\caption{{\bf Ground state energies.} Best estimates of ground state energy for the $1/8$-doped 2D Hubbard model at $U/t=8$ from DMET, AFQMC, iPEPS and DMRG in units of $t$. %Also shown (open symbol) is the AFQMC corrected for the constrained-path bias. Inset: Best estimates of ground state energy for the half-filled 2D Hubbard model at $U/t=8$. \red{Here and elsewhere, error bars indicate only the estimable numerical errors of each method; uncontrolled systematic errors are not included.} For details see \cite{supplementary}. }
\caption{{\color{red}Simulated Faraday spectrum, $F(\phi)$, as a function of Faraday depth, $\phi$,} with $B_{\rm coh} = 1~{\rm \mu G}$, $L_{\rm cell} = 10$ pc, and $L_{\rm SH} = 1$ kpc, for $N_\perp=$ 1, 9, 100, and 900 from top to bottom. For each value of $N_\perp$, four different realizations are shown.\label{fig01}}
\caption{{\color{red}Simulated Faraday spectrum, $F(\phi)$, as a function of Faraday depth, $\phi$,} with $B_{\rm coh}=0, 1, 3, 5~\mu{\rm G}$; $L_{\rm cell} = 10$ pc, $L_{\rm SH} = 1$ kpc, and $N_\perp = 100$. \label{fig03}}
\caption{{\color{red}Simulated Faraday spectrum, $F(\phi)$, as a function of Faraday depth, $\phi$,} for $N_\perp=$ 1, 9, 100, and 900 (black lines), overlaid with the analytically reproduced spectrum (gray lines); $B_{\rm coh} = 1~{\rm \mu G}$, $L_{\rm cell} = 10$ pc, and $L_{\rm SH} = 1$ kpc. The simulated spectra are the same as those in Figure \ref{fig01}, but different realizations. \label{fig08}}
\caption{{\color{red}Simulated Faraday spectrum, $F(\phi)$, as a function of Faraday depth, $\phi$,} for the turbulent magnetic field reproduced with power-law spectra; $B_{\rm coh}=0,1,5~\mu{\rm G}$ from top to bottom, and $\sigma_B = 15/\sqrt{3}\ \mu$G. The spectra with different outer scales, $L_{\rm outer}$, and power-law slopes, $\beta$, are shown. The black lines are the analytic spectra of section \ref{sec3} for $\beta=-5/3$. See the main text for further details.\label{fig10}}
\caption{Example of subgraph counts. There are $6$ copies of the edge (\protect\Textedge) in all three graphs. There are $2$ copies of the triangle (\protect\Texttriangle) in a) and b), but $0$ in c). There are $1$, $0$ and $3$ copies of the square (\protect\Textsquare) in a), b) and c) respectively.}
\caption{PSNR (in dB) and SSIM comparisons on Set5, Set14 and BSD100 for $\times 2$, $\times 3$ and $\times 4$ upscaling factors among various network architectures. \textcolor{red}{Red} indicates the best and \textcolor{blue}{blue} indicates the second best performance. }
\caption{PSNR (SSIM) comparison on three test data sets for various upscaling factors among different methods. The best performance is indicated in \textcolor{red}{red} and the second best performance is shown in \textcolor{blue}{blue}. % \textcolor{red}{Red} indicates the best and \textcolor{blue}{blue} indicates the second best performance. The performance gain of our best model over all the other models' best is shown in the last row.}
\caption{\red{(a)} Static structure factor at $\phi =$ 0.7, 0.68, 0.66 and 0.653 with the non-analytic linear model with the parameters optimized in Ref.~\cite{donev2005}. \red{(b)} A zoom on the low wavevector behavior of the top panel. The inset shows the second derivative of the static structure factor with a line highlighting the apparent linear behavior of $S(k)$. \red{(c)} The structure factor at $\phi=0.653$ fitted to various functional forms.}
\caption{\red{(a)} Behavior of the coarse-grained density correlation function $P(r)$ calculated for a non-analytic structure factor of the form $S(k) = A + B|\vec{k}|$, with $A$ and $B$ taken from the fit shown in Fig.~\ref{fig:sk}. \red{(b)} A zoom on the negative correlation near $r \approx 5$ \red{of (a)}. The inset illustrates the $r^{-4}$ power law decay of $P(r)$ from negative values as $r \to \infty$.}
\caption{The measured coarse-grained density correlation function $P(r)$ at $\varphi = 0.7$, 0.68, 0.66, 0.653. using three different \red{representations} to emphasize the peak at $r = 0$ \red{(a)}, the negative dip near $r \approx 5$ \red{(b)}, and the positive correlation near $r \approx 10$ \red{(c)}.}
\caption{Behavior of $r^4 P(r)$ for \red{(a)} the measured $r^4P(r)$ at $\varphi = 0.7$, 0.68, 0.66, and 0.653, and \red{(b)} the calculated $r^4 P(r)$ from non-analytic [Eq.~(\ref{eq:claim})] and analytic [Eq.~(\ref{eq:analytic})] models of $S(k)$. The negative correlation predicted from the non-analytic model is not observed in the measured $P(r)$, which exhibits a positive correlation which is well captured by the analytic models with $n \geq 4$. }
\caption{Measured variance of density fluctuations in a spherical cavity of radius $R$ at $\varphi=0.66$. $\ave{\Delta D(R)} $ is the variance of the microscopic number density $\rho(\vec{x})$, and $\ave{\Delta \Psi(R)} $ is the variance of the coarse-grained density $\psi(\vec{x})$. The apparent anomalous scaling of $\ave{\Delta D(R)}$ originates from short-ranged correlations which are filtered out in $\ave{\Delta \Psi(R)}$. \red{This quantity obeys the scaling expected from the central limit theorem. This shows that density fluctuations are not suppressed at large scale in jammed packings. The vertical dashed line indicates half of the dimension of the simulation box. }}
\caption{\red{(a)} System size dependence of the coarse-grained density correlation function $P_N(r)$ at $\phi =$ 0.7. \red{(b)} System size dependence of the coarse-grained density correlation function with the finite size effect correction $P_N(r) + S(0)/N$ at $\phi =$ 0.7. }
\caption{Dependence on the number of \red{independent} samples of \red{(a)} $S(k)$ at $\varphi=0.7$ and \red{(b)} $r^4 P(r)$ at $\phi =0.66$. These results show that the low-$k$ behavior of $S(k)$ and the large-$r$ behavior of $P(r)$ are statistically significant.}
\caption{Specifications of decaying simulations with maximal initial magnetic helicity \cite{data15a,Linkmann15a}. $R_L$ denotes the integral-scale Reynolds number, $R_{\lambda}$ the Taylor-scale Reynolds number, $R_{-}$ the generalized Reynolds number, $\mu$ the magnetic resistivity, $k_0$ the peak wavenumber of the initial energy spectra, $k_{max}$ the largest resolved wavenumber, \blue{$\eta_{mag} = (\mu^3/\vep_{mag})^{1/4} $ and $\eta_{kin} = (\nu^3/\vep_{kin})^{1/4}$ the magnetic and kinetic Kolmogorov microscales, respectively, at the peak of total dissipation}, \# the ensemble size,$\Ceps$ the dimensionless total dissipation rate, $\sigma$ the standard error on $\Ceps$ and $\rho_c(0)$ the initial relative cross helicity. All Reynolds numbers are measured at the peak of total dissipation. }
\caption{Specifications of decaying simulations \cite{data15a,Linkmann15a} for magnetic fields with negligible initial magnetic helicity. $R_L$ denotes the integral-scale Reynolds number, $R_{\lambda}$ the Taylor-scale Reynolds number, $R_{-}$ the generalized Reynolds number, $\mu$ the magnetic resistivity, $k_0$ the peak wavenumber of the initial energy spectra, $k_{max}$ the largest resolved wavenumber, \blue{$\eta_{mag} = (\mu^3/\vep_{mag})^{1/4} $ and $\eta_{kin} = (\nu^3/\vep_{kin})^{1/4}$ the magnetic and kinetic Kolmogorov microscales, respectively, at the peak of total dissipation}, \# the ensemble size,$\Ceps$ the dimensionless total dissipation rate, $\sigma$ the standard error on $\Ceps$ and $\rho_c(0)$ the initial relative cross helicity. All Reynolds numbers are measured at the peak of total dissipation. }
\caption{Specifications of stationary simulations. ND, SF and HF refer to the forcing schemes $\vec{f}_1$, $\vec{f}_2$ and $\vec{f}_3$, respectively. $R_L$ denotes the integral-scale Reynolds number, $R_{\lambda}$ the Taylor-scale Reynolds number, $R_{-}$ the generalized Reynolds number given in \eqref{eq:gen_Rey}, $\mu$ the magnetic resistivity, $k_{max}$ the largest resolved wavenumber, \blue{$\eta_{mag} = (\mu^3/\vep_{mag})^{1/4} $ and $\eta_{kin} = (\nu^3/\vep_{kin})^{1/4}$ the magnetic and kinetic Kolmogorov microscales, respectively}, $\Ceps$ the dimensionless total dissipation rate defined in \eqref{eq:ceps_defn}, $\sigma_{\Ceps}$ the standard error on $\Ceps$ and $t/T$ the run time in stationary state in units of large-eddy turnover time $T=L_u/U$. All Reynolds numbers are time averages. }
\caption{ \blue{(a)} Eq.~\eqref{eq:model} fitted to the different datasets H, NH, CH06H, CH06NH (free decay) and ND (stationary). Data from the stationary series SF and HF are shown for comparison purposes. The black lines refer to fits using Eq.~\eqref{eq:model} up to first order, while the gray lines use Eq.~\eqref{eq:model} up to second order in $1/R_-$. The error bars show one standard error. As can be seen, the respective asymptotes differ between the data sets H, NH, CH06H, CH06NH and ND, while data from series SF and HF is compatible with data from series ND. \blue{(b) Fit of the expression $C/R_-^n$ corresponding to eq.~\eqref{eq:model} after subtraction of the asymptote $\Cinf$ on a logarithmic scale to the datasets H (free decay, top line) and ND (stationary, bottom line). Data from the series NH, CH06H, CH06NH, SF and HF are shown for comparison purposes. The resulting exponents are $n=1.00 \pm 0.01$ for free decay and $n=1.00 \pm 0.02$ for the stationary case.} }
\caption{Examples illustrating entailment relation composition. (a) for Figure~\ref{fig:align-example} (b); (b) for Figure~\ref{fig:align-example} (d). For each hypothesis tree node, the dashed line shows to its most confident alignment. The three color stripes in each node indicate the confidences of the corresponding entailment relation estimation: red for {\color{red} contradiction}, green for {\color{green} neutral}, and blue for {\color{blue} entailment}. The colors of the node borders show the dominant estimation. Note: there is no strong alignment for hypothesis word ``are'' in (a). \label{fig:ent-example}}
\caption[Compression/tension and shear waves emanating from a dipping fault]{Compression/tension and shear waves emanating from a $1$km, $1$s uniform slip across a fault with dip $0.2$rad, top-edge depth $100$km, and width $50$km, shown at four different times (\textcolor{red}{red}: compression, \textcolor{blue}{blue}: tension, black lines: contours of velocity in the direction tangent to the fault).}
\caption[Compression/tension waves and resulting vertical surface displacement evolving to the Okada solution]{Results for a $1$km, $1$s uniform slip across a fault with dip $0.2$rad, top-edge depth $20$km, and width $50$km. The upper plot at each time shows the surface deformation (\textcolor{blue}{blue solid line} - numerical surface displacement, \textcolor{red}{red dashed line} - Okada solution), and the lower plot shows the compression/tension waves (\textcolor{red}{red}: compression, \textcolor{blue}{blue}: tension).}
\caption[Vertical and horizontal surface displacement for various fault parameters matching that of Okada]{Vertical (top) and horizontal (bottom) surface displacement at $t=90$s matching that of Okada for fault parameters varying from the baseline of a $1$km, $1$s uniform slip across a fault with dip $0.2$rad, top-edge depth $100$km, and width $50$km \textcolor{blue}{Blue solid line}: numerical solution, \textcolor{red}{red dashed line}: Okada solution).}
\caption[Vertical cross-section of compression/tension waves and plane view of surface deformation]{The upper plot at each time shows a plane view of the surface deformation (\textcolor{bluegreen}{blue-green}: uplift, \textcolor{brown}{brown}: subsidence, black lines: contours of numerical solution, \textcolor{red}{red lines}: contours of the Okada solution), and the lower plot shows a vertical cross-section of the compression/tension waves through the three-dimensional computation (\textcolor{red}{red}: compression, \textcolor{blue}{blue}: tension) }
\caption{ Distributions of cell sizes $v$ and the appropriate variable in log-space $a=\ln\left(v\right) $ for the adder model. Three different noise strengths $\langle \eta^2 \rangle$ were used: $\langle \eta^2 \rangle=0.1$ (circles \textcolor{red}{$\bigcirc$} ), $\langle \eta^2 \rangle=0.1$ (dimonds \textcolor{blue}{$\Diamond$} ) and $\langle \eta^2 \rangle=0.1$ (squares \textcolor{black}{ $\Box$} ). The parameter $\Delta$ is $1$ in all plots. Panel {\bf(a)} displays the distributions of $a$, panel {\bf(b)} displays the distributions of $v$. Thick lines are appropriate analytic approximation by the virtue of Eqs.~(\ref{boltzmn}) and (\ref{adderEnrg}). Panel {\bf(c)} displays the asymptotic power-law decay of the $v$ distributions. The decay is consistent with Eq.~(\ref{powerlaw}). }
\caption{ Stable behavior of the cell size in the adder model. Panel {\bf (a)}: distribution of the cell size collapse when normalized by the mean. The symbols present different $\Delta$s: $\Delta=0.1$ (circles \textcolor{red}{$\bigcirc$} ), $\Delta=1.9$ (dimonds \textcolor{blue}{$\Diamond$} and $\Delta=5.0$ (crosses \textcolor{black}{$\times$}). The thick line is the analytic solution. Panel {\bf (b)} presents the behavior of the moments of the cell size for $\langle \eta^2 \rangle = 0.5$ (circles \textcolor{red}{$\bigcirc$}) and $\langle \eta^2 \rangle = 0.1$ (squares \textcolor{black}{$\Box$}), $\Delta$ equals $1.0$ for both cases. Thick lines represent the analytic solution for the adder model and dashed lines are solutions of the linearized adder model.}
\caption{From Example ~\ref{ex:graph-from-skeleton}: Top: the one-out graphs $G_f$, $G_g$ and $G_h$ corresponding to the skeletons $\nu$, $\sigma$, and $\tau$. In each graph, the representative of $X$ is indicated in \textcolor{red}{red} and the representative of $Y$ is in \textcolor{blue}{blue}. Bottom: the graph $G=G_\nu=G_\sigma=G_\tau$. Graphs $G_f$ and $G_g$ are subgraphs of $G$, but $G_h$ is not.}
\caption{\emph{Color online} (a) Equilibrium speed $v^\star$ of Eq.~\ref{Equation2D} for $v_x=0$ as a function of $f$. Stable solution $v^-$ (dashed black line), saddle node solution $v^+$ (solid black line) and unstable solutions (dotted black line). The critical force value $f^\star$ is defined graphically as the merging of the saddle-node and the unstable fixed point. (b) Penetration depth reached by the particle as a function of the angle of incidence $\theta_0$ (log scale). Low force regime $f=10^{-3}<f^{\star}$ ({\color{blue} $\triangledown$}) and high force regime $f=10^{3}>f^{\star}$ ({\color{mygray} $\circ$}). Solid gray line: maximal penetration depths for an Hamiltonian particle. Solid blue line: maximal penetration depth for a self-propelled particle in the limit $f\ll f^\star$ predicted by Eq.~\ref{thetaC}. Dashed red: limit encountered by Hamiltonian particles. Inset: Same analysis performed with a harmonic force field (identical color code). (c) and (d) flow representation in the ($v_x, v_y$)-plane respectively for $f=0.1<f^{\star}$ (blue line in Fig.~\ref{Fig2}(a)) and for $f=0.7>f^{\star}$ (red line in Fig.~\ref{Fig2}(a)) of Fig.~\ref{Fig2}(c) and (d). White arrows indicate the direction of the flow. }
\caption{\emph{Color online} Evolution of the equivalent Boltzmann factor $\beta$ (log. scale along $x$) for a harmonic force field of natural frequency $\omega$ ({\color{mygray}\Large $\circ$}) and a constant force field of magnitude $f$ ({\footnotesize $\blacklozenge $}). Vertical dashed lines indicate $f^\star$ and $\omega^\star$ in gray and black respectively. {\bf Inset} Probability $\mathcal{P}$ to cross the potential barrier (log. scale) as a function of the maximal potential energy $u_{\mathrm{max}}$ of the energy potential barrier. Uniform distribution and harmonic force field ($\blacktriangle$), uniform distribution and constant force field ({\color{Red} $\bigstar$}). Gaussian distribution with $\sigma = \pi/8$ ({\color{mygray} $\times$}), $\pi/4$ ({\color{Orange} $\bullet$}), $\pi/2$ ({\color{ForestGreen} $\blacksquare$}) respectively and constant force field.}
\caption{An instance $(G,k)$ of \probIndSet with $k=2$ (highlighted in gray) transformed to an instance of \probrFixSwap. Thick edges at the bottom correspond to $k+1$ pendant vertices. A gadget has been expanded for $u_1$: the three dots hide a similar gadget for each of $u_2$, $u_3$, and $u_4$. The colors are ``\protect\tikz[baseline=0.25ex] \protect\fill[draw,fill=black] (1ex,1ex) circle (1ex); $= 1$'', ``\protect\tikz[baseline=0.25ex] \protect\fill[draw,fill=white] (1ex,1ex) circle (1ex); $= 2$'', and ``\protect\tikz[baseline=0.25ex] \protect\fill[draw,fill=gray] (1ex,1ex) circle (1ex); $= 3$''.}
\caption{\textbf{(a)} In the spread gadget, fixing the conflict at the root $r$ corresponds to choosing an instance $\ell_h$, for $h \in [t]$. \textbf{(b)} The clause gadget with its initial vertex-coloring, where the white vertices are colored such that $(P_1)$ and $(P_2)$ are respected. In both figures, thick edges correspond to $k+1$ pendant vertices. The colors are ``\protect\tikz[baseline=0.25ex] \protect\fill[draw,fill=black] (1ex,1ex) circle (1ex); $= 1$'', ``\protect\tikz[baseline=0.25ex] \protect\fill[draw,pattern=north east lines] (1ex,1ex) circle (1ex); $= 2$'', and ``\protect\tikz[baseline=0.25ex] \protect\fill[draw,fill=gray] (1ex,1ex) circle (1ex); $= 3$''.}
\caption{\textbf{(a)} A partially precolored input graph $G$ of \probkPrExtShort. \textbf{(b)} To reduce clutter, the reduction of Theorem~\ref{thm:3chrom_vil_planar} expanded for only two vertices $x \in X$ and $w \in W$. The colors are ``\protect\tikz[baseline=0.25ex] \protect\fill[draw,fill=black] (1ex,1ex) circle (1ex); $= 0$'', ``\protect\tikz[baseline=0.25ex] \protect\fill[draw,pattern=north east lines] (1ex,1ex) circle (1ex); $= 1$'', and ``\protect\tikz[baseline=0.25ex] \protect\fill[draw,fill=gray] (1ex,1ex) circle (1ex); $= 2$''.}
\caption{The sample selection: The blue diamonds represent the \hix\sample, the red circles the control sample. The light blue, wide diamond marks the data of galaxy\gal. The grey shade in the background represents the distribution of the \hipass\parent sample. The black solid line gives the\citet{Denes2014} scaling relation and the black dashed line marks, where the measured \hi\mass is 2.5\,times larger or smaller than the expected\hi\mass.}
\caption{The \hi\gas mass fraction as a function of the stellar mass. The blue diamonds present the\hix\sample and red circles the control sample. The light blue diamond is\gal, which will be discussed in more detail further below. Blue and red errorbars in the top right corner give the median errorbars for the \hix\and control sample respectively. The dark grey, solid line presents the running average of the\hipass\parent sample along with the$1\,\sigma$ scatter as the grey shaded region. The light grey, dashed line and shaded area represents the running average and $1\,\sigma$ scatter of the GASS sample. Open symbols present other surveys of (\hi\rich) galaxies:\highmass\(squares), \citet{Lemonias2014} (circles), HIGHz (asterisks) and \bluedisk\(crosses).}
\caption{The star formation efficiency as a function of stellar mass. Symbols are as in Figure \ref{fig:fhi_vs_mstar}. The red, solid line is a fit to the \hix\data and the blue, solid line a fit to the control data. The black, dotted line is the average SFE for$\rm log\,M_{\star} / M_{\odot} > 10$ galaxies found by \citet{Schiminovich2010} and the grey, dotted lines are model predictions by \citet{Wong2016}. For comparison, the light blue, dashed lines are fits to the \highmass\and\citet{Lemonias2014} samples respectively (colours as for their symbols).}
\caption{Comparison of \hi\spectra as measured from the\hipass\data cube (light blue), the ATCA data cubes (dark blue for robust=0 and medium blue for robust=0.5) and the\tirific\model data cube (red dashed line, see Section\ref{sec:tirific}).}
\caption{\textbf{(a)} The optical SuperCOSMOS \bjband\image of\gal. \textbf{(b)} The velocity field (moment 1 map) of \gal\, solid and dashed contours are separated by 20\kms and indicate the receding and approaching side respectively. The thick black contour is drawn at 10613\kms indicating the systemic velocity. \textbf{(c)} The column density (moment 0) map of ESO075-G006 overlaid on the SuperCOSMOS optical B-band image. \hi\column density contours are at (0.3, 1.3, 2.3, 3.3)\msunpcsq. \textbf{(d)} The velocity dispersion (moment 2) map of \gal. Contours are separated by 15\kms. The red ellipses in the bottom left corners of the moment maps indicate the synthesised beam size. All moment maps are created from the rob=0.5 cube.}
\caption{Radial profiles of HI properties: \textbf{(a):} The radial \hi\column density profile of\gal\in azimuthally averaged rings (black points and dashed line) and along the major axis from the centre going north-westerly (south-easterly) in light (dark) blue lines. The bottom grey dashed line marks the lower limit on column densities of$\rm 0.4 \cdot 10^{20}\,cm^{-2}$, the top grey dashed line marks the 1\msunpcsq\level. The vertical grey, solid line marks the measured$\rm R_{HI}$. \textbf{(b):} Rotation curve from \tirific, the black dashed line connects the measurements and the black solid line is a fit to the data. \textbf{(c):} The radial variation of the inclination (black) and the position angle (grey) as fitted by \tirific.}
\caption{The atomic to baryonic mass fraction as a function of the global stability parameter q. The blue diamond presents our measured data for \gal. The red circles present the data of THINGS galaxies \citep{Walter2008} as measured by \citet{Obreschkow2014}. The black line and grey shaded area mark the analytical model by \citet{Obreschkow2016}. Blue squares present \highmass\galaxies. Filled square indicate galaxies for which resolved\hi\maps have been published\citep{Hallenbeck2014,Hallenbeck2016}.}
\caption{The gas-phase oxygen abundance distribution in \gal: The grey shaded image in the background is the \galex\\nuv\image. The coloured circles indicated star forming regions within which a metallicity has been measured. The colour of the circle indicates the measure metallicity.}
\caption{\small Schematic view of neural machine translation. The \textcolor{red}{red} source words are first mapped to word vectors and then fed into a recurrent neural network (RNN). Upon seeing the $\langle$eos$\rangle$ symbol, the final time step initializes a target \textcolor{blue}{blue} RNN. At each target time step, \textit{attention} is applied over the source RNN and combined with the current hidden state to produce a prediction $p(w_t| w_{1: t-1}, x)$ of the next word. This prediction is then fed back into the target RNN.}
\caption{ Average Kendall's tau distance between the word alignments obtained after translating the test set with each MT system being evaluated and a monotone alignment (left); and average Kendall's tau distance between the word alignments obtained %of \textcolor{purple}{for} each MT system's translation and the word alignments of the reference translation (right). Larger values represent more similar alignments. If the difference between the distances depicted in the two last columns is statistically significant according to paired bootstrap resampling~\cite{koehn2004statistical} with $p = 0.05$ and $1\,000$ iterations, the largest distance is shown in bold.}
\caption{\label{f:length_chrf_enfi}NMT and PBMT chrF1 scores on subsets of different sentence length for \textcolor{purple}{the language direction} EN$\rightarrow$FI.}
\caption{Band structures of silicene on F-doped monolayer WS$_2$ for the 2 K and 2 K$'$ points not covered by Fig.\\ref{fig:fig5}(a/b). The SOC is neglected (left) or taken into account (right). The F impurity band is shown in black color. \textcolor{red}{ The different Brillouin zone paths are highlighted by dashed arrows (bottom).}}
\caption{Theoretical peak transmission deviation from unity, $1-T_{\textrm{max}}$ as a function of the wedge angle $\epsilon$ between the membranes for $w_0=50$ $\mu$m and for 35\%-reflectivity membranes ($\lambda\sim 900$ nm) with $n=1.99$, $d=8.57$ $\mu$m ({\color{red}$\square$}) and $d=200$ $\mu$m ({\color{blue}$\bullet$}). The yellow line shows the approximate expression of Eq.~(\ref{eq:Tmax}).}
\caption{(a) Transmission spectrum in the range 800-960 nm of a 100 $\mu$m-spacing array measured with broadband ({\color{blue}$\circ$}) and monochromatic ({\color{red}$\bullet$}) illuminations. The solid lines represent the results of fits of the data with the theoretical model. (b) Same data over the range 898-920 nm.}
\caption{Mechanical quality factors for the lowest frequency modes of single membranes ($\square$, {\color{defgray}$\square$}: 500 $\mu$m-thick frame without spacer, $\diamond$, {\color{defgray}$\diamond$}: 500 $\mu$m-thick frame with 8.5 $\mu$m spacer, $\circ$, {\color{defgray}$\circ$}: 200 $\mu$m thick frame) and double-membrane arrays ({\color{red}$\blacktriangledown$}, {\color{blue}$\blacktriangle$} 8.5$\mu$m-spacing, {\color{green}$\bullet$}: 200 $\mu$m-spacing). {\color{blue}$\blacktriangle$} corresponds to the array whose spectrum is shown in Fig.~\ref{fig:9um}. The corresponding square drum mode indices are indicated below the data points.}
\caption{A decision tree diagram. Denote $\mathcal{T}\triangleq\{t\in\mathbb{N}_{0}^{}; t<N : \x(t)\in\mathcal{C}_{0}^{}\}$. This means that $\mathcal{T}$ represents a set of time instances at which a new control signal does not need to be computed and transmitted from the controller to the actuator. In the figure, dark colored circles represent the case $\x(t)\in\mathcal{C}_{0}^{}$ for any $t\in\mathcal{T}$ whereas light colored circles represent the case $\x(t)\in\mathbb{R}^{n}\setminus\mathcal{C}_{0}^{}$ for any $t\notin\mathcal{T}$. For instance, when $N=4$, the set $\mathcal{T}$ gets values in $\mathsf{P}(\{1,2,3\})=\big\{\emptyset,\{1\},\{2\},\{3\},\{1,2\},\{1,3\},\{2,3\},\{1,2,3\}\big\}$.} \label{fig:Decision_tree} \end{figure} As the feasible region $\mathcal{S}$ can be expressed as a finite union of polyhedra, the disjunctive formulation can be applied. In particular, assuming $n$-dimensional problem instance~\eqref{eqn:opt_problem}, its feasibility region can be divided into $2n+1$ disjunctive polyhedra\footnote{See Fig.~\ref{fig:Example_1} for an example of two dimensional problem where each disjunctive set is shown in different colors.}. Following the construction above, we write: % \begin{align} \mathcal{S} \triangleq \bigcup_{p = 1}^{2n} \mathcal{C}_{p}^{} \;, \label{eqn:Union_set} \end{align} % where $n$ is the state dimension, and $\mathcal{C}_{p}^{}\subset\mathbb{R}_{}^{n}$ are full-dimensional polyhedra, i.e., % \begin{align*} \mathcal{C}_{p}^{} \triangleq \big\{ \x\in\mathbb{R}_{}^{n} : T_{p}^{}\x \leq \boldsymbol{d}_{}^{} \big\} \;, \end{align*} % for some $T_{p}^{}\in\mathbb{R}_{}^{(2n-1)\times n}$ and $\boldsymbol{d}_{}^{}\in\mathbb{R}_{}^{2n-1}$. We define a piece-wise constant function of time $\sigma(t)$, which takes on values in $\{0,1,\cdots,2n\}$ and whose value $p$ determines, at each time $t\in\{0,\cdots,\mathrm{T}-1\}$, the state variable $\x(t)$ to belong to the interior of the polyhedron $\mathcal{C}_{p}^{}$. We divide the non-convex set $\mathcal{S}$ into convex subsets $\mathcal{C}_{p}$ with $p\in\{0,1,\cdots,2n\}$ and, at each time step $t\in\{0,\cdots,\mathrm{T}-1\}$, we choose only one active set. The switching signal, % \begin{align} \sigma(t) = \begin{cases} 0 & \text{if} ~ t\in\mathcal{T}, \\ p & \text{otherwise}, \end{cases} \end{align} % gives a sequence $\Sigma=\{\sigma(0),\cdots,\sigma(N-1)\}$. Then, for given $\mathcal{T}$ and $\Sigma$, we rewrite the optimization problem~\eqref{eqn:opt_problem_noncvx} as % \begin{align} \begin{array}{rll} J_{\Sigma}^{\star}(\x(0)): & \stackrel[\boldsymbol{\pi}]{}{\text{minimize}} & J(\x(0),\boldsymbol{\pi}) \\ & \text{subject to} & \x(t+1) = A\x(t) + B\bu(t), \\ & & \hspace{6mm} \forall t\in\{0,1,\cdots,N-1\}, \\ & & \x(t)\in\mathcal{C}_{\sigma(t)}, ~ \forall \sigma(t)\in\Sigma, \\ & & \bu(t) = \0_{m}^{}, ~ \forall t\in\mathcal{T}, \\ & & \x(0) = \x_{0}. \end{array} \label{eqn:opt_problem_cvx} \end{align} % It is necessary to solve all possible combination of problem~\eqref{eqn:opt_problem_cvx} to determine the global optimal solution. Particularly, the problem~\eqref{eqn:opt_problem} can be solved as a sequence of $(2n+1)^{N-1}$ quadratic programming (QP) problems. Hence, we have the following series of optimization problems: % \begin{align}\label{eqn:opt_sequence_problem} \begin{array}{rll} J_{}^{\star}(\x(0)): & \stackrel[\Sigma\in\{0,\cdots,2n\}_{}^{N-1}]{}{\text{minimize}} & J_{\Sigma}^{\star}(\x(0)) \;. \end{array} \end{align} It is worth noting that even though the number of QP sub-problems grows exponentially with $N$ and $n$, all these problems are independent of each other; therefore, one can parallelize the problems and reduce the computation time. \begin{corollary} The global optimal solution of the control problem~\eqref{eqn:opt_problem} can be computed by solving a set of convex quadratic programming problems of the form~\eqref{eqn:opt_problem_cvx}, in parallel, and selecting the solution that provides the lowest control cost among all feasible solutions. \end{corollary} The following examples should give the reader a better understanding of the problem. \begin{example} Let $n=2$ and $\mathcal{S}=\mathbb{R}^{2}\setminus\mathcal{C}_{0}^{}$ with $\mathcal{C}_{0}^{} = \{ (x_{1}^{},x_{2}^{})\in\mathbb{R}^{2} : -\varepsilon < x_{1}^{} < \varepsilon, -\varepsilon < x_{2}^{} < \varepsilon \}$. The non-convex set $\mathcal{S}$ can be divided into four convex subsets: % \begin{align*} \mathcal{C}_{1}^{} &= \{ (x_{1}^{},x_{2}^{})\in\mathbb{R}^{2} : x_{1}^{} \geq x_{2}^{}, x_{1}^{} \geq -x_{2}^{}, x_{1}^{} \geq \varepsilon \} \;, \\ \mathcal{C}_{2}^{} &= \{ (x_{1}^{},x_{2}^{})\in\mathbb{R}^{2} : x_{1}^{} \leq x_{2}^{}, x_{1}^{} \geq -x_{2}^{}, x_{2}^{} \geq \varepsilon \} \;, \\ \mathcal{C}_{3}^{} &= \{ (x_{1}^{},x_{2}^{})\in\mathbb{R}^{2} : x_{1}^{} \leq x_{2}^{}, x_{1}^{} \leq -x_{2}^{}, x_{1}^{} \leq -\varepsilon \} \;, \\ \mathcal{C}_{4}^{} &= \{ (x_{1}^{},x_{2}^{})\in\mathbb{R}^{2} : x_{1}^{} \geq x_{2}^{}, x_{1}^{} \leq -x_{2}^{}, x_{2}^{} \leq -\varepsilon \} \;, \end{align*} % which are illustrated by different colors in Fig.~\ref{fig:Example_1}. The optimal control inputs can be computed by solving quadratic programs for different combinations of convex sets. In order to compute optimal control inputs, it is necessary to solve $5_{}^{N-1}$ convex quadratic programming problems. \end{example} \begin{example} Let us consider the second-order system: % \begin{align} \x(t+1) = \begin{bmatrix} 0.9 & 0.2 \\ 0.8 & 1.5 \end{bmatrix} \x(t) + \begin{bmatrix} 0.6 \\ 0.8 \end{bmatrix} \bu(t) \;, \label{eqn:example_system} \end{align} % with the initial condition $\x_{0}^{}=\left[\begin{smallmatrix} 0 & -1 \end{smallmatrix}\right]_{}^{\intercal}$. The event-triggered controller transmits control messages if the $\ell_{\infty}^{}$-norm of the state variable $\x(t)$ is greater than $\varepsilon = 0.25$. The performance indices are given by $Q=\mbox{diag}\{2,2\}$ and $R=5$. The control horizon is chosen as $N=7$. The state trajectories, depicted in Fig.~\ref{fig:Example_1}, can be obtained via solving the optimization problem~\eqref{eqn:opt_problem_cvx} for given sets $\Sigma_{1}^{}=\{4,4,4,1,1,0,0\}$, $\Sigma_{2}^{}=\{4,1,1,2,2,0,0\}$, and $\Sigma_{3}^{}=\{4,1,2,2,2,0,0\}$. Fig.~\ref{fig:Example_1} shows the state trajectories for two different sets. The set $\Sigma_{1}^{}$ leads to the global optimal solution. \end{example} \begin{figure}[!t]\centering \includegraphics[scale=0.9]{paper_bd_16_figure3} \caption{The non-convex set $\mathcal{S}\triangleq\mathbb{R}^{2}\setminus\mathcal{C}_{0}^{}$ can be divided into four convex regions, which are illustrated by different colors. For a given scheduling sequence, i.e., $\mathcal{T} = \{6, 7\}$, the optimal control inputs can be computed by solving the optimization problem~\eqref{eqn:opt_problem_cvx} for all possible combinations of five convex sets. For $N=7$, finding optimal control actions requires solving $15,625$ problems, and selecting the minimum control cost among $2,650$ feasible solutions. In Fig.~\ref{fig:Example_1}, the blue curve represents the state trajectory for $\Sigma_{1}^{}=\{ 4, 4, 4, 1, 1, 0, 0 \}$, the red curve represents the state trajectory for $\Sigma_{2}^{}=\{ 4, 1, 1, 2, 2, 0, 0 \}$, and the magenta curve denotes the state trajectory for $\Sigma_{3}^{}=\{ 4, 1, 2, 2, 3, 0, 0 \}$.} \label{fig:Example_1} \end{figure} Example~1 demonstrates that obtaining the optimal solution to the finite-horizon optimal control problem, mentioned earlier, requires solving an exponential number of QPs. Therefore, in the next section, we propose heuristic algorithms that significantly reduce computational complexity, by compromising the optimality guarantees slightly. \section{Computation of sub-optimal control actions} \label{sec:subopt_control} In this section, we propose two heuristic algorithms that usually achieve a low control loss while significantly decreasing the computation time. We, firstly, develop a simple algorithm that takes the initial state $\x_0$ as input and finds a suitable sequence $\Sigma$ in a greedy manner by increasing the horizon of the problem from $1$ to $N$. Secondly, we apply an algorithm based on ADMM to compute the heuristic scheduling sequence $\Sigma$. This sequence is then used to solve the disjunctive problem \eqref{eqn:opt_problem_cvx} and obtain a good approximate control sequence $\pi$. Both techniques are of polynomial time complexity, hence, reducing the computational efforts significantly. Numerical investigations (see Section~\ref{sec:numerical}) indicate that these heuristic algorithms usually provide close to optimal solutions. % -------------------------------------------------------------------------------------------------------------------------- % % Greedy Algorithm % % -------------------------------------------------------------------------------------------------------------------------- \subsection{Greedy heuristic method} % We construct an intuitive greedy algorithm to solve the optimal control problem, proposed in~\eqref{eqn:opt_problem}, efficiently. As discussed earlier, given a feasible switching sequence $\Sigma$, a corresponding optimal control problem can be formulated as a quadratic programming problem of the form \eqref{eqn:opt_problem_cvx}. One, then, solves the combinatorial problem~\eqref{eqn:opt_sequence_problem} to find the optimal switch sequence $\Sigma^\star$, thereby solving the optimal control problem \eqref{eqn:opt_problem}. A good way to deal with the combinatorial nature of this type of algorithms and reduce the computational complexity is to employ a greedy algorithm, which solves a global problem by making a series of locally optimal decisions. Greedy algorithms do not necessarily provide optimal solutions; however, they might determine local optimal solutions, which can be used to estimate a globally optimal solution, in a reasonable time. Our proposed greedy algorithm works as follows. Since $\x_{0}^{}$ is known, we can compute corresponding switching signal $\sigma(0)$. Then, we set $\Sigma = \{ \sigma(0), p \}$ for all $p \in\{0,1,\dots,2n\}$ and solve $2n+1$ number of QPs of the form~\eqref{eqn:opt_problem_cvx} to obtain the optimal switch signal $\sigma(1)$. The same procedure then repeats for $k=2,\dots, N-1$, where at each step $k$, we solve $2n+1$ times the QP problem~\eqref{eqn:opt_problem_cvx} to find the best $\sigma(k)$ keeping the previously found scheduling sequence $\Sigma = \{ \sigma(0), \dots, \sigma(k-1) \}$ unchanged. Algorithm~\ref{alg:greedy_heuristic} provides the detailed steps of discussed greedy heuristic method. Algorithm~\ref{alg:greedy_heuristic} executes $(2n+1)(N-1)$ number of QPs compared to $(2n+1)_{}^{N-1}$ as in optimal control procedure. As a result, it runs in polynomial time. %%%%%%%%%%%%%%% %%%%%%%%%%%%%%%% \begin{algorithm}[t]\label{alg:greedy_heuristic} %\begin{algorithmic} \SetAlgoLined \KwData{ $\x_{0}$, $n$, $N$\;} Compute $\sigma(0)$ based on $\x_{0}$\; \For{Iteration count $k=2, \ldots, N$}{ Obtain $\tilde{\x}=\x^\star$ and $\tilde{\bu}=\bu^\star$ by solving \eqref{eqn:opt_problem_cvx} with $\Sigma = \{ \sigma(0), \cdots, \sigma(k-2), p \}$ for all $p \in\{0,1,\cdots,2n\}$\; Compute $p_{}^{\star} = \mbox{argmin}_{\Sigma}^{}J_{\Sigma}^{\star}(\x(0))$\; Set $\sigma(k-1) = p_{}^{\star}$\; } return $\Sigma$, $\tilde{\x}$, and $\tilde{\bu}$\; %\end{algorithmic} \caption{Greedy heuristic} % for solving problem \ref{e-prob}} \end{algorithm} % -------------------------------------------------------------------------------------------------------------------------- % % ADMM % % -------------------------------------------------------------------------------------------------------------------------- \subsection{A heuristic method based on ADMM }\label{sec:ADMM} % In this section, we develop a computationally efficient method for solving~\eqref{eqn:opt_problem}. Our approach is based on, Alternating Direction Method of Multipliers (ADMM), a well developed technique to solve large-scale disciplined problems~\cite{BPC+:11}. The idea of utilizing ADMM as a heuristic to solve non-convex problems has been recently considered in literature~\cite{BPC+:11, TMB+:16, DTB:16, DBE+:13}. In particular, \cite{DTB:16, TMB+:16} employ ADMM to approximately solve problems with convex costs and non-convex constraints. The approximate ADMM solutions are improved by multiplicity of random initial starts and a number of local search methods applied to ADMM solutions. In this paper, using some of the techniques introduced above, we first apply ADMM to original non-convex problem to achieve an intermediate solution. This algorithm, for a general non-convex problem, does not necessarily converge; however, when it does, the corresponding solution gives a good initial guess for finding the optimal solution to this problem. In a second phase of our heuristic method, we perform a polishing technique with the aim of achieving a feasible and hopefully closer to the optimal solution (see \cite{DTB:16} for an overview of polishing techniques). Our polishing idea is based on disjunctive programming problem~\eqref{eqn:opt_problem_cvx} with the fixed event index $\mathcal{T}$ and trajectory sequence $\Sigma$. Initially developed to solve large-scale convex structured problems, ADMM has been recently advocated, to a great extent, for being effective approximation technique to solve non-convex optimization problems (see~\cite{BPC+:11, TMB+:16, DTB:16} and references therein). We start with casting~\eqref{eqn:opt_problem} to the ADMM form. Essentially, we have \begin{equation}\label{eqn:opt_problem_ADMM} \begin{array}{ll} \underset{\z}{\mbox{minimize}}& \dfrac{1}{2}\z^\top F \z + \mathcal{I}(\w)\\ & G \z = \h,\\ & \z = \y, \end{array} \end{equation} % where $\z$ collects the state and control components; i.e., ${\z\triangleq[\x(0)^\top,\dots, \x(N)^\top,\bu(0)^\top, \dots, \bu(N-1)^\top]^\top}$, $F$ represents the positive definite Hessian of the quadratic cost. That is $F\triangleq \mbox{blkdiag}(Q_s,R_s)$ with $Q_s\triangleq I_{N+1}\otimes Q$ and $R_s\triangleq I_N \otimes R$. Moreover, $\mathcal{I}(\y)$ is the indicator function enforcing the event-triggering control law; that is \begin{equation} \label{eqn:opt_ADMM_indicator} \mathcal{I}(\y)=\left\{ \begin{array}{ll} \infty & \mbox{if}\; \eqref{eqn:opt_ADMM_violation} \;\mbox{holds},\\ 0& \mbox{otherwise}, \end{array} \right . \end{equation} % \begin{align} \label{eqn:opt_ADMM_violation} \{\exists t\in \{0, \dots, N-1\} \vert \y(t+1)\in \mathcal{C}_0\; \mbox{and} \;\y(t+N+2)\neq 0_m\}, \end{align} % where \eqref{eqn:opt_ADMM_violation} detects the constraint violation in~\eqref{eqn:opt_problem} in terms of having $\x(t)\in \mathcal{C}_0$ while corresponding $\bu(t)\neq \mathbf{0}_m$. The constraints in~\eqref{eqn:opt_problem_ADMM} are twofold. The first constraint in~\eqref{eqn:opt_problem_ADMM} describes the LTI system dynamics~\eqref{eqn:Discrete_time_system}. Here $G\in \mathbb{R}^{nN \times \left(n(N+1)+mN\right)}$ have the following $i=\{1,\dots,N\},\; j=\{1,\dots 2N+1\}$- blocks, \begin{equation}\nonumber G_{ij} = \left\{ \begin{array}{ll} A^{i-1}& \;\mbox{for } i=1\dots N, j=1,\\ -I_n & \;j=i \;\mbox{and}\; j>1,\\ A^{i+N-j}B & \; \mbox{if}\; j\geq N+2\; \mbox{and}\; i+N\geq j,\\ 0 & \; \mbox{otherwise}. \end{array} \right. \end{equation} and $\h\in\mathbb{R}^{n(N+1)+mN}\triangleq[\x(0)^\top, \mathbf{0}_n^\top,\dots, \mathbf{0}_m^\top]^\top$. Finally, the last constraint in~\eqref{eqn:opt_problem_ADMM} is of consensus-type to ensure that non-convex constraint in~\eqref{eqn:opt_problem} is asymptotically satisfied. After formulating the problem, each iteration of ADMM algorithm consists of~(see \cite{BPC+:11} for a complete treatment of the subject): \begin{equation}\label{eqn:ADMM_algorithm} \begin{aligned} \z^{k+1/2} &= \underset{\z}{\mbox{argmin}} \left( \dfrac{1}{2}\z^\top F \z\right. \\ &\left. +\dfrac{\rho}{2}\left\Vert \left[ \begin{array}{l} G \\I\end{array}\right] \z -\left[ \begin{array}{l} 0 \\I\end{array}\right] \z^k-\left[ \begin{array}{l} \h \\0\end{array}\right]+ \bu^k\right\Vert^2 \right) \\ \z^{k+1} &= \Pi (\z^{k+1/2}+\left[ \begin{array}{ll} 0& I \end{array}\right] \bu^k)\\ \bu^{k+1}&= \bu^k + \left[\begin{array}{l} G\\ I\end{array}\right] \z^{k+1/2}-\left[ \begin{array}{l} 0 \\I\end{array}\right] \z^k-\left[ \begin{array}{l} \h \\ 0\end{array}\right]. \end{aligned} \end{equation} Here, $k$ denotes the iteration counter, $\rho>0$ is the step-size (penalty) parameter, and $\Pi$ denotes the projection onto non-convex constraint and is given by \begin{equation} \Pi_i(\x) = \left\{\begin{array}{ll} 0& \mbox{if}\; i\geq N+2 \mbox{ and } \x(i-N-1)\in \mathcal{C}_{0}^{}\\ \x(i)& \mbox{otherwise}. \end{array}\right. \end{equation} When the cost function in hand is convex and the constraints set is closed and convex then the ADMM algorithm converges to the optimal solution of the problem. In our case, since $\mathcal{I}(\y)$ is non-convex the ADMM algorithm~\eqref{eqn:ADMM_algorithm} may not converge to optimal point. Moreover, examples can be found in which~\eqref{eqn:ADMM_algorithm} does not even converge into a feasible point. To further improve the ADMM solution, we utilize an additional convex problem to polish the intermediate results. That is, we restrict the search space to a convex set that includes the ADMM solution pattern. Let ${\tilde{\z}\triangleq[\tilde{\x}(0)^\top,\dots, \tilde{\x}(N)^\top,\tilde{\bu}(0)^\top, \dots, \tilde{\bu}(N-1)^\top]^\top}$ be the output of the ADMM algorithm and the sets $\widetilde{\Sigma}$ and {${\widetilde{\mathcal{T}}\triangleq\{t\in\mathbb{N}_{0}^{}; t<N : \tilde{\x}(t)\in\mathcal{C}_{0}^{}\}}$} denote the trajectory sequence associated with {$[\tilde{\x}_1^\top, \dots, \tilde{\x}_N^\top]^\top$} and the restriction index of $\tilde{\x}$ into $\mathcal{C}_0$, respectively. Then our heuristic polishing procedure formulates~\eqref{eqn:opt_problem_cvx} with $\mathcal{T}=\mathcal{\widetilde{T}}$ and $\Sigma=\widetilde{\Sigma}$. Algorithm~\ref{alg:ADMM_heuristic} provides a summary of our heuristic method. The input variables include the error-tolerance threshold $\epsilon^{tol}$, randomized initial state $\z^0$ as well as $\z_{best}$ and $f_{best}$ to store the final solution. % \iffalse the following problem \begin{equation} \label{eqn:ADMM_disjunctive} \begin{array}{ll} \underset{z}{\mbox{minimize}}& \dfrac{1}{2}z^\top P z, \\ &\left[ \begin{array}{l} F\\G \end{array}\right] z = \left[ \begin{array}{l} b\\0_{\vert \widetilde{\mathcal{T}}\vert m} \end{array}\right], \end{array} \end{equation} where $G\in\{0,1\}^{\vert\widetilde{\mathcal{T}}\vert m \times \left(n(N+1)+mN\right)} $ and constructed such that for $t\in \widetilde{\mathcal{T}}$ it has the following block structure \begin{equation}\nonumber [0_{m\times (n(N+1)+(t-1)m)}, I_m, 0_{m\times (N-t)m }]. \end{equation} \fi % A few comments related to Algorithm~\ref{alg:ADMM_heuristic} are in order. \begin{enumerate} \item \EG{In contrast to the ADMM algorithms for convex problems that converge to the optimum for all positive range of the step-size parameter $\rho$, the stability of the ADMM algorithm for non-convex problems is sensitive to the choice of $\rho$. Moreover, similar to the convex case, the convergence speed is also affected by the choice of $\rho$.