{2}}}$) of three rat strains and change ratios ($R^{\mathrm{\Rmnum{1}}}$ and $R^{\mathrm{\Rmnum{2}}}$) of two groups. In this paper, we use ``{\it Hist1h2ai}'' as the abbreviation of gene symbol ``{\it Hist1h2ai\_predicted /// Hist1h4a\_predicted}''.}
\caption{\chandra 0.5 - 6.0 keV X-ray spectrum of IC\,2497 (total emission, r = 4 arcsec) for ObsID 14381 (black) and ObsID 13966 (grey). The spectra are binned so that each bin contains at least 3 counts. The lines correspond to the best fit model{\texttt{phabs}}$\times$({\texttt{zpow}}+{\texttt{APEC}}), where {\texttt{zpow}} corresponds to the AGN emission and {\texttt{APEC}} to the emission from a hot diffuse gas (see text). The absorption is fixed to the Galactic value. The residuals of the fit are shown in the bottom panel. \textcolor{black}{The dotted and dashed lines correspond to the \texttt{phabs}$\times${\texttt{zpow}} and \texttt{phabs}$\times${\texttt{APEC} models, respectively, both normalised by a factor 10.}}}
\caption{Comparison of the ideal point ranks of legislators in the 107$^{\mbox{th}}$ US Senate obtained by fitting independent one-dimensional Bayesian spatial voting models \citep{jackman2001multidimensional} to the motions voted on before and after the defection of senator James M.\Jeffords from the Republican party. Democratic senators appear as blue circles ({\color{blue} $\circ$}), Republicans appear as red plus signs ({\color{red} $+$}), and party-switcher Jeffords appears as a green triangle ({\color{green} $\triangle$}).}
\caption{Senators with the 40 largest probabilities of a change in estimated ideal points under our joint unidimensional model with a probit link. Democratic senators appear as blue circles ({\color{blue} $\circ$}), Republicans appear as red plus signs ({\color{red} $+$}), and party-switcher Jeffords appears as a green triangle ({\color{green} $\triangle$}). There is at least weak evidence of a change in revealed preferences for 15 legislators, including party leaders Daschle and Lott.}
\caption{Comparison of the posterior mean of the ideal points (left panel) and the rank order of senators (right panel) before and after the switch in Senate control under our joint unidimensional model with a probit link. Democratic senators appear as blue circles ({\color{blue} $\circ$}), Republicans appear as red plus signs ({\color{red} $+$}), and party-switcher Jeffords appears as a green triangle ({\color{green} $\triangle$}). Note that the estimates of the change of the order rank of legislators under our joint model are much more conservative than those shown in Figure \ref{fig:independent_ideal_107s}.}
\caption{Estimated posterior mean ideal points densities for before and after the majority party switch. The solid line (\solidrule) is based on the estimates after Jeffords' switch, while the dashed line (\dashedrule) is based on the estimates before Jeffords' switch. Note that the two modes seem to become more pronounced (i.e., the parties become more polarized) after senator Jeffords' switch.}
\caption{Senators with the 40 largest probabilities of a change in estimated ideal points under our joint unidimensional model with a logit link. Democratic senators appear as blue circles ({\color{blue} $\circ$}), Republicans appear as red plus signs ({\color{red} $+$}), and party-switcher Jeffords appears as a green triangle ({\color{green} $\triangle$}). In this case there is at least weak evidence of a change in revealed preferences for only 12 legislators (note that Republican leader Lott is not among them).}
\caption{Posterior mean of the rank order of senators before and after the switch in Senate control under our joint unidimensional model with a logit link. Democratic senators appear as blue circles ({\color{blue} $\circ$}), Republicans appear as red plus signs ({\color{red} $+$}), and party-switcher Jeffords appears as a green triangle ({\color{green} $\triangle$}). These estimates are very similar to those shown in Figure \ref{fig:comparebeta_probit}.}
\caption{Senators with the 40 largest probabilities of a change in estimated ideal points under our joint unidimensional model with a probit link under the alternate prior $a=1$ and $b=1$. Democratic senators appear as blue circles ({\color{blue} $\circ$}), Republicans appear as red plus signs ({\color{red} $+$}), and party-switcher Jeffords appears as a green triangle ({\color{green} $\triangle$}).}
\caption{Mean proportion of false positives (black solid line, \solidrule) and proportion of simulations with at least one false positive (red dashed line, {\color{red} \dashedrule}) for different posterior probability thresholds in our first simulation study. A false positive rate of 0.05 is shown in grey. Note that for thresholds above 0.5 both of these error rates are below 1\%.}
\caption{Receiver operating characteristic (ROC) curves for our second simulation study. Panel (a) presents the ROC curves constructed from our method that uses posterior probabilities, while panel (b) shows the ROC curves constructed from the methods used in \cite{clinton2004statistical} and \cite{roberts2007statistical}. We show both the individual ROC curves associated with each of our simulated datasets (grey dashed lines, {\color{gray} \dashedrule}), as well as the average ROC curve (black solid line, \solidrule). The area under the curve (AUC) for the average ROCs curves are 0.98 and 0.86 respectively.}
\caption{Dendrogram of similarities. It shows the hierarchical clustering of the pathway-based gene network. In this paper, the following nine genes (as recorded in Ref.~\cite{LLW2008pg}), {\it Rhod\_predicted}, {\it Polr3e\_predicted}, {\it Hist1h2ai\_predicted /// Hist1h4a\_predicted}, {\it Sdhb\_predicted}, {\it Actg1 /// LOC295810}, {\it Col6a1\_predicted}, {\it Col4a2\_predicted}, {\it Prcp\_predicted}, and {\it Cd36 /// RGD1562323\_predicted}, are abbreviated as {\it Rhod}, {\it Polr3e}, {\it Hist1h2ai}, {\it Sdhb}, {\it Actg1}, {\it Col6a1}, {\it Col4a2}, {\it Prcp}, and {\it Cd36}, respectively.}
\caption[]{ Three-dimensional visualization of vertical magnetic field, $B_{z}$ at the surface (color-coded) together with \blue{three-dimensional volume rendering of the vertical component of} the magnetic field for Run~RM1. %Run \url{dynamo256k30VF_relhel1_roff-1_morediag_proc_var_R50} }
\caption{Cost (value of the potential) shown as a function of the level of supervision for 100 Monte Carlo replicates. Shading indicates $\pm$ two standard deviations. Colors indicate algorithm:\\ \textcolor[rgb]{.62, .65, .03}{gold}: {\tt Constrained-KMeans} (without Lloyds iterations);\\ \textcolor[rgb]{0.0, .72, .96}{blue}: {\tt ss-k-means++} (without Lloyds iterations);\\ \textcolor[rgb]{1.0, .44, .42}{red}: {\tt Constrained-KMeans};\\ \textcolor[rgb]{0.0, .76, .53}{green}: {\tt ss-k-means++}; and\\ \textcolor[rgb]{.91, .40, .94}{pink}: {\tt Constrained-KMeans} initialized at true centroids of labels.}
\caption{Fractional cost (value of the potential over an estimate of the optimal) plotted as a function of the level of supervision for 100 Monte Carlo replicates. Shading around the lines indicates $\pm$ two standard deviations. The shaded region is the region corresponding to the theoretical cost in expectation from Section \ref{sec:theory}. Colors indicate algorithm:\\ \textcolor[rgb]{.62, .65, .03}{gold}: {\tt Constrained-KMeans} (without Lloyds iterations);\\ \textcolor[rgb]{0.0, .72, .96}{blue}: {\tt ss-k-means++} (without Lloyds iterations);\\ \textcolor[rgb]{1.0, .44, .42}{red}: {\tt Constrained-KMeans};\\ \textcolor[rgb]{0.0, .76, .53}{green}: {\tt ss-k-means++}; and\\ \textcolor[rgb]{.91, .40, .94}{pink}: {\tt Constrained-KMeans} initialized at true centroids of labels.}
\caption{Lloyd's iterations before convergence plotted as a function of the level of supervision for 100 Monte Carlo replicates. Shading indicates $\pm$ two standard deviations. Colors indicate algorithm:\\ \textcolor[rgb]{.62, .65, .03}{gold}: {\tt Constrained-KMeans} (without Lloyds iterations);\\ \textcolor[rgb]{0.0, .72, .96}{blue}: {\tt ss-k-means++} (without Lloyds iterations);\\ \textcolor[rgb]{1.0, .44, .42}{red}: {\tt Constrained-KMeans};\\ \textcolor[rgb]{0.0, .76, .53}{green}: {\tt ss-k-means++}; and\\ \textcolor[rgb]{.91, .40, .94}{pink}: {\tt Constrained-KMeans} initialized at true centroids of labels.}
\caption{Average ARI shown as a function of the level of supervision for 100 Monte Carlo replicates. Shading indicates $\pm$ two standard deviations. Green beats red. Colors indicate algorithm:\\ \textcolor[rgb]{.62, .65, .03}{gold}: {\tt Constrained-KMeans} (without Lloyds iterations);\\ \textcolor[rgb]{0.0, .72, .96}{blue}: {\tt ss-k-means++} (without Lloyds iterations);\\ \textcolor[rgb]{1.0, .44, .42}{red}: {\tt Constrained-KMeans};\\ \textcolor[rgb]{0.0, .76, .53}{green}: {\tt ss-k-means++}; and\\ \textcolor[rgb]{.91, .40, .94}{pink}: {\tt Constrained-KMeans} initialized at true centroids of labels.}
\caption{Power coefficient of WTS and deviations due to $\gamma$:\\[0.1cm] % {\protect\tikz{\protect\draw[very thick,draw={rgb,1:red,0.0;green,0.8;blue,0.8}] (0,-0.5ex)(0,0)--(4ex,0);}} $0.9 \frac{\cP(\lambdaoptstar,\betaopt)}{\left(\lambdaoptstar\right)^3} \lambda^3$ \\[0.1cm] % {\protect\tikz{\protect\draw[very thick,draw=blue] (0,-0.5ex)(0,0)--(4ex,0);}} $1.0 \frac{\cP(\lambdaoptstar,\betaopt)}{\left(\lambdaoptstar\right)^3} \lambda^3$ \\[0.1cm] {\protect\tikz{\protect\draw[very thick,draw={rgb,1:red,0.42;green,0.65;blue,0.8}] (0,-0.5ex)(0,0)--(4ex,0);}} $1.1 \frac{\cP(\lambdaoptstar,\betaopt)}{\left(\lambdaoptstar\right)^3} \lambda^3$}
\caption{NIKEL\_AMC board picture. It has dimensions of$\rm 18 \times 15 \times 3\,cm^3$. \label{PictureNikelAMC}}
\caption{NIKEL\_AMC board overview. Radio-frequency parts are colored in orange.\label{NIKEL_AMC_Hw}}
\caption{Pictures of the two boards composing the CCSB mounted on the NAT\textregistered \MCH. The main CCSB board is the top board in the left picture. The extension PCB providing the SMA connectivity with the front panel for the reference clock and the PPS is not mounted.\label{CCSBhw}}
\caption{Illustration of the PCB effects on the band 0 frequency response. It can be seen that a variation of 500\,MHz induces an amplitude drop of 6\,dB. The plot was obtained by varying the LO between 1\,GHz and 2\,GHz in 100\,MHz steps and recording the system response for each measurement. The 11 data sets are plotted in this figure (one color per set).\label{band0_resp-crop}}
\caption{ Average noise per tone between 10\,Hz and 11\,Hz without (red) and with (green) amplitude compensation. The 400 tones are evenly distributed over the 500\,MHz bandwidth.\label{bruit10Hz}}
\caption{Architectural diagram for \anywaredc{}.}
\caption{Energy efficiency of \anywaredc{} compared to unmodified Anyware and a typical desktop-dominated computing environment.}
\caption{Hours of generator use by \anywaredc{}, \anywareups{} and a desktop-based enterprise for varying power outage probabilities.}
\caption{Effect of laptop and UPS battery backup on the energy efficiency of \anywaredc{}. Each hour, the probability of power availability is 50\%.}
\caption{ Monte Carlo simulations confirm that, in \textit{absence} of dipolar interactions, the entropic probability distribution for a unique pair of topological defects to be separated by a vector $\vec R$ follows an exponential scaling law as in Eq.~(\ref{eq:entropicP}). This scaling law for the ``2 in - 2 out'' spin-ice phase (black solid line) is known exactly~\cite{Castelnovo11a} and serves as a validity check of our simulations (\textcolor{green}{$\triangle$}). As for the FCSL (\textcolor{blue}{$\square$} and \textcolor{red}{$\bullet$}), the exponential scaling law is found to be characterized by a parameter $\alpha_{\rm FCSL}=0.473\pm0.005$ (dashed line). For the sake of completeness, the two cases where the initial pair of defects out of the FCSL were ``2 in - 2 out'' (\textcolor{red}{$\bullet$}) and ``4 in / 4 out'' (\textcolor{blue}{$\square$}) were considered separately, but no pertinent difference was noticed. The $y-$axes are on a logarithmic scale, while the $x-$axes are on a a) linear and b) logarithmic scale to respectively emphasize the exponential scaling law and the long distance behavior. All data have been arbitrarily normalized to 1 at $R=1$, $r_{d}$ being the shortest distance between charges. The error bars are smaller than the data symbols. }
\caption{Averaged potential energy between a) two ``2 in - 2 out'' defects (\textcolor{red}{$\circ$}), b) ``4 in'' and ``4 out'' defects (\textcolor{blue}{$\triangle$}), c) ``2 in - 2 out'' and ``4 in / 4 out'' defects (\textcolor{green}{$\square$}), as illustrated in the inset of each panel, obtained from Monte Carlo simulations with dipolar interactions. ``2 in - 2 out'' tetrahedra are colorless, while the ``4 in / 4 out'' tetrahedra are represented by thick lines. The green spins in panel c) represent the double-spin motion responsible for the creation of the ``2 in - 2 out'' and ``4 in / 4 out'' defects. To compare the simulations with Eq.~(\ref{eq:Vpot}) (dashed lines), the data have been shifted by a reference energy, serving as fitting parameter, whose origin is discussed in Eqs.~(\ref{eq:DEh}$-$\ref{eq:DEhm}). There are no multiplicative prefactors, meaning that all defects carry a unit of magnetic charge. The dipolar energy scale $D$ is set to 1. The error bars, computed from the results of 100 independent samples, are smaller than the data symbols.}
\caption{At very low energy scale, dipolar interactions ultimately lift the FCSL degeneracy at $T_{c}/D=0.166\pm 0.003$. a) The first-order nature of the transition is clearly visible in the specific heat $C_{h}$ (\textcolor{red}{$\bullet$}) and in the discontinuity of the R-state order parameter (\textcolor{blue}{$\blacktriangle$}). The data are obtained from Monte Carlo (MC) simulations within the FCSL ensemble, averaged over 10$^{7}$ worm updates and 10 independent initial configurations, for a system size of 3456 pyrochlore sites. b) The hysteresis of the transition makes it difficult to precisely estimate the transition temperature. This is why in order to show convergence of the simulations, two kinds of equilibration processes were used. The system was either \textit{quenched} into one of the the R-states ($\circ$) or slowly annealed from high temperature ($\triangle$), which respectively over- and under-estimate the transition temperature. The MC time $t_{MC}$ of the $x-$axis represents the number of worm updates which are accepted/rejected based on a Metropolis argument accounting for the dipolar interactions. For both equilibration processes, the simulations were equilibrated at the temperature of measurements during $t_{MC}/10$. The results are given for different system sizes, showing convergence of the simulations towards $T_{c}/D=0.166\pm 0.003$. }
\caption{\label{fig:leed_structure}% (a) Micro-LEED pattern taken on the edge of a \typed{1} island grown on a Ge(001) epilayer, averaged over an electron energy range from 3 to 33\,eV. (b) LEED spot positions expected for different unit meshes: large black dots ($\bullet$) dots correspond to Ge(001)-(\mesh11); plus signs~({\bfseries+}) mark superstructure spots from Ge(001)-(\mesh12) and (\mesh21); large \reddish{} diamonds~(\protect\symbolip) originate from the small oblique unit mesh labeled ``1p'' in the top view onto the Co-terminated \CoGeii(111) surface in frame (c); small blue dots~(\protect\symboliiip) mark the additional spots expected for the large oblique unit mesh labeled ``3p'' in (c). (c) Top view and (d) side view of the bulk-terminated \CoGeii(111) surface according to the bulk structure reported in Ref.~\citen{Schubert_NW1948}. Blue (dark) spheres correspond to Co, amber (bright) spheres to Ge atoms. }
\caption{Predictions for $T_{c}$ in a dipolar Fermi gas. $\lambda$ is the dimensionless average distance between dipoles given by $(6\pi^{2})^{1/3}/(k_{F}\alpha C_{dd}m\hbar^{-2})$. The straight line at the top is the position of the Fermi energy, $T=T_{F}$. The solid blue line is our full numeric calculation. The dashed red line is the theoretical prediction given in Ref.~\citep{2002-PRA_Baranov-Shly_BCSDipRenorm}. The markers are the locations of current experiments: $\protect\mr{^{40}K^{87}Rb}$\citep{2010-Nature_JILA_UColdDipFermiGas} at the theoretical maximum polarisation (\textifsymbol[ifgeo]{49}), $\protect\mr{^{40}K^{87}Rb}$\citep{2010-Nature_JILA_UColdDipFermiGas} at current experimental polarisations (\textbigcircle ) $\protect\mr{^{23}Na^{40}K}$\citep{2015-PRL_Harvard_NaKGstate} at the theoretical maximum polarisation (\textifsymbol[ifgeo]{97}), $\protect\mr{^{161}Dy}$\citep{2012-PRL_Lu-Burdick-Lev_DyFermiGas} (\textifsymbol[ifgeo]{54}),$\protect\mr{^{167}Er}$ \citep{2014-PRL_Innsbruck_ErDeg} (\textifsymbol[ifgeo]{102}), and $\protect\mr{^{53}Cr}$\citep{2015-PRA_CNRS_DegenCr} is not plotted as it sits too far off to the right. Inset: $T_{c}$ vs $\lambda$ zoomed in at the peak in $T_{c}$. The bottom two lines (Red) are with just $l=1$. The top two lines (blue) are with $l=1,3$. The dashed lines are the results with exchange interactions switched off. The upper solid (blue) line in the inset corresponds to the solid (blue) line in the outer plot.\label{fig:Tc}}
\caption{Comparison among the energy detector, matched filter detector, and feature extraction detector.}{\includegraphics[scale=0.7]{Tab1.eps}\label{Tab1}}
\caption{{System Parameters Used in Numerical Evaluation.}}{\includegraphics[scale=0.6]{Tab2.eps}\label{Tab2}}
\caption{{\bf Spin-spin correlations in the vicinity of the spin liquid phases} for $x=-0.9$ (a,b) $-1$ (c,d) and $-1.05$ (e,f), obtained from Monte Carlo simulations. The temperatures considered are $T=0.01$ (\textcolor{green}{$\blacklozenge$}), 1 (\textcolor{orange}{$\blacksquare$}) and 891.25 (\textcolor{blue}{\Large \textbullet}). Because of the anisotropy of the lattice, we want to separate correlation functions which start on A-sites (a,c,e) and B-sites (b,d,f). The radial distance is given in units of the unit-cell length. The agglomeration of data points around $C\sim 2. 10^{-5}$ is due to finite size effects. The y-axis is on a logarithmic scale. }
\caption{ Figure adapted from \citet[][]{Casassus2015ApJ...812..126C} . {\bf Upper row:} multi-frequency data of HD~142527 brought to a common $uv$-coverage. The ALMA data have been filtered for the ATCA response with Monte-Carlo simulations of ATCA observations on deconvolved models of the ALMA data. \textcolor{blue}{\bf a)}: restored ATCA image at 34~GHz. \textcolor{blue}{\bf b)}: average of Monte-Carlo (MC) simulations of ATCA observations on the ALMA band~7 data. \textcolor{blue}{\bf c)}: MC simulations of ATCA observations on the ALMA band~9 data. {\bf Lower row:} Grey-body diagnostics inferred from the multi-frequency data. \textcolor{blue}{{\bf a}}: optical depth map at the reference frequency of 345~GHz. \textcolor{blue}{{\bf b}}: line of sight emissivity index map $\beta$, with ATCA specific intensity contours in red. \textcolor{blue}{{\bf c}}: root-mean-square uncertainties on the emissivity index map. {\bf d}: line of sight temperature, $T_s$. \label{fig:Ttaubet} }
\caption{ Gallery of optical/IR spirals. From left to right, we show scattered-light images of AB~Aur \citep[][part of their Fig.~3, \copyright AAS, reproduced with permission]{Fukagawa2004ApJ...605L..53F} , HD~141569A \citep[][part of their Fig.~8, \copyright AAS, reproduced with permission]{Clampin2003AJ....126..385C}, HD~135344B \citep[][part of their Fig.~1, \copyright ESO, reproduced with permission]{Garufi2013AA...560A.105G}, and HD~100453 \citep[][part of their Fig. 2, \copyright AAS, reproduced with permission]{Wagner2015ApJ...813L...2W}. All figures have been reproduced by permission of the AAS or A\&A. In HD~100453 we have highlighted two intensity dips where the two-armed spiral pattern seem to stem from. These dips are very reminiscent of the HD~142527 shadows, which also seem to be at the root of spirals.}
\caption{\red{A spectrogram depicting a segment of the speech signal emitted from Cnut's mouth as recorded by the monopole sensors.} \label{fig:spect}}
\caption{Parameter values in post\red{-}processing.}
\caption{Spectral energy distributions of {\gray}s obtained for the northern (left panel) and southern (right panel) extended residual templates. % The points with error bars are derived from the fitting procedure with (black symbols) and without (red symbols) subtracting the contribution from the unassociated 3FGL catalogue sources. The curves represent the best spectral fits for the hadronic ($\pi^0$ - decay) channel (for details see the text). % % }
\caption{Spectral energy distribution of {\gray}s obtained for the NFW templates. % % % }
\caption{ Radial profiles of the \gray emissivity for the southern parts of the excess. For comparison, we show the constant (solid curves), $1/r$ (dotted curves), and $1/r^2$ (dot-dashed curves) profiles, which are predicted by the impulsive injection, continuous injection, and the transport through wind, respectively. Left panel: the emissivity per hydrogen atom related to interactions with gas (i.e. inelastic interactions of protons and ions or bremsstrahlung of electrons). Right panel: the \gray emissivity due to IC normalized to the density of ISRF $1~\rm eV/cm^3$. % % }
\caption{The anomaly detection algorithm is applied to a real performance metric collected over three months. \textcolor{red}{Red} boxes highlight the periods in which the incidents occurred. \textcolor{green}{Green} diamonds are the data points flagged as anomalous by our tool. The upper and lower bounds are determined by $\mu \pm \beta \sigma$ in Algorithm~\ref{alg:anomaly-detect}.}
\caption{\label{MonoPart_HcEvol_azimuth} Azimuthal behavior of the reduced coercive field \Hcr for a single F particle at $T =$~\SI{0.01}{K}~(\rond), \SI{10}{K}~(\carre) and \SI{30}{K}~(\pentagon). The arrows identify the coercive apexes positions. }
\caption{ \label{Distrib_FMpart_azimuth} On the right quarter, azimuthal behavior of size-distributed particles for $T =$ \SI{0.01}{K} (\rond), \SI{20}{K} (\carre), \SI{50}{K}(\pentagon) and \SI{100}{K}(\trianglemark). On the left quarter, for comparison, are plotted \Hcr for a $\SI{8}{nm}$ single particle for $T =$~\SI{0.01}{K}~(\trait{red}), \SI{10}{K}~(\traitun{blue}) and \SI{30}{K}~(\traitdeux{magenta}). }
\caption{ \label{fig:S_FMAF_az} (a) \Hc and (b) \He angular thermal evolution of size-distributed core-shell particles with the (\Kaf,\Jeb) set of parameters. (c) \Hc and (d) \He for the second set $(\Kaf^*,\Jeb^*)$. Temperatures are $T = $ \SI{0.01}{K}~(\rond), \SI{10}{K}~(\carre), \SI{20}{K}~(\pentagon), \SI{50}{K}~(\trianglemark), \SI{100}{K}~(\pentagondown) and \SI{200}{K}~(\carrebis). }
\caption{\colornote{}Sphere radii $R^*$ that minimise the potential energy for $N$ particles interacting either through a Lennard-Jones (LJ) or a Morse (Morse) potential, as fraction of the estimated radius that would tightly pack $N$ hard disks of diameter $d_0$, $d_0\sqrt{N/\phi_M}/4$. Note that the LJ data coincides well with the results of Voogd.\cite{voogd-thesis} The largest difference between the two is no more than $0.02r_0$ ($<2\%$). \label{fig:min-radii}}
\caption{\colornote{}Temperature dependence of the excess defect fraction, the number of particles with other than six nearest neighbours beyond the first twelve, for $N=10$ to $N=100$ Lennard-Jones particles, using the distance criterion. \label{fig:lj-dist-defects}}
\caption{\colornote{}The five Lennard-Jones packings for $N=72$ with the lowest potential energy found using the GMIN program \cite{gmin} (a): $D_{5h}$ packing with energy per particle $U/N = -3.0564\epsilon,$ (b): $D_3$ packing with $U/N = -3.0559\epsilon,$ (c): icosahedral packing with $U/N = -3.0548\epsilon,$ (d): tetrahedral packing with $U/N = -3.04636\epsilon$ and (e): packing with two times three rectangular patches that wrap around the sphere similar to the seam on a baseball, with $U/N = -3.04630\epsilon.$ Top and bottom show different orientations. The colour coding indicates the coordination numbers five (blue) or six (red). \label{fig:struct-mismatches-72}}
\caption{\colornote{}Excess defect fraction for $N=10$ to $N=100$ Lennard-Jones particles determined by means of the convex hull, as described in Section \ref{sec:methods}. \label{fig:lj-hull-defects}}
\caption{\colornote{}Excess defect fraction for $N=38,~44,~48$, 60, 72 and 92. Lennard-Jones particles as function of temperature. Note the clear re-entrance of excess defects for with increasing temperature for $N\neq 24.$ \label{fig:lj-some-defects}}
\caption{\colornote{}Global potential energy minimum for $N=24$ Lennard-Jones particles, shown from two vantage points (a) and (b), obtained using the GMIN program.\cite{gmin} The colour indicates the number of nearest neighbours. Red particles have 6 and blue 5 nearest neighbours, as identified by the convex hull and the distance criterion. % Particles are shown at half their effective size and are connected to their nearest neighbours with grey bonds. % The convex hull identifies some particles as having six-fold coordination that the distance criterion considers five-fold defects. \label{fig:struct-mismatches-24} }
\caption{\colornote{}Probability of encountering an icosahedral, $D_{5h}$ or $D_3$ packing with $N=72$ Lennard-Jones particles on a sphere with radius $R=2.55037 r_0,$ where $r_0$ is the equilibrium spacing of the pair potential, as function of the dimensionless temperature $k_B T / \epsilon.$ The frequencies do not sum to unity because for some time frames the packing could not be identified. \label{fig:freqs}}
\caption{\colornote{}Free energy differences between packings of $N=72$ particles on a radius $R=2.55037 r_0.$ At low temperatures the $D_3$ and $D_{5h}$ packings are nearly equal in free energy, but an increasing importance of entropy destabilises the $D_3$ packing more than the $D_{5h}$ at higher $T.$ Both the $D_{5h}$ and $D_3$ packing are destabilised at higher $k_B T$ in favour of the icosahedral packing. At $k_B T \approx 0.032\epsilon$ the three packings appear to be equally probable. \label{fig:dFs}}
\caption{\colornote{}Temperature dependence of the excess defect fraction using the distance criterion for $N=10$ to $N=100$ Morse particles with a range parameter of $\alpha = 60/r_0.$ \label{fig:morse-dist-defects}}
\caption{\label{fig:semiparinf} Treatment effect estimates in four A/B trials from eBay. The Bayesian posterior is the region in {\bf\color{blue}blue}, Winsorized estimation is in {\bf black}, and the naive sample estimator is in {\bf\color{red}red}. Points and lines are point estimates and intervals are $\pm1\mathrm{sd}$ or $\pm1\mathrm{se}$, as appropriate. }
\caption{\small Our two verification strategies for some images of the dog class. Yes/No verification (left): verify a detection as either correct (\textcolor{green}{Yes}) or incorrect (\textcolor{red}{No}). YPCMM verification (right): label a detection as \textcolor{green}{Yes}, \textcolor{yellow}{Part}, \textcolor{cyan}{Container}, \textcolor{magenta}{Mixed} or \textcolor{red}{Missed}.}
\caption{\small Visualisation of \textcolor{blue}{search space} reduction induced by YPCMM verification on some images of the cat class (\textcolor{yellow}{part}, \textcolor{cyan}{container}, \textcolor{magenta}{mixed}, and \textcolor{red}{missed}). In the last row, the search space reduction steers the re-localization process towards the small cat on the right of the image and away from the dog on the left.}
\caption{\emph{Upper panel:} longitudinal (\color{blue}$\circ$\color{black}) and transverse ({\footnotesize{\color{red}$\times$\color{black}}}) time-dependent diffusivities $D_\|$ and $D_\perp$ in the particle frame, as measured in simulation (a) and experiment (b). Their common plateau gives the diffusion constant $D$, and the late-time slope of $D_\|(t)$ the propulsion speed $v_\text{p}$. \emph{Lower panel:} histograms of the corresponding short-time displacements. }
\caption{The results of the $R$ = 1 simulated spectral fitting from Test A, where all 6 parameters were free to vary. Plots show input parameter values on the abscissa and the measured values on the ordinate. The dashed 1:1 line represents a perfect measurement and the dotted lines denote allowed input ranges. Simulated spectra were produced using the \xmm pn response and fit from 2.5 -- 10\keV\using the XSPEC model\textsc{relxill} for a collection of randomly-generated input parameters. Each data point corresponds to a modelled spectrum with \redchisq $<$ 1.1. The data were binned by input value and overplotted as blue bands, the widths of which represent 1$\sigma$ error.}
\caption{Summary of the $R$ = 1 results for all four tests in the 2.5--10\keV\band. Spectra were binned with respect to input values and plotted with the central solid lines showing the data. The 1$\sigma$ errors for the Test A (small blue dots), Test B (large green hexes), Test C (small pink crosses), and Test D (small cyan hexes) data are illustrated as opaque colored bands. Tests where certain parameters remained fixed are not plotted for that respective parameter (e.g. Test C, D for $q_{1}$ and Test B, D for $\xi$).}
\caption{$N_L(\Gamma)$ at the three largest values of $L$ studied for $n_v=0.05$. Circles demarcate the crossover region centered at the crossover scale $\Gamma_c$. Data for $\Gamma \lesssim \Gamma_c$ fits well to power-law form $N_{\rm 1D}(\Gamma)$ \textcolor{blue}(see text for details) with the value of $y$ indicated in the figure, while the large-$\Gamma$ regime fits well to the modified Gade-Wegner form $N_{\rm GW}(\Gamma)$.}
\caption{Dependency relationships between various functions in \texttt{AM}: an arrow pointing from A to B means that A depends on B. Color of a function box denotes its type - \textcolor{green}{green}: initializing; \textcolor{red}{red}: updating; \textcolor{blue}{blue}: interfacing and \textcolor{yellow}{yellow}: accessor function. Shape of a function box represents the part of \texttt{AM} it belongs to - rectangle: Image Operations; rounded rectangle: Similarity Functions; ellipse: Distance Feature. The numbers attached to some of the nodes refer Table \ref{tab_am_func_specification}.}
\caption{\label{fig:fig5} (a) SEM image of a STOF. (b): Microscope image of a trapped fluorescent particle with an outline of the STOF for clarity \textcolor{blue}{(see visualization 1 \& 2)}. (c) and (d) show SEM images of the fiber after the experiment was performed. Particles can be seen inside the slot was well as on the surface. (e): Particle position versus time along the \textit{z}-axis of the STOF. The particle is seen to spend most of its time near the slot center. Each pixel was found to correspond to a 100 nm $\times$ 100 nm area and Gaussian fits to the particle center enable high resolution tracking. (f): Histogram of the particle positions given in (e) showing bunching at regular intervals.}
\caption{{\it \red{Ray-traced} $\nu L_\nu$ spectra: impact of directionality}. The local spectra have been computed with GR illumination. The \textbf{top row} shows emission-angle-averaged (dotted) and angle-dependent (solid) ray-traced spectra. \red{The straight black dashed line shows the level of the direct-component power law (see text for details).} The \textbf{middle row} shows the relative difference between averaged and angle-dependent ray-traced spectra. The \textbf{bottom row} shows the same quantity as the middle row for the local spectra, evaluated at $r=r_{\mathrm{min}}$. The spin is $0$ (left column) or $0.98$ (right column). Red curves refer to spectra ray traced from an inclination of $\theta=5^\circ$, i.e. close to face-on. Green curves are computed at an inclination of $\theta=85^\circ$, i.e. close to edge-on. Note that "face-on" and "edge-on" in the two upper rows (ray-traced quantities) refer to the inclination angle $\theta$, while in the bottom row (local quantities) these words refer to the emission angle $i$. }
\caption{\label{fig:spectrum_observables_microcanonical}(Color online) Dependence of the algebraic connectivity (a) and the spectral radius (c) on the energy of the triangulation normalized by the energy of the random triangulation for $4\times4$-triangulations (\protect\includegraphics{point_triangulation_size_4.pdf}), $8\times 8$-triangulations (\protect\includegraphics{point_triangulation_size_8.pdf}), $16\times 16$-triangulations (\protect\includegraphics{point_triangulation_size_16.pdf}) and $32\times 32$-triangulations (\protect\includegraphics{point_triangulation_size_32.pdf}). The solid lines are fits for $E/\langle E_{\mathrm{rnd}} \rangle \in [0,1.5]$. The eigenvalues are normalized by their values at ground states of the respective system size. (b,d): Gradients of the fits are plotted against the lattice size of the triangulations, the purple points are the approximation for the spectral radius using first order perturbation theory with periodic ground state from Eq.~\eqref{eq:sr_approximation_microcanonical_gradient}.}
\caption{\label{fig:spectrum_observables_random}(Color online) Algebraic connectivity $\lambda_1$ (a) and spectral radius $\lambda_{MN-1}$ (b) of random triangulations (\protect\includegraphics{point_triangulation.pdf}), \erdoes (\protect\includegraphics{point_erdoes_reyni.pdf}), Newman-Watts (\protect\includegraphics{point_newman_watts.pdf}) and \barabasi (\protect\includegraphics{point_barabasi_albert.pdf}) random graphs for different number of vertices $MN$. The displayed lines are power law fits for the numerical data.}
\caption{\label{fig:inverse_participation_ratio_random}(Color online) Scaling of the inverse participation ratio (IPR) of the Laplacian spectrum of random triangulations. Expectation values of the average IPR $\langle \overline \chi \rangle$ (a), the IPR $\langle \chi_1 \rangle$ of the algebraic connectivity $\lambda_{1}$ (b) and the IPR of $\langle \chi_{MN-1} \rangle$ of the spectral radius $\lambda_{MN - 1}$ of random triangulations (\protect\includegraphics{point_triangulation.pdf}), maximal ordered triangulations (\protect\includegraphics{point_triangulation_gs.pdf}), \erdoes (\protect\includegraphics{point_erdoes_reyni.pdf}), Newman-Watts (\protect\includegraphics{point_newman_watts.pdf}) and \barabasi (\protect\includegraphics{point_barabasi_albert.pdf}) random graphs for different number of vertices $MN$. The displayed lines are power law fits for the numerical data. Probability density funtion $p(\chi)$ of the IPR for random triangulations for different system sizes (d) and for the different random graphs for $32^2$ vertices (e). }
\caption{\label{Tab1}The estimated observables $\theta _{\infty }$, $s$, $r$ and the SDL coefficients $u_m$, $\bar{a}$, $\bar{b}$ for the Schwarzschild, the charged Galileon, the tidal RN and the RN black holes. We assume these black holes have the same mass and distance as Sgr A* with $M_{\bullet}=4.31\times10^6$ $M_\odot$ and $D_{\mathrm{OL}}= D_{\bullet}=8.33$ kpc \cite{Gillesse2009ApJ707.L114}. \textcolor[rgb]{1,0,0}{$\tilde{Q}$ is the tidal charge parameter which has a positive value for the RN black hole and has a negative value for the tidal RN black hole.} $\theta_{\infty}$ and $s$ are respectively in the units of micro-arcsecond ($\mu$as) and nano-arcsecond (nas). The unit of $u_m$ is Schwarzschild radius $R_s=2GM_{\bullet}/c^2$. $r$, $\bar a$ and $\bar b$ are dimensionless.}
\caption{\label{Pic2} The estimated observables $\theta_\infty$, $s$ and $r$ of the charged Galileon black hole, the RN black hole \red{and the tidal RN black hole}. It is assumed that they have the same mass and distance as Sgr A*. Diagram (a), (b) and (c) represent $\theta_\infty$, $s$ and $r$ against the parameter $\Gamma$, where the solid lines refer to the charged Galileon black hole, \red{the dashed and dot dashed lines refer to the RN black hole and the tidal RN black hole when the tidal charge parameter $\tilde{Q}=\Gamma$. The charged Galileon, the RN and the tidal black holes have identical $\theta_{\infty}$ so that their curves coincide with each other in diagram (a).}}
\caption{\label{Pic3} The SDL coefficients $u_m$, $\bar{a}$, $\bar{b}$ of the charged Galileon black hole, the RN black hole \red{and the tidal RN black hole}. These black holes are also assumed to have the same mass and distance as Sgr A*. Diagram (a), (b) and (c) represent the SDL coefficients against the parameter $\Gamma$, where the solid lines refer to the charged Galileon black hole, \red{the dashed and dot dashed lines refer to the RN black hole and the tidal RN black hole where the tidal charge parameter is set as $\tilde{Q}=\Gamma$. The charged Galileon, the RN and the tidal black holes have identical $u_m$ so that their curves coincide with each other in diagram (a).}}
\caption{\label{Pic4} Diagrams (a), (b) and (c) show $\Delta T_{n,m}$, $\Delta T_{n,m}^{\ 1}$ and $\tilde\eta_{n,m}$ of the charged Galileon black hole, the RN black hole and \red{the tidal RN black hole} with respect to $\Gamma$. These black holes are also assumed to have the mass $M_{\bullet}$ and the distance $D_{\bullet}$. The solid lines refer to the charged Galileon black hole, the dashed and dot dashed lines represent the RN black hole \red{and the tidal RN black hole} when the tidal charge parameter $\tilde{Q}=\Gamma$. The color of the lines and the corresponding values of $m$ and $n$ can be seen in the legend chart.}
\caption{{(a) Expert-defined hierarchy (classes with high degree of similarities are marked with symbols \begingroup\color{orange}$\bigstar$\endgroup, \begingroup\color{red}$\CIRCLE$\endgroup) modified using various methods: (b) Agglomerative clustering with cluster cohesion to restrict the height to original height \cite{li2007hierarchical} (c) Global-INF flattening method \cite{naik2016inconsistent} (d) Proposed rewiring method. Modified structure changes are shown in green color.}}
\caption{ECM for autonomous robotic playing\textcolor{red}{todo}}
\caption{Left: Simulated SED (histogram) at three different time epochs for our benchmark case (Case 1). The data points include a \gray flare of March 19, 2001, observed by Whipple, and the simultaneous \xray data from \rxte (pre-flare in blue, flare peak in cyan). \xray and optical high/low states data taken by \rxte and Mt Hopkins 48-inch telescope in 2001 (dark gray), and \xray high/low states data from BeppoSAX in 1998 and 2000 (light gray) are also shown \citep{fossati_2008:xray_tev}. In other bands there are \wise infrared data from May 21, 2010 in magenta, and \fermi 4-year average \gray data \citep{fermi_2015:4year_catalog:218.23} in yellow. Right: Samples of normalized simulated light curves in 6 different frequency bands. Observational light curves are plotted as connected dots. They include simultaneous 2-4 keV \rxte X-ray (grey) and TeV Whipple/HEGRA \gray (dark blue) data. The time stamp 100\ks in the figure corresponds to MJD 59186.9 or March 18, 2001.}
\caption{Left: Full light curves for the entire simulation Case 1. Right: PSD based on the light curves excluding the first 100\ks. The vertical lines indicate various timescales in the system (times$2\pi$), namely acceleration decay as blue dashed line, light crossing as blue dot-dashed line, and a red/green long dashed lines representing synchrotron cooling for electrons that emit synchrotron radiation primarily in 2-keV \xray/ 1-eV optical band. The gray dashed line indicates a power-law with index of -2. We also plot the observational PSDs of \mrk. They include \xray data from \maxi (squares) and \asca (circle) as maroon points in the lower panel \citep{isobe_2015:maxi_421_psd:798.27} and GeV \gray from \fermi as brown points in the upper panel.}
\caption{Correlation analysis for Case 1. Left: Flux-flux correlation between the \xray and \gray bands during the time period 1200ks-1500ks. The first 30\ks are plotted in red, and the last 30\ks are plotted in cyan. The grey connected circles are observational data points for\mrk on March 19, 2001 from \citet{fossati_2008:xray_tev}. Three grey dashed lines are also plotted so that the readers can compare them to the trend of the flux-flux relationship. Right: Cross-correlation between \xray and TeV \gray fluxes based on the light curves, excluding the first 100\ks. The confidence interval for the lag based on the correlation function is marked with the horizontal solid blue line, while the estimated lag is identified with the dashed blue vertical line.}
\caption{Left: Quantum efficiency as a function of wavelength for the \twentysquare (Na$_2$KSb) photocathode on the first fully-assembled \LAPPDTM module while still under vacuum in the SSL tank. The tube was fully functional, with good gain and uniformity. When brought up to air one corner of the seal between the module assembly and the window leaked due to misalignment. }
\caption{Collusion Matrix of the Prolog program. \textcolor{truepositive}{$\clubsuit$} = Information theft. \textcolor{truepositive}{\textdollar} = Money theft. \textcolor{truepositive}{$\spadesuit$} = Service misuse. \textcolor{falsepositive}{$\clubsuit$}, \textcolor{falsepositive}{\textdollar}, \textcolor{falsepositive}{$\spadesuit$} = False positives.}
\caption{Our proposed model: given an image, a CNN is first applied to produce the attribute-based representation \textcolor{red}{$\Att(I)$}. The internal textual representation is made up of image captions generated based on the image-attributes. % The hidden state of the caption-LSTM after it has generated the last word in each caption is used as its vector representation. These vectors are then aggregated as \textcolor{green}{$\Cap(I)$} with average-pooling. The external knowledge is mined from the KB and the responses are encoded by Doc2Vec, which produces a vector \textcolor{blue}{$\Know(I)$}. The 3 vectors $\mathbf{V}$ are combined into a single representation of scene content, which is input to the VQA LSTM model that interprets the question and generates an answer. }
\caption{BLEU-1,2,3,4 and $\mathcal{PPL}$ metrics compared to other state-of-the-art methods and our baseline on Flickr8k and Flickr30K dataset. $\ddagger$~indicates ground truth attributes labels are used, which (in \colorbox[rgb]{0.7,0.7,0.7}{gray}) will not participate in rankings. Our $\mathcal{PPL}s$ are based on Flickr8k and Flickr30k word dictionaries of size 2538 and 7414, respectively.}
\caption{BLEU-1,2,3,4, METEOR, CIDEr and $\mathcal{PPL}$ metrics compared to other state-of-the-art methods and our baseline on MS COCO dataset. $\ddagger$~indicates ground truth attributes labels are used, which (in \colorbox[rgb]{0.7,0.7,0.7}{gray}) will not participate in rankings. Our $\mathcal{PPL}s$ are based on MS COCO word dictionaries of size 8791.}
\caption{We matched DXL bands (dotted) as closely as possible to the ROSAT bands (solid) for direct comparison. The {\it DXL} bands are labeled as {\color{black} D1 (black)}, {\color{red}D2 (red)}, {\color{green}D4 (green)}, {\color{blue}D5 (blue)}, {\color{cyan}D6 (cyan)} and {\color{magenta}D7 (magenta)} in correspondence to {\it ROSAT}'s {\color{black} R1 (black)}, {\color{red}R2 (red)}, {\color{green}R4 (green)}, {\color{blue}R5 (blue)}, {\color{cyan}R6 (cyan)} and {\color{magenta}R7 (magenta)} bands. The bands are defined by the grasp (a product of effective area to solid angle in cm$^2$~sr as a function of energy.}
\caption{Properties of the set of states $\{ 26,27,28 \}$ in the interval $t[-0.014,-0.2]$: (a) Energies, (b) Fidelity, (c) Entropy and (d) Maximum local and normal components. \textcolor{red}{ In panel (d)} full figures corresponds to maximum local components, while the corresponding empty figure refers to the normal component.}
\caption{Semi-supervised verified feedback generation: {\color{green}\ding{51}} is a submission verified to be correct, {\color{red}\ding{55}} is a faulty submission for which feedback is generated and ? is an unlabeled submission for which feedback is not generated.}
\caption[Stacked radial profiles.]{Radial profile of the stacked sample up to $r_{200}$, using all galaxies (black) and red sequence galaxies (red). These profiles are well fit by NFW profiles with the red subsample somewhat more concentrated than the full sample, with concentrations of \RPredstacksol\and\RPallstacksol, respectively. \label{fig:rpall}}
\caption[Evolution of the concentration parameter.]{Cluster NFW concentration parameter evolution for red sequence selected galaxies (top panel) and all galaxies (bottom panel) within projected $r_{200}$. Grey points represent the individual cluster fits using Cash statistics, and the five black points are representative of the concentrations found by simultaneously fitting to ensembles of 5 clusters each. The central open circle corresponds to the concentration extracted from the fit of the stacked sample up to $r_{200}$ (see Fig.~\ref{fig:rpall}). Open squares in the bottom panel correspond to values found in the literature. There is some evidence for redshift evolution in the total sample given a slope of \CGallevosolslope, and for the red subsample given the slope of \CGredevosolslope. The apparent trend is consistent between both the stacked and the individual data.\label{fig:cg_evol} }
\caption[Individual and stacked luminosity function.]{ We plot 24 of the 26 individual LFs (top) versus $m-m^*_{\rm model}$, where $m^*_{\rm model}$ is the predicted CSP characteristic luminosity at the redshift of the cluster. Each individual LF is extracted using the band redward of the 4000\AA~break. The two excluded clusters included the lowest redshift system where our imaging is not adequate and another system that has a foreground star field, making it difficult to identify the faint galaxy population. The BCGs are excluded. The weighted averaged luminosity function appears below. In black the total population is shown, and in red the red-sequence population is displayed. Bins with at least two contributing clusters are shown. The fit for the all galaxies stacked is \phiallsol\and\alphaallsol\(\chiallsol). The fit for red-sequence galaxies is \phiredsol\and\alpharedsol\(\chiredsol). \label{fig:lf_stack}}
\caption[Evolution of red luminosity function parameters.]{Same as~Fig.\ref{fig:evolution} for red sequence galaxies. {\it Panel (a):} There is no significant evolution in $\Delta mag=(m^*_{\rm model} - m^*)$, indicating the CSP model provides a good description of cluster galaxies over this redshift range. {\it Panel (b):} Evolution of $\alpha$ is suggested by the data with best fit line having intercept \alpharedevointer\and slope\alpharedevoslope. {\it Panel (c):} $\phi^*/E^2(z)$ extracted when fixed \mstar\is consistent with no evolution at\phiredsignificancemfix $\sigma$ level. \label{fig:evolution_red}}
\caption{The thin-stack algorithm operating on the input sequence $\mathbf x = \text{(\word{Spot}, \word{sat}, \word{down})}$ and the transition sequence shown in the rightmost column. $S$[$t$] shows the top of the stack at each step $t$. The last two elements of $Q$ (underlined) specify which rows $t$ would be involved in a \reduce\operation at the next step.}
\caption{\protect\label{tab:results}Results on SNLI 3-way inference classification. Params.\is the approximate number of trained parameters (excluding word embeddings for all models). Trans.\acc.\is the model's accuracy in predicting parsing transitions at test time. Train and test are SNLI classification accuracy.}
\caption{Responses generated by the baseline (LSTM-MMI) and the Speaker Model for ten randomly selected users, without cherry picking. {\color{red} $\#$} indicates poor-quality responses produced by the system.}
\caption{{\color{blue}Shocks along the time evolution of three given profiles in 2-D.}}
\caption{\color{blue}Case 1: $\rho_0$ concentrated around $0.2$ with support on $0~\leq~ r_0~\leq~0.6$ and maximum value of $0.35$.}
\caption{\color{blue}Case 2: $\rho_0$ concentrated around $0.75$ with support on $0.5~\leq~ r_0~\leq~1.0$ and maximum value of $0.8$.}
\caption{\color{blue}Case 3: $\rho_0$ with support on $0~\leq~ r_0~\leq~1.0$ and maximum value of $0.4$.}
\caption{{\color{blue}Shocks along the time evolution of three given profiles in 3-D.}}
\caption{\color{blue}Case 1: $\rho_0$ concentrated around $0.2$ with support on $0~\leq~ r_0~\leq~0.6$ and maximum value of $0.35$.}
\caption{\color{blue}Case 2: $\rho_0$ concentrated around $0.75$ with support on $0.5~\leq~ r_0~\leq~1.0$ and maximum value of $0.8$.}
\caption{\color{blue}Case 3: $\rho_0$ with support on $0~\leq~ r_0~\leq~1.0$ and maximum value of $0.4$.}
\caption{{\color{blue}Onset of congestion.}}
\caption{{\color{blue}$\varepsilon = 1$} and $j$ from 0 to 1.5,}
\caption{{\color{blue}$\varepsilon = 0.01$} and $j$ from $0$ to $0.25$.}
\caption{{\color{blue}Dependence of congestion on viscosity.}}
\caption{\color{blue}Supercritical case,}
\caption{\color{blue}Subcritical case.}
\caption{{\color{blue}Critical current $j$ as a function of viscosity.}}
\caption{{\color{blue}Stationary and time-dependent solutions.}}
\caption{\color{blue}Dominance,}
\caption{\color{blue}Trend to the equilibrium.}
\caption{{\color{blue}Numerical example 1.}}
\caption{\color{blue}Density $\rho$,}
\caption{\color{blue}Eikonal solution $u$.}
\caption{{\color{blue} Numerical example 2.}}
\caption{\color{blue}Density $\rho$}
\caption{\color{blue}Eikonal solution $u$.}
\caption{{\red (Color online)} Dispersion of {\blue longitudinal} acoustic phonons propagating along the {\blue $[100]$} direction ({\red top} panel) and {\blue transverse} acoustic phonons propagating along the $[011]$ direction ({\red bottom} panel). The data have been collected at the scattering vector $\bm{Q} = \hbar (\bm{k}_{i}-\bm{k}_{f})$ specified in each panel, $\bm{k}_{i}$ and $\bm{k}_{f}$ being the wavevector of incident and scattered photons, respectively. %{\red if we have this in one column we have no problem placing the figs.!} \label{3D}}
\caption{(Color online) Constant-pressure heat capacity of NpO$_{2}$ (black solid line) obtained as the sum of the calculated vibrational contribution (blue dashed line), the dilation contribution calculated by the Gr\"{u}neisen relation (green dashed line), Ref.\[\onlinecite{serizawa01}], and the Schottky contribution due to crystal field (CF) excitations (blue dot-dashed line), calculated on the {\red basis} of the CF level scheme derived from inelastic neutron scattering experiments \cite{magnani05,fournier91,caciuffo92}. Experimental data are taken from Ref.\[\onlinecite{magnani05a}] (a, red solid circles), Ref.\[\onlinecite{westrum53}] (b, red open circles), and Ref.\[\onlinecite{benes11}] (c, red dot line). The lambda-type anomaly at 25 K is associated with the transition to the multipolar ordered phase of NpO$_{2}$. Experimental data for ThO$_{2}$ (d, blue open squares) are taken from Ref.\[\onlinecite{osborne53}]. \label{Cp-temperature}}
\caption{(Color online) Phonon linewidth distribution calculated along the high-symmetry lines for \textit{fcc} NpO$_2$ at 300 K (left panel) and accumulated phonon thermal conductivity $\tilde{\kappa}$ as a function of the phonon energy (right panel). Different colors in the linewidth distribution indicate different phonon branches, {\red whilst the $q$-dependent linewidth is depicted by the width of the branch.} The accumulated phonon thermal conductivity is {\red shown} {\cyan (right panel)} by the red line whilst its derivative with respect to the energy is depicted by the blue line. \label{acc-kappa}}
\caption{(Color online) GGA+$U$ at 0 K (red line), GGA+$U$+SOC at 0 K (green dashed line) and GGA+$U$ at 300 K (blue dashed line) calculated phonon dispersions (left panel) and corresponding phonon density of states (right panel) of NpO$_2$. {\red Note that the spin-orbit coupling has a relatively small influence on the \textit{ab initio} computed phonon dispersions, comparable to that of an increased temperature.} %The $\bm{q}$-point labels in the left panel are those for the standard high-symmetry positions of the fcc Brillouin zone \label{fig-SOC}}
\caption{{\color{\revA}I) A path graph with $N=27$, and II)-IV) generalized path graphs with $N=17$.}}
\caption{{\color{\revA}The same as in Fig~\ref{fig:boundary_effect_comp} but for the agents defined by (\ref{eq:mat_sim_1}) and (\arabic{multi_eq1}b).}}
\caption{ $H^1$ semi-norm errors: \textcolor{red}{$\circ$} is for $|u - u_h|_{H^1(\Omega)}$, \textcolor{blue}{$\star$} is for $|u - \widetilde u_h|_{H^1(\Omega)}$, {\smaller \smaller \smaller \textcolor{green}{$\triangle$}} is for $|u_h - \widetilde u_h|_{H^1(\Omega)}$ for Example 1 (top row), 2 (middle row), and 3 (bottom row) using linear (left column), quadratic (middle column), and cubic (right column) CGFEM. }
\caption{\label{Fig1}(a) Time evolution of $\msd$ for various values of $\eps$, where $N=32\times32$. Here $K=10$, $\eta=10^{-6}$ and \red{$T=1$ throughout this paper.} (b) Sharp transition observed by the averaged \lifetime as a function of $\eps$ for $N=16\times16,~32\times32$, and $64\times64$, where $\hbar=2\pi/N_1=2\pi/\sqrt{N}$. Note that the averaged \lifetime above the critical $\eps$ is much larger than the order of magnitude of $N$. }
\caption{The Roche-lobe radius of the donor star in \source\as a function of its mass,$M_2$ (black solid line), compared to a selection of the measured masses and radii of stars on the MS (blue points). The dashed red line shows the relation of $R_2=\rsun (M_2/\msun)^{0.9}$ for $M_2\geq 0.1\msun$, which approximates the middle of the MS, and $R_2=0.1\rsun$ for $M_2<0.1\msun$, which approximately describes the case of old brown dwarfs. \blue{The green and cyan curves show the models of stars on the ZAMS and TAMS, respectively. The intersections of these curves with the Roche-lobe radius show that MS stars in the mass range of $\simeq (0.52$--$0.84)\msun$ are acceptable models of the donor star.} }
\caption{\blue{The evolution of partially stripped giants in the $M_{\rm c}$--$R_2$ diagram. The total star masses are 0.15, 0.16, 0.175, 0.2, 0.3, 0.5 and $0.9\msun$ for the solid black, blue, green and red curves, and dashed cyan, magenta and black curves, respectively. The evolution proceeds (from left to right) at the constant total mass, during which the H-burning shell is moving outwards. This increases the mass of the He core and decreases the mass of the H-rich envelope. The black and red dotted lines show the radius of the Roche-lobe around the donor (and so, to a good approximation, also the radius of this star) for $M_2= 0.15\msun$ and $M_2= 0.2\msun$, respectively. Possible solutions for the donor in \source\are given by the intersections between the dotted and the solid lines on the left-hand side of the plot.}}
\caption{\color[HTML]{0000FF}{\label{cat} The quantum network model reveals the information flow in the thought experiment. Single lines represent quantum systems (qubits) and double lines classical systems (bits); blue (red) lines denote the atom (cat). Quantum (classical) information flows along the single (double) lines. (a) Quantum network entangling the atom and the cat. A quantum cat in the initial state $\ket{alive}$ is coupled to the atom via a quantum CNOT gate resulting in the state $a\ket{0} \ket{alive}+ b \ket{1} \ket{dead}$. (b) Quantum network for the standard setup of Schr\"odinger's cat. The atom evolves freely to$a\ket{0}+ b\ket{1}$ and is coupled to the cat via a classical information channel: if the detector clicks ($m=1$), a poison is released and kills the cat.}}
\caption{% Predictive accuracy of original estimator (\green{green}) and M-estimator (\blue{blue}) over \# of training examples, with standard error bars taken over 100 simulations. Top:$m=5$ hidden states with $n=10$ hidden states. Bottom: $m=10$ hidden states with $n=20$ observed states. }
\caption{Comparison of prediction performance of proposed \tool~and alternatives (best viewed in color). The proposed \tool~outperforms the compared alternatives in terms of all the evaluation metrics used (\textcolor{\bhlcolor}{F-measure}, specificity, sensitivity, accuracy, negative predictive value, and positive predictive value; see the footnote on \textcolor{\bhlcolor}{page 5} for the definitions of these metrics \textcolor{\bhlcolor}{and Table~\ref{tbl:results-accuracy-tool} for the values of each metric})}
\caption{The dynamical spin susceptibility and spin-fluctuation linewidths for the strange metal phase of cuprates at various ${\bf q}$ in the Brillouin zone \textcolor{orange}{(0,0)}\textcolor{red}{($\pi/2,0$)} \textcolor{purple}{($\pi,0$)}\textcolor{green}{($\pi,\pi/2$)}\textcolor{black}{($\pi,\pi$)}\textcolor{blue}{($\pi/2,\pi/2$)} of the reduced Brillouin zone at $\beta$ = 100 (left panel) and 1000 (right panel).}
\caption{Renormalized Phonon lineshapes and linewidths as functions of various ${\bf q}$ in the Brillouin zone. \textcolor{orange}{(0,0)}\textcolor{red}{($\pi/2,0$)}\textcolor{purple}{($\pi,0$)}\textcolor{green}{($\pi,\pi/2$)}\textcolor{black}{($\pi,\pi$)}\textcolor{blue}{($\pi/2,\pi/2$)} of the reduced Brillouin zone at $\beta$ = 100 (left panel) and 1000 (right panel). Asymmetry in the lineshapes as well as a ``Fano''-like peak at low energy is also clearly visible.}
\caption{The dynamical spin susceptibility and spin-fluctuation linewidths for the underdoped cuprates at various ${\bf q}$ in the Brillouin zone \textcolor{orange}{(0,0)}\textcolor{red}{($\pi/2,0$)}\textcolor{purple}{($\pi,0$)}\textcolor{green}{($\pi,\pi/2$)}\textcolor{black}{($\pi,\pi$)}\textcolor{blue}{($\pi/2,\pi/2$)} of the reduced Brillouin zone at $\beta$ = 100 (left panel) and 1000 (right panel). Phenomenological introduction of a nematic-plus-$d$-wave modulation electronic order suppresses the infra-red singularity in spin-fluctuations in a momentum-dependent way.}
\caption{Renormalized Phonon lineshapes and linewidths at different ${\bf q}$ points \textcolor{orange}{(0,0)}\textcolor{red}{($\pi/2,0$)}\textcolor{purple}{($\pi,0$)}\textcolor{green}{($\pi,\pi/2$)}\textcolor{black}{($\pi,\pi$)}\textcolor{blue}{($\pi/2,\pi/2$)} of the reduced Brillouin zone at $\beta$ = 100 (left panel) and 1000 (right panel).}
\caption{Images that were assigned an incorrect sense in the \pred setting.}
\caption{An example narrative for a EMBERS alert message. Here, color {\color{red}red} indicates named entities, {\color{green} green} refers to descriptive protest related keywords. Items in {\color{blue} blue} are historical or real time statistics and those in {\color{magenta}magenta} refer to inferred reasons of protest. }
\caption{Stress field at the contact region $\sigma$ within the primitive cell of the system delimited by the discontinuous lines [\textcolor{red}{- $\cdot$ -}], obtained for $E_p=E_s=1.6\times10^{6}$ Pa, $\zeta=0.3$ $\mu$m, different values of $\Phi$ -- \subref{Fig:Sig2} $\Phi=0.03$ and \subref{Fig:Sig4} $\Phi=0.82$ -- and the remaining parameter values given in Table~(\ref{Tab:Param}). The dotted curves [\textcolor{red}{$\cdots$}] delimitate the computed contact region for each case. Figures on the left-hand side show a 2D representation in the plane $xy$, whereas figures on the right-hand side show a 1D representation in the axis $r=\sqrt{x^2+y^2}$. In the $\sigma$ \textit{vs} $r$ representation, curves indicating the Hertz and soft-flat-punch stress fields, given respectively by eqs.\eqref{Eq:SFHertz} and \eqref{Eq:SFSFP}, and a modified Hertz-like behavior, for which the stress $\sigma_0$ at $r=0$ and the apparent contact radius $a$ were obtained from the numerical simulations, are also depicted.}
\caption{Stress at the center of the pillar $\sigma_{0S}$, for the soft-flat-punch contact, as a function of the ratio of Elastic moduli $E_l/E_p$, for $\Phi=2.3\times10^{-3}$ and the remaining values given in Table~(\ref{Tab:Param}). The behavior described by the continuous curve [\textcolor{red}{---}] was computed with eq.\eqref{Eq:S0BvsE}.}
\caption{Multi-tier drone-cell networks can be used for many scenarios: \mytikzdot{1} Providing service to rural areas (macro-drone-cell), \mytikzdot{2} Deputizing for a malfunctioning BS (macro-drone-cell), \mytikzdot{3} Serving users with high mobility (femto-drone-cell), \mytikzdot{4} Assisting a macrocell in case of RAN congestion (pico-drone-cell), \mytikzdot{5} Assisting a macrocell in case of core network congestion or malfunctioning (macro-drone-cell), \mytikzdot{6} Providing additional resources for temporary events, e.g., concerts and sports events, \mytikzdot{7} Providing coverage for temporary blind spots, and \mytikzdot{8} Reducing energy dissipation of sensor networks by moving towards them (femto-drone-cell).}
\caption{DMF mechanism and potential business and information model demonstrating partitioning of the traditional MNO into InP (cloud, server, drone-BS etc.) and MVNO: \mytikzdot{1} Collect and store global data; \mytikzdot{2} Process data for network monitoring and creating intelligence; \mytikzdot{3} Provide guidance for drone-cell's operation (placement, content to be loaded, access technology, service duration, coverage area, moving patterns); \mytikzdot{4} Re-configure the virtual network of MVNO for drone-cell integration by SDN and NFV technologies, e.g., introduce another gateway to handle busy traffic and create new paths among the new and existing functions; \mytikzdot{5} Drone-cell assists the network; \mytikzdot{6} SP can continue delivering services successfully.}
\caption{Intrinsic rest-frame UV luminosity function (LF) at $z\in\{10,11,12\}$ predicted by the \bluetides\simulation alongside observational constraints at$z\approx 10.4$ from Bouwens et al. (2015b) and at $z\sim 9.9$ from Oech et al. (2014). The shaded regions on the curves represent the 1$\sigma$ uncertainty on the LF computed from the entire \bluetides\volume.}
\caption{The cumulative volume density of galaxies at $z=11$ as a function of absolute magnitude. The observation of GN-$z11$ is shown as a point, with a horizontal error bar denoting the uncertainty in the magnitude. The vertical error is estimated using a simple Poisson error. The right hand axes show the number of galaxies in both the \bluetides\volume and the volume probed by the O16 analysis.}
\caption{The average predicted SED of bright ($M_{\rm UV}\approx -22$) galaxies at $z=11.1$ in \bluetides\compared with the observed fluxes of GN-$z11$. The two SEDs shown denote both the pure stellar case ($f_{\rm esc, LyC}=1$) and the case in which the Lyman continuum escape fraction is effectively zero. In the latter case the Lyman-$\alpha$ line has also been damped.}
\caption{Stellar mass, star formation rate and stellar ages versus UV luminosity for the galaxies at $z=11$ in \bluetides. The large black data point denotes the inferred constraints on GN-$z11$ from O16. The green line shows the mean value for \bluetides\in UV magnitude bins. The 2D histogram shows the distribution of galaxies in\bluetides.}
\caption{The stellar surface density color coded by stellar age (blue to red) for a sample of $M_{\rm UV} <-22$ galaxies selected from the \bluetides\simulation at$z=11$. We show three galaxies most closely matching the magnitude of GN-$z11$ and the two brightest galaxies in \bluetides\at$z=11$.}
\caption{Mean stellar and star forming gas metallicity versus UV luminosity for galaxies at $z=11$ in \bluetides. The green line shows the mean value for \bluetides\in UV magnitude bins. The 2D histogram shows the distribution of galaxies in\bluetides. The shaded bands show the luminosity of GN-$z11$ and its observational uncertainty.}
\caption{Top panel: Prediction for the magnitude difference between the AGN and its host galaxy for galaxies and black holes in \bluetides\at$z=11$. Bottom panel: the predicted black hole masses as a function of host galaxy magnitude. Black points denote the mean and standard deviation in UV magnitude bins. The gray area shows the region corresponding to the magnitude of GN-$z11$.}
\caption{Plots of the fraction of unfixed galaxies assigned correctly and the time required for the optimization to finish as a function of number of bins for a series of optimization runs with 2000 galaxies in $V$, 80\% of which were fixed, and a binning scheme of 100 pairs per bin. {\color{blue}{In the above plots, the circles}} correspond to optimization runs that completed in under $10^5$ seconds. The {\color{blue}{x's}} correspond to optimization runs that had not completed after $10^5$ seconds and were terminated after this amount of time; the reported fraction of unfixed galaxies assigned correctly was calculated using the optimal solution found before termination. }
\caption{{\bf Barnes-Hut-SNE visualization of source word representations} -- shown are sample words from the {\it Rare Word} dataset. We differentiate two types of embeddings: {\color{blue} frequent} words in which encoder embeddings are looked up directly and {\it{\color{magenta} rare}} words where we build representations from characters. Boxes highlight examples that we will discuss in the text. We use the hybrid model \model{} in this visualization.}
\caption{The ionized gas line ratios of the ELG (grayscale) and SPOG samples (green dots \citealt{a16_sample}), including \nii/\ha\vs.\oiii/\hb\(left; \citealt{bpt}), \sii/\ha\vs.\oiii/\hb\(middle), and \oi/\ha\vs.\oiii/\hb\(right; \citealt{veilleux+87}), overlaid with the line diagnostic models of \citet{kauffmann+03,kewley+06}. The purple line defines the boundaries of the shock models, SPOG criterion \citep{allen+08,rich+11,a16_sample}. {\em WISE} 22$\mu$m-detected SPOGs are shown in dark green. CO-SPOGs (light blue stars), and 1.4\,GHz radio-matched CO-SPOGs (dark blue stars) are also shown. CO-SPOGs span the ionized gas diagnostic space of the larger SPOG sample. There is also little difference between the radio matched and unmatched 22$\mu$m-detected SPOGs.}
\caption{SDSS {\em g\,r\,i} 3-color thumbnails of the 52 CO-SPOGs. The field of view of all thumbnails is 30$''$. Objects detected by CARMA are labeled in green, with the CARMA beam shown as an ellipse in the bottom right. Objects observed with the IRAM~30m are labeled in pale blue. The SPOG index from \citet{a16_sample} are labeled in white (bottom left). Non-detections are demarcated with a small \textcolor{Red}{\bf X} in the top right. Objects that have been morphologically classified as clearly disrupted have a small yellow symbol (\textcolor{Dandelion}{$\mathbf{\infty}$}) below the object name. Those that are classified as possibly disrupted, a (\textcolor{Dandelion}{$\mathbf{\sim}$}). Many SPOGs show signs of interaction, including tidal tails, dust lanes, and morphological disruption. Most galaxies appear to be red in color, with many also showing signs of peaked, bright nuclei, possibly due to the presence of an AGN.}
\caption{Continuation of the SDSS {\em g\,r\,i} 3-color images of CO-SPOGs. The last panel shows the 22.3$''$ single dish primary beam at $\nu_{\rm obs}$\,=\,110\,GHz for the SPOGs observed with the IRAM~30m (labeled in pale blue).}
\caption{{\bf (Left):} The 22$\mu$m and CO(1--0) fluxes are compared for many samples, including star-forming galaxies such as COLD GASS (black dots; \citealt{coldgass}), EGNoG (gray dots; \citealt{bauermeister+13}) and AMIGA (dark gray squares; \citealt{lisenfeld+11}), early-type galaxies (yellow diamonds; ATLAS$^{\rm 3D}$: \citealt{young+11,alatalo+13,davis+14}), AGNs (red triangles; \citealt{evans+05}), poststarburst galaxies (green crosses; \citealt{french+15}), NGC\,1266 (outlined yellow star;\citealt{alatalo+11,a14_stelpop,a15_sfsupp}), and CO-SPOGs (blue stars). Light blue dots inside of the stars denote CO-SPOGs that have detections with S/N between 3--5. The 22$\mu$m and CO(1--0) fluxes of star-forming galaxies are well correlated (shown as a black line), with AGNs showing the strongest divergence from the relation. CO-SPOGs sit below the relation as well, though not as extremely as the AGNs. {\bf (Right):} The 22$\mu$m vs. CO(1--0) fluxes are shown, emphasizing the SPOGs with particular properties (symbols for other samples from lefthand plot have been changed to grayscale). This includes those galaxies with 3$\sigma$ excess \nad\absorption (a; top right, red), galaxies that were detected with FIRST (b; middle right, green), and galaxies that were classified as clearly or possibly morphologically disrupted (c; bottom right, pink). Although radio-detected CO-SPOGs account for most of the objects that fall the farthest from the$F_{22}$--$F_{\rm CO}$ relation, they do not universally exhibit quasar-like 22$\mu$m excess relative to their CO fluxes.}