} A variety of techniques including proper step-size selection, over relaxation, constraint matrix pre-conditioning and caching can be employed to accelerate the ADMM procedure~\eqref{eqn:ADMM_algorithm} (see \cite{ GTS+:15,GiB:16} for a reference on the topic). \item One can utilize an adaptive iteration count procedure to improve the probability of finding feasible solution to~\eqref{eqn:opt_problem}. The procedure works as the following. Run the ADMM algorithm for a fixed number of iterations and then check if the disjunctive QP problem has feasible solution. If not, then increase the number of iterations (e.g., double it) and repeat the procedure (run ADMM and disjunctive QP) until finding a feasible solution or reaching to a point where the current ADMM solution does not improve the previous one in terms of constraint violations and quadratic cost. \EG{\item The variable $z^0$ is initialized randomly using a normal distribution. In general, repeating the ADMM algorithm for a multiple of initial points may improve the approximate solution at the cost of extra computational complexity (as suggested in e.g.,~\cite{TMB+:16}). However, in our application we did not experience significant performance improvement by repeating the ADMM algorithm (steps $1-6$ in Algorithm~\ref{alg:ADMM_heuristic}) with different initializations. One reason could be that in Algorithm \ref{alg:ADMM_heuristic}, we employ the solution trajectories $\widetilde{\Sigma}$ and $\widetilde{\mathcal{T}}$ based on the ADMM solution $\tilde{z}$ and not $\tilde{z}$ itself in the disjunctive problem~\eqref{eqn:opt_problem_cvx}.} \item Using caching and $LDL^\top$ matrix factorization techniques in the first-step of~\eqref{eqn:ADMM_algorithm} \cite{BPC+:11}, each $\z$-update in Algorithm~\ref{alg:ADMM_heuristic} costs on the order of $O\left((n(N+1)+mN)^2\right)$ for dense $P$ matrices. Furthermore, if we assume $N\geq \max\{n,m\}$ then each ADMM iteration costs on the order of $O(N^2)$. This means that overall cost of the ADMM algorithm is of $O(K N^2)$ where $K$ is the number of ADMM iterations. In our application, with proper algorithm parameter selection, the ADMM algorithm converges almost always within a few hundreds of iterations (see the results presented in Section~\ref{sec:numerical}). \item The computational complexity of the disjunctive step in our heuristic method falls into the generic complexity of convex QPs on the order of $O(\left(\vert \mathcal{T} \vert 2n +(N-\vert \mathcal{T}\vert)(2n-1)\right)^2)$ \cite{BoV:04}. \end{enumerate} %%%%%%%%%%%%%%% %%%%%%%%%%%%%%%% \begin{algorithm}[t]\label{alg:ADMM_heuristic} \SetAlgoLined \KwData{ $\epsilon^{tol}$, $\z^0\sim\mathcal N(0,\sigma^2I)$, $\z_{best}=\emptyset$, and $f_{best}=\infty$\;} \For{Iteration count $k=1, 2, \ldots, K$}{ update $\z$ from~\eqref{eqn:ADMM_algorithm}\;\If{$\Vert G\z^k-\h\Vert\leq \epsilon^{tol}$ \mbox{and} $(1/2)\z^{k\top} F \z^k < f_{best}$}{ $\z_{best}= \z^k$\;} } \If{$f_{best} < \infty$}{Obtain $\z_{best}$ from disjunctive problem~\eqref{eqn:opt_problem_cvx}\;} return $\z_{best}$\;\caption{ADMM-Disjunctive heuristic} \end{algorithm} % -------------------------------------------------------------------------------------------------------------------------- % % Receding Horizon Control % % -------------------------------------------------------------------------------------------------------------------------- \section{Receding horizon control} Except a few special cases, finding a solution to an infinite horizon optimal control problem is not possible. Receding horizon control is an alternative scheme to infinite horizon problem that repeatedly requires solving a constrained optimization problem. At each time instant $t\geq 0$, starting at the current state $\x(t)$ (assuming that a full measurement of the state $\x(t)$ is available at the current time $t\in\mathbb{N}_{0}^{}$), the following cost function % \begin{multline*} J(\x(t),\boldsymbol{\pi}_{t}^{\star N-1}) = V_{f}^{}(\x(t+N\mid t)) \\ + \sum_{k=t}^{t+N-1}\ell(\x(k\mid t),\bu(k\mid t)) \end{multline*} % subject to system dynamics and constraints involving states and controls is minimized over a finite horizon $N$. Here, the function $\ell$ defines the stage cost, and $V_{f}^{}$ defines the terminal cost. Denote the minimizing control sequence, which is a function of the current state $\x(t)$, by % \begin{align*} \boldsymbol{\pi}_{t}^{\star N-1}=\big[\bu^{\star}(t\mid t)^{\intercal}, \cdots, \bu^{\star}(t+N-1\mid t)^{\intercal} \big]\;,\end{align*} % then the control applied to the plant at time $t\geq 0$ is the first element of this sequence, that is, % \begin{align*} \bu(t) = \big[I ~ 0 ~ \cdots ~ 0 \big]\boldsymbol{\pi}_{t}^{\star N-1} \;.\label{eqn:receding_horizon_strategy} \end{align*} % Time is then stepped forward one instant, and the procedure described above is repeated for another $N$-step ahead optimization horizon. \begin{figure}[!t]\centering \includegraphics[scale=0.9]{paper_bd_16_figure4} \caption{Evolution of the state trajectories of the system~\eqref{eqn:example_system} under the receding horizon strategy~\eqref{eqn:receding_horizon_strategy} and constraints~\eqref{eqn:event_triggered_set}--\eqref{eqn:event_triggered_constraint_time} for a set of initial conditions $\x_{0}^{}$.} \label{fig:RHC_example} \end{figure} It is worth noting that, due to the fundamental limitation of the optimization problem~\eqref{eqn:opt_problem}, the state $\x(t)$ never converges to the origin if the system is unstable. It can only converge into a set including the origin. The following result shows the existence of this set: \begin{theorem}\label{thm:Stability} % Suppose that $\mathcal{D}_{\mu}^{} \triangleq \{ \x\in\mathbb{R}^{n} \,:\, \parallel \x \parallel_{\infty}^{} \leq \mu \}$ is a neighborhood of the origin, where % \begin{align} \mu &\triangleq \sqrt{\frac{\kappa\lambda_{\max}(A^{\intercal}PA+Q) n \varepsilon^{2}}{\lambda_{\min}^{2}(Q)}} \end{align} % for some $\kappa > 0$. Then, the system~\eqref{eqn:Discrete_time_system} with event-triggering constraints~\eqref{eqn:event_triggered_constraint} and \eqref{eqn:event_triggered_constraint_time} is uniformly practically asymptotically stable, i.e., $\displaystyle{\lim_{t\rightarrow\infty} \parallel \x(t) \parallel_{\infty}^{} \leq \mu}$. % \end{theorem} Theorem~\ref{thm:Stability} establishes practical stability of the system~\eqref{eqn:Discrete_time_system} with event-triggering constraints \eqref{eqn:event_triggered_constraint} and~\eqref{eqn:event_triggered_constraint_time}. It shows that if provided conditions are met, then the plant state will be ultimately bounded in an $\ell_{\infty}^{}$-norm ball of radius $\mu$. It is worth noting that, as the other stability results, which use Lyapunov techniques, this bound will not be tight. Next, we would like to exemplify the convergence of the state $\x(t)$ to the set $\mathcal{D}_{\mu}^{}$ from a set of initial conditions $\x_{0}^{}$. \begin{example} Consider the second-order system, given by~\eqref{eqn:example_system}, with initial conditions $\x_{0}^{}=1.2\left[\begin{smallmatrix} \text{sin}(\frac{\pi k}{6}) ~ \text{cos}(\frac{\pi k}{6}) \end{smallmatrix}\right]^{\intercal}$ for all $0 \leq k \leq 12$. The controller transmits a control message whenever the event-triggering condition, i.e., $\parallel\x(t)\parallel_{\infty}^{} > 0.25$, holds. The performance indices are chosen as $Q=\mbox{diag}\{2,2\}$ and $R=5$ while the prediction horizon is set to $N=6$. Fig.~\ref{fig:RHC_example} shows the state trajectories of the receding horizon implementation of the event-triggered control system, described by \eqref{eqn:Discrete_time_system} -- \eqref{eqn:event_triggered_constraint_time}, for different initial conditions $\x_{0}^{}$. As can be seen in the same figure, the number of transmitted control commands depends on the initial condition. \end{example} %==================================================================================================== % % NUMERICAL EXAMPLES % %==================================================================================================== \section{Numerical examples}\label{sec:numerical} \subsection{Performance of sub-optimal event-triggered control actions} % To illustrate performance benefits of the proposed heuristic algorithms, we consider a third-order plant with the following state-space representation: \BD{ % \begin{multline} \x(t+1) = \begin{bmatrix} 0.53 & -2.17 & 0.62 \\ 0.22 & -0.06 & 0.51 \\ -0.92 & -1.01 & 1.69 \end{bmatrix} \x(t) + \begin{bmatrix} 0.4 \\ 0.7 \\ 0.9 \end{bmatrix}\bu(t) \\ + B_{\w}^{}\w(t) \;.\label{eqn:example_3d} \end{multline} % } % Assume that there is no disturbance acting on the plant, i.e., $B_{\w}^{} = 0$. Our aim, here, is to minimize~\eqref{eqn:cost_function} with performance indices $Q_{}^{} = P_{}^{} = \mathrm{diag}\{2,2,2\}$ and $R_{}^{} = 5$, and the horizon length $N = 8$. The initial condition is chosen as % \begin{align*} \x_{0} = \begin{bmatrix} \sin\theta\cos\phi & \sin\theta\sin\phi & \cos\theta \end{bmatrix}^{\intercal} \;,\end{align*} % where $0 \leq \theta \leq \pi$ and $0 \leq \phi \leq \pi$. To perform simulations, we selected $577$ different data points, which are equidistant over a half spherical surface. We evaluated the performance of the greedy and the ADMM-based heuristic algorithms for all these initial conditions. For each data point on the half sphere, we run $823,543$ number of QPs to obtain the global optimal solution. In total, we run more than $475$ million number of QPs in our evaluation. We used a linux machine with $32$ cores in Intel Core $i7$ Extreme Edition $980X$ to solve these QPs in parallel threads. Within this computing resources, it takes approximately $6$ days to perform simulations for one problem setup. To compare the (true) optimal control performance with our heuristic method, we performed Algorithm~\ref{alg:ADMM_heuristic} on the same problem set. In particular, we evaluated three sets of problem with $\varepsilon = 0.2$, $\varepsilon=0.4$, and ${\varepsilon=0.6}$. It is generally expected that the control problem becomes more challenging with increasing the event-threshold $\varepsilon$. We run the ADMM-based heuristic algorithm with adaptive iteration count procedure described in Section \ref{sec:ADMM}. Moreover, we picked $\rho=9.8$, $\rho=5.8$, and $\rho=6.9$ respectively, for the aforementioned problem scenarios. Under this setup, in both cases (i.e., $\varepsilon = 0.2$ and $\varepsilon=0.4$), the ADMM method produced always feasible solutions for maximum $300$ number of ADMM-iterations. However, when $\varepsilon = 0.6$ and $\rho=6.9$, the method created infeasible solutions for two initial conditions: $\x_{0}^{}=\left[ 0.3036,\, 0.2330,\, 0.9239 \right]^\top$ and $\x_{0}^{}=\left[0.6830, \, 0.1830, \, 0.7071 \right]^\top$. To compute feasible solutions, we needed to adjust $\rho$ accordingly for these initial conditions. Fig.~\ref{fig:comparison_admm_vs_optimal} shows the outcome of the comparison. As it shows, for $\varepsilon=0.2$, the ADMM-based heuristic algorithm is able to find near optimal solutions (with relative error less than $5\%$) in almost all cases, and for $\varepsilon=0.4$ and $\varepsilon=0.6$, the algorithm finds solutions with relative error less than {$5\%$} in more than $80\%$ of problem instances. The average optimality gap $(J_{\text{admm}}-J^\star)/J^\star$ for the three cases were $0.0035$, $0.0237$, and $0.0716$, respectively. Finally, by comparing the cardinality of control event index $\mathcal{T}$ for the optimal solutions and our ADMM method, one concludes that the heuristic method in more than $50\%$ of cases finds similar-sparse solution as the optimum ones, and in the rest of cases it tends to find more sparse control actions. We also compared the true optimal control performance with a greedy heuristic method, shown in Algorithm~\ref{alg:greedy_heuristic}, for three sets of problems, mentioned above, with $\varepsilon = 0.2$, $\varepsilon=0.4$, and ${\varepsilon=0.6}$. The greedy algorithm detected $22$, $90$, and $135$ optimal solutions of the aforementioned $577$ problems, respectively. Also, it computed the optimal solution with relative error less than $5\%$ in more than $65\%$ of problems at hand for all three problem setups. It is important to note that the greedy heuristic method always provided feasible solutions for all problem cases. The average optimality gaps $(J_{\text{greedy}}-J^\star)/J^\star$ in all three cases were $0.0514$, $0.0780$, and $0.0826$, respectively. As compared to the ADMM-based heuristic method, the optimality gaps became larger. % \begin{figure}[!t]\centering \includegraphics[scale=0.55]{paper_bd_16_figure5} \caption{Relative error $(J_{\mathrm{admm}}^{}-J_{}^{\star})/J_{}^{\star}$ and control-action cardinality difference of the (sub-optimal) ADMM-based heuristic solution from the optimal cost for $577$ problem instances of third order plant model~\eqref{eqn:example_3d}. % } \label{fig:comparison_admm_vs_optimal} \end{figure} % % \begin{figure}[!t]\centering \includegraphics[scale=0.55]{paper_bd_16_figure6} \caption{Relative error $(J_{\mathrm{greedy}}^{}-J_{}^{\star})/J_{}^{\star}$ and control-action cardinality difference of the greedy solution from the optimal cost for $577$ problem instances of third order plant model~\eqref{eqn:example_3d}. % } \label{fig:comparison_greedy_vs_optimal} \end{figure} % % \begin{figure}[!t]\centering \includegraphics[scale=0.6]{paper_bd_16_figure7} \caption{\BD{States and input versus time for the event-triggered model predictive controller presented in Section~\ref{sec:numerical_mpc}. In this figure, $\x_{1}^{}(t)$, $\x_{2}^{}(t)$ and $\x_{3}^{}(t)$ represent state variables whereas $u(t)$ and $u_{\mathrm{admm}}^{}(t)$ denote the optimal and the heuristic control inputs computed based on the exhaustive search and the ADMM algorithm, respectively.} % } \label{fig:receding_horizon_control} \end{figure} % \BD{ \subsection{Receding horizon control implementation}\label{sec:numerical_mpc} % We consider the open-loop unstable discrete-time system represented by~\eqref{eqn:example_3d} with $B_{\w}^{} = 0$. The prediction horizon is chosen to be $N = 6$ with performance indices $Q_{}^{} = P_{}^{} = \mathrm{diag}\{2,2,2\}$ and $R_{}^{} = 5$. The event-threshold $\varepsilon$ for the event-triggered receding horizon implementation is set to $0.4$. Fig.~\ref{fig:receding_horizon_control} demonstrates the time responses of the event-triggered receding horizon control system for the initial condition $\x_{0}^{}=\left[0, \, \frac{\sqrt{2}}{2}, \, -\frac{\sqrt{2}}{2} \right]^\top$. Here, two different approaches for computing control input trajectories are considered: the exhaustive search algorithm and the heuristic ADMM algorithm with $\rho=4.8$. As can be seen in Fig.~\ref{fig:receding_horizon_control}, while the state trajectories associated to two control approaches deviate from one another (except for the few first sampling instants that they match together), the corresponding control inputs follow a rather similar sparsity pattern. The event-triggered receding horizon control via exhaustive search algorithm uses the communication channel $14$ times over the $50$ sampling instants, which results in a reduction of $72\%$ in communication and computational burden, whereas the event-triggered receding horizon control via ADMM algorithm uses the channel $16$ times over the $50$ sampling instants, which leads to a reduction of $68\%$ in communication and computational burden. For the aforementioned initial value, the exhaustive search achieves a better control cost: $65.42$ compared to $77.72$ as of the control cost of the ADMM-based heuristic. However, this conclusion does not hold in general. In particular, we experienced cases where starting from a different initial value, the receding horizon controller derived by exhaustive search performs worse than the one based on ADMM-based heuristic. } \subsection{Trade-off between communication rate and control performance} % We consider the case where a random disturbance $\w(t)\in\mathbb{R}_{}^{n}$, which is modeled by a discrete-time zero-mean Gaussian white process with co-variance $\Sigma_{\w}^{}$, is acting on the plant characterized by the injection matrix $B_{\w}^{} = \mbox{diag}\{1,1,1\}$. The initial condition $\x_{0}^{}$ is modeled as a random variable having a normal distribution with zero mean and co-variance $\Sigma_{\x_{0}^{}}^{}$. The process noise $\w(t)$ is independent of the initial condition $\x_{0}^{}$. The control performance is measured by a quadratic function: % \begin{align} J_{\infty}^{} = \frac{1}{500}\sum_{t=0}^{499} \big(\x_{}^{\intercal}(t)Q\x_{}^{}(t) + \bu_{}^{\intercal}(t)R\bu(t)\big)\;, \label{eqn:empirical_cost} \end{align} % while the communication rate is measured by % \begin{align} \pi_{\infty}^{} = \frac{1}{500}\sum_{t=0}^{499} \mathbbm{1}_{\{\parallel\x(t)\parallel_{\infty}^{}\geq 0\}}^{} \;.\label{eqn:empirical_rate} \end{align} The visualization of the trade-off between the control loss and the communication rate is demonstrated in Fig.~\ref{fig:tradeoff}. Different data points on the curve correspond to different event-threshold values ranging from $0.5$ to $4.0$, and the control loss~\eqref{eqn:empirical_cost} and the communication rate~\eqref{eqn:empirical_rate} are evaluated via Monte Carlo simulations. Note that the communication rate decreases dramatically with an increased control loss as the threshold $\varepsilon$ varies between $0$ and $2.25$ (dark colors). For $\varepsilon > 2.25$ (lighter color), both quantities become less sensitive to changes in the threshold value. We can also identify $0\leq\varepsilon\leq 1.5$ as a particularly attractive region where a large decrease in the communication rate can be obtained for a small loss in control performance. % \begin{figure}[!t]\centering \includegraphics[scale=0.7]{paper_bd_16_figure8} \caption{A comparison of the control performance and the communication frequency for different event-threshold values $\varepsilon > 0$ (shown in gray scale). The marked curve with is obtained by averaging $2,500$ Monte Carlo simulations for the horizon length $500$ samples with the process noise $\{\w(t)\}_{t\in\mathbb{N}_{0}^{}}$and the initial condition $\x_{0}^{}$ generated randomly.} \label{fig:tradeoff} \end{figure} % %==================================================================================================== % % CONCLUSION % %==================================================================================================== \section{Conclusion} In this paper, we considered a feedback control system where an event-triggering rule dictates the communication between the controller and the actuator. We investigated the optimal control problem with an additional constraint on the event-triggered communication. While this optimization problem, in general, is non-convex, we used a disjunctive programming formulation to obtain the optimal solution via solving an exponential number of quadratic programs. Then, we proposed a heuristic algorithms to reduce the computational efforts while achieving an acceptable performance. Later, we provided a complete stability analysis of receding horizon control that uses a finite-horizon optimization in the proposed class. Our numerical study confirmed the theory. %==================================================================================================== % % APPENDIX % %==================================================================================================== \section{Appendix} We begin with the definition of the \textit{practical stability} of control systems. \begin{definition}[practical stability~\cite{JiW:02}] The system~\eqref{eqn:Discrete_time_system} is said to be uniformly practically asymptotically stable in $\mathcal{A}\subseteq\mathbb{R}^{n}$ if $\mathcal{A}$ is a positively invariant set for~\eqref{eqn:Discrete_time_system} and if there exist a $\mathcal{KL}$-function $\beta$, and a nonnegative constant $\delta\geq 0$ such that % \begin{align*} \parallel \x(t) \parallel \leq \beta(\parallel \x_{0} \parallel, t) + \delta \;.\end{align*} % \end{definition} %\EG{@Burak: I guess it is good to cite above definition, besides what is $\beta (\cdot,t)$?} \begin{definition}[practical-Lyapunov function~\cite{JiW:02}] A function $V:\mathbb{R}^{n}\rightarrow\mathbb{R}_{\geq 0}$ is said to be a practical-Lyapunov function in $\mathcal{A}$ for the system~(1) if $\mathcal{A}$ is a positively invariant set and if there exist a compact set, $\Omega\subseteq\mathcal{A}$, neighbourhood of the origin, some $\mathcal{K}_{\infty}$-functions $\alpha_{1}, \alpha_{2}$ and $\alpha_{3}$, and some constants $d_{1}, d_{2}\geq 0$, such that % \begin{align} V( \x ) &\geq \alpha_{1}(\parallel \x \parallel), \forall\x\in\mathcal{A} \;,\\ V( \x ) &\leq \alpha_{2}(\parallel \x \parallel) + d_{1}, \forall\x\in\Omega \;,\\ V( f(\x) ) - V( \x ) &\leq -\alpha_{3}(\parallel \x \parallel) + d_{2}, \forall\x\in\mathcal{A} \;.\end{align} % \end{definition} % \EG{@Burak: what is $f(x)$ in above definition?} % \EG{@Burak:I could not find the exact definition in [34], would you please refer to the exact place within the reference. Like [34, Definition xxx]} %~\cite{AgQ:13} \noindent\textbf{Proof of Theorem~\ref{thm:Stability}.} % To prove the practical stability of the system~\eqref{eqn:Discrete_time_system} with a set of control actions $\boldsymbol{\pi}_{}^{\star} = \{ \bu^{\star}(t\mid t)^{}, \cdots, \bu^{\star}(t+N-1\mid t)^{} \}$, which fulfills constraints~\eqref{eqn:event_triggered_constraint} and~\eqref{eqn:event_triggered_constraint_time}, we will analyze the value function, % \begin{align*} V_{N}^{\star}(\x(t)) \triangleq \min_{\boldsymbol{\pi}}^{}J(\x(t),\boldsymbol{\pi}) , \end{align*} % where $J(\x(t),\boldsymbol{\pi})$ is as seen in~\eqref{eqn:cost_function}. We borrow the shifted sequence approach, described in~\cite{RaM:09}, and we use a feasible control sequence, i.e., $\tilde{\boldsymbol{\pi}} = \{\bu_{}^{\star}(t+1 \mid t), \cdots, \bu_{}^{\star}(t+N-1 \mid t), \hat{\bu} \}$. By the optimality property, we obtain the following bound (with $\x_{}^{}(t \mid t) = \x(t)$ and $\bu_{}^{\star}(t \mid t) = \bu(t)$): \begin{align} V_{N}^{\star}(\x(t+1)) & - V_{N}^{\star}(\x(t)) \leq V_{N}^{}(\x(t+1),\tilde{\boldsymbol{\pi}}) - V_{N}^{\star}(\x(t)) \nonumber \\ & = - \ell(\x(t),\bu(t)) + V_{f}\big(A\x(t+N \mid t)+B\hat{\bu}\big)\nonumber\\ & - V_{f}(\x(t+N \mid t)) + \ell(\x(t+N \mid t),\hat{\bu}) \label{eqn:optimality_bound} \end{align} % for all $\x(t)\in\mathbb{R}^{n}$. % Thus,~\eqref{eqn:optimality_bound} holds if the following one is satisfied: % We now investigate the quantity of \begin{align*} \Delta V_{f}^{}(t) + \ell(\x(t+N \mid t),\hat{\bu}) \;,\end{align*} % where $\Delta V_{f}^{}(t) \triangleq V_{f}\big(A\x(t+N \mid t)+B\hat{\bu}\big) - V_{f}(\x(t+N \mid t))$. Based on $\x(t+N \mid t)$, there are two possibilities: % \begin{itemize} \item Suppose that $\parallel \x(t+N \mid t) \parallel_{\infty}^{} > \varepsilon$, then for any feasible control law, i.e., $\hat{\bu}=K\x(t+N \mid t)$, the following inequality holds: \begin{align}\label{eqn:deltaV_case_1} & \Delta V_{f}^{}(t) + \ell(\x(t+N \mid t),\hat{\bu})\nonumber \\ & = \x(t+N \mid t)_{}^{\intercal} \big(A_{K}^{\intercal}PA_{K}^{} - P + Q_{}^{*} \big)\x(t+N \mid t)_{}^{} < 0 \;,\end{align} % where $A_{K}^{} = A + BK$ (i.e., Schur) and $Q_{}^{*} = Q + K_{}^{\intercal}RK$. From~\eqref{eqn:optimality_bound} and \eqref{eqn:deltaV_case_1}, it follows that \begin{equation}\nonumber V_{N}^{\star}(\x(t+1)) - V_{N}^{\star}(\x(t)) < 0 \;.\end{equation} \item Suppose that $\parallel \x(t+N \mid t) \parallel_{\infty}^{} \leq \varepsilon$, then $\hat{\bu} = 0$. Hence, it follows that: \begin{align*} \Delta V_{f}^{}(t) + & \ell(\x(t+N \mid t),\hat{\bu}) \\ & = \x(t+N \mid t)_{}^{\intercal} \big(A^{\intercal}PA - P + Q \big)\x(t+N \mid t)_{}^{} \\ & \leq \x(t+N \mid t)_{}^{\intercal} \big(A^{\intercal}PA + Q \big)\x(t+N \mid t)_{}^{} \\ & \leq \lambda_{\max}^{}\big(A^{\intercal}PA + Q \big)\parallel \x(t+N \mid t)_{}^{} \parallel_{2}^{2} \\ & \leq \lambda_{\max}^{}\big(A^{\intercal}PA + Q \big)n \parallel \x(t+N \mid t)_{}^{} \parallel_{\infty}^{2} \\ & \leq \lambda_{\max}^{}\big(A^{\intercal}PA + Q \big)n \varepsilon{}^{2} \;,\end{align*} % where we used the norm inequality $\parallel \v \parallel_{2}^{} \leq \sqrt{n} \parallel \v \parallel_{\infty}^{}$ for any $\v \in \mathbb{R}_{}^{n}$; see~\cite{Ber:09}. \end{itemize} % As a result, we conclude that, for any $\x(t+N \mid t)_{}^{} \in \mathbb{R}^{n}$, the following inequality holds: % \begin{align} % V_{f}\big(A\x(t+N \mid t)+B\tilde{\bu}\big) - V_{f}(\x(t+N \mid t)) \\ \Delta V_{f}(t) + \ell(\x(t+N \mid t),\hat{\bu}) \leq \eta\;,\label{eqn:Bound_on_terminal_cost} \end{align} % where $\eta = \lambda_{\max}^{}\big( A^{\intercal}PA + Q \big) n \varepsilon{}^{2}$. % Inserting~\eqref{eqn:Bound_on_terminal_cost} and $a_{1}^{}\parallel \x(t) \parallel_{2}^{2} \leq \ell(\x(t),\bu(t))$ where $a_{1}^{}\triangleq\lambda_{\min}(Q)$ into~\eqref{eqn:optimality_bound} yields: % \begin{align}\label{eqn:Lyapunov_ineq2} V_{N}^{\star}(\x(t+1)) & - V_{N}^{\star}(\x(t)) \leq -\lambda_{\min}(Q) \parallel \x(t) \parallel_{2}^{2} + \eta \;.\end{align} % Now, let the binary variable $\delta(t)\in\{0,1\}$, for each time instant $t$, represent the transmission of the control action as follows: % \begin{align} \delta_{}^{}(t) = \begin{cases} 0 & \text{if} ~ \parallel \x(t) \parallel_{\infty}^{} \leq \varepsilon \;,\\ 1 & \text{otherwise} \;.\end{cases} \end{align} % For any feasible control sequence, i.e., $\boldsymbol{\pi}_{}^{\prime} = \{ K \x(t), \cdots, K \x(t+N-1 \mid t) \}$, there exists an associated scheduling sequence, denoted by $\Delta_{}^{}(t) = \{\delta_{}^{}(t), \cdots, \delta_{}^{}(t+N-1)\}$. Applying these feasible control inputs, the upper bound of $V_{N}^{\star}(\x)$ becomes: % \begin{align}\label{eqn:V_upperbound} V_{N}^{\star}(\x(t)) &\leq \x^{\intercal}(t) S_{\Delta_{}^{}(t)}^{} \x(t) \leq \lambda_{\max}(S_{\Delta_{}^{}(t)}^{})\parallel \x(t) \parallel_{2}^{2} \;,\nonumber \\ & \leq a_{3}^{} \parallel \x(t) \parallel_{2}^{2} \;,\end{align} % where $a_{3}^{} \triangleq \displaystyle{\max_{\Delta\in\mathcal{S}}^{}\lambda_{\max}(S_{\Delta_{}^{}}^{})}$ and % \begin{align*} S_{\Delta_{}^{}}^{} = Q_{\delta(0)}^{} + \sum_{i=1}^{N-1}\prod_{j=0}^{i-1} A_{\delta(j)}^{\intercal}Q_{\delta(j)}^{}A_{\delta(j)}^{} + \prod_{k=0}^{N-1}A_{\delta(k)}^{\intercal}QA_{\delta(k)}^{} \;,\end{align*} % with $Q_{\delta(i)}^{} \triangleq \delta(i)Q_{}^{*} + (1-\delta(i))Q_{}^{}$ and $A_{\delta(i)}^{} \triangleq \delta(i)A_{K}^{} + (1-\delta(i))A_{}^{}$ for all $i\in\{0, \cdots, N-1\}$. Here, $\mathcal{S}$ denotes a set of all possible feasible schedules generated by a stabilizing control sequence $\boldsymbol{\pi}_{}^{\prime}$. To obtain the lower bound, we have $\x_{}^{\intercal}(t) Q \x(t) \leq V_{N}^{\star}(\x(t))$ and hence $a_{2}^{} \parallel \x(t) \parallel_{2}^{2} \leq V_{N}^{\star}(\x(t))$ where $a_{2}^{}\triangleq\lambda_{\min}^{}(Q)$. From, \eqref{eqn:Lyapunov_ineq2} and \eqref{eqn:V_upperbound} we establish the following relation: % \begin{align} V_{N}^{\star}(\x(t+1)) \leq \bigg(1 - \frac{a_{2}^{}}{a_{3}^{}} \bigg)V_{N}^{\star}(\x(t)) + \eta \;,\label{eqn:Lyapunov_difference} \end{align} % which implies that % \begin{align} \parallel \x(t+1) \parallel_{2}^{2} \leq \gamma\frac{a_{3}^{}}{a_{2}^{}} \parallel \x(t) \parallel_{2}^{2} + \frac{\eta}{a_{2}^{}} \;.\label{eqn:State_difference} \end{align} % Since $0 < a_{2}^{} \leq a_{3}^{}$, it follows that $\gamma = 1 - \frac{a_{2}^{}}{a_{3}^{}}$, i.e., $0 < \gamma \leq 1$. % Therefore, by iterating~\eqref{eqn:State_difference}, it is possible to exponentially bound the state evolution via: % \begin{align*} \parallel \x(t) \parallel_{2}^{2} \leq \gamma_{}^{t} \frac{a_{3}^{}}{a_{2}^{}} \parallel \x(0) \parallel_{2}^{2} + \frac{1 - \gamma^{t}}{1 - \gamma}\frac{\eta}{a_{2}^{}} \;.\end{align*} % Using the norm inequality $\parallel \v \parallel_{\infty}^{} \leq \parallel \v \parallel_{2}^{} \leq \sqrt{n} \parallel \v \parallel_{\infty}^{}$ for any $\v \in \mathbb{R}_{}^{n}$ (see~\cite{Ber:09}), we get: % \begin{align*} \parallel \x(t) \parallel_{\infty}^{2} \leq \gamma_{}^{t} n \frac{a_{3}^{}}{a_{2}^{}} \parallel \x(0) \parallel_{\infty}^{2} + \frac{1 - \gamma^{t}}{1 - \gamma}\frac{\eta}{a_{2}^{}} \;.\end{align*} % Thus, $\displaystyle{\lim_{t\rightarrow\infty} \parallel \x(t) \parallel_{\infty}^{} \leq \mu}$. $\hfill\blacksquare$ %------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- % Bibliografia %------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- \bibliographystyle{IEEEtran} \bibliography{acc2013} \end{document}}}
\caption{Dependence of order parameters $q_2$ and $q_4$ on radius of averaging circular region $\xi/L$, scaled with particle length $L$. Left panels (a), (c), (e), (g) and (i) refer to $q_2$, while right panels (b), (d), (f), (h) and (j) depict $q_4$. Aspect ratio $\kappa$ increases from top to bottom, as indicated by labels. Different symbols indicate different densities, in the sequence \tikzcircle[white, fill=red]{3.5pt}, \tikzcircle[white, fill=green]{3.5pt}, \tikzcircle[white, fill=blue]{3.5pt}, \tikzsquare[white, fill=red]{6.5pt},\hspace{0.1cm}\tikzsquare[white, fill=green]{6.5pt}, \hspace{0.0cm}\tikzsquare[white, fill=blue]{6.5pt} and \textcolor{myred}{$\blacktriangle$} for increasing density. For each value of $\kappa$, the sequences in $\varphi$ are: for $\kappa=4$, $0.52$, $0.59$, $0.64$, $0.68$, $0.70$ and $0.75$; for $\kappa=6$, $0.55$, $0.60$, $0.64$, $0.69$, $0.74$, $0.78$ and $0.80$; for $\kappa=8$, $0.65$, $0.67$, $0.70$, $0.74$, $0.78$ and $0.81$; for $\kappa=10$, $0.59$, $0.64$, $0.70$, $0.73$, $0.78$ and $0.82$; and for $\kappa=12$, $0.65$, $0.69$, $0.73$, $0.77$ and $0.81$. }
\caption{\textcolor{vino}{\huge${\bullet}$}$^i$ weights %$\beta_i$ , \textcolor{black}{\tiny$\bullet$} CM weights, \textcolor{ruta}{\LARGE$\bullet$}\textcolor{ruta2}{\LARGE$\bullet$} $\Lscr$}
\caption{The temperature dependence of the width and center frequency of the 350 GHz phonon is shown in panel a. The effects of spin-phonon coupling are clear in the changes of the behavior of the frequency of this mode at the magnetic ordering transition. Panel b. shows the ratio $\sigma_1(\text{T}<30K)/\sigma_1(\text{T}=30K)$, which emphasizes the temperature behavior of the three observed absorptions. We have adjusted this ratio as described in the text to isolate the response of the absorptions from the background. The absorptions are seen to occur around 350 GHz, 600 GHz, and 1.2 THz, as indicated by arrows. The bump at 900 GHz is an artifact of the analysis and not a feature. Panel c. shows the normalized spectral weight for the 520-720 GHz (\protect\includegraphics[scale=0.45]{GreenDot_v3}, multiplied by 2) and .85-1.44 THz (\protect\includegraphics[scale=0.45]{BlueDot_v2}) regions and also the interpolated peaks from panel b. (\protect\includegraphics[scale=0.45]{RedTriangle_v2}: 350 GHz absorption; \protect\includegraphics[scale=0.45]{BlackTriangle_v2}: 1.2 THz absorption; error bars are from Gaussian fits to $\sim\pm$ 100 GHz around each peak). The N\'{e}el temperature is marked by a vertical dashed line to emphasize the significance of the magnetic transition. Panel d. shows selected fixed frequency cuts of panel b. The emergence of the magnon is clearly seen below the N\'{e}el temperature marked as a vertical dash line.}
\caption{The error for the Strang splitting/discontinuous Galerkin ($o=5$) and the exp2 scheme as a function of the computational cost is shown for the KPI equation (left) and the KPII equation (right) using the Schwartzian initial value (the setup from section \textcolor{blue}{\ref{subsec:numerical-KP}} is employed). The computational cost is in arbitrary units and is determined according to the model described in this section.\label{fig:comparison}}
\caption{Comparison of our method with a FCN on the ISPRS Vaihingen benchmark. Building detection is not impeded by shadows anymore and cars are more finely segmented.\\ (white: roads, {\color{blue!80!black} blue}: buildings, {\color{cyan!80!black} cyan}: low vegetation, {\color{green!80!black} green}:~trees, {\color{yellow!80!black} yellow}: cars)}
\caption{(a) Geometric structure of ZSiCNR-4 unit cell shown in the red dashed box is considered for the calculation\textcolor{red}{. }(b) Calculated spin-density of ZSiCNR-4. Blue and yellow colored spin densities denote opposite spin orientations. (c) Computed edge formation energy of ZSiCNRs as a function of width. }
\caption{\textcolor{black}{Total energies per unit cell for non-magnetic (NM) and spin-polarized (SP) ZSiCNRs for widths N$_{z}$= 4 to 6. Last column gives energy difference between both the configurations}\textcolor{green}{. \label{tab:mvsnm}}}
\caption{User study. {\red Red} represents the normal seam-cutting framework wins. {\blue Blue} represents our perception-based seam-cutting framework wins. {\textcolor[rgb]{1,0.76,0}{Yellow}} represents an even.}
\caption{Monte Carlo scheme for Sedov radius (in the log), for N=100000. Note the loss of log-normality when sample grows arbitrarily large, thereby granting additional information on the statistics, such as kurtosis or skewness (clearly present).}{\includegraphics[width=\columnwidth]{PValues.pdf}}
\caption{Mean particle density $\mu_{\log{n}} = -1.16$ and variance $\sigma^2_{\log{n}} = 0.77$. The small variance in these simulated distributions appears to be consistent with the analysis of the SNR data. \label{fig:lowmass}}{% \includegraphics[width=0.4\textwidth]{CroppedHighMassStats.pdf}% }
\caption{\textbf{Scheme of the experimental structure and LLP classifier.} \color{darkgray} \\ Top to bottom: The sentence "Franzy jagt ..." is spelled three times. To spell a single character in one trial, 68 highlighting events occur, with 32 belonging to sequence 1 and 36 belonging to sequence 2. The resulting 68 ERP responses (epochs) are averaged for each sequence, and these averages are exploited to reconstruct the mean target and non-target ERP responses.}
\caption{\textbf{Classification results of LLP applied on artificial data sets generated from an auditory (A) and a visual (B) ERP paradigm.} \\ \color{darkgray} For each artificial data set, the target vs non-target accuracy result for different mixing matrices $\boldsymbol{\Pi}_1$-$\boldsymbol{\Pi}_4$ is shown. In the notation of the mixing matrices, individual sequences are separated by commas. Per sequence, the first entry indicates the target ratio whereas the second one indicates the non-target ratio. NAF = Noise Amplification Factor.}
\caption{\textbf{Grand average (N=13) visual ERP response.}\\ \color{darkgray} Top row: Average responses evoked by visual target (blue) and non-target (green) stimuli in the occipital channel O1 (thick) and the central channel Cz (thin). Prior to averaging, a baseline correction was performed based on data within the interval [-200,0]\,ms. The signed$R^2$ values for channels O1 and Cz over time are provided by two horizontal color bars. Their scale is identical to the scale of plots in the last row. Middle rows: Scalp plots visualising the spatial distribution of mean target and non-target responses within four selected time intervals: [50 120], \mbox{[120 200]}, \mbox{[201 380]} and [381 700]\,ms relative to stimulus onset. Bottom row: Scalp plots with signed$R^2$ values indicate spatial areas with high class-discriminative information.}
\caption{\textbf{Overview of neurophysiological features and supervised classification performance.} \color{darkgray} The amplitude and latency of peak amplitudes were derived after epoch-wise baseline removal and class-wise averaging of epochs. Values reported for N150 were determined as the minimum of channel O1 of the interval [100, 200]\,ms, while the late positivity (P300) was derived as the maximum of channel Cz in the interval[250 500]\,ms. The last column lists the AUC values estimated via cross-validation from a supervised classifier (see text).}
\caption{\textbf{ERP responses for S11 of sequence 1 \textbf{(A)} and sequence 2 \textbf{(B)}.} \\ \color{darkgray} Top row: Average responses evoked by target (blue) and non-target (green) stimuli in the occipital channel O1 (thick) and the central channel Cz (thin). Prior to averaging, a baseline correction was performed based on data within the interval [-200,0]\,ms. Bottom rows: Scalp plots visualising the spatial distribution of mean target and non-target responses within four selected time intervals:[50 120], [120 200], [201 380] and \mbox{[381 700]}\,ms.}
\caption{\textbf{ERP responses for S6 of the reconstructed class-wise means using LLP (\textbf{A}) and original labelled data (\textbf{B}). C shows the LLP target estimations in [120 200]\,ms for different numbers of training points}. \\ \color{darkgray} For details, see description of Fig~\ref{Fig:ERP_S1_vs_S2}.}
\caption{{\bf Spelling performance as seen online using the LLP (A) and after the post hoc re-analysis with LLP (B).} \\ \color{darkgray} Top: Each row represents a single spelling of the test sentence "Franzy jagt ...", with yellow squares indicating incorrectly spelled characters and blue squares indicating correctly spelled characters. Bottom: Average number of correctly classified characters.}
\caption{{\bf Comparison of LLP to an unsupervised EM-based classification approach for each sentence (A) and after 5 characters (B)} \\ \color{darkgray} \textbf{A}: Thin lines represent the binary target vs. non-target AUC performance of the two learning models with every line corresponding to the spelling of a single sentence. Each of the subjects ($N=13$) spelled each sentence three times resulting in 39 lines. Dashed lines depict average performances. \textbf{B}: Each dot represents the EM and LLP performance after 5 characters for the same subject and sentence. }
\caption{(Color on-line only) Mean evacuation time for 100 individuals as a function of the desired velocity. $\ \ $\protect\tikz\protect\draw[black,fill=black] (0,0) rectangle (0.3cm,0.3cm); No fallen pedestrians (``faster is slower''); \mydiamond{Blue} Narrow pathway; \protect\tikz\protect\draw[black,fill=red] (0,0) rectangle (0.3cm,0.3cm); Wide pathway; \protect\tikz\protect\draw[black,fill=green] (0,0) circle (.9ex);~v$_p$= 1~m/s; \mytriangleleft{cyan} v$_p$= 3~m/s; \mytriangleright{magenta} v$_p$= 6~m/s; \protect\hwplotB limit between those situations with no fallen (unconscious) individuals and thew ones with fallen pedestrians; mean values were computed from 30 realizations; the error bars corresponds to $\pm\sigma$ (one standard deviation).}
\caption{Cracked links computed with all approaches A1-A5. \textcolor{red}{SET \#34}}
\caption{ Schematic picture of the parameter region in the case~(a). The inflaton is dominated by the scalaron for $\xi_h^2/\lambda_h \lesssim \xi_s$ (the \textcolor{dark_red}{red region}), while it is dominated by the Higgs for $\xi_h^2/\lambda_h \gtrsim \xi_s$ (the \textcolor{dark_blue}{blue region}). The \textcolor{dark_yellow}{yellow band} corresponds to the parameter region consistent with the CMB observation. If $\xi_h^2/\xi_s \gtrsim 4\pi$ (the black region), the UV theory is non-perturbative so that we might not obtain the standard model effective theory even below the energy scale of the inflaton mass. For more details on this point, see Sec.~\ref{sec:cut-off}. }
\caption{ The time evolution of $\tau$ calculated from $\phi$ and $h$. The x-axis is time in the unit of the Planck mass inverse, and the y-axis is the values of $\tau$ and $\tau_\mathrm{min}$. The \textcolor{dark_blue}{blue line} corresponds to our numerical calculation, while the \textcolor{dark_red}{red dashed line} is $\tau_\mathrm{min}$ given in Eq.~\eqref{eq:tau_min}. We take the parameters as follows. Top: $\xi_s = 5\times10^8$, $\xi_h^2 = 2\times10^7$, $\lambda_h = 0.01$, $\phi_\mathrm{ini} = 6\,M_P$, and $h_\mathrm{ini} = 0.01\,M_P$ (Higgs-inflation like). Bottom: $\xi_s = 2\times10^{9}$, $\xi_h^2 = 5\times10^6$, $\lambda_h = 0.01$, $\phi_\mathrm{ini} = 6\,M_P$, and $h_\mathrm{ini} = 0.05\,M_P$ ($R^2$-inflation like). Here $\phi_\mathrm{ini}$ and $h_\mathrm{ini}$ are the initial field values of the inflaton and Higgs, respectively. We take the initial velocities of the fields as zero. $\tau$ starts to oscillate when $\xi_hM_P^2 t^2/\xi_s \sim 1$. After hundreds or thousands of oscillations $\tau$ eventually settles down to its potential minimum $\tau_\mathrm{min}$. Note that it happens within a few e-foldings because of the inequality~\eqref{eq:tau_heavy}, and hence $\tau$ has almost no effect on the inflationary dynamics. }
\caption{ The dynamics of the Higgs and scalaron in the field space. The x-axis is the Higgs squared $h^2$, and the y-axis is the scalaron $s$ calculated from $\phi$ and $h$. We take $M_P = 1$ in this plot. The \textcolor{dark_blue}{blue line} is our numerical calculation, while the \textcolor{dark_red}{red dashed line} is the direction of $\phi$ that is orthogonal to $\tau$. We take the parameters as follows. Top: $\xi_s = 5\times10^8$, $\xi_h^2 = 2\times 10^7$, $\lambda_h = 0.01$, $\phi_\mathrm{ini} = 6\,M_P$, and $h_\mathrm{ini} = 10^{-2}\,M_P$ (Higgs-inflation like). Bottom: $\xi_s = 2\times10^{9}$, $\xi_h^2 = 5\times10^6$, $\lambda_h = 0.01$, $\phi_\mathrm{ini} = 6\,M_P$, and $h_\mathrm{ini} = 0.05\,M_P$ ($R^2$-inflation like). Here $\phi_\mathrm{ini}$ and $h_\mathrm{ini}$ are the initial field values of the inflaton and Higgs, respectively. We have taken the initial velocities of the fields as zero. We have followed the inflationary dynamics until $M_Pt = 2\times10^7$. The two fields oscillate very frequently in the $\tau$ direction at first, but soon $\tau$ settles down to its potential minimum $\tau_\mathrm{min}$, and the two field dynamics follows the trajectory of $\phi$, which is expressed by the \textcolor{dark_red}{red line}. For the time evolution of $\tau$, see Fig.~\ref{fig:tau}. }
\caption{Comparison of LSRO, ``All in one'', and ``Pseudo label'' under different numbers of GAN-generated images on Market-1501. We show that LSRO is superior to the other two methods whose best performance is highlighted in \textcolor{blue}{blue} and \textcolor{red}{red}, respectively. Rank-1 accuracy (\%) and mAP (\%) are shown.}
\caption{Examples of pose estimation errors and the confidence scores. \textbf{Upper:} four pedestrian bounding boxes named with (a), (b), (c), and (d), and their pose estimation results. \textbf{Lower:} pose estimation confidence scores of the four images. A confidence vector consists of 14 numbers corresponding to the 14 body joints. We highlight the correctly detected joints in \textcolor{green}{green}. }
\caption{{\it Left panels:} Narrow band images of the \ion{Fe}{II}* emission optimally extracted from the MUSE cube shown with a 1\arcsec-wide slit oriented -20$^{\circ}$ (top) and +70$^{\circ}$ (bottom). The +70$^{\circ}$ slit orientation in the bottom panels follows the \ion{Fe}{II}* major kinematic axis. The white crosses indicate the galaxy major and minor axes from the \cite{2016A&A...591A..49C} kinematic analysis of [\ion{O}{II}]. {\it Right panels:} Position-velocity diagrams of the \ion{Fe}{II}* emission from a slit oriented -20$^{\circ}$ (top) and +70$^{\circ}$ (bottom). Zero velocity is relative to the galaxy systemic redshift, $z = $\redshift, measured from the [\ion{O}{II}] emission. Solid black contours trace \ion{Fe}{II}* flux levels at 2, 6, 10, 14, and $18\times 10^{-20}$~erg~s$^{-1}$~cm$^{-2}$. Dashed black contours show the negative flux levels at -2 and -4~$\times 10^{-20}$~erg~s$^{-1}$~cm$^{-2}$. The red gradient indicates the continuum intensity. The right side panel of each p-v diagram compares the spatial profile of the \ion{Fe}{II}* emission (black) with that of the stellar continuum (red). The bottom sub-panel of each p-v diagram shows the total 1D flux spectrum integrated across the spatial region between the solid green horizontal lines in the p-v diagram. In the bottom p-v diagram, the white solid line follows the velocity gradients. This panel also reveals a `C'-shape pattern in the blue wing of the \FeIIs\emission extending to -400~km~s$^{-1}$. The blue wing of the \ion{Fe}{II}* emission is more pronounced along the slit orientation corresponding to the minor-axis (PA$= +70^{\circ}$). }
\caption{ A distortion comparison among different warps in the non-overlapping region, which use the same homography alignment and the same seam-cutting composition in the overlapping region. In the stitching results, we use {\blue blue} rectangles to highlight the comparison of projective distortion, and {\red red} rectangles to highlight the comparison of perspective distortion. (a) Homography. (b) SPHP. (c) Our warp. Homography is of global consistency but suffers from projective distortion. SPHP is of local consistency but suffers from perspective distortion. Our warp is partially of global and local consistencies and balances projective and perspective distortion.}
\caption{Perspective distortion v.s. Slope preservation. (a) Mesh of the target image. (b) Warped mesh of homography. (c) Warped mesh of SPHP. (d) Warped mesh of our result. The {\red red line} is the special horizontal line that remains horizontal under a homography. Note that homography preserves arbitrary straight lines but incrementally increases areas of meshes along the red line, while SPHP maintains shapes of meshes by a uniform scaling factor but gradually changes the slope of straight lines. Our warp relaxes arbitrary line-preserving to only preserving the slope of the mesh, while relaxes uniformly-scaling everywhere to only uniforming the density of the mesh on the red line, to show a balance of perspective distortion and projective distortion in the non-overlapping region.}
\caption{Projective distortion v.s. Scale linearization. It demonstrates scaling functions of different warps on the special horizontal line that remains horizontal under a homography (marked {\red red} in Fig. \ref{meshsmall}), where $x_{*}$ corresponds to the closest vertical partition line that isolates $\mathcal{O}$, and $u_1$, $u_2$ correspond to two partition lines that divide $\mathbb{R}^2$ into three regions in (\ref{sphp}). Note that the scaling function is a rational one in homography (\ref{criticalx}), while it is a piece-wise function in SPHP which consists of a rational one as same as homography, a quadratic one and a linear one. The scaling function in our warp consists of a rational one as same as homography and a linear one.}
\caption{ Quasi-homography v.s. Homography. (a) Target image. (b) Reformulation of homography. (c) Derivation of quasi-homography. In the target image, any point $(x,y)$ can be expressed as the intersection point of a horizontal line and a vertical line, which is corresponding to a point $(x^{\prime},y^{\prime})$ under a homography as the intersection point of two lines with slopes $\mathrm{slope}(x,y,0)$ and $\mathrm{slope}(x,y,\infty)$. The location of $(x^{\prime},y^{\prime})$ can be controlled by the density on a horizontal line (marked in {\red red}) and a vertical line (marked in {\blue blue}). Note that quasi-homography linearizes the density on the red line but without changing the density on the blue line, such that it combines the homography by a single partition line and squeezes the meshes of homography without varying the shape.}
\caption{\label{fig:Cp_validation} (a) Pressure coefficient $C_p$ for the nominal case with grids listed in the table \ref{table:grids}. (\circle) experiments by \cite{coles1979flying}, \legendbaseline. (b) Evolution of the first off-wall cell size in wall units along the suction side. Legend: (\trianopd) $\Delta x^+$, (\triansold) $\Delta y^+$, and (\solidcircle) $\Delta z^+$. }
\caption{Streamwise velocity profiles in the separated region at locations $x=0.675c$, $x=0.731c$,$x=0.7860$, $x=0.842c$,$x=0.897c$, $x=0.953c$ (from left to right). Symbols are experimental values by \cite{coles1979flying} and lines are from large-eddy simulations for \legendbaseline, (\dotted) case A05(M=0.3). \label{fig:validation_case_u_and_v}}
\caption{ (a)Evolution of the pressure coefficient $-C_p$ near the trailing edge for $R=0.2$ \legendone and (b) Unsteady pressure contribution to the lift and drag coefficients for cases \ftwortwo~(\solid) and \ftworfive(\dashed). \label{fig:cpcl} }
\caption{A population of randomly initialized pathways (\textcolor{purple}{purple lines} in Box 1) are evolved whilst learning task A, Pong. At the end of training, the best pathway is fixed (\textcolor{darkred}{dark red lines} in Box 5) and a new population of paths are generated (\textcolor{lightblue}{light blue lines} in Box 5) for task B. This population is then trained on Alien and the optimal pathway that is evolved on Alien is subsequently fixed at the end of training, shown as \textcolor{darkblue}{dark blue lines} in Box 9.}
\caption{The PathNet used for learning Atari and Labyrinth games consists of a 4 layer network with 10 (or sometimes 15) modules in each layer. The first three layers' modules are convolutional 2D kernels with 8 kernels per module (\textcolor{fig2_green}{green boxes} in figure), kernel sizes (8, 4, 3), and strides (4, 2, 1) from 1st to 3rd layers respectively, and with the final layer consisting of fully connected linear layers of 50 neurons each (\textcolor{fig2_purple}{purple boxes}). After each module there is a rectified linear unit. Between layers the feature maps are summed before being passed to the modules of the subsequent layer (\textcolor{fig2_blue}{blue boxes}). Typically a maximum of 4 modules per layer are permitted to be included in a pathway (shown as \textcolor{fig2_red}{red boxes}), otherwise evolution would simply grow the pathway to include the whole network as this would improve fitness by growing the number of parameters that can be tuned by learning.}
\caption{PathNet outperforms the de novo control and the fine-tuning control on learning seekavoid\_arena for the first time, on average over all hyperparameters explored, out of 243 experiments. It also outperforms fine tuning when seekavoid\_arena\cite{jaderberg2016reinforcement} is learned a second time, with a new value and policy readout, with the same set of 243 hyperparameter settings.}
\caption{Average bounding box overlap ratio on individual sequence. \textcolor{red}{Red}: best, \textcolor{blue}{blue}: second best.}
\caption{Verification of the filter responsivities with FC2 {\it ICO Calibration} observations of Saturn. The intensity was integrated over a circle of 9 pixel radius surrounding the target ($n=8$). The Saturn albedo was adopted from \citet{K94}, assuming a disk radius of 60550~km, shown along with the averages in the filter passband ($\square$). For clarity, error bars are shown only for one of the symbol types. {\bf A}:~The observed flux (in DN~s$^{-1}$) compared to that expected, using the correction factors derived from standard star observations (\textcolor{red}{$\blacktriangle$}) and the revised factors in Table~\ref{tab:filters} ($\blacktriangledown$). The albedo beyond 1000~nm is not available (assumed zero), which underestimates the expected flux for filter F5. {\bf B}:~The observed flux (in pW m$^{-2}$ nm$^{-1}$) compared to that expected. (\textcolor{red}{$\bullet$}): Responsivities calculated for a target with a solar spectrum with the original correction factors. (\textcolor{LimeGreen}{$\bullet$}): Responsivities calculated for Saturn albedo with the original factors. ($\bullet$): Responsivities calculated for Saturn albedo with the revised factors. {\bf C}:~Reconstruction of Saturn's albedo through Eq.~\ref{eq:reflectance}. Plot symbols as for (B).}
\caption{ \label{fig:HSSComparisonGauss} Pseudotransparency plots of performance measures for the family $f_{j}$ and varying $n$ using a Gaussian kernel. {\color{red}Data for CV is shown in red}, while {\color{blue}data for TDE is shown in blue}. Each panel represents data on a particular performance measure (top, $\hat h - h_{opt}$; middle, $\text{ISE}(\hat h)$; bottom, $c_{45}$) and $f_j$ (from left to right: $j = 1,\dots,6$). For each value of $n$, the corresponding vertical strip in a panel depicts the histogram (with precisely the same bins for all values of $n$) of a performance measure using a normalized linear transparency scale.}
\caption{ \label{fig:fkmEval500Matlab} (Left panels) Pseudotransparency plots of performance measures for the family $f_{km}$ for $n = 500$ using a Gaussian kernel. {\color{red}Data for CV is shown in red}, while {\color{blue}data for TDE is shown in blue}. Each panel represents data on a particular performance measure (left, $\hat h - h_{opt}$; middle, $\text{ISE}(\hat h)$; right, $c_{45}$) and value of $k$ (top, $k = 1$; middle, $k = 2$; bottom, $k = 3$), with $m$ varying from 1 to 10. The vertical axis limits are the same within each column. For each value of $m$, the corresponding vertical strip in a panel depicts the histogram (with precisely the same bins for all values of $m$) of a performance measure using a normalized linear transparency scale. (Right panels) As in the left panels, but with an Epanechnikov kernel.}
\caption{ \label{fig:fkmEval500MatlabTop} (Left panels) Pseudotransparency plots of performance measures relating to the unimodal category for the family $f_{km}$ for $n = 500$ using a Gaussian kernel. {\color{red}Data for CV is shown in red}, while {\color{blue}data for TDE is shown in blue}. Each panel represents data on a particular performance measure (left, $\text{ucat}$ for TDE alone; middle, $\text{ucat}$ for CV alone; right, the empirical probability that the estimate of $\text{ucat}$ is correct) and value of $k$ (top, $k = 1$; middle, $k = 2$; bottom, $k = 3$), with $m$ varying from 1 to 10. (Right panels) As in the left panels, but for the number of local maxima instead of the unimodal category.}
\caption{ \label{fig:fkmEval500MatlabFitdist} As in figure \ref{fig:fkmEval500Matlab} for $n = 500$, but with {\color{red}red indicating data for MATLAB's \texttt{fitdist}} instead of CV.}
\caption{Ground state energy $E_N(S^z)$ for the triangular lattice as a function of $t'$, and for different $N$-site clusters. Dashed (solid) lines correspond to the $S^z =S_{max}^z$ ($S^z_{min}$) sector energy. For smaller $t'$ both energies overlap. $E^{\triangle}_{Nag}$ is the thermodynamic limit of the Nagaoka state energy. Inset: zoom of the critical region, showing the intersection of $E_N\left(S^z_{min}\right)$ with the ferromagnetic energy $E^{\triangle}_{Nag}$.} \label{ene-t} \end{figure} To determine the critical $t'_c,$ firstly we have looked for the $t'$ value where the degeneracy between $E_N\left(S^z_{min}\right)$ and $E_N\left(S_{max}^z\right)$ is broken for each lattice. In Figs. \ref{ene-t} and \ref{ene-c} we show these energies as a function of $t'$ and for different cluster sizes, for the triangular and square lattices, respectively. We can see that the degeneracy is broken in the region around $t' \sim 0.2 - 0.3$, signaling, as we explained above, that the ground state of the systems moves from the highest-spin Nagaoka state to another one with minimal total spin. As the conventional DMRG algorithm uses the $S^z$ quantum number without discriminating between different total spin $S$, we cannot compute the energy of the excited $S=\frac{1}{2}$ state below $t'_c$ (notice that, in this case, the calculated $E_N\left(S^z_{min}\right)$ corresponds to the $S^z=\frac{1}{2}$ sector energy of the $S_{max}$ Nagaoka ground state). Therefore, we do not have access to the expected energy level crossing between the highest- and lowest-spin sectors, that would facilitate the determination of $t'_c$. For the square lattice model, the lack of the level crossing is not so important as there is an appreciable kink in the ground state energy at $t'_c$ (see main panel of Fig. \ref{ene-c}). However, for the triangular case, the transition seems to be much smoother, as is shown in the main panel of Fig. \ref{ene-t}, and, consequently, it is more difficult to estimate the critical point where the degeneracy between $E_N\left(S^z_{max}\right)$ and $E_N\left(S^z_{min}\right)$ is lost. To avoid this difficulty, we have evaluated $t'_c$ extrapolating the level crossing between $E_N\left(S^z_{min}\right)$ and the infinite-lattice Nagaoka energy, $E_{Nag}$ (see insets of Figs. \ref{ene-t} and \ref{ene-c}). $E_{Nag}$ can be computed exactly as the problem of one hole moving in a ferromagnetic background is identical to the spinless tight-binding system \cite{brinkman70} \begin{table}[t] \begin{center} \begin{tabular}{lcccccc} \hline \hline $N$ & $18$ &$36$ & $54$ & $72$ & $90$ & $\infty$\\ $t'^{\triangle}_c$ & $0.387$ & $0.295$ & $0.268$ & $0.256$ & $0.249$ & $0.222$\\ \hline \hline \label{tab-t} \end{tabular} \end{center} \caption{Critical $t'_c$ values for different $N$-site triangular clusters, and its $N\to \infty$ extrapolation limit.} \label{table_ene-t} \end{table} \begin{figure}[hb] \centering \includegraphics[width=0.45\textwidth]{fig3.pdf} \caption{Ground state energy $E_N(S_z)$ for the square lattice as a function of $t'$, and for different $N$-site clusters. Dashed (solid) lines correspond to the $S_z =S^{max}_z$ ($S_z = \frac{1}{2}$) sector energy. $E^{\square}_{Nag}$ is the thermodynamic limit of the Nagaoka state energy. Inset: zoom of the critical region, showing the intersection of $E_N\left(\frac{1}{2}\right)$ with the ferromagnetic energy $E^{\square}_{Nag}$.}
\caption{Local magnetization $M_s$ as a function of $t'$ for (a) the $N=54$ triangular lattice without a magnetic field (open triangles) and with a magnetic field $h=0.1$ applied to one $120^{\circ}$ sublattice (solid triangles); (b) the $N=60$ square lattice without a magnetic field (open squares) and with a magnetic field $h=0.1$ applied to one of the two sublattices of the $(\pi,\pi)$ N\'eel order. The dashed lines correspond to the classical local magnetization,$M_s^{clas}=\frac{1}{2}-\frac{1}{2N}$.}
\caption{Theoretical (solid and dotted lines) and simulation (solid and open symbols) results for the MSDs, $W_{i}(t)$, and ISFs $F^s_{i} (k=7.18;t)$ of a highly asymmetric ($\delta=0.2$) binary mixture of HS, for the three volume fractions, in the control parameter space $(\phi_b,\phi_s)$. (a) and (b) Corresponds to the point \textbf{I}$=(0.45,0.05)$. (c) and (d) Corresponds to \textbf{II}$=(0.45,0.2)$. (e) and (f) Corresponds to \textbf{III}$=(0.60,0.05)$. In all cases, solid symbols ($\CIRCLE$,\textcolor{red}{$\blacktriangle$}, \textcolor{blue}{$\blacksquare$}) and solid lines corresponds to the dynamical properties of big particles, whereas empty symbols ($\Circle$, \textcolor{red}{$\bigtriangleup$}, \textcolor{blue}{$\square$}) and dashed lines accounts for small particles.}
\caption{ Time history of pressure amplitudes of a freely-traveling wave at 300 K, with 1 kHz propagation with zero bulk viscosity \legenddashed[black]{} and with effective bulk viscosity \legendline[black]{}, and for 10 kHz propagation with zero bulk viscosity \legenddashed[gray]{} and with effective bulk viscosity \legendline[gray]{}. Simulation relative humidity in percentage is 20\%. }
\caption{ Pressure and flow rate fluctuation amplitudes along the curvilinear abscissa $x_s$, as defined by an $x$ segment through the tubes and stack, an abscissa $j$ segment through the junction, and an $r$ segment through the radial cavity. Simulation data \legenddot{} vs Rott's theory \legenddashed{} along defined axis. }
\caption{ Time-series of pressure amplitudes predicted by Navier--Stokes calculations: no bulk viscosity \legendline{} and with bulk viscosity, corresponding to relative humidity levels of 1\% \legenddashed{}, 5\% \legenddotted{}, and 20\% \legenddasheddotted{}. }
\caption{ Dimensionless effective bulk viscosity $\mu_B^*$ versus frequency at $T=$ 300 K \legendline{}, 450 K \legenddashed{}, 600 K \legenddotted{}, and 750 K \legenddasheddotted{}. Results shown for $h_r=5\%$ and atmospheric pressure (a). Dimensionless effective bulk viscosity $\mu_B^*$ versus relative humidity, $h_r$, at $T=$ 300 K \legendline{}, 450 K \legenddashed{}, and 600 K \legenddotted{}. Chosen humidity levels of 1\%, 5\%, and 20\% are highlighted with vertical dotted lines (b). }
\caption{Free opening energies of A{\textperiodcentered}T and G{\textperiodcentered}C base pairs, calculated in the corresponding works, the difference between the opening forces of A{\textperiodcentered}T and G{\textperiodcentered}C base pairs $\Delta F=F_{GC}-F_{AT} $ and the averange opening force $F_{av}=(F_{GC} +F_{AT})/2 $. $^*$fitting parameters in~\cite{3Bockelmann}, $^{**}$opening energies of A{\textperiodcentered}T- and G{\textperiodcentered}C- containing polymers, averaged over the nearest neighbors~\cite{mfk2006}, $^{***}$our estimations using (\ref{for:op_force}), $^{****}$calculated in~\cite{3Bockelmann}.}
\caption{Schematic representation of the adenine-thymine nucleic base pair in DNA macromolecule: a) Watson-Crick configuration~\cite{Saenger}. The arrow shows the {\textquoteleft}opening{\textquoteright} pathway: the bases rotation around the axis passing through atoms $C_{1}'$ perpendicular to the plane of the pair; b) {\textquoteleft}pre-opened{\textquoteright} A{\textperiodcentered}T pair mediated by the water molecule~\cite{VolkovKryachko}.}
\caption{Parameters for A{\textperiodcentered}T and G{\textperiodcentered}C Watson-Crick pairs. The hydrogen bond distance values and $R$ distances are taken from~\cite{Saenger}. The angles $\alpha_1$, $\alpha_2 $ and binding energy $E$ are calculated in this paper.}
\caption{Binding energy dependencies from distance calculated for A{\textperiodcentered}T (dashed line) and G{\textperiodcentered}C (solid line) base pairs for {\textquoteleft}stretch{\textquoteright} scenario. $E_{AT},E_{GC}$ - calculated opening energies as the difference between the Watson-Crick configuration (minimum) and configuration when central hydrogen bond ($N_1$ {\textbullet}{\textbullet}{\textbullet} $N_3$ distance) reaches $3.7$ {\AA} (red points), $\Delta E= E_{GC}-E_{AT}$.}
\caption{Additional \aastex\symbols}
\caption{\label{ROC-example} Examples of stories from the story cloze task. The table shows a story prefix with three contrastive endings: The {\color{blue}{\bf original}} ending, a {\color{forestgreen}{\it coherent}} ending and a {\color{red}{\uline{incoherent}}} one. }
\caption{The distribution of five frequent POS tags (\ref{roc_pos_distribution}) and words (\ref{roc_word_distribution}) across {\color{blue}{\emph{original}}} endings (horizontal lines) from the story cloze training set, and {\color{forestgreen}{\emph{right}}} (diagonal lines) and {\color{red}{\emph{wrong}}} (solid lines) endings, both from the story cloze task development set.}
\caption{Single- and two-qubit gate counts for the circuits on the superconducting (star-shaped) and the ion trap (fully connected) system after mapping to the respective hardware using the respective gate libraries. For comparison, the gate counts for a linear nearest-neighbor (LNN) architecture as implemented in \cite{Barends14} are included. We also note the gate count for the Quantum Fourier Transfrom (QFT) for $3$ and $5$ qubits. The latter was implemented in \cite{Debnath16} using a sequence of modular gates that was not optimized for gate count. The QFT-5 cannot be implemented exactly using the current IBM gate library. If we assume Z$^a$ operations are possible, the counts shown as {\textasteriskcentered } are $47$ for single- and $29$ for two-qubit gates.}
\caption{Halo mass functions (HMFs) for the different baryon physics runs (in the absence of neutrino physics) from cosmo-OWLS (left) and the different collisionless massive neutrino runs from \calsim\(right). Colours denote the various runs (see the legend and Table \ref{tab:sims}), while the different linestyles denote different redshifts. In the bottom left panel, the cosmo-OWLS HMFs have been normalised by the \dmonly\case, whereas in the bottom right panel the HMFs have been normalised by the massless neutrino case. Suppression of the HMF due to AGN feedback (orange, yellow and green, left) is important at intermediate (group) masses but becomes less important at high halo masses, where it begins to converge towards the\dmonly\case (with the mass scale where the convergence occurs depending on the AGN heating temperature\deltaTheat). The suppression due to feedback is only a weak function of redshift. In the collisionless neutrino simulations (right panel), the suppression is strongest for the highest mass haloes and, in contrast to the effects of feedback, exhibits a strong redshift dependence. }
\caption{\label{fig:a_CC_1} Evolution of the comoving halo space densities above different mass thresholds [i.e., $n(M_{200,{\rm crit}}>M_{\rm threshold},z)$] for the different baryon physics runs in the absence of neutrino physics using cosmo-OWLS (left) and in the absence of baryon physics using the different collisionless massive neutrino runs from \calsim\(right). Solid, dashed, and dotted curves correspond to threshold masses of $10^{12}~{\rm M_{\odot}}$, $10^{13}~{\rm M_{\odot}}$, and $10^{14}~{\rm M_{\odot}}$, respectively. The bottom left panel shows the halo space densities for the baryon physics models normalised to the \dmonly\case, while the bottom right panel shows the massive neutrino models normalised to the dark matter-only massless neutrino (\nua~DM) case. The introduction of AGN feedback results in a suppression of the halo space density that is nearly independent of redshift, while the suppression above a fixed mass threshold due to \nufs\increases strongly with increasing redshift, particularly for models with high values of the summed neutrino mass.}
\caption{\label{fig:a_MC_1} Best fit total mass \mc\relations for different baryon physics models in the absence of neutrino physics in the\wmapa\cosmology (left) and for different\mnu\values in the absence of baryonic physics in\wmapb\cosmology (right) at$z=0$. Halo masses are those from each of the individual simulations (i.e., not the matched DM-only masses). Stars mark the locations of the mean concentration value in each 0.5 dex mass bin, solid lines display the best-fit power laws to these means with the functional form of equation \ref{eq:Neto}. The upper panel displays these in log(c) - log(M) space, while the lower panel displays the same data normalised to the best fit for the the relevant DM only model. Increasing \mnu\results primarily in a reduction of the amplitude of\mc\with respect to the\dmonly~\nuadm\model with minimal alteration to the gradient. Conversely,\fdbk\alters both the amplitude and the gradient.}
\caption{\label{fig:cl_1} Real-space 2-point halo autocorrelation functions ($\xi$)\for the different baryonic physics runs in the absence of neutrino physics from cosmo-OWLS (left) and the different collisionless massive neutrino runs from BAHAMAS (right). Solid, dashed and dotted curves correspond to$\xi$ for haloes in mass bins of $10^{12}-10^{13} {\rm M}_\odot$, $10^{13}-10^{14} {\rm M}_\odot$ and $10^{14}-10^{15} {\rm M}_\odot$ in $M_{200,{\rm crit}}$ respectively. The bottom left panel shows $\xi$ for each of the baryonic physics models normalised to the \dmonly~case, while the bottom right panel shows the massive neutrino models normalised to the dark matter-only massless neutrino (\nua\DM) case. The introduction of AGN feedback results in a$\sim10\%$ increase in the amplitude of the autocorrelation function, with the precise shift depending on the halo mass range and AGN heating temperature under consideration. Neutrino free-streaming also increases the amplitude, with the strength of the effect depending sensitively on the precise value of the summed neutrino mass.}
\caption{\label{fig:cl_2} Comparison of the real space 2-point halo autocorrelation functions ($\xi$)\arising when simulating baryonic feedback and neutrino free-streaming simultaneously (curves) and those calculated by multiplying the separate effects of baryonic feedback in the absence of neutrinos and the effects of neutrino free-streaming in the absence of baryon physics (crosses). Solid, dashed and dotted lines correspond to$\xi$ for haloes in mass bins of $10^{12}-10^{13} {\rm M_\odot}$, $10^{13}-10^{14} {\rm M_\odot}$ and $10^{14}-10^{15} {\rm M_\odot}$ in $M_{200,{\rm crit}}$ respectively. The bottom panel shows each of the multiplicative models normalised by the corresponding combined case. The multiplicative treatment recovers the combined result with a few percent accuracy for $r > 1 {\rm h^{-1}Mpc}$ independent of the chosen summed neutrino mass in all but the highest mass bin.}
\caption{ The energy spectrum (a) and the enstrophy spectrum (b) of a synthesized turbulence field generated on a $256 \times 256 \times 256$ grid using an integral length of $L = 0.1$ and a covariance tensor of $\boldsymbol{\Sigma} = \boldsymbol{I}$. The re-sampled spectra of stochastically generated turbulence field:~({\color{blue} \textbf{--------}}) is here compared to the target spectra used to generate the field:~({\color{red} \textbf{--------}}). The wave-number of the smoothing radius $\sigma$:~({\color{black} \textbf{--~--~--~--}}) corresponding to minimum Kolmogorov length of $\eta/L = 0.0141$ or equivalently a kinematic viscosity of $\nu = 2.6 \times 10^{-3}$. A 10th order smoothing filter is used. }
\caption{\label{tab1} Constraints on free parameters and two derived parameters $H_0$ and $\sigma_8$ of the $\nu$IVCDM, $\nu \Lambda$CDM, $\nu_s$IVCDM and $\nu_s \Lambda$CDM models. For each parameter, the four entries (mean values with 1$\sigma$ errors) are the constraints from the data combinations \textcolor{red}{CMB + BAO + JLA}, \textcolor{blue}{CMB + BAO + JLA + CFHTLenS}, \textcolor{magenta}{CMB + BAO + JLA + SZ} and \textcolor{darkgreen}{CMB + BAO + JLA + CFHTLenS + SZ}, respectively. The parameter $H_0$ is in the units of km s${}^{-1}$ Mpc${}^{-1}$, while $\sum m_{\nu}$ and $m_{{\nu}_s}$ are in the units of eV. The $\chi^2_{\rm min}$ values of the fit are also displayed.}
\caption{Behaviour of absolute value of \textcolor{myblue}{$H_V^+$}, \textcolor{myorange}{$H_V^-$} and \textcolor{mygreen}{$H_V^0$}. The ``SM'', ``SM + power correction'' and ``SM + NP'' are shown with solid, dashed and dotted lines, respectively. Assuming $h_{+}^{(0)}$ to be constrained in the fit for power corrections. \label{fig:HVallAbsConstrained}}
\caption{Behaviour of absolute value of \textcolor{myblue}{$H_V^+$}, \textcolor{myorange}{$H_V^-$} and \textcolor{mygreen}{$H_V^0$}. The ``SM'', ``SM + power correction'' and ``SM+NP'' are shown with solid, dashed and dotted lines, respectively. No constraint is assumed for $h_{\lambda}^{(0,1,2)}$ in the fit for power corrections. \label{fig:HVallAbsFree}}
\caption{ Correlation of the temporal dynamics of cities and heart beats. Left panels: An example of heart beat ECG signal at approximately 80 beats per minute (red) and the average week of activity for three cities: S\~ao Paulo (yellow), Jakarta (blue) and London (green). Vertical black lines show the time of synchronization (see text). Right panel: Correlation of the heartbeat with 50000 random series (gray curve), other heartbeats (red line), periodic signals (magenta lines, from left to right: sawtooth, squared and sinusoid), and all urban areas colored by group determined by clustering analysis (see Supplement).}
\caption{Energies of the low-lying flux tube states with the quantum numbers shown, versus the length of the flux tube. Results for $SU(3)$ at $\beta=6.0635$ taken from \cite{AABBMT_d4}, compared to our new results for the $J^{P_{\shortparallel}P_{\perp}} = 0^{--}$ flux tube at $\beta=5.825$ ({\color{red} $\circ$}).}
\caption{Architecture for Bilateral Multi-Perspective Matching (BiMPM) Model, where {\color{red} $\otimes$} is the multi-perspective matching operation described in sub-section \ref{sub:multi-func}.}
\caption{The intensity-dependent index of refraction of ITO at $\lambda=1240\,\textrm{nm}$, where the real part of the linear permittivity $\epsilon^{(1)}$ vanishes. Equation~(\ref{eq:n_smallI}) performs poorly at describing the refractive index at most intensities (dashed blue line). Using Eq.~(\ref{eq:intensity-dependent index withchi}), we obtain much-improved agreement with the measurement without additional fit parameters (red line). \redline{$I_0$ and $E$ refer to the free-space intensity and the corresponding field magnitude for a plane wave inside the material, respectively.}}
\caption{Interference patterns versus focusing magnetic field for two values of the quantum point contact apertures $W$. The spin orbit interaction is so strong that there are two spin orbit split focusing peaks. The separation between the source and the detector is 2000nm. The black dashed curves are obtained with the semiclassical kernel (\ref{kso}) and the red curves are obtained with the full ``Airy'' kernel (\ref{kso1}). {\cred This figure corresponds to the unpolarized source/detector} }
\caption{({\em color online}) Predicted high-pressure phase diagram of \bbo. Up to 100 GPa, we predict three structural phase transitions: monoclinic (\monoclinic) to triclinic (\triclinic) at about 20~GPa, triclinic to \textit{``clustered''} (\clustered) at about 28~GPa and \textit{``clustered''} to ``distorted$_1$'' (\distorted) at about 87~GPa. EP (Evolutionary Prediction) means that structures were obtained using the evolutionary algorithms structural prediction method and SA (Symmetry Analysis) -- using the group-theoretical approach.}
\caption{({\em Color online:}) Average number of O neighbors of Bi atoms for a pool of two formula units \bbo\structures obtained using the evolutionary algorithms approach. Each dot is associated with a specific structure and its color denote the average number of neighbors. For the perovskite structure the average number of neighbors is 6 and the deviation from this value shows the amplitude of distortion presented in the system. Details on the neighbor analysis can be found in the supplemental material in subsection ``C. Neighbors analysis''.}
\caption{\textcolor{Black}{PV generation real output power and predictions.}}
\caption{\textcolor{Black}{Nominal total microgrid load profile (for bus voltages of 415V) and predictions.}}
\caption{Additional \aastex\symbols}
\caption{Phase-space plot of the vertical and horizontal sphere velocity components for case C ($Ga=190$) with $D/\Delta x=36$ (left) and $48$ (right). {\color{blue} \textbf{---}} BB, {\color{OliveGreen} \textbf{---}} CLI, {\color{red}\textbf{---}} MR, {\color{Turquoise} \textbf{---}} M1B1, {\color{Mulberry}\textbf{---}} M2B2, {\color{GreenYellow}\textbf{---}} M3B2, {\color{black}\textbf{---}} reference.}
\caption{Normalized probability density function of translational sphere velocity for D-CLI-36 (left) and D-M3B2-36 (right) with $Ga=250$. {\color{black} \textbf{---}} vertical component $u_{pV}$, {\color{red} \textbf{---}} horizontal component $u_{pr}$, {\color{blue} \textbf{---}} Gaussian reference curve. The reference data from \cite{uhlmann_motion_2014} is shown as dashed lines.}
\caption{Normalized probability density function of rotational sphere velocity for D-CLI-36 (left) and D-M3B2-36 (right) with $Ga=250$. {\color{black} \textbf{---}} vertical component $\omega_{pV}$, {\color{red} \textbf{---}} horizontal component $\omega_{px}$, {\color{blue} \textbf{---}} Gaussian reference curve. The reference data from \cite{uhlmann_motion_2014} is shown as dashed lines.}
\caption{Temporal auto-correlation of translational sphere velocity for D-CLI-36 (left) and D-M3B2-36 (right) with $Ga=250$. {\color{black} \textbf{---}} vertical component $u_{pV}$, {\color{red} \textbf{---}} horizontal component $u_{pr}$. The reference data from \cite{uhlmann_motion_2014} is shown as dashed lines.}
\caption{Temporal auto-correlation of rotational sphere velocity for D-CLI-36 (left) and D-M3B2-36 (right) with $Ga=250$. {\color{black} \textbf{---}} vertical component $\omega_{pV}$, {\color{red} \textbf{---}} horizontal component $\omega_{px}$. The reference data from \cite{uhlmann_motion_2014} is shown as dashed lines.}
\caption{Computer model of SKALA-2 antenna on top of a metallic ground plane (the ground plane is a mesh of wires with 30 cm pitch). The antenna is a log-periodic dipole array made of 4 identical metallic arms forming 2 polarizations with 9 dipoles each. The gray ring is the base of the antenna. The cabling coming down vertically in the center of the antenna \highlighttext{is} the optical fiber and power cable. The top enclosure with a green cap hosts the low-noise amplifiers and \highlighttext{an} RF-over-Fiber transmitter. The antenna has a footprint of 1.2 x 1.2 m and a height of 1.8 m.}
\caption{\highlighttext{Simplified RF schematic model of the first amplification stage of the LNA. The S-parameters and noise parameters for the transistors used in this design are obtained from the manufacturers. This diagram highlights the key components in the input side of the LNA that have been used for the improvement of the design (the DC blocking capacitors C1 and C2, the bias inductors L1, L2 and the ESD shunt inductors L3 and L4). The ESD diodes D1 and D2 are also shown here. For the simulations in this paper the full model of the LNA also included the second stage transistor although this has a small effect on the matching with the antenna.}}
\caption{Antenna and LNA impedances; (a) LNA input impedance for power match, (b) LNA noise circles for noise match and (c) Antenna input impedance. \highlighttext{The Smith chart is plotted on the complex reflection coefficient plane in two dimensions. The horizontal axis corresponds to a purely real reflection coefficient and the center is the reference impedance, 50 $\Omega$.}}
\caption{Requirements from~\citet{Trott2016}. $\delta$ is the normalized fractional residual after \highlighttext{a third-order} polynomial fitting.}
\caption{\highlighttext{Amplitude of the LNA voltage passband when connected to SKALA-3 (blue) vs when connected to SKALA-2 (red). The plot shows the sharp feature found in the passband of SKALA-2 at 60 MHz.}}
\caption{\highlighttext{Phase of the LNA voltage passband when connected to SKALA-3 (blue) vs when connected to SKALA-2 (red).}}
\caption{\highlighttext{Residuals obtained when fitting wider bandwidths.}}
\caption{Column 1, 2 and 3 represent Bibtex, Delicious, and EURLex datasets. Results for standard datasets: Bibtex, Delicious and EURLex. Legend: \textcolor{red}{RIPML (- -$\blacktriangle$- -)}, \textcolor{black}{LEML (--$\blacksquare$--)}, \textcolor{blue}{CPLST (--$\bullet$--)}, \textcolor{green}{CSSP (--$\blacklozenge$--)}. Y-axis: m = [50,~100,~ 20\% of $L$,~ 40\% of $L$,~ 80\% of $L$]. Here $k = 5$.}
\caption{Column-1, 2 and 3 correspond to $d = 200, 300,$ and $400$ respectively. Results on Chinese Relevance Modeling for different ambient dimensions. Legend: \textcolor{red}{RIPML (- -$\blacktriangle$- -)}, \textcolor{black}{LEML (--$\blacksquare$--)}, \textcolor{blue}{CPLST (--$\bullet$--)}, \textcolor{green}{CSSP (--$\blacklozenge$--)}. X-axis: m = [50,~100,~ 20\% of $L$,~ 40\% of $L$,~ 80\% of $L$]. Here $k = 5$.}
\caption{\label{fig:patent} \emph{UltimateWalk scales linearly.} The proposed algorithm UltimateWalk scales well, linearly with the number of edges. (U.S. patent 1975--1999 dataset; see Lemma{\color{red}~\ref{thm:complexity}}.) }