\caption{The equivalent widths EW(Mg\,b) vs. EW(\nad) of the ELG sample (grayscale; \citealt{a16_sample}) compared with the entire SPOGS sample (green contours). The dashed line represents the empirical relation found in \citet{a16_sample}. The dotted line represents a 3$\sigma$ departure from the empirical relation. The CO-SPOG sample is shown (stars), including FIRST radio detections (dark blue) and non-detections (light blue), as well as NGC\,1266 (gold;\citealt{davis+12}). CO-SPOGs with a light blue dot in their centers are detected with S/N between 3--5. SPOGs have \nad\compared to Mg\,b that is enhanced beyond what is seen in normal star-forming galaxies in the ELG sample. The CO(1--0) detected objects show even more enhanced\nad\characteristics compared to non-CO observed SPOGs, with 19 (37\%) objects beyond the 3$\sigma$ boundary defined by the normal (ELG-defined) relation. There does not appear to be a difference between the objects observed by CARMA and those observed by the IRAM~30m. Clearly and possibly morphologically disrupted SPOGs are outlined in pink. Radio-detected SPOGs tend to have the largest \nad\excess, and radio non-detected, non-disrupted SPOGs have the least excess. Most radio non-detected SPOGs that do show\nad\excess at the$>$3$\sigma$ level also show morphological disruptions.}
\caption{A comparison between $I_{\rm CO}$ determined in both of the methods: fitting a single Gaussian vs. integrating the total intensity of the line for detected IRAM CO-SPOGs. Strong (S/N\,$>$5) detections are shown as blue circles and tentative (S/N\,=\,3--5) are shown as red triangles. The average disagreement between the two methods is 5\%, and the vast majority of objects agree within errors.}
\caption{ The green box demonstrates how a pedestrian forms beliefs about others $\mu_{\neg n}^{(t)}$ and encode such information into social compliance feature $f_{n,soc}^{(t)}(\vec{x})$. {\color{red}Red} indicates high reward and {\color{blue}blue} indicates low reward.}
\caption{(Top) Features invariant of time. Neighborhood occupancy feature $f_{occ}(\vec{x})$ used to encode how close a pedestrian will walk near static objects. Distance-to-goal feature $f_{dog}(\vec{x})$ describes pedestrians' desire to reach destination. Body orientation feature $f_{bod}(\vec{x})$ captures pedestrian tendencies to walk aligned with body direction. (Bottom) The green box demonstrates how a pedestrian forms beliefs about others $\mu_{\neg n}^{(t)}$ and encodes such information into the social compliance feature $f_{soc}^{(t)}(\vec{x})$. {\color{red}Red} indicates high reward and {\color{blue}blue} indicates low reward.}
\caption{\red{(a) Radius of gyration $R_g$, (c) end-to-end distance $R_e$, (d) hydrodynamic radius $R_h$, (e) the ratio $R_e^2/R_g^2$, (f) the ratio $R_h/R_g$, (g) the asphericity $b$, and (h) the acylindricity $c$, as a function of the B-monomer concentration. The radius of gyration for multivalent salts is shown in (b) for comparison. } The A-monomer concentration is $10^{-4}\, \sigma^{-3}$.}
\caption{ {\color{red}Left panel:} shows the intrinsic black-body spectra of individual Pop-III stars per unit stellar mass, multiplied by the main sequence lifetime, for a set of stellar masses (in units of solar mass) indicated in the legend. Also shown as the vertical magenta dashed line is the photon energy of $2$eV for the photo-detachment process. The LW band is indicated by the two black vertical dashed lines. {\color{red}Right panel:} shows the radiation intensity at $2$eV (magenta solid curves) and $11.2$eV (black solid curves) for a single star of mass shown on the x-axis. The star is assumed to be located at the center and the intensity is measured at the core radius of the minihalo (see text for definition of core radius). Two cases are shown, one for a minihalo at $z=25$ with virial temperature of $10^3$~K (upper solid curves) and the other at $z=7$ with virial temperature of $10^4$~K (lower solid curves). For both IR and LW photons, no absorption is assumed for this illustration. The horizontal dashed lines with the same corresponding colors are the threshold intensity for complete suppression of \h2 formation by the respective processes. }
\caption{ {\color{red}Left panel:} shows the intrinsic black-body spectra of individual Pop-III stars per unit stellar mass, multiplied by the main sequence lifetime, for a set of stellar masses (in units of solar mass) indicated in the legend. Also shown as the vertical magenta dashed line is the photon energy of $2$eV for the photo-detachment process. The LW band is indicated by the two black vertical dashed lines. {\color{red}Right panel:} shows the radiation intensity at $2$eV (magenta solid curves) and $11.2$eV (black solid curves) for a single star of mass shown on the x-axis. The star is assumed to be located at the center and the intensity is measured at the core radius of the minihalo (see text for definition of core radius). Two cases are shown, one for a minihalo at $z=25$ with virial temperature of $10^3$~K (upper solid curves) and the other at $z=7$ with virial temperature of $10^4$~K (lower solid curves). For both IR and LW photons, no absorption is assumed for this illustration. The horizontal dashed lines with the same corresponding colors are the threshold intensity for complete suppression of \h2 formation by the respective processes. }
\caption{ {\color{red}Left panel:} shows the critical cumulative stellar mass for complete suppression of \h2 formation, as a function of the lower mass cutoff of the IMF ${\rm M_{low}}$, via either the photo-dissociation process by LW photons with (blue open squares) and without \h2 self-shielding of LW photons (blue solid squares) or the photo-detachment process by infrared photons (red solid dots). In this example, we assume that a minihalo of virial temperature $T_v=10^3$K is formed at $z=25$ when star formation commences, and the critical stellar mass (i.e., upper limit on total stellar mass) is evaluated at $z=7$ when the minihalo has grown to a virial temperature of $T_v=10^4$K. {\color{red}Right panel:} shows the upper bound on the mean gas metallicity, corresponding to the critical stellar mass shown in the left panel, evaluated at $z=7$ when the minihalo has grown to a virial temperature of $T_v=10^4$K. In both panels, the errorbars indicate the dispersions obtained by Monte Carlo realizations of different star formation histories, described in the text. }
\caption{ {\color{red}Left panel:} shows the critical cumulative stellar mass for complete suppression of \h2 formation, as a function of the lower mass cutoff of the IMF ${\rm M_{low}}$, via either the photo-dissociation process by LW photons with (blue open squares) and without \h2 self-shielding of LW photons (blue solid squares) or the photo-detachment process by infrared photons (red solid dots). In this example, we assume that a minihalo of virial temperature $T_v=10^3$K is formed at $z=25$ when star formation commences, and the critical stellar mass (i.e., upper limit on total stellar mass) is evaluated at $z=7$ when the minihalo has grown to a virial temperature of $T_v=10^4$K. {\color{red}Right panel:} shows the upper bound on the mean gas metallicity, corresponding to the critical stellar mass shown in the left panel, evaluated at $z=7$ when the minihalo has grown to a virial temperature of $T_v=10^4$K. In both panels, the errorbars indicate the dispersions obtained by Monte Carlo realizations of different star formation histories, described in the text. }
\caption{{\it The space of possible paths, and the payoff of the American claim}. The labels at the nodes on the graph consist of a quadruple, the elements of which are time, price level, {\cblue node probability} and {\cred payoff of the American option} respectively. Here node probabilities are equivalent to the prices of Arrow-Debreu securities. {\cgreen Also labelled on the figure by $\frac{1}{2}, \frac{1}{2}$ and $p,q,r,s,t,u$ are conditional probabilities of transitions between nodes.} }
\caption{A comparison of six models on four data sets. Model performance on the test set is quantified by AUC, a measure of how well the model discriminates (predicts) correct and incorrect student responses. The models are trained on one set of students and tested on another set. Note that the AUC scale is different for each graph, but tic marks are always spaced by .03 units in AUC. On \assistments\and\synthetic, DKT results are from Piech et al.\\cite{Piechetal2015}; on \statics\and\spanish\, DKT results are from our own implementation.\color{figr}BKT\color{black} = classic Bayesian knowledge tracing; \color{figy}BKT+A\color{black} = BKT with inference of latent student abilities; \color{figg}BKT+F\color{black} = BKT with forgetting; \color{figc}BKT+S\color{black} = BKT with skill discovery; \color{figb}BKT+FSA\color{black} = BKT with all three extensions; \color{figm}DKT\color{black} = deep knowledge tracing \label{fig:results} }
\caption{Performance comparison with the state of the art. Results reported with the use of AlexNet or VGG are marked by (\ebA) or (\ebV), respectively. Use of fine-tuned network is marked by (\ebf), otherwise the off-the-shelf network is implied. \mbox{D:~Dimensionality}. Our methods are marked with \our and they are always accompanied by \cpl2. New state of the art highlighted in \nb{red}, surpassed state of the art in \ob{bold}, state of the art that retained the title in \protect\sb{outline}, and our methods that outperform previous state of the art on a \fcolorbox{gray!15}{gray!15}{gray background}. Best viewed in color. \label{tab:stateofart}}
\caption{Identified open challenges in previous work on human action recognition for streaming data. To the best of our knowledge, FIVER is the only algorithm equipped with all these features, crucial for dealing with streaming data. \emph{LEGEND}: \surdgreen: exhibits the feature; \approxblu: partially exhibits the feature; \crossred: does not possess the feature.}
\caption{Ensemble averaged Macroscopic Von Mises equivalent stress $\langle \bar{\sigma}_\mathrm{eq} \rangle$ (\markeri) versus the macroscopic applied equivalent strain $\bar{\varepsilon}$, respectively normalized by the initial yield stress and strain, $\sigma_\mathrm{y0}^\mathrm{soft}$ and $\varepsilon_0^\mathrm{soft}$, of the soft matrix. The homogeneous stress of the soft phase (\markerj) and the hard phase (\markerk) act as a lower and upper bound respectively.}
\caption{Simulation results of three randomly generated cells. From left to right, the maximum damage in the cell is the (a) lowest, (b) intermediate, and (c) highest among all cells. From top to bottom: the microstructure, equivalent plastic strain $\varepsilon_\mathrm{p}$, hydrostatic stress $\sigma_\mathrm{m}$, and damage $D$ at $\lambda = 1.2$. The hard phase (i.e.\inclusions) is highlighted by a black outline.}
\caption{Comparison of the onset of ductile fracture for different combinations of hard phase volume fraction $\varphi^\mathrm{hard}$ and hardness $\chi^\mathrm{hard}$, resulting in the same macroscopic hardening behavior. Different volume fractions are denoted by different markers. The onset of ductile fracture, i.e.\point where each curve has been truncated, is highlighted using a bigger marker.}
\caption{\label{fig:reconstructions} Semantic 3D Reconstructions: we improve in \blue{weakly observed areas} and resolve \yellow{unary potential artifacts} at the same time. Five semantic labels are used: ground, building, vegetation, clutter, free space. \vspace*{-0.15cm}}
\caption{a) Schematic of the experimental system. b) The primary results of this work summarized in a phase diagram as a function of the smectic alignment parameter $C$ and the twist-composition coupling parameter $a$. The lines indicate the phase boundary between the microphase separated and bulk phase separated states as obtained from linear stability analysis (black/solid) and numerical integration of Eqs. \ref{dyn1}-\ref{dyn2} (green/dashed). The snapshots show configurations at steady state obtained from numerics at the indicated parameter values ({\color{red} o}). c) Illustration of the evolution to steady state for two parameter sets. The results shown here and in the rest of the paper are for a 60-40 mixture with $\lambda_\psi=0.1$, and $q_{0}=0.1$}
\caption{The maximum separation ratio from Equation~\eqref{eq:6} attained by the reservoirs on sample training sets, adapted for Algorithm~\ref{alg:Wout t} in blue `{\color{blue}$*$}' and for Algorithm~\ref{alg:cluster} in red `{\color{red}$\circ$}', for several values of $\alpha$ and $N$. Results for the ESN-style reservoirs are in the left plot, and results for TDR-style reservoirs are in the right plot.}
\caption{A comparison of the performance of the proposed method using Algorithm~\ref{alg:cluster} and the method of trained linear output weights using Algorithm~\ref{alg:Wout t} on ESN and TDR reservoirs for various parameters $\alpha$ and reservoir size $N$. The top two figures present the classification accuracy on the test set, and the bottom two figures present the total time required in seconds to classify the entire test set of 7000 images. The plots with red '{\color{red}$\circ$}' denote results from Algorithm~\ref{alg:cluster}, and the plots with blue `{\color{blue}$*$}' or cyan `{\color{cyan}$\square$}' denote results from Algorithm~\ref{alg:Wout t} for different regularization parameters.}
\caption[caption]{Architecture of the proposed training algorithm. Given an image, and selective search RoIs, the conv network computes a conv feature map. In (a), the \protect\emph{readonly} RoI network runs a forward pass on the feature map and all RoIs (shown in {\protect\color{green}green} arrows). Then the Hard RoI module uses these RoI losses to select $B$ examples. In (b), these hard examples are used by the RoI network to compute forward and backward passes (shown in {\protect\color{red}red} arrows).}
\caption{(a) The shape of the capillary focusing for Ca $=0.01$ with no pressure boundary condition correction at the outflow, $A=1$ (red), and for Ca $=0.004$ with a pressure boundary condition correction at the outflow, $A=0.4$ (blue); $w_1(L)\approx b$. (b) Computed interface curvature normalized by $2/b$, the magnitude of the curvature at the outflow ($x=L$), as a function of $x/w$ for Ca $=0.01$ and $A=1$ ({\color{red}$\blacktriangledown$}), and for Ca $=0.004$ and $A=0.4$ ({\color{blue}$\blacklozenge$}); the dashed line, $\kappa=0$, is plotted for visual assistance for when the curvature changes sign. $w_{1}(0)/b=10$.}
\caption{ Back-gate dependence of the WAL effect of an in-situ Al$_2$O$_3$-capped Bi$_2$Te$_3$ thin film. (a) The gate-voltage dependence of the low field magnetoconductance $\Delta G_\text{xx}(B)$. All data were taken at 100 mK. The inset highlights the observation of a parabolic background contribution to $\Delta G_\text{xx}$ at high positive $V_{bg}$. (b) Gate dependence of the prefactor $\alpha$ and the dephasing length $l_{\varphi}$, obtained from best fits to equation \textcolor{blue}{(\ref{HLN})}. (c) A schematic band diagram illustration relevant to the thin films used in this study. BCB is the bulk conduction band, BVB is the bulk valence band, and $E_{f}$ is the position of the Fermi level. We access the transport regimes I, II and III of (b) as the Fermi level moves through the respective shaded regions of the band structure in (c) and Fig.~\ref{fig:TransportCharacteristics}\textcolor{blue}{(e)} when the gate voltage is lowered.}
\caption{Transport characteristics of a Bi$_2$Te$_3$ film capped in-situ with a $6$ nm Te layer. (a) Back-gate voltage ($V_{bg}$) dependence of the 2D carrier density extracted from the zero-field slope of $R_\text{xy}(B)$ and of the sheet resistance. (b) Back-gate dependence of the WAL effect of an in-situ Te-capped Bi$_2$Te$_3$ film. The prefactor $\alpha$ and the dephasing length $l_{\varphi}$ as function of $V_{bg}$, extracted from best fits to equation \textcolor{blue}{(\ref{HLN})} as illustrated in the inset.}
\caption{LDL concentration profiles across the arterial wall obtained for transmural pressure equal to $70 \, \mathrm{mmHg}$: (a) stationary state and (b) after 30 minutes of incubation. '\textit{Fluid concentration}'(\textcolor{Blue}{$\boldsymbol{-}$}) is the normalized LDL concentration in the plasma, obtained directly from calculations. '\textit{Tissue concentration}' (\textcolor{Blue}{\textbf{\texttt{-{}-}}}) is the normalized LDL concentration profile in the wet tissue. '\textit{Averaged tissue concentration}' ($\filledcirc{blue}$) is the tissue concentration averaged in the $20\mbox{-}\mu \mbox{m-}$slices as in experiments. Results after 30 minutes (b) are put together with experimental results of Meyer et al. \cite{meyer} ({\color{red}$\boldsymbol{\times}$}). Subsequent layers are marked with colors: magenta for endothelium, green for intima, blue for IEL and white for media. }
\caption{ LDL concentration profiles across the arterial wall obtained for transmural pressure equal to $160 \, \mathrm{mmHg}$: (a) stationary state, (b) after 30 minutes of incubation, (c) after 5 minutes of incubation. Fluid concentration with fraction of leaky junction dependent on pressure and time is marked by (\textcolor{OliveGreen}{$\boldsymbol{-}$}), averaged tissue concentration with $\phi$ dependent on pressure and time is marked by ($\filledcirc{black!50!green}$), fluid concentration with fraction of leaky junction independent on pressure and time is marked by (\textcolor{Blue}{$\boldsymbol{-}$}) and averaged tissue concentration with $\phi$ independent on pressure and time is marked by ($\textcolor{Blue}{\blacktriangle}$). Results are put together with experimental results of Meyer et al. \cite{meyer} ({\color{red}$\boldsymbol{\times}$}) and Curmi et al \cite{curmi1990effect} ({\color{blue}$\boldsymbol{\times}$}). Subsequent layers are marked with colors: magenta for endothelium, green for intima, blue for IEL and white for media. Time dependence of fraction of leaky junctions $\phi$ ($\filledcirc{red}$) is illustrated by the red curve (\textcolor{red}{---}) in Panel (d).}
\caption{Summary of LDL concentration profiles across the arterial wall obtained for transmural pressure equal to $120 \, \mathrm{mmHg}$: (a) stationary state and (b) after 30 minutes of incubation. The fluid concentration and averaged tissue concentration obtained with pressure dependent fraction of leaky junctions (\textcolor{OliveGreen}{---} $\filledcirc{black!50!green}$) and without pressure dependent fraction of leaky junctions ({\color{blue}--- $\blacktriangle$}) are put together with the experimental results of Meyer et al. \cite{meyer} ({\color{red}$\boldsymbol{\times}$}). The subsequent layers are marked with colors: magenta for endothelium, green for intima, blue for IEL and white for media.}
\caption{Fluid concentration profiles across the arterial wall in the stationary state under different pressure conditions: $\Delta P = 70 \, \mathrm{mmHg}$ (\textcolor{red}{$\boldsymbol{-}$}), $\Delta P = 120 \, \mathrm{mmHg}$ (\textcolor{OliveGreen}{$\boldsymbol{-}$}), $\Delta P = 160 \, \mathrm{mmHg}$ (\textcolor{Blue}{$\boldsymbol{-}$}): (a) fraction of leaky junction $\phi$ dependent on the $\Delta P$ and (b) fraction of leaky junction $\phi$ independent on the $\Delta P$.}
\caption{Profiles of normalized LDL concentration across the arterial wall obtained with included intima compression (averaged tissue concentration ($\filledcirc{black!50!green}$), fluid concentration (\textcolor{OliveGreen}{ $\boldsymbol{-}$}) and tissue concentration (\textcolor{OliveGreen}{ \textbf{\texttt{-{}-}}})), and without intima compression (averaged tissue concentration ($\filledcirc{blue}$), fluid concentration ({\color{blue} $\boldsymbol{-}$}) and tissue concentration (\textcolor{blue}{ \textbf{\texttt{-{}-}}})), compared with experimental results of Curmi et al. \cite{curmi1990effect} ({\color{red}$\boldsymbol{\times}$}). }
\caption{ (a) Test (\protect\tikz[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \protect\draw (0,0) -- (1em,0);) and training loss (\protect\tikz[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \protect\draw[dotted,thick] (0,0) -- (1em,0);) loss for approximation of a multivariate polynomial for different net structures. % schroeder/non_scaled/00024% (b) Polynomial data (grayscale coded) to be learned ranging with training points in red. (c) Relative error for networks with $\expn$ and $\tanh$ as transfer functions.}
\caption{\label{vi} \color{mycol} a) Schematics of HS+VI+TC/GVL1 cable inside the cryostat. Arrows represent thermalizations. A room temperature cable HS (gray) was used to carry signal from vacuum feedthrough to vacuum insulated cable VI (black), since the geometry of the cryostat required flexible cables. Cable VI was used between room temperature and the 2nd pulse tube stage (PT2). Its outer conductor was thermalized to respective flanges of the cryostat. We used TC or GVL1 (red) between the PT2 and mixing chamber (MC) flange. b) Shematics of the VI cables (not to scale). Connector 1 had an opening through which the inner conductor could be tightened. The conductor was later cut to fit inside the connector and the opening was shielded with a hat. The connectors were connected to the CuNi tube by first crimping and then soldering.}
\caption{Average success rates (SR) in percentages. The best and second best results are presented in {\color{red}red} and {\color{blue}blue} fonts respectively.}
\caption{Average COL errors in pixels. The best and second best results are presented in {\color{red}red} and {\color{blue}blue} fonts respectively.}
\caption{Average SR and COL errors for hand-crafted representation trackers and DST. The best results are presented in {\color{red}red} font.}
\caption{A time-space diagram that shows traffic flow speed on a 13-mile stretch of I-55. Cross-section speed measured on from 18 February 2009 (Wednesday). \textcolor{red}{Red} means slow speed and light \textcolor{yellow}{yellow} corresponds to free flow speed. The direction of the flow is from 0 to 13.}
\caption{Comparison of the residuals forecasts over time. On all plots \textbf{black} solid line is the measured data (cross-section speed), {\color{red} \textbf{red}} dots are absolute values of residuals from our model's forty minute forecast. First column compares models for data from Thursday October 10, 2013, the day when Chicago Bears team played New York Giants. The game starts at 7pm and lead to an unusual congestion starting at around 4pm. Second column compares models for data from Wednesday December 11, 2013, the day of light snow. The snow leads to heavier congestion during both, the morning and evening rush hours. Third column compares models for data from Monday October 7, 2013. There were no special events, accidents or inclined weather conditions on this day.}
\caption{Comparison of the forecasts. On all plots \textbf{black} solid line is the measured data (cross-section speed), {\color{red} \textbf{red}} dashed line is our model's forty minute forecast and dashed {\color{blue} \textbf{blue}} line is naive forecast. Green dashed horizontal line is the speed limit (55 mi/h) and vertical orange line is the morning peak hour (8am). First column compares models for data from Thursday October 10, 2013, the day when Chicago Bears team played New York Giants. The game starts at 7pm and lead to an unusual congestion starting at around 4pm. Second column compares models for data from Wednesday December 11, 2013, the day of light snow. The snow leads to heavier congestion during both, the morning and evening rush hours. Third column compares models for data from Monday October 7, 2013. There were no special events, accidents or inclined weather conditions on this day.}
\caption{\label{gofr_force} (a) Normalized force in the $y$-direction versus particle separation for two different system sizes ($R/\sigma$ = 1.4, $\phi = 0.855$. $N=1000$ {\color{black}$\square$}, $N=10,000$ {\color{red}$\fullmoon$}). Each data point represents the average of 800 systems. (b) Data with $N=10,000$ compared with the prediction from Eq.\\ref{pmf}. }
\caption{\label{sizerat_force} Normalized force along the $y$-direction versus particle separation for different size ratios: ($\blacksquare$) $R/\sigma$ = 1.6 , ({\color{red}$\newmoon$}) $R/\sigma$ = 4, ({\color{blue}$\blacktriangle$}) $R/\sigma$ = 8. The packing fraction, $\phi = 0.89$, and the number of particles, $N=1000$, are constant. Each data point represents the average of 800 systems.}
\caption{\label{back_and_forth} Mean force in the $y$-direction between two pinned particle upon repeated decompressions and compressions for $D/R=1$ and $R/\sigma=21$. Each data point represents the average of 200 systems with $N$ = 10,000 particles each. Starting from $\Delta\phi=0.15$ ({\color{black}$\square$}), the systems are decompressed towards the jamming transition, then re-compressed. The final state of the system is lowest curve ({\color{Purple}$\bigstar$}). }
\caption{\label{other_comp} (a) Comparison of the normalized force versus $\Delta\phi$ for both Harmonic ($\blacksquare$) and Hertzian ($\fullmoon$) interactions with $D/R=0.4$ and $R/\sigma=21$. (b) Data for two sequential sets of decompression and compression where all free particles are initially placed outside the boundary of the pinned particles.The order of the events is given by {\color{red}$\newmoon$}, {\color{red}$\fullmoon$}, {\color{blue}$\blacksquare$}, {\color{blue}$\square$}. All data points represent the average of 200 systems with $N$ = 10,000 particles each.}
\caption{\label{force_vs_phi} (a) Normalized force along the $y$-direction versus $\Delta\phi$ for different surface-to-surface distances: ($\square$) $\frac{D}{R}$ =0 , ({\color{Dandelion}$\fullmoon$}) $\frac{D}{R}$ = 0.2, ({\color{red}$\triangle$}) $\frac{D}{R}$ = 0.4, ({\color{green}$\triangledown$}) $\frac{D}{R}$ = 1, ({\color{blue}$\triangleleft$}) $\frac{D}{R}$ = 1.4, ({\color{violet}$\varhexagon$}) $\frac{D}{R}$ = 2.4, ({\color{CarnationPink} \ding{80}}) $\frac{D}{R}$ = 3. (b) Force versus the standard deviation of the force distribution for the data shown in (a). (c) The same data collapsed onto a universal functional form: $\langle F_y/(PR)\rangle=B-A\Delta\phi^{1/2}$. Each data point represents the average of 200 independent systems with $N$ = 10,000. The inset shows the fitting parameters $A$ and $B$ for each value of $D/R$. For all data, $R/\sigma$ = 21.}
\caption{\label{collapse} Normalized fitting parameters $B(L/R)^{1/2}$ (a) and $A(L/R)^{3/4}$ (b) versus $D/D^*$, as defined in Eqn.\\ref{dstar}. Data is shown for three different values of $R/\sigma$: $R/\sigma=21$ ({\color{red} $\blacktriangle$}), $R/\sigma=26$ ({\color{blue} $\blacksquare$}), and $R/\sigma=32$ ({\color{black} $\newmoon$}). Open symbols are used in (b). The insets show the same data on a logarithmic scale. The dashed lines correspond to fits to the data where $D/\sigma>3.5$ and $D/D^*<6$: $B(R/L)^{1/2}\approx 0.66 (D^*/D)^{0.50\pm0.04}$ and $A(R/L)^{3/4}\approx 1.66 (D^*/D)^{0.78\pm0.05}$. }
\caption{\label{b_cond} (a-b) Coarse-grained color map of the mean contact number per particle, $\langle z \rangle$, around an isolated pinned particle. Data is shown for $R/\sigma=32$ and $D/R=1.9$, and for two different packing fractions: (a) $\Delta\phi=10^{-5}$, and (b) $\Delta\phi=0.15$. The data is averaged over 200 independent systems with $N$ = 10,000 particles each. (c) Plot of $\langle z \rangle$, averaged around the perimeter of the pinned particle, versus the normalized distance, $s/\sigma$, from boundary of the pinned particles. Data is shown for $\Delta\phi=10^{-5}$ ({\color{blue} blue line}), and $\Delta\phi=0.15$ ({\color{red} red line}). For both low and high packing fractions, $\langle z \rangle$ is reduced near the boundary of the pinned particles. Large variations near $s/\sigma=0$ are due to structural layering of particles at the boundary.}
\caption{\label{pressure} Coarse-grained map of $\langle f_z\rangle$ for $D/R=0.4$ with $\Delta \phi=10^{-5}$ (a) and $\Delta \phi=0.15$ (b). The data is averaged over 200 independent systems with $N$ = 10,000 and $R/\sigma=32$. The colors represent mean force ({\color{orange} orange}) and number of standard deviations, $\delta$, from the mean force. }
\caption{The statistic pie chart of preceding CMEs in twin-CME scenario using different time threshold $\tau$ ahead of 126 fast main CMEs in solar cycle 23. The value of $\tau$ are shown on top of each chart. The blue section indicates ``twin CMEs'' that generate SEP events (Group I), the green section indicates the ``twin CMEs'' that fail to produce SEP events (Group III), the red section indicates ``single CMEs'' that generate SEP events (Group II), and the yellow section indicates ``single CMEs'' that fail to produce SEP events \red{(Group IV)}. }
\caption{ Plots of $r_1$, $r_2$, $r_3$, \red{and $r_4$} as defined in equations~(\ref{eq:r1}) to \red{(\ref{eq:r4})}. The figure shows the percentage of events in each of the four different groups as a function of the time interval $\tau$ between the two CMEs. The red line with crosses is $r_1$; the blue line with circles is $r_2$; the cyan line with triangle is $r_3$; \red{and the green line with squares is $r_4$}. }
\caption{The peak flux of all $59$ western SEP events from Table~\ref{table.1}. The $x$-axis is the flare class. The $y$-axis is the $>$ 10 MeV peak proton flux measured by the GOES spacecraft. Twin CME events using \red{$\tau=13$} hrs are labeled as blue crosses (group I) and single CME events are labeled as red dots (group II). }
\caption{The peak flux of all $59$ western SEP events from Table~\ref{table.1}. The $x$-axis is the CME speed. The $y$-axis is the $>$ 10 MeV peak proton flux measured by the GOES spacecraft. Twin CME events using \red{$\tau=13$} hrs are labeled as blue crosses (group I) and single CME events are labeled as red dots (group II). }
\caption{The peak flux of all $59$ western SEP events from Table~\ref{table.1}. The $x$-axis is the the 1-day prior daily average proton intensity from the Ultra Low Energy Isotope Spectrometer (ULEIS) instrument ($0.64$ to $1.28$ MeV channel) on board the Advanced Composition Explorer (ACE) spacecraft. The $y$-axis is the $>$ 10 MeV peak proton flux measured by the GOES spacecraft. Twin CME events using \red{$\tau=13$} hrs are labeled as blue crosses (group I) and single CME events are labeled as red dots (group II). The arrows indicate the median values of 1-day prior daily average proton intensity of all $59$ events and of daily average proton intensity from 1998 to 2006 in $0.64-1.28$ MeV channel. }
\caption{The figure shows the sample frames of the actions from UCF101 dataset~\cite{Soomro12}. The color of frame borders specifies the action type to which they belong: {\color{NavyBlue} Human-Object Interaction}, {\color{red} Body-Motion Only}, {\color{RoyalPurple} Human-Human Interaction}, {\color{YellowOrange} Playing Musical Instruments}, {\color{ForestGreen} Sports} (c.f.\\ref{sec:appendixA}).}
\caption{(colour online) (a) Angle averaged hydrogen NCP $\bar F(y,q)$ for SW at $T$ = 271 K {\color{blue}blue error bars}. The best fit of this function using M2 is plotted as {\color{red}red line}. The experimental resolution $R(y,q)$ is plotted as {\color{green}green line}. (b) Radial momentum distributions $4\pi p^2n(p)$ for SW ({\color{blue}blue line}) and ice ({\color{red}red line}) at $T$ = 271 K. The difference between SW and ice line-shapes (magnified by a factor of 10) is plotted as a {\color{green}green line}.}
\caption{(colour online) (colour online) (a) $\ek$ values for SW at T=271 K ({\color{blue}blue full square}), ice at T=271 K ({\color{red}red full square}), and water at 300 K ({\color{green}green full circle}) from present DINS study; (b) $\ek$ values for ice at T=5 K, T= 71 K\cite{senesir_2013} and T=271 K ({\color{black}black full square})\cite{2011Flammini} and water at 300 K and above ({\color{black}black full square}) from previous DINS measurements\cite{2008Pantalei_prl}. The insert shows a magnified picture in the temperature range 270 K - 305 K.}
\caption{(colour online) Values of $\ek$ for SW, ice, and water. a) $\ek^{DINS}$ values for SW from present study at T=271 K ({\color{blue}blue}), from path integral molecular dynamics simulation at T=270 K\cite{ramirez2011kinetic} and from previous DINS experiments at T=269 K\cite{2008_pietropaolo_prl}, T=271 K\cite{2008_pietropaolo_prl} and T= 273 K \cite{BZ_2009}; $\ek^{INS}$ values from previous INS experiment at the same corresponding temperatures\cite{2013senesi_OH_str_jcp}; b) $\ek^{DINS}$ for ice from present study at T=271 K ({\color{red}red}), from path integral molecular dynamics simulation at T=270 K\cite{ramirez2011kinetic} and from previous DINS experiments at T=271 K\cite{2011Flammini}; $\ek^{INS}$ value from previous INS experiment at T=271 K and $\ek^{M}$ value from an harmonic theoretical model at T=271 K\cite{2014moreh}; c) $\ek^{DINS}$ for water from present study at T=300 K({\color{green}green}) and from previous DINS experiments at T=300 K\cite{2008Pantalei_prl}, $\ek^{INS}$ and $\ek^{M}$ values for water from previous studies\cite{2013senesi_OH_str_jcp,2008Pantalei_prl,BZ_2009,2014moreh}.}
\caption{\babar\NLO measurements: Vector dominance model (VDM) fits of the squared form-factors using a Gounaris-Sakurai (GS) parametrization. Left:$\pip\pim$ \cite{Aubert:2009ad,Lees:2012cj}. Right: $\Kp\Km$ \cite{Lees:2013gzt}. \label{fig:VDM} }
\caption{Contributions to $a_\mu$ for recent \babar\publications: comparison of the measured value to the previous world average on the energy range$\sqrt{s'}< 1.8\,\gev$ (units $10^{-10}$). \label{tab:comparison} }
\caption{\color{red}Change in fluorescence spectrum as a function of time during decay (top) and recovery (middle) of sample C in the one measurement that shows a small amount of recovery. The light dashed blue curve on the decay plot shows the spectrum after the sample has recovered for the time specified. The bottom graph shows the decay and recovery fluorescence signal at peak fluorescence with exponential fits.}
\caption{Summary of the different scattering behaviours observed for $AB$, $BA$, $B\overline{A}$ and $A\overline{B}$ kink collisions: repulsion, annhilation, formation of a true or false domain wall (``sticking") and reflecting off each other. Here, the symbol ``\color{green}\tickYes\color{black}" indicates that the respective behaviour is observed, whereas ``\color{red}\tickNo\color{black}\,'' denotes that the respective behaviour is not exhibited.}
\caption{$(a)$ A diamond in a cubic graph, and $(b), (c)$ the only two ways to decompose it in a {\clawtriangledecomp}.}
\caption{Growth of \textcolor{blue}{$a(n)$} (odd partitions) against \textcolor{red}{$b(n+2)$} (chiral partitions) on a logarithmic scale}
\caption{Posterior parameters estimates and 95\% Credible Intervals (CI) for the selected model \filltriangle[0.17] and for the traditional joint model \fillpentagon[0.17] with a baseline adjusted with P-Splines.}
\caption{Histogram of the posterior means for the $\sigma_i$'s (model \filltriangle[0.17]). There's a considerable heterogeneity among these right skewed estimates.}
\caption{Posterior mean estimates, together with the corresponding 95\% Credible Bands (CB), for the selected model \filltriangle[0.17]. The top left panel shows $g_0=\log{(h_0)}$ and the subsequent panels have the time-varying regression coefficients as a function of time in years, $t$. }
\caption[]{% A robot (\subref{fig:intro-all})~uses real-time feedback from contact sensors to manipulate an bottle on a cluttered table. (\subref{fig:intro-relative})~Our approach uses policies from the obstacle free, hand-relative contact manipulation problem to a search in the full state space. (\subref{fig:intro-lattice})~The configuration of the robot is represented as a point in a state-space lattice consisting of feasible (\feasibleswatch), infeasible (\infeasibleswatch), and unevaluated (\unevaluatedswatch) edges. The probability distribution over the pose of the bottle is shown as a collection of semi-transparent renders. Best viewed in color. }
\caption[]{% Comparison of the (\subref{fig:planner-lazy-full})~full lattice to the (\subref{fig:planner-lazy-partial})~subset visited by DESPOT, with feasible (\feasibleswatch), infeasible (\infeasibleswatch) and unevaluated (\unevaluatedswatch) edges. We interleave planning with lattice construction to construct only the subset of the lattice visited by the search. This figure is best viewed in color. }
\caption[]{% Performance of the Rel-QMDP (RM \mdpswatch), Rel-SARSOP (RS \sarsopswatch), and Rel-DESPOT (RD \despotswatch) policies on Rel-POMDP. (Left)~Value $V_\text{rel}$ achieved by each policy computed over 100 timesteps. Note that the $y$-axis is inverted; lower (less negative) is better. (Right)~Probability that the movable object is in $\Xgoal$ as a function of timestep. Error bars denote a 95\% confidence interval. Best viewed in color. }
\caption[]{% Performance of the Rel-QMDP (RM \mdpswatch), Rel-SARSOP (RS \sarsopswatch), Rel-DESPOT (RD \despotswatch), Lift-QMDP (LM \liftmdpswatch), Lift-SARSOP (LS \liftsarsopswatch), and Lat-DESPOT (LD \latdespotswatch) policies on Lat-POMDP on four environments. (Top)~Value achieved by each policy after 100 timesteps. Note that the $y$-axis is inverted; lower (less negative) is better. (Middle)~Probability $p = \Prob(s_t \in \Xgoal)$ that the movable object is in the goal region at each timestep. (Bottom)~Probability that the execution is feasible as a function of time. Lift-QMDP, Lift-SARSOP, and Lat-DESPOT are omitted because they do not take infeasible actions. In all cases, error bars denote a 95\% confidence interval. Best viewed in color. }
\caption{Parameter values applied in \textbf{Case A2}. Parameters in \textcolor{red}{RED} are different from Case A1.}
\caption{Parameter values applied in \textbf{Case B}. Parameters in \textcolor{red}{RED} are different from Case A2.}
\caption{Space-time evolution of a one-dimensional random walk that clears out an interval---the desert---where food (shaded) has been eaten. Whenever the walk reaches a food-containing site, the food is consumed and the desert grows by one lattice spacing $a$. The walk starves (\textcolor{red}{$\boldsymbol{\times}$}) when it travels $\mathcal{S}$ steps without encountering food.}
\caption{\Fermi-LAT adaptively smoothed $>$100 MeV counts maps of the first four \gray\detected novae,\ncyg\2010,\nsco\2012,\nmon\2012, and\ndel\2013. Each nova is placed at the map centers (marked with yellow 1\deg\radius circles) and observed near bright diffuse Galactic\gray\emission (\nmon\is seen directly through the plane). This Figure is taken directly from\cite{ack14}.}
\caption{\Fermi-LAT $>$100 MeV 1-day binned light curves of the the first four \gray\detected novae (cf., Figure~1). Vertical bars indicate 1$\sigma$ uncertainties for $>3\sigma$ points (black) and $2$-$3\sigma$ (gray) significances, otherwise, 2$\sigma$ upper limits are indicated with gray arrows. Start times (from top to bottom panels) of 2013 August 16, 2012 June 19, 2012 June 15, and 2010 March 10 were defined as the day of the first LAT detection (there was a $2.4\sigma$ detection of \ndel\in 0.5-day binned data from August 16.5-17.0; see\cite{ack14}). The epochs of the observed optical peaks are marked with a blue diamond in each panel except for \nmon, where the optical peak was unobserved due to proximity to the Sun. This Figure is taken directly from \cite{ack14}.}
\caption{[left] \Fermi-LAT $>$100 MeV average \gray\spectrum of the classical nova\nmon\2012 over 22 days of data. Figure taken directly from\cite{ack14}. Vertical bars indicate 1$\sigma$ uncertainties for data points with significances $>2\sigma$, and arrows indicate 2$\sigma$ limits and the best-fit hadronic (from \p0-decay) and leptonic (from inverse Compton and lower-level bremsstrahlung emission) model curves are overlaid. See \cite{ack14} for details and for the plots for the other two LAT detected classical novae (\nsco, \ndel) and for the symbiotic nova \ncyg\(see \cite{abd10} as well). [right] A `composite' early-time \gray\spectrum of an ONe-type classical nova consisting of the\Compton\OSSE$0.05-0.3$ MeV spectrum from \nvel\1999 and the\Fermi-LAT $>$100 MeV spectrum of \nmon\2012 from the left panel.}
\caption{Preliminary \Fermi-LAT $>$100 MeV \gray\light curves (top panels) of\ncen\(left) and \nsgr\(right). Vertical bars indicate $1\sigma$ uncertainties for data points with TS $> 4$ (black), while gray arrows indicate $2\sigma$ upper limits when TS $< 4$. In each case, the \gray\light curve is compared to the AAVSO Visual light curve in 0.25 day bins for the same time interval (bottom panels).}
\caption{The optical depth of the universe from the IBL and the CBR as well as the total optical depth as a function of energy, given for redshifts of 0.1, 0.5, 1, 3, 5. It can be seen that the contribution to the optical depth from the IBL dominates at lower \gray\energies and redshifts and that from the CBR photons dominates at the higher energies and redshifts. The optical depth from CBR photons is an exact function of energy as given by equation (\ref{cbr}) and therefore the confidence band is becomes a thin line. The dashed lines indicate the opacities $\tau = 1$ and $\tau = 3$. }
\caption{A $\tau = 1$ energy-redshift plot (Fazio \& Stecker 1970) showing our uncertainty band results compared with the{\it Fermi} plot of their highest energy photons from FSRQs (red), BL Lacs (black) and and GRBs (blue) {\it vs.} redshift (from Abdo et al. 2010).}
\caption{Comparison of our results for $z = 1$ with those obtained from an analysis of blazar \gray spectra (Ackermann et al. 2012)}
\caption{ Fluxes of total energy and magnetic helicity normalized with the corresponding dissipation rates for the run with $k_fL = 40$. \blue{Black: normalized total energy flux $\Pi_E(k)/\vep$. Blue (dark gray): normalized magnetic helicity flux $\Pi_{H_b}(k)/\vep_{H_b}$. } The inset presents the scaling $\vep^-/\vep \propto \vep_{H_b}/(\vep L) \propto (k_fL)^{-1}$. \blue{ Black: $\vep^-/\vep$. Red (light gray): $\vep_b^-/\vep$. Blue (dark gray): $\vep_{H_b}/(\vep L)$. } }
\caption{ (a) Magnetic and kinetic energy spectra compensated by $k^{-2}$, where $E_u(k)$ has been shifted by a factor of $10^{-3}$. (b) Absolute value of magnetic helicity spectra. The spectra are collapsed by rescaling with $k/k_f$. The dashed lines correspond to the predictions from absolute equilibria, i.e. Eqs.~\eqref{eq:statmech}. \blue{The green (light gray) curves refers to the simulation with $k_fL=40$, the blue (dark gray) to $k_fL=20$ and the red (gray) to $k_fL=10$.} The inset shows the absolute value of the relative magnetic helicity $\rho_b(k)= H_b(k)/(\langle|\vec{a}_{\vk}|^2\rangle^{1/2}\langle|\vec{b}_{\vk}|^2\rangle^{1/2})$}
\caption{A schematic of general integral problems. $\Omega$ is the exterior region and the interior region $\Gamma$ consists of different arbitrary-shaped bodies. $\Gamma=\bigcup\limits_{i=1}^{P}p_{i}$, where $P$ is the total number of bodies and $p_{i}$ is the region occupied by $i$th body. The boundary the interior region is denoted as $\partial\Gamma$. \textcolor{red}{Confined geometries can also be considered as shown in Fig.~\ref{fig_color_dielectric}(b).}}
\caption{(a) \textcolor{red}{Relative errors} of the stray field in the unit sphere, and the stray field energy in the unit cube calculated from simulations compared to their analytical values, as a function of surface DoFs of the magnetic bodies. The inset numbers indicate convergence rates. (b) Computational time spent on the boundary integral as a function of surface DoFs, tested using a single CPU core on Intel Xeon E5-1607 v3 @ 3.1GHz. The inset numbers 1 and 2 indicate the scaling of $O(N)$ and $O(N^2)$ respectively. (c) Strong scaling of our computational approach for modeling a unit sphere with 7.4M volume DoFs and 70K surface DoFs, tested on high performance computer Blues at Argonne National Laboratory (ANL).}
\caption{\textcolor{red}{Relative errors of the stray-field energy in the unit sphere as a function of surface DoFs of the magnetic body using FMM and a direct method with linear shape functions, and FMM with quadratic shape functions.}}
\caption{\textcolor{red}{FMM computational time and relative errors of the stray-field energy for the unit cube as a function of octree height (left panel) and the Chebyshev order (right panel). The volume and surface DoFs of the unit cube considered here are 373248 and 30248 respectively.}}
\caption{\textcolor{red}{Relative errors of the body forces on unit dielectric spheres calculated from simulations compared to their analytical values for different DoFs/Sphere, as a function of center to center distance between two spheres.}}
\caption{ \textbf{\red{Basis states and gate operations of the charge quadrupole qubit.}} \textbf{(a)} A triple-dot charge qubit can be described in terms of a position basis, corresponding to the single-electron occupations of dots 1, 2, or 3. For a symmetrized triple dot, it is preferable to adopt, instead, an even-odd basis $\{\ket{C},\ket{E},\ket{L}\}$, referring to center, even, and leakage states, respectively. Here, $\ket{C}$ and $\ket{E}$ have even symmetry, $\ket{L}$ has odd symmetry, and the half-filled circles represent average occupations of 1/2. \textbf{(b)} The \red{charge quadrupole} eigenstates, $\ket{\tilde 0}$, $\ket{\tilde 1}$, and $\ket{\tilde L}$, obtained by solving \red{the system Hamiltonian,} equation~(\ref{eq:HCQ}), as a function of quadrupolar detuning, $\epsilon_\text{q}$. Here, we set \red{the tunnel couplings,} $t_\text{A}=t_\text{B}\, (\equiv t/\sqrt{2})=2.5$~GHz and \red{the dipolar detuning,} $\epsilon_\text{d}=0$. The insets depict the effect of $\epsilon_\text{q}$ on the triple-dot confinement potential. \textbf{(c)} A cartoon depiction of \red{microwave (AC)} and \red{pulsed (DC)} gate sequences useful for qubit manipulation. Initialization and readout are implemented in the far-detuned regime, $\epsilon_\text{q} \ll 0$, with $t\gtrsim 0$. In the DC scheme, $\epsilon_\text{q}$ is suddenly pulsed to the double sweet spot, $\epsilon_\text{q}=0$. Free evolution then yields an $X$ rotation in the logical basis $\{\ket{ C},\ket{E}\}$. For a $Z$ rotation, $t$ is suddenly pulsed to 0, while $\epsilon_\text{q}$ is pulsed away from zero (either positive or negative). In the AC scheme, $t\gtrsim0$ is held fixed, while an adiabatic ramp of $\epsilon_\text{q}$ to its sweet spot leaves the qubit in its logical ground state. $X$ and $Y$ rotations in the rotating frame are implemented by applying resonant microwave bursts with appropriate phases to $\epsilon_\text{q}$, centered at the sweet spot. Alternatively, microwaves may be applied to $t$, although we do not consider that possibility here.}
\caption{ \textbf{\red{Preserving the symmetry of the quadrupole qubit with external couplings.}} Schematic of the geometry of \textbf{(a)} a microwave stripline resonator coupled to \textbf{(b)} a quadrupole qubit. The stripline geometry shown is similar to those suggested in refs~\onlinecite{Frey2012,Petersson2012,Stockklauser2015,Mi2017}. Both accumulation-mode gates \red{that control the dot occupations} (with local potentials labeled $U_1$, $U_2$, and $U_3$) and depletion-mode gates \red{that control the tunnel couplings} (labeled $t_L$, $t_\text{A}$, $t_\text{B}$, and $t_R$) are included here (see refs~\onlinecite{Wu2014,Eng2015,Zajac2015,VeldhorstUnp} for a discussion). The corresponding dots are labeled 1, 2, and 3. The coupling occurs through the middle gate 2, which is connected to the resonator. The qubit can be coupled, similarly, to other qubits or charge sensors.}
\caption{ \textbf{\red{Simulated process fidelities for charge qubit gate operations.}} Simulations of bare $X_\uppi$ rotations and composite pulse sequences, $\tilde X_\uppi$, for \red{charge dipole (CD)} and \red{charge quadrupole (CQ)} qubits are performed in the presence of \red{dipolar ($\delta\epsilon_\text{d}$)}, and \red{quadrupolar ($\delta\epsilon_\text{q}=\delta\epsilon_\text{d}/40$) detuning fluctuations}. Plots show the infidelity ($=1-$fidelity) as a function of the standard deviation $\sigma_\epsilon$ of $\delta\epsilon_\text{d}$. (The simulations, charge noise averages, and process fidelity calculations are described in Methods.) Here, the blue and orange curves correspond to bare, single-pulse $X_\uppi$ rotations of CQ and CD qubits, respectively. The infidelity follows the same scaling in both cases, even though it arises from different mechanisms: pure dephasing for the CD qubit vs.\leakage for the CQ qubit. The gray curve corresponds to a composite, three-pulse sequence,\red{$\tilde X_\uppi \equiv Z_{2\uppi}X_{3\uppi}Z_{-2\uppi}$}, which removes the leading order $\delta \epsilon_\text{d}$ noise in the CQ qubit; no comparable sequence exists for CD qubits. The simple form of the CQ pulse sequence derives from the quadrupole geometry, which transfers some of the overhead for noise protection from the control pulse sequence to the qubit hardware. All simulations assume the same tunnel couplings ($t$ for the CD qubit; $t_\text{A,B}$ for the CQ qubit) of \SI{10}{GHz}. }
\caption{The hydrodynamic force effect of the upper fish row (presented by {\color{magenta}$\bullet$}) on the remaining fish school. interaction intensity: blue (low) to red (high). Upper subplot: diamond pattern. Remaining subplots: random patterns.}
\caption{The hydrodynamic force effect of the upper fish layer (presented by dark blue {\color{darkblue}$\bullet$}) on the remaining fish school. interaction intensity: blue (low) to red (high). {\bf Left}: side views (fish move to the right). {\bf Middle}: front views (fish move into page). {\bf Right}: Isometric view. First row: 27 fish, in a $6\times3\times3$ box. Second row: 64 fish, in a $8\times4\times4$ box. Third row: 125 fish, in a $10\times5\times5$ box. Fourth row: 216 fish, in a $12\times6\times6$ box.}
\caption{Upper and lower bounds on the capacity of a 2-user GS-CIC with respect to the cross-gain $g=\rho e^{i\phi}$. $1 < \rho\leq10^3$ and $\phi$ is chosen randomly in $\left[-\pi/4,+\pi/4\right]$ for each vaiation of $\rho$.} \label{fig:Fig_A} \end{figure} In second part of this work we introduce channel model, lattice structures and the complex case is defined for $2-$ user Gaussian symmetric interference channels (GS-IFC). In third part, the minimum mean square error (MMSE) estimator, the correspondent Hermitian form and the achievable sum-rate are defined by using the main framework of \cite{IEEE:Gastpar} and \cite{IEEE:Erez}. After showing the benefit of using a phase precoding scheme at transmitters, the new protocol of CoF combined with phase precoders is introduced. The phase precoders will use CSI as a feedback information to calculate the best phase precoder coefficients to improve the final achievable sum-rate. Our proposed precoding scheme also belongs to the general framework introduced by A. Sakzad \textit{et al.} in \cite{IEEE:Sakzad}; however, this scheme has not been analyzed in detail. Finally, at the last part of this work, the numerical results are presented for high $\SNR$ and strong interference channels. \subsection{Notational Convention} Through this paper, we use $\C$, $\Z$ and $\Z\left[i\right]$ to denote the field of complex numbers, the set of integers and Gaussian integers, respectively. $\F_q$ to denote the finite field of size $q$. Here in all equations, bold letters are for vectors and bold capital letters are for matrices, e.g., \textbf{x} and \textbf{X}, respectively. We define $\log_2^{+}\left(x\right)\triangleq\max\left (\log_2\left(x\right),0\right)$. We refer to lattices over complex integers as $\Z\left[i\right]-$lattices. We let $\left|z\right|$ and $\arg(z)$ denote the modulus and the phase of the complex number $z$, respectively. \section{channel model, lattice structure and The complex case} \subsection{Channel Model} The $K-$user GS-CIC is the general model of this paper. The transmitters reliably communicate linear functions to $M$ relays over a complex-valued symmetric interference channel. In this model each relay $m$ observes a noisy linear combination of the transmitted signals through the channel, \begin{equation} \label{Channel_Mod}{y_m}=\sum \limits_{l=1}^K h_{ml}\cdot{}{x}_{l}+{z}_{m} \end{equation} where $h_{ml}\in\C$ are the channel coefficients, $\mathbf{x}=\left[x_{1},\cdots,x_{K}\right]$, $\mathbf{y}=\left[y_{1},\cdots,y_{K}\right]$ and $\mathbf{z}=\left[z_{1},\cdots,z_{K}\right]$ are used to denote the input-vector, the output-vector and the complex i.i.d Gaussian noise-vector all of size \emph{K}, respectively. Furthermore, ${\mathbf{z}_m}\sim\mathcal{CN}\left(0,{\sigma}^2\right)$ and $\mathbf{x}_{l}\in\Lambda_{c}$. Let $\mathit{\Lambda_{c}}$ be the coding lattice which is common to all users. In this paper all users have the same power constraint i.e., $P_\mathrm{i} = P$, so the $\SNR$ is defined as $\SNR=\frac{P}{\sigma^2}$. By considering a simple lattice IA defined in \cite{IEEE:Erez} this $K-$user case will be approximately equivalent to a $2-$user case showed in Fig.~\ref{Fig:GS-CFC}, this means that, \begin{equation} \mathbf{H}=\left[\begin{array}{cc} 1 & g\\ g & 1 \end{equation}\end{array}\right]\end{equation} which the channel coefficients are considered to be in the field of complex numbers. The simplified expression of (\ref{Channel_Mod}) for $2-$user case is written as, \begin{equation} \label{Sim_Chan_mod} \left[\begin{array}{c}y_1\\y_2\end{equation}\end{array}\right]= \mathbf{H}\cdot\left[\begin{array}{c}x_1\\x_2\end{array}\right] +\left[\begin{array}{c}z_1\\z_2\end{array}\right]\end{equation} What we see in Fig.~\ref{Fig:GS-CFC}, the channel is symmetric, it means $\mathbf{H}(i,j) = g$ for all $i \not= j$, and $\mathbf{H}(i,i) = 1$ for all \emph{i}'s. \begin{figure}[h] \centering \begin{tikzpicture} \draw (0,0) rectangle (0.7,-0.7); \node at (-0.75,-0.35) {$w_1$}; \draw[->] (-0.52,-0.35)--(0,-0.35); \node at (0.35,-0.35) {$\mathcal{E}_1$}; \draw[->] (0.7,-0.35)--(2.25,-0.35); \node at (0.4,-0.95) {$\mathrm{TX}_{1}$}; \node at (0.9,-0.15) {${x}_1$}; \draw (2.4,-0.35) circle [radius=0.15]; \draw (2.3,-0.35)--(2.5,-0.35); \draw (2.4,-0.25)--(2.4,-0.45); \node at (1.8,-0.15) {$1$}; \node at (2.45,0.25) {${z}_1$}; \draw[->] (2.4,0.18)--(2.4,-0.2); \draw[->] (2.6,-0.35)--(3.25,-0.35); \node at (3,-0.15) {${y}_1$}; \draw (3.25,0) rectangle (3.95,-0.7); \node at (3.6,-0.35) {$\mathcal{D}_1$}; \node at (3.7,-0.95) {$\mathrm{RX}_1$}; \draw[->] (3.95,-0.35)--(4.47,-0.35); \node at (4.7,-0.3) {$\hat{w}_1$}; \draw (0,-2) rectangle (0.7,-2.7); \node at (-0.75,-2.35) {${w}_2$}; \draw[->] (-0.52,-2.35)--(0,-2.35); \node at (0.35,-2.35) {$\mathcal{E}_2$}; \draw[->] (0.7,-2.35)--(2.25,-2.35); \node at (0.4,-2.95) {$\mathrm{TX}_{K=2}$}; \node at (0.9,-2.15) {${x}_2$}; \draw (2.4,-2.35) circle [radius=0.15]; \draw (2.3,-2.35)--(2.5,-2.35); \draw (2.4,-2.25)--(2.4,-2.45); \node at (1.8,-2.15) {$1$}; \node at (2.45,-1.75) {${z}_2$}; \draw[->] (2.4,-1.82)--(2.4,-2.2); \draw[->] (2.55,-2.35)--(3.25,-2.35); \node at (3,-2.15) {${y}_2$}; \draw (3.25,-2) rectangle (3.95,-2.7); \node at (3.6,-2.35) {$\mathcal{D}_2$}; \node at (3.7,-2.95) {$\mathrm{RX}_2$}; \draw[->] (3.95,-2.35)--(4.47,-2.35); \node at (4.7,-2.3) {$\hat{w}_2$}; \draw[->] (1,-0.35)--(2.3,-2.25); \draw[->] (1,-2.35)--(2.3,-0.45); \node at (1.27,-1) {$g$}; \node at (1.3,-1.65) {$g$}; \end{tikzpicture} \caption{2-user Gaussian Symmetric complex-valued Interference Channel.}\label{Fig:GS-CFC} \end{figure} In addition, each transmitter must satisfy the power constraint, which for \emph{n} channel uses and user \emph{i} is given by: \begin{equation} \label{Chan_power_cont}{\frac{1}{n}} \sum_{j=0}^{n} \vert{\mathbf{x}_\mathrm{\emph{i}}}\vert^{2} \leq P_\mathrm{\emph{i}} \end{equation} In order to perform CoF, we need a couple of ingredients: \begin{itemize} \item \emph{Each} transmitter is equipped with an encoder $\mathcal{E}_{l}$, that maps length$-K$ messages over a \emph{lattice code}. \item \emph{Each} receiver will recover the transmitted messages by using a decoder $\mathcal{D}_{l}$, that maps the received messages over a \emph{lattice code} to its transmitted version. This mapping is possible with \emph{K} integers $\{a_{1},\cdots,a_{K}\}$ which will \textit{approximate} the channel gains $h_{ml}$. The main idea is that, if $\mathbf{x}_{l}\in\Lambda_{c}$ and if $\Lambda_{c}$ is a coding complex lattice, then \begin{equation} \lambda\triangleq\sum \limits_{l=1}^K a_{l}\mathbf{x}_{l}\in\Lambda_{c} \end{equation} as well. \end{itemize} \subsection{Lattice structure} We recall some fundamental notation of lattice codes which are essential throughout the paper. The definition of a \emph{N-dimension complex lattice $\Lambda$} with \emph{generator matrix} $\mathbf(G)\triangleq\left(\mathbf{g}_{1}^T| \mathbf{g}_{2}^T|\dots|\mathbf{g}_{N}^T\right)$, for all $\mathbf{g}_{j}\in\C^{m}$ and $1\le j\le N$ is the set of points in $\C^{m}$ (the $\mathbf{g}_{i}^T$ vectors are considered to be the columns of $\mathbf{G}$) \begin{equation} \label{Lattice_Model} \Lambda = \{\mathbf{x}=\mathbf{u}\mathbf{G}|\mathbf{u}\in{\Z\left[i\right]^{m}}\} \end{equation} when $m=N$, the lattice is called \emph{full rank}. Around each lattice point $\mathbf{x}\in\Lambda$ there is a \emph{Voronoi region} $\nu\left(\mathbf{x}\right)$ defined as: \begin{equation} \label{Ver_region} \nu\left(\mathbf{x}\right)= \{\mathbf{y}\in\C^{m}:\|\mathbf{y}-\mathbf{x}\|\le \|\mathbf{y}-\lambda\|,\forall\lambda\in\Lambda\} \end{equation} If $\Lambda'$ is a lattice itself, then a subset $\Lambda'\subseteq\Lambda$ is called a \emph{sublattice}. By having a sublattice $\Lambda'$, the \emph{lattice code} $\Lambda/\Lambda'$ can be define. $\Lambda/\Lambda'$ includes a finite constellation of lattice points from the lattice $\Lambda$. An intelligent choice \cite{IEEE:Sloane} for the sublattice $\Lambda'$ is $a\Lambda$ for some $a\in\Z\left[i\right]$. The shape of the lattice constellation is determined by the Voronoi region of the lattice $\Lambda'$. we can also define the \emph{lattice modulo operation} as \begin{eqnarray*} {Q_\Lambda} {\left(\mathbf{x}\right)}\triangleq\argmin_{\lambda\in\Lambda} \|\mathbf{y}-\mathbf{\lambda}\| \\ \mathbf{y}\mod\Lambda\triangleq\mathbf{y}-{Q_\Lambda} {\left(\mathbf{x}\right)} \end{eqnarray*} The generator matrix of a lattice can be found by applying the Cholesky decomposition to a definite positive Hermitian form of a complex lattice. \subsection{The Complex Case} In the complex case, we have another degree of freedom. Now, instead of dealing with $\Z-$lattices, we consider an $\mathcal{O}_{\K}$-lattice $\Lambda_{c}$ where $\mathcal{O}_{\K}$ is the ring of integers of a complex quadratic number field $\K=\Q(\sqrt{-b})$ for some square free positive integers $b$. When $\mathcal{O}_{\K}$ is a Principal Ideal Domain (PID), then the definition of an $\mathcal{O}_{\K}$-lattice is similar to the one of a $\Z-$lattice, and there exists lattice basis. Unfortunately, when $\K$ is a complex quadratic number field, then $\mathcal{O}_{\K}$ is not in general a PID. It is a PID when $b\in\{1,2,3,7,11,19,43,67,167\}$. The most natural choice, in the complex case , is to mimic the real case, that is take $b=1$, which implies $\mathcal{O}_{\K}=\Z[i]$. A more judicious choice has been made in \cite{IEEE:Narayan} with $b=3$ giving rise to $\mathcal{O}_{\K}=\Z[\omega]$ where $\omega=\zeta_{3}$ \footnote{Here, as usually, $\zeta_{q}$ denotes a primitive $q^{th}-$root of unity.}. As a lattice, $\Z[\omega]\simeq A_{2}$, that is the hexagonal lattice, when $\Z[i]\simeq\Z^{2}$. This is explained by the fact that $A_{2}$ is better for quantization than $\Z_{2}$. \section{MMSE estimator and The Compute-and-Forward With Phase precoding} Suppose that we code over \emph{n} time-slots where \emph{n} is the degree of the relative number field $\K$ over quadratic complex field $\F$ which can be, for instance, $\Z[i]$ or $\Z[\omega]$. In this part first we aim at computing the MMSE, $n\times n$ matrix B, Then we define our proposed phase precoding scheme to improve the behavior of the achievable sum-rate. \subsection{the minimum mean square error estimator} We suppose, for sake of simplicity, that all transmitted signals $\mathbf{x}_{k}$ are coming from a 1-dimensional $\mathcal{O}_{\K}-$lattice, which is obviously $\mathcal{O}_{\K}$ itself. Let $Gal_{\K/\F}=\{\sigma_0,\sigma_1\,\dots,\sigma_{n-1}\}$ denotes the Galois group of the extension $\K$ over $\F$ and set \begin{equation} \mathbf{A}_{k}=diag(\sigma_0(a_k),\sigma_1(a_k),\dots,\sigma_{n-1}(a_k)) \end{equation} Where $a_k\in\mathcal{O}_{\K}$. In the same way, the transmitted vectors are $\mathbf{x}_{k}=diag(\sigma_0(x_k),\sigma_1(x_k),\dots,\sigma_{n-1}(x_k))$, with $x_k\in\mathcal{O}_{\K}$. The received vector is expressed in (\ref{Channel_Mod}). Now we multiply (\ref{Channel_Mod}) by the MMSE matrix B and get, \begin{equation} \label{MMSE_mult} \mathbf{B}\mathbf{y}=\sum \limits_{k=1}^K h_{k}\mathbf{B}\mathbf{x}_{k}+\mathbf{B}\mathbf{z} \end{equation} But we want to decode the equation $\sum\limits_{k=1}^K a_{k}\mathbf{x}_{k}$, which means, the lattice point $\sum\limits_{k=1}^K h_{k}\mathbf{B}\mathbf{x}_{k}$ through canonical embedding. So, equation (\ref{MMSE_mult}) becomes, \begin{equation} \label{MMSE_mult_0} \mathbf{B}\mathbf{y}=\sum\limits_{k=1}^K \mathbf{A}_{k}\mathbf{x}_{k}+\tilde{\mathbf{z}} \end{equation} where $\tilde{\mathbf{z}}=\sum\limits_{k=1}^K (h_{k}\mathbf{B}-\mathbf{A}_{k}) \mathbf{x}_{k}+\mathbf{B}\mathbf{z}$ is the equivalent noise. Minimizing the mean square error is equivalent to minimize, \begin{equation} \begin{array}{c} \label{MMSE_mult_1} \varepsilon=\frac{1}{n}\E\left[\tilde{\mathbf{z}}^{\dagger}\tilde{\mathbf{z}}\right]\\ =\frac{P_{x}}{n}\sum\limits_{k=1}^K \Tr\left[(h_{k}\mathbf{B}-\mathbf{A}_{k})(h_{k} \mathbf{B}-\mathbf{A}_{k})^{\dagger}\right]+N_{0}\Tr\left[BB^{\dagger}\right] \end{array} \end{equation} where $P_{x}$ is the per user power and $N_0$ is the per component noise variance. Let us now rewrite the equation (\ref{MMSE_mult_1}), \begin{equation} \begin{array}{c} \varepsilon=\frac{P_{x}}{n}\alpha \Tr\left[\mathbf{B}\mathbf{B}^\dagger- \mathbf{B}\mathbf{C}^{\dagger}-\mathbf{C}\mathbf{B}^{\dagger}+\frac{1}{\alpha} \sum\limits_{k=1}^K \mathbf{A}_{k}\mathbf{A}_{k}^\dagger\right]\\ =\varepsilon=\frac{P_{x}}{n}\alpha \Tr\left[(\mathbf{B}-\mathbf{C}) (\mathbf{B}-\mathbf{C})^{\dagger}+\frac{1}{\alpha} \sum\limits_{k=1}^K \mathbf{A}_{k}\mathbf{A}_{k}^{\dagger}- \mathbf{C}\mathbf{C}^{\dagger}\right] \end{array} \end{equation} with $\alpha=\|h\|^{2}+\frac{1}{\rho}, \rho=\frac{P_{x}}{N_{0}}$ and \begin{equation} C=\frac{1}{\alpha}\sum\limits_{k=1}^K h_{k}^{*}\mathbf{A}_{k} \end{equation} The minimization of $\varepsilon$ with respect to $\mathbf{B}$ is equivalent to set $\mathbf{B}=\mathbf{C}$, in which case the expression of $\varepsilon$ is \begin{equation} \begin{array}{c} \label{MMSE_mult_2} \varepsilon=\frac{P_{x}}{n} \Tr\left[\sum\limits_{k=1}^K \mathbf{A}_{k}\mathbf{A}_{k}^{\dagger} -\frac{1}{\alpha}\left(\sum\limits_{k=1}^K h_{k}^{*}\mathbf{A}_{k}\right) \left(\sum\limits_{k=1}^K h_{k}\mathbf{A}_{k}^\dagger\right)\right]\\ =\frac{p_{x}}{n} \{\sum\limits_{k=1}^K\Tr_{\K/\F}\left(|a_{k}|^{2}\right) -\frac{1}{\alpha}\sum\limits_{i,j=1}^K h_{i}h_{j}^{*}\Tr_{\K/\F}\left(a_{i}a_{j}^{*}\right)\}. \end{array} \end{equation} \subsection{The Compute-and-Forward With Phase Precoding} In terms of making IA, we aim to not divide the general GS-CIC to two separate channels in parallel (create from real and imaginary parts of the original complex-valued channels). The objective is to keep the GS-CIC in its original shape, this choice will intercept of having interference between parallel channels. The same CoF protocol introduced in \cite{IEEE:Erez} is used in this section for GS-CIC. We propose a phase precoding scheme which is using channel coefficients as feedback to improve the results of previous works. In CoF protocol the computation rate is known as a maximal rate which users can send their codewords to a destination, this rate is given by: \begin{equation} \label{Computation_rate} R(\mathbf{h},\mathbf{a})=\frac{1}{2}\log_2^{+}\left\{\left({\parallel \mathbf{a} \parallel^{2}}-\frac{\mathsf{SNR}|\mathbf{h}^{*}\mathbf{a}|^{2}}{1+\mathsf{SNR} {\parallel \mathbf{h}\parallel^{2}}}\right)^{-1}\right\} \end{equation} where $\mathbf{a}$ is a vector of Gaussian integers characterizing the equation we want to decode. We are interested to improve the behavior of achievable sum-rate given by CoF protocol in the \emph{strong} and \emph{very strong} interference regimes. For $2-$user GS-CFC, $\mathbf{h}=[1,g]$ and the interference-to-noise ratio is defined as $\mathsf{INR}\triangleq |g|^{2}\mathsf{SNR}$. Here, $g=\rho e^{i\phi}\in\C$ is the cross channel coefficient and $\mathbf{a}=[a_1,a_2]$ with $a_1,a_2\in\Z[i]$. The bloc fading channel model is considered, then the simplified expression of (\ref{Computation_rate}) is \begin{equation} \label{Computation_rate_form} R(\mathbf{h},\mathbf{a})=\frac{1}{2}\log_{2}^{+}\left\{ \frac{\left(\frac{1} {\mathsf{SNR}}+(1+|g|^{2})\right)}{q\left(a_1,a_2\right)}\right\} \end{equation} where for a complex-valued channel $q(a_1,a_2)$ is a definite positive Hermitian form equal to: \begin{equation} \label{quadratic_form} \begin{array}{c} q\left(a_1,a_2\right)=(|a_1g|-|a_2|)^{2}+\frac{1}{\mathsf{SNR}}(|a_1|^{2}+|a_2|^{2})\\ -2\Re({a_{1}a_{2}^{*}g})+2|a_{1}a_{2}g| \end{array} \end{equation} We can deduce the Hermitian matrix of (\ref{quadratic_form}) as, \[G= \left(\begin{array}{ c c } |g|^{2}+\frac{1}{\mathsf{SNR}} & -g \\ -g^* & 1+\frac{1}{\mathsf{SNR}} \end{array} \right)\] As a definite positive Hermitian form, $q(a_1,a_2)$ defines a rank $2$ complex-lattice, $\Lambda_{\mathrm{CF}}$. we aim at finding the two successive minima of (\ref{quadratic_form}). The Hermitian form is of dimension $2$, an algorithm for finding the two successive minima is the Complex-LLL algorithm defined by \cite{IEEE:Cong}. This algorithm will reduce the basis of a lattice in complex field. Let define matrix $B=\mathbf{Cholesky}(G)$ be a basis of $\Lambda_{\mathrm{CF}}$ and $B_{\mathrm{red}}$ be the reduced basis after complex-LLL reduction. $U$ is the Unimodular basis change matrix. Call $G_{\mathrm{red}}=B_{\mathrm{red}}^T B_{\mathrm{red}}$ the reduced Gram matrix, then the two successive minima are the diagonal entries of $G_{\mathrm{red}}$. If we fix the $\rho$ part of the cross channel coefficient, and varying its phase between $[0,\pi]$, it can be observed in Fig.~\ref{fig:Fig_B} when the phase is approximately around $0$ or $\pi$ the achievable sum-rate will be maximal. In this scenario $\arg(g)$ should be aligned to reach the maximal achievable sum-rate. \begin{figure}[!h] \centering \includegraphics[width=3in]{Fig_B} \caption{Upper and lower bounds on the capacity of a 2-user GS-CIC with respect to the phase variation $\phi$. $g=\rho e^{i\phi}$ with $0\le\phi\le\pi$ and $\rho=7.0770$ is set to be unchangeable.} \label{fig:Fig_B} \end{figure} After analyzing the method described in \cite{IEEE:Sakzad}, we propose our precoding scheme. In our scheme each codeword should be multiplied by a precoder before transmitting them to destinations, this will let us to align their phase to maximize the final achievable sum-rate. To realize this scheme we assume that the precoder for each transmitter is a complex scalar $e^{i\theta_k}$, for some $-\pi/4\le\theta_k\le+\pi/4$, multiplying the lattice codeword. Finding the optimal precoder factors is realized by using the CSI available as feedback information at transmitters. We apply the phase precoding function as, \begin{equation} \label{Phase_precoder_1} \begin{array}{c} D_{k}:\C^n\to\C^n\\ D_{k}\left(\mathbf{x}_k,\theta_{k}\right)\triangleq e^{i\theta_{k}}\mathbf{x}_k \end{array} \end{equation} for $\theta_{k}\in\left[-\pi/4,\pi/4\right]$ and $1\le k\le K$. In the $2-$user case, we send $e^{i\theta_1}x_{1}$ and $e^{i\theta_2}x_{2}$ instead of $x_1$ and $x_2$. The phase precoded $e^{i\theta_{k}}\mathbf{x}_k$ continues to satisfy the power constraint (\ref{Chan_power_cont}). In general the equivalent channel coefficients vector is, \begin{equation} \label{Phase_precoder_2} \begin{array}{c} \mathbf{h}'=\left[h_{1}e^{i\theta_1},h_{2}e^{i\theta_2}\right]\\=\left[\rho_{1}e^{i\times\left(\phi_{1}+\theta_1\right)},\rho_{2}e^{i\times\left(\phi_2+\theta_2\right)}\right]\end{array} \end{equation} When the channel is symmetric, $h_{1}=1$. In our proposed precoding scheme after applying precoders, the new channel coefficients are $\mathbf{h}'= \left[e^{i\left(\theta_1\right)},\rho_{2}e^{i\left(\phi_2+\theta_2\right)}\right]$ , the optimal choice for precoders is therefore, \begin{equation} \begin{array}{l} \theta_1=0\\ \theta_2=-(\phi_{2}\pm\eta) \end{array} \end{equation} $\eta\in\left[-0.2~\radian,+0.2~\radian\right]$ is the precoder precision factor, it is used to keep $\left(\phi_2+\theta_2\right)$ approximately in an area which the rate shown in Fig.~\ref{fig:Fig_B} is maximal. We dispose the CSI, then calculating $\theta_2$ is feasible by using four quadrant $\arctan$ function, $\theta_2=\arctan_2(\Im(g)/\Re(g))$. The numerical result of this scheme is presented in the next section to show its performance. \section{Simulation Results} The proposed scheme defined in section I\hspace{0.08mm}I\hspace{0.08mm}I was implemented with following parameters: The channel is considered as a 2-user GS-CIC with $h_1=1$ and $h_2=\rho e^{i\phi}$. To calculate the sum-rate, $\SNR$ is chosen to be 65 dB and the lattice reduction is done in the complex field. For evaluating the performance of our proposed scheme and being in strong and very strong interference regimes, $\rho$ is set to vary in $\left[1.1,10^3\right]$ with a very high precision. Then $\phi$ is selected ramdonly in $\left[-\pi/4,\pi/4\right]$ for each variation of $\rho$. The reason for choosing $\phi$ in this interval, is because of using $\Z\left[i\right]$-lattices. \begin{figure}[!h] \centering \includegraphics[width=3in]{Fig_C} \caption{Upper and lower bounds on the capacity of a 2-user GS-CIC with respect to the cross-gain. Our proposed Phase precoding technique is used.} \label{fig:Fig_C} \end{figure} It can be observed from Fig.~\ref{fig:Fig_C} that for this configuration, the phase precoding scheme defined in previous section improve the achievable sum-rate of the CoF protocol \cite{IEEE:Erez} showed in Fig.~\ref{fig:Fig_A} for the case of 2-user GS-CIC. As we can see in Fig.~\ref{fig:Fig_C}, the deep fadings have been limited to 0.3 [Bits/Channel Use], and there is no more gap between the upper bound and the achievable sum-rate. It means by using this scheme the interference have been aligned and also the performance of the CoF protocol has been improved. \section{Conclusion} In this paper, after studying the channel model, $\Z\left[i\right]-$lattice structure and the MMSE estimator matrix for the complex-case, we have showed the existence of a phase precoding scheme with partial feedback for the CoF protocol defined in \cite{IEEE:Gastpar}, which will increase the finale communication rate between users. This proposed precoding scheme is enabling the relays to decode reliably to linear combination of lattice points with complex integer coefficients. Numerical results suggest that using our precoding scheme will limit the fading-like behavior and also achieve significant gain improvement in the case of 2-user GS-CIC as presented in \cite{IEEE:Erez}. Choosing the best precoder factor with a very high precision is essential step to improve performance of the CoF protocol when the interference is available. Future work will include a more detailed study of precoding schemes in the case of lattices over Eisenstein integers. \begin{thebibliography}{1} \bibitem{IEEE:Ashlswede} R.~Ahlswede, ``Multi-way communication channels", in \emph{Proc. IEEE International Sympos. on Inform. Theory,} 2:23-52, 1974. \bibitem{Berkeley:Shannon} C. E.~Shannon, ``Two-way communication channels," in \emph{Proc. Berkeley Sympos. on Mathematical Statistics and Probability}, 1962. \bibitem{Book:Syed} Syed A.~Jafar, ``Foundations and Trends in Communications and Inform. Theory", Vol. 7, No. 1, USA, 2011. \bibitem{IEEE:Motahari} A. S.~Motahari, S. O.~Gharan, M. A~Maddah-Ali, and A. K.~Khandani, ``Real interference alignment: Exploiting the potential of single antenna systems", in \emph{IEEE Trans. Inform. Theory}, 2009. \bibitem{IEEE:Carlerial} A.Carlerial, ``A case where interference does not reduce capacity," in \emph{IEEE Trans. on Inform. Theory}, 21(5):569-570, 1975. \bibitem{IEEE:Jafarian} S.~Sridharan, A.~Jafarian, S.~Vishwanath, and S. A.~Jafar,``Capacity of symmetric k-user Gaussian very strong interference channels," in \emph{Proc. IEEE Global Telecommunications Conf. GLOBECOM}, pages 1-5, 2008. \bibitem{IWCIT:Ehsan} E. Ebrahimi Khaleghi and JC. Belfiore, ``Compute-and-forward for the interference channel: Diversity precoding", in \emph{Iran Workshop on Communication and Inform. Theory}, 1-6, 2014. \bibitem{IEEE:Gastpar} B.~Nazer and M.~Gastpar, ``Compute-and-forward: Harnessing interference with structured codes", in \emph{IEEE International Sympos. on Inform. Theory}, pp.~772-776, 6-11 July 2008. \bibitem{IEEE:Erez} O.~Ordentlich, U.~Erez and B.~Nazer, ``The approximate sum capacity of the symmetric Gaussian K-user interference channel", in \emph{Proceedings of the IEEE International Sympos. on Inform. Theory}, pp. 2072-2076, 1-6 July 2012. \bibitem{IEEE:Narayan} N.E. Tunali, K.R. Narayanan, J. Boutros and Y.C. Huang, ``Lattices over Eisenstein integers for CoF", in \emph{Annual Allerton conf. on communication, control and computing}, 2012, pp. 33-40. \bibitem{IEEE:Sakzad} A. Sakzad, E. Viterbo, J. Boutros and Y. Hong, ``Phase Precoded Compute-and-Forward with Partial Feedback", in \emph{IEEE International Sympos. on Inform. Theory}, 2117-2121, 2014. \bibitem{IEEE:Sloane} J. Conway and N. Sloane, ``A fast encoding method for lattice codes and quantizers", in \emph{IEEE Trans. on Inform. Theory}, vol. 29, pp. 820-824, 1983. \bibitem{IEEE:Cong} Y.H. Gan, C. Ling and W.H. Mow, ``Complex lattice reduction algorithm for low-complexity full-diversity MIMO detection", in \emph{IEEE Trans. Signal Processing}, vol. 57, pp. 2701-2710, July 2009. \end{thebibliography} \end{document} }}\end{array}\end{equation}}