\caption{\label{fig:MethylAcetate}The PACs (top panels) and the ESP correlation (bottom panels) for the methyl-acetate molecule, using standard population analysis methods (left panels) and other PAC methods (right panels). The mean absolute relative deviation (MARD) of Eq.~\ref{eq:MARD} appears in parenthesis near each PAC method. Each point in the ESP correlation plot describes a pair of potentials $\left[\varphi^{PAC}\left(\boldsymbol{r}\right),\varphi^{QM}\left(\boldsymbol{r}\right)\right]$, the abscissa is the PAC potential (Eq.~\eqref{eq:PotPac}) and the ordinate the quantum potential (Eq.~\eqref{eq:QMPot}) where $\boldsymbol{r}$ is taken from a subset of grid points of spacing $\Delta x=0.3\mathring{A}$ around the molecule (see description of the grid in Section~\eqref{sec:Results}). } \end{figure} Another source of PACs are the quantum mechanical population analysis (PA) techniques, such as the Mulliken (MPA),\cite{Mulliken1955}, Loewdin (LPA),\cite{Loewdin1950}, Hirshfeld (HPA)\cite{Hirshfeld1977}, and natural population (NPA)\cite{Foster1980} analyses. These PAs reflect not only the charge distribution but also aspects of the quantum mechanical wave function. In Fig.~\ref{fig:MethylAcetate} (top left) we show as bar-plot the PACs produced by these methods applied to the methyl acetate molecule. It is seen that the different methods produce sometimes significantly different sets of PACs, even PAC signs are not preserved! For example, the LPA assigns positive charges to oxygen atoms, which seems awkward given their high electronegativity. Furthermore, standard PAs do not reproduce the MOL-ESPs closely, as shown in the ESP correlation plot of Fig.~\ref{fig:MethylAcetate} (bottom left), where several PAC-ESPs, \begin{equation} \varphi^{PAC}\left(\boldsymbol{r}\right)=\frac{e}{4\pi\epsilon_{0}}\sum_{a}\frac{q_{a}}{\left|\boldsymbol{r}-\boldsymbol{R}_{a}\right|},\label{eq:PotPac} \end{equation} are plotted vs. the MOL-ESP $\varphi\left(\boldsymbol{r}\right)$ calculated from the QM density (Eq.~\ref{eq:QMPot}) at a grid point $\boldsymbol{r}$. The thin red line in the plot corresponds to the perfectly correlated condition $\varphi^{PAC}=\varphi^{QM}$. In order to quantify the quality of $\varphi^{PAC}\left(\boldsymbol{r}\right)$ we define the mean absolute relative deviation (MARD) from $\varphi\left(\boldsymbol{r}\right)$ as \begin{equation} MARD\left(\varphi^{PAC},\varphi\right)=\left\langle\left|\frac{\varphi^{PAC}\left(\boldsymbol{r}\right)-\varphi\left(\boldsymbol{r}\right)}{\varphi\left(\boldsymbol{r}\right)}\right|\right\rangle ,\label{eq:MARD} \end{equation} where an average is taken over all grid-points $\boldsymbol{r}$ for which: 1) $\boldsymbol{r}$ is ``outside of the molecule'', i.e. its distance from any nucleus $a$ is larger than the atomic van-der-Waals radius $R_{a}^{vdW}$ \cite{Singh1984}) and 2) $\boldsymbol{r}$ is not too far from the molecule, so that its potential $\left|\varphi\left(\boldsymbol{r}\right)\right|$ is not smaller than the threshold value of $e\varphi_{thresh}=0.3eV$.\footnote{Note that the expression in Eq.~\ref{eq:MARD} cannot become singular due to this requirement.} PACs obtained by ``standard'' PAs have large MARDs: ranging from 0.37 for HPA up to a whopping 1.76 for NPA. On the right panel of the figure we show data concerning the same molecule, but using the iterated-Hirshfeld method (iHPA),\cite{Bultinck2007,Bultinck2009,VanDamme2009} the CM5 method \cite{Marenich2012}, which is a parameterized database correction to HPA charges, and the ChElPG method\cite{Breneman1990}, which selects PACs that reconstruct the \emph{ab initio }ESP on a set of grid points as close as possible. The latter approach is taken here as representative of a class of methods routinely used for PACs determination. Other members of this method class are the ``charge from ESPs'' (ChElP)\cite{Chirlian1987}, the Merz-Kollman\cite{Momany1978determination,Singh1984,Besler1990a}, the charge-restraint ESPs (RESP)\cite{Bayly1993,Cornell1993}, atomic multipoles ESPs\cite{Williams1988}, in combination with molecular multipoles\cite{Sigfridsson1998} (related to the method proposed here), the dynamical RESP (D-RESP)\cite{Laio2002a} and Hu-Yang fitting\cite{Hu2007}.The iHPA, CM5 and ChElPG methods yield much improved description of the ESP with MARD going from 0.3 for iHPA and CM5 down to 0.08 for ChElPG. Despite the close ESP fit, ChElPG produces PACs that are usually not invariant under transformations preserving the point symmetry of the molecule or under rotations or translations of the nuclei with respect to the real space grid used to perform the fit. Furthermore, in larger molecules the PACs of atoms distant from the molecular surface can become unwieldy large. Both of these issues are discussed in the literature\cite{francl2000pluses,Hu2007}This instability is likely linked to the fact that the number of parameters derivable from the ESP in a statistically significant way is considerably less than the number of atoms.\cite{jakobsen2016searching} Therefore, iHPA and CM5 are often considered preferred approaches for PACs, although as seen in the figure, both methods leave ample room for improvement. Note that the iHPA charges for this molecule are close to the ChElPG PACs. Here, we study a new idea: take PACs which are as close as possible to a reference set, for example the MPA, HPA or iHPA PACs, but insist that they reproduce exactly the components of the lowest ESM tensors (dipole and quadrupole) characterizing the molecular charge distribution. We formulate a straightforward method to determine such ``minimally-corrected PACs'' (section~\ref{sec:Method}) and then benchmark the results using a subset of molecules taken from the database of ref. \cite{Marenich2012} (section \ref{sec:Results}). Final conclusions are summarized in section~\ref{sec:Summary-and-conclusions}. All MPA, HPA and iHPA PACs, as well as the associated MOL-ESPs and MOL-ESMs were computed using developer versions of Q-Chem 4.3 and 4.4 \cite{Shao2015a} at the M06-L DFT level \cite{zhao2006new} and using the MG3 semi-diffuse (MG3S) basis set \cite{Lynch2003}. This functional/basis set combination was used for developing of the CM5 approach. The CM5, NPA and LPA results were taken from ref. \citenum{Marenich2012}. \section{\label{sec:Method}Method} Consider a molecule having $A$ nuclei at given Cartesian positions $\boldsymbol{R}^{a}=\left(R_{x}^{a},\,R_{y}^{a},\,R_{z}^{a}\right)$ ($a=1,\dots,A$), for which a QM calculation has determined the charge density $\rho\left(\boldsymbol{r}\right)$ of the molecule and from it, low order moments the charge $Q$, the dipole $\mu_{i}$ and the symmetric traceless quadrupole moment tensor $\Theta_{ij}$. Note that below, we use the notation $\Theta_{i}^{D}\equiv\Theta_{ii}$ for the diagonal elements of $\Theta$ and $\Theta_{i}^{OD}\equiv\Theta_{jk}$ where $i=x,y,z$ and $ijk$ is a cyclic permutation of $xyz$. For any set of PACs $\boldsymbol{q}=\left(q_{1},\dots,q_{A}\right)$ we define the PAC-ESMs as: the monopole (total charge) $Q^{PAC}\equiv e\sum_{a}q_{a}$, the dipole $\mu_{i}^{PAC}\equiv e\sum_{a}q_{a}R_{i}^{a}$ and the quadrupole $\Theta_{ij}^{PAC}\equiv e\sum_{a}q_{a}\left(3R_{i}^{a}R_{j}^{a}-\delta_{ij}\left(R^{a}\right)^{2}\right)$, ($i,j=x,y,z$) . Given a set of reference PACs $\boldsymbol{q}^{ref}$ we seek to determine a ``minimally-corrected'' set of PACs $\boldsymbol{q}^{mc}=\boldsymbol{q}^{ref}+\Delta\boldsymbol{q}$ such that that the size of the correction $\left\Vert \Delta\boldsymbol{q}\right\Vert ^{2}=\Delta\boldsymbol{q}\cdot\Delta\boldsymbol{q}$ is \emph{minimal }but the multipoles are equal to the QM determined multipoles, i.e. the following constraints are satisfied: \begin{align} c & =Q-Q^{PAC}=0,\nonumber \\ c_{i} & =\mu_{i}-\mu_{i}^{PAC}=0,\label{eq:Constraints}\\ c_{ij} & =\Theta_{ij}-\Theta_{ij}^{PAC}=0.\nonumber \end{align} Note that the number of constraints (denoted $C$) in Eq.~\eqref{eq:Constraints} is $9$ and not $13$ since the electric quadrupole tensor is symmetric and traceless. Point symmetries can reduce this number of constraints further. If, for example, both positive and negative charge densities are symmetric against the reflection through a plane (the x-y plane, for example) then there are 3 constraint less (one from the z component of the dipole and and 2 from XZ and YZ components of the quadrupole, which are zero by symmetry). Only when the number of atoms $A$ in the molecule is greater than the number of constraints $C$ can we hope to reproduce the constraints exactly. We therefore demand that $A>9$ and use the $A-C$ additional ``degrees of freedom'' to minimize the deviance $\Delta\boldsymbol{q}$. When $2\le A\le9$ we avoid the quadrupole moment constraint and use only the dipole moment constraint. We are led to consider the Lagrangian \begin{align} L_{mcDQ} & =\frac{1}{2}\sum_{a}\left(q_{a}^{mc}-q_{a}^{ref}\right)^{2}-\lambda c-\lambda_{i}c_{i}\label{eq:Lagrangian}\\ & -\sum_{xy,yz,zx}\lambda_{ij}^{OD}c_{ij}-\sum_{x,y,z}\lambda_{ii}^{D}c_{ii}\nonumber \end{align} as a function of the $A$ $q's$ and the ten Lagrange multipliers: one $\lambda,$ three $\lambda_{i}$'s , three diagonal $\lambda_{ii}^{D}$ and three off-diagonal$\lambda_{jk}^{OD}$ where $i=x,y,z$ and $ijk$ is a cyclic permutation of $xyz$. Taking derivatives with respect to these variables and equating to zero leads to the following set of $\left(10+A\right)$ linear equations in $\left(10+A\right)$ unknowns, given here in block-matrix/vector form\footnote{Since the matrix is dominated by zero's one can formulate the linear equation in a more concise way. However, this form is straightforward to derive and manipulate when there are instabilities, discussed later.}: \begin{align} S & \left(\begin{array}{c} q_{1}^{mc}\\ \vdots\\ q_{A}^{mc}\\ \lambda_{1\times1}\\ \lambda_{3\times1}\\ \lambda_{3\times1}^{D}\\ \lambda_{3\times1}^{OD} \end{array}\right)=\left(\begin{array}{c} q_{1}^{ref}\\ \vdots\\ q_{A}^{ref}\\ Q_{1\times1}\\ \mu_{3\times1}\\ \Theta_{3\times1}^{D}\\ \Theta_{3\times1}^{OD} \end{array}\right).\label{eq:Ax=00003Db} \end{align} The $\left(A+10\right)\times\left(A+10\right)$ matrix $S$ is of the following form: {\scriptsize{} \begin{equation} S=\left(\begin{array}{ccccccccccccc} \text{\textSFi} & \cdots & \text{\textSFiii} & -1 & \text{\textSFi} & & \text{\textSFiii} & \text{\textSFi} & & \text{\textSFiii} & \text{\textSFi} & & \text{\textSFiii}\\ \vdots & I_{A\times A} & \vdots & \vdots & \vdots & -D_{A\times3} & \vdots & \vdots & -T_{A\times3}^{D} & \vdots & \vdots & -T_{A\times3}^{OD} & \vdots\\ \text{\textSFii} & \cdots & \text{\textSFiv} & -1 & \text{\textSFii} & & \text{\textSFiv} & \text{\textSFii} & & \text{\textSFiv} & \text{\textSFii} & & \text{\textSFiv}\\ 1 & \cdots & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \text{\textSFi} & \cdots & \text{\textSFiii} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ & D_{3\times A} & & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \text{\textSFii} & \cdots & \text{\textSFiv} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \text{\textSFi} & \cdots & \text{\textSFiii} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ & T_{3\times A}^{D} & & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \text{\textSFii} & \cdots & \text{\textSFiv} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \text{\textSFi} & \cdots & \text{\textSFiii} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ & T_{3\times A}^{OD} & & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \text{\textSFii} & \cdots & \text{\textSFiv} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \end{equation}\end{array}\right)\label{eq:Smat} \end{equation} }and depends \emph{only }on the location of the atomic nuclei. The matrix is composed of blocks: the $I_{A\times A}$ block is a $A\times A$ unit matrix, $D_{A\times3}$, $T_{A\times3}^{D}$ and $T_{A\times3}^{OD}$ are matrices of dimension $A\times3$ ($3$ columns each of length $A$) of matrix elements: $D_{ai}=eR_{i}^{a}$ and $T_{ai}^{D}=e\left(3R_{i}^{a}R_{i}^{a}-\Vert R^{a}\Vert^{2}\right)$ for $a=1,\dots,A$ and $i=x,y,z$ and and $T_{ai}^{OD}=3eR_{j}^{a}R_{k}^{a}$ (where the ordered set $i,j,k$ is a cyclic permutation of $x,y,z$). The $D_{3\times A}$, $T_{3\times A}^{D}$, and $T_{3\times A}^{OD}$ blocks are respectively the transposed matrices. The $A+10$ column-vector on the left-hand-side of Eq.~\ref{eq:Ax=00003Db} includes the unknowns, the $A$ partial charges $q_{a}^{mc}$ and the $\lambda's$, the ten Lagrange multipliers for the ten constraints. The $10+A$ column-vector on the right-hand-side has $A$ values of the reference charges $q_{a}^{ref}$, followed by the total charge on the molecule $Q$, then the three values of the QM dipole moment $\mu_{i}$ followed by the three values of the diagonal elements of the given QM quadrupole tensor $\Theta_{i}^{D}=\Theta_{ii}$ and finally the QM values of the three off-diagonal elements $\Theta_{i}^{OD}=\Theta_{jk}$ where $i=x,y,z$ and $ijk$ is a cyclic permutation of $xyz$). A similar equation holds for the mcD method, where the six last rows are erased from $S$ and from the column vectors and the six right columns are erased from $S$ as well. This leaves us with a $\left(A+4\right)\times\left(A+4\right)$ system of equations. The structural matrix $S$ may become singular or rank deficient. One trivial source for singularity the use of 3 diagonal constraints while their sum is composed to be zero. The use of the singular-value-decomposition pseudo-inverse \cite{Golub1996} for solving Eq.~\ref{eq:Ax=00003Db} helps to bypass such a singularity. A more delicate source of singularities may arise from symmetry. For example, when the molecule is perfectly planar (or has a plane of symmetry) in the x-y plane then the row corresponding to the dipole in the z direction $D_{az}=eR_{z}^{a}$ must be identically zero and the matrix $S$ will be rank deficient. In this case the $T_{ai}^{OD}$ with $i=x$and $y$ must also be zero). In these cases the SVD pseudoinverse will automatically eliminate constraints that cannot be met due to this kind of symmetry. But for near-symmetrical configurations, instabilities may exist. In cases such as these we can still spot problems by examining the values of the Lagrange multipliers $\lambda,$$\lambda_{i}$ and $\lambda_{ij}$in the solution vector of Eq.~\ref{eq:Ax=00003Db}. The Lagrange multiplier is equal to the derivative of the minimal value of the Lagrangian $L$ with respect to the constraint value ($Q$, $\mu_{i}$ and $\Theta_{ij}$, respectively). Thus if the \emph{ab initio }dipole moment $\mu_{x}$ is given to precision $\delta\mu_{x}$, the product $\left|\lambda_{x}\delta\mu_{x}\right|$is expected to be the error in the minimal value of $L$. Clearly, the minimizing procedure is meaningless unless this error is much smaller than 1. Hence, it is important to eliminate ``offending'' constraints from the matrix equation (the corresponding row and column in the matrix and the entry in the column vectors) for those having large Lagrange multipliers. We know that \emph{ab initio }multipole properties are usually given to 3 digits hence we eliminate constraints corresponding to Lagrange multipliers large than 1000. The reduced equation is then solved and the remaining Lagrange multipliers are examined again. We repeat such elimination until all Lagrange multipliers have proper magnitudes. This pruning procedure helps avoid cases where small inaccuracies of the input data dominate the final result. Within the molecules studied here such a pruning procedure was used only for few cases of molecules having a near plane symmetry. When symmetry is active, our procedures reduce the number of constraints and hence the number of \emph{independent }$q_{a}$'s (called number of degrees of freedom (NDOFs)). For example, the water molecule has 3 nuclei but due to symmetry the two H nuclei will have the same PACs and so NDOF=2. Due to the symmetry only the dipole moment in the direction of the $C_{2}$ axis is a constraint (the components perpendicular to the C2 axis are zero by symmetry). Together with the charge of water (0) we already have 2 constraints so one must give up imposing the quadrupole moment for water. \begin{figure}[h] \includegraphics[width=1\columnwidth]{1-fluoro-4-nitrobenzene-MPA-mcD-MPA-mcDQ-MPA-graphs} \caption{\label{fig:Efficacy}The PACs (top panels) and and ESP correlation plots (bottom panels) for MPA, mcD-MPA and mcDQ-MPA (left panels) and iHPA, mcD-iHPA and mcDQ-iHPA (right panels) in the 1-fluoro-4-nitrobenzene molecule. The mean absolute relative deviations (MARD) appears in parenthesis near each PAC method. } \end{figure} \section{\label{sec:Results}Results } To demonstrate the efficacy of the method we show in Fig.~\ref{fig:Efficacy} the MPA and iHPA PACs and their ESP correlation plots before and after applying the minimal corrections required for imposing dipole and quadrupole moments (denoted mcD/mcDQ-MPA and mcD/mcDQ-iHPA respectively).\footnote{Minimally-corrected PACs that reproduce only the dipole moment are designated mcD and those that reproduce the components of the dipole and the quadrupole ESMs are designated mcDQ.} Notice that the MPA-ESP has low correlation with the MOL-ESP, as can be evident visually and also by the reported MARD of 2. The mcD corrections improve the ESP but only mcDQ corrections show high quality ESP (with MARD of 0.05). In accordance with previous reports,\cite{VanDamme2009} the iHPA ESP already correlates nicely with the MOL-ESP (MARD of 0.16) but the mcDQ-iHPA improves the correlation significantly and the MARD reduces by a factor of 4. For this molecule, both mcDQ-MPA and mcDQ-iHPA have similar MARDs but this is not typical, for most molecules the mcDQ-iHPA MARDs are much smaller than those of mcDQ-MPA (see Fig.\ref{fig:HPA-Histog}). The mcDQ-MPA PACs are not drastically different from the MPA PACs yet their MARDs are considerably lower. This shows the power of the minimally-corrected PACs, where a small change in PACs can improve the PAC based ESP considerably. In Fig.~\ref{fig:HPA-Histog} we display a log-scale bar-plot of MARDs of several PAC-based potentials on selected molecules containing 10-18 atoms. Each PAC method can be characterized by a pair of numbers (shown in parenthesis within the legend box) indicating the median/maximal MARD taken over the given set of molecules. The PACs obtained by minimally-correcting the $q\equiv0$ reference (called 0PA) are actually the minimal PACs that give the dipole and quadrupole of the molecules. It is seen that their correlation with the exact ESP is considerably higher than that of MPA and HPA, somewhat similar to that of mcDQ-MPA and mcDQ-HPA, close to that of CM5. This goes to show that the fit of just the dipole and quadrupole, keeping the charges as small as possible gives a reasonably behaved ESP, although in general, for very large molecules the mcDQ-0PA performance may degrade with size compared to the PA methods. We see that MPA and HPA have similar MARDs while iHPA seems to give considerably smaller MARDs (by a factor 2-3). The minimal corrected (mcDQ) to MPA and HPA yield smaller MARDs by a factor 4 and for iHPA by a factor 2. Altogether the mcDQ significantly improves the ESP. The mcDQ-iHPA median MARD is 7\% is similar to that of ChElPG (5\%). It is worthwhile to examine the sensitivity of the MARD estimation with respect to the distance of grid points from the nearest nuclei. In Fig.~\ref{fig:HPA-Histog} all sampling grid points were at a distance larger than $1.5\times r_{vdW}$ from any atom. When MARD is estimated using points further way (distance larger than a value of $2\times r_{vdW}$) the iHPA MARD dropped from 0.14 to 0.09 and mcDQ-iHPA MARD dropped from 0.07 to 0.03. ChElPG MARD also reduced, from 0.05 to 0.03. This finding is consistent with the fact that the MCDQ methods provide an asymptotically exact far-field ESP resulting from their reconstruction of the molecular dipole and quadrupole moments. \begin{figure} \includegraphics[width=1\textwidth]{histog-10-18_mcDQ_MPA_mcDQ-MPA_HPA_mcDQ-HPA_iHPA_mcDQ-iHPA_CM5_ChElPG} \caption{\label{fig:HPA-Histog}The MARD (Eq.~\eqref{eq:MARD}) of various PACs schemes for a subset of molecule taken from ref. \cite{Marenich2012}. Numbers in parenthesis appearing near the molecule names indicate the number of atoms in that molecule. The pair of numbers (median/max) appearing in the legend box near each scheme is, respectively, the median and maximum of the relative deviance taken over the shown set molecules.} \end{figure} In Table~\ref{tab:dq} we show, for each set of PACs the magnitude of the charge correction $\left\Vert \Delta q\right\Vert _{\infty}$. For a given molecule the mcD correction is largest for 0PA and then for MPA and HPA and it is smallest for iHPA. mcDQ corrections are in general considerably larger than mcD but in both methods $\left\Vert \Delta q\right\Vert _{\infty}$ decreases as the number of atoms in the molecule grows. This is due to the fact that in large systems even small charge shifts have a large affect on the dipole and the quadrupole moments. \begin{table} {\scriptsize{}}% \begin{tabular}{|l|l|l|l|l|r@{\extracolsep{0pt}.}l|r@{\extracolsep{0pt}.}l|r@{\extracolsep{0pt}.}l|r@{\extracolsep{0pt}.}l|r@{\extracolsep{0pt}.}l|r@{\extracolsep{0pt}.}l|r@{\extracolsep{0pt}.}l|r@{\extracolsep{0pt}.}l|} \hline \multirow{2}{*}{{\scriptsize{}Molecule}} & \multirow{2}{*}{{\scriptsize{}Sym}} & \multirow{2}{*}{{\scriptsize{}$A$}} & \multirow{2}{*}{{\scriptsize{}$F$}} & \multirow{2}{*}{{\scriptsize{}$C$}} & \multicolumn{8}{c|}{{\scriptsize{}$\left\Vert \Delta q\right\Vert _{\infty}$ (mcD)}} & \multicolumn{8}{c|}{{\scriptsize{}$\left\Vert \Delta q\right\Vert _{\infty}$ (mcDQ)}}\tabularnewline \cline{6-21} & & & & & \multicolumn{2}{c|}{{\scriptsize{}0PA}} & \multicolumn{2}{c|}{{\scriptsize{}MPA}} & \multicolumn{2}{c|}{{\scriptsize{}HPA}} & \multicolumn{2}{c|}{{\scriptsize{}iHPA}} & \multicolumn{2}{c|}{{\scriptsize{}0PA}} & \multicolumn{2}{c|}{{\scriptsize{}MPA}} & \multicolumn{2}{c|}{{\scriptsize{}HPA}} & \multicolumn{2}{c|}{{\scriptsize{}iHPA}}\tabularnewline \hline \hline {\scriptsize{}Pyridazine} & {\scriptsize{}$C_{2v}$} & {\scriptsize{}10} & {\scriptsize{}7} & {\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}11} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}06} & {\scriptsize{}0}&{\scriptsize{}04} & {\scriptsize{}0}&{\scriptsize{}27} & {\scriptsize{}0}&{\scriptsize{}29} & {\scriptsize{}0}&{\scriptsize{}17} & {\scriptsize{}0}&{\scriptsize{}12}\tabularnewline \hline {\scriptsize{}Ethylamine} & {\scriptsize{}$C_{s}$} & {\scriptsize{}10} & {\scriptsize{}7} & {\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}04} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}1}&{\scriptsize{}17} & {\scriptsize{}1}&{\scriptsize{}61} & {\scriptsize{}1}&{\scriptsize{}22} & {\scriptsize{}1}&{\scriptsize{}28}\tabularnewline \hline {\scriptsize{}Acetoacetic acid} & {\scriptsize{}$C_{1}$} & {\scriptsize{}10} & {\scriptsize{}9} & {\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}07} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}54} & {\scriptsize{}0}&{\scriptsize{}26} & {\scriptsize{}0}&{\scriptsize{}25} & {\scriptsize{}0}&{\scriptsize{}21}\tabularnewline \hline {\scriptsize{}Acetone} & {\scriptsize{}$C_{2v}$} & {\scriptsize{}10} & {\scriptsize{}5} & {\scriptsize{}5} & {\scriptsize{}0}&{\scriptsize{}17} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}52} & {\scriptsize{}0}&{\scriptsize{}48} & {\scriptsize{}0}&{\scriptsize{}34} & {\scriptsize{}0}&{\scriptsize{}08}\tabularnewline \hline {\scriptsize{}Ethylene-glycol} & {\scriptsize{}$C_{1}$} & {\scriptsize{}10} & {\scriptsize{}10} & \textbf{\scriptsize{}9} & {\scriptsize{}0}&{\scriptsize{}09} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}67} & {\scriptsize{}0}&{\scriptsize{}88} & {\scriptsize{}0}&{\scriptsize{}64} & {\scriptsize{}0}&{\scriptsize{}54}\tabularnewline \hline {\scriptsize{}Cyclopentadienone} & {\scriptsize{}$C_{2v}$} & {\scriptsize{}10} & {\scriptsize{}6} & \textbf{\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}09} & {\scriptsize{}0}&{\scriptsize{}04} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}00} & {\scriptsize{}0}&{\scriptsize{}19} & {\scriptsize{}0}&{\scriptsize{}22} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}02}\tabularnewline \hline {\scriptsize{}Oxetane} & {\scriptsize{}$C_{s}$} & {\scriptsize{}10} & {\scriptsize{}7} & {\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}08} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}1}&{\scriptsize{}09} & {\scriptsize{}1}&{\scriptsize{}29} & {\scriptsize{}1}&{\scriptsize{}09} & {\scriptsize{}1}&{\scriptsize{}09}\tabularnewline \hline {\scriptsize{}3-Imino-2,3-dihydroisoxzole} & {\scriptsize{}$C_{s}$} & {\scriptsize{}10} & {\scriptsize{}8} & \textbf{\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}79} & {\scriptsize{}0}&{\scriptsize{}83} & {\scriptsize{}0}&{\scriptsize{}69} & {\scriptsize{}0}&{\scriptsize{}41}\tabularnewline \hline {\scriptsize{}Lithium-dimethylamine} & {\scriptsize{}$C_{2v}$} & {\scriptsize{}10} & {\scriptsize{}6} & \textbf{\scriptsize{}4} & {\scriptsize{}0}&{\scriptsize{}31} & {\scriptsize{}0}&{\scriptsize{}16} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}04} & {\scriptsize{}0}&{\scriptsize{}31} & {\scriptsize{}0}&{\scriptsize{}16} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}04}\tabularnewline \hline {\scriptsize{}Methylacetate} & {\scriptsize{}$C_{s}$} & {\scriptsize{}11} & {\scriptsize{}9} & \textbf{\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}10} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}53} & {\scriptsize{}0}&{\scriptsize{}22} & {\scriptsize{}0}&{\scriptsize{}26} & {\scriptsize{}0}&{\scriptsize{}01}\tabularnewline \hline {\scriptsize{}Pyridine} & {\scriptsize{}$C_{2v}$} & {\scriptsize{}11} & {\scriptsize{}7} & \textbf{\scriptsize{}4} & {\scriptsize{}0}&{\scriptsize{}06} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}15} & {\scriptsize{}0}&{\scriptsize{}23} & {\scriptsize{}0}&{\scriptsize{}09} & {\scriptsize{}0}&{\scriptsize{}03}\tabularnewline \hline {\scriptsize{}2-Cyanopyridine} & {\scriptsize{}$C_{s}$} & {\scriptsize{}12} & {\scriptsize{}12} & \textbf{\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}12} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}04} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}19} & {\scriptsize{}0}&{\scriptsize{}19} & {\scriptsize{}0}&{\scriptsize{}07} & {\scriptsize{}0}&{\scriptsize{}03}\tabularnewline \hline {\scriptsize{}1-fluoro-4-nitrobenzene} & {\scriptsize{}$C_{2v}$} & {\scriptsize{}14} & {\scriptsize{}9} & \textbf{\scriptsize{}6{*}} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}06} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}13} & {\scriptsize{}0}&{\scriptsize{}20} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}04}\tabularnewline \hline {\scriptsize{}Morpholine} & {\scriptsize{}$C_{s}$} & {\scriptsize{}15} & {\scriptsize{}9} & \textbf{\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}47} & {\scriptsize{}0}&{\scriptsize{}04} & {\scriptsize{}0}&{\scriptsize{}33} & {\scriptsize{}0}&{\scriptsize{}08}\tabularnewline \hline {\scriptsize{}Quinoline} & {\scriptsize{}$C_{s}$} & {\scriptsize{}17} & {\scriptsize{}17} & \textbf{\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}12} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}10} & {\scriptsize{}0}&{\scriptsize{}26} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}02}\tabularnewline \hline {\scriptsize{}Fluorocyclohexane (A)} & {\scriptsize{}$C_{s}$} & {\scriptsize{}18} & {\scriptsize{}12} & \textbf{\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}02} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}17} & {\scriptsize{}0}&{\scriptsize{}06} & {\scriptsize{}0}&{\scriptsize{}05} & {\scriptsize{}0}&{\scriptsize{}02}\tabularnewline \hline {\scriptsize{}Fluorocyclohexane (E)} & {\scriptsize{}$C_{s}$} & {\scriptsize{}18} & {\scriptsize{}12} & \textbf{\scriptsize{}6} & {\scriptsize{}0}&{\scriptsize{}04} & {\scriptsize{}0}&{\scriptsize{}03} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}01} & {\scriptsize{}0}&{\scriptsize{}17} & {\scriptsize{}0}&{\scriptsize{}08} & {\scriptsize{}0}&{\scriptsize{}07} & {\scriptsize{}0}&{\scriptsize{}04}\tabularnewline \hline \multicolumn{5}{|l|}{\textbf{\scriptsize{}Median}} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}07} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}03} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}03} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}01} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}31} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}23} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}17} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}04}\tabularnewline \hline \multicolumn{5}{|l|}{\textbf{\scriptsize{}Max}} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}31} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}16} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}06} & \textbf{\scriptsize{}0}&\textbf{\scriptsize{}04} & \textbf{\scriptsize{}1}&\textbf{\scriptsize{}17} & \textbf{\scriptsize{}1}&\textbf{\scriptsize{}61} & \textbf{\scriptsize{}1}&\textbf{\scriptsize{}22} & \textbf{\scriptsize{}1}&\textbf{\scriptsize{}28}\tabularnewline \hline \end{tabular}{\scriptsize \par} \caption{\label{tab:dq}The PAC change $\left\Vert \Delta q\right\Vert _{\infty}=max_{1\le a\le A}\left|\Delta q_{a}\right|$ induced by mcD and mcDQ for 0PA (where the reference PACs are all zero), MPA, HPA and iHPA for the set of molecules of Fig.~\ref{fig:HPA-Histog}. Also shown the number of atoms $A$ the number of degrees of freedom $F$ and the number of constraints $C$ for each molecule.} \end{table} In table~\ref{tab:Bitzuim} we summarize the MARD statistics (median and maximal) for for four sets of reference charges: 0PA (reference charges are equal to zero) and MPA, HPA, iHPA. The efficiency of the mc procedure is apparent for MPA, HPA and iHPA, where the mcD reduces the median/maximal MARD by about a factor of 2. mcDQ reduces the MARD further, by a factor of 3 for 0PA and \textasciitilde{}2 for MPA and HPA and only 1.1 for iHPA. We thus see that iHPA reconstruction of the ESP strongly benefits from a dipole correction and, interestingly, much less a quadrupole correction. \begin{table} \begin{tabular}{|c|c|c|c|c|} \hline & 0PA & MPA & HPA & iHPA\tabularnewline \hline \hline no-correction & NA & 0.39/1.90 & 0.38/0.78 & 0.14/0.32\tabularnewline \hline mcD & 0.41/0.67 & 0.18/0.60 & 0.18/0.53 & 0.08/0.17\tabularnewline \hline mcDQ & 0.12/0.26 & 0.10/0.28 & 0.11/0.19 & 0.07/0.15\tabularnewline \hline \end{tabular} \caption{\label{tab:Bitzuim}The median/maximal MARD (for the set of molecules used above) determined for each PAC reference found for: non-corrected, and minimally-corrected schemes, mcD and mcDQ . } \end{table} PACs are sometimes used when molecules distort. In this case, it is important that the they remain continuous under the distortion, so as to enable force calculations. The MPA/HPA/iHPA do not show non-smooth behavior and the mcDQ which is a minimization procedure does not show it as well.\footnote{We cannot rule out possible issues if the matrix $S$ of Eq.~\ref{eq:Smat} becomes rank deficient. However, we believe this is an unlikely or quite rare event.} In \ref{fig:q_O(phi)} we show the MPA, mcDQ-MPA and ChElPG PACs of the oxygen atom in N-methylethanamide\cite{Hu2007} the as a function of the dihedral angle $\phi$. It is seen that as the angle increases from 0 the mcDQ-MPA PAC slightly decreases and then increase rapidly followed by a rapid yet continuous drop near $\phi_{c}=0.25$ from a value of $q_{O}=-0.25$ to $q_{O}\approx-0.64$. An additional very sharp feature is seen near $\phi=\pi$. We have checked that this sharp feature is not discontinuous (see inset in Fig.~\ref{fig:q_O(phi)}) and that the matrix $S$ of Eq.~\ref{eq:Smat} does not become rank deficient. Similar behavior is seen for the PACs of other atoms. We thus conclude that the charges change continuously although sometimes very rapid charge fluctuations can occur. \begin{figure} \includegraphics[width=0.7\columnwidth]{MPA-angle} \caption{\label{fig:q_O(phi)}The partial charge, determined by MPA, mcDQ-MPA and ChElPG on the oxygen atom as a function of the O-C-N-H dihedral angle $\phi$ in N-methylethanamide (NMA). Inset om the right shows the sharp feature near $\phi=\pi$. } \end{figure} \section{\label{sec:Summary-and-conclusions}Summary and conclusions} We have studied a new scheme for minimally correcting reference PACs so that they reproduce the exact dipole and quadrupole moments of a molecule and we found that such a minimal correction greatly improves the correlation of the PAC-ESP with respect to the MOL-ESP. The minimal correction scheme does not alter symmetry properties of the reference PACs. Hence, minimally-corrected PACs (mc-PACs) based on MPA, HPA, iHPA fully respect the point-symmetry and rotational/translational symmetries of the molecule. An additional benefit of the mc-PACs is their stability for inner (or buried) atoms of large molecules. This rises from the stability of the standard population schemes themselves and the fact that mc-PACs involve rather small corrections. As an example, consider the 2-(Dimethylamino)-2-propanol molecule: \noindent\begin{minipage}[t]{1\columnwidth}% \begin{center} \includegraphics[width=0.4\textwidth]{c20-fig} \par\end{center}% \end{minipage}\\ for which the ChElPG, HPA and mcDQ-HPA PACs are shown in Fig.~\ref{fig:Burial}. Here, ChElPG tends to polarize the molecule: the oxygen and nitrogen share between them a negative unit charge and this is counteracted by the positive unit charge of the central carbon atom C3. On the other hand, MPA assigns a low charge for C3 and spreads rather evenly the remaining positive charge on the 12 terminal hydrogen atoms. mc-MPA charges are very close to those of MPA and thus yet they improve significantly the ESP description for this molecule: the MPA MARD is 0.35 while that of the mc-MPA is 0.1. It is worthwhile to note that the PACs assigned by ChElPG also have a MARD of 0.1. When the underlying reference is the iHPA set of PACs the resulting ESP is of similar quality to that of the ChElPG set of PACs resulting from a best-fit to ESPs. The dependence of the PACs on the molecular distortion was demonstrated to have sometimes very sharp features however all the changes were smooth, hence forces can be calculated on the atoms of the molecule. The method here bears a similarity to the optimal point-charge model of Ref.~\cite{Simmonett2005} which determines PACs that reproduce as many low-order moments as possible. The crucial difference is best seen when systems grow, model of Ref.~\cite{Simmonett2005} would target increasingly higher electrostatic moments as more atoms are included while the present method targets multipoles up to second order and not beyond, thereby avoiding the numerical instabilities described in see Ref.~\cite{Gilbert2006}. On the other hand. the optimal point-charge model treats the multipole constraints in a more systematic way by minimizing the error over unused moments in the last incomplete spherical shell. \begin{figure} \includegraphics[width=0.75\columnwidth]{c20} \caption{\label{fig:Burial}The ChElPG, MPA and mc-MPA PACs for the 2-(Dimethylamino)-2-propanol } \end{figure} \section*{Acknowledgments} Authors express special thanks to Dr. Yihan Shao from Q-CHEM Inc. for his advice and critical assistance in performing the iHPA calculations. We also gratefully acknowledge the support of the Israel Science Foundation Grant No. 189/14. %\bibliographystyle{unsrt} \bibliography{EfratBibLib} \end{document} }\end{equation}}
\caption[]{Sketch of the continuous (left) and discrete (right) Riemannian de Casteljau algorithm defining a continuous and a discrete cubic B\'ezier curve, respectively (the algorithm proceeds as follows:\tikz{\node[circle, fill=black, inner sep=2pt] at (0,0) {}; \node[circle, fill=blue, inner sep=2pt] at (0.3,0) {}; \node[circle, fill=red, inner sep=2pt] at (0.6,0) {}; \node[circle, fill=green, inner sep=2pt] at (0.9,0) {};}).}
\caption[]{Sketch of the construction of the control shells $\control_{3j-1}$ and $\control_{3j+1}$ for the cardinal spline in the continuous (left) and discrete (right) setting (the algorithm proceeds as follows: \tikz{\node[circle, fill=black, inner sep=2pt] at (0,0) {}; \node[circle, fill=blue!30, inner sep=2pt] at (0.3,0) {}; \node[circle, fill=red, inner sep=2pt] at (0.6,0) {}; \node[circle, fill=green, inner sep=2pt] at (0.9,0) {};}).}
\caption[]{Sketch of the construction for the interpolatory binary four-point scheme in the continuous case (left) and the discrete case (right) (the algorithm proceeds as follows: \tikz{\node[circle, fill=black, inner sep=2pt] at (0,0) {}; \node[circle, fill=blue, inner sep=2pt] at (0.3,0) {}; \node[circle, fill=red, inner sep=2pt] at (0.6,0) {}; \node[circle, fill=green, inner sep=2pt] at (0.9,0) {};}).}
\caption{Results of triaging content severity. Numbers are percentages. \flagged category is \amber $\cup$ \red $\cup$ \crisis. \urgent category is \red $\cup$ \crisis. F1 is F1-Score and Acc is Accuracy. Baseline is the SVM classifier on post body (unigram and bigram features). Table (a) presents classification results and comparison with the baseline and state-of-the-art on the test set. Table (b) shows classification results on training set based on 10-fold stratified cross validation. For Table (b), $\dagger$($\ddagger$) shows statistically significant improvement over the baseline (all other methods in the Table) according to the Student's t-test ($p<0.02$).}
\caption{Effect of each set of features on triaging based on the test set. Numbers show percentages of macro averaged results for the \flagged categories (\crisis $\cup$ \red $\cup$ \amber). Acc: Accuracy, F1: F1-score, P: Precision, R: Recall. Body is the textual body of the post; ``skip thought'' is dense representation of text using skip thought vectors, ``meta'': forum metadata features; ``subj'': subjectivity features; ``topic'': Topic modeling features extracted using LDA, ``\liwc'': Linguistic Inquiry and Word Count features. Plus (+) signs show that the feature is added to the features in the above row and minus ($-$) signs show that the feature is eliminated from the above row. The row shown with ($\ast$) indicates the features (listed in Table \ref{tab:features}) used in the single model in tables \ref{tab:classification-res} and \ref{tab:per-category}. Accordingly, the last row is the ensemble model.}
\caption[Entrainment]{ Time-averaged entrainment, defined as vertical velocity conditional averaged over negative velocities ($\widetilde{U_y^-}/U_\infty$) for the $St_H = 0.21$ case (a). The entrainment through the line $y/H=0.9$ for all cases as function of $x/H$ (b). $St_H=0.21$~({\color[rgb]{0,0.447,0.741} \rule[.4ex]{1em}{2pt}}), $St_H=0.48$~({\color[rgb]{0.850,0.325,0.098} \rule[.4ex]{1em}{2pt}}), $St_H = 0.85$~({\color[rgb]{0.929,0.694,0.125} \rule[.4ex]{1em}{2pt}}) and ${St_H=1.98}$~({\color[rgb]{0.494,0.184,0.556} \rule[.4ex]{1em}{2pt}}) (colour online), the arrow indicates increasing Strouhal number. \label{fig:entr_circ_1} }
\caption[]{Reattachment length as function of total entrainment for ${St_H=0.21}$~({\color[rgb]{0,0.447,0.741}\tiny $\bm{\mathrm{\Delta}}$}), $St_H=0.48$~({\color[rgb]{0.850,0.325,0.098}\tiny $\bm{\nabla}$}), $St_H = 0.85$~({\color[rgb]{0.929,0.694,0.125}$\bm{\triangleleft}$}) and ${St_H=1.98}$~({\color[rgb]{0.494,0.184,0.556}$\bm{\triangleright}$}). \label{fig:xr_entr} }
\caption[Vorticity]{ Time-averaged positive vorticity ($\widetilde{\omega'_+}/\omega_0$) with swirling strength applied for the $St_H = 0.21$ case (a). Time-averaged circulation per unit length ($\Gamma^*$) as function of $x/H$ for all cases (b). $St_H=0.21$~({\color[rgb]{0,0.447,0.741} \rule[.4ex]{1em}{2pt}}), $St_H=0.48$~({\color[rgb]{0.850,0.325,0.098} \rule[.4ex]{1em}{2pt}}), $St_H = 0.85$~({\color[rgb]{0.929,0.694,0.125} \rule[.4ex]{1em}{2pt}}) and ${St_H=1.98}$~({\color[rgb]{0.494,0.184,0.556} \rule[.4ex]{1em}{2pt}}) (colour online), the arrow in (b) indicates increasing Strouhal number. \label{fig:entr_circ_2} }
\caption[Vortex tracking]{ Comparison of tracked location (a) and absolute circulation (b) between positive~({\color[rgb]{0,0,1}\tiny $\bm{+}$}) and negative~({\color[rgb]{1,0,0}\tiny $\bm{\times}$}) vortex for the $St_H = 0.21$ case. \label{fig:Track_posneg} }
\caption[]{ Tracked vortex trajectories and size for ${St_H=0.21}$~({\color[rgb]{0,0.447,0.741}\tiny $\bm{\mathrm{\Delta}}$}), $St_H=0.48$~({\color[rgb]{0.850,0.325,0.098}\tiny $\bm{\nabla}$}), $St_H = 0.85$~({\color[rgb]{0.929,0.694,0.125}$\bm{\triangleleft}$}) and ${St_H=1.98}$~({\color[rgb]{0.494,0.184,0.556}$\bm{\triangleright}$}). Horizontal vortex location as function of time (a). Vortex radius as used for scaling (b) the black line indicates a decrease with 1/f. Scaled vortex location and time (c) the black line indicates the trajectory given by Equation~\ref{eq:x_t}. Normalized horizontal convection velocity of the vortices as function of scaled horizontal location (d) the black line indicates the convection velocity given by Equation~\ref{eq:V_t}. \label{fig:Trajectory} }
\caption[]{ (Absolute) circulation tracked as function of time for $St_H=0.21$~({\color[rgb]{0,0.447,0.741}\tiny $\bm{\mathrm{\Delta}}$}), $St_H=0.48$~({\color[rgb]{0.850,0.325,0.098}\tiny $\bm{\nabla}$}), $St_H = 0.85$~({\color[rgb]{0.929,0.694,0.125}$\bm{\triangleleft}$}) and ${St_H=1.98}$~({\color[rgb]{0.494,0.184,0.556}$\bm{\triangleright}$}). The data is plotted on (a) linear and (b) log-log axes. The black trend-lines in (b) show a decay of $t^{-2}$. Normalized circulation as function of scaled time on linear (c) and log-log axes (d). The black lines indicate the decay of circulation as given by Equation~\ref{eq:Gscaling}. Dashed lines indicate the values at $t/T = 0.5$, 1.0 and 1.5. \label{fig:Circulation} }
\caption[Normalized vorticity and entrainment]{ Time-averaged circulation per unit width (a) and entrainment through $y/H=0.9$ (b) as function of $x/L$. $St_H=0.21$~({\color[rgb]{0,0.447,0.741} \rule[.4ex]{1em}{2pt}}), $St_H=0.48$~({\color[rgb]{0.850,0.325,0.098} \rule[.4ex]{.35em}{2pt}\hspace{.2em}\rule[.4ex]{.35em}{2pt}}), $St_H = 0.85$~({\color[rgb]{0.929,0.694,0.125} \rule[.4ex]{.4em}{2pt}\hspace{.1em}\rule[.4ex]{.15em}{2pt}\hspace{.1em}\rule[.4ex]{.25em}{2pt}}) and ${St_H=1.98}$~({\color[rgb]{0.494,0.184,0.556} \rule[.4ex]{.2em}{2pt}\hspace{.2em}\rule[.4ex]{.2em}{2pt}\hspace{.2em}\rule[.4ex]{.2em}{2pt}}) (colour online). \label{fig:circ_entr_norm} }
\caption{Time series plot of R{\"o}ssler Attractor: (a) t-x plane, (b) t-y plane, and (c) t-z plane. Shows sensitivity to initial conditions with $x_0=1$ (\textcolor{blue}{blue}) and $x_0=1.001$ (\textcolor{red}{red}).}
\caption{Colloidal rollers in random obstacle lattices. (a) Top panel: Superimposed pictures taken at {\em equal time intervals} of a colloidal roller deflected by a lithographied post of radius $b=10\,\rm\mu m$. Note that the direction of motion is changed at constant speed. Bottom panel: Radial density of colloidal rollers propelling around an isolated obstacle. Circles: Experiments. Dark lines: Simulated radial densities. Solid line: $B\ell/v_0=5$, dashed line: $B\ell/v_0=100$ as defined in \eqref{equationv}. {\color{bleu}$\ell$ is defined as the value where the density plateaus. In all our experiments we find $\ell\sim 2b$}. {\color{bleu}Error bars: binning size.} (b), (c) and (d) Trajectories of colloidal rollers (red and yellow) superimposed to the pictures of the obstacle lattices. Scale bar: $500\,\rm\mu m$. Total time: $300\,s$. (b) $\rho= 0.21/(\pi b^2)$, the trajectories form a single percolating cluster. (c) $\rho = 0.45/(\pi b^2)$, the trajectories form disconnected clusters. The largest cluster (in red) percolates through the observation region. (d) $\rho = 0.89/(\pi b^2)$, none of the disconnected clusters percolate, and no macroscopic transport is observed. The largest cluster of maximal dimension $l_{\rm c}$ is colored in red. (e) Variations of the normalized maximal cluster size with the obstacle density. $L$ is the width of the observation window. The dashed line indicates the critical density $\rho_{\rm c}$. {\color{bleu}Experimental errors on the determination of the cluster sizes are smaller than the figure markers. Defining a statistical error on this extremal quantity would require a number of independent realisations beyond our experimental reach.} }
\caption{ {From active diffusion to localization.} (a) Probability distribution function of the roller speed for three obstacle densities. $\rho= 0.21/(\pi b^2)$, $\rho = 0.45/(\pi b^2)$ and $\rho = 0.89/(\pi b^2)$. The three distributions are peaked at the same typical speed value: $v_0=250\,\rm \mu m/s$. (b) The angular diffusivity $D$ increases with the obstacle fraction. {\color{bleu} $D$ is measured as the inverse of the time of half decorrelation of the velocity autocorrelation shown in the inset. Error bars: 1 sd.} Inset: Autocorrelation function of the instantaneous orientation of the rollers velocity ${\rm VACF}=\langle\hat {\mathbf v}(t_0)\cdot\hat {\mathbf v}(t_0+t) \rangle_{t_0}$. (c) Mean squared displacements of the rollers as a function of time. The colors indicate the obstacle density. The rollers are localized in finite regions at high obstacle densities. (d) Variations of the dynamical exponent $\alpha$ defined as $\Delta r^2\sim t^\alpha$. $\alpha$ is estimated using power-law fits of the mean squared displacements. {\color{bleu} For each obstacle density, independent fits have been performed in 7 intervals of width $v_0 t/(2b)=50$ in the shaded region in (c). $\alpha$ represents the mean of the fitted exponents and error bars represents one $\sigma$.} The vertical dashed line indicates the value of the critical density $\rho_{\rm c}$ defined in the last section. The horizontal dashed lines indicates the value $\alpha=0.66$ corresponding to an ideal overlapping Lorentz gas.}
\caption{Median predictive performance of each model created. Each row corresponds to the performance statistics of a dataset split by a certain time threshold. Cells marked with \textcolor{carmine}{{\bf red}} indicate ``bad'' results; i.e. false alarms over 33\% or precision or recall results under 33\%.}
\caption{\label{figure1}Image features computation: (\textit{a}) length and width (red \dotted and thick \full), core-azimuthal length and width (same blue lines); (\textit{b}) core-azimuthal half-image width.}
\caption{\label{figure2}Selection quality factor vs shower core distance. (\textit{a})~O~500~TeV, (\textit{b})~Fe~500~TeV, (\textit{c})~Fe~200~TeV. Green \fullcircle -- image kurtosis, red \fullcircle -- image width, blue \fullcircle -- half-image width, blue \opencircle -- core-azimuthal half-image width. Solid and dotted lines are guide to the eyes.}
\caption{Methodology of Scalable Secondary Studies (MOSSS). Note the red bubbles denoting the location of our three research questions \textcolor{red}{{\bf RQ1, RQ2, RQ3}} within this framework.}
\caption{The computed errors for $E_1((\nabla_s\sigma)_x)$ $({\color{black} \bullet})$, $E_{\infty}((\nabla_s\sigma)_x)$ $({\color{black} \blacktriangle})$, $E_1((\nabla_s\sigma)_y)$ (\Smblacksquare), and $E_{\infty}((\nabla_s\sigma)_y)$ $({\color{black} +})$. The order of convergence for $(\nabla_s\sigma)_x$ is $1.9$ and $1.5$ for $E_1$ (\solidrule) and $E_\infty$ (\dotsrule) errors respectively, and the order of convergence for $(\nabla_s\sigma)_y$ is $1.7$ and $1.4$ for $E_1$ (\dashedrule) and $E_\infty$ (\dashdotrule) errors respectively. The symbols represent the errors from the computations and the lines show the linear fits.}
\caption{ The computed error for $E_1((\nabla_s \sigma)_x)$ $({\color{black} \bullet})$, $E_{\infty}((\nabla_s \sigma)_x)$ $({\color{black} \blacktriangle})$, $E_1((\nabla_s \sigma)_y)$ (\Smblacksquare), and $E_{\infty}((\nabla_s \sigma)_y)$ $({\color{black} +})$. The order of convergence for $(\nabla_s \sigma)_x$ is $0.85$ and $0.019$ $E_1$ (\solidrule) and $E_\infty$ (\dotsrule) errors respectively, and the order of convergence for $(\nabla_s \sigma)_y$ is $0.86$ and $0.045$ for $E_1$ (\dashedrule) and $E_\infty$ (\dashdotrule) errors, respectively. The symbols represent the errors from the computations, and the lines show the linear fit of those points.}
\caption{The computed errors for $L_1((\nabla_s \sigma)_x)$ $({\color{black} \bullet})$, $L_1((\nabla_s \sigma)_y)$ $({\color{black} \blacktriangle})$, and $L_1(T)$ (\Smblacksquare). The order of convergence for $(\nabla_s \sigma)_x$ and $(\nabla_s \sigma)_y$ is $1.2$ (\solidrule) and $1.4$ (\dashedrule), respectively. The order of convergence of $T(x,y)$ at the interface is $1.2$ (\dotsrule). The symbols represent the errors from the computations and the lines show the linear fit of those points.}
\caption{ Errors of the $x$ component of the surface gradient at the interfacial cells for $\Delta=a/8$ $({\color{black} \bullet})$, $a/32$ $({\color{black} \blacktriangle})$, $a/128$ (\Smblacksquare), and $a/512$ $({\color{black} +})$. Initializing the temperature using (a) equation \eqref{eq:temp45} and (b) equation \eqref{eq:temp_centroid}. $\theta$ is defined to be zero at the positive $x$ axis and increasing counterclockwise. }
\caption{The computed errors for $E_1((\nabla_s \sigma)_x)$ $({\color{black} \bullet})$, $E_{\infty}((\nabla_s \sigma)_x)$ $({\color{black} \blacktriangle})$, $E_1((\nabla_s \sigma)_y)$ (\Smblacksquare), and $E_{\infty}((\nabla_s \sigma)_y)$ $({\color{black} +})$. The order of convergence for for $(\nabla_s \sigma)_x$ is $0.94$ (\solidrule) and $0.65$ (\dashedrule) for $E_1$ and $E_\infty$ errors, respectively, and the order of convergence for for $(\nabla_s \sigma)_y$ is $0.89$ (\dotsrule ) and $0.58$ (\dashdotrule), respectively. The symbols represent the errors from the computations, and the lines show the linear fit of those points.}
\caption{Drop migration velocity for $\text{Re} = \text{Ma} = 0.72$ and $\text{Ca} = 0.0576$ for $\Delta=$ $1/64$ ({\color{red}\solidrule}), $1/256$ ({\color{green_darker}\solidrule}), and $1/256$ ({\color{cyan}\solidrule}) compared with the result given in \cite{Ma2011} ({\color{purple}\solidrule}) for 2D simulations. }
\caption{Convergence of the migration velocity as a function of the time step for $\Delta t=$ $10^{-4}$ ({\color{red}\solidrule}), $5\times10^{-6}$ ({\color{yellow}\solidrule}), $10^{-5}$ ({\color{green_darker}\solidrule}), and $5\times10^{-6}$ ({\color{cyan}\solidrule}) compared with the results in the results\cite{herrmann2008} ({\color{purple}\solidrule}) for 2D simulation, for $\text{Re} = \text{Ca} = 0.066$ and $\text{Ma} = 0 $. The velocity is rescaled by $v^*_{ygb}$.}
\caption{Convergence of the migration velocity in a 3D simulation with mesh refinement for $\Delta =$ $1/16$ ({\color{red} \solidrule}), $1/32$ ({\color{green}\solidrule}), and $1/64$ ({\color{cyan}\solidrule}), and $1/128$ ({\color{purple}\solidrule}); $\text{Re} = \text{Ma} = 0.72$ and $\text{Ca} = 0.0576 $.}
\caption{Upper panels show magnetization $M$ as a function of magnetic field at different temperatures for the six samples. Lower panels show the extracted penetration fields $H_p$ for opposite orientations of magnetic field by taking the crossing point of M(H) at a given temperature with the dashed lines shown in the panels. Black dash lines are \textcolor{green}{empirical} fits to $H_{p}(T) = H_{p}(0)\,(1-(T/T_c)^{\alpha})$.}
\caption{\small Test set accuracy and std error on \textbf{left:} WDW when initialized with off-the-shelf GloVe embeddings of different sizes, \textbf{right:} CBT-NE when initialized with embeddings trained on BT corpus after removing a fraction of stopwords ({\color{red} red}), or using different window sizes ({\color{green} green}).}
\caption{The lowest three moments $\mu_n^{u-d}$ with $n=1, 3$ and $5$. The blue pluses ($\textcolor{blue}{+}$) are the target numbers, $6\,\mu_n=\int_0^1 dx x^n \left[u(x)-d(x)\right]$. The red crosses ($\textcolor{red}{\times}$) are the solution of the OPE (\ref{va}). For the error estimate see the text. At $n=3,5$ the error is smaller than the symbol.} \label{fig3} \end{figure} The mathematical methods described above enable us to obtain the lower moments of the structure function, as well as $F_1(x)$ as a whole, rather accurately from a relatively small set of values of $T_{33}(\omega)$. To demonstrate that, we start from the experimental nonsinglet structure function $F_1^{u-d}(x)$, which we parameterize by \begin{equation} 6\hspace*{0.2mm} x\hspace*{0.25mm} F_1^{u-d}(x) = x\left[u(x)-d(x)\right] \label{mstw} \end{equation} with $u(x)$ and $d(x)$ taken to be the LO parton distributions at the scale $q^2=1\,\mbox{GeV}^2$~\cite{Martin:2009iq}. From (\ref{mstw}) we compute $T_{33}(\omega)$ for $\omega=0.1$, $0.2$, $\cdots$, $0.9$ using (\ref{opess2}). %The desired $\omega$ values may be obtained, for example, from photon and %nucleon momenta $\vec{q}=(2,6,0)$ and $\vec{p}=(1,0,0)$, $(2,0,0)$, %$(0,1,0)$, $(1,1,0)$, $(2,1,0)$, $(0,2,0)$, $(1,2,0)$, $(2,2,0)$, $(3,2,0)$, %both in units of $2\pi n/aL, n \in \mathbf{N}$, with $n$ tuned to the %requested value of $q^2$. The result is shown in Fig.~\ref{fig2}. This we consider our `data', from which we want to retrieve the moments $\mu$ and structure function $F_1^{u-d}(x)$, and to which the lattice results (see Fig.~\ref{fig6}) will aspire to. %A calculation of $T_{33}(\omega)$ for $O(10)$ values of $\omega$ spread over %the interval $[0,1]$ appears to be realistic. %The `experimental' (target) values of moments and structure function are %shown by the blue pluses in Fig.~\ref{fig2} and the solid curve in %Fig.~\ref{fig3}, respectively. \begin{figure}[!t] \begin{center} \epsfig{file=F1e.ps,width=10cm,clip=} \end{center} \vspace*{-0.75cm} \caption{The structure function $F_1^{u-d}(x)$. The solid line is the target structure function, $\displaystyle 6xF_1(x) = x \left[u(x)-d(x)\right]$. The crosses ($\textcolor{red}{\times}$) are the solution of the SVD (\ref{svd}). For the error estimate see the text.}
\caption{The structure function $F_1^{u-d}(x)$ obtained from the Mellin transform of (\ref{me}) fitted to the moments (\textcolor{red}{$-$}), compared with the target structure function $\displaystyle 6xF_1(x) = x \left[u(x)-d(x)\right]$ (\textcolor{blue}{$-$}). For the error estimate see the text.} \label{fig5} \end{figure} To retrieve the structure function $F_1^{u-d}(x)$, we apply the SVD (\ref{svd}) to our `data'~\cite{ma}. Three out of nine eigenvalues of $W$ turned out to be zero. In Fig.~\ref{fig4} we show the result for $M = 19$. It shows that the structure function $F_1^{u-d}$ can be well reproduced from a relatively small set of data, except perhaps for $x \lesssim 0.05$. Similar results are obtained for the singlet structure function $F_1^{u+d+\bar{u}+\bar{d}+\bar{s}}$. We have not made any attempts to optimize the SVD. It can be improved in several respects. A Bayesian approach~\cite{Liu:2016djw} to alleviate overfitting, for example, might lead to particularly robust results. There are other possibilities as well to compute the structure function from the Compton amplitude. A particularly promising approach is to fit the moments, for example in the interpolating polynomial (\ref{poly}), by an appropriate function $\mu(s)$ with $\mu(n)=\mu_n$ and employ an inverse Mellin transform on $\mu(s)$ to obtain $F_1(x)$. It turns out that the moments can be fitted surprisingly well by the simple expression \begin{equation} \mu(s)=A\,(s+\alpha)^{-\beta}\,, \label{me} \end{equation} for which the inverse Mellin transform is known analytically~\cite{er}. Starting from the moments $6\,\mu_n=\int_0^1 dx x^n \left[u(x)-d(x)\right]$, the result of the Mellin transform is shown in Fig.~\ref{fig5}. %To estimate the error, we have added an error of $10\%$ to the individual %moments and performed the Mellin transform on the error envelope. The result %is shown by the shaded area. %Alternatively, one can invert (\ref{opess2}) directly using the polynomial %fit (\ref{poly}) as input. That can be done either by an inverse Stieltjes %transform~\cite{er,db} or by using the Maximum Entropy Method %~\cite{Nakahara:1999vy}. The latter method has proven to be very successful %in computing hadron spectral functions from lattice two-point functions. The analysis so far has been limited to $\omega \in[0,1]$. The SVD method can be extended to larger values $\omega > 1$ without problem. This will allow us to probe the small-$x$ region of $F_1(x)$, which is not accessible through moments of the structure function. Indeed, by extending the calculation to $\omega = 2$, we were able to retrieve the singlet structure function $F_1^{u+d+\bar{u}+\bar{d}+\bar{s}}(x)$~\cite{Martin:2009iq} down to fractional momenta $x \lesssim 0.005$, which was not possible before. Odd moments of the structure functions can be obtained by also including the local axial vector current $\bar{\psi}_f(x)\gamma_3 \gamma_5 \psi_f(x)$ to (\ref{add}) and studying the interference with the vector current. This is achievable through a simple extension of the procedure described above. The method can be generalized to nonforward Compton scattering as well. That will allow us to derive generalized parton distribution functions (GPDs). \begin{figure}[!b] \begin{center} \epsfig{file=T33L.ps,width=10cm,clip=} \end{center} \vspace*{-0.75cm} \caption{The proton Compton amplitude $T_{33}(p,q)$ for momenta $\vec{p}=(2,-1,0)$, $(-1,1,0)$, $(1,0,0)$, $(0,1,0)$, $(2,0,0)$, $(-1,2,0)$, $(1,1,0)$, $(0,2,0)$, $(2,1,0)$, $(1,2,0)$, from left to right, and $\vec{q}=(3,5,0)$, in lattice units. The current has been attached to the $d$ quark, leading to the `handbag' diagram in Fig.~\ref{fig1}. $Z_V$ has been taken from~\cite{Constantinou:2014fka}. The solid line shows a sixth order polynomial fit (giving $\chi^2/{\rm dof}=0.9$), and the shaded area shows the error.} \label{fig6} \end{figure} There is the question what accuracy can be achieved with real data. It turns out that the second derivative of the nucleon energy can be computed rather accurately. In a proof-of-principle study we have computed (\ref{fh}) from $O(900)$ configurations generated at the SU(3) symmetric point~\cite{Bietenholz:2011qq} on a $32^3\times 64$ lattice with lattice spacing $a\approx 0.074\,\mbox{fm}$. First results are presented in Fig.~\ref{fig6}, where the contribution from $\displaystyle \mathcal{F}_1(0,q^2) = \left.- 2\,E_\lambda(0,q)\, \partial^2 E_\lambda(0,q)/\partial \lambda^2\,\right|_{\lambda=0}$ has been subtracted. The precision for lattice momenta $\vec{p}^2=1$ and $2$ is already quite impressive. We should be able to improve on the precision of the data at higher momenta by employing `momentum smearing' techniques~\cite{Bali:2016lva}, which has not been attempted here. Based on this result, we consider an overall projected error of $10\%$ on the Compton amplitude $T_{33}$, marked by the shaded area in Fig.~\ref{fig2}, a conservative estimate. The resulting errors on $\mu_n$ and $F_1^{u-d}(x)$, shown in Figs.~\ref{fig3}, \ref{fig4} and \ref{fig5}, have been obtained from replacing the initial values of $T_{33}(\omega_i)$ by the corresponding numbers on the confidence envelopes. This leads us to conclude that the entire structure function, including its moments, can be reconstructed from a lattice calculation of the Compton amplitude with unprecedented accuracy, devoid of any renormalization and mixing issues. %To answer that, we add a $10\%$ overall error to our `data' points %$T_{33}(\omega_i)$, as indicated in Fig.~\ref{fig2}. We consider this a %reasonable estimate in view of our first real result on $T_{33}$ shown in %Fig.~\ref{fig6}. To assess the error on the moments and structure function, %we replace the initial values of $T_{33}(\omega_i)$ by the values on the %error envelope and proceed as before. The resulting errors on $\mu_n$ and %$F_1^{u-d}(x)$ are shown in Figs.~\ref{fig3} and \ref{fig4}, respectively. %In the case of real and wiggly data a higher order polynomial (\ref{poly}) %might not be a suitable fit function. In this case a smooth function might be %the better choice. %The moments are rather insensitive to errors and number of the data. If one %is only interested in the lowest three moments, $\mu_1, \mu_3$ and $\mu_5$, %five data points with step size $\epsilon \approx 0.2$ would be sufficient. \section*{Acknowledgement} GS thanks Akaki Rusetsky for useful discussions. The numerical calculations were carried out on the IBM BlueGene/Qs at Edinburgh and J\"ulich using DIRAC 2 and NIC resources, on the Cray XC30 at HLRN, Berlin and Hannover, and on the NCI National Facility at Canberra. HP and GS are supported by DFG Grant Nos.\SCHI 422/10-1 and SCHI 179/8-1. PELR is supported by the STFC under contract ST/G00062X/1. RDY and JMZ are supported by the Australian Research Council Grant Nos.\FT120100821, FT100100005 and DP140103067.%Allows to compute higher twist contributions by varying $q^2$. %\section{The hypercubic group} \begin{thebibliography}{99} \bibitem{Martinelli:1996pk} G.~Martinelli and C.T.~Sachrajda, %``On the difficulty of computing higher twist corrections,'' Nucl.\Phys.\B{\bf 478} (1996) 660 [hep-ph/9605336]. \bibitem{dislat} For pioneering work see: G.~Martinelli and C.T.~Sachrajda, %``A Lattice Study of Nucleon Structure,'' Nucl.\Phys.\B{\bf 316} (1989) 355, M.~G\"ockeler, R.~Horsley, E.-M.~Ilgenfritz, H.~Perlt, P.E.L.~Rakow, G.~Schierholz and A.~Schiller,%``Polarized and unpolarized nucleon structure functions from lattice QCD,'' Phys.\Rev.\D{\bf 53} (1996) 2317 [hep-lat/9508004].\\ For recent reviews of the subject see: M.~Constantinou, %``Recent progress in hadron structure from Lattice QCD,'' PoS CD {\bf 15} (2015) 009 [arXiv:1511.00214 [hep-lat]]; H.-W.~Lin, %``From C to Parton Sea: Bjorken-x Dependence of the PDFs,'' arXiv:1612.09366 [hep-lat]; and references cited therein. \bibitem{Martinelli:1998hz} G.~Martinelli, %``Hadronic weak interactions of light quarks,'' Nucl.\Phys.\Proc.\Suppl.\{\bf 73} (1999) 58 [hep-lat/9810013]. \bibitem{op} S.~Capitani, M.~G\"ockeler, R.~Horsley, H.~Oelrich, D.~Petters, P.E.L.~Rakow and G.~Schierholz,%``Towards a nonperturbative calculation of DIS Wilson coefficients,'' Nucl.\Phys.\Proc.\Suppl.\{\bf 73} (1999) 288 [hep-lat/9809171]; S.~Capitani, M.~G\"ockeler, R.~Horsley, D.~Petters, D.~Pleiter, P.E.L.~Rakow and G.~Schierholz,%``Higher twist corrections to nucleon structure functions from lattice QCD,'' Nucl.\Phys.\Proc.\Suppl.\{\bf 79} (1999) 173 [hep-ph/9906320]; W.~Bietenholz, N.~Cundy, M.~G\"ockeler, R.~Horsley, H.~Perlt, D.~Pleiter, P.E.L.~Rakow, G.~Schierholz, A.~Schiller, T.~Streuer and J.M.~Zanotti,%``A Non-Perturbative Operator Product Expansion,'' PoS LAT {\bf 2009} (2009) 138 [arXiv:0910.2437 [hep-lat]]. \bibitem{Caracciolo} S.~Caracciolo, A.~Montanari and A.~Pelissetto, Nucl. Phys. B (Proc.~Suppl.) {\bf 73} (1999) 273, ibid.\{\bf 83-84} (2000) 875, JHEP {\bf 9} (2000) 45; G.C.~Rossi, Chin.\J.\Phys.\{\bf 38} (2000) 721. \bibitem{Detmold:2005gg} W.~Detmold and C.J.D.~Lin, %``Deep-inelastic scattering and the operator product expansion in lattice QCD,'' Phys.\Rev.\D{\bf 73} (2006) 014501 [hep-lat/0507007]. \bibitem{dis} R.~Devenish and A.~Cooper-Sarkar, {\it Deep Inelastic Scattering}, Oxford University Press (2003, Oxford, UK); A.V.~Manohar, %``An Introduction to spin dependent deep inelastic scattering,'' in {\it Lake Louise 1992, Symmetry and Spin in the Standard Model}, p.\1-46[hep-ph/9204208]; K.-F.~Liu, %``Parton degrees of freedom from the path integral formalism,'' Phys.\Rev.\D{\bf 62} (2000) 074501 [hep-ph/9910306]. \bibitem{Agadjanov:2016cjc} See, for example: A.~Agadjanov, U.-G.~Mei{\ss}ner and A.~Rusetsky, %``The nucleon in a periodic magnetic field,'' arXiv:1610.05545 [hep-lat]. \bibitem{Horsley:2012pz} R.~Horsley, R.~Millo, Y.~Nakamura, H.~Perlt, D.~Pleiter, P.E.L.~Rakow, G.~Schierholz, A.~Schiller, F.~Winter, J.M.~Zanotti, %``A Lattice Study of the Glue in the Nucleon,'' Phys.\Lett.\B{\bf 714} (2012) 312 [arXiv:1205.6410 [hep-lat]]; A.J.~Chambers, R.~Horsley, Y.~Nakamura, H.~Perlt, D.~Pleiter, P.E.L.~Rakow, G.~Schierholz, A.~Schiller, H.~St\"uben, R.D.~Young and J.M.~Zanotti,%``Feynman-Hellmann approach to the spin structure of hadrons,'' Phys.\Rev.\D{\bf 90} (2014) 014510 [arXiv:1405.3019 [hep-lat]]; A.~J.~Chambers, R.~Horsley, Y.~Nakamura, H.~Perlt, D.~Pleiter, P.E.L.~Rakow, G.~Schierholz, A.~Schiller, H.~St\"uben, R.D.~Young and J.M.~Zanotti,%``Disconnected contributions to the spin of the nucleon,'' Phys.\Rev.\D{\bf 92} (2015) no.11, 114517 [arXiv:1508.06856 [hep-lat]]; A.J.~Chambers, J.~Dragos, R.~Horsley, Y.~Nakamura, H.~Perlt, D.~Pleiter, P.E.L.~Rakow, G.~Schierholz, A.~Schiller, K.~Somfleth, H.~St\"uben, R.D.~Young and J.M.~Zanotti,%``Electromagnetic form factors at large momenta from lattice QCD,'' arXiv:1702.01513 [hep-lat]. \bibitem{Bakeyev:2003ff} T.~Bakeyev, M.~G\"ockeler, R.~Horsley, D.~Pleiter, P.E.L.~Rakow, G.~Schierholz and H. St\"uben,%``Nonperturbative renormalization and improvement of the local vector current for quenched and unquenched Wilson fermions,'' Phys.\Lett.\B{\bf 580} (2004) 197 [hep-lat/0305014]. \bibitem{nr} W.H. Press, B.P. Flannery, S.A. Teukolsky and W.T. Vetterling, {\it Numerical Recipes}, Cambridge University Press (1989, Cambridge, UK). \bibitem{ma} Wolfram Research, Inc., {\it Mathematica}, Version 11.0 (2016, Champaign, USA). \bibitem{Martin:2009iq} A.D.~Martin, W.J.~Stirling, R.S.~Thorne and G.~Watt, %``Parton distributions for the LHC,'' Eur.\Phys.\J.\C{\bf 63} (2009) 189 [arXiv:0901.0002 [hep-ph]]. \bibitem{Liu:2016djw} K.-F.~Liu, %``Parton Distribution Function from the Hadronic Tensor on the Lattice,'' PoS LATTICE {\bf 2015} (2016) 115 [arXiv:1603.07352 [hep-ph]]. \bibitem{er} A. Erd\'{e}lyi, {\it Table of Integral Transforms}, Vol. I and II, McGraw-Hill (1954, New York, USA). \bibitem{Bietenholz:2011qq} W.~Bietenholz, V.~Bornyakov, M.~G\"ockeler, R.~Horsley, W.G.~Lockhart, Y.~Nakamura, H.~Perlt, D.~Pleiter, P.E.L.~Rakow, G.~Schierholz, A.~Schiller, T.~Streuer, H.~St\"uben, F.~Winter and J.M.~Zanotti,%``Flavour blindness and patterns of flavour symmetry breaking in lattice simulations of up, down and strange quarks,'' Phys.\Rev.\D{\bf 84} (2011) 054509 [arXiv:1102.5300 [hep-lat]]. \bibitem{Constantinou:2014fka} M.~Constantinou, R.~Horsley, H.~Panagopoulos, H.~Perlt, P.E.L.~Rakow, G.~Schierholz, A.~Schiller and J.M.~Zanotti, %``Renormalization of local quark-bilinear operators for $N_f$=3 flavors of stout link nonperturbative clover fermions,'' Phys.\Rev.\D{\bf 91} (2015) no.1, 014502 [arXiv:1408.6047 [hep-lat]]. \bibitem{Bali:2016lva} G.~S.~Bali, B.~Lang, B.~U.~Musch and A.~Sch\"afer,%``Novel quark smearing for hadrons with high momenta in lattice QCD,'' Phys.\Rev.\D{\bf 93} (2016) no.9, 094515 [arXiv:1602.05525 [hep-lat]]. %\bibitem{db} %L. Debnath and D. Bhatta, {\it Integral Transforms and their Applications}, %CRC Press (2015, Boca Raton, USA). %\bibitem{Nakahara:1999vy} % Y.~Nakahara, M.~Asakawa and T.~Hatsuda, % %``Hadronic spectral functions in lattice QCD,'' % Phys.\ Rev.\ D {\bf 60} (1999) 091503 % [hep-lat/9905034]. \end{thebibliography} \end{document} \begin{table}[!b] \begin{center} \vspace*{0.25cm} \begin{tabular}{rrrrl|r} \multicolumn{5}{c|}{$\vec{p}$} & \multicolumn{1}{c}{$\omega$} \\ \hline (\hspace*{-0.35cm} & 2 & -1 & 0 &\hspace*{-0.35cm}) & $\displaystyle 2/34$ \\[0.75em] (\hspace*{-0.35cm} &-1 & 1 & 0 &\hspace*{-0.35cm}) & $\displaystyle 4/34$ \\[0.75em] (\hspace*{-0.35cm} & 1 & 0 & 0 &\hspace*{-0.35cm}) & $\displaystyle 6/34$ \\[0.75em] (\hspace*{-0.35cm} & -2 & 2 & 0 &\hspace*{-0.35cm}) & $\displaystyle 8/34$ \\[0.75em] (\hspace*{-0.35cm} & 0 & 1 & 0 &\hspace*{-0.35cm}) & $\displaystyle 10/34$ \\[0.75em] (\hspace*{-0.35cm} & 2 & 0 & 0 &\hspace*{-0.35cm}) & $\displaystyle 12/34$ \\[0.75em] (\hspace*{-0.35cm} & -1 & 2 & 0 &\hspace*{-0.35cm}) & $\displaystyle 14/34$ \\[0.75em] (\hspace*{-0.35cm} & 1 & 1 & 0 &\hspace*{-0.35cm}) & $\displaystyle 16/34$ \\[0.75em] (\hspace*{-0.35cm} & -2 & 3 & 0 &\hspace*{-0.35cm}) & $\displaystyle 18/34$ \\[0.75em] (\hspace*{-0.35cm} & 0 & 2 & 0 &\hspace*{-0.35cm}) & $\displaystyle 20/34$ \\[0.75em] (\hspace*{-0.35cm} & 2 & 1 & 0 &\hspace*{-0.35cm}) & $\displaystyle 22/34$ \\[0.75em] (\hspace*{-0.35cm} & -1 & 3 & 0 &\hspace*{-0.35cm}) & $\displaystyle 24/34$ \\[0.75em] (\hspace*{-0.35cm} & 1 & 2 & 0 &\hspace*{-0.35cm}) & $\displaystyle 26/34$ \\[0.75em] (\hspace*{-0.35cm} & 3 & 1 & 0 &\hspace*{-0.35cm}) & $\displaystyle 28/34$ \\[0.75em] (\hspace*{-0.35cm} & 0 & 3 & 0 &\hspace*{-0.35cm}) & $\displaystyle 30/34$ \\[0.75em] (\hspace*{-0.35cm} & 2 & 2 & 0 &\hspace*{-0.35cm}) & $\displaystyle 32/34$ \end{tabular} \end{center} \caption{Possible hadron momenta for $\vec{q}=(3\; 5\; 0)$, with $q^2 = 34$, in units of $2\pi n/aL$, $n \in \mathbbm{N}$.} \label{tab1} \end{table} \end{document} }
\caption{\label{fig:detcontainment} The distribution of relative efficiency of neutron capture and energy deposition of the positron and the annihilation \gray shown across the detector volume for each detector plane and for one detector plane at position 11 close the middle of the volume.}
\caption{Demonstration of a pulse discrimination method based on the number of samples above a chosen threshold and the pulse amplitude for a detector cell exposed to an AmBe source. The horizontal band corresponds to PVT and ZnS signals produced by recoil protons and \gray. Neutron events can be seen on the upper part of the figure. \label{fig:neutronPID}}
\caption{The residual energy correction $E_{\mathrm{corr}}$ for positrons of kinetic energy (1, 3, 5)\,MeV. The red histograms show the correction to the visible energy obtained by integrating over all detector voxels in a 1m$^3$ detector and the blue histograms show the corrections when the visible energy is reconstructed by summing only the two most energetic deposits in the detector. \label{fig:EresComp} }
\caption{Panoramic view of the MUSE spectra. Top panels: the zoomed region centered on ID14a,b,c,d,e is shown ($6.8''$ wide). From left to right we show the MUSE images at the \lya\(panel A), at 1548 \AA~(panel B), in the HST/F814W band (panel C), and the gray-coded magnification map (panel D), in which the yellow contours enclose the region with $\mu>25$. The arc-shaped black line marks the aperture used to extract the MUSE spectrum as the sum of images ID14a,b,c. The black dashed circle identifies the image ID14e and the magenta ellipse the diffuse \lya\emission at the same redshift as ID14 on top of the critical line of the galaxy cluster. Panel (B) shows the rectangular red aperture with the same width and position angle of the X-Shooter slit (also shown in panel C, top-right, with black dotted thin lines). The orange dotted circle in panel C marks the position of a possible counterpart of the diffuse\lya\emission (magenta ellipse). The blue circle ($1.2''$ diameter) marks the aperture used to derive the spectrum for image ID14a only. Panel (E), shows the MUSE spectrum of ID14a,b,c (black line) extracted using the arc-shaped aperture. The continuum of the elliptical galaxies E1+E2 has been subtracted (white thick line) and the high ionization lines are shown in the corresponding insets. The colors of the spectra in the insets corresponds to the colors of the apertures shown in panel B. The flux scale of all images is $F(\lambda)$ [$10^{-18}$ \erg]. \label{muse}}
\caption{The continuum-subtracted data-cube at the \civblue\wavelength position is shown on the left panel ($6.5''$ wide). The blue circles mark the multiple images of ID14 detected on the HST color image, reported in the right panel. In particular, faint line emissions arise at the position of multiple images ID14d and ID14e. \label{CIV}}
\caption{HST/ACS and WFC3 cutouts ($1.5''$ wide) centered on the diffuse \lya\emission marked with the magenta ellipse. The dashed-orange circle marks the position of the faint$m \sim 30$ (or $m>34$ de-lensed) underlying object if placed at the same redshift of the \lya\emission.\label{orange}}
\caption{User Reviews by Categories. N = 7,396,551. {\small\textcolor{caribbeangreen}{\ensuremath\blacksquare}} = Positive, {\small\textcolor{lightgray}{\ensuremath\blacksquare}} = Neutral, {\small\textcolor{carminepink}{\ensuremath\blacksquare}} = Negative, {\small\textcolor{darkgray}{\ensuremath\blacksquare}} = Undefined.}
\caption{{\bf Illustration of double- and triple-slit interference and generic atomic analogues.} {\bf (a)} Optical double-slit interference experiment where the interference pattern has \emph{one} sinusoidal component, {\bf (b)} Double-slit-type interference setup in an atomic system. The time-domain measurement of fluorescence intensity will show show oscillations with \emph{one} beat note {\bf (c)} Optical triple-slit interference experiment where the interference pattern has \emph{three} sinusoidal components. {\bf (d)} \blue2{Simplest configuration of an analogue} triple-slit-type interference setup in an atomic system. The time-domain measurement of fluorescence intensity will show oscillations with \emph{three} beat notes. In (b) and (d), the photon emerges from a superposition of excited states with different energies, where $|e_i\rangle$ ($i=1,2,3$) represent the excited states and $|g\rangle$ the ground state. The energy level structure depicted here is for illustrative purposes only and is not an exact representation of the multilevel configuration in \blue2{our~QD}. }
\caption{{\bf Double-slit-type excitonic interference: FSS quantum beats. (a)} Resonance-fluorescence (RF) detuning spectra in CW mode for $X^{0}$, and {\bf (b)} $X^{1-}$ for comparison. The filled black circles and lines are experimental data and Lorentzian fits, respectively. The insets show the corresponding electron and hole spin configurations which have degenerate energies at zero external magnetic field for $X^{1-}$, effectively making it a two-level system. For $X^{0}$ the electron-hole exchange interaction causes a fine-structure splitting (FSS) of \blue2{$\sim13\mu$eV} at $B=0$ in the QD under study. {\bf (c) } Quantum erasure scheme. Excitation and detection polarisation configuration leading to quantum erasure of which-path information originally encoded in the polarization of photons emerging from $|e_1\rangle$, $|e_2\rangle$. % {\bf (d)} Quantum-beat frequency versus external magnetic field. {\bf (e$-$h)} {$\bf X^0$ %Time-resolved resonance fluorescence (TRRF) quantum beats} under $\pi$-pulse excitation with 100-ps and $\sim5\mu$eV temporal and spectral width, observed at various FSS values manipulated by an external magnetic field $B$. The fits to the data are of the form $I(t)\propto e^{t/T_1}[A+B \cos(\delta_0 t)]$ \cite{Flissikowski:2001hx}. \blue2{Time}-resolved measurements were performed with a $\sim$$150$-ps-resolution detection setup.}
\caption{{\bf Triple-slit-type excitonic quantum interference: FSS and RS combined quantum beats}. \blue2{Time}-resolved resonance fluorescence (TRRF) for {\bf (a,b)} driven $X^0$ exciton (V-system) showing triple-slit-type interference, combining beats due to both Rabi splitting (RS) and fine-structure splitting (FSS), and {\bf (c,d)} the case of a driven $X^{1-}$ (2-level system) showing double-slit-type interference based only on RS quantum beats, shown for comparison. %(b) and (d) are higher-excitation-power versions of (a) and (b), respectively. The solid red lines are fits of sinusoidal oscillations to the data having one beat component in (c,d) and three beat components with respective frequencies $|\Omega\pm\delta_0/2|$ and $\Omega$ in (a,b). The insets illustrate the effective three- and two-slit energy level configurations\blue2{, with the dressed states defined as $|\pm\rangle \equiv |g,n+1\rangle\pm |e_{(1)},n\rangle,~|\pm_g\rangle \equiv |g,n\rangle\pm |e_{(1)},n-1\rangle $, where $n$ is the photon number.} {\bf (e,f)} show Rabi frequencies $\Omega$ extracted from corresponding fits to the data as a function of excitation power for $X^{0}$ (with FSS \blue2{$\hbar\delta_0=13 \mu$eV}) and $X^{1-}$, respectively. The presence of multiple sinusoidal components evidenced by clear deviation from single sinusoidal oscillations in the time-resolved fluorescence for $X^{0}$ indicates genuine multi-slit type interference. }
\caption{{\bf Numerical simulation of TRRF showing RS and FSS quantum beats as well as their combination}. Simulation of TRRF using the master-equation method with {\bf (a)} 100-ps $\pi$ pulse, $X^{1-}$ {\bf (b)} 100-ps $\pi$ pulse, $X^{0}$ {\bf (c)} 2-ns pulse, $X^{1-}$ {\bf (d)} 2-ns pulse, $X^{0}$. Here \blue2{$\Omega/(2\pi)=1.3$} GHz. }
\caption{Parameter sets used in the simulations {\color{red} update numbers}.}
\caption{We show in a) qualitative results of our person matcher on PathTrack. False positives are even challenging for humans. b) Evolution of matching accuracy for different amounts of training trajectories. Training on the 15,380 trajectories of PathTracks results in an accuracy of 88\%, reducing the misclassification rate by 45\%, compared to MOT15. c) Person-matching accuracy for different annotation times using path supervision (\textcolor{blue}{blue}) or exhaustive LabelMe annotation (\textcolor{red}{red}). A high-quantity annotation strategy with our path supervision provides the best accuracy for the same annotation-time budget.}
\caption{Partial snapshots of (a) forward-moving and (b) backward-moving sphere-dimer motors comprising catalytic (\textcolor[rgb]{0.9,0.6,0.0}{orange}) and noncatalytic spheres (\textcolor[rgb]{0.8,0.4,0.0}{vermillion}) in a solvent consisting of fuel (not shown for clarity) and product molecules (\textcolor[rgb]{0.0,0.45,0.7}{blue}).}
\caption{The set $\set{C}^\varphi$ of all feasible affine disturbance feedback policies ({\color{blue}\textbf{solid}}) and its inner approximation $\overbar{\set{C}}^\varphi$ ({\color{lightred}\textbf{dashed}}). Both are parametrized by $\theta$.}
\caption{Trajectories of controlled car corresponding to random disturbances ({\color{lightred}\textbf{solid}}), best-case ({\color{blue}\textbf{dotted}}) and worst-case ({\color{blue}\textbf{dashed}}) disturbance.}
\caption{Control effort $\|\mathbf{u}\|_2$ for different disturbance realizations ({\color{blue}\textbf{solid}}), best ({\color{red}\textbf{circle}}), worst ({\color{red}\textbf{square}}) and open-loop case ({\color{blue}\textbf{dashed}}).}
\caption{Flow of information across three time steps: The \textcolor{perception_color}{perception modules} continuously infer the state of the robot and the world $\boldsymbol{s}$ from sensory input $\boldsymbol{y}$. The \textcolor{lrc_color}{\lrc} immediately translates this world state into a local policy $\boldsymbol{\pi^l}$. The \textcolor{cmo_color}{continuous motion optimization} computes a plan $\boldsymbol{\pi^g}$ for some time-horizon at a slightly lower rate. \textcolor{rp_color}{\Rp} combines these two policies into one policy $\boldsymbol{\pi}$, which enables it to immediately react to local changes, and to look ahead in time to react to larger changes. Finally, this policy $\boldsymbol{\pi}$ produces a control output $\boldsymbol{u}$ (omitted for readability) for which is sent to the robot. \label{fig:system} }
\caption{Variance $V_N^{DAPI}$ by the filter constant $\tau$ for a ring graph with 50 nodes. \textcolor{Green}{Edit figure and caption!}}
\caption{ (Color online) The linear dependence of global polarization, $2\langle\Pi_{0y}\rangle_{p}$, as a function of impact parameter ratio $b_0$ at 11.5 GeV, 27.0 Gev and 62.4 GeV. {\color{red} %Panel (b) is the extensive quantity type of global polarization, %$P = R_v \cdot 2\langle\Pi_{0y}\rangle_{p}$, %as a function of impact parameter ratio $b_0$ at 11.5GeV and 27.0Gev. %It shows a quadratic curve similar to the Figs. in Refs. \cite{BPR08,GCD08} }%end color red }
\caption{Success rates ($\%$ at IoU $>$ 0.50) of BACF versus HOG-based trackers. The {\color{red}first}, {\color{green}second} and {\color{blue}third} best methods are shown in color.}
\caption{Success rates ($\%$ at IoU $>$ 0.50) of BACF compared to CF trackers with deep features. The {\color{red}first}, {\color{green}second} and {\color{blue}third} highest rates are highlighted in color. }
\caption{Calculated $\kappa$ of 3C-SiC including respectively anharmonic phonon scattering, isotope scattering, grain boundary scattering (\noindent\textcolor{black}{\rule{0.4cm}{1.5pt}} $L_{\text{grain}}=0.8~\mu$m) and different concentrations of defects [\noindent\textcolor{blue}{\rule{0.4cm}{1.5pt}} $2.2\times10^{20}$ cm$^{-3}$ N$_{\text{C}}$ (0.45$\%$), \noindent\textcolor{blue}{\rule{0.15cm}{1.5pt}}\textcolor{white}{\rule{0.1cm}{1.5pt}}\textcolor{blue}{\rule{0.15cm}{1.5pt}} $8\times10^{20}$ cm$^{-3}$ N$_{\text{C}}$ (1.63$\%$), \noindent\textcolor{forestgreen}{\rule{0.4cm}{1.5pt}} $6.6\times10^{20}$ cm$^{-3}$ N$_{\text{C}}$ (1.3$\%$), $2.6\times10^{20}$ cm$^{-3}$ B$_{\text{C}}$ (0.5$\%$), and \noindent\textcolor{forestgreen}{\rule{0.15cm}{1.5pt}}\textcolor{white}{\rule{0.1cm}{1.5pt}}\textcolor{forestgreen}{\rule{0.15cm}{1.5pt}} $1.32\times10^{21}$ cm$^{-3}$ N$_{\text{C}}$ (2.7$\%$), $1.23\times10^{21}$ cm$^{-3}$ B$_{\text{C}}$ (2.5$\%$)]. The experiments are from Morelli \textit{et al.} \cite{Morelli_1993} and Ivanova \textit{et al.} \cite{Ivanova_im2006}. }
\caption{ (Color online) Switching in a 3:2 layer sample after 40 ps. Color coding: {\color{cyan} $\blacksquare$} Magnetization switches, {\color{red} $\blacksquare$} magnetization does not switch and returns to the initial state, {\color{orange} $\blacksquare$} both magnetizations change their sign twice, ending up in the initial state (back-switching). The compensation temperature $T_\text{comp}$ is indicated by a solid black line, the Curie temperature $T_\text{C}^\text{Gd}$ of the isolated Gd layer by a black dashed line. The three points indicate the chosen values for the scenarios in Fig.~\ref{f:var_mag}.}
\caption{An example graph and its representation in the graph format. The IDs of the vertices are drawn within the cycle, the vertex weight is shown next to the circle ({\color{red}red}) and the edge weight is plotted next to the edge ({\color{blue}blue}).}
\caption{Reference energies (\bluecircleopen,\redsquareopen) and second-order NCSM-PT energies (\bluecircle,\redsquare) with $N_\text{max}^{\text{ref}}=0$ and $2$, respectively, for the ground states of $\elem{C}{11-20}$, $\elem{O}{16-26}$, and $\elem{F}{17-31}$ for the NN+3N-full interaction with $\alpha=0.08\,\text{fm}^4$ and model-space truncation $e_{\text{max}}=12$. All calculations are performed with a Hartree-Fock optimized single-particle basis at $\hbar \Omega =20\,\text{MeV}$. Importance-truncated NCSM calculations (\greentriangleup) are shown for comparison~\cite{TiMu18}. Experimental values are indicated by black bars~\cite{AME16}.}
\caption{Reference energies (\bluecircleopen,\redsquareopen) and second-order NCSM-PT energies (\bluecircle,\redsquare) with $N_\text{max}^{\text{ref}}=2$ for the ground states of $\elem{C}{11-20}$, $\elem{O}{16-26}$, and $\elem{F}{17-31}$ for the NN+3N-ind (squares) and NN+3N-full interaction (circles). The SRG flow parameter is given by $\alpha=0.08\,\text{fm}^4$ and all calculations are performed within a $e_{\text{max}}=12$ truncated model space. We use Hartree-Fock optimized single-particle basis at $\hbar \Omega =20\,\text{MeV}$. Experimental values are indicated by black bars~\cite{AME16}.}
\caption{Comparison with state of the art on Market-1501 (single query) and CUHK03. * denotes unpublished papers. Base networks are annotated. C: CaffeNet, R: ResNet-50, A: AlexNet, G: GoogleNet \cite{szegedy2015going}. The best, second and third highest results are in \textcolor{blue}{blue}, \textcolor{red}{red} and \textcolor{green}{green}, respectively.}
\caption{The distribution of the phase factor in the path integral (\ref{PartFunc2}) for the calculation with CDW mass term $m_{CDW}=40 \meV$ on the $6 \times 6$ lattice at temperature $T=0.09 \eV$. Calculations were performed for superlattice of vacancies shown in the Fig.~\textcolor{red}{3c}.}
\caption{Comparing trackers on three tracking scenarios including higher frame rate tracking (240 FPS), lower frame rate tracking with synthesized motion blur (30 FPS MB) and lower frame rate tracking without motion blur (30 FPS no MB). Results are reported as the AUC of success plots. We also show the speed of each tracker on CPUs and/or GPUs if applicable. The {\color{red}first}, {\color{green}second}, {\color{blue}third}, {\color{cyan}forth} and {\color{Purple}fifth} highest AUCs/speeds are highlighted in color.}
\caption{Rows(1-3) show tracking performance of three trackers icluding a CF tracker with HOG (SRDCF), a CF tracker with deep features (HDT) and a deep tracker (MDNets), comparing lower frame rate ({\color{green}green} boxes) versus higher frame rate ({\color{red}red} boxes) tracking. Ground truth is shown by {\color{blue}blue} boxes. Last row visualizes a failure case of higher frame rate tracking caused by non-rigid deformation for {\color{red}BACF}, {\color{green}Staple}, {\color{blue}MDNet} and {\color{cyan}SFC}.}
\caption{{\bf{\color{SMblack} Schematic of a 2D diffusive photonic circuit (double chain).}} {\color{SMblack}The circuit consists of two parallel dissipatively coupled chains. Here, the squares represent reservoirs (R), the circles are bosonic modes (the red circles indicate possible initial excitations).} \label{fig3-suppl}}