\caption[Principle of MIMO radar]{ \label{fig:geomrad} Principle of MIMO radar: $\times$ and \tikz \node (r1) at (0,-0.05cm) [shape=circle,draw,inner sep =0.1cm] {}; correspond to transmit and receive antennas. Throughout, we assume the spacing of the $N_T$ transmit and $N_R$ receive antennas to be $d_T= \frac{1}{2f_c}$ and $d_R= \frac{N_T}{2f_c}$, where $f_c$ is the carrier frequency. }
\caption{Visualization of error statistics for generating novel views from a single input view on the \emph{car} category. The heatmaps ({\color{blue} blue}--low, {\color{red} red}--high) depict the mean pixel error for obtaining the target view (columns) from the input view (rows) for the baseline~\cite{tatarchenko2015single} (left) and our approach(middle). Some common failure modes of our method are visualized on the right.}
\caption{View synthesis examples using our multi-view network. Each input view makes independent prediction of a candidate target view as well as a selection/confidence mask ({\color{blue} blue}--low, {\color{red} red}--high). The final prediction is obtained by linearly combining the single-view predictions with weights normalized across the selection masks. Typically, the final prediction is more similar to the ground-truth than any independent prediction.}
\caption{ Topological invariants assigned to the coin angles of the 1D split-step walk (\figlab{a}) and the 2D protocol (\figlab{b}). Due to the form of the coin operator, $\op{C}(\theta)$, the walk possesses a $4\pi$-periodicity in the coin angles. At the phase boundaries, the gap closes at quasienergy $\epsilon = 0$ (dotted), $\epsilon=\pi$ (dashed), or both at $\epsilon=0$ and $\epsilon=\pi$ (dash-dotted). The coin angle pairs chosen in the numerical examples in this work, and the corresponding phase transitions defined in Eqs.~\eqref{eq:coin_angles_1D}, \eqref{eq:coin_angles_2D} are also displayed (\starSym\!\textbf{---}\!\starSym). The 1D Hadamard walk $(\theta_1, \theta_2) = (\pi/2, 0)$, which is discussed in Fig.~\ref{fig:winding}, is also shown (\diamondSym).\vspace{-2mm} }
\caption{ Object localization using Hand Information. Visualization of object location probability map (\textcolor{red}{\textbf{red}}) and object bounding box (\textcolor{green}{\textbf{green}}). (a: GTEA, b: Gaze+) }
\caption{Audio description (AD) and movie script samples from the movies ``Harry Potter and the Prisoner of Azkaban'', ``This is 40'', and ``Les Miserables''. Typical mistakes contained in scripts marked in \redtext{red italic}.}
\caption{Numerical simulations: (a) Comparison between the chain shape observed experimentally (images in the background) and the prediction of numerical simulations (overlaid dashed curve). Here there are no adjustable parameters ($L=54\mathrm{~cm}$, $R=4\mathrm{~cm}$, time between two frames $\Delta t=0.9\mathrm{~ms}$). (b) At each instant before snapping, the chain may be divided into four zones. (c) Plot of the curvature $\kappa$ (dashed curve) and the chain speed $v$ (solid curve) as functions of the arc length. (d) A numerical reconstruction of the tension $T$ within the chain shows that it varies linearly with arc length in regions (I), (III) \& (IV). Region (II) bridges regions (I) and (III), each of which is characterized by a different acceleration. For comparison, the dashed line shows the tension as a function of arc length in the idealized solution with uniform and constant acceleration$a$ where lift-off is prevented (this artificial solution would give rise to a negative contact pressure).}
\caption{ Searching for self-similar solutions in the snapping (for this particular simulation, the linear mass is $\rhol = 5$ and the radius of pulley is $R=1$ and $L=2\pi$ such that $\pi_2=R/L=1/(2\pi)$. There is no gravity and no bending rigidity. The simulation timestep is $\Delta t=2\times10^{-4}$ and the mesh size is $\Delta s = 3.14\times10^{-3}$. Imposed acceleration at endpoint is $a=0.5$). (a) Identifying the singularity exponent $\alpha$ for the curvature and the critical time $\overline t^*$ using the quantity $\overline K(\overline t)$ defined in \eqref{eq:K}. From our numerical simulations we find that $(\alpha,\overline{t}^*)=(0.659,1.41531)$ (red dashed line). (b) These values may be checked in a standard log-log plot. The wells are numerical artifacts: they correspond to values of $|\overline t^*-\overline t|$ as small as a few simulation steps, implying that the curvature is very peaked and prone to discretization errors. (c) Configurations right before and after the snapping time, $1.405\leq \overline t \leq 1.42$ ($\overline t^*\approx 1.415$): darker configurations correspond to times closer to snapping and reference configurations are denoted by the symbols (\protect\tikz \protect\draw[black,fill=PineGreen!10] (-.11,0) -- (.0,.11) -- (.11,0)-- (0,-.11)--(-.11,0);), (\protect\tikz \protect\draw[black,fill=PineGreen!10] (0,0) circle (.75ex);) and (\protect\tikz \protect\draw[black,fill=PineGreen!10] (-.09,-.09) -- (.09,-.09) -- (.09,.09)-- (-.09,.09)--(-.09,-.09);) (d)~Collapse of the curvature in rescaled variables $(\tilde s, \overline \kappa/\overline K)$ using the same shading convention for $1.405\leq \overline t \leq1.414$ (before snapping occurs).}
\caption{ Decomposition of road networks by \Shredder. Percentage of long-edges, road sections in paths, trees and other components in which the road networks is decomposed. }
\caption{The cost function of rate-$1/2$ MET-LDPC code for optimizing the degree distribution, $\boldsymbol{L}$, for fixed, $\boldsymbol{\Lambda}$. $\boldsymbol{L}=[a_1 ~a_2 ~a_3 ~a_4]$ where $a_4 = 1-a_1-a_2$ and $a_3= a_4$ to satisfy constraint (\ref{c1}). Blue colour \textquoteleft \textasteriskcentered \textquoteright ~shows the locations of local maxima.}
\caption{The cost function of rate-$1/10$ MET-LDPC code for optimizing the degree distribution, $\boldsymbol{L}$, for fixed, $\boldsymbol{\Lambda}$. $\boldsymbol{L}=[a_1 ~a_2 ~a_3]$ where $a_3 = 1-a_1-a_2$ to satisfy constraint (\ref{c1}). Blue colour \textquoteleft \textasteriskcentered \textquoteright ~shows the locations of local maxima.}
\caption{The cost function for optimizing the rate-$1/2$ MET-LDPC code structure, where $\boldsymbol{\Lambda} = [([0~1], [i_1 ~0 ~0 ~0]) \hspace*{0.3cm} ([0~1], [i_2 ~0 ~0 ~0]) \hspace*{0.3cm} ([1~0], [0 ~3 ~3 ~0]) \hspace*{0.3cm} ([0~1], [0 ~0 ~0 ~1])]$. Blue colour \textquoteleft \textasteriskcentered \textquoteright ~shows the locations of local maxima. }
\caption{The cost function for optimizing the rate-$1/10$ MET-LDPC code structure, where $\boldsymbol{\Lambda} = [([0~1], [3 ~0 ~k_1 ~0]) \hspace*{0.3cm} ([0~1], [3 ~0 ~k_2 ~0]) \hspace*{0.3cm} ([0~1], [0 ~0 ~0 ~1])]$. Blue colour \textquoteleft \textasteriskcentered \textquoteright ~shows the locations of local maxima. }
\caption{The architecture of our proposed network. Trainable layers are shown in {\color{green}green}. Each hidden layer is followed by a ReLu non-linearity~\cite{nair2010rectified}. The final scoring layer is a linear function of the last hidden layer. The rank loss acts on pairs of segments and is non-zero, unless $s^+$ scores higher than $s^-$ by a margin of 1. To emphasize that the loss acts on pairs of segments, we show the two passes separately, but we use a single network.}
\caption{The average (luminosity-weighted mean) specific spectral energy distribution of galaxies (at $z=8$ with $>10^{8}\,{\rm M_{\odot}}$) in \bluetides\assuming the{\sc pegase} SPS model. The top panel shows the pure stellar emission (i.e. $f_{\rm esc, LyC}=1$). The middle panel shows the nebular continuum and line emission (assuming $f_{\rm esc, LyC}=0$) while the bottom panel shows the total emission (including both nebular and stellar components) assuming $f_{\rm esc, LyC}=0$. The coloured markers denote the broadband luminosities expected assuming a simple top-hat filter over the wavelength interval given next to each point and denoted by the horizontal error bar.}
\caption{The specific (i.e. per unit stellar mass) LyC photon production rate (top-panel) and production efficiency $\xi_{\rm LyC}$ (bottom-panel) for galaxies at $z\in\{8,9,10,11,12,13,14,15\}$ in \bluetides.}
\caption{ {\color{black!100}Error events in sequential frame synchronization problem.} }
\caption{\color{black!100}A binary input binary output model for AWGN channel with transition probabilities $\epsilon_f = \mathsf{P}(y_A(1)|x_A(0)) =\mathsf{P}(n>a\sqrt{P}) $ and $\epsilon_m = \mathsf{P}(y_A(0)|x_A(1))=\mathsf{P}(n>(1-a)\sqrt{P})$.}
\caption{Important parameters for the \bluetides\simulation.\bluetides\uses the cosmology from the Wilkinson Microwave Anisotropy Probe nine-year data release\citep{WMAP}.}
\caption{Top Panel: Comparison of the intrinsic (solid blue lines) and dust corrected (solid green lines) \bluetides\LFs to available observational data from B15 (red data points) and\citet{McLeod} (purple data points). Shaded regions show the $1\sigma$ Poisson errors. Also shown are the best DPL fits (dashed blue and green lines) to the \bluetides\LFs and the Schechter fits to observation data from B15 (red lines). The galaxies to the right of the dashed black lines are those that are UV bright enough to be detected by\wfirst. Bottom Panel: \bluetides\prediction for the number density of objects like GN-$z11$. The error bar on the \citet{Oesch16} point is a simple Poisson error.}
\caption{Surface density photometry for the \wfirst\HLS band passes. Luminosities bey ond the Lyman break limit have been set to zero for the appropriate redshift in each band.}
\caption{Top Panel: Surface density for $z=8$ and beyond. The results for $z=9,11,$ and $13$ are left out for clarity. The \wfirst\field of view is$\sim 2200 {\rm deg}^2$, so will be able to detect galaxies out to $z=15$ according to the \bluetides\LF results. Our dust model is too extreme by$z=10$, so the lower limit at these redshifts are very conservative estimates. Bottom Panel: Predicted number of galaxies in the \wfirst\HLS at or above a given redshift. The predictions shown are for the\bluetides\galaxies (blue region) and AGN (green line) for the$5\sigma$ limit. Extrapolation of the B15 $z=8$ Schecther function is also shown (black dashed line). The grey lines show the same quantities but for the $10\sigma$ limit.}
\caption{AGN LFs for \bluetides. Shaded regions show the $1\sigma$ Poisson errors. Grey dashed lines show a best fit to a power law LF. The black arrow and dashed line indicates the galaxies that are UV brighter than the \wfirst\$5\sigma$ limit.}
\caption{AGN surface densities according to \bluetides\for AGN that have an intrinsic brightness above the\wfirst\$5\sigma$ limit (no dust correction assumed).}
\caption{Stellar mass vs halo mass for the \wfirst\galaxy population as predicted by\bluetides. Dashed lines show results from abundance matching by \citet{Behroozi}. Data points show the $z=7$ results from \citet{Harikane2015}. The 2D histogram shows the distribution of galaxies in the \bluetides\volume at$z=8$. Solid lines show the mean in bins of $M_{\rm Halo}$ for the higher redshifts as indicated in the figure.}
\caption{Galaxy properties for those bright enough to be detected by \wfirst. The 2D histogram shows the distribution of galaxies in the \bluetides\volume at$z=8$. The solid lines show the mean values in bins of stellar mass. sSFR observations come from \citet{Gonzalez2014}. Best fit to metallicity measurements at $z=3.5$ come from \citet{Maiolino2008} and observations at $z=5$ come from \citet{Faisst2016}.}
\caption{Star formation rate density prediction for the \wfirst\HLS. The solid line shows the predicted result for\wfirst, without dust correction. The dashed line shows the contribution from the total \bluetides\galaxy sample. Observational data points are for the HUDF\citep{Oesch14b}, HFF A2744 \citep{Oesch14a}, and observational estimates from CLASH \citep{Bouwens14b, Zheng12, Coe13}}
\caption{Bias measurements for the galaxies visible to \wfirst\from\bluetides\(blue points). Also shown is the linear bias computed from \citet{Tinker} for a given halo mass (dashed lines).}
\caption{Comparison of the BOSS CMASS sample \citep{Anderson:2012} and the \wfirst\$z=8$ sample as predicted by \bluetides. $P_{\rm gg}$ and $P_{\rm dm}$ are computed at the BAO scale $k=0.2 h/Mpc$ at the mean redshift of the sample in question. Ranges for $n$ (and $nP$) are computed using galaxies brighter than the \wfirst\$5\sigma$ according to the intrinsic and dust corrected galaxy luminosities. $V$ is the comoving volume for each survey's respective area coverage, assuming the same range in redshift centered about the mean redshift $\bar{z}$.}
\caption{$nP_{\rm gg}$ computed for the \wfirst\survey using the number density and bias computed from\bluetides. Upper (lower) limit uses the number density $n$ computed with the intrinsic (dust corrected) luminosities. Black line shows the value for the BOSS CMASS sample which detected the BAO peak at the $5\sigma$ level. \citep{Anderson:2012}}
\caption{Intrinsic luminosity functions for \bluetides. Dashed lines show the best fit DPL defined in equation \ref{dpl}.}
\caption{Dust corrected luminosity functions for \bluetides. Dashed lines show the best fit DPL defined in equation \ref{dpl}.}
\caption{{\color{blue}Blue:} $\mathbf{E}\left[|\pi\left(f^+\right)|\right]$ ($y$-axis) vs. $\xi^{\min}$ ($x$-axis) on synthesized experimental data. \\ {\color{red}Red:} $\gamma$ ($y$-axis) vs. the average of $\xi^{\min}$'s correspond to that $\gamma$ ($x$-axis). \\(a) $p={1-0.01\over0.01}$, $c=-\log \left(1-0.5\right)$. (b) $p={1-0.001\over0.001}$, $c=-\log \left(1-0.6\right)$. (c) $p={1-0.002\over0.002}$, $c=-\log \left(1-0.7\right)$. (d) $p={1-0.005\over0.005}$, $c=-\log \left(1-0.9\right)$.} \label{fig:gamma} \end{figure} \section{Related work} Exact inference on NOBN is fundamentally intractable~\citep{cooper90}. Brute force inference on NOBN is $O(|F|\cdot2^{|D|})$ as it calculates $P(F)$ by summing up $P(F\mid D')\cdot P(D')$, where $D'$ can be the combination of the presence or the absence of any subsets of $D$. Junction tree algorithms~\citep{pearl88} can be more efficient in practice at $O(2^{|M|})$, where $|M|$ is the maximal clique size of the moralized network. Quickscore~\citep{qs} reduces the temporal complexity to some exponential function of a quantity substantially smaller than $|D|$ or $|M|$ and make the inference practical for common usage. Quickscore~\citep{qs} achieves $\widetilde{O}\left(|D|\cdot 2^{|F|}\right)$ by exploiting marginal and conditional independence\footnote{the soft-$O$ bound is derived from $O\left(|D|\cdot |F^-| \cdot 2^{|F^+|}\right)$ given in \citep{qs}.}. \begin{table}[ht!] \centering \caption{Overall temporal complexities for exact and variational inferences on NOBN in terms of $|D|$, $|M|$, $|F|$, and $|F'|$ (note that all results are independent of $|S|$). In practical applications like \texttt{QMR-DT}, $|D|=534$, $|M|\approx 151$, and $|F|\approx 43$~\citep{j99,jj99}.} \label{tab:nobn} \vspace{-0.2cm} {\setlength{\extrarowheight}{6pt}% \begin{tabular}{ c | c | c | c } Brute force & Junction tree & Quickscore & Variational \\ \hline \small $O\left(|F|\cdot2^{|D|}\right)$ & \small $O\left(|D|\cdot2^{|M|}\right)$ & \small $\widetilde{O}\left(|D|\cdot 2^{|F|}\right)$ &\small $\widetilde{O}\left(|D|\cdot \left[|F'| +2^{|F-F'|}\right]\right)$ \\ \end{tabular} } \end{table} Various approximate inference methods are proposed in place of Quickscore when processing expensive inference cases in NOBN (particularly \texttt{QMR-DT}). Variational inference for NOBN developed in \citep{jj99} reduces the cost in computing $P(F)$ by applying variational transformation to a subset of $F'\subset F$. The variational evidence is incorporated as posterior probability when performing quickscore on the remaining findings. The running time is then \small$\widetilde{O}\left(|D|\cdot \left[|F'| +2^{|F-F'|}\right]\right)$\normalsize. Other general approximation methods that can be applied to NOBN include loopy belief propagation~\citep{loopy}, mean field approximation~\citep{ng00}, and importance sampling based sampling methods~\citep{uai10}. Some have also considered processing each finding in $F$ sequentially~\citep{pami13}, which is arguably more similar to the style of a realistic patient-to-doctor diagnosis. \section{Experiments} We evaluate the proposed inference algorithms on a real-world symptom-disease NOBN called \texttt{F120}. \texttt{F120} is a QMR-like medical NOBN constructed from multiple reliable medical knowledge sources and is amended by medical experts. Unlike \texttt{QMR-DT}, \texttt{F120} focuses on symptoms and diseases related to maternal and infant care. Due to the anonymous submission, the authors refrain from discussing \texttt{F120}'s details other than listing its vital statistics in Table \ref{tab:data}. Due to the unavailability of the proprietary \texttt{QMR-DT} network~\citep{uai06}, an anonymized version (\texttt{aQMR}) is available~\citep{uai13}. However, \texttt{aQMR} anonymizes the symptom and disease node names and randomizes \texttt{QMR-DT}'s $P(f^+\mid d^+)$ probabilities. With the medical connotation removed, it is difficult to confidently generate user queries (a user query is a tuple $\langle F^+ , F^- , d_l\rangle$, where $d_l$ is the label disease: the most likely disease given the symptoms according to medical experts). Previous works working with \texttt{aQMR} do not face this issue since they do not require use-cases. For example, \citep{uai13,nips13} focus on recovering the network structure and parameters; \citep{uai10} focuses on the inference time and the \emph{relative} divergence between approximate inference outcome and the exact inference outcome. We also evaluate the algorithm's scalability on the artificially generated \texttt{S1} that is much larger in scale than \texttt{F120} and \texttt{QMR-DT}. \texttt{S1} has 40,000 hidden disease nodes, which is approximately the total number of diseases in ICD-10 classification. Figure \ref{fig:speed} compares various inference algorithms against the baseline in \citep{jj99} (JJ99+CVX). The proposed variational-first hybridization (VFH) is consistently faster than other methods. Despite having the same variational cost as VFH (shown in Table \ref{tab:variational}), Joint hybridization (JH) is the slowest due to its repeated negative evidence computation of Equation \ref{e11}. JJ99+PPF is significantly faster than JJ99+CVX due to the simplified $\Xi^{\min}$ estimation. Figure \ref{fig:accuracy} compares the inference accuracies on \texttt{F120}. To simulate the wide-ranged inaccuracy in the disease priors $P(d^+)$'s, we scramble them with samples drawn from the uniform mean (Bates) distribution $\mathcal{P}$ at different ${1\over 1+p}$ values. In total, we test four sets of queries with different kinds of false positive findings. Each query in the 1st set (\texttt{random20}) contains 20\% random $f^+$'s that are not caused by the labeled disease. For the 2nd set (\texttt{chronic20}), the 20\% false $f^+$'s are symptoms caused by some common chronic diseases (e.g., asthma, hypertension). Chronic symptoms are often mentioned inadvertently by patients during doctor's visit and making the diagnosis harder. The 3rd set (\texttt{chronic40}) has the same type of false $f^+$'s as \texttt{chronic20} but the ratio is 40\% of $F^+$. For the 4th set (\texttt{confuse20}), the 20\% false $f^+$'s are symptoms caused by diseases similar to the labeled disease (e.g., influenza and common cold). Such diseases share several symptoms, but often the severity and other key symptoms are decisive in telling them apart. Each of the four sets has 800 queries and each query consists of on average eight $f^+$'s and four $f^-$'s. Shown in Figure \ref{fig:accuracy}, VFH+CVX+FDO performs better than the JJ99+CVX+GDO baseline across the wide range of $P(d^+)$ and even outperforms the exact Quickscore for certain $P(d^+)$ values. VFH+PPF+FDO suffers from its suboptimal (although fast, shown in Figure \ref{fig:speed}) $\xi^{\min}$ estimations. VFH+PPF+FDO is comparable to JJ99+CVX+JJ99 at the lower range of $P(d^+)$ values. Lastly, JH+CVX+FDO has the closest performance portfolio to that of Quickscore and is quite competitive. \begin{figure}[ht!] \begin{floatrow} \capbtabbox{% \setlength{\extrarowheight}{8pt}% \begin{tabular}{ l | l | l | l | l } \textbf{NOBN} & $|D|$ & $|S|$ & \tiny$|\{P(f^+\mid d^+) >0\}|$\normalsize & \textbf{Density}\\ \hline \texttt{F120} & 665 & 1,276 & 10,552 & 1.24\%\\ \texttt{QMR-DT} & 534 & 4,040 & 40,740 & 1.89\%\\ \texttt{S1} &40,000 & 12,000 & 384 million & 80.0\%\\ \end{tabular} }{% \captionsetup{position=above} \caption{Comparisons of NOBNs on the network size and density (measured as total number of nonzero $P(f^+\mid d^+)$ as a percentage of $|D|\cdot |F|$).}\label{tab:data}% }\hspace{-1.2cm} \ffigbox{% \includegraphics[width=4.6cm]{timing1.pdf} }{% \captionsetup{justification=centering} \caption{Runtime comparisons \\ of different algorithms \\ on the \texttt{S1} network.}\label{fig:speed}% } \end{floatrow} \end{figure} \begin{figure}[ht!] \centering \subfigure[]{ \includegraphics[width=3.65cm]{random20top1.pdf} }\hspace{-0.5cm}% \subfigure[]{ \includegraphics[width=3.65cm]{chronic20top1.pdf} }\hspace{-0.5cm}% \subfigure[]{ \includegraphics[width=3.65cm]{semantic20top1.pdf} }\hspace{-0.5cm}% \subfigure[]{ \includegraphics[width=3.65cm]{semantic40top1.pdf} }\\[-0.4cm] \subfigure[]{ \includegraphics[width=3.65cm]{random20top3.pdf} }\hspace{-0.5cm}% \subfigure[]{ \includegraphics[width=3.65cm]{chronic20top3.pdf} }\hspace{-0.5cm}% \subfigure[]{ \includegraphics[width=3.65cm]{semantic20top3.pdf} }\hspace{-0.5cm}% \subfigure[]{ \includegraphics[width=3.65cm]{semantic40top3.pdf} }\\[-0.1cm] \includegraphics[width=12cm]{legend.pdf} \vspace{-0.3cm} \caption{ Accuracy comparisons on \texttt{F120}. $x$-axis is the mean $P(d^+)$ value (i.e., ${1\over 1+p}$); $y$-axis is top-1/top-3 accuracy. All configurations (except Quickscore) transforms 2 findings variationally. \newline (a, e) \texttt{random20}. (b, f) \texttt{chronic20}. (c, g) \texttt{chronic40}. (d, h) \texttt{confuse20}. \newline (a, b, c, d) measure top-1 accuracies. (e, f, g, h) measure the corresponding top-3 accuracies.} \label{fig:accuracy} \end{figure} \section{Conclusions and future work} In this work, we study the important problem of approximate inference on noisy-or Bayesian networks (specifically, their medical applications). We introduce novel algorithms for variational hybridization and variational transformation. The proposed algorithms greatly immunize the current variational inference algorithms against the inaccuracies in widely-ranged hidden prior probabilities, a common issue that arises in modern medical applications of Bayesian networks. In the future, we plan to investigate the applicability of the proposed algorithms to more general Bayesian networks. \small \bibliography{nips_2016} \end{document} }}
\caption{Position resolution multiplied by pressure value as a function of the photon energy for: a) an infinite volume; b) 10x10x1\,cm\textsuperscript{3} detector. All data points, except (open circles \textcolor[rgb]{0.0,0.5176471,0.81960785}{O}), were obtained at 1 bar.}
\caption{(Color online) (a) 1D- and (b) 3D-like phase diagrams for $B = 940\,$G. \rdown{} (\textcolor{userblue}{\scriptsize$ \blacktriangledown$}) and \rdiff{} ($\textcolor{userred}{\bullet}$) are scaled by $N^{1/2}l_{z}$~\cite{Orso07, Liao10}, where $l_{z}=\sqrt{\hbar/m \omega_{z}}$ is the axial harmonic oscillator length and $N=N_{\uparrow} + N_{\downarrow}$. The colored regions correspond to the indicated phases. In (b), the open circle indicates the measured \pcthree{} from (d). The dotted line is an extrapolation from \pcthree{}. (c, d) The local central polarization \po{} vs. \ptube{}, used to find \pcthree{}. The insets show the central region near \pcthree{}. The solid red line is a fit to the data to find \pcthree{}, using a function with a bilinear slope~\cite{Olsen15}. The green vertical arrow indicates \pcthree{}. Each data point is the average of ${\sim}10$ experimental realizations, binned with width $\Delta \ptube{}=0.005$.}
\caption{Distribution of Phase-A sources. Among 2216 target sources (\textcolor{blue}{blue}), 1903 sources (\textcolor{green}{green}) have been observed, 363 sources (\textcolor{red}{red}) were detected in 3 or more observations.}
\caption{Distribution of Phase-B sources. Among 4802 target sources (\textcolor{blue}{blue}), 324 sources (\textcolor{green}{green}) have been observed, 72 sources (\textcolor{red}{red}) were detected in 3 or more observations.}
\caption{SEM micrograph of the Neofer\,\textregistered$ $ 25/60p filament, NdFeB spheres inside the PA11 matrix.}
\caption{Best empirically found printer and slicer parameter for Neofer\,\textregistered$ $ 25/60p.}
\caption{(a) Geometry of the permanent magnet, and area scan of $\vec{B}$ with a step size of 0.1\,mm in the middle of the printed magnet. (b) Line scan 2.5\,mm over pyramid tip (T) compared with FEMME simulation of the magnet with perfect shape.}
\caption{Printed isotropic NdFeB magnet with Neofer\,\textregistered$ $ 25/60p with optimized shape to suppress $B_x$ and $B_y$ along the x-axis $r_x$.}
\caption{Summery of the datasheet and the measured Neofer\,\textregistered$ $ 25/60p material properties.}
\caption{ (a) The disorder averaged wavepacket at different times for the Dyson II model. The central core decays quickly and saturates after initial dynamics; whereas the tail of the distribution keeps spreading with time. Inset shows the return probability for $L=4097$. (b) The growth of $\overline{\langle X^2(t)\rangle}$ with time for $L=\{513, \ldots, 4097\}$ in log-log scale. For finite systems it saturates to a value which grows linearly with the system size. Inset shows $\overline{\langle X^2(t)\rangle}_{\Delta E}$ with $\ez$ present(\pentagon) in $\Delta E$ which grows subdiffusively and absent ($\lozenge$) which saturates, hence confirming that the dynamics is governed by the states close to $E=0$~($L=4097$). (c) The entanglement entropy shows a logarithmic growth in time $\overline{\msr{S}(t)} \sim \log t$ and the saturation, $\overline{\msr{S}_\infty}$, grows logarithmically with $L$ as shown in the inset.}
\caption{Illustrating (a) shallow fusion and (b) deep fusion. The {\color{lightblue}light blue}, white, {\color{pastelviolet}violet}, {\color{lightgreen}light green} boxes correspond to the input, a block of layers (subnetwork), the fusion layer, and the classification layer. (Best viewed in color.)}
\caption{\label{attention-analysis} Examples of bilingual phrases from our translation model with both {\it phrase structures} and {\it attention visualization}. For each example, important words are highlighted in \textcolor{zb_dred}{dark red} (with the highest attention weight), \textcolor{zb_red}{red} (the second highest), \textcolor{zb_lred}{light red} (the third highest) according to their attention weights. {\it Good} = good translation pair, {\it Bad} = bad translation pair, judged according to their semantic similarity scores.}
\caption{Color image of our cluster based on Gemini i'band image (R), {\it Spitzer} {\color{blue} IRAC1} (G) and {\it Spitzer} MIPS {\color{blue} 24$\mu$m} (B). White circles represent the galaxies within the cluster redshift range used in this work, and the large yellow circle correspond to the virial radius of 1.05 Mpc. The magenta circles represent the galaxies at the cluster redshift range in \citet{PintosCastro13}. We have 11 out of 18 galaxies in common.}
\caption{\small Exemplar video summaries by \textsf{MLP-Shot} and \textsf{dppLSTM}, along with the ground-truth importance ({\color{blue}blue} background). See texts for details. We index videos as in \cite{Song2015TVSum}.}
\caption{The evolution of energy loss-gain ratio $\mathcal{R}_{lg}$ under different initial energies, pitch-angles, and radial positions. Subfigures (a) and (b) show the evolution of $\mathcal{R}_{lg}$ for $K_{0}=2.5\,\mathrm{MeV}$ and $12.5\,\mathrm{MeV}$ respectively. The initial pitch-angles in (a) and (b) are chosen as $\theta_{p}=0^{\circ},\,12^{\circ},\,24^{\circ},\,36^{\circ},\,48^{\circ}$. In (c), the initial radial position varies from $0$ to $0.3\,\mathrm{m}$ while the intial momentum is set as $p_{\parallel0}=5\,\mathrm{m_{0}c}$ and $p_{\perp0}=1\,\mathrm{m_{0}c}$. The influence of $r_{0}$ on evolution of $\mathcal{R}_{lg}$ is negligible, and a zoomed-in window showing the details of these curves is embeded in (c) to give a detailed presentation. The dashed curves in all subfigures are reference lines determined by $\mathcal{R}_{lg}^{m}\left(t\right)$, which are calculated using the initial pitch angle $\theta_{p}=0^{\circ}$ in subfigures (a) and (b) and using the initial radial position $r_{0}=0$ in (c). \label{fig:Rlg} } \end{figure} \subsection{Energy balance time} We now focus on how long it takes for a runaway electron to reach its energy limit, i.e., the energy balance time $t_{blc}$. To calculate $t_{blc}$, the start point and end point of acceleration process should be determined. The end point is defined as the moment when the loop electric field is balanced by the synchrotron radiation. At this time, because of the neoclassical pitch-angle scattering, physical quantities, such as momentum, electric acceleration power, and radiation power, show strong oscillations in the gyro-period timescale. The electric field is balanced out by the radiation loss only in the sense of long-term average. Therefore, we define the average loss-gain power ratio as \begin{equation} \eta_{RE}=\left\langle \frac{P_{R}}{P_{E}}\right\rangle _{trans}\,,\label{eq:etaRE} \end{equation} where $P_{R}=\left|\mathbf{F}_{R}\cdot\mathbf{v}\right|$ is the radiation power, $P_{E}=\mathbf{E}\cdot\mathbf{v}$ is the electric acceleration power and the bracket $\left\langle \cdots\right\rangle _{trans}$ means the average over a transit period. Then it is convenient to define the end point of $t_{blc}$ as the moment when $\eta_{RE}=1$. The end point of $t_{blc}$ can also be inferred from the relative behavior of $\mathcal{R}_{lg}$ and $\mathcal{R}_{lg}^{m}$. Because the evolutionary trend of $\mathcal{R}_{lg}$ becomes the same as that of $\mathcal{R}_{lg}^{m}$ after reaching energy limit, the moment when the curves of $\mathcal{R}_{lg}$ and $\mathcal{R}_{lg}^{m}$ begin to overlap also indicates the finish of acceleration. The settlement of the start point of $t_{blc}$, however, is more complex since runaway electrons are born with different phase-space states and origins. There is even not a clear criterion for the emergence of a single runaway electron because of its statistical essence. Fortunately, it can be verified that different\textcolor{black}{{} initial energies matters little to th}e energy balance time for runaways under several MeVs. The typical electric field can accelerate a low-energy runaway electron to several MeVs within 10\% $t_{blc}$ \cite{CollisionlessScater_NF_Letter_2016}. In this paper, we set the start-up energy of a runaway electron a\textcolor{black}{s $2.1\,\mathrm{MeV}$}. The arbitrariness of the setup of the start-up energy has small impact on the value of $t_{blc}$ within the range of several MeVs. According to Fig.\,\ref{fig:Rlg}, it can also be observed that different initial samplings in phase space has little effect on $t_{blc}$. \begin{figure} \includegraphics[scale=0.45]{TypicalEtaRE} \caption{Typical evolution curves of (a) electric acceleration power $P_{E}$, (b) radiation power $P_{R}$, and (c) the average loss-gain power ratio $\eta_{RE}$. The green bands denote the full-orbit evolution curves including fine timescale oscilations. The red curves denote the results after averaging over one transit period.\label{fig:TypicalEtaRE}} \end{figure} The typical evolutions of $P_{E}$, $P_{R}$, and $\eta_{RE}$ are plotted in Fig.\,\ref{fig:TypicalEtaRE}. All of these curves show strong oscillations. The transit-period average values are plotted using the red curves. The power of electric acceleration increases at the beginning because of the growth of runaway velocity. Once the runaway speed is close enough to $\mathrm{c}$, the electric acceleration pow\textcolor{black}{er is dominated by t}he strength of the loop electric field. Since the electric field is inversely proportional to radial position, $P_{E}$\textcolor{black}{{} decreases accompanied by }the outward drift of runway transit orbits, see Fig.\,\ref{fig:TypicalEtaRE}a. The radiation power monotonously increases with the runaway energy\textcolor{black}{{} accompanied by} stronger oscillations, see Fig.\,\ref{fig:TypicalEtaRE}b. The red curve in Fig.\,\ref{fig:TypicalEtaRE}c is a typical evolution of $\eta_{RE}$, which reflects that the loss-gain rate power ratio climbs steeply in the midterm of runaway acceleration. The energy balance time is around $3\,\mathrm{s}$ in this case. \section{Influences of tokamak Device Parameters\label{sec:Effects-of-Tokamak}} Many experiments have confirmed the existence of runaway electrons with energies ranging from $10$-$100\,\mathrm{MeVs}$ in tokamak devices\cite{Gill_JET_REenergy_1993,Jarvis_JET_RE_1989,Wongrach_TEXTOR_disruption_2014}.\textcolor{black}{{} To}\textcolor{red}{{} }\textcolor{black}{describe }and further understand the experimental results, it is necessary to study the dependence of runaway dynamical properties on the device parameters. In this section several \textcolor{black}{characteristics o}f runaways, which may be experimentally diagnosed directly or indirectly, such as the maximum energy $E_{max}$, the energy balance time $t_{blc}$, the oscillation amplitude of perpendicular momentum near the energy limit\textcolor{black}{{} $A_{sp}$, and the average maximum perpendicular momentum $p_{\perp max}$, are discussed.} The influences from three key tokamak device parameters are considered, including the strength of tokamak field, the major radius, and the safety fac\textcolor{black}{tor. The impact of magnetic field ripple is also studied through full-orbit simulations.} As vital design parameters, the intensities of equilibrium tokamak fields are reflected in $E_{l}$ and $B_{0}$. The toroidal curvature of fields is determined by the major radius $R_{0}$, while the poloidal curvature is reflected in the safety factor $q$. All these parameters influence the energy limit and collisionless pitch-angle scattering by stepping in different aspects of runaway dynamics, such as the acceleration, the synchrotron radiation, and the change rate of magnetic field during each gyro-period. Larger $E_{l}$ leads to stronger acceleration, while increasing $B_{0}$ results in the mitigation of collisionless pitch-angle scattering. The increase of magnetic curvature corresponds to the enhancement of radiation and toroidal effect. Considering the loop electric field $\mathbf{E}$ and toroidal magnetic $\mathbf{B}$ decreases radially, the neoclassical drift velocity, approximately given by $qE_{l}/B_{0}$ \cite{Guan_Qin_Sympletic_RE,Neoclassical_Drift_report}, also interferes the energy limit rule and the neoclassical scattering process. The change of one single device parameter thus may affect the runaway dynamics in several interactional mechanisms. On the other hand, the magnetic field ripple can also impose stochastic instability to runaway dynamics through the nonlinear resonance, which restricts the maximum runaway energy below the synch\textcolor{black}{rotron limit \cite{laurent_Rax_MagneticRipple1990}. I}n this section, the initial conditions are sampled in the phase space the same as in Sec.\,\ref{sec:Dynamics-in-Gyro-period}. \subsection{Influences of field strength } \begin{figure} \includegraphics{EB_Emax_tlbc_Azp_pperpmax} \caption{The plots of (a) the maximum energy of a runaway electron, (b) the energy balance time, (c) the oscillation amplitude at energy limit, and (d) the maximun perpendicular momentum against loop electric field with different magnetic field strengths. The initial condition is set to $p_{\parallel0}=5\,\mathrm{m_{0}c}$, $p_{\perp0}=1\,\mathrm{m_{0}c}$, $R=1.8\,\mathrm{m}$, and $\xi=z=0$. The major radius of tokamak is $R_{0}=1.7\,\mathrm{m}$, and the safety factor is $q=2$.\label{fig:EB_Emax_Tblc_Azp_pperpmax}} \end{figure} Figure \ref{fig:EB_Emax_Tblc_Azp_pperpmax} summarizes the influences from $E_{l}$ and $B_{0}$ on $E_{max}$, $t_{blc}$, $A_{sp}$, and $p_{\perp max}$. Curves with different colors correspond to different central magnetic field strength. The major radius is set as $1.7\,\mathrm{m}$ and the safety factor is $2$. In tokamak experiments, the loop electric field has the strongest impact on the runaway energy limit. During disruptions the energy of plasma is released through a strong inductive loop electric field. According to Figs.\,\ref{fig:EB_Emax_Tblc_Azp_pperpmax}a and \ref{fig:EB_Emax_Tblc_Azp_pperpmax}b, the runaways can reach higher energy limits in shorter time when the loop electric field increases, thus posing more severe threat in major disruption. On the other hand, the energy limit also depends on the magnetic field significantly. From Fig.\,\ref{fig:EB_Emax_Tblc_Azp_pperpmax}a, we can see that the energy limit is higher for smaller $B_{0}$. It can be noticed that the increment of $E_{max}$ due to the drop of $B_{0}$ is smaller than the result in Ref.\,\cite{CollisionlessScater_NF_Letter_2016} which assumes a uniformly distributed electric field in the radial direction. The difference is caused by the neoclassical drift and radial distribution of electric field. For smaller $B_{0}$, as the neoclassical drift is faster, the electric field witnessed by runaway electrons decreases faster. Therefore, the energy limit is reduced by several MeVs compared with that in the uniform electric field distribution. This mechanism also results in longer balance time, see Fig.\,\ref{fig:EB_Emax_Tblc_Azp_pperpmax}b. The strength of tokamak fields has small effects on the oscillation amplitude in small timescale, see Fig.\,\ref{fig:EB_Emax_Tblc_Azp_pperpmax}c. The largest relative change of oscillation amplitude caused by tokamak field\textcolor{black}{{} is about 1}0\%. However, according to Fig.\,\ref{fig:EB_Emax_Tblc_Azp_pperpmax}d, the ramp-up of perpendicular momentum in large timescale can be altered significantly by adjusting the strength of tokamak field. The growth rate of $p_{\perp max}$ versus loop electric field becomes much larger for smaller magnetic field. This embodies stronger accumulation of neoclassical pitch-angle scattering on perpendicular runaway momentum, also more violent deviation from gyro-center model, for smaller magnetic field and large electric field. \subsection{Influences of Major Radius} Next-generation tokamak devices possess larger major radius to achieve higher operation parameters. For example, the major radius of ITER is designed to be $6.2\,\mathrm{m}$. Larger tokamaks have better confinement on fusion plasma as well as runaway electrons. More energy may release through runaway currents during major disruptions. It is obvious that runaway electron can gain more energy from the stronger electric field. On the other hand, we will show that, even with the same strength of electric field, the change of major radius $R_{0}$ will influence the energy limit rule and the neoclassical pitch-angle scattering directly. In this part, the tokamak field is set as $E_{l}=0.2\,\mathrm{V/m}$ and $B_{0}=2\,\mathrm{T}$, and the safety factor is 2. \begin{figure} \includegraphics{R0_Emax_tblc} \caption{The plots of energy limit $E_{max}$, denoted by the blue solid curve and the left ordinate, and the energy balance time $t_{blc}$, denoted by the red dashed curve and the right ordinate, against the major radius. The initial condition of runaway electron in phase space is sampled as $p_{\parallel0}=5\,\mathrm{m_{0}c}$, $p_{\perp0}=1\,\mathrm{m_{0}c}$, $R=1.8\,\mathrm{m}$, and $\xi=z=0$. The tokamak field is set as $E_{l}=0.2\,\mathrm{V/m}$ and $B_{0}=2\,\mathrm{T}$, and the safety factor is chosen as $q=2$.\label{fig:Emaxtblc_R}} \end{figure} \begin{figure} \includegraphics{Azp_Pperpmax_R} \caption{The plots of magnitude of oscillation at energy limit $A_{zp}$, denoted by the blue solid curve and the left ordinate, and the maximum perpendicular momentum $p_{\perp max}$ , denoted by the red dashed curve and the right ordinate, against the major radius. The initial condition of runaway electron in phase space is sampled as $p_{\parallel0}=5\,\mathrm{m_{0}c}$, $p_{\perp0}=1\,\mathrm{m_{0}c}$, $R=1.8\,\mathrm{m}$, and $\xi=z=0$. The tokamak field is set as $E_{l}=0.2\,\mathrm{V/m}$ and $B_{0}=2\,\mathrm{T}$, and the safety factor is $q=2$.\label{fig:R_Azp_Pperpmax}} \end{figure} Figure \ref{fig:Emaxtblc_R} plots the energy limit and the balance time against different major radius. We can see that both the maximum energy and balance time increase proportional to the major radius $R_{0}$ approximately, because the synchrotron radiation closely depends on the curvature of the runaway orbit \cite{Martin_Momentum_RE_1998}. Smaller $R_{0}$ brings larger toroidal curvature and thus stronger radiation power. The runaway electrons have stronger synchrotron dissipation in small devices. As a result, their energy limit is lower and energy balance time $t_{blc}$ is shorter. The curvature of tokamak field reflects the significance of toroidal geometry. So the major radius also affects the collisionless pitch-angle scattering evidently. For devices with smaller major radius and larger toroidal curvature, the same distance traveled in toroidal direction brings more variation of the magnetic field direction. Consequently, the assumption of gyro-center model breaks down easier in smaller devices. As expected, the magnitude of oscillation at energy limit $A_{zp}$ and the maximum perpendicular momentum $p_{\perp max}$ drop with the increase of $R_{0}$, see Fig.\,\ref{fig:R_Azp_Pperpmax}. \subsection{Influences of safety factor} The safety factor $q$ is another important parameter of tokamaks, which reflects the geometric character of magnetic surface. Smaller $q$ corresponds to stronger poloidal magnetic field and more poloidal periods the magnetic line winding during each toroidal cycle. The curvature of magnetic line is determined toroidally by the major radius and poloidally by $q$. Therefore, dynamical processes related to geometry configurations, such as synchrotron radiation and the neoclassical pitch-angle scattering, will be influenced by $q$. \textcolor{black}{Larger poloidal field also brings more difficult for electric field to accelerate the runaway electrons toroidally. Meanwhile, the value of $q$ also influence the neoclassical drift and thus the change of the local strength of electric field. Generally speaking, the safety factor $q$ has compound impacts on runaway dynamics, which makes its consequences vague by analyzing any individual factor. During disruptions, large portion of the poloidal magnetic field is induced by the runaway current. So the $q$ profile in major disruption involves self-consistent evolution of runway electrons.} In this part, we use the parameters $E_{l}=0.2\,\mathrm{V/m}$, $B_{0}=2\,\mathrm{T}$, and $R_{0}=1.7\,\mathrm{m}$, while the value of $q$ is sampled from 0.2 to 10. \begin{figure} \includegraphics{q_Emax_Tblc} \caption{The plots of energy limit $E_{max}$, denoted by the blue solid curve and the left ordinate, and the energy balance time $t_{blc}$, denoted by the red dashed curve and the right ordinate, against safety factor. The initial condition of runaway electron in phase space is sampled as $p_{\parallel0}=5\,\mathrm{m_{0}c}$, $p_{\perp0}=1\,\mathrm{m_{0}c}$, $R=1.8\,\mathrm{m}$, and $\xi=z=0$. The tokamak field is set to $E_{l}=0.2\,\mathrm{V/m}$ and $B_{0}=2\,\mathrm{T}$, and the major radius $R_{0}=1.7\,\mathrm{m}$.\label{fig:q_Emax_Tblc}} \end{figure} Figure \ref{fig:q_Emax_Tblc} plots the energy limit and the energy balance time against the safety factor $q$. Both $E_{max}$ and $t_{blc}$ increases as $q$ becomes larger. This phenomenon comes from two main reasons. Firstly, for smaller $q$, the poloidal field is stronger. Therefore, the toroidal acceleration of electric field is hindered. Secondly, when the toroidal curvature determined by the major radius keeps unchanged, smaller $q$ corresponds to larger poloidal curvature and thus stronger synchrotron radiation. Even though the neoclassical drift velocity is proportional to $q$ and the electric field decreases faster for larger $q$ due to its radial distribution, the numerical results in Fig.\,\ref{fig:q_Emax_Tblc} imply that the effect of safety factor by modifying the neoclassical drift is weaker than the above two effects. \begin{figure} \includegraphics{q_Af_Pperpmax} \caption{The plots of magnitude of oscillation at energy limit $A_{zp}$, denoted by the blue solid curve and the left ordinate, and the maximum perpendicular momentum $p_{\perp max}$ , denoted by the red dashed curve and the right ordinate, against safety factor. The initial condition of runaway electron in phase space is sampled as $p_{\parallel0}=5\,\mathrm{m_{0}c}$, $p_{\perp0}=1\,\mathrm{m_{0}c}$, $R=1.8\,\mathrm{m}$, and $\xi=z=0$. The tokamak field is set to $E_{l}=0.2\,\mathrm{V/m}$ and $B_{0}=2\,\mathrm{T}$, and the major radius is $R_{0}=1.7\,\mathrm{m}$.\label{fig:q_Asp_pperpmax}} \end{figure} The influences of $q$ on collisionless pitch-angle scattering and maximum perpendicular momentum are plotted in Fig.\,\ref{fig:q_Asp_pperpmax}. When the safety factor is less than 2, the amplitude of perpendicular momentum oscillation drops evidently with the increase of $q$. Under this condition, the neoclassical scattering is extremely strong, and it is much easier for an energetic runaway electron to move cross magnetic surfaces. The rotation of magnetic field witnessed by runaway electrons becomes fast at the same time. The oscillation amplitude even exceeds the maximum perpendicular momentum at energy limit when $q$ is small enough. When $q$ is larger than 2, the geometric effects mainly come from the toroidal field. The dependence of $A_{zp}$ on $q$ is not notable. On the other hand, the plot of $p_{\perp max}$ shows the similar trend to that of $E_{max}$, which increases monotonously with $q$. \subsection{Influence of magnetic field ripples on energy limit} In tokamaks, the toroidal magnetic fields are induced by toroidal coils with finite coil number, which leads to magnetic field ripples in experiments. According to L. Laurent and J. M. Rax's paper in 1990, the stochastic instability caused by the nonlinear cyclotron resonances with magnetic field ripples can transfer the parallel energy to perpendicular direction and thus restrict runaway energy limit far below the synchrotron energy limit \textcolor{black}{\cite{laurent_Rax_MagneticRipple1990}}. In this subsection, utilizing full-orbit simulations, we study the impacts of magnetic field ripples on the runaway energy limit. Besides the equilibrium electromagnetic field given by Eqs\textcolor{black}{.\,\ref{eq:B} and \ref{eq:E}}, there also exist the radial perturbation of magnetic ripple \textcolor{black}{$\delta\mathbf{B}$,} expressed by \begin{equation} \delta\mathbf{B}=\delta B\mathbf{e}_{r}\,,\label{eq:Bripple} \end{equation} \begin{equation} \delta B\left(r,\theta,\varphi\right)=\sum_{m=0,n=1}^{m=\infty,n=\infty}\delta B_{mn}\left(r\right)\mathrm{cos}\left(m\theta\right)\mathrm{cos}\left(nN\varphi\right)\,,\label{eq:MrippleInfty} \end{equation} \textcolor{black}{where, $r$, $\theta$, and $\varphi$ are three components of the toroidal coordinates, $N$ is the number of toroidal field coils, and $m$ and $n$ denote respectively the toroidal and poloidal harmonics. Following the discussion in Ref.\,\cite{laurent_Rax_MagneticRipple1990}, we consider only the terms with $m=0,\,1$ in Eq. \ref{eq:MrippleInfty}. The amplitude of perturbation magnetic field is given by the analytical approximation for small $m$, namely, \begin{equation} \delta B_{0n}\left(r\right)=\delta B_{1n}\left(r\right)\approx\frac{B_{0}}{2}\left(1+\frac{qR_{0}}{R_{0}+b}\right)\mathrm{exp}\left[-Nn\left(\frac{b-r}{b+R_{0}}\right)\right]\,,\label{eq:DeltaB1n} \end{equation} where, $b$ is the radius of toroidal coils. Based on the Tore Supra tokamak \cite{laurent_Rax_MagneticRipple1990}, we set the simulation parameters as $N=18$, $B_{0}=1.8\,\mathrm{T}$, $E_{l}=0.1\,\mathrm{V/m}$, $q=2$, $R_{0}=2.4\,\mathrm{m}$, $a=0.75\,\mathrm{m}$, and $b=1.3\,\mathrm{m}$. The initial condition of the runaway electron is given by $p_{\parallel0}=5\,\mathrm{m_{0}c}$, $p_{\perp0}=1\,\mathrm{m_{0}c}$, $r_{0}=0.1\,\mathrm{m}$, and $\theta_{0}=\varphi_{0}=0$.} \begin{figure} \includegraphics{MagneticRipple} \caption{The energy evolution of a runaway electron in tokamak fields with different magnetic ripple perturbations. The red curve shows the energy evolution without magnetic field ripple, the green one depicts the result considering the $n=1$ components of $\delta B$, and the blue one is the result considering both $n=1$ and $n=2$ harmonics of $\delta B$. The summation over $m$ only covers two lowest components, namely, $m=0,\,1$.\label{fig:MRipple_Energy}} \end{figure} Figure \ref{fig:MRipple_Energy} depicts the energy evolution of a runaway electron affected by different harmonics of magnetic ripple. Indicated by the red curve in Fig.\,\ref{fig:MRipple_Energy}, without the magnetic field ripple, the synchrotron energy limit is about $80\,\mathrm{MeV}$. When considering the $n=1$ components of ripple field, $\delta B_{01}$ and $\delta B_{11}$, the maximum runaway energy decreases to about $60\,\mathrm{MeV}$ approximately, see the green curve in Fig.\,\ref{fig:MRipple_Energy}. The restriction of the $n=2$ components of ripple field on runaway energy is more significant. As shown by the blue curve, if we add the components of $n=2$, namely, $\delta B_{02}$ and $\delta B_{12}$, the energy limit is reduced to $22\,\mathrm{MeV}$. The results in Fig.\,\ref{fig:MRipple_Energy} exhibit that the magnetic ripples can limit the maximum runaway energy far below the synchrotron limit, which is consistent with the results in Ref.\,\cite{laurent_Rax_MagneticRipple1990}. \section{Conclusions\label{sec:Summary}} In this paper, the multi-timescale runaway dynamics in tokamak field is comprehensively exposed. The physical pictures in different timescales, from $10^{-11}\,\mathrm{s}$ to $3\,\mathrm{s}$, have been fully exhibited. The utilization of the relativistic volume-preserving algorithm is vital to this study, because the long-term numerical accuracy and stability of VPA ensures the accomplishment and correctness of the secular numerical results. In the physical model, the toroidal configuration of tokamak field and the synchrotron radiation are considered. Correspondingly, the key role of geometric effects and coupling of multi-timescale runaway dynamical processes are perfectly captured. In small timescale imposed by Lorentz force, unlike the common wisdom, the helical trajectory of energetic runaway electrons is elongated both toroidally and poloidally so much that the collisionless neoclassical scattering rises. A theoretical description of the neoclassical scattering is provided through the coupling between the rotations of magnetic field and momentum. The drift in momentum space is also analyzed based on the rotation vector of magnetic field. The micro timescale dynamics discussed in this paper has established a comprehensive picture of runaway motion. More importantly, our results have shown that the coupling between $\mathcal{T}_{c}$ and transit period plays an important role for energetic runaways. In large timescale up to several seconds, the long-term structure of momentum evolution is portrayed by four characteristic quantities. To find out the secular integral laws, we also studied the energy gain-loss ratio and the energy balance time. The initial condition is proved to have significant effects on small timescale momentum oscillation but little influence on the long-term integral behaviors, such as energy limit and energy balance time. \textcolor{black}{Meanwhile, the dynamics of runaways can also be impacted by tokamak parameters. The electromagnetic field, major radius, and safety factor have different influences on both the energy limit and the neoclassical scattering process through altering different aspects of runaway dynamics. It is also proved that the existence of magnetic field ripple can reduce the maximal runaway energy.} Considering the complex influences from many different physical processes in real tokamak discharges, we will study other factors on the energy limit of runaway electrons, such as different instabilities and resonance magnetic perturbations, in the future dynamical analysis of runaway electrons. \textcolor{black}{Meanwhile, the observed runaway effects in experiments are generally collective behaviors of large amounts of runaway electrons. Therefore, the statistical treatment of runaway }evolution with large samplings in the phase space will be carried out to obtain macroscopic results, which is convenient for experimental observation and verification. \begin{acknowledgments} This research is supported by National Magnetic Confinement Fusion Energy Research Project (2015GB111003, 2014GB124005), National Natural Science Foundation of China (NSFC-11575185, 11575186, 11305171), JSPS-NRF-NSFC A3 Foresight Program (NSFC-11261140328), and the GeoAlgorithmic Plasma Simulator (GAPS) Project. \end{acknowledgments} \bibliographystyle{apsrev} \bibliography{Refs_RunawayElectrons} \end{document} }
\caption{\label{table_group operators}List of useful quantum states and corresponding operators that may be used to implement protocols for AQD and QD. Specifically, to implement a protocol of QD using a state mentioned in the 1st column of $i$th row of this table, both the users should use one of the groups listed in Column 3 or 4 of the same row. However, to implement an AQD, using a state mentioned in the 1st column of $i$th row of this table, one of the user (say Bob) would use a group ($G_{B}$) listed in Column 3 or 4 of the same row, as was done in QD, and Alice (who is expected to communicate less) would use a\textcolor{red}{{} }smaller group $G_{A}:G_{A}\otimes I_{2}^{j-l}<G_{B},$ where Bob and Alice encode their message by using $j$ and $l$ qubit unitary operations, respectively.}
\caption{Network Alignments with Three Network Models. The highest correlation is in bold font. The results of graphlet degree distribution, RGF distance and path difference are generated by the GraphCrunch tool described in the work of \cite{graphlet} \cite{graphcrunch} and \cite{stickymodel}. The description of these metrics can be found in \textcolor{blue}{Supplementary Information 4}. In this table we can see in both datasets {\bf IAM} achieves better GDD-agreement than {\bf PA} and {\bf WM}. Additionally an interesting observation is, {\bf PA} has better GDD-agreement than {\bf WM} in DBLP while {\bf WM} has higher agreement value than {\bf PA} in Facebook. This further implies our proposition that locality influence in network and popularity influence have different effects on evolution of DBLP and Facebook. Additionally we observe that {\bf IAM} still achieves better performance than {\bf PA} and {\bf WM}.}
\caption{$\mathcal{N}_{\mbox{min}}$ vs ratio between the average shortest path-length of the network of minima for the Thomson Problem and the corresponding random network, $L/L_0$. {\color{red}Font is too small.}}
\caption{$\mathcal{N}_{\mbox{min}}$ vs ratio between the clustering coefficients of the network of minima for the Thomson Problem and the corresponding random network, $C/C_0$. {\color{red}Font is too small.}}
\caption{Performance of {\color{red}task A} and {\color{blue}task B} on {\color{red}VGG16A}, {\color{blue}VGG16B}, and {\color{green}VGG16AB}}
\caption{Performance of {\color{cyan}task C} on VGG16, {\color{red}VGG16A} and {\color{green}VGG16AB}}
\caption[Transitions of double branching annihilating random walkers with heights]{A four particle configuration with heights. A \emp{single arrow} corresponds to a flip ({\color{blue}blue}: from 0 to 1, {\color{red}red}: from 1 to 0). From left to right these can happen with rates $1,1,1,1,1,2p$ and $1$, respectively. Adjacent \emp{opposite arrows} can produce a swap in heights which can happen with rates $b,b,p\cdot b$ and $p\cdot b$, respectively.}
\caption{Phase portrait of the closed-loop system starting from various different initial points \red{using $\nu=20$ and $N=100$}.}
\caption{Average probability vs.~budget by dataset (Student Performance or ARIC) and by bound-setting method. Solid lines represent a result obtained using the logistic model, while dotted lines represent a result obtained using the SVM model. PGD denotes use of the gradient method, while Sens denotes use of the sensitivity analysis-based method. The \textcolor{cyan}{cyan} dashed line is a randomly selected individual whose recommendations will be shown and discussed in the next subsection.\label{fig:avgprobres}}
\caption{\label{fig:openmpi_reduce_vs_allreduce} Verification of \mpireduce \guidelt \mpiallreduce with (a) original and (b) new Reduce implementation (\jupiteropenmpi, \machone, $\nmpiruns=\num{5}$, $\nrep=\num{100}$).}
\caption{{\it Upper panel}: \gray counts map above 2 GeV after point source subtraction in the region Orion A (left) and Taurus (right). {\it Lower panel}: $\tau_{353}$ maps for Orion A (left) and Taurus (right) }
\caption{\label{tab:parameter} Site occupancies and model parameters of each crystallographically inequivalent atom. The spin magnetic moments, $m^s$, are calculated from the first-principles calculation code, Machikaneyama (AkaiKKR).\cite{akaikkrmachikaneyama} The anisotropy parameters $D^{\rm A}_{i_{\rm TM}}$ and $A_l^{\red{m_l}}\langle r^l \rangle $ are taken from previous results.\cite{miura_magnetocrystalline_2014,yamada_crystal-field_1988} We neglected the $D^{\rm A}$ values of B and Nd, as they are less than $0.1\,\rm meV$, and used the $\langle r^l \rangle$ values of Nd, Ref.~\citenum{freeman_theoretical_1962}, i.e., $\langle r^2 \rangle=1.001\,\red{a_B^2}$, $\langle r^4 \rangle=2.401\,\red{a_B^4}$, and $\langle r^6 \rangle=12.396\,\red{a_B^6}$\red{, where $a_B$ is the Bohr radius}. }
\caption{{(a)} Reverse-forward bias scan of a typical pre-poled sample: calculated \cred{J-V} characteristics (red lines) in comparison with experimental data (black points). The forward and reverse data corresponds to increasing and decreasing voltage, respectively. The initial pre-set voltage is $V_0=1.6$ V. {(b) and (c)} show the polarization as a function of applied bias, for both linear an non-linear terms, $P_l$ and $P_{nl}$, respectively. The dotted horizontal line marks \cred{$P_\infty=6.97$ mC/cm$^2$}. }
\caption{Decomposition of the total current in the solar cell under illumination according to Eq.\(\ref{IVeq}). \cred{The most significant currents, $J_{ph}$, $J_d$ and $J_{c}^{(nl)}$, as well as the current in the external circuit $J$ are presented in (a), while the smaller currents $J_{sh}$ and $J_{c}^{(l)}$ are shown in (b) and (c), respectively.} }
\caption{Dynamic hysteresis under different bias scan rates: (a) simulated J-V characteristics, qualitatively similar with our experimental curves and other data as well \cite{tress}; the inset shows simulated J-V characteristics, taking into account a large value for the dielectric constant, $\epsilon_r=10^5$ (dashed lines), as reported in Ref.\\cite{perez}; (b) the corresponding polarization; \cred{(c) the current component $J_{c}^{(nl)}$; (d)} experimental J-V characteristics, measured at different BSRs.}
\caption{Full-disc schematic drawing created from HMIDD data for 2014-07-06T20:00:41UT as it is described in the case of Figure 5. The images of this type are available at {\sf ftp://fenyi.solarobs.unideb.hu:2121/pub/SDO/images/Polarity\_drawings/}.}
\caption{Driving data (Left column) and its distribution (Right column) for two kinds of driving styles. Red line ({\color{red}{--}}) represents the aggressive driver and the black line ({\color{black}{--}}) represents the (normal) typical driver.}
\caption{Excitation diagnostic diagrams, comparing the emission-line intensity ratios. The points color corresponds to different locations along the slit according to the scale-box. The $3\sigma$ error-bars are shown for the points in the main part of the Red Ellipse ($r=-5...+8''$). The top panels present classical BPT diagrams. The dividing lines between the regions ionized by hot stars and shocks/AGN are shown according to \citet{Kewley2001} and \citet{Kauffmann2003}. The bottom panels show diagrams used for separation between SNR and HII regions in nearby galaxies with the dividing borders adopted from \citet{Leonidaki2013}. \blue{The red star marks the lines ratio derived from the total spectrum of the nebulae integrated in the range $r=-5...+8$ arcsec.} }
\caption{(a) Energy barrier that separates a regime of short-distance attraction and long-range repulsion for various state points using the bulk charges from the binary cell model of Fig. \ref{fig:charges} with $\eta_r=\eta_g$. The lines are used to guide the eye. The upper line is for $\kappa^{-1}=6\ \mu\mathrm{m}$, while the lower one {\mycolor(two points at the bottom)} is for a lower ionic strength, $\kappa^{-1}=10\ \mu\mathrm{m}$, and in (b) we plot the corresponding charges of red ($Z_r$) and green ($Z_g$) particles. The full lines are for $\kappa^{-1}=6\ \mu\mathrm{m}$ and the dashed lines are for $\kappa^{-1}=10\ \mu\mathrm{m}$, and the symbols match the state points of (a).}
\caption{Clusters (mostly dumbbells) formed in a suspension of locked PMMA particles (radius $a=1.12$~$\mu$m) in CHB at a distance 0.3 mm from the CHB-water interface, two days after sample preparation. \mycolor{The scale bar indicates 10 $\mu$m.}}
\caption{An overview of training the adversarial evaluator using a hierarchical neural model. {\color{Green} Green} denotes input contexts. {\color{red} Red} denotes a sentence from human-generated texts, treated as a positive example. {\color{Purple} Purple} denotes a sentence from machine-decoded texts, treated as a negative example.}
\caption[]{Log of observations. \textcolor{blue}{}}
\caption{Ratio of the C III] 1909\AA\flux on day 99 over day 248 flux (solid line) and same for the N IV] 1486\AA\(dashed line). Note the strong decrease in the strength of the +600 km s$^{-1}$ feature and the near identity of the variation for the two lines. In both cases, the weaker component of the doublet is negligible.}
\caption{ Harmonic spectra ($\bullet$ odd and {\color{blue!80}$\bullet$} even harmonics) produced in helium in the non-dipole regime for intensities of $3.2\times 10^{14}$ and $\SI{e15}{W/cm^2}$ and monochromatic fields of \SI{800}{nm} and \SI{1.6}{\micro\metre} at a beam half-angle of $\theta=4\si{\degree}$, on arbitrary scales and eliminating $z$-polarized harmonics. The intensity ratio between even and odd harmonics varies from ${\sim}10^{-3}$ for \SI{800}{nm} drivers to ${\sim}10\%$ for strong mid-IR fields at $\SI{1.6}{\micro\metre}$ and $\SI{e15}{W/cm^2}$, which are still accessible with current technology. }