\caption{{\bf{\color{SMblack} A diffusive photonic honeycomb lattice.}} {\color{SMblack} When all the bosonic modes in a hexagonal cell (indicated by the dashed line) have the same amplitude, the cell can support a stationary, compacton-like state.} \label{fig4-suppl}}
\caption{{\bf{\color{SMblack} A diffusive square lattice.}} Stationary distributions of absolute values of modal amplitudes for {\color{SMblack}a} $6\times6$ square lattice {\color{SMblack}without (a) and with (c) additional losses at the sites indicated by red crosses in the inset of (d). Eigenvalues (in the units of $\gamma$) of the systems [{\color{SMblack}Supplementary} Eq.~\ref{square eq}] without (b) and with (d) additional losses. Filled circles in the insets of (b,~d) denote the initial excitations.} \label{sfig1}}
\caption{Combined constraints on $(\xi_\ell,\bar\xi_\ell)$ under the Cheng-Sher ansatz. The dark and light gray regions are respectively the parameter space allowed by the $h \to \tau\tau$ measurement from the combined \sout{Run-I} \tblue{LHC} data and the CMS $h \to \mu \tau$ data at 95\% CL. In the case of $\bar Y_{sb}= 3.4 \times 10^{-4}$ (left plot) and $\bar Y_{sb}= 1.3 \times 10^{-4}$ (right plot), the parameter space satisfying $75\% < \mathcal B(B_s \to \mu^+\mu^-)/\mathcal B(B_s \to \mu^+ \mu^-)_\SM < 95\%$ is shown by the green region.}
\caption{\textcolor{blue}{Only robot $1$ can triangulate in this configuration. % Red triangle represent the beacon and each % circle indicate a robot with unknown location. }}
\caption{An image showing objects of different classes: \textcolor{color1}{miniskirt}, \textcolor{color2}{Hat with a wide brim}, \textcolor{color3}{sunglasses}, \textcolor{color4}{plastic bag}, \textcolor{color5}{purse}, \textcolor{color6}{person}}
\caption{\label{fig:mvsin} Contours of the cosmologically preferred region $0.3$ GeV $<T_d< 1$ GeV as a function of the twin neutrino mass and mixing, as well as current and projected bounds on sterile neutrinos. The shaded {\color[rgb]{0.53, 0.47, 0.70}\bf purple} region shows the 95\% C.L. limits from DELPHI on sterile neutrinos produced from $Z$-decays at LEP~\cite{Abreu:1996pa}. We also show projected reaches from displaced searches (see next section) as dashed curves. Projections for SHiP~\cite{Anelli:2015pba} ({\color[rgb]{0.56, 0.69, 0.19}\bf green}), DUNE~\cite{Adams:2013qkq} ({\color[rgb]{1, 0.75, 0}\bf yellow}), and FCC-ee~\cite{Blondel:2014bra,Abada:2014cca} ({\color[rgb]{.92, .38, .2}\bf red}) are taken from \cite{Deppisch:2015qwa}. The LHC reach at $\sqrt{s} = 13$~TeV with 300~fb$^{-1}$, with searches for lepton jets ({\color[rgb]{0.36,.61,.78}\bf blue}) and trileptons ({\color[rgb]{0.77,.43,.10}\bf brown}) are also shown~\cite{Izaguirre:2015pga}. The exclusion region and projection curves are only valid if the twin-neutrino cannot decay into light twin particles, for instance, in the FTH (1 generation) scenario. }
\caption{Prediction of flight trajectory: delay compensation ({\color{cyan}dashed}), \ac{mpc} prediction ({\color{magenta}solid}); reference path ({\color{green}dotted}).\vspace{0pt}}
\caption{Trajectory tracking results (Flight~1): kite position trajectory, $\xi(t)$ ({\color{blue}solid}), following a reference figure-of-eight path, $\xi^\8$ ({\color{green}dotted}), with delay compensation prediction trajectory, $\xi^{t_d}(t)$ ({\color{cyan}dashed}). The trajectories start at the position marked with crosses at time $t=0$s.\vspace{0pt}}
\caption{Velocity trajectory, $v_{\theta\phi}(t)$, estimated based on derivatives of position measurements, $r(t)\sqrt{\dot{\theta}^2(t)+\cos^2(\theta(t))\dot{\phi}^2(t)}$ ({\color{blue}solid}), and based on model~\eqref{eq:velocity} ({\color{cyan}dashed}).}
\caption{Trajectory of heading angle, $\gamma(t)$ ({\color{blue}solid}), tracking commanded signal, $\gamma^\re(t)$ ({\color{green}dotted}), with shifted tracking error, $e_{t_d}(t)$ ({\color{red}dashed}).}
\caption{Bound on the rate of change of the commanded heading angle, $l_r$ ({\color{blue}solid}), and resulting magnitude of the rate of the commanded orientation, $|\dot{\gamma}^\re(t)|$ ({\color{magenta}dotted}).}
\caption{(Color online) (a) Illustration of the anisotropic triangular lattice. The phase factor $\gamma_{ij}=1$, $e^{i2\pi/3}$ and $e^{-i2\pi/3}$ for the bond along the $\bf{a}_1$, $\bf{a}_2$ and $\bf{a}_3$, respectively. (b) The first Brillouin zone of the triangular lattice. The high symmetry points $K=\big(\frac{4\pi}{3},0\big)$~(magenta) and $K'=\big(-\frac{4\pi}{3},0\big)$~(cyan) at the corners, and $M_0=\big(0,\frac{2\pi}{\sqrt3}\big)$~(red), $M_+=\big(\pi,\frac{\pi}{\sqrt3}\big)$~(blue) and $M_-=\big(-\pi,\frac{\pi}{\sqrt3}\big)$~(green) at the middle of the edges are marked. (c)-(e) show the magnetic patterns of the classical spins. (c) shows the $120^{\circ}$ order whose peaks of the static structure factors\cite{Notes-SSF} locate at the $K$~($K'$) point. (d) and (e) show the three degenerate ground states of the stripe-B and stripe-A order, respectively. The peaks of the static structure factors of the three degenerate states locate at the $M_0$, $M_+$ and $M_-$ points, from top to bottom, respectively.} \label{FIG-Model} \end{figure} \section{Classical phase diagram}\label{SEC-CLPhaseDiagram} We firstly give glimpses of the classical phase diagram of the model \eqref{QSL-Ham} before moving to the large-scale numerical calculations. The classical phase diagram has already been obtained by Li \textit{et al}. with Luttinger-Tisza method\cite{LT-46,Bertaut-61,Litvin-74} and Monte Carlo simulation\cite{Metropolis-1953,LWC-2016}. Here, we will focus on the criticality of the phase transitions between those phases in the phase diagram. The classical spin is an $O(3)$ vector, which is given by \begin{equation}\label{classicalspin3} \boldsymbol{S}_i = S\left(\sin{\theta_i}\cos{\phi_i}, \sin{\theta_i}\sin{\phi_i}, \cos{\theta_i}\right), \end{equation} where $\theta_i$ and $\phi_i=\textbf{Q}\cdot\boldsymbol{R}_i+\varphi$ are respectively the polar and azimuthal angles at site $i$ with the position $\boldsymbol{R}_i$. The ordering wave vector $\textbf{Q}$ is determined after minimizing $\mathcal{H}$ with respect to $\{\theta_i, \phi_i\}$. %$S_{\pm\textbf{Q}}\neq0$, but $S_{\bf{q}}=0$ for all the other $\bf{q}\neq\pm\textbf{Q}$. By the Fourier transformation $S_{i}^{\alpha}=\frac{1}{\sqrt{N}}\sum_{\bf{q}}e^{i{\bf{q}}\cdot{\boldsymbol{R}}_i}S_{\bf{q}}^{\alpha}$ with $\alpha=x,~y$ and $z$, the model \eqref{QSL-Ham} takes the form \begin{eqnarray} \mathcal{H} & = & \sum_{\left<ij\right>} \sum_{\alpha\beta}S_i^{\alpha}J_{ij}^{\alpha\beta}S_j^{\beta} \nonumber \\ & = & \sum_{\alpha\beta}\sum_{{\bf{q}}} S_{\bf{q}}^{\alpha}J^{\alpha\beta}(\bf{q})S_{\bf{-q}}^{\beta} \end{eqnarray} where $J(\bf{q})$ is a $3\times3$ symmetric matrix, which is written as \begin{align}\label{JqMatrix} &J({\bf{q}})= \left[ \begin{array}{ccc} 2J_{\pm}\mathcal{F}+2J_{\pm\pm}\mathcal{G} & -2\sqrt{3}J_{\pm\pm}\mathcal{K} & -\sqrt{3}J_{z\pm}\mathcal{K} \\ -2\sqrt{3}J_{\pm\pm}\mathcal{K} & 2J_{\pm}\mathcal{F}-2J_{\pm\pm}\mathcal{G} & J_{z\pm}\mathcal{G} \\ -\sqrt{3}J_{z\pm}\mathcal{K} & J_{z\pm}\mathcal{G} & J_{zz}\mathcal{F} \\ \end{array} \right] \end{align} with \begin{align} \left\{ \begin{array}{l} \mathcal{F} = f({\bf{q}}) = \cos q_x + 2\cos\frac{q_x}{2}\cos\frac{\sqrt{3}q_y}{2} \\ \mathcal{G} = g({\bf{q}}) = \cos q_x - \cos\frac{q_x}{2}\cos\frac{\sqrt{3}q_y}{2} \\ \mathcal{K} = h({\bf{q}}) = \sin\frac{q_x}{2}\sin\frac{\sqrt{3}q_y}{2} \\ \end{array} \right.. \end{align} The smallest eigenvalue of $J({\bf{q}})$ over the first Brillouin zone~(FBZ)~(Fig. \ref{FIG-Model} (b)) provides a lower bound for the classical ground-state energy\cite{LT-46,Bertaut-61,Litvin-74}. Therefore, we obtain the magnetic order with the characteristic $\textbf{Q}$ for the given parameters. %&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&& \begin{table*} \caption{The classical phases, and the characteristic wave vectors $\textbf{Q}$, classical ground-state energy $E_{\textrm{cl}}/(NS^2)$, the allowed angles $\left(\theta,\phi\right)$, and the conditions for the phases for $J_{z\pm}\geq0$.\label{TabPhaseEnergyJzpmGEQ0}}
\caption{(Color online) Classical ground-state energy per site $E_{\rm{cl}}/N$ for $J_{z\pm} = $\;0.0 (red solid), 0.5 (cyan dashed), 1.0 (blue dot-dashed) and 1.5 (green dotted) with$S=1/2$. The filled symbols mark the phase transition points.}
\caption{Calculated rephasing one-quantum (a) and non-rephasing two-quantum (b) 2D FWM spectra including F\"{o}rster coupling. The additional peaks arising due to exciton transfer processes between the QDs are marked by white circles\textcolor{red}{.}}
\caption{Comparison of test accuracy of kernel learning with RFS and neural networks. For each architecture test we provide its ranking in terms of performance and relative rank ordering. The column designated as ``Arch'' describes the convolution size as pairs {\color{blue}$a$},{\color{red}$b$} where the first layer convolution is of size {\color{blue}$a$}x{\color{blue}$a$} and the second is {\color{red}$b$}x{\color{red}$b$}.}
\caption{If all agents start at non-collinear (non-coplanar) positions, then their positions will be non-collinear (non-coplanar) {\color{blue} for any finite time.} }
\caption{Mean reprojection error of all 3D points of all cameras (in pixels). In {\color{red}red}, fraction of calibrated cameras in case of failure. Our residual error in indoor scenes is about 4 to 10 times larger than on Strecha's dataset, but picture resolution is 3 times as large.}
\caption{ Embedding neighborhood of \textsc{iraq} extracted from a dynamic embedding fitted to the congress data. It is the word whose embedding vector has largest absolute drift. By listing the neighborhood for all the time bins, we can see how Iraq's embedding vector drifts smoothly. In \green{1858} \textsc{iraq's} embedding neighborhood contains countries and regions. In \blue{1950} a rethoric more specific to arabic countries crystalizes. In \purple{1980} Iraq invades Iran and words like \purple{invasion}, \purple{aggressor} and \purple{troops} are in the neighborhood. By \red{2008} the embedding neighborhood of iran contains words like \red{terror}, \red{terrorism} and \red{saddam}.}
\caption{Plane sweeping using three different patch similarity measures. Proposed learned multi-view similarity \vs pairwise ZNCC and pairwise LIFT. Reference images (a-d). Ground truth (e-h). Absolute deviations from ground truth for the proposed method (i-k), ZNCC (m-p) and LIFT (q-t). \textcolor{blue}{Blue}: no difference, \textcolor{darkred}{red}: high difference. }
\caption{Comparison among log-likelihood of first ({\color{green} green}) and second level ({\color{blue} blue}) of kernel decomposition algorithm, for each valid sequence.}
\caption{Comparison among loglikelihood of our kernel composition algorithm (Discretized Estimation) ({\color{blue} blue}), gradient descent exponential Hawkes ({\color{red} red}) and an Ensemble Model (Discretized Estimation + Gradient-based) ({\color{green} green}), for each valid sequence.}
\caption{\label{comparison} Summary of the information (rows 1-5) and the forecasting power (rows 6-10) for different datasets. For the first two datasets, we start the prediction from the sixth year from the beginning of the data. For the last two datasets, we forecasts all the \el ~events since 1950, since there is enough historical data to calculate the first value of ${\rm FI}(t)$ in Jan. 1950. In all the datasets, ${\rm FI}(t)$ is only based on the past data, from the beginning of the data. The ``Resolution of the data'' refers to the spatial (zonal and meridional) resolution of the data. In the row of ``$D$'' , the symbol ``*'' indicates that there are not enough events in the datasets to perform the Kolmogorov-Smirnov statistic. \textcolor{mypink1}{The numerator of hit rate is composed of two parts, ``normal''$ +$``delayed'', the denominator is the number of all the \el ~events for each dataset. The numerator of the false alarm rate is the number of false alarms, the denominator is the sum of number of false alarms and the ``correct rejections'' which equals the number of years where no \el ~started and no false alarm appeared in the past $18$ months before the year.} In the columns of lead time, we show the mean value $\pm$ 1 standard deviation in units of months. For more details on the above results see Figs. S2 in the SI.}
\caption{ (Color online) $\eta/s$ and $\zeta/s$ versus scaled temperature $T/T_C$: \textcolor{dmagenta}{ \textbf{Top:} The symbols indicate the PHSD results of $\eta/s$ from Ref.~\cite{Ozvenchuk:2012kh}, calculated using different methods: the relaxation-time approximation (red line$+$diamonds) and the Kubo formalism (blue line$+$dots); the black line corresponds to the parametrization of the PHSD results for $\eta/s$. The orange short dashed line demonstrates the Kovtun-Son-Starinets bound \cite{KSS} $(\eta/s)_{KSS}=1/(4\pi).$ For comparison, the results from the virial expansion approach (green line) \cite{Mattiello} are shown as a function of temperature, too. The orange dashed line is the $\eta/s$ of VISHNU hydrodynamical model that has been recently determined by Bayesian analysis; \textbf{Bottom:} $\zeta/s$ from PHSD simulation from Ref. \cite{Ozvenchuk:2012kh} and the $\zeta/s$ that is adapted in our hydrodynamical simulations. } }
\caption{ (\aaa) Box plot of average migration speeds in peristaltic ($N = 20$) and amphistaltic ($N = 8$) \phys fragments. Two asterisks denote statistically significant differences between medians ($p < 0.01$). % (\bbb) Simplified model schematic for the distance traveled by endoplasmic fluid particles per oscillation cycle. Top panel, peristaltic fragments; bottom panel, amphistaltic fragments. % (\ccc) Scatter plot of the distance $\Delta x_{cent}$ traveled by the centroid of the \phys fragment per oscillation cycle vs. the net distance $\Delta x_{flow}$ traveled by an endoplasmic fluid particle. \textcolor{red}{\solidcircle}, peristaltic fragments; \textcolor{blue}{\solidsquar}, amphistaltic fragments. The dashed line is $\Delta x_{cent}=\Delta x_{flow}$. % }
\caption{Time histories of longitudinal ectoplasm velocity (\textcolor{verde}{\linetri}) and longitudinal traction stresses (\textcolor{blue}{\linecir}) at two specific locations in the front (panel \aaa) and the back (panel \bbb) of the peristaltic \phys fragment shown in Figure \ref{fig:kymo2}. % The tiled bars at the top of the plots represent the time-dependent phase differences (in radians) between the ectoplasm velocity and the traction stresses. Blue and orange tiles represent phase differences near $-\pi/2$ (cell and substrate stick) and zero (cell and substrate slip) respectively, as indicated by the color scale at the top of the figure.}
\caption{Top row (\aaa, \ccc, \eee, \ggg): Time histories of endoplasmic flow velocity (\textcolor{blue}{\linecir}) and ratiometric measurement of [Ca$^{2+}$]i (\textcolor{verde}{\linetri}), averaged along the width of a peristaltic fragment (panels \aaa\; and\ccc) and an amphistaltic fragment (panels \eee\; and\ggg). % Bottom row (\bbb, \ddd, \fff, \hhh): Time histories of endoplasmic flow velocity (\textcolor{blue}{\linecir}) and peripheral traction stress (\textcolor{red}{\linesquar}), averaged along the width of a peristaltic fragment (panels \bbb\; and\ddd) and an amphistaltic fragment (panels \fff\; and\hhh). % Panels (\aaa, \bbb, \eee, \fff): Fragment front. Panels (\ccc, \ddd, \ggg, \hhh): Fragment rear.}
\caption{Left: An overview of our framework. % The inputs are an environment image and a sentence (either a navigation {\color{blue}command} or a {\color{red}question}). % The output is either a navigation {\color{blue}action} or an {\color{red}answer} to the question, respectively. % The red and blue lines in (a) indicate different tasks going through exactly the same process. % Right: The pipeline of the programmer of the language module. % The input is a sequence of word embeddings. % The output is the attention map at the final step. }
\caption{The projections from word embedding to {\color{blue}syntax embedding} and {\color{red}functionality embedding}. }
\caption{Details of computing the embedding masks. % (a) The pipeline of Question Intention in Figure~\ref{fig:overview}b, which computes an embedding mask according to a question. % (b) The two FC layers used by Mask Computation in both (a) and Figure~\ref{fig:overview} Right. % They project a {\color{red}functionality embedding} to a mask. }
\caption{A Bidirectional RNN that receives a sequence of {\color{blue}syntax embeddings}, and outputs a sequence of {\color{ForestGreen}word contexts} and an {\color{orange}RNN booting} vector. }
\caption{(color online). The quark flow diagrams for the process $\Lambda(1405) \rightarrow K^-\, p$ can be decomposed into a completely-connected part and two parts involving disconnected sea-quark loop contributions. The upper-left completely-connected diagram and the upper-right diagram (\color{pb} \textit{blue} \color{black} dots) are included in the lattice QCD calculations as the photon interacts with a valence quark. The case where a photon couples to a disconnected sea quark loop, illustrated in the lower diagram (\color{red}\textit{red} \color{black} dot), is not included in the lattice QCD calculations. \vspace*{-12pt} \label{fig:dcomp}}
\caption{\textit{Left panel}: \grid \gray intensity map in Galactic coordinates with a three-pixel Gaussian smoothing. Photon energy: 50--400 MeV. Integration time: 2015-06-20 UT 06:00:00 -- 2015-06-30 UT 06:00:00. Pixel size: $0.5^{\circ}$. Green contour: $95\%$ confidence region. White contour: statistical + systematic ($0.1^{\circ}$) containment region. White cross: optical position of \vcyg. \textit{Right panel}: The quiescent phase of \vcyg -- with the same characteristics of the left panel, but different size -- from 2015-01-01 UT 12:00:00 to 2015-06-01 UT 12:00:00. The three pulsars included in the multi-source analysis are marked with magenta crosses (from left to right: PSR J2032+4127, PSR J2021+4026, and PSR J2021+3651). The white cross is the optical position of \vcyg. The white contour is the AGILE containment region of the flaring source (stat + syst). The color bar is the same for both the maps.}
\caption{Differential \gray spectrum from \vcyg as detected by the \grid during the peak emission activity, 2015-06-24 UT 06:00:00 to 2015-06-26 UT 06:00:00 (MJD 57197.25--57199.25). The flux UL is a $68\%$ C. L. value.}
\caption{HE \gray light-curves across the \vcyg peak emission. From top to botton: \grid 48h-bin light-curve (50--400 MeV energy band); \grid TS of each time-bin; \lat 24h-bin light-curve (60--400 MeV energy band); \lat TS for each time-bin. For both AGILE and \fermi, flux error bars and flux ULs are $68\%$ C. L. values.}
\caption{\grid (blue) and \lat (red) time evolution of the \vcyg off-axis angle, for the period 2015-06-24 UT 06:00:00 and 2015-06-26 UT 18:00:00 (MJD 57197.25--59199.75), union of the \gray peak emission time intervals found by AGILE and \fermi (see Fig.5~\ref{fig:agile_fermi_june2015}).}
\caption{\small (a) The proposed algorithm, \methodall, consistently performs the best, by having the highest AUC, in predicting fair and unfair users with varying percentage of training labels. (b) \methodall\discovered a bot-net of 40 confirmed shills of one user, rating each other positively.}
\caption{This is the set of mutually recursive definitions of fairness, reliability and goodness for the proposed \methodall\algorithm. The yellow shaded part addresses the cold start problems and gray shaded part incorporates the behavioral properties.}
\caption{\underline{Unsupervised Predictions}: The table shows the Average Precision values of all algorithms in unsupervised prediction of unfair and fair users across five datasets. The {\color{blue!75}best algorithm} in each column is colored {\color{blue!75}blue} and {\color{gray!100}second best is gray}. Overall, \methodall\performs the best or second best in 9 of the 10 cases.$nc$ indicates `no convergence'.\label{tab:unsupervised_results} %%\vspace{-4mm} }
\caption{Flow rate and evolution of the two surface angles during the discharge: a) $\theta$ vs $t$ for $\varphi_s=0.59$ (-$\square$-) and $\varphi_p=0.66$ (\textcolor{blue}{$--$} $\theta_1$, \textcolor{blue}{$\bullet$} $\theta_2$). b) $\theta$ vs $t$ for different values of $\varphi_p$. Lines with dots and dashed lines of the same colour correspond to $\theta_2$ and $\theta_1$, respectively. The gray line indicates the angle of repose. The angle error due to image analysis was of the order of one pixel, corresponding to $\pm 1^{\circ}$. c) The flow rate $Q=3.82 \pm 0.05$ g/s is constant during the discharge and independent of the initial packing fraction of the bed.}
\caption{Cavity radius as a function of time for a) 2D and b) 3D discharges of standard and densely packed grains; in both cases a power-law dependence is found, see log-log plots in the insets. c) $R$ vs $t$ obtained from the solution of eq. \ref{eqR2} for different values of $\varphi_p$ (dashed lines) compared with the 2D experiments (points); the inset shows the transition coordinate $x(t)$ obtained from the analysis of videos. Colours in (a) and (c) are in correspondence with fig. \ref{fig3}b; in (b): $\varphi_s$ (\textcolor{black}{$\bullet$}) and $\varphi_p$ (\textcolor{red}{$\blacksquare$}). }
\caption{\label{fig:1dpresentposteriors}Here we show the 1d posteriors for the above scans. The colors of the posterior indicate which scan they belong to; in the same order as above, {\color{niceblue} Planck T}, {\color{nicered} Planck P}, {\color{nicegreen} Planck P+BAO}, {\color{mypurple} Planck P+BAO+$H_0$}, and {\color{niceorange} Planck P+BAO+$H_0$+LSS}.}
\caption{\label{fig:2dpresentposteriors}Here we show two different 2d posteriors for three of the five scans ({\color{niceblue} Planck T}, {\color{nicegreen} Planck P+BAO}, and {\color{niceorange} Planck P+BAO+$H_0$+LSS}). The solid lines are 1$\sigma$ contours, and the dot-dashed lines are 2$\sigma$ contours. The posteriors in the top figure exhibit degeneracy between $N_{\rm eff}$ and $N_{\rm fld}$, motivating the parametrization in terms of $N_{\rm tot}$ and $f_{\rm fs}$ in the bottom figure, and demonstrating that the strongest constraints arise on the sum $N_{\rm tot}$. }
\caption{\label{fig:2dpresenttensionposteriors}Here we show different 2d posteriors for three of the five scans ({\color{nicegreen} Planck P+BAO}, {\color{mypurple} Planck P+BAO+$H_0$}, and {\color{niceorange} Planck P+BAO+$H_0$+LSS}). The solid lines are 1$\sigma$ contours, and the dot-dashed lines are 2$\sigma$ contours. The posteriors demonstrate how free-streaming DR affects late-time cosmological observables differently than interacting DR does. In the left figures, we explore the effect of $N_{\rm eff}$; in the right figures, $N_{\rm fld}$. In the top figures, we explore the effect on $H_0$; in the bottom figures, $\sigma_8$.}
\caption{\label{fig:futureconstraints}Here we show two different 2d posteriors for two of the three future constraints ({\color{darkergreen} Planck P+S3} and {\color{darkerpurple} Planck P+S4}) alongside one of the present-day constraints ({\color{nicered} Planck P}). The solid lines are 1$\sigma$ contours, and the dot-dashed lines are 2$\sigma$ contours. The left figure shows the conventional parametrization in terms of $N_{\rm eff}$ and $N_{\rm fld}$, whereas the right figure shows our alternative parametrization in terms of $N_{\rm tot}$ and $f_{\rm fs}$.}
\caption{\label{fig:2dfuturetensionposteriors}Here we show two different 2d posteriors for one present-day and two mock scans ({\color{nicered} Planck P}, {\color{darkergreen} Planck P+S3}, and {\color{darkerpurple} Planck P+S4}). The solid lines are 1$\sigma$ contours, and the dot-dashed lines are 2$\sigma$ contours. The posteriors demonstrate how free-streaming DR will affect late-time cosmological observables differently than interacting DR will. In the left figures, we explore the effect of $N_{\rm eff}$; in the right figures, $N_{\rm fld}$. In the top figures, we explore the effect on $H_0$; in the bottom figures, $\sigma_8$.}
\caption{\label{fig:2dsig8H0posteriors}Here we show $H_0$-$\sigma_8$ 2d posteriors for six scans. On the left, we show present results with {\color{nicegreen} Planck P+BAO}, {\color{mypurple} Planck P+BAO+$H_0$}, and {\color{niceorange} Planck P+BAO+$H_0$+LSS}. On the right, we show the present {\color{nicered} Planck P} contrasted with the forecasted {\color{darkergreen} Planck P+S3} and {\color{darkerpurple} Planck P+S4}. The solid lines are 1$\sigma$ contours, and the dot-dashed lines are 2$\sigma$ contours.}
\caption{\label{fig:triangle}Here we show the 2d posteriors in the {\color{nicegreen} Planck P+BAO} run. The dark gray regions are the 1$\sigma$ allowed regions, and the light gray regions are the 2$\sigma$ allowed regions. These plots illustrate correlations between the various cosmological parameters upon digital zoom-in.}
\caption{\label{plot} (a) Magnetic period of the spin spirals in the triple layer Fe on Ir(111) at different temperatures from experiments, mean-field calculations and Monte Carlo simulations. The temperature scale for the mean-field data was rescaled by a factor of 0.71 for better comparison, since it significantly overestimates the temperature range of the period change as well as the critical temperature. The error bars reflect the resolution of the Fourier transformation (used to measure the period from the experimental data and from the simulations) as well as the thermal drift in the experiment. (b)-(c) Schematic view of the model system used to describe the triple-layer Fe on Ir(111) in the theoretical calculations, with the coupling constants used in Eq.~(\ref{eqn1}). Perfect fcc stacking was used to simplify the geometry of the model. The intralayer couplings are different in every layer, with $\left|J_{11}\right| < \left|J_{22}\right|< \left|J_{33}\right|$ for the Heisenberg couplings and $D_{11} > D_{22} > D_{33}$ for the Dzyaloshinsky-Moriya interactions. (d) Numerical values of the coupling constants used for the Monte Carlo simulations in (a), with the notations of (b)-(c). } \end{figure} % % theory part For a quantitative theoretical explanation of the remarkable increase of the spiral wavelength with temperature, we relied on a model which treats the three atomic layers of Fe separately instead of using single Heisenberg and Dzyaloshinsky--Moriya interactions for the whole ultrathin film. The model Hamiltonian reads % In order to reproduce the increase of the spiral wavelength with the % temperature in a theoretical model, an effective Hamiltonian with only % nearest-neighbor interactions in a single layer is not sufficient and % therefore we use the model Hamiltonian \begin{eqnarray} \!\!\!\!\!\!H\!=\frac{1}{2}\!\!\!\sum_{p,q,\left<i,j\right>}\!\!\!J_{pq,ij} \boldsymbol{S}_{p,i}\boldsymbol{S}_{q,j}\!+\!\frac{1}{2}\!\!\! \sum_{p,\left<i,j\right>}\!\!\!\boldsymbol{D}_{pp,ij} \left(\boldsymbol{S}_{p,i}\times\boldsymbol{S}_{p,j}\right),\label{eqn1} \end{eqnarray} also illustrated in Figs.~\ref{plot}(b)-(c). The $\boldsymbol{S}_{p,i}$ denote classical unit vectors representing the magnetic moments. The $p,q=1,2,3$ indices denote the three layers starting from the one closest to the Ir substrate, while $i$ and $j$ are intralayer indices. The summations only run over the nearest neighbors, including six intralayer neighbors and three neighbors in each adjacent layer. The $J_{pq,ij}$ coefficients denote intralayer and interlayer Heisenberg exchange interactions. The intralayer Dzyaloshinsky--Moriya vectors $\boldsymbol{D}_{pp,ij}$ are perpendicular to the nearest-neighbor bonds and are chosen to be in-plane -- see the dashed arrows in Fig.~\ref{plot}(c). Regarding the temperature-dependent spin spiral period, the crucial point about Eq.~(\ref{eqn1}) is the consideration of layer-dependent interaction parameters. Layer-dependent coupling coefficients naturally appear during \textit{ab initio} calculations~\cite{zakeri_direct_2013, meyerheim_new_2009, meng_direct_2014, dupe_engineering_2016} due to the different hybridization effects for atomic layers with different distances to the non-magnetic substrate and the vacuum interface. In the present system, this choice is further supported by the fact that the actual intralayer atomic distances may differ between the three Fe layers due to the strain relief~\cite{hsu_electric-field-driven_2017, finco_tailoring_2016}. Note that the microscopic parameters in Eq.~(\ref{eqn1}) are not explicitly temperature dependent. In an atomistic model this is generally justified for Fe which has strong localized magnetic moments~\cite{mryasov_temperature-dependent_2005}. In principle, a modification of the atomic structure with temperature could influence the coupling coefficients~\cite{Akimitsu, veber_high-field_2008}. However, we did not observe any obvious modification of the structure of the Fe film as a function of temperature in the topographic images. In addition, the mechanisms discussed in Ref.~\cite{izyumov_modulated_1984} for homogeneous bulk magnets seemed to be insufficient to quantitatively explain the exceptionally large increase of the period in the triple layer Fe. % , although changes of the interlayer %distances cannot be excluded. For obtaining such a strong temperature dependence of the period, it was necessary to assume that the $J_{pp}$ and $\boldsymbol{D}_{pp}$ intralayer couplings determine different spiral periods in the different layers. The Dzyaloshinsky--Moriya interaction is expected to get weaker when moving away from the Ir substrate, since it primarily appears at the interface between the magnetic Fe layers and the heavy metal Ir substrate with strong spin--orbit coupling. On the contrary, the Heisenberg exchange interactions should get stronger when moving away from the substrate, partly because the rearrangement of the Fe atoms into the bcc structure~\cite{hsu_electric-field-driven_2017} should make their values approximate the strong ferromagnetic coupling in bcc Fe, and partly because \textit{ab initio} calculations indicate a weakening of the Heisenberg exchange interactions at Fe/Ir interfaces due to hybridization effects~\cite{von_bergmann_observation_2006, simon_spin-correlations_2014, dupe_tailoring_2014, rozsa_skyrmions_2016}. As a net effect, the determined period will be higher in layers further away from the Ir substrate. It is also assumed that if only the intralayer couplings are considered, then the critical temperature of the layers should increase from the first layer towards the third. % ; this % is in agreement with the enhanced thermal stability with increasing coverage. The interlayer ferromagnetic couplings $J_{12},J_{23}$ contribute to the enhancement of the thermal stability, while ensuring that the layers are coupled sufficiently strongly to each other, meaning that the period of the spin spiral will be the same between the three layers at any fixed temperature, and that the triple layer will only have a single critical temperature.% The experimental results do not indicate that the magnetic % ordering should differ between the layers, in contrast to thin Fe films on % Cu(001)~\cite{meyerheim_new_2009}. How the period depends on the temperature can be understood from a mean-field model. Following the derivation given in the Supplemental Material~\cite{supp}, the free energy $F_{\textrm{MF}}$ per spin of the spin spiral state with wave vector $k$ may be expressed as \begin{eqnarray} \frac{1}{N}F_{\textrm{MF}}\left(k\right)= -\frac{1}{2}\sum_{p,q}\mathcal{J}_{pq} \left(k\right)\left<S_{p}\left(k\right)\right> \left<S_{q}\left(k\right)\right>\nonumber \\ -\sum_{p}k_{\textrm{B}}T\ln\left(4\pi\sinh\left( \frac{B_{p}\left(k\right)}{k_{\textrm{B}}T}\right) \frac{k_{\textrm{B}}T}{B_{p}\left(k\right)}\right),\label{eqn2} \end{eqnarray} with \begin{eqnarray} B_{p}\left(k\right)=&&-\Bigg[\frac{1}{N}\sum_{q,\left<i,j\right>}J_{pq,ij} \cos\left(k\left(x_{p,i}-x_{q,j}\right)\right) \left<S_{q}\left(k\right)\right>\nonumber \\ &&+D_{pq,ij}\sin\left(k\left(x_{p,i}-x_{q,j}\right)\right) \left<S_{q}\left(k\right)\right>\Bigg]\label{eqn3} \end{eqnarray} the mean field in energy dimensions in layer $p$, \begin{eqnarray} \mathcal{J}_{pq}\left(k\right)= &&\frac{1}{N}\sum_{\left<i,j\right>}J_{pq,ij} \cos\left(k\left(x_{p,i}-x_{q,j}\right)\right)\nonumber \\ &&+D_{pq,ij}\sin\left(k\left(x_{p,i}-x_{q,j}\right)\right)\label{eqn4} \end{eqnarray} the Fourier transform of the interaction coefficients, and $\left<S_{p}\left(k\right)\right>$ the order parameter of the spin spiral state. $N$ denotes the number of atoms in a single layer. The equilibrium period of the spin spiral may be obtained by minimizing Eq.~(\ref{eqn2}) with respect to the wave vector $k$. The wave vector dependence of the free energy is included in the coupling coefficients $\mathcal{J}_{pq}\left(k\right)$, which are minimized by different $k$ values in the different layers as mentioned above. The temperature dependence is encapsulated in the order parameters $\left<S_{p}\left(k\right)\right>$, which do not depend significantly on the wave vector. However, $\left<S_{p}\left(k\right)\right>$ decreases faster with temperature in the first and second layers than in the third one, gradually decreasing their relative contribution to the free energy expression (\ref{eqn2}). This effect shifts the minimum of $F_{\textrm{MF}}\left(k\right)$ towards lower wave vectors with increasing temperature, explaining the effect observed in the experiments. The interaction coefficients in Eq.~(\ref{eqn1}) were determined based on the above assumptions regarding their relative magnitudes, and the numerical values summarized in Fig.~\ref{plot}(d) were obtained by tuning the values in order to quantitatively reproduce the temperature dependence observed in the experiments by mean-field calculations and Monte Carlo simulations. Figure~\ref{plot}(c) demonstrates a good agreement between the theoretical model and the experiments. The details of the simulations are given in the Supplemental Material~\cite{supp}. Note that explaining the zigzag wavefront of the spin spirals~\cite{Hagemeister} or the range of possible periods at low temperature~\cite{finco_tailoring_2016} requires assumptions going beyond the model in Eq.~(\ref{eqn1}), but the Hamiltonian seems to capture the main mechanism behind the temperature dependence of the wavelength. % % Conclusion In summary, we have shown using SP-STM measurements that the observed magnetic period in the triple-layer Fe on Ir(111) increases significantly, by approximately a factor of 16 between \SI{8}{\kelvin} and \SI{300}{\kelvin}. Based on the different hybridization and strain relief effects in the three atomic layers, we proposed a theoretical model with layer-dependent coupling coefficients, which quantitatively reproduces the period increase observed in the experiments. Since the presence of different and coupled magnetic layers appears to be decisive, our work shows that the usual mapping of ultrathin films onto a single effective layer might fail to describe temperature effects or phase diagrams for films thicker than a single atomic layer. Our results can hence motivate further investigations in magnetic multilayers regarding the finite temperature behavior of noncollinear spin structures. \begin{acknowledgments} The authors thank T. Eelbo for technical help. Financial support by the European Union via the Horizon 2020 research and innovation programme under grant agreement No.~665095 (MagicSky), by the Deutsche Forschungsgemeinschaft via SFB668-A8 and -A11, by the Hamburgische Stiftung f\"{u}r Wissenschaften, Entwicklung und Kultur Helmut und Hannelore Greve, and by the National Research, Development and Innovation Office of Hungary under project No.~K115575 is gratefully acknowledged. \end{acknowledgments} %merlin.mbs apsrev4-1.bst 2010-07-25 4.21a (PWD, AO, DPC) hacked %Control: key (0) %Control: author (8) initials jnrlst %Control: editor formatted (1) identically to author %Control: production of article title (-1) disabled %Control: page (0) single %Control: year (1) truncated %Control: production of eprint (0) enabled \begin{thebibliography}{37}% \makeatletter \providecommand \@ifxundefined[1]{% \@ifx{#1\undefined} }% \providecommand \@ifnum[1]{% \ifnum #1\expandafter \@firstoftwo\else \expandafter \@secondoftwo\fi }% \providecommand \@ifx[1]{% \ifx #1\expandafter \@firstoftwo\else \expandafter \@secondoftwo\fi }% \providecommand \natexlab[1]{#1}% \providecommand \enquote[1]{``#1''}% \providecommand \bibnamefont[1]{#1}% \providecommand \bibfnamefont[1]{#1}% \providecommand \citenamefont[1]{#1}% \providecommand \href@noop [0]{\@secondoftwo}% \providecommand \href[0]{\begingroup \@sanitize@url\@href}% \providecommand \@href[1]{\@@startlink{#1}\@@href}% \providecommand \@@href[1]{\endgroup#1\@@endlink}% \providecommand \@sanitize@url[0]{\catcode `\\12\catcode `\$12\catcode `\&12\catcode `\#12\catcode `\^12\catcode `\_12\catcode `\%12\relax}% \providecommand \@@startlink[1]{}% \providecommand \@@endlink[0]{}% \providecommand \url[0]{\begingroup\@sanitize@url\@url}% \providecommand \@url[1]{\endgroup\@href{#1}{\urlprefix }}% \providecommand \urlprefix[0]{URL }% \providecommand \Eprint[0]{\href }% \providecommand \doibase[0]{http://dx.doi.org/}% \providecommand \selectlanguage[0]{\@gobble}% \providecommand \bibinfo[0]{\@secondoftwo}% \providecommand \bibfield[0]{\@secondoftwo}% \providecommand \translation[1]{[#1]}% \providecommand \BibitemOpen[0]{}% \providecommand \bibitemStop[0]{}% \providecommand \bibitemNoStop[0]{.\EOS\space}% \providecommand \EOS[0]{\spacefactor3000\relax}% \providecommand \BibitemShut[1]{\csname bibitem#1\endcsname}% \let\auto@bib@innerbib\@empty%</preamble> \bibitem[{\citenamefont{Parkin}\\emph{et~al.}(2008)\citenamefont{Parkin}, \citenamefont{Hayashi},\and\\citenamefont{Thomas}}]{Parkin}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{S.~S.~P.}\\bibnamefont{Parkin}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Hayashi}}, \and\\bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Thomas}},\}\href{\doibase 10.1126/science.1145799}{\bibfield{journal}{\bibinfo{journal}{Science}\}\textbf{\bibinfo{volume}{320}},\\bibinfo{pages}{190} (\bibinfo{year}{2008})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Fert}\\emph{et~al.}(2013)\citenamefont{Fert}, \citenamefont{Cros},\and\\citenamefont{Sampaio}}]{Fert}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Fert}}, \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Cros}}, \and\\bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Sampaio}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Nat. Nanotechnol.}\}\textbf{\bibinfo{volume}{8}},\\bibinfo{pages}{152} (\bibinfo{year}{2013})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Dzyaloshinsky}(1958)}]{Dzyaloshinsky}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Dzyaloshinsky}},\}\href{\doibase http://dx.doi.org/10.1016/0022-3697(58)90076-3}{\bibfield{journal}{\bibinfo{journal}{J. Phys. Chem. Sol.}\}\textbf{\bibinfo{volume}{4}},\\bibinfo{pages}{241 } (\bibinfo{year}{1958})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Moriya}(1960)}]{Moriya}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Moriya}},\}\href{\doibase 10.1103/PhysRevLett.4.228}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{4}},\\bibinfo{pages}{228} (\bibinfo{year}{1960})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Moreau-Luchaire}\\emph{et~al.}(2016)\citenamefont{Moreau-Luchaire}, \citenamefont{Moutafis}, \citenamefont{Reyren}, \citenamefont{Sampaio}, \citenamefont{Vaz}, \citenamefont{Horne}, \citenamefont{Bouzehouane}, \citenamefont{Garcia}, \citenamefont{Deranlot}, \citenamefont{Warnicke}, \citenamefont{Wohlh\"{u}ter}, \citenamefont{George}, \citenamefont{Weigand}, \citenamefont{Raabe}, \citenamefont{Cros},\and\\citenamefont{Fert}}]{moreau-luchaire_additive_2016}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Moreau-Luchaire}}, \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Moutafis}}, \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Reyren}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Sampaio}}, \bibinfo{author}{\bibfnamefont{C.~a.~F.}\\bibnamefont{Vaz}}, \bibinfo{author}{\bibfnamefont{N.~V.}\\bibnamefont{Horne}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Bouzehouane}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Garcia}}, \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Deranlot}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Warnicke}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Wohlh\"{u}ter}}, \bibinfo{author}{\bibfnamefont{J.-M.}\\bibnamefont{George}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Weigand}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Raabe}}, \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Cros}}, \and\\bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Fert}},\}\href{\doibase 10.1038/nnano.2015.313}{\bibfield{journal}{\bibinfo{journal}{Nat. Nanotechnol.}\}\textbf{\bibinfo{volume}{11}},\\bibinfo{pages}{444} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Chen}\\emph{et~al.}(2015)\citenamefont{Chen}, \citenamefont{Mascaraque}, \citenamefont{N'Diaye},\and\\citenamefont{Schmid}}]{chen_room_2015}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Chen}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Mascaraque}}, \bibinfo{author}{\bibfnamefont{A.~T.}\\bibnamefont{N'Diaye}}, \and\\bibinfo{author}{\bibfnamefont{A.~K.}\\bibnamefont{Schmid}},\}\href{\doibase 10.1063/1.4922726}{\bibfield{journal}{\bibinfo{journal}{Appl. Phys. Lett.}\}\textbf{\bibinfo{volume}{106}},\\bibinfo{pages}{242404} (\bibinfo{year}{2015})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Jiang}\\emph{et~al.