\caption{Change Detection in a pair of images. Our approach uses only image-level labels to detect and localize changed pixels. \textbf{Left:} pair of images. \textbf{Right-most} our change localization results ({\color{blue} blue} denotes a high change region, enclosed in a {\color{red}{red}} box for clarity). Note that the paired images have rich backgrounds with different motion patterns (e.g., fountain in the \emph{third row}) and subtle changes (e.g., small vehicles in the \emph{second} row), which make the detection task very challenging. \vspace{-1.5em}}
\caption{Root-mean-squared fluctuation intensities of the momentum fluxes along the three coordinate directions, computed with the fluctuation velocity equation \r{eq:NS2}. (a) Fluxes of the streamwise momentum. \dashed, $\phi_{11}$; \solid, $\phi_{12}$; \chndot, $\phi_{13}$. % (b) Fluxes of the transverse momenta. \dashed, $\phi_{22}$; \solid, $\phi_{23}$; \chndot, $\phi_{33}$. % Lines with circles are $R'_{ij}$ from \r{eq:NS2}. Those without symbols are optimum fluxes $\phi'_{ij}$ from (\ref{eq:poisF1}--\ref{eq:poisF5}). }
\caption{Higher-order moments of the fluctuation intensities of the centred momentum fluxes, computed from the fluctuation velocities. \dashed, $\phi'_{11}$; \solid, $\phi'_{12}$; \chndot, $\phi'_{22}$. % (a) Third-order skewness, $\bra \phi'^3\ket/ \bra \phi'^2\ket^{3/2}$. % (b) Fourth-order flatness. $\bra \phi'^4\ket/ \bra \phi'^2\ket^{2}$. % The horizontal dashed lines are theoretical values for different functions of gaussian-distributed variables. \dtrian, A gaussian variable $(S_3=0,\,F_4=3)$; \squar, the product of two gaussian variables with cross-correlation coefficient $-0.4$ $(S_3=-2.02,\,F_4=11.9)$. % (c) Pdf of $\phi_{12}$ at $x_2/h\in(0.1-0.2)$, normalised in wall units. The dashed line is the product of two gaussian variables with cross-correlation coefficient $-0.4$. % (d) As in (c), for $\phi_{22}$. The line with triangles is for $u'^2_2+p$. % In all figures, lines with circles are the Reynolds products, $u'_i u'_j$, and those without symbols are optimum fluxes from (\ref{eq:poisF1}--\ref{eq:poisF5}). }
\caption{% (a) Two-dimensional premultiplied spectra as functions of the wall-parallel wavelengths. \solid, $R'_{12}$; \dashed, optimal $\phi'_{12}$; \chndot, cospectrum of $u'_1$ and $u'_2$. The three contours contain 10\%, 50\% and 80\% of the spectral mass. The dashed diagonals are $\lambda_1=\lambda_3$. Flow as in figure \ref{fig:flux3}, $x_2/h=0.15$. % (b) Streamwise premultiplied spectra, as a function of $x_2$. Each horizontal section is a spectrum normalised to unit energy, and the contours are 50\% and 80\% of the global maximum. % \solid, $R'_{12}$; \dashed, $\phi'_{12}$. The dashed diagonal is $\lambda_1=1.2\, x_2$. }
\caption{Example dialog with Skylar and the corresponding dialog tree. \textbf{S:} Start, \textbf{R:} Root, \textbf{DW:} Domain weather, \textbf{RR:} Request root, \textbf{IR:} Inform root, \textbf{ME:} Misunderstanding error, \textbf{CAM:} Cambridge Remote Agent. The \textcolor{red}{red} text is from Skylar and the \textcolor{green}{green} text is from the Cambridge Remote Agent.}
\caption{Spectral extraction regions used for Figure~\ref{fig:4}. The contours correspond to the \gray image (>1000 MeV). The red lines indicate the regions of the radio and \planck aperture photometry. The contours inside the corresponding regions show the template used for the extraction of the corresponding LAT spectrum. N1, N2, and N3 are the outer, middle, and inner regions of the north lobe, and S1, S2, and S3 are the outer, middle, and inner regions of the south lobe. The circle is the core region.}
\caption{Sample detection results on the FDDB dataset, where {\color{green}green} bounding boxes are ground-truth annotations and {\color{red} red} bounding boxes are detection results of the Faster R-CNN.}
\caption{Sample detection results on the IJB-A dataset, where {\color{green}green} bounding boxes are ground-truth annotations and {\color{red} red} bounding boxes are detection results of the Faster R-CNN.}
\caption{\csentence{Rank diversity of sports and games.} Plot showing the rank diversity $d(k)$ for all datasets ({\color{blue} blue} dots), as well as the fit $\Phi$ ({\color{red} red} lines). Values of $\mu$, $\sigma$, maximum rank diversity $d_{max}$, and $R^2$ are also shown. }
\caption{\csentence{Similarity in the normalized rank diversity across sports and games.} Plot showing a comparison of the rank diversity $d(k)$ for all activities considered. With the values of $\mu$ and $\sigma$ obtained from the fit $\Phi$, we have normalised $d(k)$ by rank $k$ with $\frac{\log(k)-\mu}{\sigma}$. As reference we include the basic form of \eref{eq:Phi} (thick {\color{red} red} line), indicating that all activities have the same functional shape of rank diversity. }
\caption{\csentence{Simulated rank diversity.} Plot showing the rank diversity $d(k)$ coming from our random walk model, with $\hat \sigma$ values adjusted from empirical data. We include the values of $\mu$, $\sigma$, maximum rank diversity $d_{max}$, and $R^2$, as well as the basic form of \eref{eq:Phi} (thick {\color{red} red} line). }
\caption{A flow diagram of the two-pathogen SIRWS model. Boxes represent the compartments into which the population partitions, and arrows represent the rates at which individuals transfer between compartments. The force of infection $\lambda_i$ is the transmission coefficient times the sum of those infectious with pathogen $i = 1, 2$. For brevity, the dashed ($\DashedArrow$) and dotted ($\DashedArrow[dotted]$) arrows represent the immunity waning rates $2\kappa_1$ and $2\kappa_2$, respectively. The death rate $\mu$ is represented by a bullet (\protect\tikz \protect\fill[black] (0.4ex,0.4ex) circle (0.5ex);).}
\caption{\textbf{Contact maps for the ``two-state'' model}. In this figure we show contact maps and representative snapshots corresponding to the dynamics of the system with two different values of $\alpha$. As one can notice, while low $\alpha=\epsilon/k_BT_L$ leads to a checker-board contact map at large times, high values of $\alpha$, or deep quenches, freeze the network of contacts. \red{Each point in the contact map is coloured red, blue or black if the entry in the matrix of contacts $C_{ij}$ is between two red beads, two blue beads, or a blue and a read beads, respectively. This is represented in the figure as a colour bar. Since the contact maps correspond to individual snapshots (i.e., they are not averaged over time), each bin is either coloured or empty. One can notice that high values of $\alpha$ (bottom row) leads to rapid folding of the chain and to the appearance of many mixed contacts (black points) which are then slowly lost (in favour of coloured contacts) over time.}}
\caption{\textbf{Contact maps and snapshots for the ``two-state with intermediate state'' model before and after the perturbation.} In this figure we report the evolution of the contact map for the two-state with intermediate state model during the artificial recolouring. \textbf{(a)-(c)} refer to \textbf{(a)} $4$ $10^5$ $\tau_{Br}$ before the perturbation while \textbf{(b)} and \textbf{(c)} show the contact map immediately before and after the recolouring (in \textbf{(c)} (note the small red segment in the middle of the contact map). \textbf{(d)-(f)} refer to \textbf{(d)} $1$ $10^5$ $\tau_{Br}$, \textbf{(e)} $2$ $10^5$ $\tau_{Br}$ and \textbf{(f)} $4$ $10^5$ $\tau_{Br}$ after the perturbation. \red{The contact maps correspond to snapshots in time and each entry is coloured red, blue or black accordingly to whether the contact is red-red, blue-blue or mixed (see colour-bar in figure).} From this figure one can readily appreciate the dramatic change in conformation (phase transition) driven by the localised artificial recolouring. }
\caption{\textbf{Snapshots of the contact maps for the ``two-state model'' with broken detailed balance model}. This figure shows the instantaneous contact maps at different timesteps within the last 200 recolouring steps. \red{The colour code for the maps is based on the colours of the beads forming the contact: red, blue or black for red-red, blue-blue or mixed contacts, respectively (see colour bar)}. From this figure one can appreciate the temporary aggregation of globules, whose underlying 1D epigenetic segment is forming long-lived domains during the simulation. Temperatures were: $T_L= 1.75 \epsilon/k_B$, $T_{\rm Rec}=0.1 \epsilon/k_B$.}
\caption{\red{{\bf Classification of the regimes in the ``two-state'' model with broken detailed balance}. This figure shows the observed regimes for the two-state model with broken detailed balance. Simulations are initialised with the system in the swollen disordered phase. Here, we fix $T_{\rm Rec}=0.1 \epsilon/k_B$ and vary (from bottom to top) $T_L \in \{ 1.5, 1.75, 2, 6\} \epsilon/k_B$. We observe that $T_L \leq 1.75 \epsilon/k_B$ leads to a collapse of the coil into a single-state dominated globule, similar to that observed in the standard ``two-state'' model, but with a higher critical temperature. On the other hand, setting $T_L > 1.75 \epsilon/k_B$ eventually leads to a swollen ordered regime in steady state (see Fig.~\ref{fig:NOdb_ic}). All simulations show an interesting long-lived transient regime, where local domains coexist along the chain giving a characteristic ``block-like'' pattern to the averaged contact map.}}
\caption{ \label{fig3} \textbf{Comparison between the active sample and the control sample.} \textbf{a} and \textbf{b}, Temperature change of the probe, $\Delta T_\mathrm{probe}$, as a function of the QCR operation voltage and the bath temperature for experimental and simulated data in the case of the active sample (\textbf{a}) and the control sample (\textbf{b}). %The control sample has an added superconducting shunt across each resistor as shown in Fig.~1(\textbf{D}) to suppress heat transport through photonic channels. \textbf{c}, Temperature changes of the probe from \textbf{a} and \textbf{b} as functions of the bath temperature at the operation voltage corresponding to the maximum cooling point of the probe and the QCR, respectively. The error bars indicate the \red{maximum} 1$\sigma$ uncertainty for each dataset.}
\caption{ \label{figS2} \textbf{Optimized sample design.} \textbf{a}, Quantum circuit refrigerator with two dc bias leads and bonding pads (on the left) is capacitively coupled to the end of a high-quality co-planar waveguide resonator (on the right). \textbf{b}, View of the design in the area indicated by the red rectangle in \textbf{a}. \textbf{c}, Close view of the \red{3$\times$3-$\mu$m$^2$} copper block forming the normal metal of the QCR. The block is partially overlapping the end of the resonator centre conductor to induce capacitive coupling. Due to shadow evaporation, there is a 20-nm layer of aluminum below most of the copper parts. \textbf{d}, Close view of the bottom NIS junction. The lithographic junction size is 50$\times$70~nm$^2$, giving rise to an effective junction area of roughly 70$\times$70~nm$^2$ due to the 20-nm aluminum layer. See Supplementary Table~\ref{tab_S2} for the parameters of the optimized QCR sample. \textbf{e}, SEM image of a sample similar to the optimized design but with four NIS junctions and a larger normal-metal island.}
\caption{ \label{figS3} \textbf{Effect of operation voltage to the operation characteristics of the QCR.} \textbf{a}, Resonator excitation ($\Gamma^{\mathrm{T}}_{0\rightarrow1}$) and relaxation rates ($\Gamma^{\mathrm{T}}_{1\rightarrow0}$) as functions of the QCR operation voltage, $V_{\mathrm{QCR}}$, for the measured QCR sample (solid lines), the optimized sample using the measured QCR electron temperature (dashed lines), and the optimized sample using 50~mK lower electron temperatures (dash-dotted lines). See Supplemenraty Table~\ref{tab_S2} for the parameters of the optimized sample. \red{Each} operation voltage yielding the minimum temperature corresponding to the photon-assisted tunneling, \red{$T_{\mathrm{res,\Gamma^\mathrm{T}}}=\hbar\omega_0/[\mathrm{log}(\Gamma^{\mathrm{T}}_{0\rightarrow 1}/\Gamma^{\mathrm{T}}_{1\rightarrow 0})k_\mathrm{B}]$, is denoted by} an arrow. Here, the temperature assumes the value 60~mK (solid line), 50~mK (dashed line), and 31~mK (dash-dotted line). \textbf{b}, Resonator quality factor corresponding to the relaxation induced by the photon-assisted tunneling, $Q_{\Gamma^\mathrm{int,T}}$, as a function of the QCR operation voltage for the three cases shown in \textbf{a}. }
\caption{\label{fig:3cases}Radial distribution of the different terms in (\ref{balance}) to the streamwise momentum balance. Solid line $\Delta{\bar{u}_0}(r)$; \textcolor{blue}{-$\cdot$-} Reynolds shear stress term$RSS$; \textcolor{red}{$--$} steady streaming terms $ST$; \textcolor{magenta}{$\cdot$$\cdot$$\cdot$$\cdot$} non-homogeneous terms $NH$. (\textit{a}) case I: large drag-decrease $(A^+,k_c)=(10,10)$, (\textit{b}) case II: small drag-decrease $(A^+,k_c)=(0.7,10)$, (\textit{c}) case III: drag-increase $(A^+,k_c)=(2,2)$.}
\caption{\protect\includegraphics{unnamed-chunk-4-1}}
\caption{\protect\includegraphics{unnamed-chunk-8-1}}
\caption{\protect\includegraphics{unnamed-chunk-11-1}}
\caption{\protect\includegraphics{unnamed-chunk-24-1}}
\caption{\fontsize{11}{11}\selectfont Example discussion on Yahoo!~Answers. Besides the best answer, other answers also contain relevant information (in \textit{italics}). For example, the sentence in {\color{blue} blue} has a contrasting viewpoint compared to the other answers. }
\caption{\small Example discussion from wikipedia talk page for article ``Iraq War", where editors discuss about the correctness of the information in the opening paragraph. We only show some sentences that are relevant for demonstration. Other sentences are omitted by ellipsis. Names of editors are in \textbf{bold}. ``$>$" is an indicator for the reply structure, where turns starting with $>$ are response for most previous turn that with one less $>$. We use ``\textit{NN}", ``\textit{N}", and ``\textit{PP}" to indicate ``strongly disagree", ``disagree", and ``strongly agree". Sentences in {\color{blue}blue} are examples whose sentiment is hard to detect by an existing lexicon.}
\caption{{\bf Figure 1: {\red Four kinds of time series}}. {\red Schematic representation of the four cases considered here. The green bars represent available data for $T$ consecutive generations, whereas the blue dashed lines represent non available data. The time arrow goes from left to right with the present time called generation $T_F$. The data includes an initial condition at generation zero. We follow the trajectory of the allele $B$, whose frequency at present time is known in all cases. In cases $I$ and $II$, $T_F$ is just one generation after the measurement $T$, \ie $T_F=T+1$, so that the available data concern the recent history of the allele. In cases $III$ and $IV$ the measurement $T_F$ is made a long time after the measurement $T$. We can think of the cases $III$ and $IV$ as time-series where both the ancient history and the present frequency are known in detail but data in between are missing. Within the Wright-Fisher model we consider a variant of cases $II$ and $IV$, in which $B$ is very close to fixation but not yet fixed.} \label{fig1}}
\caption{{\bf Figure 2: {\red Selection coefficient for the Moran model.}} {\red For each of the four cases, we have generated 100 independent trajectories with $S=2$ (dashed horizontal line). For each such trajectories we have constructed the likelihoods $L_c(s)$ and $L(s)$ and found the two values of $s$ that maximize each of them separately. From the distribution of these two sets of maximizing $s$ we obtain the mode and the 95\% confidence interval (CI) shown here. The conditioned likelihood $L_c(s)$ always provides a good estimate of the true selection coefficient (red squares). The unconstrained likelihood $L(s)$ delivers a poor estimate of the selection coefficient (grey squares) except for case $III$ due to the slow dynamics of the Moran model. For each trajectory: $T=500$, $N_B(0)=27$ and $N=54$. In cases $I$ and $III$ we have set $N_B(T_F)=40$. In cases $II$ and $IV$ we have set $N_B(T_F)=N$. }}
\caption{{\bf Figure 3: {\red Selection coefficient for the Wright-Fisher model.}} {\red For each of the four cases, we have generated 100 independent trajectories with $S=2$ (dashed horizontal line). For each such trajectories we have constructed the likelihoods $L_c(s)$ and $L(s)$ and found the two values of $s$ that maximize each of them separately. From the distribution of these two sets of maximizing $s$ we obtain the mode and the 95\% confidence interval (CI) shown here. Here, the rapid dynamics of the Wright-Fisher model leads to very short trajectories in cases $II$ and $IV$ that leads to poor statistics. For this reason, in cases $II$ and $IV$ we have set $N_B(T_F)=N-1$ (very close to fixation) instead of $N$. The conditioned likelihood $L_c(s)$ always provides a good estimate of the true selection coefficient (red squares). The unconstrained likelihood $L(s)$ delivers a poor estimate of the selection coefficient (grey squares) with a CI smaller than the box size. For each trajectory: $T=100$, $N_B(0)=27$ and $N=54$. In cases $I$ and $III$ we have set $N_B(T_F)=40$. } \label{fig3}}
\caption{Comparison of maximum error over all frequencies with inf-sup constant of reduced system at selected frequencies for geometry 1. Basis generated by global solves. Increased error and reduced inf-sup constant around basis size of 900. Script: maxwell\_infsup\_during\_reduction.py}
\caption{Basis size distribution. Script: postprocessing\_draw\_basis\_sizes\_maxwell\_geochange.py}
\caption{Charge dynamics driven by multiphoton absorption. (a) NV$^-$ population versus $P_\mathrm{NIR}$, for fixed visible excitation at \SI{532}{\nano\meter} (\SI{9}{\micro\watt}$\simeq$$0.2 P_\mathrm{sat}$, \protect\includegraphics{green_diamond.png}), \SI{592}{\nano\meter} (\SI{20}{\micro\watt}$\simeq \mathrm{0.1} P_\mathrm{sat}$, \protect\includegraphics{592_square.png}), and for NIR light only (\protect\includegraphics{NIR_triangle.png}). Curves are fits described in the text. The shaded region indicates the 95\% confidence interval for a simulation of the NIR data using separately measured rates \cite{Supplemental}. (b) Recombination (\protect\includegraphics{blue_diamond.png}) and ionization (\protect\includegraphics{red_square.png}) rates versus $P_\mathrm{NIR}$ for $P_{532}=\SI{5.4}{\micro\watt}\simeq 0.13P_\mathrm{sat}$. The shaded region indicates the noise floor. (c) NV energy diagram indicating allowed optical transitions and corresponding coefficients in the rate model. Insets in (a-b) depict the experimental pulse sequences for each measurement. Except where indicated by error bars, symbol sizes exceed the experimental uncertainty. }
\caption{Spin-to-charge conversion \textit{via} singlet ionization. (a) Generalized pulse diagram for probing time-domain charge dynamics. (b) Resulting charge state as a function of the delay between a \SI{200}{\nano\second} shelving pulse at \SI{532}{\nano\meter} and a ~\SI{400}{\nano\second} NIR pulse, for a spin prepared in $m_s=-1$. (c) Final charge state as a function of \SI{532}{\nano\meter} shelf duration for spins prepared in $m_s=0$ (\protect\includegraphics[scale=1]{blue_diamond.png}) and $m_s=-1$ (\protect\includegraphics[scale=1]{red_square.png}), and $\tau_d=$ \SI{30}{\nano\second}. (d) Spin-readout SNR comparison of traditional PL (red), measured multi-SCC (blue), and the predicted upper bound for multi-SCC assuming 100\% singlet ionization (green). (e) Required total acquisition time for an SNR=1. The same color legend is used as in (d). The charge state readout duration and power is optimized at each operation time to maximize the speed up.}
\caption{\coloronline\Pairwise factorized Monte Carlo algorithm.\emph{Left:} The move of the active particle is vetoed by one target particle, so that the necessary consensus (all \quot{Y}) is not reached (see \eq{eqFactMetropolis}). \emph{Right:} In the cell-veto algorithm, vetos are provisionally solicited on the cell level (between the active cell $\CCAL_a$ and the target cell $\CCAL_t$) before being confirmed for the active particle, at $\rvec_a \in \CCAL_a$ and the target particle, at $\rvec_t \in \CCAL_t$. Nearby and surplus particles are treated differently.}
\caption{\coloronline~ Event-chain algorithm for a long-ranged dipolar potential in two dimensions, $\beta U = 10 \times (d/r)^3$. \emph{a)} The active particle takes infinitesimal moves in the $+x$ direction. At time $\Delta s$, a move is vetoed by particle $t$, at event-time distance $\rvec_{at}$. The vetoing particle $t$ becomes the new active particle and starts to move in the $+x$ direction. \emph{b)} Heatmap representation of the particle-event rate $\qpart(\rvec_{at})$. The active particle is in the center, black corresponds to $\qpart = 0$. \emph{c)} Probability distributions of the step size $\Delta s$ taken by the active particle and of particle-event distances $r_{at}$. }
\caption{\coloronline~% \emph{a)} Total cell-veto rate \qcelltot, \Eq{eqCellSelectRate}, for inverse-power-law potentials in $D=2$. The event rate for bare particles diverges as $n\to D-1$ (vertical line) because of the presence of periodic images, while screened event rates stay finite. Solid bullets are the Coulomb system ($n=D-2$; for 2D, $U(r)\propto -\ln r$). \emph{b)} Scaling of \qcelltot\with system size$N$, for the screened-lattice algorithm (green solid), and for the bare-particle algorithm (purple dashed). The inclined line is $\propto N^{1/2}$. }
\caption{(Color online) (a) The number of iterations for the Lanczos method with/without wavefunction prediction (+/\textcolor{red}{$\times$}) versus the number of iterations for the parallel iDMRG. (b) Fidelity errors $\epsilon^{(n)}_{1}$ in Eq.~(\ref{eq.fid1}), $\epsilon^{\prime(n)}_{1}$ in Eq.~(\ref{eq.fid2}) and truncation error $\varepsilon^{(n)}_{1}$ in Eq.~(\ref{eq.tru}) are denoted by black lines, block circles, and red lines with a circle, respectively. }
\caption{(color online). Experimental setup. (a)~False-colored scanning electron micrograph of the SET (brown) and the SEB (blue). Both devices are fabricated by electron beam lithography and three-angle shadow evaporation~\cite{Joonas2015}. The gate voltages $V_\text{g1}$ and $V_\text{g2}$ are used to control the tunneling rates. The bias voltage $V$ is applied across the SET and the current $I$ is measured. (b)~The current in the SET switches between the two normalized values $\langle I_-\rangle~=0$ and $\langle I_+\rangle= 1$ due to the tunneling of single electrons on and off the SEB. (c)~Distribution of the time-integrated current (\textcolor{red}{$\circ$}) for the integration time $\tau=180$~ms together with the tilted ellipse given by Eq.~(\ref{eq:P(I)}). The controllable rates for tunneling on and off the SEB are $\Gamma_+= 72$ Hz and $\Gamma_-= 37 $ Hz.}
\caption{% %(Color online) Ratio of the DY reaction cross sections for different nuclei vs. $x_T$. Data points are from the E772 experiment~\cite{E772} (\fullsquare) and E866 experiment~\cite{E866} (\opensquare). Note that the ratio is normalized to one bound nucleon and taken relative to the deuteron for E772 and $^9$Be for E866. The curves are the predictions of Ref.\cite{KP14} for the deuteron (green) and beryllium (red) ratios with (\full) and without (\dashed) the projectile energy loss effect (see text and legend). \label{fig:dy}}
\caption{\small Mean scores assigned by annotators to different systems. The best-performing baseline and experimental systems are shown in bold. {\color{blue}{this table need revamp}}}
\caption{\small Precision of different systems. The best-performing baseline and experimental systems are shown in bold.{\color{blue}{this table need revamp}}}
\caption{Comparison of some of the many determinations of the strange magnetic moment. Results in {\emr red} are from the global analysis of world data, results in {\color{green!90!blue}green} are from indirect calculations, and results in {\emb blue} are from lattice QCD calculations. }
\caption{Results by MLC classifiers trained with the spectral bands ($\omega$), with spatial features extracted from the three first \blue{ principal components, PCs} ($s$, including morphological and attribute filters) or with the proposed active set (\textsc{AS-}). In the \textsc{Houston 2013A/B} cases, features extracted from the DSM have been added to the input space of the baselines.}
\caption{\makered{(color online) PDF of the magnitude of the horizontal velocity ($V_H = \sqrt{V_x^2+V_y^2}$) normalised by the gravitational velocity. Multiple experiments for the same $\text{Ga}$ are in agreement and are combined to improve the quality of the statistics. Starred data is from numerical simulations and has $\Gamma=1.5$~\cite{doychev14,uhlmann14a}, while the experiments have $\Gamma = 2.5$.}}
\caption{{\bf Qualitative localization results compared to \cite{brubaker2013lost}:} Results of other sequences can be found in appendix. The left most column shows the full map region for each sequence, followed by zoomed in sections of the map showing the posterior distribution over time. The upper row is the result of \cite{brubaker2013lost}, and the lower row is ours. The black line is the GPS trajectory and the concentric circles indicate the current GPS position. Grid lines are every 500m. High probability is indicated with {\color{red}red} while low probability regions are shown in {\color{blue}blue}.}
\caption{{\bf Qualitative results of semantics estimated by our CNNs: } predictions of road and intersection types are under each image. The failure prediction is marked in {\color{blue} blue}. We also show predicted sun direction on the right of each image in black, while ground truth in {\color{red} red}.}
\caption{2D embedding of the Sun-CNN feature space with t-SNE. Color of the image border denotes ground truth relative sun position, \ie\{\color{red} \bf red} sun is in front of the vehicle, {\color[rgb]{0,0.7,0} \bf green} sun is on the right, {\color{cyan} \bf cyan} sun is on the back, and {\color{magenta} \bf magenta} sun is on the left. Sun-CNN not only effectively separates images showing different sun directions, but also preserves the \emph{relative relationship}, \eg\images where sun are on the{\color{red} \bf front} lie between images where sun are on the {\color{magenta} \bf left front} and {\color[rgb]{0.9,0.8,0} \bf right front}}
\caption{\footnotesize Clip from the AMI meeting corpus \cite{Carletta05theami}. A, B, C and D refer to distinct speakers; the numbers in parentheses indicate the associated meeting decision: {\sc decision 1}, {\sc 2} or {\sc 3}. Also shown is the gold-standard (manual) abstract (summary) for each decision. \textcolor{blue}{Colors} indicate overlapping vocabulary between utterances and the summary. \underline{Underlining}, {\it italics}, and [bracketing] are decscribed in the running text.}
\caption{ A compilation of the \redd{e}xperimental puzzles of heavy ion collisions described in this section. Panels (a,b) show the near-same dependence on geometry of, respectively, soft and hard $v_2(p_T)$ \cite{cms1,cms2}. Panels (c,d) show $v_2(p_T)$ independence on energy and rapidity \cite{stars,brahms}. (e,f) show that once $p_T$ was integrated, $v_2$ only depends on the transverse entropy density in the same way as $\ave{p_T}$ \cite{cms1,phobos}, while (g,h) show, respectively, the same $v_3$ for pA,dA and AA \cite{cmspa,cmspp,phenixda} and the near-equality of photon and hadron $v_2$ \cite{photon}.}
\caption{\label{figu2bessel} $u_2$ as a function of $x, Q$ (solid, dotted and dashed lines for $Q=1,2,3$ GeV) when started as a small value at a ``moderate'' $x$ (where processes such as higher\redd{-}twist can occur) and evolved to $x \rightarrow 0$. The evolution is cut off when the linear approximation becomes untenable. }
\caption{(a) SFG output power and single photon conversion efficiency as a function of the pump power with (\textcolor{red}{$\bullet$} at 135$^\circ$C with theoretical prediction as solid line and \textcolor{red}{$\blacksquare$} at 95~$^\circ$C, internal conversion) and without (\textcolor{green}{$\bullet$} at 135$^\circ$C and \textcolor{green}{$\blacksquare$} at 95$^\circ$C, external conversion) Fresnel losses compensation. Coupled IR power is 1~mW. (b). DFG output power with internal single photon conversion efficiency (\textcolor{red}{$\bullet$}) and the theoretical fit (solid line) for a coupled UV power of 30~$\mu$W at a working temperature of 135~$^\circ$C. }
\caption{Measure of the scattered Raman photons rate as a function of the coupled pump power after annealing and RPE (\textcolor[rgb]{1,0,0}{\textbullet}) and after 6h of extra annealing (\textcolor[rgb]{0,0,1}{\textbullet}) together with a linear fit (solid line). }
\caption{\label{FigWordAdj} Word adjacencies example. The colours indicate the class labels, nouns~\showcolor{graphic1} and adjectives \showcolor{graphic2}, and the magnetic eigenmaps correspond to $g = 2/5$.}
\caption{\label{FigPolBlogs} Political blogosphere example. The colours indicate the class labels, left leaning \showcolor{graphic1} and right leaning \showcolor{graphic2}, and the magnetic eigenmaps correspond to $g = 1/4$.}
\caption{Integrated information $\Phi$ for different levels of measurement noise $p$. Inset: (\textcolor{blue}{blue}) Mean and variance of the ratio $\eta$ between $\Phi$ of the corrupted and the original time series. (\textcolor{red}{red}) Exponential fit $\eta = \exp(-p/\ell)$, with $\ell \approx 0.04$.}
\caption{Examples of aggregates. Top left = BA {\color{red}(512 monomers)}, top right = BAM2 {\color{red}(512 monomers)}. Bottom = hierarchic BAM2: {\color{red}2048 pseudo-monomers, each pseudo-monomer} is an aggregate composed of a different number of monomers (see description in the text).}
\caption{Ratio of the cross-section $A$ for BCCA, BA, BAM1, BAM2 clusters made up of $N$ monomers to the cross-section of a sphere $A_{sphere}$ with the same {\color{red}mass}.}
\caption{Ratio of aggregate cross-section $A$ to cross-section of an equivalent mass sphere $A_{sphere}$ for BCCA and hierarchic aggregates [BA]$^2$ (crosses) and [BAM2]$^2$ (stars). Lines ({\color{red}dashed and dotted}) show extrapolated relations for [BA]$^2$ and [BAM2]$^2$ respectively. See also Table \ref{AoverAsphere}.}
\caption{The environments of the most massive $z=8$ black holes in \bluetides. The images show the large scale gas density around the black holes, zooming into regions of width 6 $\mpch$ (left panels) to 3 $\mpch$ (center) and 0.5 $\mpch$ (right). The gas density field is color coded by temperature (blue to red indicating cold to hot respectively). The final zoom on the top right column show the stellar density for the host galaxy. The positions of the black holes in all images are indicated with white crosses the sizes of which are proportional to black hole mass.}
\caption{Left: The disk-to-total ratio for \bluetides BH host galaxies versus $t_{1}$, the tidal field strength. Right: the Eddington accretion rate of BHs versus $t_{1}$. The inset panels in each case show the same data, but with an increased y-axis range. The points with error bars show the mean values of D/T and Eddington ratio in bins of $t_{1}$. The error bar is the error on the mean, and the shaded regions indicate the range which encloses 68\% of the galaxies in each bin. The smooth curves are linear fits of D/T and Eddington ratio to $t_{1}$, with the fit parameters given in Table \ref{table:spearman}}
\caption{(a) A multistep kinetic scheme for oxidation of aluminum as described in the Cabrera-Mott model (adapted from Ref.{\thinspace \onlinecite{Boggio66-Metal-ThinFilms-Theory})}. Directly after ionization [Eq.{\thinspace(}1.a)], electrons pass freely through the oxide interface until reaching incorporated oxygen which are then ionized to O$_{2}^{-}$ [Eq.{\thinspace(}1.b)]. This charge transfer together with the left-behind positive Al$^{3+}$ induce a local electric field which drives the slow migration of Al$^{3+}$ across the oxide interface leaving behind a vacancy \color{red}V$_{\text{Al}}$ \color{black} [Eq.{\thinspace(}1.c)]. Equation{\thinspace(}1.d) indicates a typical formation of Al$_{2}$O$_{3}$ by the combination of migrated Al$^{3+}$ and ionized O$_{2}^{-}$. Panels (b) and (c): Illustration of the reaction at metal-oxide interface (represented by the solid green line) of a directly exposed film [panel (b) is before an event of Eq.(1.a) while panel (c) is after the event of Eq. (1.c); adapted from Ref.{\thinspace\onlinecite{Atkinson85-Growth-Oxides-Films}}]. (d) We consider that, during the codeposition process of a granular film, the incorporated oxygen does not enter as an idle and neutral entity, rather it does react with Al matrix, just as in the normal oxidation process of Cabrera-Mott leading to an interaction similar to that described in panels (a)-(c). Ultimately this accumulates into an extended \nanosized grain. (e) Similar reaction occurring at the boundary of an isolated oxide grain of an irradiated film. In all cases, incorporated and ionized oxygen are represented, with no loss of generality, by O$_{2}$ and O$_{2}^{-}$. }
\caption{(\textit{a}) Representative $\rho(T,n\mathrm{th})$ curves. The solid lines are linear fit with a slope $\left( \partial\rho/\partial T\right) _{\text{100-220K}}$. Due to aging effects, a unique room-temperature resistivity for each warming-up measuring cycle is taken to be the extrapolated $\rho_{\text{300K}}^{ext}$, solid circle on the high-$T$ linear extrapolation. (\textit{b}) Evolution of $\left( \partial\rho/\partial T\right) _{\text{100-220K}}$ as a function of $\rho_{\text{300K}}^{ext}$. (\textit{c}) An expanded view of low-$T$ $\rho(T,H$, $n\mathrm{th})$ curves of panel \textit{a}. \Open (solid) symbols denote zero-field (5kOe) curve. (\textit{d}) Hall coefficients of as-prepared film and that of the same film measured 85 days after the 7$^{th}$ irradiation. (\textit{e}) $\Delta\rho (H/T$, $n\mathrm{th})=\rho(H/T$, $n\mathrm{th})-\rho(0,T$, $n\mathrm{th})$ $versus$ \$\left( H/T\right) ^{2}$: showing the breakdown of the $\left( H/T\right) ^{2}$ scaling. These curves are from $n\mathrm{th}=$6, 7 irradiation and are limited to the range of $\ H_{c2}<H<$20\thinspace kOe and $T_{c}<T<T_{\text{K}}^{mn}$ (small $\nicefrac{H}{T}$). (\textit{f}) The same as panel \textit{e} but are scaled to $\sqrt{\nicefrac{H}{T}}$. This better scaling is emphasized by the solid line fit $\Delta\rho(\nicefrac{H}{T}$, 6$\mathrm{th})\propto\left( H/T\right) ^{\nicefrac{1}{2}}$. }
\caption{Normal-state and superconducting phase diagram of different Al thin films shown as a log-log plot of $T$ vs $\rho_{300\text{K}}$. \color{red}$\blacktriangledown$, $\spadesuit$, $\clubsuit,\blacklozenge$\color{black}: $T_{\text{c}}^{zero}$, $T_{\text{c}}^{onset}(=$ $T_{\text{K}}^{\max})$, $T_{\text{K}}^{\min}$, $T_{\text{NCR}}$, resp., as obtained by this work. \color{blue}$\bigstar$, $\spadesuit$, $\clubsuit$, $\blacklozenge$\color{black}: $T_{\text{c}}^{zero}$, $T_{\text{c}}^{onset}(=$ $T_{\text{K}}^{\max})$, $T_{\text{K}}^{\min}$, $T_{\text{NCR}}$, resp., of{ granular films }deposited at 77 K (Refs. {\thinspace \onlinecite{Bachar15-Mott-granular-Al,*Bachar13-Kondo-granular-Al,Bachar14-PhD-Thesis}}){. }T{he location of each thermal event is explicitly shown at the} right-side inset.{ }\color{cyan}$\triangledown,\blacktriangledown $\color{black}: $T_{c}$ of granular films deposited at room-temperature (Ref.{\thinspace}\onlinecite{Abeles67-Hc-granular-SUCs} and {\onlinecite{Chui81-Nano-Al-MagRes-Localizatio,*Chui81-Tc-Hc-Granular-Al-SUC,*Mui84-Granular-Al-MagRes}}, resp.). \color{yellow}$\bigstar$\color{black}: $T_{c}$\after O-implantation at liquid helium (Ref.{\thinspace \onlinecite{Lamoise75-Irradiation-Al-Resistivity-annealing})}. Left-hand-side inset: A semilog plot of our irradiated-film's $\rho(T,6^{th},$0 kOe$)$ and $\rho(T,6^{th},$5 kOe$)$ (open and close circles, resp.). Right-hand-side inset: A semilog plot of $\rho(T)$ curve of a granular sample having $\rho_{300K}=$310 $\mu\Omega$-cm (taken from Refs.{\thinspace \onlinecite{Bachar15-Mott-granular-Al,*Bachar13-Kondo-granular-Al,Bachar14-PhD-Thesis}}). The main graph is extrapolated down to 2.75$\mu\Omega$-cm of bulk Al.\cite{Desai84-Al-resistivity} The lines are guides to the eye }
\caption{Inelastic scattering spectra plotted from the Brillouin zone-center to the zone-edge. \textbf{(a-d)} $M_{TA_2}$ and $X_{TA}$ illustrate what a typical dependence looks like. The dashed curve is only to guide the eye. In contrast, the $M_{TA_1}$ and $R_{TA}$ at the zone-edge have magnitude larger even than the Bragg tail. \textbf{(e)} Sketch of the motion of the observed anharmonic modes with the idealised A-site cation (\MA) position represented in blue, the B-site cation (Pb$^{2+}$) in grey at the center of the octahedra, and the X-site anion (I$^-$) in purple. Animations of these modes, and several others, are online (\textcolor{blue}{\href{https://figshare.com/s/97a6cbc033b17aa83a18}{Figshare}}). }
\caption{Local symmetry breaking in \MAPI\at 350~K.\textbf{(a,b)} Distortions from cubic symmetry generate anisotropic cavities and couple to motion of the \MA\ion, which we represent as off-centered and oriented along the long-axis of the cavity.\textbf{(c)} DFT-based lattice dynamic calculations show that the energy minimum at the $R$-point at 350~K is displaced in a double-well potential that causes local symmetry breaking. \textbf{(d)} Comparison of the experimental PDF (purple) to cubic ($Pm3m$), centrosymmetric ($I4/mcm$), and non-centrosymmetric ($I4cm$) tetragonal models (blue) show a superior fit for the low-symmetry models at low-$r$ (2-8 $\text{\AA}$). However, the models perform oppositely at high-$r$ with the high-symmetry cubic structure giving the best agreement to the data in the 12-50 $\text{\AA}$ region. The residuals (orange) are scaled $\times3$ for clarity.}
\caption{Dependence of quantum channel fidelity $\mathcal{F}$ (a) and two-photon coincidence counts (b) on initial polarization state of target qubit specified by angle $\omega$. The results are shown for three versions of the protocol: full implementation with quantum filter and feed-forward ({\color{MYgreen}{\scriptsize $\blacksquare$}}), quantum filter set on but feed-forward switched off ({\color{MYyellow}{\large $\bullet$}}), and both filtering and feed-forward switched off ({\color{red}{\normalsize $\blacktriangledown$}}). The horizontal dashed line shows the classical measure-and-prepare bound $\mathcal{F}=2/3$. The vertical dashed line indicates the setting $\omega=55^\circ$ for which the channel matrices are plotted in Fig. \ref{fig:res1cut}. The solid lines indicate predictions of a theoretical model that accounts for imperfections of the partially polarizing beam splitter where the source and target photons interfere.}
\caption{\color{red} benchmark (Vol. $0.152633$)}
\caption{\color{blue!70!black} $1^{st}$ order (Vol. $0.105885$)}
\caption{\color{blue!70!black} $2^{nd}$ order (Vol. $0.104254$)}
\caption{\color{red} benchmark (Vol. $0.152633$)}
\caption{\color{blue!70!black} $1^{st}$ order (Vol. $0.12375$)}
\caption{\color{blue!70!black} $2^{nd}$ order (Vol. $0.0949364$)}
\caption{Thresholded stresses for each scenario in the cantilever setup, $1^{st}$ and $2^{nd}$ order dominance for the equal setup (row one and two, color-coded as $0$~\protect\includegraphics[height=1ex,width=5em]{images/res/cb_stress}~$4.99$ and $0$~\protect\includegraphics[height=1ex,width=5em]{images/res/cb_stress}~$4.85$) and the varying load and varying probability setup (row three and four, colorcoded as $0$~\protect\includegraphics[height=1ex,width=5em]{images/res/cb_stress}~$9.05$ and $0$~\protect\includegraphics[height=1ex,width=5em]{images/res/cb_stress}~$6.87$). Stresses $> 4.99$,$> 4.85$, $> 9.05$ and $> 6.87$ are mapped to red.}
\caption{\color{red} Benchmark\\Vol. $0.484881$}
\caption{\color{blue!70!black} $1^{st}$ order\\Vol. $0.330986$}
\caption{\color{blue!70!black} $2^{nd}$ order\\Vol. $0.12375$}
\caption{Thresholded stresses for each scenario in the weighted carrier plate setup, $1^{st}$ and $2^{nd}$ order dominance, colorcoded as $0$ \protect\includegraphics[height=1ex,width=5em]{images/res/cb_stress} $3.21$ and $0$ \protect\includegraphics[height=1ex,width=5em]{images/res/cb_stress} $3.24$, stresses $> 3.21$ and $> 3.24$ are mapped to red.}
\caption{Additional \aastex\symbols}
\caption{Percolation. (a): diagram of a percolation event on the field. Black areas are the proper surfaces of the dunes, gray areas are their interaction surfaces. The surface surrounded in black shows the cluster of dunes that percolates through the periodic boundaries. (b): probability $p_p$ for the system to percolate as a function of the initial density $\rho_0$, with $\Phi = 0$ and $\lambda = 0$, at a fixed length $L=16$, for different widths of the field : $\ell= 32$ ($\bullet$), 128({\color{red}$\blacksquare$}), 1024 ($\color{green}\blacklozenge$). The continuous lines are given by Eq.~\ref{eq:proba} without any fitting parameter. (c): $\ln(1-p_p)$ \emph{vs} $\ell$ for a fixed system length $L=16$ and different initial densities ($\rho_0\simeq 0.0039 (\bullet)$, $0.0078 (\color{red}{\blacksquare})$, $0.0117 (\color{green}{\blacklozenge})$, $0.0156(\color{blue}{\blacktriangle})$, $0.0195(\color{orange}{\blacktriangleleft})$). The continuous lines are given by Eq.~\ref{eq:proba_scaling} without any fitting parameter. (d): finite size effects on the percolation transition. We plot the initial density $\rho_0$ needed to get a certain probability $p_p^0$ for the system to percolate : when $\ell$ varies and $p_p^0 = 0.25$ ($\bullet$), 0.5 ($\color{red}\blacksquare$), 0.75 ($\color{green}\blacklozenge$), $L=16$; when $L$ varies and $p_p^0=0.978$ and $\ell=2$ ($\color{blue}\blacktriangle$). The black continuous line shows a power law of exponent -1/2, the dashed (blue) one an exponent of 1. \label{fig:percol}}
\caption{The EAS-MSU array setup: (a)~the central part of the array; (b)~the entire array. Four empty black squares in the center represent the special \red central \black room. Green circles denote ``vans'' with scintillation counters, empty red circles denote ``vans'' without scintillation counters, blue squares denote ``boxes''. Lines represent the tetragons of the peripheral trigger system. See the text for more details.}
\caption{\label{tab:cuts} \red Number of independent and resampled CORSIKA showers and the effect of various cuts on the number of events in Monte-Carlo and in real data. Note that we are interested in $N_{e}>2\times 10^{7}$ and simulated showers with $E>10^{16.5}$~eV only, while the installation recorded also events with much lower energies, which explains a large amount of data events not passed the $N_{e}$ cut. The geometric area where the artificial showers are thrown 1.3 times exceeds the area where they are selected.}
\caption{Data versus MC comparison of the distribution in $R$. Points with error bars: data. \red (a):~green dashed hystogram: MC (protons), red dotted hystogram: MC (iron); (b):~blue hystogram: MC (the best-fit composition). }
\caption{Distribution of ratios of the reconstructed $N_{e}$ to the ``thrown'' $N_{e}$ which is the total number of charged particles from the CORSIKA output supplemented by the contribution of photons to the signal, see the text, for MC events. \red (a):~green dashed hystogram: MC (protons), red dotted hystogram: MC (iron); (b):~blue hystogram: MC (the best-fit composition). \black}
\caption{\label{fig:Cf252sf-pfgs}Average prompt fission \gray~spectrum (PFGS) for the spontaneous fission of $^{252}$Cf. The high-energy tail of the calculated PFGS reproduces very well the recent measurement by Billnert \etal~\cite{Billnert:2013} (see inset), while the low-energy part of the spectrum shows very clear structures that can be attributed to transitions between low-lying excited states in fission fragments. \CGMF~calculations do reproduce many of these structures reasonably well. The exact magnitude of these peaks depends on several factors, including our present knowledge of the nuclear structure of fission fragments. Note that no energy resolution broadening has been applied to the \CGMF~results.}
\caption{\label{fig:Cf252sf-Pnug}(top) Prompt fission \gray~multiplicity distribution calculated with \CGMF~and to compared to a negative binomial distribution~\cite{Valentine:2001} with best-fit parameters. A double Poisson distribution as proposed originally by Brunson~\cite{Brunson:1982} to fit his experimental results is shown in red on the log-plot (bottom). The negative binomial distribution is in much better agreement with the tail of the distribution as calculated by \CGMF~than the double Poisson distribution. The \gray~threshold used in the \CGMF~results is 140 keV.}
\caption{\label{fig:Ngt-Cf252SF}Average prompt \gray~multiplicity as a function of time for $^{252}$Cf (sf). Experimental data are from Verbinski~\cite{Verbinski:1973}, Billnert~\cite{Billnert:2013}, and Chyzh~\cite{Chyzh:2014}. Valentine data point~\cite{Valentine:2001} represents a weighted average of experimental data available prior to 2001.}
\caption{\label{fig:Ngt-U5T-Pu9T}Same as Fig.~\ref{fig:Ngt-Cf252SF} for the thermal neutron-induced fission of $^{235}$U (top) and $^{239}$Pu (bottom). The experimental data by Sund \etal~\cite{Sund:1974} (green solid steps) were normalized to match \CGMF-calculated average \gray~multiplicity at 10 nsec.}
\caption{\label{fig:lateGammasPerMass}The distribution of prompt fission \grays~emitted between 10 nsec and 1 $\mu$sec after fission, for $^{252}$Cf spontaneous fission, is shown as the blue histogram. Also shown are the independent fission yields (IFY), i.e., post-neutron emission fission yields, as calculated by \CGMF~(dashed cyan) and taken from the ENDF/B-VII.1 library (solid red), in arbitrary units.}
\caption{\label{fig:Cf252-pfgs-10-2000}Energy spectrum of late fission \grays~emitted in the 10 nsec to 2 $\mu$sec time window following fission, in the case of $^{252}$Cf spontaneous fission. Experimental data are from John~\etal~\cite{John:1970}.}
\caption{\label{fig:Ngt-all}Cumulative average prompt \gray~multiplicity as a function of time since fission for thermal neutron-induced fissions of $^{235}$U and $^{239}$Pu, and spontaneous fission of $^{252}$Cf. }
\caption{\label{fig:Egt-all}Cumulative average prompt total \gray~energy as a function of time since fission for thermal neutron-induced fissions of $^{235}$U and $^{239}$Pu, and spontaneous fission of $^{252}$Cf.}
\caption{\label{fig:sensitivity}Sensitivity of the time dependence of the cumulative \gray~multiplicity (red) and energy (blue) on the choice of the model input parameters $R_T$ and $\alpha$. The ``reference" results, shown in Figs.~\ref{fig:Ngt-all} and~\ref{fig:Egt-all}, are reported here as solid lines.}
\caption{\label{fig:1} Dose fractionation effect for evenly fractionated carbon-ion RT of total 40 Gy (RBE): (a) ERD ($D_{\rm ER}$, \full) and TCP ($P_{\rm TC}$, \dashed) as functions of number of fractions. (b) RBEs for low-$\alpha$ (\dotted), mid-$\alpha$ (\dashed) and high-$\alpha$ (\full) carbon-ion radiations as functions of fraction RWD. }
\caption{\label{fig:2} A multimodal RT treatment in ten fractions of 2 Gy with photons and six fractions of 4 Gy (RBE) with carbon ions over 16 days: cumulative RWD ($D_{\rm RW}$, \dotted), ERD ($D_{\rm ER}$, \full) and TCP ($P_{\rm TC}$, \dashed) as functions of number of fractions or days.}
\caption{\label{fig:3} Relation between fraction clinical dose (${D_{\rm C}}_1$) and fraction RWD (${D_{\rm RW}}_1$, \dashed), instant ERD (${D_{\rm ER}}_1$, \full) and the photon-equivalent dose (${D_{\rm x}}_1$, \dotted) in the NIRS clinical dosimetry system.}
\caption{\label{fig:5} Profiles on the patient right--left axis for the prostate-cancer treatment plans: (a) dose-mean $\alpha/\beta$ ratio for inactivation of incubative HSG tumor cells, (b) total physical dose for 12 fractions, (c) total RWD and (d) ERD with successive beam delivery, and (e) total RWD and (f) ERD with alternate beam delivery. In (a) and (b) are drawn the contributions of the left (\dashed) and right (\dotted) beams and their total (\full) with the $\alpha_{\rm ref}/\beta_{\rm ref}$ = 12.42 Gy level (\chain). In (c)--(f) are drawn the treatment-plan doses for 12 (\full), 8 (\dashed), 4 (\dotted) and 2 (\chain) fractions prescribed for 24.49 Gy in ERD.}
\caption{Reynolds-averaged particle number density (a) and filtered streamwise velocity (b), obtained with different SGS particle models: no-interpolation of LES and particle quantities ({\color{red}{$\square$}}), NGP interpolation ({\bf{\color{blue}{$\circ$}}}) and second-order interpolation ($\triangle$). Downward triangles ({\color{magenta}{$\triangledown$}}) in panel (b) refer to the filtered streamwise velocity provided by LES. The time window for averaging is $\Delta t^+ = 3000$, in wall units.}
\caption{Reynolds-averaged particle number density (a) and filtered streamwise velocity (b), obtained with different SGS models for the fluid: No-model ({\color{red}{$\square$}}, {\color{red}{$-$}}), Smagorinsky model ({\color{blue}{$\circ$}}, {\color{blue}{$- -$}}) and Germano dynamic model ({\color{black}{$\triangle$}}, {\color{black}{$- \cdot$}}). The time window for averaging is $\Delta t^+ = 3000$, in wall units.}
\caption{Comparison of particle number density distributions predicted by two different particle SGS model formulations: Simplified stochastic model (\textit{LFMDF1}, {\color{purple}{$\triangledown$}}) and complete stochastic model (\textit{LFMDF2}, {\color{black}{$\triangle$}}). See also Sec. \ref{sec:equiv_stoc_syst}. Other symbols: {\color{red}{$\square$}} DNS, {\color{blue}{$\circ$}} LES without particle SGS model. Panels: (a) $St=1$ particles, (b) $St=5$ particles, (c) $St=25$ particles. Profiles are computed at $t^+ =2130$ after particle injection into the flow.}
\caption{Effect of parameter $C_0$ on particle number density along the wall-normal coordinate (\emph{a-priori} estimate). Red symbols ({\color{red}{$\square$}}) refer to the DNS result, all other symbols refer to LES results obtained with the LFMDF model. Panels: (a) $St=1$, (b) $St=5$, (c) $St=25$. Profiles are computed at $t^+ =2130$ after particle injection. }
\caption{Effect of parameter $C_{\epsilon}$ on particle number density along the wall-normal coordinate (\emph{a-priori} estimate). Red symbols ({\color{red}{$\square$}}) refer to the DNS result, all other symbols refer to LES results obtained with the LFMDF model. Panels: (a) $St=1$, (b) $St=5$s, (c) $St=25$. Profiles are computed at $t^+ =2130$ after particle injection.}
\caption{Comparative assessment of the LFMDF model with Eq. (\ref{eq:st1}): Predictions of the instantaneous particle number density at varying Stokes numbers ({\color{black}{$\triangle$}}) are compared with DNS results ({\color{red}{$\square$}}) and with LES results with no particle SGS model ({\color{blue}{$\circ$}}). Panels: (a) $St=1$, (b) $St=5$, (c) $St=25$. Profiles are computed at $t^+ =2130$ after particle injection. }
\caption{Comparative assessment of the LFMDF model: Prediction of the particle velocity rms at varying Stokes number ({\color{black}{$\triangle$}}) are compared with DNS results ({\color{red}{------}}) and with LES results with no particle SGS model ({\color{blue}{$\circ$}}). Panels: (a),(d) $St=1$, (b),(e) $St=5$, (c)-(f) $St=25$; (a)-(c) streamwise component, (d)-(f) wall-normal component. Statistics are obtained averaging over a time window $\Delta t^+ =1800$. }
\caption{Differential (\blue{top}) and integral (\blue{bottom}) marginalized posterior for the 0.1--1 GeV energy flux for LVT151012 and the time window $0$--$8$ ks. In the \blue{top} panel, we note that the rightmost part of the distribution is affected by sampling noise. In the \blue{bottom} panel \citep[reproduced from][]{Racusin16}, the flux at which the blue curve intersects a given probability $P(F < x)$ corresponds to the upper bound at that credibility level.}
\caption{Comparison between the TS distributions of the data (solid blue histogram) and Monte Carlo simulations (dashed red \blue{histogram} ). %\blue{The theoretical curve (a $1/2\,\chi^{2}$ with one degree of freedom), is also shown (dashed green line).} }
\caption{\textcolor{blue}{System responses without (left) and with (right) a global performance index but $q=0$.} }
\caption{\textcolor{blue}{System responses with a global performance index as $q=10$ (left) and $q=20$ (right).} }
\caption{\textcolor{blue}{Hierarchical decentralized $H_{\infty}$ control of the given MAS in presence of a white noise with magnitude equals to $1$.} }
\caption{\textcolor{blue}{Fully decentralized $H_{\infty}$ control of the given MAS in presence of a white noise with magnitude equals to $1$.} }
\caption{\textcolor{blue}{Hierarchical decentralized $H_{2}$ control of the given MAS in presence of a white noise with magnitude equals to $1$.} }
\caption{\textcolor{blue}{Fully decentralized $H_{2}$ control of the given MAS in presence of a white noise with magnitude equals to $1$.} }
\caption{\small{\em (Online color).} Parameter-space showing available DNS data sets with $\re_\lambda=185$ and $\re_\lambda=400$. For values of $\st$ smaller than $0.05$ the $x$ axis is linear, while it is logarithmic for values of $\st$ larger than $0.05$. For each data set with $\re_\lambda=185$ we have analysed a total of $130000$ trajectories of duration $6T_L$ and for each data set with $\re_\lambda=400$ we have analysed a total of $200000$ trajectories of duration $2.5T_L$. Level curves $\st(1-\beta)=\epsilon$ for constants $\epsilon=\{-1,-0.1,-0.01,0,0.01,0.1,1\}$ are plotted as black lines. Parameter families: With $\re_\lambda=185$: $\beta=0$ (red,$\circ$), $\beta=0.25$ (magenta,$\vartriangle$), $\beta=0.5$ (cyan,$\triangledown$), $\beta=0.75$ (green,$\triangleright$), $\beta=1$ (black,$\Box$), $\beta=1.25$ (brown,$\Diamond$), $\beta=1.5$ (purple,$\triangledown$), $\beta=2$ (orange,$\triangleright$), $\beta=2.5$ (dark green,$\triangleleft$), $\beta=3$ (blue,$\vartriangle$). With $\re_\lambda=400$: $\beta=0$ (red,\marker{6}). Additional data from~\cite{Cal09,Pra12}: $\beta=3$ (blue,\marker{7}). }
\caption{\small{\em (Online colour).} Acceleration variance ({\bf a}) and flatness ({\bf b}) for DNS data at changing $\beta$, $\st$ and Re. {\bf a}: points correspond to the DNS data (same symbols of Fig.\ref{fig:phasespace}). Solid lines (labeled with their corresponding value of $\beta$) show the closure scheme prediction for the acceleration variance of light and heavy particles, Eq.~(\ref{aVar}), normalized with the fluid variance, for all data from RUN I; dashed line corresponds to the closure for RUN II. {\bf b}: the flatness measured on the DNS data and the one predicted by the closure scheme, Eqs.~(\ref{aQuad}), for heavy and light particles. Thin dashed black line shows the limit of normal distributed acceleration components. Additional data for $\beta=3$ (blue,\protect\includegraphics[width=2mm,clip]{markBW7.eps}) is omitted in panel {\bf b} because the flatness was not evaluated in Refs.~\cite{Cal09,Pra12}. }
\caption{\blue{Single-particle quantities for a triangular ring. (a) The $12$ lowest states arranged into $4$ energy levels, the inset shows the degeneracy of corner states. (b-d) Probability distributions associated with the energy levels shown in Fig. (a). The $x$ and $y$ coordinates are in units of $R_{\mathrm{ext}}$}. }
\caption{Energy levels for a triangular ring. The number of confined electrons ($N$) is shown in each figure and the interaction parameters shown in Fig.\(d) are valid for all figures. \blue{In the insets to panels (a-c) we show the fine structure of the in-gap states.}}
\caption{\blue{Two-particle lateral localization. Probability distributions for two electrons in corner \blue{(including the in-gap)} states (a), in mixed corner-side states (b) and in side states (c). The $x$ and $y$ coordinates are in units of $R_{\mathrm{ext}}$.}}
\caption{Configuration-space interquark potential $V(r)$ of the Fierz-symmetric kernel $K(p,q),$ for the constituent quark mass $m=0$ and mixture $\eta=0$ \cite{WL15} (black), $\eta=1$ (\textcolor{red}{red}), $\eta=2$ (\textcolor{magenta}{magenta}), $\eta=-0.5$ (\textcolor{blue}{blue}), or $\eta=-1$~(\textcolor{violet}{violet}).}
\caption{Configuration-space interquark potential $V(r)$ of the Fierz-symmetric kernel $K(p,q),$ for the constituent quark mass $m=1$ and mixture $\eta=0$ \cite{WL15} (black), $\eta=0.5$ (\textcolor{red}{red}), $\eta=1$ (\textcolor{magenta}{magenta}), $\eta=2$ (\textcolor{blue}{blue}), and $\eta=-1$~(\textcolor{violet}{violet}).}
\caption{(a) Mass function $M(\underline{k})$ deduced from the Dyson--Schwinger model of Ref.~\cite{PM97b} for the quark propagator. (b) Configuration-space interquark potential $V(r)$ numerically determined from $M(\underline{k}^2),$ for constituent~quark~mass $m=0$ (black), $m=0.35\;\mbox{GeV}$ (\textcolor{red}{red}), $m=0.5\;\mbox{GeV}$ (\textcolor{magenta}{magenta}), $m=1.0\;\mbox{GeV}$ (\textcolor{blue}{blue}), and $m=1.69\;\mbox{GeV}$ (\textcolor{violet}{violet})~\cite{WL16n}.}
\caption{Application of Louvain method to the BiCM-induced projection of the WTW in the year 2000. The identified communities can be interpreted as representing: \textcolor{Cerulean}{\textbullet} ``advanced'' economies (EU countries, USA and Japan, whose export basket practically includes all products); \textcolor{Orchid}{\textbullet} ``developing'' economies (centro-american countries and south-eastern countries as China, India, Asian Tigers, etc., for which the textile manufacturing represents the most important sector); countries whose export heavily rests upon raw-materials like \textcolor{orange}{\textbullet} oil (Russia, Saudi Arabia, Libya, Algeria, etc.), \textcolor{PineGreen}{\textbullet} tropical agricultural food (south-american and centro-african countries), etc. Australia, New Zealand, Chile and Argentina (whose export is based upon sea-food) happen to be detected as a community on its own.}
\caption{Application of Louvain method to the $\text{BiPCM}_c$-induced projection of the WTW in the year 2000, defined by constraining the products degrees only. The identified larger communities represent: \textcolor{Rhodamine}{\textbullet} fabrics, yarn, etc.; \textcolor{Magenta}{\textbullet} clothes, shoes, etc.; \textcolor{Brown}{\textbullet} wooden products; \textcolor{GreenYellow}{\textbullet} live animals; \textcolor{Cyan}{\textbullet} basic electronics; \textcolor{Cyan}{\textbullet} chemicals; \textcolor{Black}{\textbullet} machinery; \textcolor{Orange}{\textbullet} advanced electronics (all icons are available on \url{http://thenounproject.com/} - see also [41]).}
\caption{Result of the application of Louvain method to the BiCM-induced projection of the MovieLens data set. Since some genres are quite generic, our clusters are often better described by ``combinations'' of genres (readable on the radar-plots beside them) capturing users' tastes to a larger extent: \textcolor{Orange}{\textbullet} movies released in 1996; \textcolor{CadetBlue}{\textbullet} ``family'' movies; \textcolor{Orchid}{\textbullet} movies with marked horror traits; \textcolor{Blue}{\textbullet} ``cult mass'' movies; \textcolor{Magenta}{\textbullet} independent and foreign movies; \textcolor{Yellow}{\textbullet} movies inspired to books or theatrical plays; \textcolor{Cyan}{\textbullet} ``classic'' Hollywood movies (all icons are available on \url{http://thenounproject.com/} - see also [44]).}
\caption{\label{fig:2D3D} Test example for 2D (\emph{top}) and 3D (\emph{bottom}). \emph{From left to right}: moving (atlas) image, target image, deformation result by optimizing LDDMM energy, deformation result using 50 samples from probabilistic network with a stride of 14 and patch pruning, and uncertainty as square root of the sum of the variances of deformation in all directions mapped on the predicted registration result. The colors indicates the amount of uncertainty (\textcolor{red}{red} = high uncertainty, \textcolor{blue}{blue} = low uncertainty). Best viewed in color.}
\caption{\small Summaries from analysis of transitions from node $i$=Homepage $\rightarrow j$=World with Bayesian model monitoring and discount-based intervention. Data is from the February 23rd am period. {\em Upper:} Symbol + indicates observations judged consistent with the standard model; x indicates cases identified as potential outliers by low $H_t$; \red{*} indicates those flagged as potential change points via low $L_t$; \red{o} indicates cases with $l_t>1$. The vertical arrows indicate times of automatic intervention. The full line and shaded region represent one-step forecast means and 95\% intervals. {\em Center:} Tracks of $\log(H_t)$ (above center) and $\log(L_t)$ (below center) over time. {\em Lower:} Data (+) with one-step forecast means and 95\% intervals from the standard BDFM analysis in light gray (red in on-line version) compared to the analysis with monitoring and intervention in black/dark gray (gray in on-line version). }
\caption{The viable parameter space for co-decaying dark matter assuming no cannibalization ({\bf Left}), and a cannibalizing dark sector ({\bf Right}). The central white region shows the range of validity of the model. The different regions show constraints from $N_{\rm eff}$ ({\bf \color[rgb]{.5,0,.5}purple}); DM decays out of equilibrium ({\bf \color[rgb]{0.3,0.3,0.3}gray}); unitarity constraints ({\bf \color[rgb]{0,0.5,0}green}); and indirect detection assuming decays into $ e ^+ e ^- $ ({\bf \color[rgb]{1,0,0}red}/solid) or $ \gamma \gamma $ ({\bf \color[rgb]{0.,0,1}blue}/dashed), excluding the region below the curve. The gap in the $ \gamma \gamma $ limit between $ 10 - 20 \text{ GeV} $ is due to thresholds used in the two recasts. The {\bf \color[rgb]{0.7,0.7,0.7}light gray} dotted lines represent contours of constant $ \sigma $ with values indicated on the right.\label{fig:validity}}
\caption{\label{fig:schematic} Schematic demonstration of our set up. We assume a planet with two equally weighted thermal structures with a cloud-free atmosphere of uniform composition. The fluxes from both thermal profiles are then averaged to create the disk integrated spectrum \ccolor{upon which we perform the retrievals}.}
\caption{\textit{HST} WFC3 + \textit{Spitzer} IRAC 1TP vs. 2TP fit and temperature profiles (insets) retrieval summary. The left panel shows the results for the low (20\%) contrast while the right shows the results for high (80\%) contrast. The data simulated with 2 TP profiles are shown as the black diamonds with error bars (WFC3 between 1 and 2 $\mu$m and the \textit{Spitzer} IRAC points at 3.6 and 4.5 $\mu$m). The fits and temperature profiles are summarized with a median (solid line) and 68\% confidence interval (spread) generated from 1000 randomly drawn parameter vectors from the posterior. Red corresponds to the fits/temperature profiles resulting from a single profile fit, while blue represents the result of including two temperature profiles in the retrieval. The black dashed lines in the temperature profile insets are the two TP profiles used to generate the simulated data (i.e., the ``true" TP profiles). For comparison, we also include the flux-averaged TP profile ($T_{\rm avg}^4 = \frac{1}{2}(T_{\rm day}^4 + T_{\rm night}^4)$), shown as the solid black line in the insets. The dot-dashed TP profile is the coldest profile permitted by the model: a non-irradiated cooling profile governed by the 200K internal temperature. \ccolor{By eye, the 1TP vs. 2TP performances at 20\% contrast are comparable. Based on the Bayesian evidence, the detection of the second profile is not significant ($<\ 0.1 \sigma$). At 80\% contrast, the two retrieved spectra are visibly different. The second profile is detected to $2.4\sigma$ significance.} }
\caption{Summary of the posterior probability distributions of the molecular abundances for the low (20\%, left) and high (80\%, right) contrast cases under the \textit{HST} WFC3+\textit{Spitzer} IRAC observational scenario. The red and blue 1- and 2-D histograms correspond to 1TP and 2TP scenarios. The dashed lines in the 1-D histograms and intersection of the dashed lines in the 2-D histograms are the true molecular abundances used to generate the synthetic data. \ccolor{The detection significance of the second profile from the 2TP retrieval is $<\ 0.1 \sigma$ at 20\% contrast, and the posterior distributions show that invoking a second profile did not improve our abundance estimation. At 80\% contrast, where the detection significance is $2.4\sigma$, we still note the similarities in the posterior distributions for most species. However, in the case of \methane, the 1TP approach, bound by the radiative transfer properties of one profile, overestimates both its abundance and the precision. When we include a second profile, we are able to recover a more realistic and representative distribution for the \methane\, abundance.}}
\caption{\textit{JWST} 1TP vs. 2TP fit and temperature profiles (insets) retrieval summary. The left shows the results for the low (20\%) contrast while the right shows the results for high (80\%) contrast. The data simulated with 2 TP profiles are shown as the black error bars. The fits and temperature profiles are summarized with a median (solid line) and 68\% confidence interval (spread) generated from 1000 randomly drawn parameter vectors from the posterior. Red corresponds to the fits/temperature profiles resulting from a single TP profile fit, while blue represents the result of including two temperature profiles in the retrieval. The black dashed lines in the temperature profile insets are the two TP profiles used to generate the simulated data (e.g., the ``true" TP profiles). For comparison, we also include the flux-averaged TP profile ($T_{\rm avg}^4 = \frac{1}{2}(T_{\rm day}^4 + T_{\rm night}^4)$), shown as the solid black line in the insets. \ccolor{At 20\% contrast, while the retrieved fits appear similar, we find that the second TP profile is detected to $\sim 5\sigma$ significance. At 80\% contrast, the 1TP retrieved spectra poorly fit the data, especially at $2-3\ \mu$m and at longer wavelengths. }}
\caption{Summary of the posterior probability distributions of the molecular abundances for the low (20\%, left) and high (80\%, right) contrast cases under the \textit{JWST} observational scenario. The red and blue 1- and 2-D histograms correspond to 1TP and 2TP scenarios. The dashed lines in the 1-D histograms and intersection of the dashed lines in the 2-D histograms are the true molecular abundances used to generate the synthetic data. \ccolor{When the contrast is 20\%, the second profile is detected to $\sim 5\sigma$. When the contrast is 80\%, the second profile is detected to $>\ 20 \sigma$. We see that, at higher contrasts, the 1TP retrieval case is a poor representation of the abundances. We also note the over-constraint of \ammonia\, under the 1TP prescription. This behavior is analogous to the\methane\, abundance inference using one profile that we saw with WFC3+IRAC data. Once a second profile is included, we recover the true abundance of\ammonia.} }
\caption{\label{fig1}{\bf Structure and thermodynamic properties of YbMgGaO{$_4$}}. {\bf a.} Partial crystal structure, showing a triangular layer of Yb$^{3+}$ ions (large cyan spheres) and their coordination by oxygen (small red spheres). A nearest-neighbor interaction pathway $J_1$ and a next-nearest-neighbor interaction pathway $J_2$ are shown by yellow and blue lines, respectively. {\bf b.} Magnetic component of the specific heat $C_{\mathrm{m}}(T)$, showing data measured on a powder sample (hollow squares) and a single-crystal sample (filled circles). Single-crystal data are shown for applied magnetic fields along the $c$ axis of 0, 4, 7.8, and 14\,T (labelled above each curve). The orange line shows a fit of the zero-field single-crystal data to a power law,$C_{\mathrm{m}}(T)\propto T^{0.703(4)}$. {\bf c.} Magnetic entropy change $\Delta S_{\mathrm{m}}$, showing data measured in zero field and in a 7.8\,T field (labelled on each curve).{\bf d.} Dependence of the magnetization $M$ on applied field $\mu_0 H$, showing data measured at temperatures of 1.7, 5, and 10\,K (labelled on each curve). In{\bf b}, {\bf c} and {\bf d} the temperatures and applied fields at which we performed neutron-scattering measurements [Figs.~2 and 3] are indicated by small grey arrows. }
\caption{\label{fig2}{\bf Neutron-scattering data for YbMgGaO{$_4$} measured in zero applied field.}{\bf a.} Energy dependence of magnetic excitations along high-symmetry directions in reciprocal space, showing data at 0.06\,K (upper panel) and 14\,K (lower panel). Reciprocal-space points are labeled in{\bf b} and scattering intensity in arbitrary units is shown as a color scale. \red{The temperature of 0.06 K was measured at the mixing chamber of our dilution refrigerator.}{\bf b.} Wave-vector dependence of magnetic excitations at 0.06\,K, at energy transfers of 0.25\,meV (upper panel) and 0.75\,meV (lower panel).{\bf c.} One-dimensional cuts along high-symmetry directions at 0.06\,K (upper panel) and 14\,K (lower panel). Each panel shows energy transfers of 0.25, 0.45, and 0.65\,meV (labeled on the graph) shifted vertically for clarity.\red{For {\bf a} and {\bf c}, the incident neutron energy $E_\mathrm{i}=3.32$\,meV and the data have been integrated over the vertical range$-\frac{1}{2} \leq l \leq \frac{1}{2}$; for {\bf b}, $E_\mathrm{i}=12$\,meV and$-1 \leq l \leq 1$. The scattering intensity in the top panel of {\bf b} is multiplied by a factor $2/3$ to share the same color scale as {\bf a}.}}
\caption{\label{fig3}{\bf Field-polarized neutron-scattering data and evidence for next-nearest-neighbor interactions in YbMgGaO{$_4$}}. {\bf a.} Energy dependence of magnetic excitations along high-symmetry directions, measured at 0.06\,K in an applied field of 7.8\,T. The white circles show the location of the maximum intensity at each wave-vector (see text). The red lines show a fit to the spin-wave dispersion relation, Eq.~(2). The labeling of reciprocal-space positions is given in Fig.~2{\bf b}. The rightmost panel, for which \red{H=$(\frac{1}{3}\frac{1}{3}\frac{1}{2})$}, demonstrates the absence of dispersion along $l$ (perpendicular to the triangular planes). {\bf b.} One-dimensional plots of the 0.06\,K magnetic intensity, showing that the scattering peaks\red{around} the zone centers when a 7.8\,T field is applied. The range of$E$-integration is labeled on the graph. {\bf c.} Representative fit to the $E$-dependence of the data shown in {\bf a}, used to determine the position of the intensity maximum at a given wave-vector [white circles in (a)]. {\bf d.} Dependence of the goodness-of-fit parameter $\chi^2$ on the values of the exchange interactions $J_{1}^{\pm\pm}$ and $J_{2}^{\pm}$. {\bf e.} Experimental diffuse-scattering data at 0.06\,K obtained by integrating over energy transfer (top panel); calculation from classical Monte Carlo simulations of our model with nearest and next-nearest neighbor interactions (bottom left panel); and calculation from a phenomenological model of nearest-neighbor and next-nearest-neighbor dimers (bottom right panel).{\bf f.} Calculated diffuse scattering for models with nearest-neighbor interactions only, showing Monte Carlo calculation (left panel) and dimer model (right panel). The data in {\bf a} and {\bf c} were measured with $E_\mathrm{i}=3.32$\,meV, and the data in{\bf b} and {\bf e} with $E_\mathrm{i}=12$\,meV.}
\caption{Temperature dependence of magnetic diffuse scattering calculated from classical Monte Carlo simulations of the XXZ model with nearest and next-nearest neighbor interactions, showing 1.3\,K (left panel) and 0.7\,K (right panel). Intensities for the left panel are multiplied by 2 compared to the right panel. The exchange constants are$J_{1}^{zz}=0.126$, $J_{2}^{zz}=0.027$, $J_{1}^{\pm}=0.109$, and $J_{2}^{\pm}=0.024$\,meV, with$J_{1}^{\pm\pm}=0$.}
\caption{Influence of the \textbf{number of faces} on the \textcolor{red}{\textbf{Sensitvity}}, the \textcolor{blue}{\textbf{Linearity range}} and the \textbf{SD factor} with respect to the spatial frequencies in terms of Zernike Radial Orders. Left insert: small apex angle ($\theta=0.1\frac{D}{2f}$). Right insert: large apex angle ($\theta=2\frac{D}{2f}$). \label{1}}
\caption{Influence of the \textbf{apex angle} on the \textcolor{red}{\textbf{Sensitvity}}, the \textcolor{blue}{\textbf{Linearity range}} and the \textbf{SD factor} with respect to the spatial frequencies in terms of Zernike Radial Orders. Number of faces equals to 3. Left insert: no modulation. Right insert: circular modulation ($r_m=3 \lambda/D$). \label{2}}
\caption{Influence of the \textbf{modulation radius} on the \textcolor{red}{\textbf{Sensitvity}}, the \textcolor{blue}{\textbf{Linearity range}} and the \textbf{SD factor} with respect to the spatial frequencies in terms of Zernike Radial Orders. Number of faces equals to 3. The modulation is \textbf{circular}. Left insert: small apex angle ($\theta=0.1\frac{D}{2f}$). Right insert: large apex angle ($\theta=2\frac{D}{2f}$). \label{3}}
\caption{Influence of the \textbf{shape of the modulation path} on the \textcolor{red}{\textbf{Sensitvity}}, the \textcolor{blue}{\textbf{Linearity range}} and the \textbf{SD factor} with respect to the spatial frequencies in terms of Zernike Radial Orders. Number of faces equals to 4. Maximum modulation radius equals to $3\lambda/D$ Left insert: small apex angle ($\theta=0.1\frac{D}{2f}$). Right insert: large apex angle ($\theta=2\frac{D}{2f}$). \label{4}}
\caption{Influence of the \textbf{weighting function} on the \textcolor{red}{\textbf{Sensitvity}}, the \textcolor{blue}{\textbf{Linearity range}} and the \textbf{SD factor} with respect to the spatial frequencies in terms of Zernike Radial Orders. Number of faces equals to 4. The modulation is \textbf{circular}. $r_m=3\lambda/D$ Left insert: small apex angle ($\theta=0.1\frac{D}{2f}$). Right insert: large apex angle ($\theta=2\frac{D}{2f}$). \label{5}}
\caption{ Comparison of the elevation estimation by humans and the proposed methods (CNN, BOW and combination). %Green crosses show the ground truth. %Blue circles are the CNN predictions. Blue boxes show the span of the human predictions: the red mark is the median, the edges of the box are the $25^{\mathrm {th}}$ and $75^{\mathrm {th}}$ percentiles respectively, the whiskers extend to extreme human guesses that are not considered outliers, and outliers are plotted individually as '\textcolor{red}{+}'. %\todo{full width of the page}\todo{Maybe its better to write out the horizontal and vertical axes rather than describing them in the caption?} }
\caption{\textcolor{red}{Comparison of the NDCG@5 for our algorithm and \cite{bb:state_2} evaluated for different categories in the MIRFlickr collection.}}
\caption{\small Overview of our approach for a layer 5 filter. Each local maxima of the filter's feature map leads to a stimulus detection (\textcolor{red}{red}). We transform each detection with a regressor trained to map it to a bounding-box tightly covering a semantic part (\textcolor{green}{green}).}
\caption{\small Examples of stimulus detections for layer 5 filters. For each part class we show a feature map on the left, where we highlight the strongest activation in red. On the right, instead, we show the corresponding original \textcolor{red}{receptive field} and the \textcolor{green}{regressed box}.}
\caption{\textcolor{blue}{Venn diagram}: The information theoretic quantities for three random variables, $X$, $Y$, and $Z$. The total correlation, $I(X:Y:Z)=I_s(X:Y:Z)+I_0(X:Y:Z)$, and the binding information, $I_s(X:Y:Z)=I(X:Y|Z)+ I(X:Z|Y)+I(Y:Z|X)+I_0(X:Y:Z)$, where $I_0(X:Y:Z)$ is the interaction information.}
\caption{The figure shows how the dissension vectors behave as a function of mixing parameter, $p$ for the four-qubit state, $\rho_{\Omega}$ when exposed with colored (\textcolor{red}{red} dashed line), and white noise (black solid line). The subfigures [$(i)-(v)$] depict the behaviour of the dissension with subfigures $(i)$ depicts $\vec{\delta}_1^1$; $(ii)$ depicts $\vec{\delta}_2^1$; $(iii)$ depicts $\vec{\delta}_3^1$; $(iv)$ depicts $\vec{\delta}_1^2$ and $(v)$ depicts $\vec{\delta}_2^2$. (Note that we have plotted one of the elements from each dissension vectors. This is because within a vector each elements are same as the state is symmetric.)}
\caption{A peek into SH-DPP. Given the query \textcolor{red}{\textsc{flowers}}+\textcolor{red}{\textsc{wall}}, the $Z$-layer of SH-DPP is supposed to summarize the shots relevant to the query. Conditioning on those results, the $Y$-layer summarizes the remaining video.}
\caption{Graph topology. Rectangular nodes are tweets, circular nodes are users and the diamond represents the world. Some tweet nodes are \textcolor{green}{labelled} with an initial distribution over language labels and others are \textcolor{red}{unlabelled}.}
\caption{ \label{fig:T} \red Estimates of the halo temperature $T$ and X-ray luminosity $L_{X}$ versus the density profile parameters. The red diamond and the red contour represent, respectively, the best-fit point and the 68\% CL contour obtained in this work. The right scale represents $T$ estimated from $\beta$. Blue horizonthal lines give the $T$ median value (full line, $2.22 \times 10^{6}$~K) and interquantile range (dashed lines, $0.63 \times 10^{6}$~K) from \citet{1306.2312}. Thin gray lines bound the range $L_{X}=(2\dots 3)\times 10^{39}$~erg/s favoured by \citet{ApJ_485_125,a-p/9710144}. See the text for details and important notes. \black }
\caption{Off-shell strong coupling $g_{\eta_c\eta_c\psi}(x)$ as function of $x\equiv\frac{q^2}{M_R^2}$ for $\eta_c\succ\eta_c$ (\textcolor{red}{red}) or $\eta_c\succ J/\psi$ (\textcolor{blue}{blue}) transitions.}
\caption{Off-shell strong couplings $g_{DD\psi}$ and $g_{DD^*\eta_c}$ to $D^{(*)}$ mesons (with resonances indicated by circumflexes): dependences on $x\equiv\frac{q^2}{M_R^2}$ of (a) $g_{D\hat D\psi}(x)=\frac{2\,M_\psi}{f_D}\,(1-x)\, A^{D\succ\psi}_0(q^2)$ (\textcolor{blue}{blue}) and $g_{DD\hat\psi}(x)=\frac{2\,M_\psi}{f_\psi}\,(1-x)\,F_+^{D\succ D}(q^2)$ (\textcolor{red}{red}),~got with (lines) and without (symbols) interpolation, and of (b) $g_{D\hat{D^*}\eta_c}(x)$ (\textcolor{red}{red}), $g_{DD^*\hat{\eta_c}}(x)$ (\textcolor{blue}{blue}) and $g_{\hat DD^*\eta_c}(x)$ (\textcolor{green}{green}).}
\caption{Summary of Average Overlap Scores (AOS) results for all 12 trackers. The best and second best results are in {\color[HTML]{32CB00} \textbf{green}} and {\color[HTML]{FE0000} \textbf{red}} colors, respectively.}
\caption{\label{fig:Pth}Distribution $P(\vartheta)$ in trimers \Hep\and\Hem\of corner angle$\vartheta$, in top subfigure $\measuredangle\left(\eHe{4}-\eHe{4}-\eHe{4} \right)$, in middle $\measuredangle\left(\eHe{4}-\eHe{4}-\eHe{3} \right)$, and in bottom $\measuredangle\left(\eHe{4}-\eHe{3}-\eHe{4} \right)$ using different potential models (section~\ref{ch:method-V}) and compared with renormalized measured values [a]=\cite{NC}. All distributions are normalized to $\int P(\vartheta) \rmd \vartheta = 1$.} \end{figure} Angular distribution functions are presented for different potentials in figure~\ref{fig:Pth}. The top subfigure shows the distribution $P(\vartheta)$ of corner angle $\vartheta=\measuredangle\left(\eHe{4}-\eHe{4}-\eHe{4}\right)$, middle of $\vartheta=\measuredangle\left(\eHe{4}-\eHe{4}-\eHe{3} \right)$ and bottom of $\vartheta=\measuredangle\left(\eHe{4}-\eHe{3}-\eHe{4}\right)$. Experimental data are taken from~\cite{NC}, but here normalized to $1$, and shown with error bars. Error bars of the theoretical data (this work) are of the same order of magnitude as the line width. In the top subfigure all theoretical estimates are almost the same. The most significant difference with experiment is the sharp peak of experimental data that neither theoretical model predicts. In the middle subfigure, experimental data are scattered around theoretical estimates. Even the reduction of $C_6$ by $2{\rm \%}$ makes no significant difference in the predictions of $P(\vartheta)$, while differences are clearly pronounced in the case of $P(r)$ (see figure~\ref{fig:P444}). From visual inspection it is not possible to conclude which potential model leads to the distribution that fits better the experimental data. In the bottom subfigure, similar behavior is noticeable, but with more scattered experimental data. Again neither theoretical model predicts the peak to be as sharp as extracted~\cite{NC} from experimentally measured data. Recently, theoretical estimates of $P(\vartheta)$ also for the models TTY~\cite{Bress2014} and SAPTSM~\cite{Suno2015} have been published. When comparing, one needs to be careful due to wrong normalizations. When properly renormalized, agreement with present results is obtained. In order to numerically evaluate which potential model makes better predictions we chose $P_i=P(r_i)$ in \Hep, because only these experimental data $P^{{\rm exp}}$ are given~\cite{NC} with known norm and the smallest error bars $\sigma(P^{{\rm exp}}_i)$ relative to the differences between our theoretical model predictions $P^{{\rm dmc}}(r)$. We made numerical estimates of differences $\Delta P$ between experimental and theoretical predictions. Different definitions of differences were used \begin{eqnarray} \langle\Delta P\rangle &=&\frac{1}{n}\sum_i^n (P^{{\rm dmc}}_i - P^{{\rm exp}}_i)^2 \label{eq:P2}\\ \langle\Delta P\rangle_\sigma &=&\frac{1}{n}\sum_i^n \frac{(P^{{\rm dmc}}_i - P^{{\rm exp}}_i)^2}{\sigma(P^{{\rm exp}}_i)} \label{eq:P3} \\ \int{\Delta} P &=&\sum_i^n \left|P^{{\rm dmc}}_i - P^{{\rm exp}}_i\right|\cdot \frac{r_{i+1} - r_{i-1}}{2} \label{eq:P4} \end{eqnarray} where a linear interpolation was used to calculate $P^{{\rm dmc}}_i$ in each experimental point $r_i$. Digitalized experimental data are not good enough in areas where symbols, errobars and axes cannot be clearly distinguished, so we approximated $P^{{\rm dmc}} \approx P^{{\rm exp}}$ for $r>30 \angstrom$. Definitions \eref{eq:P2}, \eref{eq:P3} and \eref{eq:P4} have returned the same ascending sorted list of potential models (from the best to the worst): HFDB-1cC6, HFDB-5mC6, TTY, HFDB-2cC6, HFDB+V3AT, SAPTSM, SAPTSM+Ret, HFDB. Calculated values are given in table~\ref{tab:DP}. A similar list could be obtained just by visual comparison of different results in figure~\ref{fig:P444}. \begin{table}[t] \caption{\label{tab:DP}Differences $\Delta P$, between DMC($V$) pure estimates and experimental data digitalized from figure 1c in~\cite{NC}, calculated using \eref{eq:P2}, \eref{eq:P3} and \eref{eq:P4}.} \centering \begin{tabular}{@{}l@{}c@{}c@{}c@{}} \br Potential ($V$) & $10^{6}$\AA$^2\langle\Delta P\rangle$ & $\quad10^{3}$\AA$\langle\Delta P\rangle_\sigma$ & $\quad10^{2}\int{\Delta} P$\\ \mr HFDB-1cC6 & 1.87 & 2.22 & 2.66\\ HFDB-5mC6 & 4.62 & 4.89 & 4.22\\ TTY & 5.55 & 5.68 & 4.70\\ HFDB-2cC6 & 6.07 & 7.10 & 5.62\\ HFDB+V3AT & 8.14 & 8.35 & 5.87\\ SAPTSM & 9.03 & 9.19 & 6.22\\ SAPTSM+Ret & 9.08 & 9.26 & 6.25\\ HFDB & 10.1 & 10.2 & 6.60\\ \br \end{tabular} \end{table} \subsection{Universal scaling} In a previous work, we established~\cite{Unihalo} both the more convenient energy-size scaling and the universal lines which trimer halo states do follow. The size of a system was measured~\cite{Unihalo} by the root-mean-square hyperradius $\rho$, \begin{equation} m\rho^2 = \frac{1}{M}\sum_{i<k}m_im_k\langle r_{ik}^2 \rangle \,\label{hyper1} \end{equation} where $m$ is an arbitrary mass unit, $m_i$ the particle mass of species $i$, $M$ the total mass of the system, and $\langle r_{ik}^2 \rangle$ the mean square distance between particles $i$ and $k$. Values of $\langle r_{ik}^2 \rangle$ extracted by the pure estimators from the DMC-sampled positions are given in table~\ref{tab:r2}. \begin{table}[h!] \caption{\label{tab:r2}Mean square distances $\langle r_{44}^2 \rangle$ between $^4$He-$^4$He and $\langle r_{43}^2 \rangle$ between $^4$He-$^3$He in clusters \Hep\and\Hem. Standard deviations are $1$-$3$\% of the corresponding quantity.} \centering \begin{tabular}{@{}lc@{~}cc} \br &\Hep & \multicolumn{2}{c}{\Hem}\\ \ns Potential & \crule{1}& \crule{2}\\ & $\langle r_{44}^2 \rangle$ / \AA$^2$ & $~\langle r_{44}^2 \rangle$ / \AA$^2$ & $\langle r_{43}^2 \rangle$ / \AA$^2$\\ \mr SAPTSM & 116 & 330 & 540 \\ HFDB & 117 & 326 & 560 \\ HFDB+V3AT & 117 & 336 & 550 \\ SAPTSM+Ret & 117 & 356 & 590 \\ HFDB+Ret & 118 & 360 & 608 \\ TTY & 120 & 368 & 635 \\ TTY+Ret & 121 & 384 & 663 \\ HFDB-5mC6 & 123 & 404 & 680 \\ HFDB-1cC6 & 128 & 482 & 884 \\ \br \end{tabular} \end{table} \begin{figure}[h!] \centering \includegraphics{figure7} \caption{\label{fig:Uni}Absolute binding energy $B$ and size of trimers \Hep\and\Hem\are calculated by DMC-pure estimators using different potential models (section~\ref{ch:method-V}) and scaled using definitions $X_E$ and $Y_\rho$ given in~\cite{Unihalo}. Different symbols are used to distinguish potential models. Universal lines for Borromean, tango and all-bound trimer type are fitted through data given in~\cite{Unihalo}. Halo states are defined by the condition~\cite{RMPhalos} $Y_\rho \gtrsim 2$.} \end{figure} The mean square distance between $^{4}$He atoms in \Hep, obtained by the most recent \VBO\potential SAPTSM is in agreement with the value given in~\cite{HK2012}, estimated using the most detailed post\VBO\potential, PCKLJS. Universal lines from~\cite{Unihalo} are shown in figure~\ref{fig:Uni}. All-bound trimer type is presented by a dotted line which, when binding is decreased, passes into the Borromean type presented by solid line. Dashed line shows a departure of the tango trimer type from the joint universal line. Symbols representing scaled energy $X_E$ and size $Y_\rho$ for both helium trimers, obtained using different potential models fall on the universal lines plotted in figure~\ref{fig:Uni}. Results of all presented models predict both helium trimers \Hep\and\Hem\to be in a halo state and to follow the universal line, being spatially wider and more weakly bound the less attractive the potential is.\section{Summary and conclusions}\label{ch:conclusions} Many semi-empirical and ab initio potential models have been proposed for the interaction of helium atoms. Considering their diversity, we made a rather complete set of tests on how the interaction potentials, and their corrections, influence the ground state binding energy and structural properties of small helium clusters. The clusters most sensitive to the changes in the interaction potential were considered, dimers and trimers. The DMC method, which was used to calculate the trimer properties, gave exact values of their studied properties within statistical error bars. The achieved binding energy statistical errors $\sigma$ were few times smaller than errors caused by the SAPTSM-uncertainty $\varsigma$. Our predictions are in excellent agreement with the most recent estimates, obtained with various methods, confirming their accuracy. Structural properties were determined with $\sigma$ approximately equal to $\varsigma$. For some models, like HFDB+V3AT, HFDB+V3BM, HFDB+V3DDDJ, TTY+Ret, HFDB+Ret, HFDB-2cC6, HFDB-1cC6, and HFDB-5mC6, we made estimates of the trimer properties for the first time. In particular, the influence of the error bar $\varsigma$ of the newest and most sophisticated He-He \VBO\estimate SAPTSM and its most significant correction Ret on helium trimer structural properties, was analyzed for the first time. Influence of other post-\VBO\SAPTSM-corrections ARQ embedded in the PCKLJS model were not considered because they are smaller than Ret, and therefore give trimer properties close to estimates obtained with SAPTSM and SAPTSM+Ret, which already are not so different. Furthermore, we estimated the influence of attenuated HFDB potential by reducing dispersion coefficient$C_6$ for $2$\% (-2cC6), $1$\% (-1cC6) and $0.5$\% (-5mC6). Among structural properties calculated in this work, angular distributions $P(\vartheta)$ are the least affected by the potential model; differences are barely visible. On the one hand, measured $P(\vartheta)$ are the most cascade-like among experimental data given in~\cite{NC}. Therefore measured values cannot be used to evaluate potential models. However, even if we had very precise measurements of angular distributions, we could not use them to rate potential corrections, because many models give similar theoretical predictions. According to theoretical estimates, some potential models, attenuations of $C_6$ and correction Ret could be distinguished from the density profiles $\rho(r)$ with respect to the center of mass and from the distributions of interparticle separations $P(r)$. Unfortunately, there are no measured values of $\rho(r)$, but there are some of $P(r)$. From visual comparison, specifying indistinguishable models as one set, we can sort potential models from the lowest to the highest correlation peak of $P(r)$ in \Hep: \{HFDB-2cC6\}, \{Experiment~\cite{Sci}\}, \{Experiment~\cite{NC}, HFDB-1cC6\}, \{HFDB-5mC6, TTY+Ret\}, \{TTY\}, \{HFDB+Ret, SAPTSM+Ret, HFDB+V3AT\}, \{HFDB, SAPTSM\}. Differences are more clear when $P(r)$ are compared in \Hem; only the effect of the three-body correction V3AT becomes invisible. Unfortunately there are no normalized experimental distributions in \Hem, so comparison is made setting all peaks to $1$. In that case HFDB-5mC6 fits the experimental data best. But this adjustment is somehow unnatural because distribution differences between models become significant in some areas where they are equal when normalized to $1$. As expected, similar comparison between theoretical models follows from the mean square interparticle distances. In order to go beyond a simple visual comparison, we evaluated measured-calculated differences $\Delta P$ of $P(r)$ in \Hep. In this way we got a sorted list of potential models (from the best to the worst predictor according to table~\ref{tab:r2}): HFDB-1cC6, HFDB-5mC6, TTY, HFDB-2cC6, HFDB+V3AT, SAPTSM, SAPTSM+Ret, HFDB. However, the first model HFDB-1cC6 significantly underestimates dimer binding energies $E_2^{'}$ and $E_2^{''}$ which follow from the two analysis~\cite{He2,ARQ2} \eref{eq:E2} and \eref{eq:E2Cen} of the experimental data~\cite{He2}. The second best HFDB-5mC6 predicts $P(r)$ in \Hep\up to the error bar equal to the TTY+Ret results. Although HFDB-5mC6 and TTY+Ret predict distinguishable dimer binding energies, due to the large error bars in values\eref{eq:E2} and \eref{eq:E2Cen}, it is not possible to state which is better. All our theoretical estimates predicted both helium trimers, all-bound type \Hep\and tango type\Hem, to be in a ground halo state, although recent articles~\cite{NC,Sci2} mention \Hem\and only excited state of\Hep\as a halo. Both are structureless clouds. However, the less bound\Hem\is wider and more spread among different shapes (linear, isoceles, scaline, equilateral). With development of methods and increase of computer power, theoretical estimates of helium cluster properties have become very accurate and efficient. Theoretical uncertainties are more than an order of magnitude smaller than experimental ones. Furthermore, the discrepancies between computed and measured values are a few times larger than the theoretical uncertainties. Therefore, a higher precision of experimental measurements would be welcomed to derive a more accurate rating of theoretical models. The temperature could affect the measured values and these effects are not taken into account in our theoretical estimates. Additionaly, the ground state of\Hep\could be contaminated by a fraction of the excited state, which could explain differences between two~\cite{NC,Sci} experimental measurements. In conclusion, the whole set of available measured and deduced values, from experimental helium dimer and trimers data, is in the best agreement with the theoretical predictions obtained using the potential models TTY+Ret and HFDB-5mC6, which are up to the error bars equal. \ack \vspace{-12pt} This work has been supported in part by the Croatian Science Foundation under the project number IP-2014-09-2452. J. B. acknowledge additional support by the MICINN-Spain, Grant No. FIS2014-56257-C2-1-P. The computational resources of the Isabella cluster at Zagreb University Computing Center (Srce), the HYBRID cluster at the University of Split, Faculty of Science and Croatian National Grid Infrastructure (CRO NGI) were used. \section*{References} \begin{thebibliography}{42} \bibitem{microSupra} Toennies J P, Vilesov A F and Whaley K B 2001 {\it Physics Today} {\bf 54} 31 \bibitem{HeRev} Toennies J P 2013 {\it Mol Phys} {\bf 111} 1879 \bibitem{4HeN} Whaley K B 1994 {\it Int Rev Phys Chem} {\bf 13} 41 \bibitem{3HeM} Sola E, Casulleras J and Boronat J 2006 \PR B {\bf 73} 092515 \bibitem{GN2003} Guardiola R and Navarro J 2003 \PR A {\bf 68} 055201 \bibitem{BMHeHe} Bressanini D and Morosi G 2004 {\it Few-Body Systems} {\bf 34} 131 \bibitem{He2bond} Luo F, McBane G, Kim G and Giese C F 1993 \JCP{\bf 98} 3564 \bibitem{He2size} Luo F, Giese C F and Gentry W R 1996 \JCP{\bf 104} 1151 \bibitem{He23dif} Schöllkopf W and Toennies J P 1994 {\it Science} {\bf 266} 1345 \bibitem{He23diffJCP} Schöllkopf W and Toennies J P 1996 \JCP{\bf 104} 1155 \bibitem{He2} Grisenti R E, Schöllkopf W, Toennies J P, Hegerfeldt G C, Köhler T and Stoll M 2000 \PRL{\bf 85} 2284 \bibitem{HeHe_exp} Kalinin A, Kornilov O, Schöllkopf W and Toennies J P 2005 \PRL{\bf 95} 113402 \bibitem{NC} Voigtsberger J \etal 2014 {\it Nature Communications} {\bf 5} 5765 \bibitem{RMPhalos} Jensen A S, Riisager K, Fedorov D V and Garrido E 2004 \RMP{\bf 76} 215 \bibitem{riisager} Riisager K 2013 \PS 2013 {\bf T152} 014001 \bibitem{Unihalo} Stipanovi\'{c} P, Vranje\v{s} Marki\'{c} L, Be\v{s}li\'{c} I and Boronat J 2014 \PRL{\bf 113} 253401 \bibitem{Efimov} Efimov V 1970 \PL{\bf 33B} 563 \bibitem{Sci} Kunitski M \etal 2015 {\it Science} {\bf 348} 551 \bibitem{Sci2} Kornilov O 2015 {\it Science} {\bf 348} 498 \bibitem{V3AT} Axilrod B M and Teller E 1943 \JCP 11 {\bf 299} \bibitem{CP} Casimir H B G and Polder D 1948 \PR{\bf 73} 360 \bibitem{V3BM} Bruch L W and McGee I J 1973 \JCP{\bf 59} 409 \bibitem{HFDB} Aziz R A, McCourt F R W and Wong C C K 1987 {\it Mol Phys} {\bf 61} 1487 \bibitem{TTY} Tang K T, Toennies J P and Yiu C L 1995 \PRL{\bf 74} 1546 \bibitem{V3DDDJ} Ujevic S and Vitiello S A 2006 \PR B {\bf 73} 012511; Cohen M J and Murrell J N 1996 {\it Chem Phys Lett} {\bf 260} 371 \bibitem{SAPTSM} Jeziorska M, Cencek W, Patkowski B, Jeziorski B and Szalewicz K 2007 \JCP{\bf 127} 124303 \bibitem{ARQ} Przybytek M, Cencek W, Komasa J, Łach G, Jeziorski B and Szalewicz K 2010 \PRL{\bf 104} 183003; See also supplementary material at \url{http://linkapsorg/supplemental/101103/PhysRevLett104183003} and the errata 2012 \PRL{\bf 108} 129902 \bibitem{ARQ2} Cencek W, Przybytek M, Komasa J, Mehl J B, Jeziorski B and Szalewicz K 2012 \JCP{\bf 136} 224303 \bibitem{HeT} Stipanovi\'{c} P, Vranje\v{s} Marki\'{c} L, Boronat J and Ke\v{z}i\'{c} B 2011 \JCP{\bf 134} 054509 \bibitem{DMC2} Boronat J and Casulleras J 1994 \PR B {\bf 49} 8920 \bibitem{pure} Casulleras J and Boronat J 1995 \PR B {\bf 52} 3654 \bibitem{Roudnev2012} Roudnev V and Cavagnero M 2012 \jpb{\bf 45} 025101 \bibitem{jensen} Jensen A S, Riisager K, Fedorov D and Garrido E 2004 \RMP{\bf 76} 215 \bibitem{tango} Robicheaux F 1999 \PR A {\bf 60} 1706 \bibitem{samba} Yamashita M T, Tomio L and Frederico T 2004 {\it Nuclear Phys} A {\bf 735} 40 \bibitem{Bress2014} Bressanini D 2014 \JPhCh A {\bf 118} 6521 \bibitem{BH2003} Braaten E and Hammer H-W 2003 \PR A {\bf 67} 042706 \bibitem{Suno2013} Suno H, Hiyama E and Kamimura M 2013 {\it Few-Body Systems} {\bf 54} 1557 \bibitem{Suno2015} Suno H 2016 \jpb{\bf 49} 014003 \bibitem{V3C} Cencek W, Jeziorska M, Akin-Ojo O and Szalewicz K 2007 \JPhCh A {\bf 111} 11311 \bibitem{HK2012} Hiyama E and Kamimura M 2012 \PR A {\bf 85} 062505 \bibitem{Bress2000} Bressanini D, Zavaglia M, Mella M and Morosi G 2000 \JCP{\bf 112} 717 \end{thebibliography} \end{document} }
\caption{\babar\NLO measurements: Vector dominance model (VDM) fits of the squared form-factors using a Gounaris-Sakurai (GS) parametrization. Left:$\pip\pim$ \cite{Aubert:2009ad,Lees:2012cj}. Right: $\Kp\Km$ \cite{Lees:2013gzt}. \label{fig:VDM} }
\caption{Contributions to $a_\mu^{\VP}$ for recent \babar\publications: comparison of the measured value to the previous world average on the energy range$\sqrt{s'}< 1.8\,\gev$ (units $10^{-10}$). \label{tab:comparison} }
\caption{We learn \textbf{low-level representations} specific for each modality (white and grays) and a \textcolor{red}{\textbf{high-level representation}} that is shared across all modalities (red). Above, we also show masks of inputs that activate specific units the most \cite{zhou2014object}. Interestingly, although the network is trained without aligned data, units emerge in the shared representation that tend to fire on the same objects independently of the modality.\vspace{-1em}}
\caption{Average PSNR(SSIM) comparison on three test datasets among different methods. \R{Red} and {\color{blue}{blue}} colors indicate the best and the second best performance.}
\caption{As the number of verbs increases from \textcolor{cmuColour}{12} to \textcolor{beoidColour}{75}, the best performance changes from SVM to SEMBED. Results are obtained with $\gamma_{fv} = 10$ and $\gamma_{bow} = 256$, $k=$\{\textcolor{cmuColour}{3},\textcolor{gteaColour}{5},\textcolor{beoidColour}{5}\}, $m$ = 240, $z$=\{\textcolor{cmuColour}{2},\textcolor{gteaColour}{6},\textcolor{beoidColour}{4}\}, $t$=\{\textcolor{cmuColour}{20},\textcolor{gteaColour}{20},\textcolor{beoidColour}{8}\} for CNN and $z$=\{\textcolor{cmuColour}{4},\textcolor{gteaColour}{5},\textcolor{beoidColour}{14}\}, $t$=\{\textcolor{cmuColour}{4},\textcolor{gteaColour}{20},\textcolor{beoidColour}{10}\} for IDT. For completion, state-of-the-art results on verb-noun classes are reported under `Other Works' thus are not directly comparable to our verb only results.}
\caption{ Influence of the variational model (a) and the non-local neighborhood size $\nbh$ (b) on the \gls*{rmse}. Best results highlighted in \textcolor{clrfirst}{orange} and second best in \textcolor{clrsecond}{yellow}. }
\caption{ Quantitative results on noisy Middlebury data: We present our results on the disparity maps of the noisy Middlebury dataset~\cite{park11} as \gls*{rmse} of the disparity values. Best results highlighted in \textcolor{clrfirst}{orange} and second best in \textcolor{clrsecond}{yellow}. }
\caption{ Quantitative and qualitative results on the \tm~\cite{ferstl13} benchmark. In (a) we present our quantitative results as \gls*{rmse} in $mm$. Best results highlighted in \textcolor{clrfirst}{orange} and second best in \textcolor{clrsecond}{yellow}. In (b) and (c) we show the results of the \gls*{fcn} and the full model, respectively. For comparison, we also show in (d) the ground-truth \gls*{hr} depth. }
\caption{VPHAS$+$ index $r-H\alpha$ vs.\color$r-i$. Symbols as in Fig.~\ref{sung-halpha}. \label{vphas-excess}}
\caption{Ratio $L_X/L_{bol}$ vs.\stellar mass. The two dotted lines indicate$\log L_X/L_{bol} =-3$ and $-4$, respectively. \label{lxlbol-mass-ir}}
\caption{Mass vs.\X-ray count rate for PMS members. The dashed line is obtained from a principal-component analysis. Only non-flaring X-ray sources are shown, within 4~arcmin from cluster center.\label{mass-xrate}}
\caption{(color online). Exact solutions (lines) and TMRG results (symbols) of the specific heat $C_\nu$ are plotted as a function of temperature $T$ for a variety of $\theta$. Inset shows asymptotic behavior of the specific heat at low temperature for two different cases: 1) $\theta=\pi/4$ ({\magenta $\triangledown$}) which is in the TLL phase, the specific heat is proportional to $T$ and 2) $\theta=\pi/8$ ({\red $\square$}) which is the transition point from the TLL phase to the gapped FM phase, the specific heat behaves as $\sqrt{T}$.}
\caption[]{Development set frame error rates vs. number of epochs (left) and vs. training real-time factor (time spent training / duration of training set, right). \label{fig:frame} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {vanilla}; \draw[blue] (-0.2, 0) -- (0, 0); \draw[blue] (0, 0) -- (0.2, 0); \filldraw[fill=blue,draw=black] (-0.07, -0.04) -- (0.07, -0.04) -- (0, 0.08) -- cycle; \end{tikzpicture} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {dropout}; \draw[black!50!green] (-0.2, 0) -- (0, 0); \draw[black!50!green] (0, 0) -- (0.2, 0); \draw[black!50!green] (-0.07, 0.07) -- (0.07, -0.07); \draw[black!50!green] (-0.07, -0.07) -- (0.07, 0.07); \end{tikzpicture} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {subsampling+dropout}; \draw[red] (-0.2, 0) -- (0, 0); \draw[red] (0, 0) -- (0.2, 0); \filldraw[fill=red,draw=black,radius=0.06] (0, 0) circle; \end{tikzpicture} }
\caption[]{ \emph{Left:} Densities versus oracle error rates of lattices generated by different pruning methods. \emph{Right:} Time spent on generating the lattices with different pruning methods. % \klcomment{say something about the vertical line?} \label{fig:pruning} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {edge pruning}; \draw[blue] (-0.2, 0) -- (0, 0); \draw[blue] (0, 0) -- (0.2, 0); \filldraw[fill=blue,draw=black] (-0.07, -0.04) -- (0.07, -0.04) -- (0, 0.08) -- cycle; \end{tikzpicture} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {vertex pruning}; \draw[black!50!green] (-0.2, 0) -- (0, 0); \draw[black!50!green] (0, 0) -- (0.2, 0); \draw[black!50!green] (-0.07, 0.07) -- (0.07, -0.07); \draw[black!50!green] (-0.07, -0.07) -- (0.07, 0.07); \end{tikzpicture} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {beam pruning}; \draw[red] (-0.2, 0) -- (0, 0); \draw[red] (0, 0) -- (0.2, 0); \filldraw[fill=red,draw=black,radius=0.06] (0, 0) circle; \end{tikzpicture} }
\caption[]{Learning curve of the proposed two-pass system compared with the baseline system. The time gap between the first pass and the second pass is the time spent on pruning. \label{fig:learning} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {baseline 1st pass ($R$)}; \draw[blue] (-0.2, 0) -- (0, 0); \draw[blue] (0, 0) -- (0.2, 0); \filldraw[fill=blue,draw=black] (-0.07, -0.04) -- (0.07, -0.04) -- (0, 0.08) -- cycle; \end{tikzpicture} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {proposed 1st pass ($A_1$)}; \draw[black!50!green] (-0.2, 0) -- (0, 0); \draw[black!50!green] (0, 0) -- (0.2, 0); \draw[black!50!green] (-0.07, 0.07) -- (0.07, -0.07); \draw[black!50!green] (-0.07, -0.07) -- (0.07, 0.07); \end{tikzpicture} \begin{tikzpicture}[baseline=-\the\dimexpr\fontdimen22\textfont2\relax] \node[right] at (0.2, 0) {proposed 2nd pass ($A_2$)}; \draw[red] (-0.2, 0) -- (0, 0); \draw[red] (0, 0) -- (0.2, 0); \filldraw[fill=red,draw=black,radius=0.06] (0, 0) circle; \end{tikzpicture} }
\caption{Examples of cargo X-ray images, including: (1) triple-view backscatter and single transmission images from an AS\&E OmniView\textregistered\Gantry; (2) example of concealed cigarettes (indicated by red arrows) in a transmission X-ray image of refrigerated container; (3) transmission image of 20 ft container carried by lorry; and (4) examples of transmission images from the Stream-of-Commerce (SoC) captured by a Rapiscan Eagle\textregistered\R60 railscanner. Sources: Zheng and Elmaghraby~\cite{Zheng2013a}; Vogel~\cite{vogel2007vehicles}; Chalmers \emph{et al.}~\cite{Chalmers2007}; and Rapiscan Systems.}
\caption{Example Empty Cargo Verificaiton (ECV) results of a rule-based algorithm (1 \& 2) and a Machine Learning (ML) based algorithm (1-3). The colouring in image (1) shows empty space (purple), roof and chassis (red), cargo walls (green) and an object (blue). Image (2) shows a false positive caused by the vertical spars. The ML-based algorithm detections are shown in (3), false negatives in (4), and false positives in (5). The red dashed boxes indicate windows classified as empty, and the solid blue boxes show windows classified as containing a load. Sources: Orphan\emph{et al.}~\cite{Orphan2005a} and Rogers \emph{et al.}~\cite{Rogers2015}.}
\caption{\label{fig:kepler} (Left panel) Photospheric magnetic activity index, $S_{\text{ph}}$ (in ppm), as a function of the rotational period, $P_\mathrm{rot}$ (in days), of the 18 seismic solar analogs observed with the {\it Kepler} satellite. The mean activity level of the Sun calculated from the VIRGO/SPM observations is represented for a rotation of 25\,days with its astronomical symbol, and its mean activity levels at minimum and maximum of the 11-year cycle are represented by the horizontal dashed lines. (Right panel) Same as the left panel but as a function of the seismic age (in Gyr). The position of the Sun is also indicated for an age of 4.567\,Gyr. The size of the symbols is inversely proportional to the rotation period,$P_\mathrm{rot}$. In both panels, each star is referred by the same number as given in Tables~\ref{table:prop} and \ref{table:sph}.}
\caption{ Comparison of the solar spectrum (black) to two seismic solar analogs, KIC\,4914923 (blue) and KIC\,5774694 (red), around the Ca\,K line observed with the\Hermes spectrograph and illustrative to stars with different magnetic activity levels and ages. \label{fig:hermes_spec} }
\caption{\label{fig:hermes} Chromospheric \Ssymbol~index derived from the \Hermes observations and calibrated into the MWO system as a function of the photospheric $S_{\text{ph}}$ (in ppm) of the seismic solar analogs observed with the {\it Kepler} satellite. The corresponding activity levels of the Sun at minimum and maximum of its 11-year magnetic cycle are represented by the horizontal and vertical dashed lines. The size of the symbols is inversely proportional to the rotation period. Each star is referred by the same number as given in Tables~\ref{table:prop} and \ref{table:sph}.}
\caption{\label{fig:twiny} (Left panel) Photospheric magnetic activity proxy $S_\mathrm{ph}$ (in ppm, solid black line) of KIC\,3241581 estimated from 1422~days of{\it Kepler} observations as a function of time. The individual measurements of the \Ssymbol~index obtained with the \Hermes observations few years after the end of the {\it Kepler} mission are represented by the dots. The gray dashed line represents a sinusoid calculated using the main periodicity found in the $S_\mathrm{ph}$ by a Lomb-Scargle analysis and centered around the first modulation of the $S_\mathrm{ph}$. (Right panel) Same as the left panel but for KIC\,10644253.}
\caption{ The enrichment history of Eu. The thin solid lines show $\abra{Eu}{H}$ of proto-galaxies, with the number of NSM events in them color coded. We randomly sample $\sim 1/300$ of proto-galaxies and plot their abundances. The dash-dotted blue line denote\red{s} the abundance of high-energy Eu in intergalactic space, $\abra{Eu}{H}_{\rm HE}$, and the dashed black line is the low-energy Eu abundance in IGM. }
\caption{Comparison between GL and PU in terms of \redw{SDR (left) and inconsistency (right) in the oracle (solid lines) and informed (dashed lines) scenarios} for various window lengths.}
\caption{Error \redw{$\mathcal{C}$} over iterations within a TF bin where the sources overlap. \redw{The dotted and solid lines respectively correspond to the initializations with the mixture phase and PU algorithm}, and the dashed line corresponds to the average values over $10$ random initializations.}
\caption{Real part of the third partial ($784$ Hz) \redw{in the STFT channel at $786$ Hz} of a C4 piano note where it overlaps with another note (G4), for various initializations of Algorithm~\ref{al:unwrap_mix}.}
\caption{Real part of the third partial ($784$ Hz) \redw{in the STFT channel at $786$ Hz} of a C4 piano note where it overlaps with another note (G4), reconstructed with several methods in the oracle scenario.}
\caption[]{MACSJ1131.8-1955. (a) TGSS 150 MHz contours, (b) GMRT 235 MHz contours, and (c) GMRT 610 MHz contours on {\it Chandra} X-ray image. All radio images have a resolution of 20$''$ $\times$ 20$''$. First contour is drawn at 3$\sigma$ where $\sigma_{150 MHz}$ = 5 mJy beam$^{-1}$, $\sigma_{235 MHz}$ = 3 mJy beam$^{-1}$, and $\sigma_{610 MHz}$ = 0.30 mJy beam$^{-1}$. Contour level increases in steps of $\sqrt{2}$. {\color{red} +} sign are for the relics sources. Dashed line shows -3$\sigma$ contours.}
\caption[]{ MACSJ2243.3-0935. (a) GMRT 610 MHz contours (resolution of 20$''$ $\times$ 20$''$) for a larger region around the core of cluster, and (b) GMRT 610 MHz contours on {\it Chandra} X-ray image for the central halo region. First contour is drawn at 3$\sigma$ where $\sigma_{610 MHz}$ = 0.10 mJy beam$^{-1}$. Contour level increases in steps of $\sqrt{2}$. {\color{red} +} sign are for the candidate relic. Dashed line shows -3$\sigma$ contours.}
\caption[]{MACSJ1131.8-1955. (a) TGSS 150 MHz contours, (b) GMRT 235 MHz contours, and (c) GMRT 610 MHz contours on {\it Chandra} X-ray image. All radio images have a resolution of 20$''$ $\times$ 20$''$. First contour is drawn at 3$\sigma$ where $\sigma_{150 MHz}$ = 5 mJy beam$^{-1}$, $\sigma_{235 MHz}$ = 3 mJy beam$^{-1}$, and $\sigma_{610 MHz}$ = 0.30 mJy beam$^{-1}$. Contour level increases in steps of $\sqrt{2}$. {\color{red} +} sign are for the relics sources. Dashed line shows -3$\sigma$ contours.}
\caption[]{ MACSJ2243.3-0935. (a) GMRT 610 MHz contours (resolution of 20$''$ $\times$ 20$''$) for a larger region around the core of cluster, and (b) GMRT 610 MHz contours on {\it Chandra} X-ray image for the central halo region. First contour is drawn at 3$\sigma$ where $\sigma_{610 MHz}$ = 0.10 mJy beam$^{-1}$. Contour level increases in steps of $\sqrt{2}$. {\color{red} +} sign are for the candidate relic. Dashed line shows -3$\sigma$ contours.}
\caption{Plot indicating the strength of each of the non-ideal terms relative to each other and to the inductive term as a function of radius: `{\color{red}- -}' is H/O, `$\cdot\, \cdot$' is A/H, `{\bf \color{red}--}' is H/I, `- -' is A/I and `- $\cdot$' is O/I. The yellow region is a strong Ohmic dominated region. The blue region is a strong Hall dominated region and the green region is a weak Hall dominated region. }
\caption{Comparison of descriptive metadatabase reference settings with settings in the back-end acquisition and control unit. The decision logic is indicated in the flow diagram. FPGA, field-programmable gate array. \textcolor{red}{}}
\caption{Instanton time distribution $P_a(\Delta t)$ for $a =$ 0.46$\sigma_S$ and $\phi =$ 0.74 for two halves of the field of view. $P_a(\Delta t)$ for the half that contains the amorphous wall is shown by ({\color{red} $ \circ$}) and that for the remaining half is shown by ({\color{black} $ \square$}).}
\caption{The excitation concentration profiles $c_a(z)$ normalized by their respective bulk values $c_a^{bulk}$ for $a =$ 0.23$\sigma_S$ ({\color{black} $ \blacksquare$}) and $a =$ 0.46$\sigma_S$ ({\color{red} $ \bullet$}) for $\phi =$ 0.68 (a), $\phi =$ 0.71 (b), $\phi =$ 0.74 (c) and $\phi =$ 0.75 (d).}
\caption{a) Determination of the beta relaxation time $\tau_{\beta}$. $d\log(\langle \Delta r^2(t) \rangle)/d\log(t)$ for $\phi =$ 0.68 ({\color{black} $ \blacksquare$}), $\phi =$ 0.71 ({\color{red} $ \bullet$}), $\phi =$ 0.74 ({\color{blue} $ \blacktriangle$}), $\phi =$ 0.75 ({\color{magenta} $ \blacktriangledown$}), $\phi =$ 0.76 ({\color{green!50!black} $ \blacktriangleright$}) and $\phi =$ 0.79 ({\color{blue!50!black} $ \blacktriangleleft$}). The dashed vertical lines denote corresponding values of $\tau_{\beta}$ b) The mean squared displacement $\langle \Delta r^2(t) \rangle$ for various $\phi$. The colors and symbols are identical to those in (a). The dashed horizontal lines denote the square of the cage size $R_c$, defined as $R_c = \sqrt{\langle \Delta r^2(\tau_{\beta}) \rangle}$, evaluated using the values of $\tau_{\beta}$ obtained from (a).}
\caption{The variation of dynamic length scales $\xi_{dyn}$ ({\color{blue} $ \bullet$}), taken from \cite{nagamanasa2015direct}, $\xi_c$ ({\color{black} $ \blacksquare$}) and $\xi_j$ ({\color{green!50!black} $ \blacktriangle$}) with $\phi$. The errors bars correspond to the error on the fitting parameter. Since we have not averaged the data over multiple realizations of the amorphous wall, the actual errors in estimating the length scales are larger than those indicated by the error bars.}
\caption{Investigation of RBC shapes at different shear rates in a cone-and-plate rheometer. A) Observation of hardened cells by optical (black and white) and confocal (red) microscopy: with increasing $\dot{\gamma}$ the formation of highly deformed stomatocytes and then polylobed cells \textcolor{red}{(trilobe and hexalobe)} are detected (scale bars are 5 ${\mu}$m). B) Shape distribution of RBCs populations in samples hardened at different shear rates: the three regions in color highlight different regime of decrease of discocytes population. \textcolor{red}{The error bars represent triplicate measurements realized in the two rapidly varying regions. They illustrate the typical variance of the measurements.}}
\caption{ \textcolor[rgb]{1,0,0}{Microfluidic observations of RBC dynamics in shear flow. A) Time lapse sequence of deformation of RBCs at different $\dot{\gamma}$: Time lapses of 20 ms for the tumbling discocyte, 6 ms for the rolling discocyte, 3.25 ms for the 3 stomatocytes, 0.6 ms for the trilobe in top view, 1.75 ms for the trilobe in side view and 0.6 ms for the hexalobe. Right panel corresponds to analogous time sequences of RBCs obtained with YALES2BIO simulations: Time lapses are given in $1/\dot{\gamma}$ units: $8$ for the tumbling discocyte, $7$ for the rolling discocyte, $6$ for the stomatocyte and the tumbling deformed stomatocyte, and $7$ for the four last cases. B) Stop flow sequences of (left) a trilobed and (right) an hexlobed cell. For the trilobe, the total relaxation time is 1 s and the intermediate images are separated by respectively 0.23 and 0.56 s. For the hexalobe, relaxation occurs in 1.18 s and successive images have a time interval of respectively 0.32 and 0.71 s. (scale bars are 5 ${\mu}$m).}. }
\caption{\textcolor[rgb]{1,0,0}{Characterization of RBCs morphology as a function of increasing $Ht$. A) Shape distributions of HRBCs at 900 s$^{-1}$ in suspensions with different $Ht$. For the sake of clarity, the number densities of discocytes and stomatocytes, which are negligible, have been omitted. Inset: black frames on the right, a top view and a cross section of a creased discocyte acquired by confocal microscopy as well as an image of a polylobed irregular shape (multilobe) obtained by optical microscopy (scale bar is 5 ${\mu}$m). Physiological hematocrits in the microcirculation lie in the range $15-30$\% indicated by the light yellow region. B) Distributions of RBCs extension for different $Ht$ values measured in SDPD simulations. The extension is defined as the maximum instantaneous cell length in the flow direction, as depicted in one of the insets. The distributions include the data from multiple RBCs in a suspension and are averaged over many time instances. The insets illustrate typical shapes for different values of the cell extension. The two extension regions marked by light yellow and light blue colors schematically depict cell-extension ranges, where polylobed shapes and creased discocytes are observed.} }
\caption{Rheology of dense suspensions of RBCs ($Ht=45\%$). A) Relative viscosity of the suspensions of deformable RBCs in plasma (red) and in PBS/BSA (yellow) as a function of $\dot{\gamma}$ in comparison to the suspensions of washed cells hardened at rest (green) and at $\dot{\gamma}$=1500 s$^{-1}$ (blue). Suspensions with a deformable corpuscular phase show a typical shear-thinning for increasing $\dot{\gamma}$, whereas hardened samples have a nearly Newtonian behavior. \textcolor{red}{SDPD simulation data (black stars) are also shown for deformable cells and agree well with the experimental results for RBCs suspended in plasma or in PBS/BSA.} B) Rheology of washed blood in comparison to the suspensions of RBCs in solutions with different dextran concentrations (T=25 $^{\circ}$C). The effect of viscosity ratio $\lambda$ is highlighted in the inset by the illustration of single cells flowing in both PBS/BSA and dextran solutions in microfluidics at a comparable shear stress $\tau$. The scale bars are equal to 5 $\mu$m.}
\caption{\gls{DL} user throughput \glspl{CDF}, over random routes through the Madrid map, with channel/\gls{CSI} and location-based \gls{MU-MIMO} transmit \gls{BF} and \gls{RBF}. Relying only on location information is better on average than using only {\colorrev\gls{CSI}} measurements. The best overall performance is obtained by using channel-based \gls{BF} and location-based \gls{RBF}.}
\caption{Illustrations of {\colorrev selected} positioning prospects in 5G a) \gls{REM} generation, b) proactive \gls{RRM} for a car whose location is being tracked, and c) \gls{ITS}-based traffic control and collision avoidance with self-driving cars.}
\caption{Accuracy of tracking the arrival and departure angles with \glspl{EKF} in terms of the median error. In network-centric \gls{EKF}, the \gls{UE} transmits \gls{UL} reference signals from all antenna elements and the \gls{AN} tracks both arrival and departure angles of the \gls{LoS}-path. In decentralized \gls{EKF}, the \gls{UE} transmits \gls{UL} reference signals from a single antenna element which is used by the \gls{AN} to track the arrival angles of the \gls{LoS}-path with an \gls{EKF}. Such directional parameters are employed in order to design a \gls{DL} beamformed reference signal that {\colorrev then} allows the \gls{UE} to track and design a similar receive beamforming vector.}
\caption{{\colorrev\Glspl{CDF} for 2D positioning errors with \SI{4.8}{MHz} and \SI{9.6}{MHz} \gls{RS} bandwidths over random routes through the Madrid map. Pos\&Clock EKF refers to synchronized\glspl{AN} whereas Pos\&Sync EKF refers to unsynchronized network elements.}}
\caption{ Software used for the calculations: \color{orange}Orca 2.9.1\color{black},\color{cyan} NWChem 6.1.1\color{black}, \color{violet} G09 \color{black} Calculations were performed at functional/def2-QZVPP//B3LYP/def2-TZVPP level of theory “Pople*” means: Geometry optimization and single point energy calculation employing the 6-311+G(3d2f) Pople's basis set. “**” means: Geometry optimization and single point energy calculation employing the respective functional and the def2-QZVPP basis set.}
\caption{ Density distribution of model parameters $\kappa$ (a) and $\theta$ (b), on a sample of 17,146 cascades from the \news dataset. The dashed vertical lines show fitted parameter values for the two retweet cascades illustrated in Fig.~\ref{fig:example-cascade}: $tw_1$ in \textcolor{red}{red} and $tw_2$ in \textcolor{blue}{blue}. We note that \seismic~\cite{Zhao2015} uses a fixed value of $\theta = 0.242$ for all cascades (denoted in green). The distributions for $c, \beta$ and $n^*$ are shown in the supplement~\cite{supplemental}. }
\caption{ Illustration of the rationale behind popularity prediction. The model parameters are estimated starting from a series of observed events $(m_i, t_i)$. Part of one possible unfolding of the diffusion cascade is simulated, using the event rate defined by Eq.~\eqref{eq:phi-separable} and %\eat{a rejection point-process simulation technique} simulated by thinning~\cite{Ogata1999}. The event generations are shown: events in $Generation_1$ are shown in \textcolor{red}{red} color, $Generation_2$ in \textcolor{ForestGreen}{green} and $Generation_k$ in \textcolor{blue}{blue}. Note there is no theoretical limit to the number of generations or to the extent of time until the cascade dies out. Some of the parent-child relations between events in consecutive generations are shown. % The expected number of events in each generation is computed using the fitted values of parameters. }
\caption{ Reduction of prediction error for a subset of cascades. For each cascade, the error made when predicting final popularity using the theoretical $N_\infty$ is shown using \textcolor{red}{red} circles, and with \textcolor{blue}{blue} squares the error made when using the predictive layer $\hat N_\infty$. Each gray arrow shows the reduction of error for a single cascade, and pairs together a red circle with a blue square. The error reduction is show in relation to $A_1$ (a) and $n^\ast$ (b). }
\caption{ Distribution of Absolute Relative Error (ARE) on the \seismicdata dataset (top row) and on the \news (bottom row), for \seismic and \hawkes. A part of the diffusion cascade is observed before making the predictions: 5 min (left column), 10 min (middle column) and 1 hour (right column). The \textcolor{red}{red} line and the numeric annotations denotes the median values of the distributions. }
\caption{ Mean ARE $\pm$ standard deviation for \seismic, \hawkes, \featuredriven and \hybrid models for the \news July'15 dataset. }
\caption{Distribution of ARE on the \news dataset, split in time for July, for \seismic, \featuredriven, \hawkes and \hybrid, after observing 10 minutes (a) and 1 hour (b). The \textcolor{red}{red} line and the numeric annotations denote median value.}
\caption{ ARE distribution over popularity percentiles, on \news dataset, after observing cascades for 10 minutes. The \textcolor{blue}{blue} line denotes the overall median ARE shown in Figure~\ref{fig:hybrid-comparison}a. Note that the y-axes are not on the same scale for all three graphs. }
\caption{Sketch of wavelet compression. Solid points (\tikzcircle[black,fill=black]{3pt}), reference points; circle points ($\bigcirc$), calibration points.}
\caption{Empirical specificity ($1-$size) of the test under $H_0$ for test statistics \colorbox{sedaseda}{$\mathcal{R}_{N}(T)$} and \colorbox{seda}{$\mathcal{C}_{N}(T)$} using the asymptotic critical values, considering a~significance level of $5\%$, weight function $w(t)=t^2$, and smoothing window width $h=2$}
\caption{Empirical sensitivity (power) of the test under $H_1$ for test statistics \colorbox{sedaseda}{$\mathcal{R}_{N}(T)$} and \colorbox{seda}{$\mathcal{C}_{N}(T)$} using the asymptotic critical values, considering a~significance level of $5\%$, weight function $w(t)=t^2$, and smoothing window width $h=2$}
\caption{Empirical sensitivity of the test for small values of $\tau$ under $H_1$ for test statistics \colorbox{sedaseda}{$\mathcal{R}_{N}(T)$} and \colorbox{seda}{$\mathcal{C}_{N}(T)$} using the asymptotic critical values, considering a~significance level of $5\%$, weight function $w(t)=t^2$, and smoothing window width $h=2$}
\caption{The calculated $J/\psi$ production cross section ratios $R_{W/Be}(x_{F})$ without the initial state energy loss(dotted line), by considering the incident quark energy loss effect( dashed line), and the energy loss of incoming quark and gluon (solid line). The solid triangles are the E866 experimental data\textcolor[rgb]{1.00,0.00,0.00}{[4]}.}
\caption{Performance of \textcolor{blue}{non-overlapping} (\textcolor{blue}{--$\blacklozenge$--$\blacklozenge$--}) vs \textcolor{red}{overlapping} (\textcolor{red}{--$\blacktriangleright$--$\blacktriangleright$--}) pre-processing for the task of kNN based classification for different number of nearest neighbors. This is with best choice of other parameters -- dimension: $250$, {\tt kmer} size: $3$, context size: $5$ ($25$ for overlapping). It is clear that non-overlapping processing performs significantly better than the overlapping processing.}
\caption{kNN Classification: Performance of \textcolor{red}{{\tt seq2vec}} (\textcolor{red}{--$\blacktriangleright$--$\blacktriangleright$--}) and \textcolor{cyan}{{\tt ProtVecs}} (\textcolor{cyan}{--$\blacksquare$--$\blacksquare$--}) as compared to that of \textcolor{blue}{BLAST} (\textcolor{blue}{--$\blacklozenge$--$\blacklozenge$--}) as a function of $k$ used for kNN.}
\caption{ Tissue structure in the SPV model . {\bf \color{red} (a)} Simulation snapshots at 3 different values of the preferred cell perimeters $p_0$. Cell centers are indicated by points and cell shapes are given by their Voronoi tessellation (red outlines). {\bf \color{red} (b)} Contour plot of the structure factor $S(q_x,q_y)$ corresponding to the states shown in (a). Scale bar has length $2\pi/D$ in reciprocal space, where $D$ is the average spacing between cell centers. {\bf \color{red} (c)} Pair-correlation function $g(r)$ at different values of $p_0$. {\bf \color{red} (d)} Structure factor $S(q)$ at different values of $p_0$. }
\caption{ Characterizing photonic properties in 2D. {\bf \color{red} (a)} Photonic band structure of Transverse Magnetic(TM) for the material constructed by placing dielectric cylinders at the cell centers exhibits a PBG. Design is based on a ground state of the SPV at $p_0=3.85$. $\bm{k}_\|$ is the in-plane wave vector. % {\bf \color{red} (b)} The optical density of states of TM at different values of $p_0$. The width of the bandgap $\Delta \omega$ has a strong dependence on $p_0$, while the midgap frequency $\omega_0$ remains constant (SI Appendix Fig. S4). % }
\caption{ {\bf \color{red} (a)} The structure of the PBG as function of $p_0$. The ''TM-optimized'' structure corresponds to the size of the TM bandgap for a material constructed by placing dielectric cylinders at cell centers. The ''TM+TE Optimized'' structure corresponds to complete PBGs for a material constructed using a trivalent network design (see text). Phases are colored according to the mechanical property of the material. At $p_0 \approx 3.81$ the material under goes a solid-fluid transition where the shear modulus vanishes. The dependence of $\omega_0$ on $p_0$ is weak as shown in SI Appendix Fig.~S4(a). {\bf \color{red} (b)} Effect of heating on the PBG in the TM optimized structure. At fixed $p_0=3.7$, temperature T is gradually increased. At $T \approx 0.25$ the material begins to fluidize through melting while the bandgap still persists into the mechanical fluid phase. {\bf \color{red} (c)} The relationship between the bandgap size in the TM optimized structure and the short range order of the system. The explicit dependence of short range order on $p_0$ and $T$ is shown in SI Appendix Fig.~S2 and S3. }
\caption{ Structure characterization in the 3D Voronoi-cell model. {\bf \color{red} (a)} Cell centers and their corresponding Voronoi tessellation as well as the decorated 3D photonic structure are shown for a state at $s_0=5$ and $N=100$ cells. The cell centers are drawn with a finite size and different colors only to aid visualization. {\bf \color{red} (b)} Cell centers and their corresponding Voronoi tessellation as well as the decorated 3D photonic structure are shown for a state at $s_0=5.82$ and $N=100$ cells. {\bf \color{red} (c)} Pair-correlation function $g(r)$ at different values of $s_0$. {\bf \color{red} (d)} Structure factor $S(q)$ at different values of $s_0$. }
\caption{ Characterizing photonic properties in the 3D design. {\bf \color{red} (a)} The optical density of states is shown at various values of $s_0$. Here frequency is in units of $2\pi c/ L$, where $L$ is the average edge length in the photonic network. {\bf \color{red} (b)} The size of the gap-midgap ratio $\Delta \omega / \omega_0$ as function of $s_0$. {\bf \color{red} (c)} Midgap frequency $\omega_0$ as function of $s_0$. }
\caption{ Image a) shows the landmarks predicted by Kazemi Detector~\cite{kazemi2014one}, b) shows an example for a landmark group, c) shows the average of the group of landmarks from the left eye and the corresponding grid for features extraction, d) shows the predicted landmarks marked with \textcolor{red}{red +} and the ground truth landmarks marked with \textcolor{green}{green x} and their names. }
\caption{ Wong and Wang's 2D dynamics model for perceptual decision-making~\cite{Wong2006}. We train the model with 90 trajectories (uniformly random initial points within the unit square, 0.5 s duration, 1 ms time step) with different input coherence levels $c=\{0, 0.5, -0.5\}$ (30 trajectories per coherence). The \textcolor{yellow}{yellow} and \textcolor{green}{green} lines are the true nullclines. The black arrows represent the true velocity fields (direction only) and the \textcolor{red}{red} arrows are model-predicted ones. The black and \textcolor{gray}{gray} circles are the true stable and unstable fixed points, while the \textcolor{red}{red} ones are local minima of model-prediction (includes fixed points and slow points). The background contours are model-predicted $\log\lVert \frac{d\,\bm{s}}{d\,t} \rVert_2$. We simulated two 1 s trajectories each for true and learned model dynamics. The trajectories start from the \textcolor{cyan}{cyan} circles. The \textcolor{blue}{blue} lines are from the true model and the \textcolor{cyan}{cyan} ones are simulated from trained models. Note that we do not train our model on trajectories from the bottom right condition ($c=1$). }
\caption{150 training trajectories for the ring attractor. \textcolor{green}{Green} circles are initial states and \ct{red} circles are final states.}
\caption{The spherical image-charge approximation. The \textcolor{blue}{blue solid} line shows cross section of the tip. While the \textcolor{red}{red dashed} line shows the imaginary sphere used to calculated the image-charge partner for the \(\star\). The \(\star\) shows the location of the electron and its image-charge partner used in~\hyperref[fig:ic-sphere-theory]{\autoref{fig:ic-sphere-theory}(a)}, while the +'s show the positions of the electrons used in~\hyperref[fig:ic-sphere-theory]{\autoref{fig:ic-sphere-theory}(b)}.}
\caption{Comparison of the surface charge density on the surface of the tip for the spherical image-charge approximation (\textcolor{blue}{blue solid}) and exact solution~\cite{Peridier4888, Gil2012794} (\textcolor{red}{red dashed}). (a) Electron at (\(\xi = 1.01\), \(\eta = 0.8\)) close to the peak of the tip (Shown as a \(\star\) in~\autoref{fig:sphere-image-charge-ex}). (b) Two electrons at (\(\xi = 1.05\), \(\eta = 0.8\) (Shown as the left and right \(+\) in~\autoref{fig:sphere-image-charge-ex}). The surface charge density at the tip of the peak due to the vacuum field is \(8.2\times 10^4\, \mathrm{\mu C/m^2}\).}
\caption{The relative error (\textcolor{red}{right axis}) of the surface charge density (\textcolor{blue}{left axis}) at the peak of the tip. a) Varying \(z\) with \(\xi = 1\) fixed. b) Varying \(\xi\) with \(\eta = -0.95\) fixed.}
\caption{Current density as a function of \(\xi\), the position on the tip with the distance from the peak shown for comparison. The \textcolor{blue}{blue} line shows the simulations results, while the \textcolor{green}{green} dashed line shows the Fowler-Nordheim result of~\autoref{eq:FN-eq}.}
\caption{Detection limits, in the sky mapper, for binary objects (top) and large asteroids (bottom). Top panel: the detection is given as a function of the separation of the pair (irrespective of its position angle) and the magnitude difference between the secondary and the primary; the colour code indicates the magnitude of the primary. The detection in the binned sky mapper CCDs stops at a separation of less than approximately 0.3\,arcsec (corresponding to$\approx 2.5$ binned pixel). Bottom panel: the detection is given as function of the apparent diameter of the object. The corresponding apparent magnitude is derived for a given albedo and three different heliocentric distances. Objects larger than 0.7\,arcsec will not be systematically detected; when detected, their predicted position can show an offset from the true one by several pixels.}
\caption{$3\,\sigma$ ellipse uncertainty on the 2029 b-plane of Apophis and position of the center of primary (\textcolor{black}{$\star$}) and secondary keyholes leading to collision at ascending node (\textcolor{red}{$\blacksquare$}) and descending node (\textcolor{blue}{$\blacksquare$}). The dotted ellipse is computed using set S$_{\scriptscriptstyle 1}$ and the filled one using set S$_{\scriptscriptstyle 2}$. The coordinates are expressed in $\sigma$ units.}
\caption{relativistic acceleration calculated with Lie (\textcolor{blue}{+}) and Radau Integrator (\textcolor{red}{---}) and the absolute difference $\Delta \pmb{\gamma}$ (- - -) between the value of $\pmb{\gamma}$ calculated with Radau and Lie integrators. The y axis is on a logarithmic scale.}
\caption{relativistic effect on Mercury's perihelion calculated with Lie integrator in arcsec (\textbf{\textcolor{blue}{-}}). $\Delta(\Delta \omega)$ represents the absolute difference between the theory and Lie value in arcsec (\textcolor{red}{$+$}). We also represented the perihelion precession per revolution (\textcolor{blue}{$\bullet$}).}
\caption{ \label{fig:sred} Reduced cross section \sred\{\it{vs.}}\reduced energy\Ered\for the$^{23}$Na\rap $^{26}$Mg reaction. The reduced cross sections \sred\slightly increase towards lower target masses; the expected range is indicated by three calculations for$^{21}$Ne, $^{36}$Ar, and $^{51}$V (taken from \cite{Mohr15}; for details see there). The new data \cite{How15PRL,Tom15PRL} fit into the general systematics whereas the earlier data \cite{Alm14} are much higher. }
\caption{Time (left-hand side, as box plots) and error (right-hand side, as bars representing means and error bars representing standard error) by \textit{Highlighting} and \textit{Slider}, grouped by \textit{Task}. % \textcolor{mygreen}{$\blacksquare$} $H^0S^0$: both \textit{Highlighting} and \textit{Slider} are off; \\ %\textcolor{myred}{$\blacksquare$} $H^0S^1$: \textit{Highlighting} is off and \textit{Slider} is on; \\ %\textcolor{myblue}{$\blacksquare$} $H^1S^0$: \textit{Highlighting} is on and \textit{Slider} is off; \\ %\textcolor{mypurple}{$\blacksquare$} $H^1S^1$: both \textit{Highlighting} and \textit{Slider} are on. \textcolor{mygreen}{$\blacksquare$} $H^0S^0$ \textcolor{myred}{$\blacksquare$} $H^0S^1$ \textcolor{myblue}{$\blacksquare$} $H^1S^0$ \textcolor{mypurple}{$\blacksquare$} $H^1S^1$ }
\caption{Statistically significant differences for time and error, by \textit{Task}. An arrow means that the source is faster or, respectively, more accurate than the destination, with the reported probability. Lines represent all-pairs comparisons between factor combinations (\textcolor{mygreen}{$\blacksquare$} $H^0S^0$, \textcolor{myred}{$\blacksquare$} $H^0S^1$, \textcolor{myblue}{$\blacksquare$} $H^1S^0$, and \textcolor{mypurple}{$\blacksquare$} $H^1S^1$), as well as pairwise comparisons by \textit{Highlighting} ($H^0$--$H^1$, top) and \textit{Slider} ($S^0$--$S^1$, left). %Lines represent pairwise comparisons and all-pairs comparisons between factor combinations. }
\caption{The Bayesian SLOPE posterior mean \textcolor{red}{$\bullet$}, the SLOPE estimate $\triangle$, and $95\%$ Bayesian credible sets for the vector of regression coefficients $\beta$.}
\caption{\veritas\TeV\gray\light curves for six northern blazars subject to regular monitoring during the IC40 run, along with the maximum-likelihood interpolation resulting from our Gaussian process regression (Sec.~\ref{sub:lightcurves}). Time is in days since the start of IC40 (MJD 54562), aligned across all lightcurves.}
\caption{Times of interest for Mrk~421 (shaded intervals). These times were selected in our initial optimization as the most sensitive search for associated neutrinos (Sec.~\ref{sub:analysis}). The selection includes 45.6 days with a total TeV \gray\fluence of$4.1\times 10^{-4}$\,\phtpercm\and yields an expected background of 1.03~neutrinos.}
\caption{Times of interest for Mrk~501, 1ES~1218+304, 3C~66A, W~Comae, and 1ES~0806+524 resulting from our Mrk~421-exclusive search (shaded intervals). These times were selected as the most sensitive search for associated neutrinos that excludes data from Mrk~421 (Sec.~\ref{sub:analysis}). The selection includes 51.6 days with a total \gray\fluence of$1.4\times 10^{-4}$\,\phtpercm\and yields an expected background of 1.17~neutrinos.}
\caption{Maps at $\nu=541$ MHz ($z=1.62$) in equatorial coordinates containing the cosmological HI signal and {\bf noise only} (upper row) and including the contribution from Galactic and extra-Galactic foregrounds, (bottom row). The masks we use in our analysis are shown in the right column. The mask used for simulations without foregrounds is shown in the upper-right panel (the \textcolor{red}{\textit{cosmological mask}}) is defined by a simple cut in declination compatible with SKA the observing site. The bottom-right panel displays the mask we employ for maps containing foregrounds (the \textcolor{blue}{\textit{foregrounds mask}}): beyond the declination cuts, we remove regions of high foreground emission.}
\caption{Characteristics of the four redshift bins used in our analysis. The first and second columns show the frequency and redshift range, while the third column displays the mean redshift of the bin. The fourth column shows the comoving volume covered by each redshift bin, with the numbers displayed in \textcolor{gray}{gray}, \textcolor{red}{red} and \textcolor{blue}{blue} corresponding to simulations using \textcolor{gray}{no mask}, using the \textcolor{red}{cosmological mask} and using the \textcolor{blue}{foregrounds mask} respectively. The fifth column shows the HEALPix resolution parameter $N_{\rm side}$ us in the analysis of each bin. The radial widths and radial resolutions of all bins are constant and correspond to $0.9\,h^{-1}{\rm Gpc}$ and $5\,h^{-1}{\rm Mpc}$ respectively.}