}(2015)\citenamefont{Jiang}, \citenamefont{Upadhyaya}, \citenamefont{Zhang}, \citenamefont{Yu}, \citenamefont{Jungfleisch}, \citenamefont{Fradin}, \citenamefont{Pearson}, \citenamefont{Tserkovnyak}, \citenamefont{Wang}, \citenamefont{Heinonen}, \citenamefont{Velthuis},\and\\citenamefont{Hoffmann}}]{jiang_blowing_2015}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Jiang}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Upadhyaya}}, \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Zhang}}, \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Yu}}, \bibinfo{author}{\bibfnamefont{M.~B.}\\bibnamefont{Jungfleisch}}, \bibinfo{author}{\bibfnamefont{F.~Y.}\\bibnamefont{Fradin}}, \bibinfo{author}{\bibfnamefont{J.~E.}\\bibnamefont{Pearson}}, \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Tserkovnyak}}, \bibinfo{author}{\bibfnamefont{K.~L.}\\bibnamefont{Wang}}, \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Heinonen}}, \bibinfo{author}{\bibfnamefont{S.~G. E.~t.}\\bibnamefont{Velthuis}}, \and\\bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Hoffmann}},\}\href{\doibase 10.1126/science.aaa1442}{\bibfield{journal}{\bibinfo{journal}{Science}\}\textbf{\bibinfo{volume}{349}},\\bibinfo{pages}{283} (\bibinfo{year}{2015})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Boulle}\\emph{et~al.}(2016)\citenamefont{Boulle}, \citenamefont{Vogel}, \citenamefont{Yang}, \citenamefont{Pizzini}, \citenamefont{de~Souza~Chaves}, \citenamefont{Locatelli}, \citenamefont{Menteş}, \citenamefont{Sala}, \citenamefont{Buda-Prejbeanu}, \citenamefont{Klein}, \citenamefont{Belmeguenai}, \citenamefont{Roussign\'{e}}, \citenamefont{Stashkevich}, \citenamefont{Ch\'{e}rif}, \citenamefont{Aballe}, \citenamefont{Foerster}, \citenamefont{Chshiev}, \citenamefont{Auffret}, \citenamefont{Miron},\and\\citenamefont{Gaudin}}]{boulle_room-temperature_2016}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Boulle}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Vogel}}, \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Yang}}, \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Pizzini}}, \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{de~Souza~Chaves}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Locatelli}}, \bibinfo{author}{\bibfnamefont{T.~O.}\\bibnamefont{Menteş}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Sala}}, \bibinfo{author}{\bibfnamefont{L.~D.}\\bibnamefont{Buda-Prejbeanu}}, \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Klein}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Belmeguenai}}, \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Roussign\'{e}}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Stashkevich}}, \bibinfo{author}{\bibfnamefont{S.~M.}\\bibnamefont{Ch\'{e}rif}}, \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Aballe}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Foerster}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Chshiev}}, \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Auffret}}, \bibinfo{author}{\bibfnamefont{I.~M.}\\bibnamefont{Miron}}, \and\\bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gaudin}},\}\href{\doibase 10.1038/nnano.2015.315}{\bibfield{journal}{\bibinfo{journal}{Nat. Nanotechnol.}\}\textbf{\bibinfo{volume}{11}},\\bibinfo{pages}{449} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Heinze}\\emph{et~al.}(2011)\citenamefont{Heinze}, \citenamefont{von Bergmann}, \citenamefont{Menzel}, \citenamefont{Brede}, \citenamefont{Kubetzka}, \citenamefont{Wiesendanger}, \citenamefont{Bihlmayer},\and\\citenamefont{Bl\"{u}gel}}]{heinze_spontaneous_2011}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Heinze}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{von Bergmann}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Menzel}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Brede}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kubetzka}}, \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Wiesendanger}}, \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Bihlmayer}}, \and\\bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bl\"{u}gel}},\}\href{\doibase 10.1038/nphys2045}{\bibfield{journal}{\bibinfo{journal}{Nat. Phys.}\}\textbf{\bibinfo{volume}{7}},\\bibinfo{pages}{713} (\bibinfo{year}{2011})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Sonntag}\\emph{et~al.}(2014)\citenamefont{Sonntag}, \citenamefont{Hermenau}, \citenamefont{Krause},\and\\citenamefont{Wiesendanger}}]{sonntag_thermal_2014}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Sonntag}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hermenau}}, \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Krause}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Wiesendanger}},\}\href{\doibase 10.1103/PhysRevLett.113.077202}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{113}},\\bibinfo{pages}{077202} (\bibinfo{year}{2014})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Jensen}\\emph{et~al.}(1992)\citenamefont{Jensen}, \citenamefont{Dreyss\'{e}},\and\\citenamefont{Bennemann}}]{jensen_calculation_1992}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{P.~J.}\\bibnamefont{Jensen}}, \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Dreyss\'{e}}}, \and\\bibinfo{author}{\bibfnamefont{K.~H.}\\bibnamefont{Bennemann}},\}\href{http://iopscience.iop.org/article/10.1209/0295-5075/18/5/015/meta}{\bibfield{journal}{\bibinfo{journal}{Europhys. Lett.}\}\textbf{\bibinfo{volume}{18}},\\bibinfo{pages}{463} (\bibinfo{year}{1992})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Schneider}\\emph{et~al.}(1990)\citenamefont{Schneider}, \citenamefont{Bressler}, \citenamefont{Schuster}, \citenamefont{Kirschner}, \citenamefont{de~Miguel},\and\\citenamefont{Miranda}}]{schneider_curie_1990}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{C.~M.}\\bibnamefont{Schneider}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Bressler}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Schuster}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kirschner}}, \bibinfo{author}{\bibfnamefont{J.~J.}\\bibnamefont{de~Miguel}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Miranda}},\}\href{\doibase 10.1103/PhysRevLett.64.1059}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{64}},\\bibinfo{pages}{1059} (\bibinfo{year}{1990})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Elmers}\\emph{et~al.}(1995)\citenamefont{Elmers}, \citenamefont{Hauschild}, \citenamefont{Fritzsche}, \citenamefont{Liu}, \citenamefont{Gradmann},\and\\citenamefont{K\"{o}hler}}]{elmers_magnetic_1995}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{H.~J.}\\bibnamefont{Elmers}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hauschild}}, \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Fritzsche}}, \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Liu}}, \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Gradmann}}, \and\\bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{K\"{o}hler}},\}\href{https://journals.aps.org/prl/abstract/10.1103/PhysRevLett.75.2031}{\bibfield{journal}{\bibinfo{journal}{Physical Review Letters}\}\textbf{\bibinfo{volume}{75}},\\bibinfo{pages}{2031} (\bibinfo{year}{1995})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Sessi}\\emph{et~al.}(2009)\citenamefont{Sessi}, \citenamefont{Guisinger}, \citenamefont{Guest},\and\\citenamefont{Bode}}]{sessi_temperature_2009}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Sessi}}, \bibinfo{author}{\bibfnamefont{N.~P.}\\bibnamefont{Guisinger}}, \bibinfo{author}{\bibfnamefont{J.~R.}\\bibnamefont{Guest}}, \and\\bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Bode}},\}\href{\doibase 10.1103/PhysRevLett.103.167201}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{103}},\\bibinfo{pages}{167201} (\bibinfo{year}{2009})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{von Bergmann}\\emph{et~al.}(2006{\natexlab{a}})\citenamefont{von Bergmann}, \citenamefont{Bode},\and\\citenamefont{Wiesendanger}}]{von_bergmann_coverage-dependent_2006}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{von Bergmann}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Bode}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Wiesendanger}},\}\href{\doibase 10.1016/j.jmmm.2005.12.015}{\bibfield{journal}{\bibinfo{journal}{J. Magn. Magn. Mater.}\}\textbf{\bibinfo{volume}{305}},\\bibinfo{pages}{279} (\bibinfo{year}{2006}{\natexlab{a}})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Izyumov}(1984)}]{izyumov_modulated_1984}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{Y.~A.}\\bibnamefont{Izyumov}},\}\href{\doibase 10.1070/PU1984v027n11ABEH004120}{\bibfield{journal}{\bibinfo{journal}{Sov. Phys. Usp.}\}\textbf{\bibinfo{volume}{27}},\\bibinfo{pages}{845} (\bibinfo{year}{1984})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Rossat-Mignod}\\emph{et~al.}(1979)\citenamefont{Rossat-Mignod}, \citenamefont{Burlet}, \citenamefont{Vogt},\and\\citenamefont{Lander}}]{Rossat-Mignod}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Rossat-Mignod}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Burlet}}, \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Vogt}}, \and\\bibinfo{author}{\bibfnamefont{G.~H.}\\bibnamefont{Lander}},\}\href{http://stacks.iop.org/0022-3719/12/i=6/a=021}{\bibfield{journal}{\bibinfo{journal}{J. Phys. C: Sol. State Phys.}\}\textbf{\bibinfo{volume}{12}},\\bibinfo{pages}{1101} (\bibinfo{year}{1979})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Akimitsu}\\emph{et~al.}(1978)\citenamefont{Akimitsu}, \citenamefont{Siratori}, \citenamefont{Shirane}, \citenamefont{Iizumi},\and\\citenamefont{Watanabe}}]{Akimitsu}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Akimitsu}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Siratori}}, \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Shirane}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Iizumi}}, \and\\bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Watanabe}},\}\href{\doibase 10.1143/JPSJ.44.172}{\bibfield{journal}{\bibinfo{journal}{J. Phys. Soc. Jpn.}\}\textbf{\bibinfo{volume}{44}},\\bibinfo{pages}{172} (\bibinfo{year}{1978})}\BibitemShut{NoStop}% \bibitem[{sup()}]{supp}% \BibitemOpen \href@noop {} {\bibinfo{journal}{Supplemental Material describing the details of the experiments, the mean-field calculations and the Monte Carlo simulations. It also includes Refs.~\cite{arblaster_crystallographic_2010,rozsa_magnetic_2015,rozsa_complex_2016}}\}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Hsu}\\emph{et~al.}(2016)\citenamefont{Hsu}, \citenamefont{Finco}, \citenamefont{Schmidt}, \citenamefont{Kubetzka}, \citenamefont{von Bergmann},\and\\citenamefont{Wiesendanger}}]{hsu_guiding_2016}% \BibitemOpen \bibfield{journal}{ }\bibfield{author}{\bibinfo{author}{\bibfnamefont{P.-J.}\\bibnamefont{Hsu}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Finco}}, \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Schmidt}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kubetzka}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{von Bergmann}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Wiesendanger}},\}\href{\doibase 10.1103/PhysRevLett.116.017201}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{116}},\\bibinfo{pages}{017201} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Hsu}\\emph{et~al.}(2017)\citenamefont{Hsu}, \citenamefont{Kubetzka}, \citenamefont{Finco}, \citenamefont{Romming}, \citenamefont{von Bergmann},\and\\citenamefont{Wiesendanger}}]{hsu_electric-field-driven_2017}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{P.-J.}\\bibnamefont{Hsu}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kubetzka}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Finco}}, \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Romming}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{von Bergmann}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Wiesendanger}},\}\href{\doibase doi:10.1038/nnano.2016.234}{\bibfield{journal}{\bibinfo{journal}{Nat. Nanotechnol.}\}\textbf{\bibinfo{volume}{12}},\\bibinfo{pages}{123} (\bibinfo{year}{2017})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Finco}\\emph{et~al.}(2016)\citenamefont{Finco}, \citenamefont{Hsu}, \citenamefont{Kubetzka}, \citenamefont{von Bergmann},\and\\citenamefont{Wiesendanger}}]{finco_tailoring_2016}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Finco}}, \bibinfo{author}{\bibfnamefont{P.-J.}\\bibnamefont{Hsu}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kubetzka}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{von Bergmann}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Wiesendanger}},\}\href{\doibase 10.1103/PhysRevB.94.214402}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. B}\}\textbf{\bibinfo{volume}{94}},\\bibinfo{pages}{214402} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Hasselberg}\\emph{et~al.}(2015)\citenamefont{Hasselberg}, \citenamefont{Yanes}, \citenamefont{Hinzke}, \citenamefont{Sessi}, \citenamefont{Bode}, \citenamefont{Szunyogh},\and\\citenamefont{Nowak}}]{hasselberg_thermal_2015}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Hasselberg}}, \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Yanes}}, \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Hinzke}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Sessi}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Bode}}, \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Szunyogh}}, \and\\bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Nowak}},\}\href{http://link.aps.org/doi/10.1103/PhysRevB.91.064402}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. B}\}\textbf{\bibinfo{volume}{91}},\\bibinfo{pages}{064402} (\bibinfo{year}{2015})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Zakeri}\\emph{et~al.}(2013)\citenamefont{Zakeri}, \citenamefont{Chuang}, \citenamefont{Ernst}, \citenamefont{Sandratskii}, \citenamefont{Buczek}, \citenamefont{Qin}, \citenamefont{Zhang},\and\\citenamefont{Kirschner}}]{zakeri_direct_2013}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Zakeri}}, \bibinfo{author}{\bibfnamefont{T.-H.}\\bibnamefont{Chuang}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Ernst}}, \bibinfo{author}{\bibfnamefont{L.~M.}\\bibnamefont{Sandratskii}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Buczek}}, \bibinfo{author}{\bibfnamefont{H.~J.}\\bibnamefont{Qin}}, \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Zhang}}, \and\\bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kirschner}},\}\href{\doibase 10.1038/nnano.2013.188}{\bibfield{journal}{\bibinfo{journal}{Nat. Nanotechnol.}\}\textbf{\bibinfo{volume}{8}},\\bibinfo{pages}{853} (\bibinfo{year}{2013})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Meyerheim}\\emph{et~al.}(2009)\citenamefont{Meyerheim}, \citenamefont{Tonnerre}, \citenamefont{Sandratskii}, \citenamefont{Tolentino}, \citenamefont{Przybylski}, \citenamefont{Gabi}, \citenamefont{Yildiz}, \citenamefont{Fu}, \citenamefont{Bontempi}, \citenamefont{Grenier},\and\\citenamefont{Kirschner}}]{meyerheim_new_2009}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{H.~L.}\\bibnamefont{Meyerheim}}, \bibinfo{author}{\bibfnamefont{J.-M.}\\bibnamefont{Tonnerre}}, \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Sandratskii}}, \bibinfo{author}{\bibfnamefont{H.~C.~N.}\\bibnamefont{Tolentino}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Przybylski}}, \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Gabi}}, \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Yildiz}}, \bibinfo{author}{\bibfnamefont{X.~L.}\\bibnamefont{Fu}}, \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Bontempi}}, \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Grenier}}, \and\\bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kirschner}},\}\href{\doibase 10.1103/PhysRevLett.103.267202}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{103}},\\bibinfo{pages}{267202} (\bibinfo{year}{2009})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Meng}\\emph{et~al.}(2014)\citenamefont{Meng}, \citenamefont{Zakeri}, \citenamefont{Ernst}, \citenamefont{Chuang}, \citenamefont{Qin}, \citenamefont{Chen},\and\\citenamefont{Kirschner}}]{meng_direct_2014}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Meng}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Zakeri}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Ernst}}, \bibinfo{author}{\bibfnamefont{T.-H.}\\bibnamefont{Chuang}}, \bibinfo{author}{\bibfnamefont{H.~J.}\\bibnamefont{Qin}}, \bibinfo{author}{\bibfnamefont{Y.-J.}\\bibnamefont{Chen}}, \and\\bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kirschner}},\}\href{\doibase 10.1103/PhysRevB.90.174437}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. B}\}\textbf{\bibinfo{volume}{90}},\\bibinfo{pages}{174437} (\bibinfo{year}{2014})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Dup\'{e}}\\emph{et~al.}(2016)\citenamefont{Dup\'{e}}, \citenamefont{Bihlmayer}, \citenamefont{B\"{o}ttcher}, \citenamefont{Bl\"{u}gel},\and\\citenamefont{Heinze}}]{dupe_engineering_2016}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Dup\'{e}}}, \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Bihlmayer}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{B\"{o}ttcher}}, \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bl\"{u}gel}}, \and\\bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Heinze}},\}\href{\doibase 10.1038/ncomms11779}{\bibfield{journal}{\bibinfo{journal}{Nature Communications}\}\textbf{\bibinfo{volume}{7}},\\bibinfo{pages}{11779} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Mryasov}\\emph{et~al.}(2005)\citenamefont{Mryasov}, \citenamefont{Nowak}, \citenamefont{Guslienko},\and\\citenamefont{Chantrell}}]{mryasov_temperature-dependent_2005}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{O.~N.}\\bibnamefont{Mryasov}}, \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Nowak}}, \bibinfo{author}{\bibfnamefont{K.~Y.}\\bibnamefont{Guslienko}}, \and\\bibinfo{author}{\bibfnamefont{R.~W.}\\bibnamefont{Chantrell}},\}\href{\doibase 10.1209/epl/i2004-10404-2}{\bibfield{journal}{\bibinfo{journal}{Europhysics Letters (EPL)}\}\textbf{\bibinfo{volume}{69}},\\bibinfo{pages}{805} (\bibinfo{year}{2005})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Veber}\\emph{et~al.}(2008)\citenamefont{Veber}, \citenamefont{Fedin}, \citenamefont{Potapov}, \citenamefont{Maryunina}, \citenamefont{Romanenko}, \citenamefont{Sagdeev}, \citenamefont{Ovcharenko}, \citenamefont{Goldfarb},\and\\citenamefont{Bagryanskaya}}]{veber_high-field_2008}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{S.~L.}\\bibnamefont{Veber}}, \bibinfo{author}{\bibfnamefont{M.~V.}\\bibnamefont{Fedin}}, \bibinfo{author}{\bibfnamefont{A.~I.}\\bibnamefont{Potapov}}, \bibinfo{author}{\bibfnamefont{K.~Y.}\\bibnamefont{Maryunina}}, \bibinfo{author}{\bibfnamefont{G.~V.}\\bibnamefont{Romanenko}}, \bibinfo{author}{\bibfnamefont{R.~Z.}\\bibnamefont{Sagdeev}}, \bibinfo{author}{\bibfnamefont{V.~I.}\\bibnamefont{Ovcharenko}}, \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Goldfarb}}, \and\\bibinfo{author}{\bibfnamefont{E.~G.}\\bibnamefont{Bagryanskaya}},\}\href{\doibase 10.1021/ja710773u}{\bibfield{journal}{\bibinfo{journal}{Journal of the American Chemical Society}\}\textbf{\bibinfo{volume}{130}},\\bibinfo{pages}{2444} (\bibinfo{year}{2008})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{von Bergmann}\\emph{et~al.}(2006{\natexlab{b}})\citenamefont{von Bergmann}, \citenamefont{Heinze}, \citenamefont{Bode}, \citenamefont{Vedmedenko}, \citenamefont{Bihlmayer}, \citenamefont{Bl\"{u}gel},\and\\citenamefont{Wiesendanger}}]{von_bergmann_observation_2006}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{von Bergmann}}, \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Heinze}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Bode}}, \bibinfo{author}{\bibfnamefont{E.~Y.}\\bibnamefont{Vedmedenko}}, \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Bihlmayer}}, \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bl\"{u}gel}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Wiesendanger}},\}\href{\doibase 10.1103/PhysRevLett.96.167203}{\bibfield{journal}{\bibinfo{journal}{Physical Review Letters}\}\textbf{\bibinfo{volume}{96}},\\bibinfo{pages}{167203} (\bibinfo{year}{2006}{\natexlab{b}})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Simon}\\emph{et~al.}(2014)\citenamefont{Simon}, \citenamefont{Palot\'{a}s}, \citenamefont{Ujfalussy}, \citenamefont{De\'{a}k}, \citenamefont{Stocks},\and\\citenamefont{Szunyogh}}]{simon_spin-correlations_2014}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Simon}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Palot\'{a}s}}, \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Ujfalussy}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{De\'{a}k}}, \bibinfo{author}{\bibfnamefont{G.~M.}\\bibnamefont{Stocks}}, \and\\bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Szunyogh}},\}\href{\doibase 10.1088/0953-8984/26/18/186001}{\bibfield{journal}{\bibinfo{journal}{J. Phys.: Condens. Matter}\}\textbf{\bibinfo{volume}{26}},\\bibinfo{pages}{186001} (\bibinfo{year}{2014})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Dup\'{e}}\\emph{et~al.}(2014)\citenamefont{Dup\'{e}}, \citenamefont{Hoffmann}, \citenamefont{Paillard},\and\\citenamefont{Heinze}}]{dupe_tailoring_2014}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Dup\'{e}}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Hoffmann}}, \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Paillard}}, \and\\bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Heinze}},\}\href{http://www.nature.com/doifinder/10.1038/ncomms5030}{\bibfield{journal}{\bibinfo{journal}{Nat. Commun.}\}\textbf{\bibinfo{volume}{5}},\\bibinfo{pages}{4030} (\bibinfo{year}{2014})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{R\'{o}zsa}\\emph{et~al.}(2016{\natexlab{a}})\citenamefont{R\'{o}zsa}, \citenamefont{De\'{a}k}, \citenamefont{Simon}, \citenamefont{Yanes}, \citenamefont{Udvardi}, \citenamefont{Szunyogh},\and\\citenamefont{Nowak}}]{rozsa_skyrmions_2016}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{R\'{o}zsa}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{De\'{a}k}}, \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Simon}}, \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Yanes}}, \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Udvardi}}, \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Szunyogh}}, \and\\bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Nowak}},\}\href{\doibase 10.1103/PhysRevLett.117.157205}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{117}},\\bibinfo{pages}{157205} (\bibinfo{year}{2016}{\natexlab{a}})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Hagemeister}\\emph{et~al.}(2016)\citenamefont{Hagemeister}, \citenamefont{Vedmedenko},\and\\citenamefont{Wiesendanger}}]{Hagemeister}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hagemeister}}, \bibinfo{author}{\bibfnamefont{E.~Y.}\\bibnamefont{Vedmedenko}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Wiesendanger}},\}\href{\doibase 10.1103/PhysRevB.94.104434}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. B}\}\textbf{\bibinfo{volume}{94}},\\bibinfo{pages}{104434} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Arblaster}(2010)}]{arblaster_crystallographic_2010}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{J.~W.}\\bibnamefont{Arblaster}},\}\href{\doibase 10.1595/147106710X493124}{\bibfield{journal}{\bibinfo{journal}{Plat. Met. Rev.}\}\textbf{\bibinfo{volume}{54}},\\bibinfo{pages}{93} (\bibinfo{year}{2010})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{R\'{o}zsa}\\emph{et~al.}(2015)\citenamefont{R\'{o}zsa}, \citenamefont{Udvardi}, \citenamefont{Szunyogh},\and\\citenamefont{Szab\'{o}}}]{rozsa_magnetic_2015}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{R\'{o}zsa}}, \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Udvardi}}, \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Szunyogh}}, \and\\bibinfo{author}{\bibfnamefont{I.~A.}\\bibnamefont{Szab\'{o}}},\}\href{\doibase 10.1103/PhysRevB.91.144424}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. B}\}\textbf{\bibinfo{volume}{91}},\\bibinfo{pages}{144424} (\bibinfo{year}{2015})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{R\'{o}zsa}\\emph{et~al.}(2016{\natexlab{b}})\citenamefont{R\'{o}zsa}, \citenamefont{Simon}, \citenamefont{Palot\'{a}s}, \citenamefont{Udvardi},\and\\citenamefont{Szunyogh}}]{rozsa_complex_2016}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{R\'{o}zsa}}, \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Simon}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Palot\'{a}s}}, \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Udvardi}}, \and\\bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Szunyogh}},\}\href{http://link.aps.org/doi/10.1103/PhysRevB.93.024417}{\bibfield{journal}{\bibinfo{journal}{Phys. Rev. B}\}\textbf{\bibinfo{volume}{93}},\\bibinfo{pages}{024417} (\bibinfo{year}{2016}{\natexlab{b}})}\BibitemShut{NoStop}% \end{thebibliography}% \end{document}}
\caption{Object classification experiments in the depth domain, comparing \DECO{} and hand crafted mappings, using $5$ pre-trained networks %on ImageNet as feature extractors. Best results for each network-dataset combination are in \textbf{bold}, overall best in \textcolor{red}{\textbf{red bold}}. Extensive experiments were performed on VGG and Caffenet, while GoogleNet and ResNet act as reference.}
\caption{Overview of the \DECO\colorization network. On the left, we show the overall architecture; on the right, we show details of the residual block.}
\caption[]{\small Error rates ($\%$) on CIFAR-10 and CIFAR-100 datasets. All methods in the same group are trained for the same number of iterations. Results of our method are colored in \textcolor{blue}{blue}, and the best result for each network/dataset pair are {\bf bolded}. $^*$ indicates numbers which we take directly from \cite{huang2016densely}.}
\caption{Word error rate on the WSJ dataset. All approaches used a unidirectional encoder; results in {\color{mygray2}grey} indicate offline models.}
\caption{ The \blue{false alarm rate} of H1 single-detector triggers \blue{above a given $\rho$} during Advanced LIGO's O1 observing run. The background shown corresponds to the \texttt{PyCBC} search region (ii) \cite{Abbott2016,Canton:2014ena,Usman2016,pycbc}. All triggers occurred in times during which L1 was operating, but were not observed in coincidence with a trigger in L1. The set of triggers shown is therefore a measure of noise events specific to H1. The shaded regions show one-, two-, and three-sigma Poisson uncertainties on the measured rate of noise events. In this example, the loudest background event has detection statistic $\rho_\mathrm{thresh} = 9.3$ with a false alarm rate $\mathrm{FAR}=7.4\yr^{-1}$. }
\caption{ \textit{Left}: The measured \blue{differential rate} $\Lambda_n p_n(\rho)$ of H1 single-detector triggers during Advanced LIGO's O1 observing run (blue points), extended with several possible models for the probability density $p_n(\rho)$. Specifically, we consider a flat background model (red), as well as power-law models $p_n(\rho)\propto\rho^{-4}$ (blue) and $p_n(\rho)\propto\rho^{-8}$ (green). Each of these background models is independently normalized following Eq. \eqref{lambda_n} and then summed with a \blue{Gaussian} kernel density estimation of the measured H1 background. Also shown is the \blue{Gaussian} KDE result itself (light blue), which falls exponentially with $\rho$. For reference, we include the expected distribution $\Lambda_s p_s(\rho)$ of astrophysical signals (dashed grey), marginalized over the measured rate $\Lambda_s$ of binary black hole mergers. \textit{Right}: The \blue{false alarm rate of single-detector events above a given $\rho$}, for each of the background models considered. As in Fig. \ref{H1Background}, the black curve marks the measured H1 background during Advanced LIGO's O1 observing run, while the shaded red regions show one, two, and three-sigma Poisson uncertainties on the background rate. The sharp drops in the left-hand plot between the measured background distribution and each of the power-law models correspond to the appearance of high-SNR tails in the cumulative rate of noise events. }
\caption{The ratio of the stellar to dark matter mass as a function of dark matter mass predicted by \bluetides. The top panel shows the full distribution of sources at $z=8$ with large points denoting the median and the error-bars showing the central $68\%$ range. The lower panel shows only the median values for bins containing $>10$ galaxies at $z\in\{14,13,12,11,10,9,8\}$. Tabulated values of the median ratios are given in Table \ref{tab:DM_stellar}.}
\caption{The galaxy stellar mass function predicted by \bluetides\at$z\in\{8,9,10,11,12,13,14,15\}$. The right-hand axis shows the total number of galaxies in \bluetides\in each$\Delta\log_{10}M=0.2$ mass bin. The points and grey line show observational constraints from \citet{Song2015} at $z\approx 8$ corrected to assume a \citet{Chabrier2003} initial mass function. The inset panel shows the number of objects with $\log_{10}(M_*/{\rm M_{\odot}})>8$ in the simulation volume as a function of redshift $z=15\to 8$. Tabulated quantities from \bluetides\are given in Table\ref{tab:GSMF}.}
\caption{The star formation rate distribution function predicted by \bluetides. The right-hand axis shows the total number of galaxies in \bluetides\in each$\Delta\log_{10}SFR=0.2$ bin. Solid lines show the dust-corrected (intrinsic) star formation rate distribution functions measured by \citet{Mashian2016} at $z\in\{7.9, 6.8, 4.9\}$. The \citet{Mashian2016} curves are corrected to assume a \citet{Chabrier2003} IMF using the calibrations proposed by \citet{KE2012}. Tabulated quantities of the \bluetides\predicted are given in Table\ref{tab:SFRDF}.}
\caption{The relationship between the specific star formation rate (${\rm SFR}/M_{*}$) and stellar mass predicted for galaxies at $z\in\{14,13,12,11,10,9,8\}$ by \bluetides. The top panel demonstrates the full distribution of sources at $z=8$ with the points denoting the median and the error-bars showing the central $68\%$ range. The lower panel shows only the median values for bins containing $>10$ galaxies at $z\in\{14,13,12,11,10,9,8\}$. The median specific star formation rates in stellar mass bins predicted by \bluetides\are tabulated in Table\ref{tab:physical}.}
\caption{The relationship between the mean stellar age and stellar mass predicted for galaxies at $z\in\{14,13,12,11,10,9,8\}$ by \bluetides. The top panel demonstrates the full distribution of sources at $z=8$ with the points denoting the median and the error-bars showing the central $68\%$ range. The lower panel shows only the median values for bins containing $>10$ galaxies at $z\in\{14,13,12,11,10,9,8\}$. The median ages in stellar mass bins predicted by \bluetides\are tabulated in Table\ref{tab:physical}.}
\caption{The stellar (light points) and star forming gas (dark points) metallicities of galaxies in \bluetides. The 2D histogram in the top panel shows all objects with $M>10^{8}\,{\rm M_{\odot}}$ at $z=8$. Points denoting the median and central $68\%$. The lower panel shows only the median values for bins containing $>10$ galaxies at $z\in\{14,13,12,11,10,9,8\}$. Observational constraints from \citet{Maiolino2008}, \citet{Mannucci2009}, and \citet{Faisst2016} at $z\sim 3.5$, $z\sim 3.1$, and $z\sim 5$ respectively are also shown. Observational measurements of the stellar mass assume a \citet{Chabrier2003} initial mass function and metallicities were converted to a mass-fraction assuming $12+\log_{10}(O/H)_{\odot} = 8.69$ and $Z_{\odot}=0.02$. The median metallicities in stellar mass bins predicted by \bluetides\are tabulated in Table\ref{tab:physical}.}
\caption{The effective escape fraction of far-UV ($150\,{\rm nm}$) photons as a function of stellar mass. The median far-UV escape fractions in stellar mass bins predicted by \bluetides\are tabulated in Table\ref{tab:L_fesc}.}
\caption{The intrinsic and dust attenuated far-UV mass-to-light ratios as function of stellar mass and redshift. The median intrinsic and observed far-UV mass-to-light ratios in stellar mass bins predicted by \bluetides\are tabulated in Table\ref{tab:MTOL}.}
\caption{Intrinsic (left panel) and dust attenuated (observed, right panel) rest-frame far-UV ($150\,{\rm nm}$) luminosity functions. Observations at $z\approx 8$ and $10.4$ from Bouwens et al.\(2015) are shown for comparison. The scale of the right-hand axis shows the number of galaxies in each bin magnitude in the simulation. Tabulated quantities of the \bluetides\predicted are given in Table\ref{tab:UVLF}.}
\caption{Intrinsic (left panel) and dust attenuated (observed, right panel) rest-frame far-UV ($150\,{\rm nm}$) luminosity functions. Observations at $z\approx 8$ and $10.4$ from Bouwens et al.\(2015) are shown for comparison. The scale of the right-hand axis shows the number of galaxies in each bin magnitude in the simulation. Tabulated quantities of the \bluetides\predicted are given in Table\ref{tab:UVLF}.}
\caption{{\em Top-panel} The fraction of galaxies in halos hosting a SMBH. {\em Middle/bottom-panels} The relationship between the mass of the central super-massive black hole and the stellar mass of galaxies in \bluetides. The top panel demonstrates the full distribution of sources at $z=8$ with the points denoting the median and the error-bars showing the central $68\%$ range within uniform stellar mass bins. The lower panel shows only the median of stellar mass bins for $z\in\{14,13,12,11,10,9,8\}$. Median SMBH masses in stellar mass bins are tabulated in Table \ref{tab:SMBH}.}
\caption{ {\bf Depth fusion results} using four uniformly spaced views around the object. (\textcolor{red}{a}) TSDF fusion produces noisy results. (\textcolor{red}{b}) TV-L1 fusion smoothes the shape, but also removes thin details. (\textcolor{red}{c}) Our approach learns from large 3D repositories and significantly improves fusion results in terms of noise reduction and surface completion.}
\caption{{\bf Complementary roles of the selective and exhaustive search.} RoIs by using the selective (\textcolor{myred}{red}) and the exhaustive (\textcolor{mygreen}{green}) search are depicted. }
\caption{{\bf ME R-CNN architecture.} Per-image convolutional network computes per-image feature map while the selective and the exhaustive search generates the RoIs (R). Then, an appropriate expert is assigned to each R. (\textcolor{myblue}{R$^\mathrm{H}$}, \textcolor{mygreen}{R$^\mathrm{S}$}, and \textcolor{myred}{R$^\mathrm{V}$} for H, S, and V expert, respectively.) The associated per-RoI feature maps which are the outputs of the RoI pooling layer are fed into the assigned experts. Exhaustive search is only used for the training process.}
\caption{Plot of the effective derivative of the axial form factor for the a12m220 ensemble as a function of source-sink separation $t$ for point- ({\color{knbgreen}$\medsquare$}) and smeared-sinks ({\color{kngreen}$\medcircle$}). Bin sizes of 1 (unbinned), 2, 3, and 4 are shown with increasingly lighter shades from left to right for both choices of sink smearing. }
\caption{Analysis of the a12m220 ensemble with 5000 bootstrap samples. (\textit{top left}) effective mass (\textit{top right}) axial effective derivative (\textit{middle left}) vector effective derivative are shown with smeared-sink ($\medsquare$) and point-sink($\medcircle$) correlation functions, with corresponding reconstructed fit curves plotted in light- and dark-green respectively. The grey regions encompass data not included in the analysis. The data is staggered for clarity. (\textit{middle right}) Stability plot of $\mathring{g}_A/\mathring{g}_V$ for ensemble a12m220. The preferred fit is presented by the solid black symbol, the green band shades the 68\% confidence interval and helps guide the eye. Variations of the fit region of the two-point correlator ({\color{kngreen}$\medsquare$}), $G_A(t)$ ({\color{kngreen}$\medtriangleup$}), and $G_V(t)$ ({\color{kngreen}$\meddiamond$}) are presented. The corresponding frequentist $p$-values are plotted below, with the dashed red line at $p=0.05$ discriminating the statistical significance of the fit results. Uncertainty of the fit variations are determined by 1000 bootstrap samples. (\textit{bottom left}) Bootstrap histogram of $\mathring{g}_A/\mathring{g}_V$. Different shaded regions mark the 68\% and 95\% confidence interval. The central value of $\mathring{g}_A/\mathring{g}_V$ is consistent with the median at the sub-percent level. (\textit{bottom right}) Bootstrap histogram for $\e_\pi$, discussed in Sec.~\ref{sec:e_pi}. The shaded regions are defined similarily to the $\mathring{g}_A/\mathring{g}_V$ histogram.}
\caption{QQplots of $-log_{10}$ p-values of analyzing association evidence between 14220 X-chromosome SNPs and meconium ileus in 3199 Cystic Fibrosis patients, under the XCI $M_1$ assumption (left), the no-XCI $M_2$ assumption (middle) and using $Z_{max}$ (right). Black circle $\bullet$ for $rs3788766$, blue up-pointing triangle {\color{blue}$\blacktriangle$} for $rs5905283$, green square {\color{green}$\blacksquare$} for $rs12839137$ and red down-pointing triangle {\color{red}$\blacktriangledown$} for $rs5905284$.}
\caption{BMA-based HPD intervals for $\beta$ and corresponding $log(BF_{AN})$ for 50 top ranked SNPs, selected from analyzing association evidence between 14220 X-chromosome SNPs and meconium ileus in 3199 Cystic Fibrosis patients. SNP are ordered based on their lower bounds of the HPD intervals. The four top SNPs identified by p-values are marked here using the same symbol: black circle $\bullet$ for $rs3788766$, blue up-pointing triangle {\color{blue}$\blacktriangle$} for $rs5905283$, green square {\color{green}$\blacksquare$} for $rs12839137$ and red down-pointing triangle {\color{red}$\blacktriangledown$} for $rs5905284$.}
\caption[The dependence of the value of \Qcrit on the integration time \tInt. The contours represent the reduced \chiSquared of the model comparison to the discharge probability]{(\textit{Colour online}) The dependence of the value of \Qcrit on the integration time \tInt for Ar-CO$_2$ and \NeCOtwo. The contours represent the reduced \chiSquared of the model comparison to the discharge probability for a given \tInt and \Qcrit. See text for details.}
\caption{\red{(Upper panel) {Three copies of the Floquet spectrum vs $\alpha$, horizontally shifted with respect to each other by $1+N/2$ and divided by $\pi/\tau$. %Under a translation of $\pi/\tau$ the spectrum matches with itself: The three curves constantly differ from each other by 1 along the vertical axis: each quasi-energy has its own partner shifted by $\pm\pi/\tau$.} %(the agreement is imperfect being $N$ finite). (Lower panel) The corresponding Floquet states are organised in pairs: in each pair the quasi-energies differ by $\pi/\tau$ and the two states are even and odd superpositions of symmetry broken states (we plot the (real) amplitudes of the members of one of such pairs in the basis of the eigenstates of $\widehat{S}^z$). Numerical parameters: $N=100,\,\tau=0.006,\,h_0=0.32,\,\phi=\pi,\,\lambda=0$.} }
\caption{\red{Panel (a): (upper plot) plot of $\mean{\log_{10}\Delta_0}$ and $\mean{\log_{10}\Delta_\pi}$ (see Eq.~\eqref{logelt:eqn}) vs $\phi$ fixing $\lambda=0$; (lower plot) plot of the corresponding fitting exponents $\alpha(\Delta_0)$ and $\alpha(\Delta_\pi)$ vs $\phi$ (see Eq.~\eqref{fittola:eqn}). Panel (b): (upper plot) plot of $\mean{\log_{10}\Delta_0}$ and $\mean{\log_{10}\Delta_\pi}$ (see Eq.~\eqref{logelt:eqn}) vs $\lambda$ fixing $\phi=\pi$; (lower plot) plot of the corresponding fitting exponents $\alpha(\Delta_0)$ and $\alpha(\Delta_\pi)$ vs $\lambda$ (see Eq.~\eqref{fittola:eqn}). As expected, there is $\pi$-spectral pairing -- $10^{\mean{\log_{10}\Delta_0}}\ll10^{\mean{\log_{10}\Delta_\pi}}$ and $\alpha(\Delta_\pi)>\alpha(\Delta_0)$-- when there is time-crystal behaviour. (Numerical parameters: $\tau=0.6,\,h=0.32,\,N=1600$ in the upper plots.) } }
\caption{Three traps create three rings of magnetic nanoparticles. The rings interact with one another (see \textcolor{blue}{Visualization 3}). [From Masajada et al., Opt. Lett. \textbf{38}, 3910 (2013).]}
\caption{Three traps create three rings of magnetic nanoparticles. The rings interact with one another (see \textcolor{blue}{Visualization 3}). [From Masajada et al., Opt. Lett. \textbf{38}, 3910 (2013).]}
\caption{ Hovering over an instance subset (e.g., for the \textcolor{myPurple}{\textbf{\sffamily \small NUM}}ber class) highlights its instances (purple dots) in the t-SNE projected view.}
\caption{{\bf 2D t-SNE embedding of the hand pose space}. \textcolor[rgb]{0,0,1}{{\it BigHand2.2M}} is represented by blue, \textcolor[rgb]{1,0,0}{ICVL} by red, and \textcolor[rgb]{0,1,0}{NYU} by green dots. The figures show (left) global view point space coverage, (middle) articulation space (25D), and (right) combined of global orientation and articulation coverage. Compared with existing datasets, the {\it BigHand2.2M} contains a more complete range of variation.}
\caption{Formation energies per unit length of the boundary of the most stable graphene/MoS$_2$ junction structures, labelled as I, II, III, and IV, as a function of S chemical potential. The atomic structure of Phase III is displayed in Fig.~\ref{structure_phase3} (see {\color{blue}Supplementary Fig. S4} for other phases).}
\caption{The total and boundary DOS of the Phase III graphene/MoS$_2$ junction. The FL is set at $E_\text{F}=0$ eV; results for the other junction structures are given in {\color{blue}Supplementary Fig. S6}.}
\caption{Spin-resolved DOS of the Phase III graphene/MoS$_2$ junction. The FL is set at $E_\text{F}=0$ eV. The results for the other junction structures are shown in {\color{blue}Supplementary Fig. S9}.}