\caption{For each simulation we fit the measured radial power spectrum on 4 the redshift bins specified in the first column using the theoretical template of Eq. \ref{template}. Columns 4-6 show the mean and standard deviation of the BAO parameter $\alpha$. Constraints are shown for maps containing only the cosmological signal (column 4), the cosmological signal plus instrumental noise (column 5) and the cosmological signal, noise and foreground residuals (column 6). The radial power spectra are measured using all pixels of the maps (gray numbers) or using a mask: \textcolor{red}{the cosmological mask} ($f_{\rm sky}=0.72$) and the \textcolor{blue}{the foregrounds mask} ($f_{\rm sky}=0.58$).}
\caption{Consequences of increasing the opacity in the CSM by changing its composition according to Equation (\ref{eq:composition_metalicity_boosted}). In particular, the models \modfid, \modztwo, \modzfive, \modztwenty and \modzhundred are shown. The composition change leads to a stronger redistribution of high energy radiation generated during the ejecta--CSM interaction into the optical $UBVRI$ bands around peak (upper panel). In the bottom panel, the increase in opacity when changing $f_{Z}$ is illustrated in terms of the absorption-mean $\kappa_{J}$ (see Equation \ref{eq:absorption_mean_opacity}), evaluated in the range $\SI{500}{\angstrom} \le c/\nu \le \SI{3250}{\angstrom}$.}
\caption{The observed $H_{160} - [3.6]$ and \IRACcolour colours of the three individual components of CR7 (filled circles) and the full object (open circle), compared to predictions from both standard and exotic models. We show the predicted colours of Pop. III models derived from {\sc Yggdrasil} (small blue circles) and from S15 (blue square). The DCBH prediction from~\citet{Agarwal2016} is shown as the black star. The grey shaded regions show the range of predicted colours from~\citet{Bruzual2003} models with ages of $10\, $--$250\,{\rm Myr}$ (effectively from left to right) where rest-frame optical emission lines have been added in the range $EW_{0}$\hboiii$ = 1000$--$2000$\AA. The light grey region was calculated assuming a standard ratio of {\sc[OIII]}/{\sc H}$\beta$, whereas the dark grey region shows the results with {\sc[OIII]}/{\sc H}$\beta = 15$.}
\caption{The predicted $H_{160} - [3.6]$ and \IRACcolour colours of the BPASS models (lines), compared to the observed colours of the A component of CR7 (filled black circle). The line colours specify the metallicity mass fraction of the model. For each line the age increases from left to right, with circles denoting $1\,{\rm Myr}$, $10\,{\rm Myr}$, $50\,{\rm Myr}$ and $100\,{\rm Myr}$. Dashed lines show metallicity mass fractions from $0.02$ to $0.004$ ($\simeq\,{Z}_{\sun}$ to $1/5\,{Z}_{\sun}$), which show a trend to a more negative \IRACcolour colour as the metallicity decreases. As the metallicity drops further however, the trend is reversed (solid lines; $0.002$ to $1 \times 10^{-5}$ corresponding to $1/10\,{Z}_{\sun}$ to $1/2000\,{Z}_{\sun}$) and the predicted \IRACcolour colour eventually becomes positive. The predicted colours of Pop. III and DCBH models are shown as the points in the upper right, as described in the caption of Fig.~\ref{fig:colours}.}
\caption{(a) NF-L inter-filament spacing vs. monovalent salt concentration of NF-L at: 0 \% PEG ($\blacktriangledown$), 20 \% PEG (\textcolor{red}{$\bullet$}), and EDTA assembly buffer with 0 \% PEG (\textcolor{blue}{$\blacktriangleleft$}). Increasing concentrations beyond 100mM do not significantly modify the stress response, in agreement with previous simulations \cite{Jayanthi2013}. The inter-filament distance measured in NF-L networks formed with EDTA buffer show that the bridging mechanism does not require divalent salts. (b) $\Pi$ vs. $D$ results of NF-L at a wider range of monovalent concentration shows significant deviations from the model at lower salt concentrations. \label{fig:LowerSalt}}
\caption{ (a) GOES X-ray light curves showing the initial C-class flare at $\sim$12:35\,UT. This is followed by an M7.3 class flare peaking at$\sim$13:00\,UT. (b) RHESSI X-ray flux observations from 3-100\,keV. Data gaps due to RHESSI night and the South Atlantic Anomaly are indicated by blue and brown lines, respectively. (c) FERMI GBM light curves showing emission from 4.5-101.6\,keV. (d) Radio dynamic spectra from NDA and Orf\'{e}es covering 10--1000\,MHz. Based on the observations below, the dynamic spectrum is split into five periods, indicated by the vertical dashed lines.}
\caption{ (a) Three colour image of full sun using AIA 94, 131, 335\,\AA. The inset shows the twisted flux rope structure with emission dominating in the 94 and 131\,\AA~channels ($\sim$6 and 10\,MK, respectively). This structure exists for several hours prior to eruption. A distance-time map is constructed along the dashed blue-white line. (b) Distance-time map of the AIA hot channels{\color{black}(see text)}. The initial slow motion can be seen as the slow rise of the green-red feature from 50\,Mm. The initial C-class flare and metric type III group occur at 12:34:30\,UT, marked by vertical white-pink line. (c) Three-color map of the cooler AIA channels showing acceleration phase at$\sim$12:40\,UT.}
\caption{ (a) AIA three colour image of 94, 131, 335\,\AA~channels showing a highly twisted structure with a jet (indicated by the arrow) occurring at its center. (b) Same image as (a) but with a wider field of view and NRH contours from 327--150\,MHz over plotted and maximum and minimum brightness temperature indicated on the corresponding frequency. The sources occur at the same time as the type III group at 12:34:30\,UT and make a roughly straight line above the jet with higher frequencies at lower altitude. (c) The radio sources occur above the north-west loop of the rope. This activity last from 12:43:30--12:46:00\,UT (see movie\_2.mpg). The dashed lines indicate the frequencies of type III emission and Emission `C' shown during period 2 of the dynamic spectrum in Figure~\ref{fig:source_motion_imgs}.}
\caption{(a) Orf\'{e}es and NDA dynamic spectrum observations from $\sim$1000 to 10\,MHz during periods 2, 3, 4 and 5 as indicated by the vertical dashed lines. Emission `C' and type IIIs are observed in period 2 (imaged in Figure~\ref{fig:fluxrope_typeIII}(c)). In period 3 `Flare Continuum A' begins, while the starting frequencies of the type IIIs observed in NDA show a slow drift towards lower frequencies, indicated by the black arrows. The coloured crosses highlighted on the dynamic spectrum correspond to the frequencies and times of the NRH contours below the dynamic spectrum, overlaid on AIA 171\,\AA~running ratio images (NRH contours are in $log_{10}(T_B $\,[K])). At each frequency, two sources are identified in these images; a stationary (AR) source located above the active region and a much weaker looptop (LT) source located above the north-west loop (`NW Loop', as indicated) of the rope. During period 4 a separate `Flare Continuum B' can be observed in Orf\'{e}es around 400\,MHz.}
\caption{(a) Detail of the Flare Continuum A and B radio bursts in the Orf\'{e}es spectrogram. The dashed lines indicate the frequencies at which the light curves below are taken. The blue circles indicate the times of the first flux peak of the Looptop (LT) source from NRH images. The pink curve is FERMI GBM 11--26.9\,keV X-ray flux. (b)-(d) Comparison of normalized Orf\'{e}es flux with the normalized flux of the active region (AR) source and LT source as imaged by NRH - see Table~\ref{tab:table0} for flux normalization values. Comparing black and red curves, the stationary AR source is responsible for the majority of Flare Continuum A, while the LT source (blue dashed curves) is much more sporadic. (e) Both AR source and LT source are comparable to the Orf\'{e}es flux variation over time until 12:53\,UT, suggesting they were both involved in the flare continuum at 432\,MHz. Note the red curve changes from solid to dashed at$\sim$12:53\,UT, indicating that the AR source at this frequency diminishes and a new source appears (seen in Figure~\ref{fig:looptop}). }
\caption{AIA 193\,\AA~running ratio images over plotted with NRH 327, 408, 432 and 445\,MHz. (a) The radio contours at 408\,MHz and above are now located inside the north-west loop of the rope (this loop is visible in the AIA image but obscured by the radio contours, hence it is demarcated by the blue-white crosses). The remainder of the rope is now indistinguishable in the images. The 327\,MHz source (and lower frequencies) is located above the active region. (b) The 408--445\,MHz sources then moves in the direction of the loop as it propagates westward. The 327\,MHz contours shows a source above the active region and one now clustered with the higher frequencies at the loop location. (c) The sources then begin to shift back towards the active region and remain there. Panels (a) to (b) show the radio sources resulting in the `Flare Continuum B' feature of Figure~\ref{fig:flux_comparison}(a).}
\caption{ NRH observations with frequencies displayed in RGB triplets. (a) 445, 432 and 408\,MHz, (b) 327, 298 and 270\,MHz, (c) 228, 173, 150\,MHz each at 12:57:57\,UT. In these images, any spatially coherent levels in brightness temperature are displayed as white i.e., equal amounts of red, green and blue. An increase in any one (or two) of the three images in the triplet is displayed as a primary colour (or combination of primaries). The brightness temperature scaling is indicated on each triplet. In panels (a) to (c) all radio emission is generally situated above the active region, with the emission at the lowest frequencies grouped in a `radio bubble'. (d) The lowest frequencies then develop into a `radio arc', while (e){\color{black}displays an EUV front near the edge of the AIA field of view}, with NRH 150\,MHz contour for comparison.}
\caption{Velocity versus time for eruptive features in the event. The initial slow rise begins in the AIA hot channels to the north-west of the rope, indicated by the dashed red line (d-t trace indicated by the dashed line on the inset). At 12:34:30\,UT the metric type III group, jet and C-class flare occur, shown by the solid vertical pink line. After this is the rapid acceleration on all areas of the flux rope, first beginning with the south-west loop, traced by the solid lines. Blue lines are d-t races from the cooler AIA maps. The LT and LS radio source velocities are derived from images, while the Flare Continuum B and type II velocities are derived from frequency drift in dynamic spectra.}
\caption{(a) Mass function $M(\underline{k})$ derived from the Dyson--Schwinger model of Ref.~\cite{PM97b} for the quark propagator. (b) Associated configuration-space potential $V(r)$ from Fierz-symmetric kernel $K(\bm{p},\bm{q})$ for constituent quark mass $m=0$ (black), $m=0.35\;\mbox{GeV}$ (\textcolor{red}{red}), $m=0.5\;\mbox{GeV}$ (\textcolor{magenta}{magenta}), $m=1\;\mbox{GeV}$ (\textcolor{blue}{blue}), and $m=1.69\;\mbox{GeV}$ (\textcolor{violet}{violet})~\cite{WL16n}.}
\caption{Configuration-space potential $V(r)$ from Fierz-symmetric effective interaction, for massless constituent quarks and mixing parameter $\eta=0$ \cite{WL15} (black), $\eta=1$ (\textcolor{red}{red}), $\eta=2$ (\textcolor{magenta}{magenta}), $\eta=-0.5$ (\textcolor{blue}{blue}), or $\eta=-1$~(\textcolor{violet}{violet}).}
\caption{Configuration-space potential $V(r)$ from Fierz-symmetric effective interaction, for constituent quarks~of mass $m=\mu$ and mixing value $\eta=0$ \cite{WL15} (black), $\eta=0.5$ (\textcolor{red}{red}), $\eta=1$ (\textcolor{magenta}{magenta}), $\eta=2$ (\textcolor{blue}{blue}), and $\eta=-1$~(\textcolor{violet}{violet}).}
\caption{(Colour online) Parameter spaces $\beta\times\alpha$ for $b=3.2$, (a) $I_0=4.6$, (b) $I_0=4.4$, and (c) $I_0=4.2$. We see the regions for {\color{yellow}SC}, {\color{red} BC}, SI, and {\color{blue}IN}.}
\caption{Example for OBD data and sensor data. [\color{blue} Comment: the figure should be revised to be consistent with other figures; the tiles of the figures should also be changed]}
\caption{\colorfig The schematic setup of entanglement generation between two $2$-mode systems in \cite{OuMandel1988-TwoPhBiEnt}. Two single photons in the $\{H,V\}$ polarization modes are generated from a down conversion source, and distributed by beam splitters into two local system with the modes $\{H_{A},V_{A}\}$ and $\{H_{B},V_{B}\}$, respectively. The state $\ket{\psi_{2}}$ after the beam splitting is given in Eq. \eqref{eq::psi_2}. By post-selecting one-photon detection events in each local system, one projects the $\ket{\psi_{2}}$ onto the entangled state $\ket{\phi_{1_{A},1_{B}}}$ in Eq. \eqref{eq::phi_1A1B}. }
\caption{\colorfig Generation and evaluation of bipartite entanglement from $M$ single photon inputs. The input modes and the modes in each output port are indexed by $0,...,M-1$. The $M$ single photon inputs are split by $M$ beam splitters, generating the output state $\ket{\psi}$. For the evaluation, a selection of linear optics transformations $\hat{U}_{A}$ and $\hat{U}_{B}$ are applied to the modes, transforming the photon number operators in the different modes into sets of non-commuting observables. }
\caption{\colorfig The probability distribution of photon number detection events for the state $\ket{\phi_{2_{A},2_{B}}}$ (a) in the original input modes and (b) in the output modes of local DFTs. The axes labled $A$ and $B$ show the photon distributions detected in each local system. In (a), the photon distributions have been arranged according to the maximal number of photons in one mode (Eq. \eqref{eq::pattern_classes_2N4M}). In (b), the distributions have been organized according to the $K$-value (Eq. \eqref{eq::K-blocks}). The first three outcomes correspond to $K=0$, the next two outcomes have $K=1$, followed by another block of three outcomes with $K=2$ and finally two outcomes with $K=3$. The outcomes are clearly grouped in correlated blocks corresponding to the expected correlations in $K$.}
\caption{\colorfig The statistics of photon number detection of the separable state $\ket{1100_{A},0011_{B}}$ (a) in the input modes and (b) in the output modes of two local DFTs. Photon distributions are arranged in the same way as in Fig. \ref{fig::phi_2A2B_corr}. The red dashed squares show outputs that have the correct correlation expected from the entangled state input $\ket{\phi_{2_{A},2_{B}}}$and therefore contribute to the fidelities $F_{\boldvec{n}}$ or $F_K$, respectively. The blue dot-dashed squares indicate the complementary pattern classes $(\boldvec{p},\bar{\boldvec{p}})$ which are used to determine the tighter bounds for separable states. In (a), the separable input state corresponds to a well-defined measurement outcome with $F_{\boldvec{n}}=1$ and $P(\mathcal{E}_{1100},\mathcal{E}_{0011})=1$. In the output modes of the DFT, the photon number distributions shown by the black squares in (b) each have a probability of $1/64$. As a result, the output mode correlation fidelity, i.e. the probability for the correlations $(K,-K)$ (red dashed squares), is $F_{K}=1/4$. }
\caption{Schematic overview of the chosen coordinate system and the relevant angles. The laser light hits the sample perpendicularly.\textcolor{magenta}{{} \label{fig:angles_gaas}}}
\caption{Modification of the particle density due to the presence of anyons in the lattice Laughlin state \eref{state} at $q=3$ and half lattice filling ($\eta=q/2$). The lattice is chosen to be a kagome lattice defined on a disc with radius $27.9$. A quasielectron (quasihole) with charge $-1/3$ ($+1/3$) is placed at the position \textcolor{red}{$*$} (\textcolor{blue}{$+$}), and the color of the $j$th lattice site shows $\langle n_j\rangle_{(-1,+1)}-\langle n_j\rangle_{(0,0)}$.\label{fig:plane}}
\caption{Quasihole (\textcolor{blue}{$+$}) and quasielectron (\textcolor{red}{$*$}) on the torus in the state $|\psi_0\rangle_{(1,-1)}$ (see \eref{statetorus}) with $q=2$. The color of the lattice sites shows $\langle n_j\rangle_{(1,-1)}-\langle n_j\rangle_{(0,0)}$. When moving the quasihole around the blue curve we find numerically, using Monte Carlo simulations, that the difference in Berry phase when the quasielectron is at \textcolor{blue}{$+$} and at \textcolor{red}{$*$}, respectively, is $\phi=-3.145$ with a statistical error of order $0.003$. This is in agreement with the expected result $-\pi$.\label{fig:torus}}
\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Top-left panel:} shows the column density distributions for all five lines in the case with HM, and for the three nitrogen lines with varying $\rm N/O$ (Eq~\ref{eq:NO}). {\color{burntorange}\bf Top-right panel:} shows the covering fraction as a function of the galacto-centric impact parameter for OVI absorbers with column density above $10^{13.27}$cm$^{-2}$ with HM (solid blue curve) and with HM+local (dot-dashed blue curve), for SiIV absorbers with column density above $10^{12.38}$cm$^{-2}$ with HM (solid green curve) and with HM+local (dot-dashed green curve). {\color{burntorange}\bf Bottom-left panel:} shows the same as in the top panel but for the three nitrogen absorption lines with column density above $10^{13.43}$cm$^{-2}$ for NV (blue curves), $10^{13.50}$cm$^{-2}$ for NIII (red curves) and $10^{13.46}$cm$^{-2}$ for NII (green curves), under the assumption that $\rm N/O$ ratio is independent of metallicity. The solid curves correspond to the case with only HM radiation background, whereas the dot-dashed curves are for the case with HM+local radiation. {\color{burntorange}\bf Bottom-right panel:} shows the same as for the bottom-left panel, except that we use Eq (\ref{eq:NO}) for nitrogen abundance as a function of oxygen abundance. }
\caption{%\footnotesize %\scriptsize {\color{burntorange}\bf Top row:} shows the probability distribution function (PDF) of the ratio of ${\rm N(NV)/N(OVI)}$ for all OVI absorbers with $\rm N(OVI)>10^{14}$cm$^{-2}$ with constant N/O ratio (left) and varying N/O as a function of O/H (Eq \ref{eq:NO}). {\color{burntorange}\bf Middle row:} the same for NIII/OVI. {\color{burntorange}\bf Bottom row:} the same for NII/OVI. (Blue, red) histograms are for (HM, HM+local) radiation field. The observational data are shown for three separate types: black dots are those with both lines detected, left green arrows are those where the numerator line is an upper limit and the denominator line a detection, and brown left arrows are those where the numerator line is an upper limit and the denominator line a lower limit. The y coordinates of the points are arbitrary. }
\caption{OO code metrics used for all studies in this paper. Last lines, shown in \textcolor{gray}{gray}, denote the dependent variables.}
\caption{Training and test {\em data set properties} for Jureczko data , sorted by \% defective examples. On the right-hand-side, we show the {\em results from learning}. Data is usable if it has a recall of 60\% or more and false alarm of 30\% or less (and note that, after tuning, there are more usable data sets than before). Results \colorbox{celadon}{ marked with ``$\star$''} show large improvements in performance, after tuning (lower {\em pf} or higher {\em pd}). Data in the \colorbox{lavenderpink}{three bottom rows}, marked with ``$\times$'', are performing poorly-- that data so many defective examples that it is hard for our learners to distinguish between classes. }
\caption{Illustration of model parameterization. (a) Model I parameters are core radius \rc, mantle composition $c$ comprising the oxides Na$_2$O--CaO--FeO--MgO--Al$_2$O$_3$--SiO$_2$, mantle radius \rsolid, mass of water \mice, mass of envelope \menv, envelope Luminosity $L$, and envelope metallicity \Zenv. (b) Model II parameters are as for a), {\coco with} atmosphere parameterized by pressure at the bottom of the atmosphere \Pbatm, number of scale-heights of opaque layers $N$, mean molecular weight $\mu$, and a temperature-related parameter $\alpha$. See Section \ref{parametrization} and table \ref{momo} for more details. \label{Illustration}}
\caption{Sampled one-dimensional (1-D) marginal posterior cdfs (blue) of model I parameters for Neptune: (a) mass of envelope \menv, (b) envelope Luminosity $L$, (c) mass of water \mice, (d) mantle radius \rsolid, (e) core radius \rc, (f) $\fesima$, (g) $\mgsima$. Prior and posterior nearly completely overlap in (g). The envelope metallicity \Zenv (not shown) is fixed, \Zenv = 0. The prior cdfs are plotted in red. {\coco Gray area in plots (a-d) represent independent literature estimates (see main text).} \label{NepGas}}
\caption{{\coco Sampled two-dimensional (2-D) marginal posterior pdfs (blue) of model I parameters for Neptune: (a) mass of envelope \menv and envelope Luminosity $L$, (b) mass of water \mice and mantle radius \rsolid. {\coco Gray areas represent independent literature estimates (see main text)}.} \label{Nep2D}}
\caption{Sampled 1-D marginal posterior cdfs (blue) of model II parameters for Neptune: (a) pressure at bottom of atmosphere \Pbatm, (b) atmospheric mass fraction m$_{\rm atm}$/M (Eq. \ref{massgas}), (c) temperature-related parameter $\alpha$, (d) number of scale-heights of opaque layers $N$, (e) mean molecular weight $\mu$, (f) mass of water \mice, (g) mantle radius \rsolid, (h) core radius \rc, (i) $\fesima$, (j) $\mgsima$. The prior cdfs are plotted in red. {\coco Gray areas in (b,e,f) represent independent literature estimates (see main text).} \label{NepScale}}
\caption{Sampled 1-D marginal posterior cdfs of model I parameters for synthetic planet cases (A-D) of 7 \ME that vary in terms of radii: \mbox{1.7 \RE (A)}, 2.2 \RE (B), 2.6 \RE (C), 2.9 \RE (D); (a) mass of envelope \menv, (b) envelope luminosity $L$, (c) envelope metallicity \Zenv, (d) mass of water \mice, (e) mantle radius \rsolid, (f) core radius \rc, (g) $\fesima$, (h) $\mgsima$. \label{VaryRadii1}}
\caption{Sampled 1-D marginal posterior cdfs of model II parameters for synthetic planet cases (A-D) of 7 \ME that vary in terms of radii: \mbox{1.7 \RE (A)}, 2.2 \RE (B), 2.6 \RE (C), 2.9 \RE (D); (a) pressure at bottom of atmosphere \Pbatm, (b) atmospheric mean molecular weight $\mu$, (c) temperature-related parameter $\alpha$, (d) number of scale-heights of opaque layers $N$, (e) mass of water \mice, (f) mantle radius \rsolid, (g) core radius \rc, (h) $\fesima$, (i) $\mgsima$. Depending on the case, the upper prior bound in (c) differs, which is indicated by the vertical colored lines corresponding to the respective case. \label{VaryRadii2}}
\caption{Sampled 2-D marginal posterior pdfs of model I parameters for synthetic planet case C showing the correlation between: (a) \rc and \rsolid, (b) \rsolid and \mice, (c) \mice and \menv, (d) \menv, and \Zenv, (e) \mice and the averaged $\mu$ corresponding to \Zenv. Those model realizations that explain the data within 1-$\sigma$ are plotted in blue. Samples in (c, d) for which gas mass fractions \menv/$M$ > 0.01 are highlighted in green and should be taken with care. See main text for discussion of features B1 and B2. \label{Comp1}}
\caption{Sampled 2-D marginal posterior pdfs of model II parameters for synthetic planet case C showing the correlation between: (a) \rc and \rsolid, (b) \rsolid and \mice, (c) \mice and \Pbatm, (d) \Pbatm, and $\mu$, (e) \mice and $\mu$, (f) \mice and $\alpha$, (g) \mice and $N$. Those model realizations that explain the data within 1-$\sigma$ are plotted in blue. Samples in (c, d) for which gas mass fractions \menv/$M$ > 0.0001 are highlighted in green and should be taken with care. \label{Comp2}}
\caption{Sampled 1-D marginal posterior cdfs of model I parameters for synthetic planet cases B, E, F, G that vary in terms of data uncertainties. B is the reference case ($\sigma_M$=0.05~$M$, $\sigma_R$ = 0.02~$R$, 20~\% for both $\sigma_{\fesi}$ and $\sigma_{\mgsi}$), E has larger uncertainties in mass and radius ($\sigma_M$=0.2~$M$, $\sigma_R$ = 0.1~$R$), whereas F and G have larger uncertainties in the abundance constraints, 50~\% and 80~\%, respectively. (a) Mass of envelope \menv, (b) envelope luminosity $L$, (c) envelope metallicity \Zenv, (d) mass of water \mice, (e) mantle radius \rsolid, (f) core radius \rc, (g) $\fesima$, (h) $\mgsima$. The priors in (g) and (h) are not shown as not to overload the plot, because they differ among the cases. \label{VaryGas}}
\caption{Sampled 1-D marginal posterior cdfs of model II parameters for synthetic planet cases B, E, F, G that vary in terms of data uncertainties. B is the reference case ($\sigma_M$=0.05 M, $\sigma_R$ = 0.02~$R$, 20~\% for both $\sigma_{\fesi}$ and $\sigma_{\mgsi}$), E has larger uncertainties in mass and radius ($\sigma_M$=0.2~$M$, $\sigma_R$ = 0.1~$R$), whereas F and G have larger uncertainties in the abundance constraints, 50~\% and 80~\%, respectively. (a) Pressure at bottom of atmosphere \Pbatm, (b) atmospheric mean molecular weight $\mu$, (c) temperature-related parameter $\alpha$, (d) number of scale-heights of opaque layers $N$, (e) mass of water \mice, (f) mantle radius \rsolid, (g) core radius \rc, (h) $\fesima$, (i) $\mgsima$. The priors in (h) and (i) are not shown as not to overload the plot, because they differ among the cases. \label{VaryScale}}
\caption{Illustration of model parametrization. Model parameters are core radius \rc, mantle composition $c$ comprising the oxides Na$_2$O--CaO--FeO--MgO--Al$_2$O$_3$--SiO$_2$, mantle radius \rsolid, mass of water \mice, mass of atmosphere \menv, atmosphere Luminosity $L$, and atmosphere metallicity \Zenv. \label{Illustration}}
\caption{Sampled two-dimensional (2-D) marginal posterior pdfs for the six selected exoplanets (a-f) showing the correlation between core and mantle sizes, \rc, and \rsolid and mantle $\fesima$. Bigger dots explain the data within 1-$\sigma$ uncertainty. The straight line describes the lower limit of the mantle size where \rsolid = \rc. Earth-like parameters are depicted by the red cross. \label{4RF}}
\caption{Sampled 2-D marginal posterior pdfs for the six selected exoplanets (a-f) showing the correlation between mantle size \rsolid and mass of water \mice. Blue points explain the data within 1-$\sigma$ uncertainty. Earth-like parameters are depicted by the red cross.\label{1RW}}
\caption{{Application of V0, V1, and V2 stellar abundance estimates: cdfs} of sampled 1D marginal posterior (green to blue) and prior (red) for HD~219134b, 55~Cnc~e$^{\rm S}$, 55~Cnc~e$^{\rm L}$, and HD~97658b: (a) \mice, (c) \rsolid, (d) \rc, and (e) $\fesima$. Posterior cdfs refer to the different host star abundance estimates (V1- iron rich, V2 - iron poor) listed in Table \ref{tabledata}. The dashed lines are the 95~\% lower and upper confidence bounds, which almost completely overlap the solid lines. Depending on the stellar abundance proxy estimates, the upper prior bound in (e) differs, which is indicated by the vertical colored lines corresponding to the respective proxy. \label{Sensitivity}}
\caption{{Application of direct and non-direct abundance proxies: cdfs} of 1D marginal posterior pdfs and prior pdfs for HD~219134b, Kepler-10b, Kepler-93b, CoRoT-7b, 55~Cnc~e$^{\rm S}$, 55~Cnc~e$^{\rm L}$, and HD~97658b: (a) \menv, (b) \Zenv, (c) \mice, (d) \rsolid, (e) \rc. Prior (red), posterior pdf in blue and green for direct and non-direct stellar-planetary abundance correlation, respectively. See discussion in section \ref{nondirect}. \label{cdf2} }
\caption{Our \emph{bet-and-run} restart strategy starts with $k$ independent runs and total time budget $t$. After time $t_1$ all but the best run are terminated (marked with \protect\includegraphics[height=2mm]{illustration-skull.pdf}). The best run (marked with \protect\includegraphics[height=2.3mm]{illustration-star.pdf}) continues for $t_2$ time steps until the total time budget runs out. %Illustration of our bet-and-run restart strategy with $k$ runs; after time $t_1$ all but the best run are terminated. }
\caption{Cumulative distribution function $F\left(s\right)$ for $\delta'=-0.15$ with $\boldsymbol{B}\left(\varphi,\,\vartheta\right)=\boldsymbol{B}\left(0,\,\pi/6\right)$ and $B=3\,\mathrm{T}$. Since $\boldsymbol{B}$ is oriented in one of the symmetry planes of the lattice, only GOE statistics can be observed when neglecting phonons. ~\label{fig:fig2}}
\caption[Two use cases]{Two use cases of itinerant routing, where \BS{scale = 0.06} is a base station, \DC{scale = 0.05} is a data center, \includegraphics[scale=0.035]{bus} is a means of public transport, and A, B, C, and D are the EON nodes.}
\caption{Electricity load data: Performance comparison between the least-squares estimator~\eqref{eq.ls} when $p<n$ (\textcolor{darkgreen}{--}) and the Lyapunov-penalized model~\eqref{eq.lc} when $p>n$, with different levels of cardinality: $s=100p$ (\textcolor{blue}{$\bullet$}), $s=125p$ (\textcolor{red}{\ding{110}}), $s=150p$ (\textcolor{brown}{$\bullet$}), $s=175p$ (\textcolor{black}{$*$}), $s=200p$ (\textcolor{blue}{$\diamond$}), $s=p^2$ (\textcolor{red}{$\bullet$}).}
\caption{Left: The factorization of the {\color{blue}hard}, {\color{green!50!black}collinear} and {\color{red}soft} sectors in SCET. Right: The schematic illustration of the jet production and the measurement of the jet shape.}
\caption{Left: Illustration of the emergent scales in the medium jet formation. Middle: The factorization of the {\color{blue}hard}, {\color{green!50!black}collinear}, {\color{red}soft} and {\color{green!50!black}Glauber} sectors in $\rm SCET_G$. Right: Parton splitting kinematics at leading order.}
\caption{Similar plot as Fig. 2, but shown for the Li-adsorbed GNRs. (a) and (b), respectively, correspond to the single-side adsorptions of adatoms at (1,10)$_{single}$ and (3,7)$_{single}$. The double-side adsorptions are shown for two adtoms at (c) (1,\textcolor{red}{10})$_{double}$ \& (d) (3,\textcolor{red}{7})$_{double}$, four adatoms at (1,\textcolor{red}{4},7,\textcolor{red}{10})$_{double}$, and six adatoms at (1,\textcolor{red}{2},5,\textcolor{red}{6},9,\textcolor{red}{10})$_{double}$. Numbers correspond to the adatom adsorption position. Two colors (black \& red) denote two distinct sides.}
\caption{The spatial charge distributions of $N_A=12$ armchair systems for (a) a pristine GNR, (b) an adatom at edge (1)$_{single}$, (c) two adatoms at (1,10)$_{single}$ and (d) four adatoms at (1,\textcolor{red}{4},7,\textcolor{red}{10})$_{double}$. The $\pi$ and $\sigma$ bondings are, respectively, enclosed by the dashed and solid rectangles. The charge density differences, corresponding to (b), (c) and (d), are revealed in (f), (g) and (h), respectively. Also shown in (e) and (i) are those of $N_z=8$ zigzag system with an adatom.}
\caption{The integrated flux density (in Jy) and the quality factor for all 45 phase calibration sources. Two qualities, good, `{\color{red}{G}}' or moderate, `{\color{blue}{M}}' are provided based on strength of the calibration source and strengths of contaminants (see Sect.~\ref{qual-fac}). The typical error (1~$\sigma$) on flux density is $\lesssim$ 10~mJy. The 235 MHz and 610 MHz are our GMRT measurements and rest of the measurements are gleaned from the VLA \citet{vla-cal-manual}, which are from \citet{Kuhretal,Largeetal,Condonetal,Condonetal2,Griffithetal,Langstonetal,Beckeretal,WhiteBecker}. }
\caption{A Mollweide projection plot in equatorial coordinates showing the sky distribution of sources at 235 MHz (left-hand panel) and at 610 MHz (right-hand panel). The horizontal and vertical axes in each panel are right ascension and declination, respectively. The legends in the plots, i.e., red and blue points, correspond to the quality factor, good, `{\color{red}{G}}' and moderate, `{\color{blue}{M}}', respectively (see Sect.~\ref{qual-fac}) of the calibration source.}
\caption{Upper panel: full synthesis GMRT maps of 3C\,52 at 235 MHz (left-hand panel) and 610 MHz (right-hand panel). The data are reduced using the image model from the database to obtain the calibration for the secondary phase calibration source; the data reduction recipe is explained in Sect.~\ref{comparison}. The synthesized beams for 235 MHz and 610 MHz are 15$^{\prime\prime}$.33 $\times$ 11$^{\prime\prime}$.92 at a P.A. of 64$^{\circ}$.46 and 5$^{\prime\prime}$.92 $\times$ 4$^{\prime\prime}$.68 at a P.A. of 58$^{\circ}$.35, respectively. The peak surface brightness and the RMS noises in the immediate vicinity of the source for 235 MHz and 610 MHz maps are 4.0 Jy~beam$^{-1}$ and 2.0 Jy~beam$^{-1}$, and 4.9 mJy~beam$^{-1}$ and 0.6 mJy~beam$^{-1}$, respectively. The contour levels in these two maps are RMS noise level $\times$ $-$3, 3, 6, 12, 24, 48, 96, 192 mJy~beam$^{-1}$. Lower panel: the original published, full synthesis GMRT maps of 3C 52 at 235 MHz (left-hand panel) and 610~MHz (right-hand panel), Fig~2, upper-panel images of \citet{lalandrao}. The synthesized beams for 235 MHz and 610 MHz maps are 15$^{\prime\prime}$.4~$\times$~12$^{\prime\prime}$.0 at a P.A. of 63$^{\circ}$.7 and 5$^{\prime\prime}$.9~$\times$~4$^{\prime\prime}$.7 at a P.A. of 57$^{\circ}$.5, respectively; and the contour levels in the two maps are, respectively, $-$50, 50, 60, 80, 100, 160, 200, 400, 600, 800, 1200, 1600, 1800 mJy~beam$^{-1}$ and $-$8, 8, 20, 30, 40, 50, 60, 80, 100, 200 mJy~beam$^{-1}$, The error-bars in the full synthesis maps found at a source free location are $\sim$1.4~mJy~beam$^{-1}$ and $\sim$0.3~mJy~beam$^{-1}$ at 235 and 610~MHz, respectively. }
\caption{\label{fig:rho_mu_dis_S} Distribution of $\rho_{\mu}(100)$ in the data sample \red ($N_{e}>2 \times 10^{7}$)\black. Points with error bars \green (statistical errors only)\black: data; blue histogram: MC simulations based on the primary composition inferred from the surface-detector data (43\% protons and 57\% iron), QGSJET-II-04.}
\caption{\label{fig:rho_mu_dis_mu} Distribution of $\rho_{\mu}(100)$ in the data sample. Points with error bars \green (statistical errors only)\black: data. (a) green dashed histogram: MC, protons; red full histogram: MC, iron; (b) orange histogram: MC, best-fit primary composition inferred from these \red $\rho_{\mu}(100)$ distributions (46\% protons and 54\% iron)\black.}
\caption{\label{fig:chi2} The \red $\chi^{2}$ per degree of freedom \black for the muon enhancement coefficient $k$ determined in the text. Blue \red continuous \black line: \red assuming \black the EAS-MSU surface-detector composition; green dashed line: the KASCADE-Grande composition \cite{KASCADE-comp}; red dotted line: the Tunka-133 composition \cite{Tunka-comp}. \red The horizontal line represents 68\% CL. \black }
\caption{\label{tab:models}\red Results for different hadronic-interaction models. }
\caption{Streamwise velocity spectra $E_{uu}(k_x)$: (a) Large-scale normalisation, $E_{uu}/ ( u^{\prime2}L_z)$ as a function of $k_x L_z$. % (b) Small-scale normalisation, $E_{uu}^\ast \equiv \dis^{-2/3} l^{-5/3} E_{uu}(k_x l)$, with $l = l_S$ for LES and $l = \eta$ for DNS. \dashedtridown (grey), DNS (L32); \solid (grey), DNS (M32); \dashedcircf (blue), $R_S=52.6$; \solidcirc (red), $R_S = 101.6$; \solidtriright (green), $R_S = 203$. The slope of the chain-dotted diagonal is $-5/3$. %$E_{uu} = 0.6 \dis^{2/3}(k_x)^{-5/3}$ for (b). }
\caption{LES equilibrium solutions, $(A_{xz},A_{yz})=(3,1.33)$ and $R_S=50.5$. % In (a,b): \solid, Lower branch, as in figure~\ref{fig:UPO_LES_flat_3d}(a); \dashed, upper branch, as in figure~\ref{fig:UPO_LES_flat_3d}(c). % (a) Mean streamwise velocity. The thin diagonal is $U = Sy$. (b) Resolved strain rate. % (c) Stability eigenvalues $(\mu_r + \ii \mu_i)/S$ of the equilibria in (a,b): $\triangledown$ (blue), lower branch; $\circ$ (red), upper branch. % (d) Temporal evolution of the fluctuation velocity magnitude of symmetric LES initialised from the equilibria in (a,b). % The darker lines are initialised without any disturbance except numerical inaccuracies. \solid (dark blue), Initialised from the lower branch; \dashed (dark red), from the upper branch. The lighter grey lines are initialised from the lower-branch with a small disturbance along the unstable direction corresponding to the real unstable mode. \solid, attracted to the turbulence state; \dashed, attracted to the laminar state. % All attempts to perturb the upper branch along its unstable modes led to laminarisation. The short thick dotted line represents the fluctuation intensity of LES$_\mathrm{s}$ in table~\ref{table:LES-DNS}, using the same box and Reynolds number. }
\caption{Reynolds stress averaged over the $y=0$ plane, $R_S=38.9$, $N_x=64$ and $N_z=32$. % (a) As a function of $A_{yz}$ for $A_{xz}=3$. \solid, $N_y=48$; $\square$\dashed$\square$, $N_y=48$. % (b) As in (a), a function of $A_{xz}$ for $A_{yz}=3$, $N_y = 64$. }
\caption{Velocity fluctuations for $A_{xz} = A_{yz} = 3$. (a) $R_S=50$, (b) $R_S=62.5$ Only half of each plot is shown, using the symmetry in y. Left side is lower branch. Right side is upper branch. \dashed (black), $u^\prime$; \solid (blue), $v^\prime$; \chndot (red), $w^\prime$. }
\caption{(a,b) As in figure \ref{fig:urms_33}, for the momentum balance. \solid (blue), $\langle -u v\rangle_{xz}/u_\tau^2$; \dashed (red), $\langle 2\nu_t \sbar_{xy} \rangle_{xz}/u_\tau^2$; \solid (black), total stress. }
\caption{% (a) The local cross-flow velocity fluctuation intensity $u_\perp$ defined in (\ref{eq:ucross}) at: (black) $y=L_y/2$ ($u_t$), (grey) $y/L_z=0$ ($u_c$). The black dashed line is the mean value, $\bra u_\perp \ket$, and the red dashed lines are $ \langle u_\perp \rangle \pm \sigma$, where $\sigma$ is the standard deviation. They are defined as averages over the centre and top of the box, as explained in the text. % (b) The intermittency factor $(1 - \gamma)$, where $\gamma$ is defined in (\ref{eq:intermittency}): \solidcirc, $R_S = 50.0$; \solidcircf, $R_S = 52.6$; \solidtridown (grey), $R_S=55.5$; \dashedtri, $R_S = 62.5$; \solidsquare (grey), $R_S = 101.6$. }
\caption{(Color online). Deflection angle $\theta$ as a function of impact parameter $h$ for initial vortex-antivortex separation $d=7.8 \xi$ (top) and $d=3.9 \xi$ (bottom) when the target is an impurity (blue line and dots) and a vortex (\red{red} line and dots). The shaded blue areas represent the parameter regions where $\theta$ cannot be defined because one vortex becomes trapped in the impurity. Comparing top and bottom, notice the absence of the trapping regime and the extension of the figure to larger positive values of $h$ (negligible deflection if the target is an impurity, non-negligible if it is a vortex). }
\caption{Performance coefficients for the tandem foils shown in Figure~\ref{fig:tandem}; \textcolor[rgb]{0.8,0,0}{front foil}, \textcolor[rgb]{0,0.5,0.5}{back foil}; dashed lines are the mean values over the cycle.}
\caption{Results for four stopping maneuver cases; $\{\Xi,\theta_{final}\}$ = \textcolor[rgb]{0.5,0,0}{$\{1/2, \pi/2\}$}, \textcolor[rgb]{0,0.5,0}{$\{1/4, \pi/2\}$}, \textcolor[rgb]{0,0.5,0.5}{$\{1/2, \pi\}$}, \textcolor[rgb]{0.5,0,0.5}{$\{1/4, \pi\}$}. Points in (a) show increments of $tU_0/c=1$.}
\caption{\label{fig:AverageErasureProbability} Estimated deletion rate for a randomly selected strand of replicated RNA as a function of {\editcolor total transcription time} for $n = 20000$ and $n = 40000$. Replication times for each type of nucleotide ($\A$, $\C$, $\G$, $\U$) were assumed equal to 0.046 s. Initial deletion rate per nucleotide was $2.3 \times 10^{-7}$. Insertion rate should be the same.}
\caption{\label{fig:EachLetterErasureProbability} Estimated deletion rate for a randomly selected strand of replicated RNA as a function of {\editcolor total transcription time} for $n = 100000$, and different replication rates per nucleotide (see equations (\ref{eqn:DifferentRates1})-(\ref{eqn:DifferentRates2})). Initial deletion rate per nucleotide was $1.7 \times 10^{-5}$. The substantially longer time, compared with Figure \ref{fig:AverageErasureProbability}, is due to the lower replication rate for $\A$.}
\caption{\label{fig:AverageSubstitutionProbability} Estimated substitution rate for a randomly selected strand of replicated RNA as a function of {\editcolor total transcription time} for $n = 20000$ and $n = 40000$. Replication times for each type of nucleotide ($\A$, $\C$, $\G$, $\U$) were assumed equal to 0.046 s. Initial substitution rate per nucleotide was $9.1 \times 10^{-6}$. Indel rate was the same as in Figures 1 and 2.}
\caption{Geometric interpretation of the learning objective. The triangle represents the simplex composed of all agents' probability distributions. The observations of the agents are generated according to a joint probability distribution $\boldsymbol{f}\left(\cdot\right)$. The joint distribution for the agent observations is parametrized by $\theta$. The agent goal is to learn a hypothesis that best describes their observations, which corresponds to the distribution $\ell(\cdot|\theta^*)$ (the closest to the distribution $\boldsymbol{f}\left(\cdot\right)$).}
\caption{The method succeeds in retrieving the parallel Wiener-Hammerstein system parameters about 10\% of the cases. Typical results of successful completion are shown in (a). The output of the identified parallel Wiener-Hammerstein system ({\color{orange}$*$}) reconstructs very accurately the noiseless output signal $y_0$ ({\color{blue}$\circ$}), while the Volterra kernels were computed from the noisy data with 10 dB SNR ({\color{red}$\times$}).}
\caption{The H/L correlation versus time lag for \red{$(t_1,t_2)=(1280,4050)$} seconds. From left to right: before/after GW150914, and both.}
\caption{The H/L cross-correlator as a function of time lag for $(t_1,t_2)=(1280\,{\rm s},4050\,{\rm s})$ with a $50$--$150$\,Hz bandpass. The correlation between the black and red lines is$E(-10\,{\rm ms},10\,{\rm ms})=0.953$.}
\caption{Examples of deformation paths between high-quality eye images for identity \#1, and their Demon measures. Compare the\textbf{low} Demon measure distances for the \textcolor{green}{\textit{visually valid}} paths to the \textbf{higher} distance for the \textcolor{red}{\textit{visually non-valid}} path.}
\caption{Examples of deformation paths between high-quality mouth images for identity \#1, and their Demon distances\GG{$D_T$, Eq. \eqref{Demon_measure}}. Compare the \textbf{low} Demon distance for the \textcolor{green}{\textit{visually valid}} path to the \textbf{higher} distance for the \textcolor{red}{\textit{visually non-valid}} path.}
\caption{Correspondence between visual validity and the (normalized) Demon distance (for eyes of identity \#1). Compare the\textbf{lower} Demon distances, corresponding to the \textcolor{green}{\textit{visually valid}} interpolated images, to the \textbf{higher} Demon distances, corresponding to the \textcolor{red}{\textit{visually non-valid}} images.}
\caption{Normalized NIQE scores for different methods for 17 examples. \textbf{The closer the normalized score is to 1 - the better the quality}. Therefore it can be seen that \textcolor{green}{our method} outperforms \textcolor{orange}{the prior-based brightening method}; which outperforms \textcolor{blue}{the input image}; which outperforms \textcolor{red}{BM3D}.}
\caption{Additional \aastex\symbols}
\caption{{\bf Mean eccentricity $\bar{e}$ (blue markers) and inclination $\bar{i}$ (red markers) as a function of transiting multiplicity $N_p$.} The filled circles show the best fit and the error bars indicate the 68\% confidence interval. As can be seen, it reveals an abrupt transition of $\bar{e}$ rather than a smooth correlation with $N_p$ (see more discussion in {\blue \it{SI Appendix} section 5.1}).}
\caption{{\bf Comparisons in host star (Left Panel) and planet (Right Panel) properties with previous studies.} The LAMOST (this work; black), Seismology (Van Eylen \& Albrecht 2015\cite{VA15}; blue), Occultation (Shabram et al. 2015\cite{Sha15}; magenta) and TTV (Hadden \& Lithwick 2014; red) samples are shown. The numbers in the brackets give the sample size. (See more discussions in Supporting Information Sec. 2.3).}
\caption{Max-$\protect\lambda_2$ algorithm results for adding leaves (top) and path clusters (bottom), using the convex relaxation ($\protect\times$), exhaustive search (\protect\textasteriskcentered) and perturbation heuristic ($+$).}
\caption{Syntactic and semantic evaluation of the parsing models. Left: Simplified labeled F1 and undirected unlabeled F1 on CCGbank, Section 23. Right: Slot filling performance (by number of entities per sentence). {\color{red} Slot-filling results are updated after the camera-ready submission. In the previous version, instead of evaluating if the gold entity is same as the first predicted entity, we mistakenly evaluated if the gold entity is present in the list of predicted answer entities. However, the initial claims are still valid. All other results and discussion are revised.}}
\caption{Our iterative pose optimization in high-dimensional space, schematized here in 2D. We start at an initial pose (\textcolor{red}{$\bm{\times}$}) and want to converge to the ground truth pose (\textcolor{blue}{$\bm{\circ}$}), that maximizes image similarity. Our predictor generates updates for each pose (\textcolor{green}{$\bm{+}$}) that bring us closer. The updates are predicted from the synthesized image of the current pose estimate and the observed depth image. (Best viewed in color)}
\caption{Predicted updates around ground truth joint position, denoted as \protect\tikz\protect\draw[red,fill=red] (0,0) circle (.5ex);. We initialize noisy joint locations around the ground truth location and show the pose updates as vectors predicted by our updater. The start and end of the vectors denote the initial and the updated 3D joint location. The different updates bring us closer to the ground truth joint location. (Best viewed on screen)}
\caption{\textcolor{cyan}{(SEE NOTE ABOVE; IN DISSERTATION BUT NOT SURE IF IT'S HELPFUL SINCE IT LOOKS AT MODELS 1 - 5.) Number of females with science degrees imputed in each combination of $Y$ and $Z$, where at least one of $Y$ and $Z$ is a professional degree. Bar plots the average over $M=50$ multiple imputations, error bars show 2.5$\%$ and 97.5$\%$ quantiles.}}
\caption{\red{Grouped results of all the methods using metrics from Table~\ref{tab:metric} ($rg$ stands for rank group) along with representative values (within brackets). For easy of reading, we recall the methods as described in Table~\ref{tab:algo}. Method 1 -- uni-bi feature with $SVM^{perf}$ with b=1, and AUC. Methods 2, 3, and 4 -- uni-bi feature with $SVM^{perf}$ with b=1, and AUC, KLD, and QuadMean loss functions, respectively. Method 5 -- SVM (Default). Method 11 -- to SVM transduction. Methods 6 and 7 -- uni-bi feature with $SVM^{cost}$ with b=0 and b=1, respectively. Methods 21, 22, and 23 -- w2v row feature with $SVM^{perf}$ with b=1, and AUC, KLD, and QuadMean loss functions, respectively. Methods 31, 32, and 33 -- w2v col feature with $SVM^{perf}$ with b=1, and AUC, KLD, and QuadMean loss functions, respectively. Methods 24 and 25 -- w2v row feature with $SVM^{cost}$ with b=0 and b=1, respectively. Methods 34 and 35 -- w2v col feature with $SVM^{cost}$ with b=0 and b=1, respectively. }}
\caption{\red{P-values of all the metrics corresponding to the Table \ref{tab:rank-group}. The statistics give some insights on how conservative we are in selecting two statistically insignificant methods to put into the same group (Please see Section \ref{sec:exp_result:stat_test} for details). For example, for the Precision metric, any p-value greater than 0.3 may put two methods into the same group for some rank group, even though we choose $\alpha=0.05$ in our case. This is because of the Linear Step Up (LSU) procedure described in \ref{sec:exp_result:stat_test}.} }
\caption{\red{Performance of $SVM^{cost}$ (J, b=0) [Method 7], $SVM^{perf}$ (AUC, b=1) [Method 21], and $SVM^{cost}$ (J, b=1) [Method 25] in Precision, Recall, AUC, and AUPRC metric over three prevalence groups.}}
\caption{Performance of \alg\in different starred thresholds. The lower is the rank, the better is the performance. The three numbers inside the brackets indicate rank group for$\langle$Low, Mid, High$\rangle$ prevalence group. }
\caption{\red{Performance values of \alg\in different starred thresholds. For Precision, Recall, AUC, AUCPR, Burden, Utility, and Yield, the higher the value, the better the performance. On the other hand, for AM and QM errors, the lower the better. The three numbers inside the brackets indicate the value of the metric for$\langle$low, mid, and high$\rangle$ prevalence groups.}}
\caption{Map of the void catalogue (top panels) and the supercluster catalogue (bottom panels) in Y1A1 redMaGiC data and in the Buzzard Y1A1 \redmagic\mock. We applied$\sigma=20~h^{-1}{\rm Mpc}$ initial Gaussian smoothing to the density field as discussed in the main text. The actual area used for the analysis is marked by the dashed rectangles. Colored disks mark the full angular size of the objects, while colored points in the disk centers indicate the redshift assigned to each void's center. The intensity bar shows the redshifts.}
\caption{The right panels show the {\it true} galaxy density distributions about the selected locations (bottom-right: voids, top-right: superclusters), whereas the left panels show the {\it apparent} distributions when the true galaxy positions are distorted by photometric redshift errors (bottom-left: voids, top-left: superclusters). We performed this analysis using the Buzzard simulation of the Y1A1 \redmagic\mock catalogue. Solid circles correspond to a spherical super-structure shape while dashed ellipses mark an elongated model with$R_{\parallel}=2.6 R_{\perp}$ estimated for the Gr08 supervoids.}
\caption{\textit{Left:} Ray traced disc images for two QPO phases (as labelled) visualising our best fit model. The disc inner radius and inclination are equal to our best fit values, and the border of the multi-coloured patches is set by the best fit illumination profile for the relevant QPO phase. The colour scheme of the multi coloured patches encodes blue shifts. \textit{Right:} The best fitting reflection spectrum, zoomed in on the iron line, for the same QPO phases. The QPO phase = $0.2$ cycles iron line (blue) has a boosted blue horn and red wing because the blue and red shifted parts of the disc are preferentially illuminated (top image), in contrast to the QPO phase = $0.4$ iron line (red line), which corresponds to the bottom disc image. Animated versions of these plots can be viewed at and downloaded from \color{blue}\underline{\smash{https://figshare.com/articles/Tomographic\_modelling\_of\_H\_1743-322/3503933}}\color{black}. These animations are designed to be played together.}
\caption{Transverse displacement of the beam's trailing edge for Re$ = 200$, $\text{K}_\text{B} = 0.0015$, and $\text{Fr} = 1.4$. Left panel: $\text{M}_\rho = 0.0001$, middle panel: $\text{M}_\rho = 1.5$, right panel: $\text{M}_\rho = 100$. Present: \protect\solidrule[3mm], Lee \emph{et al.}\\cite{lee}: \protect\mysquare[none], Huang \emph{et al.}\\cite{huang}: \protect\mytriangle[none]. The insert on the left panel is a zoom-in of the $\text{M}_\rho = 0.0001$ case for $t\in [0.5, 2]$. Note the different horizontal axis values on each panels. }
\caption{Photometric redshift point estimates of spectroscopic galaxies in CFHTLS Deep compared to their spectroscopic redshifts. The plot shows all galaxies from the VVDS-Deep, VIPERS, zCOSMOS-bright, zCOSMOS-deep, VUDS and DEEP2 spectroscopic surveys with valid CFHTLS Deep and WIRDS photometry in all bands and $i'<24$. \red{Red, green and blue symbols indicate that the best-fitting templates correspond to early-type, late-type and starburst galaxies in the classification of \citet{2005ApJ...631..126D}}.}
\caption{Left: A typical pulse trace on the (red) direct coupling and (blue) capacitive coupling configurations for \oneinchsquare pads (upper panels) and 0.44 cm-wide micro-strips on a 0.69 cm pitch (lower panels). Two traces are shown for each configuration: the pad or strip directly beneath the laser head and a neighboring pad or strip. Right: The pulse traces normalized to the same amplitude, showing that the high-frequency response is unchanged by the capacitive coupling through the metal film.}
\caption{Filter response functions associated with Pad\'e filter defined by eqn.~\ref{fpade}. -- -- --: $\hat{G}(k; \alpha=0)$, --~$\cdot$~--~$\cdot$~--: $(\hat{Q}_{ADM}(k)-1)/J; \alpha=0, J=5)$, {\bf ------}: $\hat{E}(k; \alpha=0)$, {\color{blue}------}: $\hat{E}(k; \alpha=-0.2)$, {\color{red}--$\bullet$--}: $\hat{E}(k; \alpha=0.2)$}
\caption{Filtering of formulas designed by \citet{Chakravorty2010}. ------: first derivative; {\color{red}-- -- --}: interpolation; {\color{blue}--~$\cdot$~--~$\cdot$~--}: explicit filter $E$. }
\caption{Filtering of standard 6th-order compact difference formula (------) and 10th-order filter F10 with $\alpha=0.498$ ({\color{blue}--~$\cdot$~--~$\cdot$~--}).}
\caption{\textcolor{review_color}{Dimensional swimming speed in a Newtonian fluid for \textit{stiff,} $\kb=10.0,$ and \textit{moderately soft,} $\kb=1.0,$ swimmers over a range of frequencies $\per^{-1}.$ Inset figures show shapes of swimmer over a period at the highest computed frequency for both soft and stiff swimmers. }}
\caption{(a) Normalized swimming speed as a function of $\De.$ Curves generated by varying only relaxation time (for fixed period $\per=1$) or stroke period (for fixed relaxation time $\lambda=0.5$). Body stiffness is fixed: $\kb=1.0$. \textcolor{review_color}{(b) Normalized swimming speed as a function of both $\lambda$ and $T,$ with contours overlayed for constant $\De$ values. Dashed lines correspond to the locations of the data in \ref{fig_samede} (a). } % %Other relevant stroke parameters: $A_t = 5, A_h = 2$ and $\kb=1,$ see section \ref{model}. % }
\caption{Top-1 error rates (\%) on CIFAR datasets. All the results of PyramidNets are produced with additive PyramidNets, and $\alpha$ denotes the widening factor. ``Output Feat. Dim.'' denotes the feature dimension of just before the last softmax classifier. The best results are highlighted in {\bf \color{red} red}.}
\caption{Additional \aastex\symbols}
\caption{Additional \aastex\symbols}
\caption{ \label{fig:sigma} Imaginary part of the self energy calculated with the full vertex (FV) functions and with spin-fermion (SF) models for antinodal point (\textcolor{red}{$\mathbf{1}$}) and nodal point (\textcolor{blue}{$\mathbf{2}$}) for $U=14$ in (a) and (b), $U=6, 18$ in (c) and (d). Non-local corrections to the DMFT self energy, $\delta\Sigma_{FV/SF}(\bfk, i\omega) = \Sigma_{\text{FV}/\text{SF}}(\bfk, i\omega) - \Sigma(i\omega)$ are presented with the local self energy $\Sigma(i\omega)$. Also shown are the values of the effective Coulomb interaction $U_{eff}$ for the SF model (Eq. (\ref{eq:Sig_sf})). }
\caption{\label{fig:NCE-regression} Compressing the output of neural LM. We apply NCE to estimate the parameters of the \Prediction{} sub-network (dashed round rectangle). The SpUnnrmProb layer outputs a \textbf{sp}arse, \textbf{unn}o\textbf{rm}alized \textbf{prob}ability of the next word. By ``sparsity,'' we mean that, in NCE, the probability is computed for only the ``true'' next word (red) and a few generated negative samples.}
\caption{Additional \aastex\symbols}
\caption{We validate our SED-fitting method by comparing the reddening A$_{\rm v}$ using \texttt{sedFit} for~\nstarsred~OB stars in the Cep OB3 cloud (Y-axis) to the directly observed A$_{\rm v}$ for these stars using the IUE satellite \citep[See Table 2, 4th column,][]{massa84}. We find a statistically insignificant average offset of $-0.13\pm0.17$ mag. }
\caption{Additional \aastex\symbols}
\caption{Computational results for Example 10 in \cite{Ferrer2014}. The running times (in seconds) for DCECAM and DCPA are taken from \cite{Ferrer2014} and are obtained on a computer with Intel\textregistered\Core\texttrademark\i5-3470S CPU with 2.9 GHz.}
\caption{(Colour available online) \textit{(a)} Measurements of viscosity $\eta$ as a function of shear rate $\dot{\gamma}$ for the Newtonian buffer solution M9 (closed symbols), CMC solutions in M9 (blue open symbols, concentrations from bottom to top: 300, 500, 1000, 1500, 2000, and 3000 ppm,~\citet{Sznitman2010PoF}), and halocarbon oil mixtures (grey open symbols, from bottom to top: 100\% H27, 44\% H700, 61\% H700, 78\% H700, and 95\% H700 by weight,~\citet{Shen2011}). \textit{(b)} Measurements of viscosity $\eta$ as a function of shear rate $\dot{\gamma}$ for shear-thinning solutions of XG in M9 (from bottom to top: 50, 100, 200, 300, 500, 1000, 2000, and 3000 ppm). The solid black line shows a fit to the Carreau-Yasuda model (Eq.~\ref{Carreau}) \textit{(c)} Carreau timescale $\lambda_{\mbox{\textit{Cr}}}$ as a function of concentration ({\color{red}$\boldsymbol{\vartriangle}$}) and power law index $n$ ({\color{blue}$\boldsymbol{\circ}$}) as a function of concentration $c_{\mathrm{XG}}$.}
\caption{(Colour available online) \textit{(a)} Cost of swimming as a function of zero-shear viscosity $\eta_0$ using each side of Eq.~\ref{energyBalance}. For Newtonian fluids: mechanical power (${\square}$, from \citet{Sznitman2010PoF}), viscous dissipation rate ({\color{yellow}$\boldsymbol{\pentagon}$}, buffer only), and the scaling $P \sim \eta U^2$ (solid line) calculated from our kinematics data~\citep{Gagnon2014}. For shear-thinning fluids: mechanical power ({\color{blue}$\boldsymbol{\circ }$}), viscous dissipation rate ({\color{red}${\vartriangle}$}), and the scalings $P \sim \eta_0 U^2$ (dash-dot line) and $P \sim \eta_{\mathrm{eff}} U^2$ (dashed line). \textit{(b)} Mechanical power and viscous dissipation rate replotted versus effective viscosity $\eta_{\mathrm{eff}}$.}
\caption{(Colour available online) Normalized mechanical power ({\color{blue}$\boldsymbol{\circ}$}) and viscous dissipation rate ({\color{red}$\boldsymbol{\vartriangle}$}) as a function of $\mbox{\textit{Cr}}_k$; the dashed line represents the Newtonian case. The transition from $P/P_N \approx 1$ to $P/P_N < 1$ occurs at $\mbox{\textit{Cr}}_k = \mathcal{O}(1)$. The solid black line is the theoretical scaling generated from our rheology and kinematics data, given by Eq.~\eqref{prediction}~\citep{Li2015}.}
\caption{G phonon intensity on the anti-Stokes side at various delay times with pump fluence of (a) 2.1 $mJ/cm^2$ (blue) and (b) 0.15 $mJ/cm^2$ (red, same data points as in Fig. \ref{Ph_T}\color{blue}(a) \color{black}). The solid and dashed curves are the fits using a biexponential function $-exp[-(t-t_0)/\tau_{growth}]+exp[-(t-t_0)/2400]$. The fitting parameter $\tau_{growth}$ is $100\pm 9$ fs for (a) and $210\pm 38$ fs for (b), and $t_0$ is $-23\pm7$ fs for (a) and $10\pm 13$ fs for (b). The inset shows overplotting of data with different pump powers. \label{fluence_dep}}
\caption{\label{fig:setup} The electrodes are simulated as arrays with widths ${L_x=L_y \simeq 10b}$. $m_i$ denotes the typical number of voxels in the array in the $i$-direction for $i=x,y,z$. (a) The initial setup where a single ion track (track centre marked with a solid line) with Gaussian radius $b$ is rotated \red{by} an angle $\theta \in [0,\pi/2]$ to the electric field between the electrodes with spacing $d$. (b) The setup for the simulation of a continuous beam where three ion tracks have been initialised at different spatial positions ($x,y$) and times $t$. }
\caption{\label{Fig5} (Color online) Distribution of Hartree energy of the sites on the domain wall ($\varepsilon_{i}^{wall}$ for $L=32$ ($\bullet$) and $L=64$ ({\color{Green}$\blacksquare$})) and on sites inside the domain wall where domain wall sites are excluded ($\varepsilon_{i}^{inside}$ for $L=32$ ({\color{Red}$\diamond$}) and $L=64$ ({\color{Blue}$\bigtriangleup$})). Inset shows the domain energy vs perimeter of the domain. $\eta = 0.0335$ for $L=64$.}
\caption{Spectrum of WGPs (upper row) and GGPs (first two bottom panels) sustained in different graphene-covered triangular configurations, with different opening angles, $2\vp$ (indicated in the insets), as given by Eq. \ref{spectrum:wedge} [we take $E_F=0.4$ eV]. The solid black line represents the dispersion of GSPs in a flat interface and serves as reference. The colored $\bigtriangledown$, $\bigcirc$ and $\bigtriangleup$ data points figuring in the upper row correspond to the results for the WGP dispersion as obtained from full-wave numerical simulations (COMSOL's finite-element method). The insets' shading \legendsquare{v} represents an insulator with $\ep=4$, whereas the white regions denote a medium with $\ep=1$ (e.g. air). The last panel shows the spectrum of GSPs guided along a V-shaped graphene embedded in a homogeneous medium with $\ep=2.5$ (shaded in the inset as \legendsquare{b}). }
\caption{(a) Level sets {\color{Orange} $\hat{\mathcal{M}}$}($\mathfrak{c}$) residing in kernel spaces $\setM$ induced by a loss function {\color{Blue} ${\mathcal{L}}$} from $\setM$ to a set of error values {\color{Periwinkle} ${\mathcal{N}}$}. (b) Zero level sets {\color{Red} $\hat{\mathcal{M}}$}($\mathfrak{c}=0$) contain critical kernels at which the gradient vanishes (see {Definition~\ref{def:EK}}, and {Theorem~\ref{theorem:rank_dim}}). }
\caption{ $\setM$ is a smooth $m$ dim. kernel manifold. {\color{Orange} $\hat{\mathcal{M}}$} is a $k$ dim. submanifold embedded in $\setM$. ({$\color{ForestGreen} \mathring{\mathcal{M}}$} ,{$\color{ForestGreen} \phi$} ) is a smooth chart of {\color{Orange} $\hat{\mathcal{M}}$} centered at a kernel $\omega \in \setM$. ${\color{Blue} \tilde{\mathcal{M}}} ={\color{ForestGreen} \mathring{\mathcal{M}}} \cap {\color{Orange} \hat{\mathcal{M}}} $ is a $k$-slice of ${\color{ForestGreen} \mathring{\mathcal{M}}}$ (see {Definition~\ref{def:EK}}). $({\color{Blue} \tilde{\mathcal{M}}, \psi})$ is a chart of $\setM$ centered at $\omega \in \setM$, and {{\color{Cerulean} $\pi$}: {\color{ForestGreen} $\mathbb{R}^m$}{\color{NavyBlue} $\to$}{\color{Blue} $\mathbb{R}^k$}} is the projection onto the first $k$ coordinates. Two charts of $\setM$ and {\color{Orange} $\hat{\mathcal{M}}$} are associated by {\color{Blue} $\psi$}= {\color{NavyBlue} $\pi$} $\circ$ {\color{OliveGreen} $\phi$}({\color{Blue} $\tilde{\mathcal{M}}$}). ${\color{Blue} \tilde{\mathcal{M}}} $ is a level set of a loss function ${\color{red} \mathring{\setL}}: {\color{ForestGreen} \mathring{\mathcal{M}}} \to {\color{red} \mathbb{R}^{m-k}}$ (see {Proposition~\ref{prop:level_sets})}. }
\caption{Results for Cifar-10 without DA. The results marked by $\dagger$ indicate the results reproduced by our implementation of the associated algorithm using the code provided by the authors of the related work. Classification error obtained using the baseline CNN is marked by {\color{red} red}, and our best error is marked by {\color{blue} blue}. }
\caption{\label{fig:SPIKE_Order_Surro_57_Mix_Poisson_SF} SPIKE- and Spike Train Order analysis for Poisson spike train interspersed with \highlight{spike trains that contain random spikes but also} a perfect inverse synfire pattern. The order contained within the synfire pattern spike train is distinct enough to make the Synfire Indicator for the sorted spike trains statistically significant.}
\caption{\label{fig:SPIKE_Order_Surro_109_HM_5_1_SN} SPIKE-Order for real data recorded in an acute hippocampal slice from a juvenile mouse. \highlight{Note how the color-coding of the spikes according to their SPIKE-order $D$ helps to overcome the low temporal resolution of the Figure and to resolve the spike order within the GDPs.} (a) Initially the spike trains are sorted according to their firing rate starting with the most sparse spike trains. \highlight{The messy color-patterns reveal that this is completely uncorrelated to the spike order within the GDPs.} (f) After sorting, there is a fairly consistent transition from spike trains with predominantly leading spikes (red) in the GDPs to spike trains with predominantly following spikes (blue). }
\caption{\label{fig:SPIKE_Order_Surro_159_HM_5_1_SN} SPIKE-Order for the real data already analyzed in Fig. \ref{fig:SPIKE_Order_Surro_109_HM_5_1_SN} but this time the analysis of SPIKE-Order was restricted to spikes with a SPIKE-Synchronization value of at least $0.7$. This simple thresholding allows to focus the analysis on the reliable events and to disregard all spikes between the events \highlight{(these are not colored and thus remain black)}. This results in an increase of the overall value of SPIKE-Order from $0.284$ to $0.438$. }
\caption{\label{fig:2D-plot-ColorCoding-159} Projection of the optimized Spike Train Order on the 2D-\highlight{photo} of the hippocampal slice. The Regions of Interest (ROIs) which denote filled and identified cells in the CA3 region are color-coded from leader (index $1$, red) to follower (index $163$, blue) using the optimized Spike Train Order of Fig. \ref{fig:SPIKE_Order_Surro_159_HM_5_1_SN}. The very first leader (lower right) and the very last follower (upper left) are marked by filled contours. }
\caption{\elnino\3.4 region (rectangle) and the small strip around the equator used for the analysis in this work (dashed line). Color coding represents daily SST anomaly (in$\degC{}$) on November~11, 2015.}
\caption{Both panels: Example spectra of two stars which have the same temperature ($\sim4700$ K) but differing molecular band strengths, C47795 (green spectrum; weaker CH and CN bands) and C20922 (blue spectrum; stronger CH and CN bands). Indicated in both panels are the regions used for the spectral indices. For each spectral index, the closest regions in grey were used for the continua. Top panel: the \cnblue\CN spectral index and the\chold\index in red. Bottom panel: the\hkdash\and\sch\indices in red.}
\caption{Top: Comparison of the \nfe\abundances determined by this study and\citet{Carretta2014a}. The line is the one-to-one line. Bottom: \nfe\abundances from this work and\citet{Carretta2014a} compared to the \dcnblue\spectral index strength determined from our spectra. It would be expected that there would some correlation between the\nfe\and\dcnblue\strength of stars.}
\caption{The \hkdash\indices determined for the sample of stars in NGC 1851. The symbols are as defined in Section\ref{sec:newmembers} and Figure \ref{fig:sky_plot}.}
\caption{The dependence of \sch\on the magnitude of the star. The symbols are as defined in Section\ref{sec:newmembers} and Figure \ref{fig:sky_plot}. The four new peculiar AGB members are clear outliers from the rest of the cluster stars.}
\caption{The \cnblue\index results. The symbols are as defined in Section\ref{sec:newmembers} and Figure \ref{fig:sky_plot}. One of the extratidal stars is coincident on this plot with a tidal star. Top: The dependence of \cnblue\on the magnitude of the star. The black line represents an empirical minimum value of the CN index at a given J magnitude. This line is used to convert\cnblue\to\dcnblue. Bottom: The temperature correction applied ($\dcnblue = \cnblue - [ -0.078\times J+0.874]$). We have not plotted the field stars as the temperature correction cannot be applied to them.}
\caption{Top panel: Comparison of the distribution of \cnblue\determined by this work (shaded green KDE with$\sigma = 0.15$) and \citet{Campbell2012} (light blue line; $\sigma = 0.2$). \citet{Campbell2012} suggested that there were four peaks. One of those peaks at about 0.25 appears to have been the result of small number statistics. It should be noted that the shared peak at 0 is just the result of the zeroing of the \cnblue. Bottom panel: For reference the \cnblue\of the stars against the\sch\of the stars, coded with the CN-CH populations.}
\caption{Change in \cfe\with V magnitude. The symbols are as defined in Figure\ref{fig:j_dCN_kde_all}. The solid line is the best fit line for the primordial population, and the dashed line is for the intermediate population. Data points with arrows indicate those stars for which the carbon abundance is only a lower limit.}
\caption{Top: The \hkdash\indices of stars coded with their CN-CH populations. The symbols are as defined in Figure\ref{fig:j_dCN_kde_all}. The straight line is the best fit to the data. Middle: The \hkdash\values adjusted using the straight line from the top panel (to give$\delta\hkdash$). Bottom: KDE of the $\delta\hkdash$ of the four populations. The key finding is that the two most CN-rich populations both generally have stronger Calcium H \& K lines. These lines are a metallicity indicator.}
\caption{\color{Gray} \textbf{Random component analysis of the HSA structures.} The Figure reports a random component analysis on the HSA structures contained in the dataset described in the text. The HSA structures with bound fatty acids are reported as solid (black) circles, whereas the structures without bound fatty acids are reported as void (white) circles. The algorithm clearly permits to differentiate two clusters of structures in the dataset, and the discriminant is the presence of absence, respectively, of bound fatty acids. Two similar cluster of structures have been obtained in all the random component analysis calculations carried out on the HSA dataset (see Supplementary Information).}
\caption{\color{Gray} \textbf{The Iris dataset.} Principal component analysis (left) and random component analysis (right) of the Iris dataset are reported. The Iris dataset is a simple but classical benchmark for the clustering algorithms. This dataset contains 150 entries, 50 for each of the species {\it{Iris virginica}} (black), {\it{Iris setosa}} (gray) and {\it{Iris versicolor}} (white). Both algorithms easily differentiate the {\it{Iris setosa}} cluster, whereas the other species can be only partially discriminated by the algorithms of this class. The full set of random component analysis carried out on the Iris dataset shows similar clustering results and it is reported in Supplementary Information.}
\caption{Predictive analysis on the first 50 votes: In the selection phase, the CVP shows better negative log-likelihood in almost all forums. In the voting phase, the full model shows better negative log-likelihood than all subsets of features. Quality analysis at the final snapshot: Smaller residuals and bumpiness show that the order based on the estimated quality $q_{ij}$ more coherently correlates with the average sentiments of the associated comments than the order by display rank. (SOF=StackOverflow, OF=Overflow, rest=Exchange, {\color{sig001}Blue}: $p \leq 0.001$, {\color{sig01}Green}: $p \leq 0.01$, {\color{sig05}Red}: $p \leq 0.05$)}
\caption{Success score of OPE with 11 attributes. The number after each attribute name is the number of sequences. The \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green} fonts indicate the best, second and third performance. }
\caption{Precision scores of OPE with 11 attributes. The number after each attribute name is the number of sequences. The \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green} fonts indicate the best, second and third performance. }
\caption{Success scores of TRE with 11 attributes. The number after each attribute name is the number of sequences. The \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green} fonts indicate the best, second and third performance. }
\caption{Precision scores of TRE with 11 attributes. The number after each attribute name is the number of sequences. The \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green} fonts indicate the best, second and third performance. }
\caption{Success scores of SRE with 11 attributes. The number after each attribute name is the number of sequences. The \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green} fonts indicate the best, second and third performance. }
\caption{Precision scores of SRE with 11 attributes. The number after each attribute name is the number of sequences. The \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green} fonts indicate the best, second and third performance. }
\caption{Memorization time for \colorbox{ObjsColor}{Objects}, \colorbox{ImColor}{Image}, and \colorbox{WordsColor}{Words}}
\caption{Time (\emph{in second}) to login by children and by adults using \colorbox{ObjsColor}{Objects}, \colorbox{ImColor}{Image}, and \colorbox{WordsColor}{Words}}
\caption{Successful login attempts by children and by adults using \colorbox{ObjsColor}{Objects}, \colorbox{ImColor}{Image}, and \colorbox{WordsColor}{Words}}
\caption{Number of correct tiles selected on each login attempt using \colorbox{ObjsColor}{Objects}, \colorbox{ImColor}{Image}, and \colorbox{WordsColor}{Words}}
\caption{Interview Results showing the number of participants choosing \colorbox{ObjsColor}{Objects}, \colorbox{ImColor}{Image}, and \colorbox{WordsColor}{Words}}
\caption{ Experimental results in random unit hybercube scenarios for \protect\tikz{\protect\node[fill=myblue,draw=black]{};}\; vertex batching\protect\tikz{\protect\node[fill=myred,draw=black]{};}\; edge batching, and\protect\tikz{\protect\node[fill=mygreen,draw=black]{};}\; hybrid batching. The$y$-axis is the ratio between the length of the path produced by the algorithm and length of $\gamma^*$ (the shortest path on $\calG$) for that problem. The naive strategy of searching the complete graph requires the following times to find a solution - (\subref{fig:results_2d_easy}) $\mathbf{44}$s, (\subref{fig:results_2d_hard}) $\mathbf{200}$s, (\subref{fig:results_4d_easy}) $\mathbf{12}$s and (\subref{fig:results_4d_hard}) $\mathbf{56}$s. In each case this is significantly more than the time for any other strategy to reach the optimum. Figure best viewed in color.}
\caption{Entangling power $e_p$ after $n$ actions of a controlled unitary gate $U \in U(N^2)$ plotted for $N=10$. Data from one realization of $U^{(n)}$ (\textcolor{red}{$\square$}) and one realization of $[(U^A\otimes U^B)\,U]^n$ ($\triangle$) are shown. Values averaged over local gates according to Eq.~\eqref{eqn:epxi} are indicated by (\textcolor{blue}{$\times$}), while the insets show deviations from this average. Horizontal line denotes the average over the unitary group. The case $U^n$ with no interlacing gates, for which entangling power do not converge to typical values, is marked by (\textcolor{Brown}{$\Diamond$}).}
\caption{Benchmarks of the goal-driven w.r.t. complete prefix of 1-safe Petri nets. For each model, the number of places $|P|$ and transitions $|T|$ is given. The strategy decides when the model reduction should be performed; the number of calls to the reduction procedure is indicated in the column ``Nb reductions''. Computation times were obtained on an Intel\textregistered{} Core\texttrademark{} i7 3.4GHz CPU with 16GB RAM. N/A: Non Applicable; $^*$: out-of-memory computation (with mole \cite{mole}, with the same ordering for extensions as our implementation), the indicated prefix size is only a lower bound.}
\caption{\label{fig7} Density matrix at the quasi stationary regime for the finite temperature ($D=1$) and the zero temperature ($D=0$) for simultaneous kicking. The period of the kick used was {\color{red}$\tauk=4\pi$} and the kick strength was settled to $\pi/2$. The parameters of the linear chain are: $\delta_B=0.5$, $\delta_C=0.25$, $\chi=0.15$, $\chi'=0.1$ and $\beta=0.1$.}
\caption{ \coloronline (a) The single site total angular momentum is zero in both the $jj$ and $LS$ coupling schemes. (b) Schematic phase diagram of the spin-orbital model appearing in \eqref{pertLatticeHamiltonian} pitting spin-orbit coupling $\lambda$ against superexchange $J_\mathrm{SE}$ where $\lambda$ is the spin-orbit coupling energy scale and $J_\mathrm{SE}$ is the superexchange energy scale with $z$ being the coordination number. Starting with a van Vleck phase with no atomic moments at large $\lambda$ we find a triplon condensate at $\veck{k} =\vec\pi$ for all values of the Hund's coupling $J_H/U$. The intermediate regime where $\lambda \approx z J_\mathrm{SE}$ has not been explored. At large $J_\mathrm{SE}$ we obtain effective magnetic Hamiltonians that have isotropic Heisenberg spin interactions (antiferromagnetic for small $J_H/U$ and ferromagnetic for large $J_H/U$) but the orbital interactions are more complex and anisotropic. We expect novel magnetic phases arising from orbital frustration in the intermediate and large $J_\mathrm{SE}/\lambda$ regimes. \label{AfmFmSchematicPhases} }
\caption{ \coloronline (a) The $N_\mathrm{orb}=2$ model is an approximation of oxygen mediated electron hopping between $t_{2g}$ orbitals in a simple cubic lattice. Both $d_{xy}$ and $d_{yz}$ orbitals participate in hopping along the $y$ direction. (b) The $N_\mathrm{orb}=1$ model is an approximation of direct hopping between $t_{2g}$ orbitals on the face of a face-centered cubic lattice. The $d_{xy}$ orbitals are most relevant for hopping in the $xy$ plane. \label{latticeGeometriesFigure} }
\caption{ \coloronline Orbital frustration is graphically illustrated for the $N_\mathrm{orb} = 2$ model. The orbitals shown on the vertices of the plaquette are the doubly occupied orbital on each site in a square lattice. Once the first bond, labeled as 1, is chosen to be of a particular type, either (a) AF or (b) F, the next bonds, labeled as 2, are immediately fixed by this choice. The result is that the last bond on the plaquette, labeled as 3, then takes a configuration which is neither the most energetically favorable AF bond nor the most energetically favorable F bond. \label{orbital_frustration_AF_F} }
\caption{ \coloronline The triplon condensation mechanism is graphically illustrated. When there exists a triplon excitation on a site, superexchange can move the excitation to neighboring sites. This effective hopping causes the triplon's energy to disperse in $k$-space. When superexchange becomes large enough, condensation of triplon excitations occurs as the bottom of the triplon band becomes lower in energy than the original $J_i=0$ level. \iffalse (b) Schematically, the finite temperature phase diagram is that of Bose-Einstein condensation. For large values of $J_\mathrm{SE} / \lambda$, the local F spin interactions discussed in the previous sections can cause F ordering. \fi \label{CondensationMechanismFigure} }
\caption{ \coloronline The flow of angular momentum is graphically shown where ingoing arrows are incoming angular momentum and outgoing arrows are outgoing angular momenta. Wigner-3j symbols and Clebsch-Gordan coefficients are vertices with three legs while the scalar contraction of four Wigner-3j symbols (right) is a Wigner-6j symbol.\cite{YLV1962,Brink1962angular} (a) Equation \eqref{LSprojectToTDaggerT} is shown in graphical form for the $\cre{T}{\zeta}$ part of the equation. A $J=0$ state is decomposed into its $L=1$ and $S=1$ components which are acted on by the $(L_i)_m^l$ and $(S_i)_\sigma^s$ operators. The resulting $L=1$ and $S=1$ are combined together to give a $J=1$ state with quantum number $\zeta=m+\sigma$. (b) The projection of equation \eqref{generalMultipoleForm} to \eqref{generalMultipoleHopping} conserves angular momentum. Equation \eqref{generalMultipolePairing} will appear similarly except that $1m'$ and $1m$ add to yield $LM$ instead. \label{AngularMomentumFigure} }
\caption{Classification of closeups of Vahingen (1--3) and Potsdam (4--6) validation sets. Classes: Impervious surface (white), Building {\color{blue} (blue)}, Low veget. {\color{cyan} (cyan)}, Tree {\color{Green} (green)}, Car {\color{Dandelion} (yellow)}, Clutter {\color{red} (red)}.}
\caption{ \label{fig:scatteringDkernel} Molten salt (a): Incoherent intermediate scattering function for different wavevectors (circles). Punctured line is $f(k,t) = e^{-\frac{1}{6}\langle \Delta r^2 \rangle k^2 t}$, where $\langle \Delta r^2\rangle$ is the mean square displacement. (b) The diffusion kernel at different temperatures. Punctured lines are best fit to Eq.(\ref{eq:lorentz}). \color{red}Parameter values are for $T=0.0177, 1.0177$ and $T_\infty$, respectively: $D_0=0.011, 0.84, 0.92$, $\alpha=0.0073, 0.69, 0.78$, and $\beta=1.72, 1.14, 1.12$.\color{black}}
\caption{ \label{fig:cmp} [Color online] Non-equilibrium results for molten salt (a) Charge density profiles for $T=0.0177$. Lines are sine functions with amplitudes $\w{\rho}_q=0.0072$ and $0.093$, values obtained from a spectral analysis. (b) Charge density amplitudes for all three temperatures and for $m=0$. Symbols are simulation results. Lines are predictions from the theory, \color{red} Eq. (\ref{eq:FourierCompCharge}). \color{black} }
\caption{Theoretical amplitudes of emitted four-wave mixing signal intensity assuming that 1/$e^{2}$ diameters of incoming beams equal to 400 $\mu$m. Panel (a) portrays the Gaussian intensity distribution without taking into account the longitudinal phase-matching condition. Panel (b) shows the value of the phase-matching factor $\mathrm{sinc^{2}}\left(\frac{k_{3}L}{2}\delta\right)$. The black circle corresponds to a region where the longitudinal phase matching is perfect. Variations of the refractive index induced by changing laser detuning cause the radius of this circle to change as well, which in turn impacts the direction of emission. Panel (c) presents product of (a) and (b), reflecting small change in shape and radial translation. \label{fig:symul}} \end{figure} The incoming laser beams indexed by $j=1,2,4$ are assumed to be Gaussian with equal diameters of $2\sigma$. Their frequencies $\omega_{i}$ determine their wavevector lengths $k_{j}=n_{j}\omega_{j}/c$, with $n_{j}$ being the refractive index for respective beam. They propagate at small angles $\boldsymbol{\varphi}_{j}=(\varphi_{jx},\varphi_{jy})$ with respect to the $z$-axis, so it is convenient to replace wavevectors with angles $\mathbf{k}_{j\perp}\approx\boldsymbol{\varphi}_{j}k_{j}$, $k_{iz}\approx(1-|\boldsymbol{\varphi}_{j}|^{2}/2)k_{j}$. The desired final result is the angular distribution of emitted four-wave mixing signal $\widetilde{A}_{3}(\boldsymbol{\varphi}_{3},L)$ and its total intensity $I_{3}=\int|\widetilde{A}_{3}(\boldsymbol{\varphi}_{3},L)|^{2}\textrm{d}\boldsymbol{\varphi}_{3}.$ Let us assume that the amplitude of incident beam $j$ is maximal along the $z$-axis, neglecting shifts of beams centers: \begin{equation} A_{j}(x,y,z)=a_{j}\exp{\left(-\frac{x^{2}+y^{2}}{2\sigma^{2}}\right)}\exp{(i\, \mathbf{k}_{j}\cdot\mathbf{r})}. \end{equation} Corresponding nonlinear atomic polarization generated in such a system is: \begin{equation} P_{NL}(x,y,z)=\chi_{3}^{(3)}A_{1}A_{2}A_{4}^{*}=a_{3}\exp{\left(-\frac{3(x^{2}+y^{2})}{2\sigma^{2}}\right)}\exp{(i\, \mathbf{K}\cdot\mathbf{r})},\label{eq:p4wm} \end{equation} where $a_{3}=\chi_{3}^{(3)}a_{1}a_{2}a_{4}^{*}$ and the resulting polarization wavevector $\mathbf{K}=\mathbf{k}_{1}+\mathbf{k}_{2}-\mathbf{k}_{4}.$ This wavevector is oriented at an angle $\boldsymbol{\Phi}$ to the $z$-axis, so transverse coordinates are given by \begin{equation} \mathbf{K}{}_{\perp}\approx k_{3}\boldsymbol{\Phi}=k_{1}\boldsymbol{\varphi}_{1}+k_{2}\boldsymbol{\varphi}_{2}-k_{4}\boldsymbol{\varphi}_{4}\label{eq:duzefi} \end{equation} Inserting the expression for nonlinear polarization [Eq. (\ref{eq:p4wm})] with the wavevector given by Eq.~(\ref{eq:duzefi}) to the integral [Eq.~(\ref{eq:obw})], we obtain the final distribution of the four-wave mixing signal: \begin{equation} \widetilde{A}_{3}(\boldsymbol{\varphi}_{3},L)\sim a_{3}\exp\left(-\frac{k_{3}^{2}\sigma^{2}|\boldsymbol{\varphi}_{3}-\boldsymbol{\Phi}|^{2}}{6}\right)\mathrm{sinc}\left(\frac{k_{3}L}{2}\delta\right),\label{eq:amplituda} \end{equation} \begin{equation} \delta=\underbrace{\frac{k_{1}}{k_{3}}\left(1-\frac{|\boldsymbol{\varphi}_{1}|^{2}}{2}\right)+\frac{k_{2}}{k_{3}}\left(1-\frac{|\boldsymbol{\varphi}_{2}|^{2}}{2}\right)-\frac{k_{4}}{k_{3}}\left(1-\frac{|\boldsymbol{\varphi}_{4}|^{2}}{2}\right)-1}_{-\theta_{3}^{2}/2}+\frac{|\boldsymbol{\varphi}_{3}|^{2}}{2}.\label{eq:delta} \end{equation} The above formula for the distribution consists of two factors: (a) the Gaussian factor that depends only on transverse variables and is the result which is obtained if we neglect the longitudinal phase matching and (b) the $\mathrm{sinc}^{2}\left(k_{3}L\delta/2\right)$ factor that represents the longitudinal phase-mismatch. Fig. \ref{fig:symul} shows example theoretical intensity distribution of emitted beam without taking into account the longitudinal phase-matching condition (i.e. the Gaussian factor, at the left) and the ring-shaped phase-matching factor (at the right). Shape of the four-wave mixing intensity distribution depends on angles and wavevector lengths of incoming beams. The Gaussian factor is virtually independent of frequencies. Indeed, small changes of wavevector lengths shift the Gaussian factor by much less than its width $k_{3}\sigma$, so they can be neglected. However, since $k_{3}L\approx10^{5}$, we need very small $\delta$ {[}see Eq. (\ref{eq:delta}){]} for efficient phase matching. Equation (\ref{eq:delta}) can be rewritten as $\delta=|\boldsymbol\varphi_{3}|^{2}/2-\theta_{3}^{2}/2$ which for $\delta=0$ describes the cone of perfect phase matching. In the transverse plane, the best phase-matching is achieved on the ring with radius $\theta_{3}$ with angular width of approximately \begin{equation} \label{eq:deltatheta} \Delta\theta_3=\pi/(2k_{3}L\theta_{3}). \end{equation} We now consider changes of the radius $\theta_{3}$ as a function of incident and emitted beams wavevectors. The derivative over $k_{j}$ is \begin{equation} \frac{\partial\theta_{3}}{\partial k_{j}}\approx\pm\frac{1}{k_{j}\theta_{3}}. \end{equation} A typical value we find for $\theta_{3}$ in the experiment is of the order of $10^{-3}$ rad, so derivative is of the order of $10^{-4}$ m$^{-1}$. Therefore, change of refractive index of the order of $10^{-5}$ causes change of $\theta_{3}$ of the order of $10^{-2}$ rad, much more than phase-matching ring width. By analogy we can calculate the derivative over incident beam angles $\varphi_{jx,y}$: \begin{equation} \frac{\partial\theta_{3}}{\partial\varphi_{jx,y}}=\pm\frac{k_{j}\varphi_{jx,y}}{k_{3}\theta_{30}}\label{eq:dthdfi} \end{equation} Consequently, unless the spatial configuration is properly designed, the overlap between phase-matching ring and Gaussian factor from Eq. (\ref{eq:amplituda}) could potentially decrease with any change of $\boldsymbol{\varphi}_{j}$ and resulting change of the phase-matching ring radius $\theta_{3}$. We now consider linear and nonlinear polarizabilities essential to description of the four-wave mixing process. The third order polarizability can be obtained through perturbation chain for a specific case we consider in the experiment \cite{Parniak2015}: \begin{equation} \chi_{3}^{(3)}=\frac{Nd_{12}d_{23}d_{23}^{*}d_{14}^{*}}{4\hbar^{3}\widetilde{\Delta}_{1}\widetilde{\Delta}_{2}\widetilde{\Delta}_{4}^{*}},\label{eq:chi3} \end{equation} where $\widetilde{\Delta_{j}}=\Delta_{j}+i\Gamma_{j}/2$ , $\Gamma_{j}$ is the decay rate, $d_{ij}$ are dipole moments of transitions between respective states and $N$ is the atomic concentration. Additionally, since phase matching plays a crucial role, we take into account dispersion of refractive index. The linear polarizability for beam 1 is: \begin{equation} \chi_{1}=-\frac{N}{\epsilon_{0}}\left(\frac{d_{12}^{2}}{\hbar\widetilde{\Delta}_{1}}-\frac{d_{12}^{2}\Omega_{2}^{2}}{4\hbar\widetilde{\Delta}_{1}^{2}\widetilde{\Delta}_{2}}\right), \label{eq:chi1} \end{equation} where $\Omega_{2}$ is the Rabi frequency for beam 2. The linear polarizability for beam 4 takes on analogous form. For beam 2 we have the following linear polarizability: \begin{equation} \chi_{2}=-\frac{Nd_{23}^{2}\Omega_{1}^{2}}{4\epsilon_{0}\hbar\widetilde{\Delta}_{1}\widetilde{\Delta}_{1}^{*}\widetilde{\Delta}_{2}},\label{eq:chi2} \end{equation} and the polarizability for beam 3 will take on an analogous form as well. To complete our considerations we take the Doppler broadening into account. All, linear and nonlinear, polarizabilities should be averaged over the Maxwell velocity distribution \cite{Parniak2015,Mirza2015}: \begin{equation} \chi_{j}(\omega_{1,}...,\omega_{4})\to\int\limits _{-\infty}^{\infty}\chi_{j}(\omega_{1}-k_{1}v,...,\omega_{4}-k_{4}v)g(v)\mathrm{d}v, \end{equation} \begin{equation} g(v)=\sqrt{\frac{m}{2\pi k_{b}T}}\exp{\left(-\frac{mv^{2}}{2k_{\mathrm{B}}T}\right)}, \end{equation} where $m$ is the atomic mass, $k_{\mathrm{B}}$ is Boltzmann's constant and $T$ is temperature. Dominant contribution of the Doppler broadening is present in single-photon terms, which we take into account as full Voigt profiles in the above calculation. However, there is also residual Doppler broadening in the two-photon terms with $\Delta_{2}$. Since calculation of this term is time-consuming while the broadening is small, we decided to replace it by additional natural broadening, which is of quite different shape, but gives similar and consistent results. Originally our model is created for four-wave mixing in double-ladder configuration, but it is worth to mention that it can be very easily adapted to another schemes like diamond or double-$\Lambda$ configuration and other multi-wave mixing schemes. We only need to change the polarizabilities, which in any case have analogous forms. The model is also applicable to the cylindrically-symmetric case, where all beams are co-propagating. Instead of radial shift of emission direction one could observe emission to the cone where the apex angle $\theta_3$ depends on detunings. This regime can be reached if we use beams with sufficiently small diameters $\sigma<1/\theta_3$. In following sections we describe measurements of four-wave mixing signal characteristics (intensity and direction of emission) as a function of detunings and direction of all incoming beams and we compare it with theoretical prediction of model presented above. \section{Experimental} Essential components and ideas of the experiment are presented in Fig. \ref{fig:schemat}. We use a double-ladder level configuration in $^{87}$Rb [Fig. \ref{fig:schemat}(a)]. The four-wave mixing signal is generated on the transition between $5^{2}\mathrm{D}_{5/2}$ and $5^{2}\mathrm{P}_{3/2}$ manifolds. To drive the process we use three lasers: one external-cavity diode laser (ECDL, number 2) with wavelength of 776 nm and linewidth approx. 100 kHz and two distributed feedback (DFB) laser diodes with wavelengths of 780 nm (number 1 and 4) and linewidth of about 1 MHz. To minimize the two-photon Doppler broadening laser beams 1 and 2 are arranged in the counter-propagating configuration. The 776 nm laser is stabilized using a commercial wavemeter (HighFinesse WS7). One of the 780 nm lasers is locked at the vicinity of two-photon absorption peak using an auxiliary rubidium vapor cell placed in a tunable magnetic field where we crossed strong circularly polarized 776 nm laser beam and weak linearly polarized laser 1 beam at 780 nm. By measuring polarization rotation in circular basis we could generate tunable locking signal, that controls laser 1 and lock the two lasers at $\Delta_2\neq0$. More details of this method are described in \cite{Parniak2016}. Using the fact that the difference between frequencies of lasers 1 and 4 is of the order of several GHz, laser 4 is stabilized by beat-note measurement. \begin{figure}[b] \centering{}\includegraphics[width=0.75\textwidth]{ruch_mieszania_cale}\caption{Dependence of the four-wave mixing signal intensity and emission angle on detuning $\Delta_{4}$ for $\Delta_{1}/2\pi=-3500$ MHz and $\Delta_{2}=0$ at $T=145\mathrm{^{\circ}C}$. Dots correspond to the experimental result, while solid lines correspond to theoretical prediction. \label{fig:przesuwanie}} \end{figure} All laser beams intersect in the cell with warm rubidium vapors at natural abundance and no buffer gas. Fig. \ref{fig:schemat}(b) depicts the central part of the setup and a telescope used to obtain beams intersecting at approx. $11$ mrad angle and $1/e^{2}$ diameter of 400 $\mu$m. To eliminate influence of external magnetic fields the cell is placed inside a double $\mu$-metal shielding. Additionally, the cell is heated using a bifilarly-wound coil to avoid stray magnetic field from the heater. The generated four-wave mixing signal is separated from stray driving light using a half-wave plate with a polarizing beamsplitter and then using a band-pass interference filter tilted to transmit light at 776 nm. The signal is registered by an avalanche photodiode (APD) or a CCD camera situated in the far field of the rubidium cell, allowing us to measure angular distribution of the emission. A flip-mirror is used to select either the APD or the CCD camera (Fig. \ref{fig:schemat} (c)). \section{Results} \begin{figure}[b] \centering{}\includegraphics[width=0.75\textwidth]{przesuniecie_przyklad_cale2}\caption{Transmission profile of laser 1 light (blue curve and points) and the intensity of four-wave mixing signal (red curve and points) as a function of two-photon detuning $\Delta_2$ in the vicinity of the two-photon resonance. Dots correspond to experimental result for $\Delta_{1}/2\pi=-3000$ MHz and $\Delta_{4}/2\pi=-2760$ MHz, while solid lines correspond to the theoretical prediction. Vertical dotted lines mark maxima of the four-wave mixing signal intensity and two-photon absorption. The frequency shift between the two maxima is marked as $\Delta_\mathrm{S}$, while the maximum intensity of the four-wave mixing signal attained at the resonance is marked as $I_{\mathrm{{max}}}$ .\label{fig:przes}} \end{figure} \begin{figure}[b] \begin{centering} \includegraphics[width=0.75\textwidth]{4wm_symulacja_cale3} \par\end{centering} \centering{}\caption{Maximal four-wave mixing signal intensities $I_{\mathrm{max}}$ (blue curves, left axis) and corresponding frequency shifts $\Delta_\mathrm{S}$ (red curves, right axis) for a set of $\Delta_{1}$ and $\Delta_{4}$ detuning. Panels (a) and (b) correspond to experimental and theoretical results, respectively. Subsequent curves were measured/calculated for different $\Delta_{1}$ detunings in 500 $2\pi\times$MHz steps starting from $\Delta_{1}/2\pi=-5000$ MHz on the left.\label{fig:przes_cale}} \end{figure} \begin{figure} \begin{centering} \includegraphics[width=0.75\textwidth]{mieszanie_kat_cale_5} \par\end{centering} \centering{}\caption{(a) Tilts of crossing beams to the $z$-axis. Dots and crosses inside circles correspond to direction to the positive and negative values on the $z$-axis. (b) Dependence of the four-wave mixing frequency shift $\Delta_\mathrm{S}$ in relation to the two-photon absorption resonance and (c) maximum intensity $I_\mathrm{max}$ on $\Delta_{4}$ detuning. Subsequent experimental points correspond to measurements at various detuning $\Delta_{4}$ and angles as shown in panel (a) with constant $\Delta_{1}/2\pi=-2000$ MHz. Dots correspond to experimental result, while solid lines correspond to theoretical prediction. \label{fig:kat}} \end{figure} In the following section we compare our experimental and theoretical results to demonstrate the importance of the phase-matching condition and finally engineer the optimal geometry for the process. First, we observe the dependence of the average emission angle $\alpha_{3}$ of the four-wave mixing signal on detuning $\Delta_{4}$, while remaining detunings are kept constant. In particular, driving beam 1 is far detuned from the resonance ($\Delta_{1}/2\pi=-3500$ MHz), while being simultaneously kept in the two-photon resonance, i.e. $\Delta_{2}=0$, where the four-wave mixing signal is close to the optimum. From the theoretical point of view, the phase-matching ring, as described in Sec. \ref{sec:TEORYJA}, has finite width and overlaps with the Gaussian factor from Eq.~(\ref{eq:amplituda}) at either side. For that reason, the emission angle $\alpha_{3}$ changes as a function of detuning (in this case $\Delta_{4}$). In the experiment, the full signal beam profile at the far field is measured with the CCD camera and the average emission angle is inferred from the image after electronic background subtraction. In the theoretical framework this corresponds to the angle calculated as: \begin{equation} \alpha_{3}=\frac{\int\mathrm{d\boldsymbol\varphi_{3}|\boldsymbol\varphi_{3}|}|\tilde{A}_{3}(\boldsymbol\varphi_{3},L)|^{2}}{\int\mathrm{d\boldsymbol\varphi_{3}}|\tilde{A}_{3}(\boldsymbol\varphi_{3},L)|^{2}}.\label{eq:aldadef} \end{equation} Measurement results and theoretical predictions results are presented in Fig. \ref{fig:przesuwanie}. We observe a strong correspondence between the two, which confirms that the reason of dependence between emission direction and lasers detunings is indeed dispersion and resulting change in phase matching. As long as the change of emission angle is perhaps the most direct consequence of dispersion and phase matching, another significant effect is the shift in optimum two-photon detuning $\Delta_{2}$ for the four-wave mixing signal generation in relation to two-photon absorption maximum. This effect, sometimes also visible as a strong modification of spectrum, has been observed in several previous works \cite{Zibrov2002,Brekke2015,Sell2014,Parniak2016a}, but its origin has not been studied thoroughly. Simple theory, which neglects dispersion of refractive indices, predicts that maximum intensity of measured signal is exactly at the two-photon absorption maximum, as the third order polarizability $\chi_{3}^{(3)}$ and the two-photon absorption coefficient proportional to $\chi_{2}$ depends on the two-photon detuning $\Delta_{2}$ in the very same way {[}see Eqs. (\ref{eq:chi3}) and (\ref{eq:chi2}){]}. In all cases the magnitude of the effect is limited by the two-photon linewidth. In our model not only large polarizability $\chi_{3}^{(3)}$ is required, but also overlap between the phase-matching ring and Gaussian factor from Eq. (\ref{eq:amplituda}). The latter condition is not necessarily met at $\Delta_{2}=0$. From this reasoning, the frequency shift $\Delta_\mathrm{S}$ of the maximum follows. Similar situation in double-$\Lambda$ configuration was described in \cite{PhysRevA.88.033845} where this effect was also presented as a consequence of dispersion. An exemplary plot of the two-photon absorption and the four-wave mixing signals as a function of the two-photon detuning are presented in Fig. \ref{fig:przes}. Note a significant shift $\Delta_\mathrm{S}$ between the extrema of the two signals, marked with dashed lines. Imperfect fit to theory is due to residual two-photon Doppler broadening, accounted for in numerical calculations as a broadened Lorentzian instead of a full Voigt profile. The shift was measured for multiple values of detunings $\Delta_{1}$ and $\Delta_{4}$. Experimental results, together with theoretical prediction, are presented in Fig. \ref{fig:przes_cale}. For different detunings the system exhibits different dispersion profiles, and thus we observe different frequency shifts $\Delta_{S}$. They confirm that dispersion and phase matching accurately explain dependence between the four-wave mixing signal intensity and laser detunings. According to Eqs. \eqref{eq:chi3}--\eqref{eq:chi2} all polarizabilities depend linearly on the atomic concentration. In consequence not only the four-wave mixing becomes more efficient at larger atomic concentrations, but also the dispersive effects become significantly more important. As the atomic concentration strongly depends on cell temperature, the size of the measured effect will strongly depend on temperature as well, as in e.g \cite{Brekke2015}. In our experiments we set the temperature to 145~$^\circ$C, which corresponds to the atomic concentration of the order of $10^{19}$ $\mathrm{m}^{-1}$. This setting allows us to attain large signal intensities together with clearly visible dispersive effects. Finally, we confirm that as originally intended the geometry of the experiment may be changed to tailor the four-wave mixing signal properties. From theory it follows that the overlap between the phase-matching ring and the Gaussian factor decreases with any change of incident beams tilts $\boldsymbol\varphi_{1}$, $\boldsymbol\varphi_{2}$ and $\boldsymbol\varphi_{4}$ {[}see Eq. (\ref{eq:dthdfi}){]}. However, this overlap can be fixed by adjusting detunings. As an example, we study how the maximum of the four-wave mixing signal changes as a function of $\Delta_{4}$ detuning if we tilt the incident beam 3. We performed the measurement for three different tilts presented in Fig. \ref{fig:kat}(a). Figures \ref{fig:kat}(b) and \ref{fig:kat}(c) portray obtained experimental and theoretical results for the frequency shift $\Delta_\mathrm{S}$ and maximum intensity $I_\mathrm{max}$, respectively. We observe that the maximum of the four-wave mixing signal changes quite significantly. Note, that without taking into account the phase-matching condition we only obtain $1/\Delta_{4}$ dependance for the field amplitude [as in Eq.~\ref{eq:chi3}], which would clearly give a completely incorrect theoretical prediction in this case. Instead, the proper choice of geometry is critical when laser 4 is tuned closer or further from its respective single-photon resonance. These results demonstrate that our theoretical model may be used to precisely predict the behavior of the four-wave mixing signal, explaining a variety of intricacies. In particular, we may predict the optimum geometry for a desired set of detunings, or vice-versa. \section{Conclusions} We have shown that a simple model of four-wave mixing in an atomic medium that neglects propagation effects could be easily extended by considering propagation equation with nonlinear polarization to account for phase matching. In the spatial Fourier domain, the phase-matching condition implies that only some component wavevectors of the drive beams lead to effective wave-mixing. Moreover, due to dispersion, the phase-matching conditions strongly depend on laser detunings. This approach allowed us to explain some phenomena which were unexplainable earlier, like the frequency shift of the four-wave mixing signal maximum in relation to two-photon absorption maximum \cite{Zibrov2002,Brekke2015,Sell2014,Parniak2016a}. The most direct consequence of dispersion seems to be the change of the four-wave mixing signal emission angle as a function of laser detunings. These results show that our theoretical description is much more robust than an atomic model neglecting the influence of dispersion on phase matching in atomic medium and allows precise predictions of the four-wave mixing signal behavior. Moreover, we have shown that our model facilitates proper choice of geometry and consequently stronger signal may be obtained. Apart from providing a recipe for engineering of effective interaction, we believe that our results may also help study other subtle effects in four-wave mixing \cite{DeMelo2015}, where multiple shift effects might contribute to the total frequency shift. Finally, we note that our approach is generally valid for a variety of systems - one needs to simply supply appropriate expressions for linear and nonlinear susceptibilities. In particular, effects treated in this work become even more critical in high-density ensembles, that facilitate strong light-matter interactions due to collective effects \cite{1367-2630-15-8-085027,Srivathsan2013a}, in waveguides that support modes with various longitudinal wavevectors \cite{Donvalkar2014}, in cavity-enhanced processes \cite{Offer:16} or finally in schemes that rely on phase-matching control, such as multimode quantum memories \cite{Mazelanik:16}. Our model is also applicable to the case of spontaneous four-wave mixing \cite{Srivathsan2013a}: if we post-select on one of the fiber-coupled photons, we may predict the correlated emission direction. With further extensions our theory may also be used to model the full biphoton joint spectral/spatial properties. If the four-wave mixing signal is weak as in quantum light-atom interfaces \cite{Parniak2016a,Radnaev2010a} or processes involving Rydberg states \cite{Brekke2008,DeMelo2014a,Magno2001} proper choice of geometry may be even more critical and yield a difference between having a strong signal or not obtaining the four-wave mixing signal at all. %% FUNDING \section*{Funding} The project was financed by the Polish Ministry of Science and Higher Education ``Diamentowy Grant'' Project No. DI2013 011943 and by the National Science Centre (Poland) Grants No. 2015/16/S/ST2/00424 and 2015/17/D/ST2/03471. \section*{Acknowledgments} We acknowledge generous support of T.~Stacewicz, K.~Banaszek and R.~\L{}apkiewicz as well as careful proofreading of the manuscript by M. D\k{a}browski. \end{document} }\end{equation}}
\caption{Details of the CBT device structure and package. The large arrow shows the direction of applied magnetic field $B$. Panel (\textbf{a}) is a photograph of the CBT chip, showing the array of $20 \times 32$ copper islands in the bottom $3/4$ of the image and the on-chip filters at the top. Panel (\textbf{b}) is a micrograph showing five of the islands. The package is located in the bore of a superconducting solenoid and the magnetic field is aligned parallel to the long axis of the islands. Panel (\textbf{c}) shows the design of the sample package used to allow effective precooling of the CBT device. It consists of a closed silver box (lid not shown here) in which the CBT is bonded in place using a silver conductive coating (Electrodag\textregistered). The package also contains a small PCB (green) with RLC low-pass filters to which the CBT bond wires are attached. For thermalisation, the package is mounted on a thin copper strip and attached to a silver wire, both of which are attached to the mixing chamber plate of the dilution refrigerator. Four copper wires pass from the package to additional filters at the mixing chamber plate.}
\caption{Additional \aastex\symbols}
\caption{Variance of the protocol in Fig.~\ref{fig:Concatenate1} plotted against the effective transmission of the channel. This protocol takes a channel of direct transmission \(\eta\) to one of effective transmission \(\sqrt{\eta}\) using \(\chi = 0.7\) (blue) and \(\chi=0.1\) (green). Also shown is the entanglement breaking bound (red). The point \textcolor[rgb]{0,0,0}{\ding{108}} is achieved with a success probability of \(P=0.06\) and the point \textcolor[rgb]{0,0,0}{\ding{110}} is achieved with \(P= 0.001\). \label{fig:VarOne}}
\caption{Variance of the protocol for two links of the repeater shown in Fig.~\ref{fig:Concatenate2}(purple) plotted against the effective transmission of the channel. This protocol takes a channel of direct transmission \(\eta^2\) to one of effective transmission \(\sqrt{\eta}\). The nested error correction protocols use \(\chi=0.01\) and the larger protocol used \(\chi=0.7\). Also shown is the entanglement breaking bound (red). The point \textcolor[rgb]{1,0.5,0}{\ding{108}} is achieved with success probability \(4\times10^{-8}\). \label{fig:VarConc} }
\caption{Additional \aastex\symbols}
\caption{% A demonstration of the strengths and weaknesses of the 2-projection (vertical and horizontal) multiplicative algebraic reconstruction technique (ART). The plots in the upper row are based on the function $f(x,y) = e^{-(2x^2+10y^2)}$ with the bottom row based on the function $g(x,y) = 1.0 - f(x,y)$. As is evidenced by \protect\subref{fig:rorg} and \colorme{\protect\subref{fig:riorg}}, the 2-projection method cannot capture the superimposed rotations present in \protect\subref{fig:org} and \protect\subref{fig:iorg}, respectively, as the principal axes of the function and the projection directions do not coincide. Note: the reconstruction in \protect\subref{fig:rog} is an exact reconstruction of the image in \protect\subref{fig:og}.% }
\caption{A virtual experiment where a representative density field is extracted from a ceramic powder compaction simulation and reconstructed from two orthogonal projections using multiplicative ART. The \colorme{two-projection} ART reconstruction method adequately reconstructs the \colorme{general trends of the simulated density} field, although it does smooth along the projection directions. Because the method is insensitive to the magnitude of the density, the density field of the simulated green body is normalized according to the average density of the slice.}
\caption{Comparative benchmark results for the \texttt{particle} and \texttt{saddle} examples \citep{tr-ece-08-01}, the \texttt{probabilistic-lambda-calculus} and \texttt{probabilistic-prolog} examples \citep{siskind2008} and an implementation of backpropagation in neural networks using AD.\@ Column labels are for AD modes and nesting: F for forward, Fv for forward-vector aka stacked tangents, RF for reverse-over-forward, etc.% All run times normalized relative to a unit run time for \Stalingrad\on the corresponding example except that run times for\texttt{backprop-Fv} are normalized relative to a unit run time for \Stalingrad\on\texttt{backprop-F}. % Pre-existing AD tools are named in {\blue blue}, others are custom implementations. % Key: {\green\protect\rule{1ex}{1ex}}~not implemented but could implement, including \Fortran, \Clang, and \Cplusplus; {\blue\protect\rule{1ex}{1ex}}~not implemented in pre-existing AD tool; {\red\protect\rule{1ex}{1ex}}~problematic to implement. % All code available at \url{http://www.bcl.hamilton.ie/~qobi/ad2016-benchmarks/}. }
\caption{\small Example frames of our two applications. (a) for generating dancing and karaoke, (b) for neural story-singing. Please visit {\color{blue}some link} to watch the full videos.}
\caption{Experiments on aircraft data set. Note that each row contains the results for a different test image. In the PR plots, \lq\textcolor{blue}{$\times$}\rq and \lq\textcolor{red}{$\times$}\rq mark the samples produced by our approach where \lq\textcolor{red}{$\times$}\rq indicates the sample with the best F-measure value, and \lq$\times$\rq marks that of segmentation of Kim et al.~\cite{kim2007nonparametric}.\label{fig:experiments_aircraft}}
\caption{Experiments on walking silhouettes data set. In the PR curves, the \lq\textcolor{red}{$\times$}\rq marks the sample having the best F-measure value obtained using the proposed approach (with either global or local shape priors), and the \lq$\times$\rq marks that of segmentation of Kim et al.~\cite{kim2007nonparametric}.\label{fig:experiments_walking}}
\caption{ Illustration of (a) a subdivided pumpkin gadget and of (b) a subdivided slice gadget encoding integer $7$. (c) A wedge $W_j$ of width~$24$, its transversal path $\pi_j$, and subdivided slices $S_{j1}$, $S_{j2}$, and $S_{j3}$ encoding integers~$7$,~$8$ and $9$, respectively. Shared edges are colored black, those of the subdivided pumpkin with thick lines, while private edges of $G_1$ and of $G_2$ are colored blue and red, respectively. (d) A (vertically stretched) \gracsimdraw{} of the transformed instance of \threep{}, when $m = 3$, $B = 24$ and $A = \{7,7,7,8,8,8,8,9,10\}$. Subdivided slices are drawn within wedges according to the following solution of \threep{}: $A_1 = \{7,7,10\}$, $A_2 = \{7,8,9\}$ and $A_3 = \{8,8,8\}$. }
\caption{The temperature dependence of the resistivity along the $c$-axis ($\rho_{c}$) at various pressures. The upper left inset shows the pressure dependence of the resistivity ratio between 300 K and 2 K [$\rho_{c}$(300 K) / $\rho_{c}$(2 K)]. {\color{red}{ The upper right inset shows the pressure dependence of the resistivity along the $c$- ($\rho_{c}$) and $a$- ($\rho_{a}$) axes at 2 K. }} The lower image shows the crystal structure of BP and the corresponding crystal axes. \label{fig1}}
\caption{ {\color{red}{(a)The pressure dependence of $-d^2\rho_{xx}/dB^2$ at 2 K in magnetic fields along the $b$-axis. The data were vertically offset for clarity. The inset shows the temperature dependence of the FFT spectra at 1.43 GPa. (b)The pressure dependence of the FFT spectra of $-d^2\rho_{xx}/dB^2$ from 1.43 to 2.24 GPa at 2 K. The major and minor peaks are indicated by the red and blue arrows, respectively. Several additional peaks which appear at 1.43 and 1.54 GPa are indicated by the green arrows. Each magnitude of the spectrum is normalized by the amplitude of each major peak and vertically offset for clarity. The pressure dependence of the peak frequencies in magnetic fields along the (c)$a$-, (d)$b$-, and (e)$c$-axes. In all of the field directions, the major peak with a large FFT amplitude (red), and the minor peak with smaller one (blue), were observed.}} \label{fig4}}
\caption{Example of the solar analogue KIC\,61160048. The top panel shows the Power Spectral Density (PSD) in grey while a smoothed PSD is shown in black with logarithmic scaling in both axis. The centre of the excess of oscillation power is marked with a vertical red line at$\nu_{\rm max}$$\simeq$2050\,$\mu$Hz. The middle panel depicts a zoom into the frequency range of the power excess, revealing the comb-like pattern of individual oscillation modes. The large-frequency separation, \dnu, between consecutive radial modes is indicated through the horizontal red line. The bottom panel shows the \'echelle diagram that is produced by folding the PSD of the power excess with the large-frequency separation. Each vertical bright ridge corresponds to a set of oscillation modes with a given$\ell$-value. %Frequency spectrum of KIC\,6116048. In the multicomponent fit \red{[TBD]}, the granulation and convective background components are described as three power laws, while the constant background describes the frequency independent photon noise. The oscillation modes are shown in the excess of oscillation power, described through the Gaussian component. \label{fig:PSD}}
\caption{Calculation of the \Sindex in KIC\,3241581. The blue triangle indicate the triangular filter which is used to weight the flux at the core of the Ca\,H\&K line, left and right, respectively. It should be noted that the H line is blended with the H$_\gamma$ line. The grey shaded regions are marking the blue and red normalisation passbands. The \Sindex is the ratio of the sum of flux from the cores of the lines, divided by the sum of the flux in the comparison passbands, times the multiplicative factor to scale the instrumental value onto the instrumental reference frame of the Mount Wilson Observatory. }
\caption{Comparison of the photospheric \Sph-index (x axis) with the chromospheric \Sindex, calibrated into the MWO reference (y axis) for the sample of the 18 solar-analogue stars. For both \Sph and \Ssymbol, the maximum and minimum of the solar activity levels during the solar cycle are indicated by the red and blue dashed lines respectively. The size of the symbols is inversely proportional to the rotation period \citep[adopted from Figure\,3 from][]{Salabert2016Activity}.}
\caption{ \label{fig:fit} The Bose-ghost propagator $Q(p^2)$ as a function of the (improved) lattice momentum squared $p^2$. Here we used as sources $B^{ec}_{\mu}(z)$ the formula reported in Eq.\(32) of Ref.\\cite{Cucchieri:2016czg}. We plot data for $\beta_2 \approx 2.44$, $V = 96^4$ (\protect\marksymbol{*}{red}) and $\beta_3 \approx 2.51$, $V = 120^4$ (\protect\marksymbol{diamond*}{green}), after applying a matching procedure \protect\cite{rescaling} to the former set of data. We also plot, for $V = 120^4$, a fit using the fitting function (\protect\ref{eq:fit}). Note the logarithmic scale on both axes. }
\caption{ \label{fig:props-beta0} The Bose-ghost propagator $Q(p^2)$ (\protect\marksymbol{*}{red}) and the product $g_0^2 \, G^2(p^2) \, D(p^2)$ (\protect\marksymbol{diamond*}{green}) as a function of the (improved) lattice momentum squared $p^2$ for the lattice volume $V = 64^4$ at $\beta_0$. Here we used as sources $B^{ec}_{\mu}(z)$ the formula reported in Eq.\(32) of Ref.\\cite{Cucchieri:2016czg}. Note the logarithmic scale on both axes. }
\caption{ \label{fig:props} The Bose-ghost propagator $Q(p^2)$ (\protect\marksymbol{*}{red}) and the product $g_0^2 \, G^2(p^2) \, D(p^2)$ (\protect\marksymbol{diamond*}{green}) as a function of the (improved) lattice momentum squared $p^2$ for the lattice volume $V = 120^4$ at $\beta_3 \approx 2.51$. Here we used as sources $B^{ec}_{\mu}(z)$ the formula reported in Eq.\(32) of Ref.\\cite{Cucchieri:2016czg}. Note the logarithmic scale on both axes. }
\caption{Location Word Type 2 (\textcolor{red}{Incorrect}) Vs. Type 1 (\textcolor{blue}{Correct}); ICEWS story ID: \texttt{17929606}, DRC data; and OEDA story ID: \texttt{2054559 v0.1.0}, Syria data }
\caption{Location Word Type 3 (\textcolor{red}{Incorrect}) Vs. Type 1 (\textcolor{blue}{Correct}); ICEWS story ID: \texttt{18141520}, China data; ICEWS story ID: \texttt{10672170}, DRC data.}
\caption{Location Word Type 4 (\textcolor{red}{Incorrect}) Vs. Type 1 (\textcolor{blue}{Correct}); OEDA story ID: \texttt{1160025 v0.2.0}, Syria data; ICEWS story ID: \texttt{18922379}, DRC data.}
\caption{Examples of pivot and target translations using the pivot-based translation strategy. We observe that our approaches generate better translations for both pivot and target sentences. We italicize \protect \textcolor{blue}{\textit{correct translation segments}} which are no short than 2-grams.}
\caption{ Propagation of uncertainties in the zero-point calibration of the TRGB. Note that the previous calibration based on $\omega$ Centauri gives a zero-point uncertainty of 0.124 mag \citep{bel01, bel04}. When NGC 4258 and LMC are used as a distance anchor, uncertainties are estimated to be 0.073 and 0.096 mag, respectively. Combining these two anchors (NGC 4258 and LMC) provides the most accurate calibration with an uncertainty of 0.058 mag (red starlet), 53\% smaller than the value given by \citet{bel01, bel04}. Individual error values are listed in {\color{red} \bf Table \ref{tab_error}}. }
\caption{Deluge Network components: (a) a composite layer, (b) a block transition component, and (c) a block. \textcolor{red}{Red}-colored arrows indicate $1\times1$ cross-layer depthwise convolutions.}
\caption{ Average radial profiles of merger models at the final times. The data for He0609, CO0812, and CO0909 are shown in the left, middle, and right columns of panels, respectively. The average radial profiles of hydrodynamic state variables are shown on the top row. The plotted quantities are density, $\rho \left(\si{\g\per\cubic\cm}\right)$, temperature, $T \left(\si{\K}\right)$, velocity magnitude, $\lvert v \rvert \left(\si{\cm\per\s}\right)$, and the $z$ component of the specific angular momentum, $h_z \left(\si{\cm\squared\per\s}\right)$. The velocity and specific angular momentum are measured in the corotating frame of reference. The average radial compositional profiles for major isotopes are shown in the bottom row. All profiles were calculated using the system's center of mass as the origin, and were suitably averaged over spherical shells at given radii. Note that the scaling of variables and size of the region shown change between panels. See text for discussion. \label{f:finalCond} }
\caption{The top 5 re-id matches on the CUHK03 test data for a test of 4 queries from the 100 image gallery setting. This shows the local mask based model. Note that the person image in the {\color{green} green} box has the same person id as the query image. Identities of similar appearances are near in re-id search space.}
\caption{In this plot we see p-values for the CSD-DTI test. {\color{blue} Blue }denotes CSD $>$ DTI by as significant margin, while {\color{red} Red} denotes DTI $>$ CSD by a significant margin. Gray areas denote regions that were unable to reject the null hypothesis of ``no difference''. From \textbf{top} to \textbf{bottom} we have $\sigma =\{$0.01,0.005, 0.001, 0.0005, 0.0001$\}$, and from \textbf{left} to \textbf{right} we have the left hemisphere and right hemisphere.}
\caption{Qualitative results of the CNN based trackers (~\textcolor{red}{red: ours};~\textcolor{yellow}{yellow: HCF}~\cite{Ma2015ICCV};~\textcolor{blue}{blue: DLT}~\cite{wang2013learning}.) on sequences: (a)~\textit{freeman4}; (b)~\textit{doll}; (c)~\textit{skiing}; (d)~\textit{singer2}; (e)~\textit{lemming}. }
\caption{Molecular model has 5 levels per layer, all other parameters detailed in \protect\gls{SI}.\cite{Supp:ref} Panel a): Adiabatic and diabatic potentials (same as in Fig.~\ref{fig:pot_surf}). Horizontal lines labelled as \protect\numcircledtikz{\footnotesize 1} and \protect\numcircledtikz{\footnotesize 2} show two different initial energies. Trajectories for the setups \protect\numcircledtikz{\footnotesize 1} and \protect\numcircledtikz{\footnotesize 2} are on panels b) and c), respectively. }
\caption{T1 simulations. Variations with the axes ratio $\theta$ in pressure at $z=0$ plane (in N/m$^2$) exerted by bottom half of ring on its top half for (a) non-gravitating and (b) self-gravitating simulations. Symbols correspond to the different components of the total pressure at $z=0$ plane: `{\large\color{blue}$\cdot$-{$\circ$}-$\cdot$}' for collisional component ($p_{z,coll}$), `{\color{red}$\cdot\cdot\square\cdot\cdot$}' for streaming component ($p_{z,str}$) and `-$\times$-' for the total pressure ($p_{z}$).}
\caption{T2 simulations. Variations with the axes ratio $\theta$ in pressure at $x=0$ plane (in N/m$^2$) exerted by the radially interior half of ring on its exterior half for (a) non-gravitating and (b) self-gravitating simulations. Symbols correspond to the different components of the total pressure at $x=0$ plane: `{\large\color{blue}$\cdot$-{$\circ$}-$\cdot$}' for collisional component ($p_{x,coll}$), `{\color{red}$\cdot\cdot\square\cdot\cdot$}' for streaming component ($p_{x,str}$) and `-$\times$-' for the total pressure ($p_{x}$).}
\caption{Variation in Toomre parameter ($Q$) with the axes ratio $\theta$. Figure (a) corresponds to simulations with $\tau_0=0.05\;(1432 \textrm{ particles})$, while (b) $\tau_0=0.5\;(14328 \textrm{ particles});$ all other parameters are same as in \Cref{table:sim_char}. T1 ({\large \color{red}-$\circ$-}) and T2 ({\color{blue}$\cdot\cdot \square\cdot \cdot$}) simulations are shown.}
\caption{Self-gravitating T2 simulations: Variations with characteristic frequencies in (a) granular temperature ($T_g$, in {m$^2$/s$^{2}$}), (b) impact frequency ($\omega_c$, in s$^{-1}$), (c) effective vertical thickness ($t_v$, in particle radii), (d) effective radial width ($w_r$, in particle radii). Symbols correspond to the different characteristic frequencies: `{\color{red}- -$\ast$- -}' for $\kappa^*$, `{\color{green}$\cdot\cdot\circ\cdot\cdot$}' for $\Omega^*$ and `{\color{blue}$\cdot$ -$\diamond$- $\cdot$}' for $\nu^*$.}
\caption{Self-gravitating T2 simulations: Variations with characteristic frequencies in the pressure (in N/m$^2$) at $z=0$ ($p_{z}$) and $x=0$ plane ($p_{x}$), and their streaming ($p_{z,str}$, $p_{x,str}$) and collisional ($p_{z,coll}$, $p_{z,coll}$) components. Left and right columns report results for the $z=0$ and $x=0$ plane, respectively. Symbols correspond to different characteristic frequencies: `{\color{red}- -$\ast$- -}' for $\kappa^*$, `{\color{green}$\cdot\cdot\circ\cdot\cdot$}' for $\Omega^*$ and `{\color{blue}$\cdot$ -$\diamond$- $\cdot$}' for $\nu^*$.}
\caption{Self-gravitating T2 simulations: Variations with characteristic frequencies in (a) number density ($n_s$, in m$^{-3}$) and (b) Toomre parameter ($Q$). Symbols correspond to the different characteristic frequencies: `{\color{red}- -$\ast$- -}' for $\kappa^*$, `{\color{green}$\cdot\cdot\circ\cdot\cdot$}' for $\Omega^*$ and `{\color{blue}$\cdot$ -$\diamond$- $\cdot$}' for $\nu^*$.}
\caption{Self-gravitating T2 simulations: Variations with characteristic frequencies in (a) radial tidal force ($\xi$, in s$^{-2}$) and (b) effective radial spreading rate ($s_{w_r}$, in particle radii/s). Symbols correspond to the different characteristic frequencies: `{\color{red}- -$\ast$- -}' for $\kappa^*$, `{\color{green}$\cdot\cdot\circ\cdot\cdot$}' for $\Omega^*$ and `{\color{blue}$\cdot$ -$\diamond$- $\cdot$}' for $\nu^*$.}
\caption{The average scores and ranks of accuracy and robustness of different methods on VOT2015 \cite{kristan2015visual}. The top three scores are highlighted in \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green}, respectively. }
\caption{Experimental stability curve of a 70 \textmu L ethanol drop on a PTFE substrate. Filled green circles (\textcolor{green}{$\bullet$}): breakdown of the axisymmetry. Blue squares: transition to the worm-like instability, obtained by keeping the frequency fixed and increasing (filled blue squares (\textcolor{blue}{$\blacksquare$})) or decreasing (open blue squares (\textcolor{blue}{$\square$})) the forcing amplitude $A_v$. (C): circular state; (E): two-eyes state; (W): worm-like shape. Open black triangles ($\bigtriangleup$): length of the worm measured at a fixed value of $A_v= \unit{105}{\micro\meter}$ for different frequencies (following the horizontal dashed line). Vertical dotted line corresponds to the experiments performed at constant frequency (see Fig.\,\ref{figure3}). The scale bars are \unit{5}{\milli\meter} long.}
\caption{Results of experiments at constant frequency ($f=\unit{130}{\hertz}$) following the vertical dotted line of Fig.\,\ref{figure2}. Top: filled blue circles (\textcolor{blue}{$\bullet$}) show worm length {\it vs} amplitude of vibration $A_v$. The blue dotted line corresponds to the circular drop (C) diameter. Bottom: open red squares (\textcolor{red}{$\square$}) show amplitude of the Faraday waves $A_F$ {\it vs.} amplitude of vibration $A_v$. The dashed lines correspond to the thresholds between circular (C), two-eyes (E) and worm-like shapes. Pictures correspond to (a) top-view in (W) state, (b) side view in (C) state and (c) side-view in (W) state. The scale bars are \unit{1}{\milli\meter} long.}
\caption{Amplitude of the Faraday waves as a function of the driving amplitude for a fixed frequency of \unit{130}{\hertz}. Open red square (\textcolor{red}{$\square$}): amplitude of the Faraday waves directly measured using a side-view camera. Deduced values of $A_F$ from the measurements of the length of the worm using the usual dispersion relation (blue star (\textcolor{blue}{$\star$}), dashed line) and the renormalized one, eq. \ref{masterequation}, with respectively $\Psi=\pi$ (open blue circles (\textcolor{blue}{$\circ$}), dotted line) or $\Psi=0$ (filled blue circles (\textcolor{blue}{$\bullet$}), solid line). Lines are guides for the eye. }
\caption{Surface tension of the Mie fluid as a function of the temperature. The symbols are simulation results from Galliero et al. \cite{GPMMLB09} and the lines are correlations from Eq. (\ref{eq:Mie_Gamma}): $n = 8$ ($\cdots$), $n = 10$ (\solidrule) and $n = 20$ (\dashedrule).}
\caption{Saturated densities of the Mie fluid. The symbols are simulation results from Galliero et al. \cite{GPMMLB09} and the lines are correlations from Eq. (\ref{eq:Mie_Gamma}): $n = 8$ ($\cdots$), $n = 10$ (\solidrule) and $n = 20$ (\dashedrule).}
\caption{Example images from our \textcolor{\highlightcolor}{TUM-LSI dataset}. At each capture-location, we provide a set of six high-resolution wide-angle pictures, taken in five different horizontal directions and one pointing up.}
\caption{Median localization accuracy on \textcolor{\highlightcolor}{TUM-LSI}. }
\caption{Failed \textcolor{\highlightcolor}{SfM} reconstruction\textcolor{\highlightcolor}{s} for the \textcolor{\highlightcolor}{TUM-LSI dataset, obtained with COLMAP}. The first reconstruction contains two stairwells collapsed into one due to repetitive structures. Due to a lack of sufficient matches, the method was unable to connect a sequence of images and therefore creates a second separate \textcolor{\highlightcolor}{model} of one of the hallways.}
\caption{Example\textcolor{\highlightcolor}{s from} the Old Hospital sequence. \textcolor{\highlightcolor}{The } 3D \textcolor{\highlightcolor}{scene} model is \textcolor{\highlightcolor}{projected into the image using the poses estimated by } (b) PoseNet~\cite{kendall2015posenet} and (c) our \textcolor{\highlightcolor}{method}. Active Search~\cite{Sattler2016PAMI} \textcolor{\highlightcolor}{did} not localize the image due to the occlusion \textcolor{\highlightcolor}{caused} by the tree. Note the inaccuracy of PoseNet compared to the proposed method (check the top of the building for alignment).}
\caption{\textcolor{\highlightcolor}{Our approach achieves a}ccurate outdoor image-based localization %achieved by the proposed method, even in challenging lighting conditions where other deep architectures fail.}
\caption{ {\color{blue} Breakdown of the traffic in the WeFi trace, according to the type of user originating it (vehicular or static) and the technology serving it (3G, LTE, Wi-Fi). } %blue \label{fig:mosaic} }
\caption{Recommendation system: price-of-fog {\color{blue} when caches are deployed at base stations } %blue (a); total cache size averaged over the different operators, as a function of~$p$ (b); per-operator breakdown when~$p=0$ (solid bars) and~$p=0.5$ (bars with pattern) (c). }
\caption{Location-specific content: price-of-fog {\color{blue} when caches are deployed at base stations } %blue (a); total cache size averaged over the different operators, as a function of~$q$ (b); per-operator breakdown when~$q=0$ (solid bars) and~$q=0.5$ (bars with pattern) (c). }
\caption{\textbf{Pascal VOC 2012} testing mean IOU (\%). The preceding symbol {$^{\dagger}$} indicates that the network is pre-trained with COCO \cite{lin2014microsoft} dataset. \textcolor{blue}{\textbf{Note that COCO pre-training is not used to improve the predictions of IFCN-8s.}}}
\caption{List of features from {\tt LOG} and {\tt WEB} that are used to predict video popularity ({\color{blue}all of them are novel in the context of video popularity prediction}). \label{FactorTable_LOGWEB}}{% \begin{tabular}{|l|l||p{1.5cm}|c|p{6.5cm}|p{1.1cm}|} \hline \multicolumn{2}{|c||}{\bf Feature set } & \centering{\bf Name } & {\bf Values } & \multirow{2}{*}{\bf Description (all of them are {\color{blue}new})} & \multirow{2}{*}{\bf Modes}\\ \multicolumn{2}{|c||}{\bf $\mathbf{\Psi}$} & \centering{\bf $\psi$} & {\bf $\mathbb{D}_\psi$ } & & \\ \hline \hline \multirow{7}{*}{{\tt LOG}} & & \multicolumn{4}{|l|}{ \bf \emph{ 3. Dynamic features from logs of Yandex} } \\ \cline{2-6} \cline{2-6} & \multirow{4}{*}{$\mathtt{LOG_S}$} & \multicolumn{4}{|l|}{\emph{ 3.a. Dynamic features from search logs of Yandex}} \\ \cline{3-6} & & {\tt ShowURL} & $\mathbb{Z}_+$ & { \scriptsize The number of shows of the video URLs on SERP } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt ClickURL} & $\mathbb{Z}_+$ & { \scriptsize The number of clicks on the video URLs on SERP } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt CTR} & $[0,1]$ & { \scriptsize The click--through rate of the video URLs on SERP } & {\small\tt c/d} \\ \cline{3-6} \cline{2-6} & \multirow{2}{*}{$\mathtt{LOG_B}$} & \multicolumn{4}{|l|}{\emph{ 3.b. Dynamic features from browsing logs of Yandex}} \\ \cline{3-6} & & {\tt BrowVisit} & $\mathbb{Z}_+$ & { \scriptsize The number of visits of the video URLs registered in browsing logs } & {\small\tt n/l,c/d} \\ \hline \hline \multirow{21}{*}{{\tt WEB}} & & \multicolumn{4}{|l|}{ \bf \emph{ 4. Dynamic features from the Web } } \\ \cline{2-6} \cline{2-6} & \multirow{17}{*}{$\mathtt{WEB_{ag}}$} & \multicolumn{4}{|l|}{\emph{ 4.a. Dynamic aggregated features from the Web}} \\ \cline{3-6} & & {\tt EmbCnt} & $\mathbb{Z}_+$ & { \scriptsize The number of all embeds of the video } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt EmbHCnt} & $\mathbb{Z}_+$ & { \scriptsize The number of all hosts with embeds of the video } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt MaxEPerH} & $\mathbb{Z}_+$ & { \scriptsize The maximum number of embeds of the video per host } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt AvgEPerH} & $\mathbb{R}_+$ & { \scriptsize The average number of embeds of the video per host } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt MaxEPerP} & $\mathbb{Z}_+$ & { \scriptsize The maximum number of embeds of videos per page } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt AvgEPerP} & $\mathbb{R}_+$ & { \scriptsize The average number of embeds of videos per page } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt FirstEmb} & $\mathbb{Z}_+$ & { \scriptsize The number of days passed since the first embed of the video} & {\small\tt n/l} \\ \cline{3-6} & & {\tt LastEmb} & $\mathbb{Z}_+$ & { \scriptsize The number of days passed since the last embed of the video} & {\small\tt n/l} \\ \cline{3-6} & & {\tt AvgEmb} & $\mathbb{R}_+$ & { \scriptsize The average number of days passed since any embed of the video} & {\small\tt n/l} \\ \cline{3-6} & & {\tt LinkCnt} & $\mathbb{Z}_+$ & { \scriptsize The number of all links to the video } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt LinkHCnt} & $\mathbb{Z}_+$ & { \scriptsize The number of all hosts with links to the video } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt MaxLPerH} & $\mathbb{Z}_+$ & { \scriptsize The maximum number of links to the video per host } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt AvgLPerH} & $\mathbb{R}_+$ & { \scriptsize The average number of links to the video per host } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt FirstLink} & $\mathbb{Z}_+$ & { \scriptsize The number of days passed since the day of the first link } & {\small\tt n/l} \\ \cline{3-6} & & {\tt LastLink} & $\mathbb{Z}_+$ & { \scriptsize The number of days passed since the video was linked last time} & {\small\tt n/l} \\ \cline{3-6} & & {\tt AvgLink} & $\mathbb{R}_+$ & { \scriptsize The average number of days passed since there was any link to the video } & {\small\tt n/l} \\ \cline{2-6} \cline{2-6} & \multirow{3}{*}{$\mathtt{WEB_{nag}}$} & \multicolumn{4}{|l|}{\emph{ 4.b. Dynamic non-aggregated features from the Web}} \\ \cline{3-6} & & {\tt EmbedHost} & $\mathbb{E}$ & { \scriptsize The host list with embed timestamps of the video (preprocessed into the outcomes of the LIM) } & {\small\tt n/l,c/d} \\ \cline{3-6} & & {\tt LinkHost} & $\mathbb{L}$ & { \scriptsize The host list with link timestamps of the video (preprocessed into the outcomes of the LIM) } & {\small\tt n/l,c/d} \\ \hline \end{tabular} }
\caption{Generation example in summarization. We use colors to distinguish the word source, i.e., \colorbox{yellow}{copying}, \colorbox{cyan}{alignment} or \colorbox{white}{common words}.}
\caption{DSSIM error (less is better) for different variants \emph{(rows)} when applied to different subsets of our test set \emph{(columns)}. The best alternative is shown in \textbf{bold}. Oracle analysis using ground-truth information are shown in {\color{gray}gray}. Variant images are seen in \refFig{Alternatives_small}. }
\caption{Top 3 retrieved images in SUN RGB-D. Ground truth images appear with {\color[rgb]{0,0.2,0.8}\bf blue} bars on top. {\color[rgb]{0,0.618,0} Green} bounding boxes are detection outputs matching the generated 2D layouts. {\color[rgb]{0.618,0,0} Red} boxes are missing objects (not detected) w.r.t. the expectation of generated 2D layouts.}
\caption{Matched 3D and 2D layouts based on our 2D layout matching for three ground truth images in SUN RGB-D. {\color{blue}Blue} crosses represent camera locations. {\color[rgb]{0,0.618,0} Green} bounding boxes are object detection outputs that match the 2D layouts generated from the text queries. {\color[rgb]{0.618,0,0} Red} bounding boxes represent a missing object (not detected by the object detector) within the expected region proposed by 2D layouts.}
\caption{Learned distribution of 2D spatial relations in \textbf{subject-relation-object} relationships. {\color{red} Red} bounding boxes represent the subject and blue bounding boxes represent the sampled objects in the annotations corresponding to each relation. The subject is normalized to $1\times 1$ squares (with bottom-left $(0,0)$ and top-right $(1,1)$) and all objects are rescaled with the same normalization factors in $x$-$y$ coordinates.}
\caption{Matched object layouts based on our greedy 2D layout matching for three ground truth images that are ranked top 5 among all candidate images. {\color[rgb]{0,0.618,0} Green} bounding boxes are object detection outputs that match the 2D layouts generated from the text queries. {\color[rgb]{0.618,0,0} Red} bounding boxes represent a missing object (not detected by the object detector) within the expected region proposed by 2D layouts.}
\caption{\textbf{Emergent Phenomena from Spin\textendash Orbit Coupling (SOC) at Surfaces and Interfaces. }A schematic illustration of the connection between the presence of strong SOC at material surfaces and interfaces (inner ellipse) and the resulting emergence of new interactions and electronic states (middle ellipse), such as Dzyaloshinskii\textendash Moriya interaction (DMI; see \ref{fig:Sk-Form}\textcolor{blue}{a, e} for details), Rashba interfaces (\textcolor{blue}{\ref{fig:SP-Bands}b, d}) and topological surface states (\textcolor{blue}{\ref{fig:SP-Bands}a, c}). These emergent phenomena can in turn be used to generate new 2D spintronics effects (outer ellipse), such as spin\textendash charge conversion (\textcolor{blue}{\ref{fig:SP-Bands}e, f} and \textcolor{blue}{\ref{fig:SC-Conv}}), the photogalvanic effect, enhanced SOC in 2D materials, such as graphene (\textcolor{blue}{\ref{fig:SC-Conv}d, e}), magnetic skyrmions (\ref{fig:Sk-Form}\textcolor{blue}{b}) and chiral domain walls (\ref{fig:Sk-Form}\textcolor{blue}{c}), which have direct device applications (periphery). FM, ferromagnet; NM, non-magnetic material.\label{fig:SOC-Schematic}}
\caption{The ten tasks our simulator implements, which evaluate different forms of teacher response and binary feedback. In each case the same example from WikiMovies is given for simplicity, where the student answered correctly for all tasks (left) or incorrectly (right). {\color{red}Red} text denotes responses by the bot with {\color{red}S} denoting the bot. {\color{blue}Blue} text is spoken by the teacher with {\color{blue}T} denoting the teacher's response. For imitation learning the teacher provides the response the student should say denoted with {\color{blue}S} in Tasks 1 and 8. A {(\color{blue}{+})} denotes a positive reward. }
\caption{Visualization of grounded relationship expressions in the Visual Genome dataset, trained with weak supervision (subject-GT). In each example, the first and the second column show ground-truth region pairs and our predicted region pairs respectively (subject in \textcolor{red}{red} solid box and object in \textcolor{green}{green} dashed box). The third column visualizes attention weights in Eqn. \ref{eqn:attention_1}--\ref{eqn:attention_3} for subject, relationship and object (darker is higher).}
\caption{Example pointing questions in the Visual-7W dataset. The left column shows the 4 multiple choices (ground-truth answer in \textcolor{bananayellow}{yellow}) and the right column shows the grounded subject region (predicted answer) in solid box and the grounded object region in dashed box. A prediction is labeled as correct if the predicted subject region matches the ground-truth region.}
\caption{% Time evolution of the terms in the equation for kinetic energy $E_k=\mathbf{u}\cdot\mathbf{u}/2$, averaged over the computational domain: $(a)$ case D5; $(b)$ case D11. As shown by \cite{chouippe:15a} the evolution equation reads as follows: $0= -\mbox{d}\langle E_k\rangle_\Omega/\mbox{d}t -\varepsilon_\Omega +\Psi^{(t)} +\Psi^{(p)}$, where $\langle\cdot\rangle_\Omega$ indicates averaging over the spatial domain, $\varepsilon_\Omega$ is the box-averaged instantaneous dissipation rate (in the graph: {\color{red}\dashed}), $\Psi^{(t)}$ the power-input due to turbulence forcing ({\color{black}\solidthick}), and $\Psi^{(p)}$ the two-way coupling term ({\color{red}\solidthick}); the time-rate-of-change term is plotted in blue ({\color{blue}\solidthick}). % }
\caption{% % % % % Normalized difference between the variance of the fluid velocity and that of the particle velocity. % % The symbols correspond to: {\color{red}$\mathsmaller{\lozenge}$}, \cite{homann:10}; {\color{red}\soliddiamond}, \cite{cisse:15}; {\color{red}$\times$}, density-matched particles (cases ``N1'', ``N2'') of \cite{YeoClimentMaxey:10}; {\color{black}$\times$}, particles with density ratio $\rho_p/\rho_f=1.4$ (cases ``S1'', ``S2'') of \cite{YeoClimentMaxey:10}; {\color{black}$\bullet$}, present simulations; {\color{black}$\circ$}, case ``A-G0'' from \cite{chouippe:15a}. % The dashed line ({\color{red}\dashed}) indicates a power-law proportional to $(D/\eta)^{2/3}$. % }
\caption{% % Normalized p.d.f.\of the particle Reynolds number computed from the sphere-averaged relative velocity$\mathbf{u}_{pr}$ as defined in the text. % Line styles correspond to: {\color{black}\solidthick},~case~D5; {\color{red}\solidthick},~case~D11. % The blue dashed line indicates a Gamma distribution with shape parameter $k=3$ which for a quantity $x$ with variance unity reads: $f(x)=(3^{3/2}/2)x^2\exp(-x\sqrt{3})$. % % Mean and standard-deviation of the particle Reynolds number are listed next to the graph. }
\caption{% Normalized p.d.f.s of the linear particle acceleration. Samples corresponding to particles in contact with other particles have been eliminated. Line styles correspond to: {\color{black}\solidthick},~case~D5; {\color{red}\solidthick},~case~D11; {\color{blue}\dashed}, fit of experimental data proposed by \cite{qureshi:07}. }
\caption{% % % % % % % Variance of the single-component particle acceleration normalized by the variance of fluid particles, plotted as function of the particle diameter. % % The present data is represented by the filled circular symbols ($\bullet$); additionally, numerical values are listed next to the graph. % The open circle ({\color{black}$\boldsymbol{\circ}$}) indicates the case ``A-G0'' from \cite{chouippe:15a}. % The dark green triangles ({\color{darkgreen}$\mathsmaller{\blacktriangledown}$}) mark the experimental results of \cite{voth:02} which were obtained in a von K\'arm\'an flow at different Reynolds numbers ($Re_\lambda=140\ldots970$). % % % % % The orange colored triangles ({\color{niceorange}$\mathsmaller{\blacktriangleright}$}) indicate the more recent measurements in the same flow, carried out at Reynolds numbers $Re_\lambda=400\ldots800$, as reported by \cite{brown:09}. % The blue squares ({\color{blue}\solidsquare}) correspond to the wind-tunnel data of \cite{qureshi:07} for neutrally-buoyant particles at $Re_\lambda=160$. The magenta-colored triangles ({\color{magenta}$\mathsmaller{\blacktriangle}$}) are experimental measurements in a von K\'arm\'an flow by\cite{volk:11}, performed at different Reynolds numbers ($Re_\lambda=590\ldots1050$). Red diamonds represent the simulation data of \cite{homann:10} ({\color{red}$\mathsmaller{\lozenge}$}) and those of \cite{cisse:15} ({\color{red}\soliddiamond}). % The numerical data for density-matched particles (cases ``N1'', ``N2'') of \cite{YeoClimentMaxey:10} are indicated by the symbol ``{\color{red}$\boldsymbol{\mathsmaller{\times}}$}'', and their particles with density ratio $\rho_p/\rho_f=1.4$ (cases ``S1'', ``S2'') are shown as ``{\color{black}$\boldsymbol{\mathsmaller{\times}}$}''. % The guiding lines are proportional to $(D/\eta)^{-2/3}$ ({\color{black}\dashed}) and to $(D/\eta)^{-4/3}$ ({\color{red}\dashed}). % }
\caption{% $(a)$ Lagrangian auto-correlation coefficient of one-component, linear particle acceleration, normalized with the large-eddy turnover time $T$. The correlations have been computed during collision-free intervals only. The inset shows the same data for smaller separation times $\tau$, and normalized in Kolmogorov time units $\tau_\eta$. Line styles correspond to: {\color{black}\solidthick},~case~D5; {\color{red}\solidthick},~case~D11, % $(b)$ The integral scale $\tau_{int}$ of the auto-correlation (from integration up to the first zero-crossing) shown in $(a)$, plotted as a function of the particle size. The present data is indicated by solid circles ($\bullet$); the black open circle ({\color{black}$\circ$}) indicates the data of case ``A-G0'' from \cite{chouippe:15a}. The magenta-colored triangles ({\color{magenta}$\mathsmaller{\blacktriangle}$}) correspond to data from the von K\'arm\'an flow experiments of\cite{volk:11}; the red diamonds ({\color{red}$\mathsmaller{\lozenge}$}) represent the simulation data of \cite{homann:10}. The dashed line (\dashed) indicates a linear increase $\tau_{int}/\tau_\eta=1+0.08(D/\eta)$ }
\caption{% $(a)$ Time evolution of the standard deviation of the normalized volume of the cells of the three-dimensional Vorono\"i tesselation with the particle centers as ``sites''.% Line styles correspond to: {\color{black}\solidthick},~case~D5; {\color{red}\solidthick},~case~D11, % The dashed lines indicate the average value for a sequence of random distributions of non-overlapping spheres computed for the same number of particles and domain sizes as in the respective DNS. % $(b)$ The time-average of the standard-deviation of the Vorono\"i cell volumes (normalized by the corresponding value for a random particle arrangement), plotted as a function of the particle diameter (normalized by the Kolmogorov length scale). The DNS data is indicated by circular symbols ({\color{black}$\bullet$},~case~D5; {\color{red}$\bullet$},~case~D11), with the errorbars corresponding to the standard-deviation in time. The errorbars attached to the random data (open circles) indicate the standard-deviation over the respective random sequence. Note that the red open circle has been shifted slightly in the horizontal direction for clarity. % % % }
\caption{% % % $(a)$ Normalized radial distribution function $g(r)$ of particle positions in case D5 ({\color{black}\solidthick}) and D11 ({\color{red}\solidthick}) in the statistically stationary regime. % $(b)$ The same data as in $(a)$, but showing the excess value $g(r)-1$ in double-logarithmic scaling. The blue dashed line ({\color{blue}\dashed}) corresponds to a $-1$ power-law. % }
\caption{% P.d.f.\of the Vorono\"i cell volumes, sampled over the entire statistically stationary interval in case D5 ({\color{black}\solidthick}), as compared with the distribution obtained from a random Poisson process ({\color{black}\dashed}), % % sampled over a large ensemble. % % The cross-over points between the two curves are marked with solid circles ({\color{black}$\bullet$}). % }
\caption{% % % % % % % % % $(a)$ P.d.f.\of the size distribution of particle clusters expressed in terms of the number of member cells per cluster. The black solid line ({\color{black}\solidthick}) corresponds to case D5; the dashed line ({\color{black}\dashed}) shows the distribution for a corresponding random particle arangement. % $(b)$ P.d.f.\of the distribution of the volume of particle clusters. The blue dashed line ({\color{blue}\dashed}) corresponds to a $-16/9$ power-law. % }
\caption{% The 7 largest clusters in some snapshot of case D5, with the Vorono\"i cell faces colored uniformly per cluster.% % % % % % % % % % % % % The color code is in descending order (the number of member cells is given in parenthesis): black (90), blue (80), red (77), magenta (76), green (64), cyan (59) and yellow (56). The particles inside those cells are colored in black, while all remaining particles % have been omitted from the figure. Graphs $(a)$ and $b)$ show the same data from two different perspectives. }
\caption{% The p.d.f.\of the volume of the vortical structures, educed with the criterion of\cite{hunt:88} (using a threshold equal to $1.5$ times the standard deviation), and normalized in Kolmogorov length scales. % Line styles correspond to: {\color{black}\solidthickbullet},~case~D5; {\color{red}\solidthickbullet},~case~D11; % % {\color{blue}\solidthick}, single-phase data. % % The vertical lines indicate the particle volume of cases D5 and D11. % Note that the different lower ends of the three curves correspond to the specific values of $\Delta x^3$ in each case. }
\caption{% P.d.f.\of the distance from the particles to the nearest point on any vortical structure\citep[as educed with the criterion of][]{hunt:88}. Note that the distance $d_{CS}^{(i)}$ (defined in the text) is measured from the particle surface. % Line styles correspond to: {\color{black}\solidthickbullet},~case~D5; {\color{black}\solidthick},~case~D5 and removing coherent structures with a volume inferior to $V_p$; {\color{blue}\solidthick}, distance from randomly chosen points to the coherent structures in case~D5. . % % % % }
\caption{% Radial distribution function of 'sticky points' as defined in (\ref{equ-def-sticky-points}). The number density $n_S(r)$ is normalized with the global value $n_{SG}$. The lines correspond to: {\color{red}\solidthick} case D11; {\color{blue}\solidthick} single-phase flow. % }
\caption{% Particle-conditioned, relative number density of 'sticky points' % (as defined in \ref{equ-def-sticky-points}). % The number density $n_{PS}(r)$ is normalized with the global value $n_{PSG}$. In $(a)$ the distance is scaled with the Kolmogorov length scale $\eta$, in $(b)$ it is scaled with the particle diameter $D$. % Line styles: {\color{black}\solidthick} case D5; {\color{red}\solidthick} case D11. % }
\caption{% % % % % % % % % % % % % % $(a)$ Comparison of the particle-conditioned radial distribution function in case D5 according to different criteria: {\color{black}\solidthick}, criterion (\ref{equ-def-sticky-points}); {\color{black}\chndot}, zero-acceleration points. $(b)$ Particle-conditioned, relative number density of 'sticky points' (as in figure~\ref{fig-stickyPoints-partCondDist}), showing the full set of samples versus only those samples corresponding to clustering particles (having Vorono\"i cell volumes smaller than the lower cross-over point in figure~\ref{fig-voronoi-cell-vol-pdf}). % Line styles: {\color{black}\solidthick} case D5; {\color{blue}\solidthick} only clustering particles of case D5. % % }
\caption{Location: {\color{cyan}before} vs. {\color{violet}after}}
\caption{Qualitative results on building structured contour prediction: {\color{blue} ResNet} vs {\color{red} GT}}
\caption{Examples of road segmentation. Left: panoramic view; right: top-down view. (TP: {\color{yellow} yellow}, FP: {\color{red} red}, FN: {\color{green} green})}
\caption{\label{fig} The gluon propagator $D(p^2)$ as a function of the lattice momentum $p$. {\bf Top}: $d=2$ case with $\beta = 10.0$, considering original $\Lambda_x$ lattice volumes $80^2$ (\protect\marksymbol{oplus*}{blue}) and $1280^2$ (\protect\marksymbol{diamond*}{green}), and an extended $\Lambda_z$ lattice volume $80^2 \times 16^2 = 1280^2$ (\protect\marksymbol{square*}{red}). {\bf Bottom}: $d=3$ case with $\beta = 3.0$, considering original $\Lambda_x$ lattice volumes $32^3$ (\protect\marksymbol{oplus*}{blue}) and $256^3$ (\protect\marksymbol{diamond*}{green}) and an extended $\Lambda_z$ lattice volume $32^3 \times 8^3 = 256^3$ (\protect\marksymbol{square*}{red}). }
\caption{Calibration results on the \yearpred\dataset:$x$-axis denotes the expected fraction and $y$-axis denotes the observed fraction; ideal output is the dashed blue line. {Predicted variance (left) is significantly better calibrated than the empirical variance (right).} See main text for further details.}
\caption{Sketch of the basic picture in CSM interacting SNe. Four different zones are noted with numbers: (1) the pre-shock CSM, (2) the shocked CSM, (3) the shocked SN ejecta, and (4) the freely expanding SN ejecta. These zones are divided by boundaries corresponding to the {\color{green} forward shock}, the {\color{green} reverse shock}, and the {\color{green} contact discontinuity} between the shocked CSM and shocked ejecta where material cools, mixes via Rayleigh-Taylor instabilities, and piles up. This is often called the {\color{green} cold dense shell (CDS)} in a SN IIn or Ibn. The squiggly radial lines are meant as a reminder that X-rays and UV radiation generated in the shock can propagate out to the CSM or inward to the unshocked ejecta, changing the physical state of the gas there. A zoom-in of zones 2 and 3 is shown at the right. In practice, efficient radiative cooling can cause these two zones to collapse to very thin layers, and mixing can make them merge into one thin clumpy shell. This figure is adapted from Smith et al. (2008). %If the width of the figure is less than 7.8 cm use the % \texttt{sidecapion} command to flush the caption on the left side of % the page. If the figure is positioned at the top of the page, align % the sidecaption with the top of the figure -- to achieve this you % simply need to use the optional argument \texttt{[t]} with the % \texttt{sidecaption} command }
\caption{Green\textcolor{green}{$\ast$} (Red\textcolor{red}{$ \times $}): allowed (excluded) regions from LEP, Tevatron and LHC experiments at 95\% CL in all the 2HDMs. The solid, dashed and dotted curve display the contour for $ \Delta \chi^{2}=$ 2.30 (68.27\% CL), 6.18 (95.45\% CL) and 11.83 (99.73\% CL), respectively. Here, $m_{h}=125$ GeV and $m_{H}=m_{H^{\pm}}=m_{A}=500$ GeV. }
\caption[2D]{Example: subsampled holographic reconstruction of {\it in vivo Peranema} in motion (subsampling factor vs. time). Horizontal axis depicts time and vertical axis represents subsampling factor. Reconstructions of both amplitude and phase using all pixels on the detector are shown at top, while reconstructions from subsampled pixel array data, using a factor of 4 and 9, are in middle and bottom, respectively. Consecutive frames show {\it Peranema} motion from left to right. Frame rate: first row -- 4.4FPS, second row -- 13.6FPS, third row -- 24.8FPS. Scale bar is 22 $\mu$m. See \textcolor{blue}{Visualization 1, 2 and 3} for the full videos.}
\caption{Histograms of the ratio $R = \left\langle|\varepsilon_{C09}|\right\rangle/\left\langle|\varepsilon_{I}|\right\rangle$ using the $C09$ model (blue) and the corrected one (red), in the fast (a) and the slow wind (b).} \label{Ratio_Carbone_Mod} \end{figure} %================================= \subsection{The mean plasma density}\label{dis_rho} %================================= Another point that deserves enlightment is the influence of the mean density $\rho_0$ in the BG13 model. Indeed, the original form of ${\cal{F}}_{3}$ includes the total density $\rho= \rho_{0} + \rho_{1}$ as the following~\citep{Banerjee13}: \ba \nabla_{\boldsymbol{\ell}} \cdot \pmbmath{\cal F}_{3}(\ell) &=& \nabla_{\boldsymbol{\ell}} \cdot \left\langle 2 {\overline{\delta} \left[ \left(1 + \frac{1}{\beta} \right) e + { v_A^2 \over 2}\right] \delta ( \rho {\bf v})} \right\rangle \nonumber \\ &=& \nabla_{\boldsymbol{\ell}} \cdot \left\langle 2 {\overline{\delta} \left[ \left(1 + \frac{1}{\beta} \right) e + {v_A^2 \over 2}\right] \delta ( \rho_{0} {\bf v})} \right\rangle \label{rho_rho1} \\ &+& \nabla_{\boldsymbol{\ell}} \cdot \left\langle 2 {\overline{\delta} \left[ \left(1 + \frac{1}{\beta} \right) e + { v_A^2 \over 2}\right] \delta ( \rho_{1} {\bf v})} \right\rangle \nonumber . \, \, \, \, \ea In the incompressible limit ($\rho_1 \rightarrow 0$ and $\nabla \cdot{\bf v} =0$) the divergence of $\pmbmath{\cal F}_{3}(\ell)$ vanishes. However, since in the estimation of flux terms ${\cal F}_1$, ${\cal F}_2$ and ${\cal F}_3$ using spacecraft data, we do not explicitely apply the divergence operator $\nabla_{\boldsymbol{\ell}}$, but rather $\nabla_{\boldsymbol{\ell}} \rightarrow 1/\ell$, it is practically impossible to ensure that ${\cal F}_3$ vanishes in the incompressible limit. To guarantee the convergence of BG13 and PP98 models in the limit of incompressibility, we kept only the second term of Equation~(\ref{rho_rho1}), while the first term can be easily transformed into source terms since \ba 2\nabla_{\boldsymbol{\ell}} \cdot \left\langle {\overline{\delta} X \delta ( \rho_{0} {\bf v})} \right\rangle &=& 2 \rho_{0} \nabla_{\boldsymbol{\ell}} \cdot \left\langle {\overline{\delta} X \delta {\bf v}} \right\rangle \nonumber \\ &=& \rho_{0} \nabla_{\boldsymbol{\ell}} \cdot \left\langle X {\bf v}' - X {\bf v} + X' {\bf v}' - X' {\bf v} \right\rangle \nonumber \\ &=& \rho_{0} \nabla_{\boldsymbol{\ell}} \cdot \left\langle X {\bf v}' - X' {\bf v} \right\rangle \nonumber \\ &=& \left\langle \rho_{0} X \nabla' \cdot {\bf v}' \right\rangle + \left\langle \rho_{0} X' \nabla \cdot {\bf v} \right\rangle , \label{divrel} \ea where $X=\overline{\delta} \left[\left(1 + \frac{1}{\beta} \right)e + {v_A^2 \over 2}\right]$. It is easy to see that both expressions (\ref{divrel}) and the flux term $\pmbmath{\cal F}_{3}(\ell)$ of Equation~\ref{fcphi} converge to zero in the incompressible limit (i.e., $\nabla \cdot {\bf v}=0$ and $\rho_1$ = 0). %============================================== \subsection{The influence of the angle $\Theta_{\bf VB}$}\label{dis_theta} %============================================== In Section~\ref{data} we emphasized the importance of having relatively stationary angles $\Theta_{\bf VB}$ in order to have a more reliable estimate of the energy cascade rate (both its sign and its absolute value) when dealing with single spacecraft data, and regardless of the theoretical model used. Here we discuss two possible effects of the non stationarity of the angle $\Theta_{\bf VB}$ that may influence the estimation of the cascade rate. Let us first start with the case of the presence of sharp variations (i.e., discontinuities) in the angle $\Theta_{\bf VB}$ as in the example of Figure~\ref{AngleRotationF}. Such discontinuities may be due to different reasons such as the crossings of strong current sheets frequently observed in the solar wind and the magnetosheath~\cite{Gosling08,Chasapis15}. We estimated the energy cascade rate using BG13 from a long but non stationary time interval (04:40-06:00) that contained two discontinuities in $\Theta_{\bf VB}$ (about 05:00 and 05:40) and from shorter one (05:05-05:38) where such discontinuities were excluded. The results are shown in Figure~\ref{AngleRotationF} (bottom). As one can see the long non stationary time interval yields a non uniform energy cascade rate which changes its sign, whereas the shorter one where the $\Theta_{\bf VB}$ sharp discontinuities were excluded is more uniform and has a constant sign. This result should balance the usual wisdom arguing to use long time intervals (i.e., large number of data points) to guarantee the statistical convergence of the third-order moments estimates~(e.g., \cite{Podesta09}): the existence of a very few (i.e. statistically minor) sharp discontinuities as those in Figure~\ref{AngleRotationF} can significantly influence the estimates of the cascade rate as we showed here. \begin{figure} \includegraphics[width=1\columnwidth]{fig17.pdf} \caption{Top: sample of the fast solar wind data with sharp variations of $\Theta_{\bf VB}$ on 2010-12-14 from 04:40 to 06:00. Bottom: cascade rates computed for the entire signal (green) and after filtering (blue) to exclude the $\Theta_{\bf VB}$ rotation.} \label{AngleRotationF} \end{figure} The second possible effect of the angle $\Theta_{\bf VB}$ can come from its steady but significant variation in a single time interval. Indeed, as we argued in Section~\ref{data}, the Taylor frozen-in-flow assumption generally used on single spacecraft data allows one to convert the time sampling of the data into a 1D spatial sampling of the turbulent fluctuations along the flow direction. In anisotropic turbulence, the direction of the spatial sampling carries therefore a particular importance since the sampling can be either parallel ($\Theta_{\bf VB}\sim 0^\circ$) or perpendicular ($\Theta_{\bf VB}\sim 90^\circ$) to the mean field. These two directions, as demonstated in Figure~\ref{anisotropy}, have different values of the energy cascade rate. Therefore, if $\Theta_{\bf VB}$ oscillates strongly between $0^\circ$ and $90^\circ$ then the analysis would mix between the two cascade rates estimated along the direction parallel and perpendicular to the local magnetic field, and would lead to higher uncertainty in the estimated values. This might be the reason that explains the discrepancy in the cascade rate in the slow solar wind found between our results and those of \cite{macbride08}. %============================================== \subsection{Mean value of cascade rate and sign change}\label{dis_sign} %============================================== As explained in Section~\ref{data}, among the criteria that we used to select our statistical samples is the constant sign of the estimated cascade rate $\varepsilon_C$ over the time lag $\tau \in[10,1000]$s. This step is necessary in order to get reliable estimate of the mean cascade rate $\langle \varepsilon_C \rangle$ averaged over all the time lags $\tau$. Indeed, if the sign of $\varepsilon$ changes, the resulting average will yield (by cancellation) lower values of the cascade rates. Another alternative to this approach has been used in previous works based on performing statistical studies of the cascade rate obtained at a given value of the time lag $\tau$~\citep{macbride08,Smith09}. The choice of the particular $\tau$ value has not been justified apart from the fact that it belongs to the inertial range. The drawback of this approach is that, since the sign of $\varepsilon$ can vary within the inertial range as can be seen in Figure~\ref{AngleRotationF} and in e.g. in \cite{luca}, the choice of the value of $\tau$ may influence the conclusion regarding the nature (direct versus inverse) of the turbulent cascade. Figure~\ref{dis_tau} shows the histogram of $\varepsilon_{I}$ computed using PP98 at different values of $\tau$. For $\tau = 21$ s (Figure~\ref{dis_tau}-(a)) $\langle \varepsilon_{I} \rangle$ is positive, implying a direct cascade, whereas for $\tau = 81$ s (Figure~\ref{dis_tau}-(b)), $\langle \varepsilon_{I} \rangle$ is negative indicating an inverse cascade. This result underlines the need to be cautious when interpreting statistical results about cascade rates estimated at a single value of the time lag $\tau$ even when it belongs to the inertial range. \begin{figure} \includegraphics[width=1\columnwidth]{fig18.pdf} \caption{Histogram of the estimated cascade rate $\varepsilon_I$ in the fast wind from PP98 at two different values of the time lag, (a) $\tau = 21$ s and (b) $\tau=81$ s. The red line represents the mean value of $\varepsilon_{I}$ for the given value of $\tau$.} \label{dis_tau} \end{figure} %----------------------------------------------------------------------- \section{SUMMARY AND CONCLUSIONS} \label{conclusion} %----------------------------------------------------------------------- In this paper we provided the first statistical study of the compressible energy cascade rate in fast and slow solar wind MHD turbulence using a large survey of the THEMIS/ARTEMIS spacecraft data. The work is based on the reduced form of the isothermal compressible MHD turbulence model recently derived in~\cite{Banerjee13}. Several new results have been obtained, which include the amplification of the cascade rate and its slight isotropization (in particular in the slow wind) due to compressible fluctuations and a better definition of the inertial range thanks to a steadier (in value and sign) of the estimated compressible cascade rate over more than two decades of scales in comparison to the incompressible PP98 model. The new flux terms contained in the BG13 model were shown to play a leading role in amplifying the compressible energy cascade rate rather than the modified compressible Yaglom term. This result desagrees with the finding of \cite{Carbone09} who used an heurtistic compressible model based on a modification of the Yaglom term in PP98 model via density fluctuations. That discrepancy motivated a comparative study with the C09 model, which eventually showed that the origin of the cascade rate amplification found in \cite{Carbone09} is due to the mean solar wind velocity included in that estimation through the modified (compressible) Els\"asser variables. Other important results have been obtained such as the new empirical scaling laws relating the new compressible cacade rate to the sonic turbulent Mach number, and to the different components (magnetic, kinetic and internal) of the fluctuating energy. Interpreting those empirical laws requires further theoretical investigations. Several caveats related to the data selection and to the role angle$\Theta_{\bf VB}$ on the convergence of the energy cascade rate were highlighted. While this works based on the new BG13 model undoubtfully sheds light onto new features of solar wind turbulence, it remains however a perfectible model. Two particular aspects require to be improved. The first one is related to the source terms that could not have been estimated in this work using single spacecraft data (as they involve local divergences of the Alfv\'en and the plasma velocity fluctuations). Reliable estimation of those terms can be done using multispacecraft observations. Cluster spacecraft offer that possibility but the plasma data (density, velocity and temperature) are available only on two (out of four) spacecraft which does not allow us to obtain 3D estimation of the source terms. The recently launched MMS mission offers a more interesting alternative as both the magnetic field and plasma data are available on the four spacecraft. However, the mission in its current phase explores only the magnetopause and the magnetosheath regions (with a focus on the former) and will reach out in the solar wind only in 2018. Another possible shortcoming is the spacecraft separation ($\sim 10$km), which would not allow accurate estimation of the gradients at scales of the inertial range $>100$km)~\citep{Robert98}. Other than spacecraft data, numerical simulation of isothermal compressible MHD turbulence should allow for straightforward estimation of the source terms and their comparison to the flux terms. This task is planned for the upcoming months. On a longer run, the BG13 model needs to be extended to more general closure equations such as the polytropic one to go beyond the current simplified isothermal closure. \paragraph*{Acknowledgment.} The THEMIS/ARTEMIS data come from the AMDA data base (http://amda.cdpp.eu/). We are grateful to Dr. O. Le Contel and Dr S. Banerjee for useful discussions. FS acknowledges financial support from the ANR project THESOW, grant ANR-11-JS56-0008. The french participation in the THEMIS/ARTEMIS mission is funded by CNES and CNRS. \begin{thebibliography}{} \bibitem[Auster et al.(2009)]{Auster09} Auster, H.~U., Glassmeier, K.~H., Magnes, W., et al. 2009 pp. 235--264. New York, NY: Springer New York. \bibitem[Banerjee \& Galtier(2013)]{Banerjee13} Banerjee, S., \& Galtier, S. 2013, Phys. Rev. E, 87, 013019\bibitem[Banerjee et al.(2016)]{Banerjee16} Banerjee, S., Hadid, L.Z., Sahraoui, F. \& Galtier, S. 2016, submitted.\bibitem[Bruno \& Carbone(2005)]{Bruno05} Bruno, R., \& Carbone, V. 2005, Living Rev. Solar Phys., 2, 4\bibitem[Carbone et al.(2009)]{Carbone09} Carbone, V., Marino, R., Sorriso-Valvo, L., et al. 2009, Phys. Rev. Lett., 103, 061102 \bibitem[Chasapis et al.(2015)]{Chasapis15} Chasapis, A., Retino, A., Sahraoui, F., et al. 2015, The Astrophysical Journal Letters, 804~(1), L1. \bibitem[Cho \& Lazarian(2002)]{Cho02} Cho, J. \& Lazarian, A. 2002, Phys. Rev. Lett., 88, 245001.\bibitem[Coburn et al.(2012)]{Coburn2012} Coburn, J.T., et al. 2012, Astrophys. J., 754, 2 \bibitem[Coburn et al.(2014)]{Coburn14} Coburn, J.T., et al. 2014, Astrophys. J., 786, 52 \bibitem[Coburn et al.(2015)]{Coburn15} Coburn, J.T., et al. 2015, Phil. Trans. R. Soc. A, 373, 20140150 \bibitem[Frisch(1995)]{frisch} Frisch, U. 1995,Turbulence; the legacy of A.N. Kolmogorov (Cambridge: Cambridge Univ. Press) \bibitem[Galtier(2006)]{jltp} Galtier, S. 2006, J. Low Temp. Phys., 145, 59 \bibitem[Galtier \& Banerjee(2011)]{GB11} Galtier, S., \& Banerjee, S. 2011, Phys. Rev. Lett., 107, 134501\bibitem[Galtier(2016)]{G16} Galtier, S. 2016, Introduction to modern magnetohydrodynamics (Cambridge: Cambridge Univ. Press) \bibitem[Goldstein, Roberts \& Fitch(1994)]{Goldstein94} Goldstein, M. L., Roberts, D. A. \& Fitch, C. A. 1994, Journal of Geophysical Research: Space Physics, 99, 11519-11538\bibitem[Gosling \& Szabo(2008)]{Gosling08} Gosling, J.~T. \& Szabo, A. 2008, jgr, 113, A10103.\bibitem[Isenberg, Smith \& Matthaeus(2003)]{Isenberg03} Isenberg, P.A., Smith, C.W., \& Matthaeus, W.H. 2003, Astrophys. J., 592, 564\bibitem[Kritsuk et al.(2007)]{kritsuk07} Kritsuk, A.~G., Norman, M.~L., Padoan, P., \& Wagner, R. 2007, Astrophys. J., 665, 416\bibitem[Kritsuk, Wagner \& Norman(2013)]{kritsuk13} Kritsuk, A.~G., Wagner, R., \& Norman, M.~L. 2013, J. Fluid Mech., 729, R1\bibitem[Leamon et al.(1998)]{Leamon98} Leamon, R.J., et al. 1998, J. Geophys. Res., 103, 4775 \bibitem[MacBride, Forman \& Smith(2005)]{macbride05} MacBride, B., Forman, M., \& Smith, C.~W. 2005, ESA, 592, 613\bibitem[MacBride, Smith \& Forman(2008)]{macbride08} MacBride, B., Smith, C.~W., \& Forman, M. 2008, Astrophys. J., 679, 1644\bibitem[McFadden et al.(2009) ]{McFadden09} McFadden, J.~P., Carlson, C.~W., Larson, et al. 2009, pp. 277--302. New York, NY: Springer New York. \bibitem[Marino et al.(2008)]{Marino08} Marino, R., Sorriso-Valvo, L., Carbone, V. et al. 2008, Astrophys. J. Lett., 677, L71 \bibitem[Marino et al.(2011)]{Marino11} Marino, R., Sorriso-Valvo, L., Carbone, V., et al. 2011, Planet. Space Sci., 59, 592 \bibitem[Marsch et al.(1982)]{Marsch} Marsch, E., Schwenn, R., Rosenbauer, H., et al. 1982, J. geophys. Res., 87, 52 \bibitem[Marsch(2006)]{Marsch06} Marsch, E. 2006, Living Rev. Sol. Phys., 3, 1 %\bibitem[Matthaeus, Goldstein \& Roberts(1990)]{Matthaeus90} %Matthaeus, W.~H., Goldstein, M.~L., \& Roberts, D.~A. 1990, J. Geophys. Res., 95, 20673 \bibitem[Matthaeus et al.(1999)]{Matthaeus99} Matthaeus, W.~H., et al. 1999, Phys. Rev. Lett., 82, 3444 % %\bibitem[Osman et al.(2011)]{Osman} %Osman, K.~T., Wan, M., Matthaeus, W.~H., et al. 2011, Phys. Rev. Lett., 107, 165001 \bibitem[Podesta et al.(2007)]{Podesta2007} Podesta, J.J., et al. 2007, Astrophys. J., 16, 99 \bibitem[Podesta et al.(2009)]{Podesta09} Podesta, J.J., et al. 2009, Nonlin. Processes Geophys., 664, 1 \bibitem[Politano \& Pouquet(1998)]{PP98a} Politano, H., \& Pouquet, A. 1998, Phys. Rev. E, 57, 21\bibitem[Roberts et al.(1987)]{Roberts87} Roberts, D.~A., Goldstein, M.~L., Klein, L.~W. et al. 1987, Journal of Geophysical Research: Space Physics, 92~A11, 12023--12035. \bibitem[Robert et al.(1998)]{Robert98} Robert, P., Dunlop, M.~W., Roux, A. et al. 1998, ISSI Scientific Reports Series, 1, 395-418. \bibitem[Sahraoui et al.(2009)]{Sahraoui09} Sahraoui, F., Goldstein, M.~L., Robert, P., et al. 2009, Phys. Rev. Lett., 102, 231102 \bibitem[Sahraoui et al.(2010)]{Sahraoui10} Sahraoui, F., Goldstein, M.~L., Belmont, G., et al. 2010, Phys. Rev. Lett., 105, 131101 \bibitem[Servidio(2015)]{Servidio15} Servidio, S. 2015, Private Communication \bibitem[Schmidt, Federrath \& Klessen(2008)]{Schmidt08} Schmidt, W., Federrath, C., \& Klessen, R. 2008, Phys. Rev. Lett., 101, 194505\bibitem[Smith et al.(2001)]{Smith01} Smith, C.W., et al. 2001, J. Geophys. Res., 106, 8253 \bibitem[Smith et al.(2006)]{Smith06} Smith, C.W., et al. 2006, Astrophys. J. Lett., 645, L85 \bibitem[Smith et al.(2009)]{Smith09} Smith, C.W., et al. 2009, Phys. Rev. Lett., 103; 201101 \bibitem[Sorriso-Valvo et al.(2007)]{luca} Sorriso-Valvo, L., Marino, R., Carbone, V., et al. 2007, Phys. Rev. Lett., 99, 115001 \bibitem[Stawarz et al.(2009)]{Stawarz2009} Stawarz, J. E. et al. 2009, Astrophys. J., 697, 2 \bibitem[Vasquez et al.(2007)]{Vasquez07} Vasquez, B., Smith, C.~W., Hamilton, K., et al. 2007, J. Geophys. Res., 112, 7101 \end{thebibliography} \end{document} }
\caption{\textbf{Mask setups.} Illustration of the setups used for: (a) object enhancement, (b) distractor attenuation and (c) decluttering. We increase the saliency in \textcolor[rgb]{1,0,0}{red}, decrease it in \textcolor[rgb]{0,0,1}{blue} and apply no change in \textcolor[rgb]{0.3,0.3,0.3}{gray}.}
\caption{\textbf{Qualitative assessment}: The template marked in green (a), is detected in the target image (b) using three Template Matching methods: \textcolor{bbs}{\textbf{BBS}}, \textcolor{dis}{\textbf{DIS}} and \textcolor{ddis}{\textbf{DDIS}} (all using the RGB features). (c-e) The corresponding detection likelihood maps show that DDIS yields more peaked maps that more robustly identify the template. Going over the rows from top to bottom: (1) BBS prefers a target location where the background matches the template over the location where the motorcycle is at. This happens because the motorcycle deforms and hence there are few bi-directional correspondences between its template appearance and target appearance. DIS and DDIS use more information -- they consider all the one-directional correspondences. Therefore, they locate the motorcycle correctly. (2) The trophy won by the band \emph{One Direction} is fully seen in the template, but occluded in the target. Nonetheless, DDIS finds it (as Section~\ref{sec:youOnlyNeedOneDirection} said, we only need one direction...). (3) Complex deformations together with occlusion confuse both DIS and BBS, but not DDIS.}
\caption{\textbf{Find the optimal $\tau$.} \textcolor[rgb]{1,0,0}{TODO}.}
\caption{{\it left} \grays counts map above 10~GeV in the inner $5^{\circ}$ around NGC 3603. The identified 3FGL catalog sources are labeled as red crosses. The two unassociated catalog sources are labeled as green cross. The position of NGC 3603 is marked as a magenta diamond. {\it middle} : The residual map after subtracting all the identified catalog source and the diffuse background. {\it right} : The residual map after subtracting all the identified catalog source and the diffuse background, as well as the unassociated catalog source 3FGL 1111.9-6038. Also shown is the best fit gaussian disk (white circle, the radius is corresponding to the $1~\sigma$ of the Gaussian) and the position of the five point sources (white "x") used to test the hypothesis that the extended emission comes from several independent point sources. }
\caption{The SEDs of the extended emission towards NGC 3603 together with the predicted \gray emission in the \ion{H}{ii} region assuming assuming the CR density therein is identical to the density in the solar system measured by AMS-02 \citep{ams02proton}. The \ion{H}{ii} density are derived by using equation.5 of \citet{sodroski97} and free-free intensity from \citet{planck15-10}. The electron density $n_e$ are chosen to be $2 ~\rm cm^{-3}$ and $10 ~\rm cm^{-3}$, as shown in black and red curve, respectively. % % }
\caption{Distributed executor's intermediate results of field attention. \textbf{Top}: Trained in an end-to-end fashion (a--d). \textbf{Bottom}: One-round co-training of distributed and symbolic executors (e--h). The {\color{red} red} plot indicates incorrect field attention.}
\caption{\label{code:lassoc-args} The \inline{lassoc} abstraction taking the rule arguments into the account. On the right, an example derivation of the abstraction for the element rule \inline{|(S)->|Value|->(S)|}. Almost all term arguments are obtained through simple substitution, but the input arguments for \inline{Value} (marked in {\color{langdslargdefault}pink}) are added later as the default arguments. The input parameter for \inline{N} is added too, because otherwise the first \inline{S} would be undefined in the production. Note that two stacks are passed into \inline{action}, while only the second one should actually be used. }
\caption{A semantic action with $n$ input and $m$ output attributes, which are mapped to DeepCPS function and its continuation. For convenience, the action body is put after, and not in between input and output parameters. The DeepCPS section between the curly braces ({\color{langdsldeepcps}green}) skips the head of a lambda, as it is auto-generated as: \inline{(i$_1$, i$_2$, ..., i$_n$, return)'[parse]'}. \label{code:action} }
\caption{ In this detailed illustration of the architecture, we show the contraction throughout the encoder in the first row, followed by the two decoder branches, showing the expansion in the network. Captions above/below the boxes show the layer names, as well as the number of outputs/feature maps. In the first row we also show the kernel sizes for each layer, while we do not display the fixed kernel size of 4x4 for the Upconv layers of rows 2 and 3 (see table~\ref{table:arch_details}). The fully-convolutional architecture can be used with different input sizes, and thus no resolution is displayed in this figure. Each \inlinegraphics{arrow.png} represents a convolutional layer with a kernel size of 3x3, and stride and padding values of 1, which preserves the spatial dimensions, and maps its higher dimensional input blob to a flow or frame prediction. These low resolution predictions are then up-sampled and concatenated to the input of the next layer, along with the corresponding features from the encoder. The ``On-Off'' losses of the flow prediction branch are all synchronized to (de)activate the branch when necessary. }
\caption{Registration results of brain images using transformation $\tau_{2,7/2}$ with various support sizes $c.$ The source landmarks are marked by a circle (\textcolor{red}{$\circ$}), while the target ones by a star (\textcolor{green}{$\ast$}).}
\caption{Scattering intensity $I(q)$ of deuterated silk samples at $c_P=37\ \mu$M: 0.5 mM NaCl ($\triangle$), 1.0 mM NaCl ($\blacktriangle$), 0.5 mM HCl (\textcolor{red}{$\Circle$}), and 1.0 mM HCl (\textcolor{red}{$\CIRCLE$}). }
\caption{(a) Lithium (\textcolor{red}{$\Circle$}) and bromine (\textcolor{blue}{$\CIRCLE$}) ions per protein during twelve extra dialysis iterations beyond the normal reconstitution protocol. Iteration number one represents the ion concentrations present after the standard reconstitution protocol. (b) Lithium ions in the elution with the addition of either HCl (\textcolor{red}{$\Circle$}) or NaCl ($\bigtriangleup$). (c) A common segment of the protein sequence (G-A-G-A) with lithium ions drawn to associate with carbonyl oxygens.}
\caption{$I(q)$ changes monotonically as ME increases along the arrow from $0\to58$ except for the two lowest values of ME: $\mathrm{ME}=0$ (\textcolor{YellowOrange}{$\CIRCLE$}) and $\mathrm{ME}=1.9$ (\textcolor{Magenta}{$\CIRCLE$}). $I(q)$ for each value of ME are fit with a double Guinier Porod model and are plotted as the solid lines.}
\caption{The emission line diagnostic diagrams used in this paper. In the \textit{top} row, we show the classic BPT diagram (\OIIIHb\vs.\NIIHa) (\textit{left}) and the \HeII\diagram (\textit{right}). AGN candidates selected on the BPT diagram are shown as blue diamonds, while AGN candidates selected on the \HeII\diagram are shown as red points. The grey shading represents the total population of AGN candidates. In the\textit{bottom} row, we show how the respective AGN selections correspond to each other: in the \textit{bottom-left}, we show the \HeII\diagram with the BPT-selected AGN candidates shown as blue diamonds. 95\% of the BPT-selected AGN candidates are also selected as AGN candidates using the \HeII\diagram. In the\textit{bottom-right}, we show the BPT diagram with the \HeII-selected AGN candidates as red points. While some \HeII\AGN candidates lie in the Seyfert region of the BPT diagram, many do not and scatter across the transition region and into the purely star-forming locus.}
\caption{We show galaxy colour-mass diagrams with the various AGN selections shown in the different panel. In the top panel of the left column, the BPT AGN candidates are shown as blue diamonds. They are predominantly located in the green valley. In the middle panel of the left column, the \HeII\selected AGN candidates are shown as red points. They are predominantly located in the green valley and scatter into the blue cloud. In the bottom panel of the left column we show those\HeII\selected AGN candidates\textit{which are not identified by the BPT method} as red triangles. They are concentrated in the blue cloud. The \HeII\\textcolor{red}{-} only AGN candidates host galaxies are less massive than those selected by BPT diagram. In the right hand column we show the respective fractions of AGN from the various selection methods. These AGN fraction colour-mass diagrams highlight the concentration of AGN hosts in the green valley, with the\HeII\-only AGN being the only population with an elevated fraction in the blue cloud. The fraction plots also highlight that even the prevalence of\HeII\- only AGN in the blue cloud do not lead to an elevated overall AGN fraction in highly star-forming galaxies -- there are too many underlying star-forming galaxies.}
\caption{We show the $u-r$ colour histograms of the BPT selected and \HeII\AGN host galaxies, and compare them to the total galaxy sample. The\HeII\selected AGN candidates are bluer than the BPT selected AGN host galaxies.}
\caption{Sample dialogues for the two {\it Question Clarification} tasks (rows) using both the traditional {\it QA} setting (left column) and {\it AQ} setting (right column). In each case the same example is given for simplicity. Black text prefixed by ``kb:'' denotes KB knowledge that the student has access to. {\color{blue}Blue} text is spoken by the teacher, denoted by {\color{blue}T}. {\color{blue}{(+)}} denotes a reward of 1 (and 0 otherwise) that the teacher assigns to the bot. {\color{red}Red} text denotes responses or questions posed by the bot, denoted by {\color{red}S}. {\color{brown}{Brown}} denotes typos deliberately introduced by the authors. For the {\it Question Verification} setting, the student can either ask a correct (pertinent) question (as in this example) or an incorrect (irrelevant) one. The teacher will give positive or negative feedback based on the correctness of the student's question. In our offline superised learning experiments, the probability of asking pertinent questions and correctly answering the original question from the teacher is set to 0.5. Finally, {\color{blue}T}/{\color{red}S } denotes 5 pairs of questions and answers that are irrelevant to the rest of the conversation. %{\color{yellow} JIWEI: COULD YOU PLEASE CHANGE THESE EXAMPLES OF BOT'S ANSWERS/QUESTIONS TO BE INCORRECT AS WELL? APPLY THIS TO ALL THE TASK FIGURES.} }
\caption{(a) The electron heating rate profile and (b) the effective electron temperature profile for a parallel plate capacitively coupled oxygen discharge at 10 ({\color{blue}blue}), 50 ({\color{red}red}) and 200 (black) mTorr with a gap separation of 4.5 cm driven by a 222 V voltage source at 13.56 MHz. The four cases explored are: detachment neither by O$_2$(a$^1\Delta_{\rm g}$) nor O$_2$(b$^1\Sigma_{\rm g}^+$) (dashed line), only detachment by O$_2$(b$^1\Sigma_{\rm g}^+$) included (dotted line), only detachment by O$_2$(a$^1\Delta_{\rm g}$) included (dash-dot line), and both detachment by O$_2$(a$^1\Delta_{\rm g}$) and O$_2$(b$^1\Sigma_{\rm g}^+$) included (full reaction set) (solid line). \label{JdotE_all}}
\caption{ (a) The center electronegativity and (b) the average electronegativity as a function of pressure for a parallel plate capacitively coupled oxygen discharge at 50 mTorr with a gap separation of 4.5 cm driven by a 222 V voltage source at 13.56 MHz. The four cases explored are: detachment neither by O$_2$(a$^1\Delta_{\rm g}$) nor O$_2$(b$^1\Sigma_{\rm g}^+$) ({\color{black}{\bf o}}), only detachment by O$_2$(a$^1\Delta_{\rm g}$) included ({\color{green} ${\bf \times}$}), only detachment by O$_2$(b$^1\Sigma_{\rm g}^+$) included ({\color{blue}{\bf +}}), and both detachment by O$_2$(a$^1\Delta_{\rm g}$) and O$_2$(b$^1\Sigma_{\rm g}^+$) included (full reaction set) ({\color{red}${\bf \Box}$}). \label{alpha0}}
\caption{\textbf{ScenePoseNet:} An illustration of our network architecture. Our network is comprised of three basic units: \textcolor{blue}{convolutional} block, \textcolor{brown}{residual} block and \textcolor{violet}{spatial-belief} block. Our network uses information from multiple different spatial contextual regions via skip connections ($\boxplus$ denotes the concatenation operation). The input image is mapped to the ideal heat maps for part localization, pedestrian center and segmentation mask.}
\caption{Qualitative results of our approach predicting \textcolor{green}{bounding box}, body pose in terms of part locations \textcolor{blue}{(skeleton)} and a \textcolor{red}{(segmentation mask)}. The first row shows examples where the pedestrian is occluded.}
\caption{Two real example results of our proposed model. Given an image-question pair, our model generates not only an answer, but also a set of reasons (as text) and visual attention maps. The colored words in the question have Top-3 weights, ordered as {\color{red}red}, {\color{blue}blue} and {\color{cyan}cyan}. The highlighted area in the attention map indicates the attention weights on the image regions. The Top-3 weighted visual \facts are re-formulated as human readable reasons. The comparison between the results for different questions relating to the same image shows that our model can produce highly informative reasons relating to the specifics of each question.}
\caption{Some qualitative results produced by our complete model on the Visual Genome QA test split. Image, QA pair, attention map and predicted Top-3 reasons are shown in order. Our model is capable of co-attending between question, image and supporting \facts. The colored words in the question have Top-3 identified weights, ordered as {\color{red}red}, {\color{blue}blue} and {\color{cyan}cyan}. The highlighted area in the attention map indicates the attention weights on the image regions (from red: high to blue: low). The Top-3 identified \facts are re-formulated as human readable reasons, shown as bullets.}
\caption{Calculated electrical conductivity of SnSe vs. temperature. At each temperature, calculations were performed using the same chemical potential as for the Seebeck coefficient (\textcolor{blue}{Fig. \ref{fig:seebeck}}).}
\caption{a) Variation of the relaxation time with temperature for single crystals and polycrystals. Calculations were performed following the methodology explained in \textcolor{blue}{Sec.~\ref{method}}. b) Conductivity of single crystal SnSe calculated with the relaxation times of \textcolor{blue}{Fig. \ref{subfig:tau}}.}
\caption{Cumulative plots of the accuracy of pose matching between randomly drawn query patches containing crowded organisms and a gallery of training patches. The accuracy is judged using symmetric Hausdorf distance between centerlines (using ground truth annotations). For matching, we compare distances between SON-descriptors ({\color{blue}blue line}), SIFT descriptors with optimally picked radii ({\color{red}red line}) and the $L_2$-distance between raw image patches ({\color{green}green line}). The plots show the number of samples (y-axis) with the pose distance less than threshold (x-axis). In all three datasets, the SON-distance yields better performance than SIFT and raw patches. }
\caption{The paradigm of the strong German verb {\sc treffen}, which exhibits an irregular ablaut pattern. Different parts of the paradigm make use of one of four bolded theme vowels: \textcolor{darkgreen}{{ e}}, \textcolor{darkred}{{\bf i}}, \textcolor{darkpurple}{{\bf a}} or \textcolor{darkblue}{{\bf{\"a}}}. In a sense, the verbal paradigm is partitioned into subparadigms. To see why multi-source models could help in this case, starting only from the infinitive \textcolor{darkgreen}{tr{\bf e}ffen} makes it difficult to predict subjunctive form \textcolor{darkblue}{tr{\bf{\"a}}fest}, but the additional information of the fellow subjunctive form \textcolor{darkblue}{tr{\bf{\"a}}fe} makes the task easier. }
\caption{Four possible input configurations in multi-source morphological reinflection (MRI). In each subfigure, the target form on the right is {\bf \textcolor{darkpurple}{purple}}. The source forms are on the left and are {\bf \textcolor{darkgreen}{green}} if they can be used to predict the target form (also connected with a dotted line) and {\bf \textcolor{darkred}{red}} if they cannot. There are four possible configurations: (i) {\sc AnyForm} is the case where one can predict the target form from any of the source forms. (ii) {\sc SingleForm} is the case where only one form can be used to regularly predict the target form. (iii) {\sc MultiForm} is the case where multiple forms are {\em necessary} to predict the target form. (iv) {\sc NoForm} is the case where the target form cannot be regularly derived from any of the source forms. Multi-source MRI is expected to perform better than single-source MRI for the configurations \textsc{SingleForm} and \textsc{MultiForm}, but not for the configurations \textsc{AnyForm} and \textsc{NoForm}. \figlabel{sample}}
\caption{Visual depiction of our multi-source encoder-decoder RNN. We sketch a two encoder model, where the left encoder reads in the present form \textcolor{darkgreen}{tr{\bf e}ffen} and the right encoder reads in the past tense form \textcolor{darkblue}{tr{\bf a}fen}. They work together to predict the subjunctive form \textcolor{darkpurple}{tr{\bf{\"a}}fen}. The shadowed red arcs indicate the strength of the attention weights---we see the network is focusing more on \textcolor{darkblue}{{\bf a}} because it helps the decoder better predict \textcolor{darkpurple}{{\bf{\"a}}} than \textcolor{darkgreen}{{\bf e}}. We omit the source and target tags as input for conciseness.}
\caption{Qualitative results for (a) atherosclerotic histology and (b) BraTS datasets. Query (Q) image along with `ground truth' (GT) and top fetched cross-modal images (\textcolor{green}{green} box - similar annotation as Query and \textcolor{red}{red} box - dissimilar annotation ).}
\caption{Meshes of PH models for polycrystalline aggregates: (a) parallelepipedic tessellation of (10 $\times$ 10 $\times$ 10) 1000 grains with the aspect ratio $\ell/d$=1; (b) Vorono\"i tessellation of 100 grains with$\ell/d$=1.00$\pm$0.34; (c) parallelepipedic tessellation of (10 $\times$ 10 $\times$ 10) 1000 grains with $L/D$=$\ell/d$=2, $L$ and $D$ being defined as the length and width of the volume element respectively; tessellations of $L/D$=$\ell/d$=1,2,5,10 and 20 are considered in this work; (d) Individual columnar grain for parallelepipedic tessellations in the case of $\ell/d$=2, elongated along the wire direction $x_1$; (e) parallelepipedic tessellation of (31 $\times$ 31) 961 grains with $\ell/d \rightarrow \infty$. Periodic boundary conditions, denoted \#, are considered.}
\caption[Sc_moments]{\label{fig:Sc_moments} Comparison between benchmark (black dashed lines) and reconstructions (colored lines) of (a) number density with respect to $s_{\text{c}}$ and (b) $N_{\text{CCN}}(s)/N$ \blue{for different combinations of moments of $n(s_{\text{c}})$, computed directly from the benchmark population, reveals that only~six moments $\mu_l$ are needed, but these moments must be carefully selected.}}
\caption[DiaKap_moments]{\label{fig:Sc_reconstruct} (a--c) The number distribution with respect to $s_{\text{c}}$ for the benchmark (dashed curves) are compared with reconstructions from the quadrature (blue curves), where the projection of the underlying quadrature abscissas and weights, \blue{indicated by the location and magnitude of the vertical stems.} (d--f) The bivariate moments used to construct the quadrature do not effectively constrain $N_{\text{CCN}}(s)/N$, where bounds on possible solutions are shown by the shading, but the continuous distributions reconstructed from the quadrature (blue curves) accurately represent the benchmark (black dashed curves). Results are shown for the three particle-resolved populations from Figure~\ref{fig:DiaKap_quad}, which were sampled at (a,d) 7:00~am, after 1~hour of simulation, (b,e) 12:00~pm, after 6~hours of simulation, and (c,f) 6:00~am the following day, after 24~hours of simulation.}
\caption{All words of a text contribute to the predicted location \includegraphics[width=0.02\textwidth]{reddot.png}.}
\caption{Only words filtered through the distributional model contribute votes to yield a prediction \includegraphics[width=0.02\textwidth]{reddot.png} very close to the correct position \includegraphics[width=0.02\textwidth]{greendot.png}.}
\caption{Stability plots for the two-point correlation function. Stability plot under varying $t_{min}$ with fixed $t_{max}=10$ is shown for fits with two states ({\color{kngrey}\ding{116}}), three states ({\color{knred} \ding{110}}), four states ({\color{knorange} \ding{108}}), five states ({\color{knyellow} \ding{117}}), six states ({\color{kngreen} \ding{115}}), seven states ({\color{knblue} \ding{54}}), and eight states ({\color{knpurple}\ding{58}}). The corresponding $Q$-values of the fits are plotted below with the dashed line set at $Q=0.1$. The solid symbols mark the fit with the largest Bayes factor at fixed $t_{min}$. The blue band highlights the preferred fit with seven states at $t_{min}=3$ and guides the eye for observing stability.}
\caption{Results of different history information encoding approaches on MS COCO. $\text{CNN}_{\mathcal{L}_{N \text{words}}}$ takes $N$ previous words as inputs, \textcolor{\MARK}{where} we set $N$ to 2, 4, and 8. $\text{Avg}_{\text{history}}$ computes an average over history word embeddings. $\text{CNN}_{\mathcal{L}_{\text{16 words}}^{\ast}}$ replaces the $2^{\text{nd}}$ and $4^{\text{th}}$ convolutional layers in $\text{CNN}_{\mathcal{L}}$ with the \textit{max-pooling} layer.}
\caption{The grey dots show the relation of \Deltam\vs. the predictions obtained by using PLS on the PCA space of spectra for observed SNe\,Ia. The models are deflagrations (top-right), delayed detonation (black) and modified-mixing models (cyan) (top-left), sub-Chandra (bottom-left), mergers (bottom-right). Errors on the predictions come from k-folding\citep{2015MNRAS.447.1247S} and are smaller than the size of the diamonds. The variance on the observed \Deltam\of mergers and delayed-detonation models comes from the variability due to line-of-sight effects.}
\caption{The absolute $B$-band maximum without reddening corrections and the predicted $B$max using PCA and PLS. The models are W7 (top-right, green), 3D deflagrations (top-right, blue), delayed detonation (black) and modified-mixing models (cyan) (top-left), sub-Chandra (bottom-left), mergers (bottom-right). Errors on the predictors come from k-folding on the PLS regression analysis \citep{2015MNRAS.447.1247S}. (See Fig. \ref{fig:BmV_models}). The models that are far away from the blue strip can not explain normal SNe\,Ia. However, some can be good candidates for some peculiar SNe\,Ia (e.g. SNe marked in yellow).}
\caption{\emph{Nearest-Neighbor Correlation in Dynamics.} Space-time evolution of the $z$ spin-spin correlation during the quench into the antiferromagnetic phase where the correlation between sites $i$ and $i + 1$ builds up as the external field vanishes. As expected for staggered order in an antiferromagnetic state the correlation is negative. Further we point out more strongly negative correlation due to boundary effects at the end of the quench, where the affected region is indicated with arrows ({\color{blue}$\uparrow$}). \label{fig:DynMeasAntiferroCorrxx3d}}
\caption{(Color online) The relaxation processes of the sine shear at a amplitude of 94 $s^{-1}$ : The shear rates vary following sine function. The experiments were performed under 20 ${}^\circ$C. The amplitudes of sine shear are defined as their absolute values of shear rate, $|\dot\gamma|$. We tried five different frequencies, 0 Hz (+), 6.25$\times$$10^{-2}$ Hz (\textcolor{red}{$\circ$}), 3.13$\times$$10^{-2}$ Hz (\textcolor{blue}{$\bullet$}), 1.56$\times$$10^{-2}$ Hz ({\color{yellow}{$\diamondsuit$}}), 7.81$\times$$10^{-3}$ Hz ({\color{green}{$\blacklozenge$}}). 0 Hz corresponds to the constant shearing. Solid lines are each fitting lines obtained from each plots. }
\caption{(Color online) The relaxation processes of constant, sine and square shears at a amplitude of 70 $s^{-1}$ : The experiments were performed under 20 ${}^\circ$C. constant shear : (0 Hz) (+), sine shear : 6.25$\times$$10^{-2}$ Hz (\textcolor{red}{$\circ$}), 3.13$\times$$10^{-2}$ Hz (\textcolor{blue}{$\bullet$}), 7.81$\times$$10^{-3}$ Hz({\color{green}{$\diamond$}}), square shear : 6.25$\times$$10^{-3}$ Hz (\textcolor{red}{$\vartriangle$}). 3.13$\times$$10^{-2}$ Hz (\textcolor{blue}{$\blacktriangle$}) 7.81$\times$$10^{-3}$ Hz ({\color{green}{$\triangledown$}}). Solid lines are each fitting lines obtained from each plots. }
\caption{(Color online) The plots of fitting parameters of the stretched exponential function Eq. (\ref{eq:FittingEq}) with the frequency of shear. From top to bottom : $q_1$ initial scattering vector, $t_0$ time delay, $q_1+q_2$ scattering vector at stationary state, $\tau_K$ relaxation time, $\beta$ power exponent. At $|\dot\gamma|$ of 94 $s^{-1}$ : const ($\diamond$), sine ({\color{red}{$\circ$}}). At $|\dot\gamma|$ of 70 $s^{-1}$ : const ($\blacklozenge$), sine ({\color{red}{$\bullet$}}), square ({\color{blue}{$\blacktriangle$}}). All parameters were obtained by fitting the plots of time evolution of scattering vector (Fig. \ref{fig:Fig3} and \ref{fig:Fig9}) by the least-squares method. $\beta$, $\tau_K$ are listed on Table \ref{tb:table1} and \ref{tb:table2} with $\chi^2$ value. In this paper, the data is obtained from a single process. However, we tried to estimate the statistical stability of the experiments in the same condition in the same set up. At $|\dot\gamma|$ of $94~s^{-1}$ at the constant shear, the result were $q_1+q_2$=$10.9 \pm 0.9~\mu m^{-1}$, $\beta$ = $0.46 \pm 0.06$. $\tau_K$=$4015 \pm 938~s$ in the $95\%$ confidence limit.}
\caption{(Color online) The Weibull distributions of relaxation times of the relaxation processes in Fig. \ref{fig:Fig3} at a amplitudes of the shear rate of 94 $s^{-1}$. $G(\tau)=\tau D(\tau)$. These were obtained by using the obtained fitting parameters Table \ref{tb:table1}. The frequencies are : 0 Hz (---), 6.25$\times$$10^{-2}$ Hz (\textcolor{red}{- -}), 3.13$\times$$10^{-2}$ Hz (\textcolor{blue}{- -}), 1.56$\times$$10^{-2}$ Hz ({\color{yellow}{- -}}), 7.81$\times$$10^{-3}$ Hz ({\color{green}{- -}}).}
\caption{(Color online) The Weibull distributions of relaxation times of the relaxation processes in Fig. \ref{fig:Fig9} at a amplitudes of the shear rate of 70 $s^{-1}$. $G(\tau)=\tau D(\tau)$. These were obtained by using the obtained fitting parameters Table \ref{tb:table2}. The frequencies are : black circle 0 Hz (---), 6.25$\times$$10^{-2}$ Hz (\textcolor{red}{- -}), 3.13$\times$$10^{-2}$ Hz (\textcolor{blue}{- -}), 7.81$\times$$10^{-3}$ Hz ({\color{green}{- -}}). square : 6.25$\times$$10^{-2}$ Hz (\textcolor{red}{$- -\cdot\cdot$}), 3.13$\times$$10^{-2}$ Hz (\textcolor{blue}{$- -\cdot\cdot$}), 7.81$\times$$10^{-3}$ Hz ({\color{green}{$- -\cdot\cdot$}}).}
\caption{(Color online) The frequency dependence of Shannon entropies of Weibull distribution, $D_{W}(\tau)$. The plots were obtained by using the fitting parameter on Table \ref{tb:table1}, \ref{tb:table2} and Table 1 from Ref \cite{Yatabe2}. At a amplitude of 94 $s^{-1}$ : const ($\diamond$), sine ({\color{red}{$\circ$}}). At a amplitude of 70 $s^{-1}$ : const ($\blacklozenge$), sine ({\color{red}{$\bullet$}}), square ({\color{blue}{$\blacktriangle$}}). At a amplitude of $47~s^{-1}$ : square ({\color{blue}{$\times$}}), which was calculated from the data of Ref \cite{Yatabe2}. }
\caption{ \textbf{Persistent revival resonances of pseudospin-1 particles from a weak circular scatterer at low energies}. (a) Contour map of transport cross section \textcolor{blue}{in unit of $1/R$} (on a logarithmic scale) versus the scatterer strength $\rho=V_0R$ and size $x=kR$ for relativistic quantum scattering of 2D massless pseudospin-$1$ particles. Revival resonances occur, which can lead to superscattering (see Fig.~\ref{fig:SSFig2} below). (b) Similar plot for pseudospin-$1/2$ particles for comparison, where no resonances occur, implying total absence of superscattering. The scatterer is modeled as a circular step like potential $V(r)=V_0\Theta(R-r)$, representing a finite size scalar impurity or an engineered scalar-type of scatterers. The markers correspond to the theoretical prediction, where the black circles ($\circ$) and crosses ($\times$) are from $\rho\approx\rho_{0,m},\rho_{1,n}$ (for $x\ll 1$), and the red stars (\textcolor[rgb]{1,0,0}{$*$}) follow the revival resonant condition given by $\rho = 2x$ for $\rho\ll 1$.}
\caption{ \textbf{Wavefunction patterns associated with superscattering.} For $V_0R=0.5$ and $kR=0.2485$, distribution of the real part of one component of the spinor wavefunction $\Re{(\Psi_2)}$ for (a) pseudospin-$1/2$ and (b) pseudospin-$1$ wave. (c,d) Magnification of part of (a) and (b), respectively. \textcolor{blue}{Both axies in (a)-(d) are in unit of the incident wavelength $\lambda$. The color code denotes the quantity $\Re{(\Psi_2)}$.}}
\caption{MCMAC input parameters based on our previous analysis. From the \reds we can estimate the current relative radial velocity of the subclusters, $v_{\rm rad}(t_{\rm obs})$ \citep[for more details see][]{dawson}. The projected distance $d_{\rm proj}$ corresponds to the mass peak separation.}
\caption{Inverse of the critical surface density $\Sigma_{\rm cr}^{-1}$ as a function of the lens \red. The {\it black} line represents the case in which the background sources are fixed in $z_{\rm back}=1.2$ whereas in the {\it red} line the same galaxies are the closest, in $z_{\rm back}=0.9$. While the weak lensing signal ($\propto \Sigma_{\rm cr}^{-1}$) remains nearly the same for a lens located at the A3376 redshift ($\bar{z}=0.046$, {\it dotted} line), the cut of higher \red sources acts to reduce the signal of higher \red lenses. In the absence of photometric redshifts, we tried to achieve the goal of removing the higher \reds lenses by cutting the fainter sources, most of them related to higher redshift.}
\caption{Runtime comparison of the simulator from~\cite{haenerSC2016} to the simulator in ProjectQ. The timed circuit consists a Hadamard-transform, a chain of controlled z-rotations, and a final Hadamard transform. The lazy evaluation of gates in combination with intrinsics instructions allows the ProjectQ simulator to execute this circuit between 3 and 5 times faster. Both simulators were run on both cores of an Intel\textregistered{} Core\texttrademark{} i7-5600U CPU.}
\caption{Three-color composite images of the BYF\,40 area. Figure (a) (red=IRAC2, green=IRAC1, blue=Br$\gamma$) includes N$_{2}$H$^{+}$ contours shown in green and HCO$^{+}$ clumps' emission HPW shown by ellipses (From Paper 1 and \citep{b2}). It has been transformed from RA/Dec in order to show Galactic coordinates, and the grid is tilted with respect to the image frame. \color{black} The scale bar is for an adopted distance of 6.6 kpc. Figure (b) presents a composite of Mopra-only maps of the seven BYF\,40 clumps a--g, where r=N$_{2}$H$^{+}$, g=HCN, and b=HCO$^{+}$ integrated intensities, along with the same HCO$^{+}$ HPW ellipses shown in (a). The HCO$^{+}$ and HCN emission track each other very closely (overall cyan coloring), while the N$_{2}$H$^{+}$ distribution is quite different. }
\caption{The time series of message volume for a sample event reported on Twitter. The event corresponds to Twitter discussions, where each tweet contained all three terms: ``\#chapellhillshooting'', ``muslim'' and ``white''. The\textcolor{cyan}{$\medbullet$} dot corresponds to the time window having maximum \emph{message volume} while the \textcolor{maroon}{$\medbullet$} dots correspond to the minor peaks observed in this volume. The inset diagram on the right side zooms in on one of the minor peaks, along with the rule triggering its designation.}
\caption{The allowed regions in ($\lambda_{1},\lambda_{4}$) and ($\lambda_{1},m_{H^{\pm\pm}}$) plans. (\textcolor{cyan}{cyan}) : Excluded by $\mu$ constraints, (\textcolor{red}{red}) : Excluded by $\mu$+Unitarity constraints, (\textcolor{green}{green}) : Excluded by $\mu$+Unitarity+BFB constraints, (\textcolor{blue}{blue}) : Excluded by $\mu$+Unitarity+BFB+$R_{\gamma\gamma}$ constraints, (\textcolor{yellow}{yellow}) : Excluded by $\mu$+Unitarity+BFB $R_{\gamma\gamma}$\&$T_d=0$ $\land$ $T_t=0$ constraints. Only the brown area obeys ALL constraints. Our inputs are $\lambda = 0.52$, $-2 \le \lambda_1 \le 12$, $\lambda_2 = -\frac{1}{6}$, $\lambda_3 = \frac{3}{8}$, $-12 \le \lambda_4 \le 2$, $v_t = 1$ GeV and $\mu = 1$ GeV.}
\caption{\textbf{Exploring predictron variants.} Aggregated prediction errors over all predictions (20 for mazes, 280 for pool) for the eight predictron variants corresponding to the cube on the left (as described in the main text), for both random mazes (top) and pool (bottom). Each line is the median of RMSE over five seeds; shaded regions encompass all seeds. The full $(r, \gamma, \lambda)$-prediction (\red{\textbf{red}}) consistently performed best.\label{predictron_results_first_cube}}
\caption{\textbf{Comparing predictron to baselines.} Aggregated prediction errors on random mazes (top) and pool (bottom) over all predictions for the eight architectures corresponding to the cube on the left. Each line is the median of RMSE over five seeds; shaded regions encompass all seeds. The full $(r, \gamma, \lambda)$-predictron (\red{\textbf{red}}), consistently outperformed conventional deep network architectures (\textbf{black}), with and without skips and with and without weight sharing.\label{predictron_results_second_cube}}
\caption{\textbf{Comparing depths.} Comparing the $(r, \gamma, \lambda)$-predictron (\red{\textbf{red}}) against more conventional deep networks (\textbf{black}) for various depths (2, 4, 8, or 16 model steps, corresponding to 10, 16, 28, or 52 total layers of depth). Lighter colours correspond to shallower networks. Dashed lines correspond to networks with skip connections.}
\caption{\textbf{Comparing depths.} Comparing the $(r, \gamma, \lambda)$-predictron (\red{\textbf{red}}) against more conventional deep networks (\blue{\textbf{blue}}) for different numbers hidden nodes in the fully connected layers, and therefore different total numbers of parameters. The deep networks with 32, 128, and 512 nodes respectively have 381,416, 1,275,752, and 4,853,096 parameters in total. The predictrons with 32 and 128 nodes respectively have 1,275,752, and 4,853,096 parameters in total. Note that the number of parameters for the 32 and 128 node predictrons are exactly equal to the number of parameters for the 128 and 512 node deep networks.}