\caption{Dependence of the size-ratio, $R_f/R_i$, on the viscosity ratio, $\eta_{in}/\eta_{out}$, for three fluid systems with different interfacial tensions, $\sigma$. ($\blacktriangle$) - silicone oils displaced by glycerin-water (immiscible - high interfacial tension: $24 mN/m <\sigma< 29 mN/m$) (\textcolor{blue}{$\blacksquare$}) - silicone oils displaced by mineral oils (immiscible - low interfacial tension: $1 mN/m <\sigma< 1.2 mN/m$). (\textcolor{red}{$\circ$}) - glycerin-water mixtures (miscible - negligible interfacial tension). Dotted vertical lines show boundaries of different regimes in miscible displacements as identified by \cite{irmgard}. Immiscible displacements are similar irrespective of the different interfacial tensions; miscible fluids differs from immiscible ones at high viscosity ratios}
\caption{ Dependence of the onset radius, $R_{onset}$, in immiscible fluids on the viscosity ratio, $\eta_{in}/\eta_{out}$ and the most unstable wavelength, $\lambda_c$. (a) shows the variation of $R_{onset}$ on $\eta_{in}/\eta_{out}$ for pairs of silicone-oil/mineral-oil fluids at $\lambda_c = $ 1.2 $\pm$0.1 mm, (b) shows the dependence of $R_{onset}$ on $\lambda_c$ when $\eta_{in}/\eta_{out} =$ 0.52 for a silicone-oil/mineral-oil pair. $R_{onset}$ has an approximately linear dependence on $\lambda_c$ (c) shows $R_{onset}/\lambda_c$ as a function of $\eta_{in}/\eta_{out}$ for two sets of immiscible fluids: ($\diamond$) - silicone oils/mineral oils with $1 mN/m <\sigma< 1.2 mN/m$ and (\textcolor{red}{$\ast$}) - silicone oils/glycerin-water mixtures with $24 mN/m <\sigma< 29 mN/m$. $R_{onset}/\lambda_c$ has a power-law dependence with an exponent 1.5$\pm $ 0.2}
\caption{Precision and recall for \MOTOLD (left) and \MOTNEW. The accompanying detector (ACF\cite{Dollar:2014:PAMI} for \MOTOLD, DPM \cite{Felzenszwalb:2010:PAMI} for \MOTNEW), applied on each frame of the sequence, is plotted as a solid line. We consider all annotated pedestrians, even fully occluded ones, explaining the relatively low recall of both detectors. Trackers typically do not provide a confidence and are thus plotted with dots. Note that, when compared to the full detector set (\textcolor{detcolor}{$\bullet$}), most trackers are only able to improve precision, but struggle to reduce the number of missed targets. % \fixmes{I am not sure where we see this.} \fixmea{You're right, I guess this conclusion is only true under the assumption that the trackers use \emph{all} detections, i.e. if you consider the rightmost tip of the curve} }
\caption{\small This table lists predicate recognition results for some object pairs. Images containing these pairs are listed in the first row, where the red and green boxes respectively correspond to the subjects and the objects. The most probable predicate predicted by different methods are listed in the following rows, in which \textbf{black} indicates wrong prediction and \textcolor{red}{red} indicates correct prediction.}
\caption{\small This figure illustrates some images and their corresponding scene graphs. The scene graphs are generated according to section \ref{sec:sg}. In the scene graphs, the \textbf{black} edges indicate wrong prediction, and the \textcolor{red}{red} edges indicate correct prediction.}
\caption{Phenotypes from \textsc{Trip}, central model on UCSD1+UCSD2, \textcolor{blue}{UCSD1}, and \textcolor{red}{UCSD2}. Some phenotypes appear in \textcolor[HTML]{9600FF}{both UCSD hospitals}}
\caption{\color{oxfordfootnote}{A plot of the pairs of values $(E_k,N_k)$ in the factorization ensemble $\mathbf{\mathfrak{F}}(10000)$ showing the typical band spectrum of a quantum system. This cannot be expected from number theory alone. As an example, the point $N=10969262131 = 47297\cdotp 231923$, $E=1.00441815$ is marked with a larger dot.}}
\caption{\color{oxfordfootnote}{Comparative plot of the density distribution of values $(E_k,x_k)$ calculated quantum mechanically for the simulator (left column) and counting exactly the primes (right column). For the quantum calculation, we have used a wide range of $j$ values : $\mathbf{\mathfrak{F}}(10000)$, $\mathbf{\mathfrak{F}}(400000)$ and an extreme value of $\mathbf{\mathfrak{F}}(4 \cdotp 10^{10})$(from top to bottom) versus the same figures computed counting the primes for the respective $j$. As expected, both distributions are similar. It is to be noted that to achieve this precision, the number of $\zeta(s)$ function zeroes required for the truncated series of $\pi(x)$ is reasonably small. The actual number of zeroes is different for each figure: $100$, $300$ and $1000$ (again from top to bottom). Due to the very high computational cost involved, the last classical calculation (bottom right) was not done using all primes. }}
\caption{ \color{oxfordfootnote}{Plot of the exact integrable probability density $d\mathcal{D}(\varrho,\psi)/ d\varrho =\varrho\psi(\varrho)^2$ for the factoring simulator Hamiltonian (in blue) versus the one calculated for the Penning trap Hamiltonian (in orange) with the same boundary condition at $E=1$. In order to become a physical realization of the factoring simulator both systems must have the same zeroes for $q\geq 1$ which is indeed the case.}}
\caption{\label{fig:gsmf} \textit{Top panels:} galaxy stellar mass function at $z=2-4$ compared to the observational data sets ({\color{gray}{$\bigstar$}}\citealt{Marchesini2009ApJ...701.1765M}, {\color{gray}{$\bullet$}}\citealt{Mortlock2011MNRAS.413.2845M}, {\color{gray}{{\hdashrule[0.6mm]{6mm}{3pt}{1.5mm 0.5mm}}}}\citealt{Santini2012}, {\color{gray}{\hdashrule[0.6mm]{6mm}{1.5pt}{1.5mm 0.5mm}}}\citealt{Ilbert2013}, {\color{gray}{\hdashrule[0.6mm]{6mm}{3pt}{1.5mm 0.4mm 0.4mm 0.4mm}}}\citealt{Muzzin2013}, {\color{gray}{$\blacktriangledown$}}\citealt{Tomczak2014}, {\color{gray}{\CIRCLE}}\citealt{Duncan2014}, {\color{gray}{$\blacktriangle$}}\citealt{Grazian2015}, {\color{gray}{\hdashrule[0.6mm]{6mm}{3pt}{0.4mm 0.4mm 0.4mm 0.4mm}}}\citealt{Huertas-Company2016}, {\color{gray}{$\pentagon$}}\citealt{Stefanon2016arXiv161109354S}, {\color{gray}{\hdashrule[0.6mm]{6mm}{1.5pt}{0.4mm 0.4mm 0.4mm 0.4mm}}}\citealt{Davidzon2017}). The shaded regions represent the 1$\sigma$ Poisson uncertainties. \textit{Bottom left panel:} number density of galaxies with masses above $M_*$ at $z=3.7$. Dashed, dash-dotted, solid and dotted thick lines represent samples of i) all galaxies; ii) galaxies with $\mathrm{SFR}<100\Msol\mathrm{yr}^{-1}$; iii) $\mathrm{SFR}<30\Msol\mathrm{yr}^{-1}$ and iv) $\mathrm{SFR}<1\Msol\mathrm{yr}^{-1}$. The black circle indicates the observationally estimated value and uncertainties from the S14 sample \citep{Straatman2014}. \textit{Bottom middle panel:} evolution of the number density of massive galaxies with $\log_{10}[M_*/{\rm M}_\odot]>10.6$. The observational data at lower redshifts ($z<3$; \citealt{Straatman2015}) is indicated with grey circles. \textit{Bottom right panel:} evolution of the MQG fraction. The MQG fraction with $\mathrm{SFR}<1\Msol\mathrm{yr}^{-1}$ of the model running with Eddington ratio, $\epsilon=0.1$ is shown with the dotted thin line for comparison.}
\caption{Plot of $f(n) = 4A(n,P_d)$ for $2 \leq n \leq 100$ with $d=1$ (\textcolor{red}{red circles}) below and $d=100$ (\textcolor{blue}{blue diamonds}) above.}
\caption{\textbf{\textit{Recognition} and \textit{co-occurrence} based failure cases:} \textit{Left:} A special aircraft is recognized as multiple `airplanes' (two sets of wings instead of one may cause the confusion), therefore, erroneous questions (marked in \textcolor{blue}{blue}) arise. \textit{Right:} Due to very frequent co-occurrence of green vegetable/food/fruit in food images, our VQG model generates questions (marked in \textcolor{green}{green}) about green vegetables even when they are missing. The five small images are few examples of how training set food images almost always contain greens.}
\caption{ Sample %training images from the MS-COCO dataset~\cite{lin2014microsoft}. Symmetry ground-truths (mid-column) are computed from human labels (Figure \ref{fig:Max Number of Labeled Symmetries}): line segments for reflection symmetry axes and red dots for rotation symmetry centers. %From left to right: Original Image, symmetry Ground Truth (labeled by at least 2 human raters, see Figure , and the Sym-Resnet output (training set). Right column: heatmaps for predicted symmetries, \textcolor{green}{green} for reflection symmetry axes, and \textcolor{red}{red} for rotation symmetry centers. }
\caption{The progression (left to right) of converting the human labeled ground truth symmetries into symmetry heatmaps ($ H $). The human labels are clustered to find the reflection symmetry axes and rotation symmetry centers. Reflection symmetry heatmap: \textcolor{green}{green}, rotation: \textcolor{red}{red}. }
\caption{Comparison of quantitative results including maximum F-measure (larger is better) and MAE (smaller is better). The best three results on each dataset are shown in \color[HTML]{FE0000}\textbf{red}, \color[HTML]{3166FF}\textbf{blue}\color{black}, and \color[HTML]{32CB00}\textbf{green} \color{black}, respectively. Note that the training set of DHSNet~\cite{liu2016dhsnet} includes the testing set of MSRA-B and Dut-OMRON, and the entire MSRA-B dataset is used as part of the training set of RFCN~\cite{wang2016saliency}. Corresponding test results are excluded here.\vspace{-3mm} }
\caption{ Quantitative evaluation of state-of-the-art SR algorithms: average PSNR/SSIM/IFC for scale factors $2\times$, $4\times$ and $8\times$. \red{\textbf{Red}} text indicates the best and \blue{\underline{blue}} text indicates the second best performance. }
\caption{Sketch of number density $n$, dust optical depth $\tau_{\rm{d}}$, and attenuation of ionizing radiation $\phi$ as a function of radius. The red dashed line shows the pressures of the wind bubble $P_{\rm{b}}$, the thermal gas pressure $P_{\rm{therm}}$ of the shell and lastly of the ambient medium. At very early and late times when the column density of the shell and/or the pressure from winds is low, the shell may be fully ionized (not shown). \blue{See also \citet{Martinez-Gonzalez2014}.}}
\caption{Summary of 1D shell dynamical models. \blue{Included and neglected physical processes are marked with \chk and -, respectively.} \label{tab:ModelSummary}}
\caption{Comparison of momentum $p$ deposited by the various feedback terms. The red line labeled ``hot" corresponds to feedback from hot shocked wind material during the adiabatic phase, the other terms are as in eq. (\ref{MomODE}), i.e.\ram pressure in blue, radiation pressure in yellow, and gravity, which has a negative contribution, in black. The parameters of the clouds examined in the two panels are the same as in Figure~\ref{fig:force_comparison}.}
\caption{\textbf{Arp\,305: Integrated\hi\contours from the GMRT low resolution map} overlaid on a FUV (\textit{GALEX}) image. The \hi\column density contour levels are N$_{HI}$ = 10$^{20}$ atoms cm $^{-2}$ (0.5, 1.7, 2.9, 4.1, 5.3, 6.4, 8.8, 11.1, 12.9). Major \hi\tidal features referred in the text are marked. The ellipse at the bottom left shows the size of low resolution\textcolor{black}{(31.8$^{\prime\prime}$ $\times$ 29.5$^{\prime\prime}$)} synthesised beam.}
\caption{\textbf{Arp\,305:\textcolor{black}{\textbf{\textit{Main figure}}} \hi\velocity field} from the GMRT low resolution cube for \hi\emission$>$ 3 $\sigma$. The areas in red \textcolor{black}{with} black contours have velocities $>$ the NGC\,4017 systemic velocity of 3439\km\and areas in blue\textcolor{black}{with} grey contours have velocities $<$ 3439 \km. The contours are in steps of 7 \km. The thick white contour is from the FUV (\textit{GALEX}) image. The GMRT low resolution beam (31.8$^{\prime\prime}$ $\times$ 29.5$^{\prime\prime}$) is shown in bottom right corner. \textcolor{black}{\textbf{\textit{Inset }}Zoom in on the Bridge TDG region showing velocity field of the GMRT high resolution \hi\cube for emission$>$ 3 $\sigma$. \textcolor{black}{The GMRT high resolution beam (14.3.$^{\prime\prime}$ $\times$ 11.8$^{\prime\prime}$) is shown in bottom left corner.}} }
\caption{Example translations. We italicize some {\em \color{blue} mis-translated} errors and highlight the {\bf \color{red} correct} ones in bold.}
\caption{ The overview of proposed prediction framework \textit{DESIRE}. First, DESIRE generates multiple plausible prediction samples $\hat{Y}$ via a CVAE-based RNN encoder-decoder (\textcolor{Purple}{Sample Generation Module}). Then the following module assigns a reward to the prediction samples at each time-step sequentially as IOC frameworks and learns displacements vector $\Delta{\hat{Y}}$ to regress the prediction hypotheses (\textcolor{Blue}{Ranking and Refinement Module}). The regressed prediction samples are refined by iterative feedback. The final prediction is the sample with the maximum accumulated future reward. Note that the flow via \textcolor{Aquamarine}{aquamarine-colored paths} is only available during the training phase.}
\caption{(a) A driving scenario: The white van may steer into left or right while trying to avoid a collision to other dynamic agents. \textit{DESIRE} produces accurate future predictions (shown as \textcolor{blue}{blue} paths) by tackling multi-modaility of future prediction while accounting for a rich set of both static and dynamic scene contexts. (b) \textit{DESIRE} generates a diverse set of hypothetical prediction samples, and then ranks and refines them through a deep IOC network. %\caption{Consider a driving scenario shown in (a). The white van may steer into left or right (shown as \textcolor{blue}{blue} lines) to follow either side of the road, while trying to keep away from the other dynamic agents to avoid a collision. In order to predict the dynamics of agents accurately, a model must explain for the multi-modality of future prediction while accounting for a rich set of both static and dynamic scene contexts. We present \textit{DESIRE} that naturally tackles the challenges in future prediction. \textit{DESIRE} generates a diverse set of hypothetical prediction samples that are ranked and refined through a deep IOC network (b). The IOC network is equipped with a Scene Context Fusion module that naturally combines both static and dynamic contextual information. % {(a) Predicting viable paths (\textcolor{Blue}{\textbf{blue arrows; thicker has higher chance}}) in a complex traffic scene. (b) Simplified overview of the proposed approach. (c) Sample prediction results under multiple plausible futures. }%\textcolor{blue}{Intro figure has three parts: (a) show the challenges or the problem (b) a small graphic illustration for the key ideas that help overcome those challenges (c) a sample output that shows those challenges have been successfully addressed.}} }
\caption{Improved \textit{DESIRE-SI} prediction samples (\textcolor{red}{red}) over iterations. Iterative regression refines the predictions closer to the ground truth future trajectory (\textcolor{blue}{blue}) matching with scene context.}
\caption{Execution time for selected layers of \gls{googlenet}, \gls{vgg} and \gls{alexnet} on the ARM\textregistered ~Cortex\textregistered-A57 CPU. \textbf{Lower is better}. }
\caption{Execution time for selected layers of \gls{googlenet}, \gls{vgg} and \gls{alexnet} on the Intel\textregistered ~Core\texttrademark ~i5-4570 CPU. \textbf{Lower is better}. }
\caption{\label{figure:results} Accuracy in \% (train/validation/test) of four selected models on our datasets, with a detailed evaluation of their ability to correctly understand specific instance types. Cell color indicates whether the corresponding instances were relatively \colorbox{red!35}{harder} or \colorbox{green!35}{easier} in comparison to the overall accuracy on the dataset, or whether the tendency is \colorbox{blue!25}{inconsistent} across train/validation/test accuracies.}
\caption{\label{tbl:results} Aggregate simulation results by degree of centralization and market design variant. Averages and some standard deviations, each based on a sample of size $500$ (10 networks with 50 time steps). \textbf{Boldface} and \textbf{\color{red} red boldface} mark best and extreme values in a row, \textbf{\em boldface italics} 2nd best values or values considered to be a good compromise, {\color{blue} blue} values considered too large. }
\caption{ (color online) Crystal nucleation initiated by the interference of density waves ahead of the solidification front on the right hand side of the crystal shown in Fig. \ref{fig:GFN1}: Snapshots of the particle density field were taken at dimensionless times $t = 1900, 2000$, and $2100${\color{blue}. omit ''that show''} The color scale was chosen so that it enhances the visibility of the density waves at the solid-liquid interface. (For computer animations see http://phasefield.hu/pages/2017pre/)}
\caption{Misclassification rates and Frobenius norm error averaged over 100 replications with $p=200$ for Models 1 and 2. The methods displayed are the estimator we proposed in Section \ref{proposed} (dashed and $\blacksquare$), the $L_1$-penalized Gaussian likelihood estimator (dashed and $\blacktriangle$), the Ledoit-Wolf-type estimator from \eqref{Ledoit_Wolf_Estimator} (dashed and $\newmoon$), Bayes (solid and {\Large \textasteriskcentered}), the method proposed by \citet{guo2010simultaneous} (dots and $\fullmoon$), the method proposed by \citet{mai2015multiclass} (dots and $\triangle$), and the method proposed by \citet{witten2011penalized} (dots and $\square$).}
\caption{The Pull model of information-update systems. \blue{Note that the arrows in the figure denote logical links rather than physical connections. The updates, requests, and responses are all transmitted through (wired or wireless) networks.}}
\caption{The average reward of three online RL methods as discount factor $\gamma$ rises from $0$ to $0.95$: (a) Random Warm Start RL (RWS-RL) when $T_{0}=5$ and $T_{0}=10$ respectively; (b) the proposed New Warm Start RL (NWS-RL) is able to start the online learning when the $1^{\text{st}}$ tuple is available, i.e., $T_{0}=1$. The\textbf{\textcolor{red}{{} Red value}} is the best and the \textcolor{blue}{\emph{blue value}} is the $2^{\text{nd}}$ best. \label{tab:AverageRwd_RWS_NWS}}
\caption{Dynamic power spectra of the first 5 seconds of the Type 1 burst shown in color map (top panel) and contour representation (bottom panel). {\color{blue} The power spectrum are ``Leahy'' normalized and the contour lines are drawn for power values of 10, 20 and 25.} A coherent feature at 363 Hz is seen. }
\caption{\label{fig:adv-training-not-working} Adversarial training \cite{huang2015-learning,kurakin2016-adversarial-1} does not work. This is a church window plot \cite{warde-farley2016-adversarial}. Each pixel \((i, j)\) (row index and column index pair) represents a data point \(\tilde{x}\) in the input space and \(\tilde{x} = x + \vb{h}\epsilon_j + \vb{v}\epsilon_i\), where \(\vb{h}\) is the direction computed by FGSM and \(\vb{v}\) is a random direction orthogonal to \(\vb{h}\). The \(\epsilon\) ranges from \([-0.5, 0.5]\) and \(\epsilon_{(\cdot)}\) is the interpolated value in between. The central black dot \tikz[baseline=-0.5ex]{\draw[fill=black] (0,0) circle (0.3ex)} represents the original data point \(x\), the orange dot (on the right of the center dot) \tikz[baseline=-0.5ex]{\draw[fill=orange,draw=none] (0,0) circle (0.3ex)} represents the last adversarial sample created from \(x\) via FGSM that is used in the adversarial training and the blue dot \tikz[baseline=-0.5ex]{\draw[fill=blue,draw=none] (0,0) circle (0.3ex)} represents a random adversarial sample created from \(x\) that cannot be recognized with adversarial training. The three digits below each image, from left to right, are the data samples that correspond to the black dot, orange dot and blue dot, respectively. \tikz[baseline=0.5ex]{\draw (0,0) rectangle (2.5ex,2.5ex)} ( \tikz[baseline=0.5ex]{\draw[fill=black,opacity=0.1] (0,0) rectangle (2.5ex,2.5ex)} ) represents the data samples that are always correctly (incorrectly) recognized by the model. \tikz[baseline=0.5ex]{\draw[fill=red,opacity=0.1] (0,0) rectangle (2.5ex,2.5ex)} represents the adversarial samples that can be correctly recognized without adversarial training only. And \tikz[baseline=0.5ex]{\draw[fill=green,opacity=0.1] (0,0) rectangle (2.5ex,2.5ex)} represents the data points that were correctly recognized with adversarial training only, i.e., the side effect of adversarial training.}
\caption{(color online) Evolution of the single particle energies in the core $^{100}$Sn as a function of the number of the valence neutrons. The following labels \textcolor{black}{$\circ$}$g_{7/2}$, \textcolor{red}{$\diamond$}$d_{5/2}$, \textcolor{green}{$\square$}$s_{1/2}$, \textcolor{blue}{$\triangle$}$d_{3/2}$, \textcolor{brown}{$\lhd$}$h_{11/2}$, \textcolor{Fuchsia}{$\triangledown$}$f_{7/2}$, \textcolor{cyan}{$\rhd$}$p_{3/2}$, \textcolor{magenta}{$\plus$}$p_{1/2}$, \textcolor{orange}{$\ast$}$h_{9/2}$, {\scriptsize \textcolor{OliveGreen}{A}}$f_{5/2}$ identify each single particle state.}
\caption{(color online) Evolution of the quasiparticle energies in the BCS (upper) and LN (lower) approximations as a function of the number of the valence neutrons in the $^{100}$Sn. The particle states are identify with the following labels: \textcolor{black}{$\circ$}$g_{7/2}$, \textcolor{red}{$\diamond$}$d_{5/2}$, \textcolor{green}{$\square$}$s_{1/2}$, \textcolor{blue}{$\triangle$}$d_{3/2}$, \textcolor{brown}{$\lhd$}$h_{11/2}$, \textcolor{Fuchsia}{$\triangledown$}$f_{7/2}$, \textcolor{cyan}{$\rhd$}$p_{3/2}$, \textcolor{magenta}{$\plus$}$p_{1/2}$, \textcolor{orange}{$\ast$}$h_{9/2}$, {\scriptsize \textcolor{OliveGreen}{A}}$f_{5/2}$.}
\caption{\label{tbl:results} Overall accuracy of the cross-lingual approaches for the target languages \textcolor{lightgray}{and English}.}
\caption{Brightness temperature FOV for non-isothermal large-scale prominences with gas pressure of 0.1 dyn $\mathrm{cm}^{-2}$ at $\lambda = 1.3$~mm. The \emph{top left} figure shows a map of the absorption coefficient, the \emph{top right} shows the source function and the \emph{bottom left} shows the optical thickness attenuation term. The resulting contribution function map is in the \emph{bottom right} hand panel. Integrating over each horizontal line of sight results in the brightness temperature (K) curve, \textcolor{green}{solid green line}.}
\caption{Brightness temperature FOV for non-isothermal large-scale prominences with gas pressure of 0.1 dyn cm$^{-2}$ at $\lambda = 9$~mm. The \emph{top left} figure shows a map of the absorption coefficient, the \emph{top right} shows the source function and the \emph{bottom left} shows the optical thickness attenuation term. The resulting contribution function map is in the \emph{bottom right} hand panel. Integrating over each horizontal line of sight results in the brightness temperature (K) curve, \textcolor{green}{solid green line}. The \textcolor{red}{dashed red line} shows the $\tau$ = 1 line.}
\caption{Variation of estimated (\textbf{solid black line}) and computed (\textcolor{red}{dot-dashed red line}) optical thickness with the field of view, for six isothermal prominence models at $\lambda$=1.3~mm. The field of view is orientated vertically in the solar atmosphere with the positive x-axis directed radially away from the Sun.}
\caption{Variation of estimated (\textbf{solid black line}) and computed (\textcolor{red}{dot-dashed red line}) mean emission measure with the field of view, for six isothermal prominence models. The field of view is orientated vertically in the solar atmosphere with the positive x-axis directed radially away from the Sun.}
\caption{Variation of estimated \textbf{solid black line}) and computed (\textcolor{red}{dot-dashed red line}) mean emission measure with the field of view, for six non-isothermal large-scale prominence models. The field of view is orientated vertically in the solar atmosphere with the positive x-axis directed radially away from the Sun.}
\caption{Training iterations vs test negative log-likelihoods on JSB Chorales dataset for full and diagonal models. Top row is for the Adam optimizer and the bottom row is for RMSProp. \Diagcolor curves are for the diagonal models and \fullcolor (gray in grayscale) curves are for full (regular) models. Left column is for VRNN, middle column is for LSTM and right column is for GRU. Legends show the average number of parameters used by top 6 models (F is for Full, D is for Diagonal models). This caption also applies to Figures \ref{fig:Piano-midi}, \ref{fig:Nottingham}, \ref{fig:MuseData}, with corresponding datasets.}
\caption{\label{fig:months_1_12} Months 1--12. Plots showing $\P[\exp\{Y\}>1.5|\text{data}]$ i.e. the posterior probability that the relative risk is greater than 1.5. Colour key: \protect\includegraphics[width=1em]{col1.png} [0.0,0.2], \protect\includegraphics[width=1em]{col2.png} (0.2,0.4], \protect\includegraphics[width=1em]{col3.png} (0.4,0.6], \protect\includegraphics[width=1em]{col4.png} (0.6,0.8], \protect\includegraphics[width=1em]{col5.png} (0.8,1.0].}
\caption{\label{fig:months_13_24}Months 13--24. Plots showing $\P[\exp\{Y\}>1.5|\text{data}]$ i.e. the posterior probability that the relative risk is greater than 1.5. Colour key: \protect\includegraphics[width=1em]{col1.png} [0.0,0.2], \protect\includegraphics[width=1em]{col2.png} (0.2,0.4], \protect\includegraphics[width=1em]{col3.png} (0.4,0.6], \protect\includegraphics[width=1em]{col4.png} (0.6,0.8], \protect\includegraphics[width=1em]{col5.png} (0.8,1.0].}
\caption{\label{fig:election_results} Showing voting results of 2010 (a) and 2015 (b) elections arond Manchester. Key: \protect\includegraphics[width=1em]{colCON.png} Conservative Party majority, \protect\includegraphics[width=1em]{colLAB.png} Labour Party majority, \protect\includegraphics[width=1em]{colLIB.png}Liberal Democrat Party majority}
\caption{\label{fig:predicted}Predicted party majorities under new boundaries, see text for details on how this was calculated. Key: \protect\includegraphics[width=1em]{colCON.png} Conservative Party majority, \protect\includegraphics[width=1em]{colLAB.png} Labour Party majority.}
\caption{\label{fig:winprob} Left to Right, illustrating the probability of a Conservative, Labour or Liberal Democrat win. Colour key: \protect\includegraphics[width=1em]{col1.png} [0.0,0.2], \protect\includegraphics[width=1em]{col2.png} (0.2,0.4], \protect\includegraphics[width=1em]{col3.png} (0.4,0.6], \protect\includegraphics[width=1em]{col4.png} (0.6,0.8], \protect\includegraphics[width=1em]{col5.png} (0.8,1.0].}
\caption{\label{fig:LowFreqActE2Re2&E3Re1} Momentum thickness growth (top) and modal amplitude growth (bottom) with low-frequency forcing. Cases with $\theta_0/w = 0.25$ (left) and $\theta_0/w = 0.05$ (right). In bottom figures, {\color{red2}$\square$.} represents forcing mode for $f^+/f_n = 0.500$ and {\color{green2}$\circ$} represents $f^+/f_n = 0.575$.}
\caption{\textcolor{red}{Proportion of pixels with negative powers (\%)}}
\caption{Some challenging images in PSUm \cite{Rauschert2011,Liu2013} (1st and 2nd columns) and NYm \cite{Cicconet2016} (3rd column) datasets with groundtruth in 1st row (a,b,c). our method in 2nd row (d,e,f) produces better results among El2016 \cite{Elawady2016} in 3rd row (g,h,i) and Loy2006 \cite{Loy2006} in 4th row (j,k,l). For each algorithm, the top five symmetry results is presented in such order: {\color{red} red}, {\color{yellow} yellow}, {\color{green} green}, {\color{blue} blue}, and {\color{magenta} magenta}. Each symmetry axis is shown in a straight line with squared endpoints. Best seen on screen.}
\caption{A visualization of the decoding process for an actual example from our English-German MT experiments. The output token at each timestep appears at the top of the figure, with lexical constraints enclosed in boxes. {\color{blue}\textit{Generation}} is shown in {\color{blue} blue}, {\color{dark_green}\textit{Starting}} new constraints in {\color{dark_green} green}, and {\color{red}\textit{Continuing}} constraints in {\color{red} red}. The function used to create the hypothesis at each timestep is written at the bottom. Each box in the grid represents a beam; a colored strip inside a beam represents an individual hypothesis in the beam's $ k $-best stack. Hypotheses with circles inside them are \textit{closed}, all other hypotheses are \textit{open}. (Best viewed in colour).}
\caption[ ]{Prediction error and training time for Decision Stream algorithm (\tikzcircle[gray,fill=DSColor]{4.4pt}) and decision trees with depths ranging from 3 to 15 (are denoted on the labels). X-axis~--- training time, Y-axis~--- prediction error for classification (left) and regression (right) tasks. Decision tree metrics: \tikzcircle[gray,fill=IColor]{4.2pt}~--- information gain, \tikzcircle[gray,fill=GColor]{4.2pt}~--- Gini impurity, \tikzcircle[black,fill=white]{4.2pt}~--- variance reduction, \tikzcircle[gray,fill=VGSColor]{4.2pt}~--- variance reduction and Gini impurity for continuous and categorical features, respectively.}
\caption{The optical setup for the measurement. Both pump and probe beams are incident from the backside (without grating) of the sample. Signal is detected on the first order diffraction emitted from the backside of the sample. Insert: Schematic diagram of the periodic gold stripes formed on a fused silica substrate. The nominal period $p$ is 587.0 nm or 479.2 nm, corresponding to first order optical diffraction of the probe light in air at around 45 and 60 degrees, respectively. {\red The grating spacing is significantly exagerated: in reality several gold strips are accommodated within the laser spot so that the light is well diffracted.}}
\caption{Comparison of learning algorithms in terms of robustness (best viewed in color). Mean and variance of accuracy are measured for the fixed \textcolor{\fix}{DNA sequence length of 6000} for 500 cases by changing one percent of the hidden message. The shaded line represents the standard deviation of the inference accuracy.}
\caption{\textcolor{\revisionOne}{DNA hiding scheme using synonymous codons. A watermark is a scheme \textcolor{\ediThree}{used} to deter unauthorized dissemination by marking hidden symbols or texts.} %For example, the three nucleotide bases (e.g., \texttt{ ACU, AUU, GGU, CCU}) are changed to amino acids (e.g., \texttt{T, I, G, P} ) according to the genetic code. %All of the amino acids except Methionine (\texttt{M}) and Tryptophan (\texttt{T}) have overlapping codons. For the conservation of amino acids, DNA watermarking can be changed to one of the synonymous codons.}
\caption{Projection of global mode 1 onto spatial local modes of frequency $\omega=1.111 - 0.052\i$: absolute values of the projection coefficients as functions of $x$. Legend: ({\color{blue}$\bullet$}) $k^+_1$; ({\color{green!60!black}$+$}) $k^+_2$; ({\color{red}$\times$}) $k^-_1$; (\protect\rule[1.5pt]{4mm}{1pt}) global mode component $\hat{u}_x(0,x)$ on axis; colored lines indicate the growth rates according to local analysis. }
\caption{Projection of global mode 2 onto spatial local modes of frequency $\omega=0.912 - 1.211\i$: absolute values of the projection coefficients as functions of $x$. Legend: ({\color{blue}$\bullet$}) $k^+_2$; ({\color{green!60!black}$+$}) $k^+_3$; ({\color{red}$\times$}) $k^\pm$; (\protect\rule[1.5pt]{4mm}{1pt}) global mode component $\hat{u}_x(0,x)$ on axis; colored lines indicate the growth rates according to local analysis. }
\caption{Pressure signal amplitudes along $x$ at the critical point, $r_c=0.92$. ($\bullet$) total pressure $|\hat{p}|$; ({\color{blue}\protect\rule[1.5pt]{4mm}{0.75pt}}) pressure associated with $k^+_1$ wave; ({\color{red}\protect\rule[1.5pt]{4mm}{1.5pt}}) `stripped' pressure $|\tilde{p}|$. Note that all signals are numerically zero at $x=0$.}
\caption{Influence of the numerical box length on eigenvalues. In all figures, different symbols denote different box lengths: (\protect\raisebox{0.5pt}{\tiny\color{red}$\blacksquare$}) $L=25$; ({\color{blue}$\bullet$}) $L=20$; ({\scriptsize\color{green!60!black}$\blacklozenge$}) $L=15$; (\protect\raisebox{0.5pt}{\scriptsize$\blacktriangle$}) $L=10$. \textit{a}) Arc branch eigenvalues. \textit{b}) Effective reflection coefficients according to (\ref{eqn:reflcoeff}), absolute values. \textit{c}) Effective reflection coefficients, phase. \textit{d}) Effective reflection coefficients, absolute value scaled with $L^3$. }
\caption{Arc branch spectra obtained in calculations with ({\color{blue}$\times$},{\color{green!60!black}$\bullet$}) and without ({\color{red}$+$}) absorbing layer. Circles mark the individual modes for which eigenfunctions are shown in Fig.~\ref{fig:evecs_downstreambc}.}
\caption{Embedding of the new flight modes \textcolor{magenta}{\tt LOITER\_3D} and \textcolor{magenta}{\tt EIGH\_SPHERE} into the {\tt ardupilot} code for fixed-wing aircraft ({\tt ArduPlane}). Speed and height are controlled by the \textcolor{blue}{total energy control system (TECS)}. Lateral navigation is provided by an \textcolor{red}{L1-controller} \cite{park04} in case of straight flight paths or a \textcolor{red}{PD-controller} in case of circles or circle segments. \label{fig:tasks}}
\caption{\label{fig:WedgeShapes} (Color online) Shape invariance of droplets trapped in SLIPS wedges. % (a) A 4-$\mu{\rm L}$ water droplet equilibrates at different positions within a SLIPS wedge by adjusting the opening angle, $\beta$. % (b) The radius of the droplet (normalised to eliminate variations in the apparent contact angle, $\theta = 100\pm 5^\circ$), is invariant upon changes in the opening angle of the wedge. The inset shows the dispensed and measured droplet volumes. % Error bars correspond to the standard deviation of the sample. The scale bar is 1 mm. % (c) Manual actuation of a droplet by reconfiguration of the SLIPS geometry. The actuation signal shifts the position the apex of the wedge (\textcolor{red}{$\blacktriangle$}). % The new prescribed equilibrium position (\textcolor{YellowOrange}{$\circ$}) is followed by the centre of the droplet (\textcolor{blue}{$\times$}). % (d) Equilibrium position and droplet trajectory for the sequence shown in (c). The droplet trajectory, here tracked by measuring the position of the centre of the osculating sphere in the frame of reference of the lab, $X_{\rm L}$, follows the imposed signal with a lag determined by the friction force acting on the liquid. }
\caption{Qualitative Results of top ranked keyframes on RAD. a) Liu~\etal b) Video2GIF c) Our model (from left). Video titles are shown on the left. Ground truth relevance labels are shown in {\color{blue}{Blue}}. P$=$Positive, N$=$Negative.}
\caption{We show video summaries created by Hecate~\cite{song2016click}, MMR~\cite{carbonell1998use}, our similarity model and our full summarization approach. The {\color{green} Green} number on the images depicts the frame number. We plot the ground truth relevance scores, marking the selected frames for the shown methods, and cluster annotations over the video in the bottom two rows. For cluster annotation, each color represents a unique cluster. Additional examples are provided in supplementary. }
\caption{Recall-Precision curve of the RAD testet for different methods. Our method ({\color{blue}Blue}) performs high in terms of mAP.}
\caption{Examples and corresponding sentence BLEU scores of translations using the pivot and likelihood methods in \cite{Cheng2016NeuralMT} and the proposed word-sampling method. We observe that our approach generates better translations than the methods in \cite{Cheng2016NeuralMT}. We italicize \protect \textcolor{blue}{\textit{correct translation segments}} which are no short than 2-grams.}
\caption{\textcolor{red}{C$_2$H$_3^+$ product yield (obtained from the average of 600 experimental cycles) as a function of the H$_2$ valve opening trigger time with respect to the VUV laser pulse (vertical line). To induce the reaction for this measurement, the C$_2$H$_2^+$ ions were accelerated to a kinetic energy sufficient to overcome the barrier without IR laser.}}
\caption{(a) Schematic of an interface having springs with modified stiffness $\chi k_1$. All other spring stiffnesses remain unchanged. (b) \tblue{Variation of the interface frequency as a function of $\gamma$ for the $3$ distinct values of $\chi$ listed in the legend. } Dashed curves show frequencies bounding the bandgap.}
\caption{{\em Examples of annotations in the \datasetshort{} dataset.} \textcolor{green}{{\bf Green}} indicates regions of near-constant shading (but with possibly varying reflectance). \textcolor{red}{{\bf Red}} indicates edges due to discontinuities in shape (surface normal or depth). \textcolor{cyan}{{\bf Cyan}} indicates edges due to discontinuities in illumination (cast shadows). Using these annotations, we can learn to classify regions of an image into different shading categories.}
\caption{Water level response surface at Marmande computed at $\mathcal{D}_{N_{\text{ref}}}$. Top: pGP using (a)~$N = 49$ snapshots and (b)~$N = 121$ snapshots. Bottom~: PC using (c)~$N=49$ snapshots and (d)~$N=121$ snapshots. Black dots represent the design of experiments used to construct the surrogate models. The colour map corresponds to the evaluation of the surrogate over the full data set $\mathcal{D}_{N_{\text{ref}}}$.{\color[rgb]{0.9,0.2,0.16}\protect\footnotemark}}
\caption{\red{Parameters forming the input to our calculations of $\Phi$, $\vec{a}$ and $K$, and the source of the distributions from which they are drawn. In this work we take approximate maximum-likelihood values for each; in future work full uncertainties will be estimated by Monte Carlo marginalisation. Subscripts `t' and `o' in the final row denote true and observed values respectively.}}
\caption{Plots in the $xy$-plane of ordered pairs {\color{red} $(p,\phi(p-1))$} (in red) and {\color{cyan}$(p+2,\phi(p+1))$} (in cyan) for twin primes $p,p+2$. There are no exceptional pairs visible in Figure \ref{fig:TP03}; that is, $\phi(p-1) \geq \phi(p+1)$ in each case. The exceptional pairs $2381, 2383$ and $3851, 3853$ are visible in Figure \ref{fig:TP06}. A smattering of exceptional pairs emerge as more twin primes are considered.}
\caption{ History of the observable universe with thermal inflation. Shaded regions: epochs dominated by {\color{red}radiation-like}, {\color{blue}matter-like}, or {\color{OliveGreen}vacuum-like} energy components. Solid lines: the evolution of $k$-modes (colors) and the Hubble radius (black). Subscripts denote the time when each $k$-mode is at the Hubble radius. }
\caption{\textcolor{\highlightcolor}{D2D performance is sensitive to sensor group size. %Simulated under $\SI{0.01}{\meter}^{-2}$ UE density, 20 dBm Bluetooth transmitter gain, -65 dBm noise level and 64 Byte package size. }}
\caption{\textcolor{\highlightcolor}{The quality of intra-group D2D communications significantly affected by the CSI estimation error. %Simulated under $\SI{0.01}{\meter}^{-2}$ UE density, 20 dBm Bluetooth transmitter gain and -90 dBm noise level. }}
\caption{\textcolor{\highlightcolor}{Average RACH delay per RA attempt with different number of served UEs under EAB and our D2D-based grouped RA approach}}
\caption{\label{Famp}Amplitude of the energy kick $\Delta E$ as a function of the small body rotation frequency $\omega$, computed by direct integration of the dynamics of a massless particle around a rotating dumb-bell, for \textcolor{black}{$q=3d$}, \textcolor{blue}{$q=5d$}, and \textcolor{red}{$q=7d$}, and for different reduced masses $\mu=0.01$ ($\square$), $\mu=0.1$ ($\ocircle$), $\mu=0.4$ ($\vartriangle$), and $\mu=0.5$ ($\triangledown$). Plain lines give amplitudes of the analytically determined kick function $\Delta E$ (\ref{DE}) using (\ref{W1}) and (\ref{W2}). Left panel: for the sake of clarity of the figure, amplitudes of the energy kick, $\underset{\phi}{\max}\arrowvert\Delta E\arrowvert$, are presented divided by the parameter $\mu(1-\mu)$. Inset: example of kick function $\Delta E(\phi)$ for $q=5d$, $\mu=0.1$, and $\omega=0.3\omega_0$ computed from direct integration of the dynamics of a massless particle around a rotating dumb-bell (\ding{53}). The green solid line gives the kick function $\Delta E(\phi)$ (\ref{DE}). Right panel: ratio $\underset{\phi}{\max}\arrowvert\Delta E/W_2\arrowvert$ with the same data as in the left panel. }
\caption[\red{Friendship network at T1 with weak, strong, and partnership ties}]{Friendship network at wave one. Partnership ties are marked red, weak ties with thin dotted lines, and strong ties as solid lines. Continuous node color represents well-being scores (red = low score, green = high score, white = missing). \red{Squared node shapes are male participants, whereas round node shapes are female.}}
\caption{\red{Descriptive statistics of aggregated emotional well-being scores}}
\caption{\red{Results of ordered SAOMs on weak and strong friendship networks, and emotional well-being}}
\caption[Ego-alter selection table]{\red{Heatplot of the ego-alter selection table. Only estimates for observed emotional well-being values (ranging from 2.0 to 7.0) are displayed. The colors of the cells indicate the relative gain to ego's objective function when choosing alters with various levels of emotional well-being (red = negative, yellow/white = positive).}}
\caption[\red{Friendship network at T3 with weak, strong, and partnership ties}]{Friendship network at wave three. Partnership ties are marked red, weak ties with thin dotted lines, and strong ties as solid lines. Continuous node color represents well-being scores (red = low score, green = high score, white = missing). \red{Squared node shapes are male participants, whereas round node shapes are female.}}
\caption{Cell Proposal Network (CPN): Top half shows the first network, which proposes \textit{N} bounding boxes and their scores. Bottom half shows the second network which generate segmentation masks for the \textit{N} proposals. {\color{ProcessBlue}Convolutional (blue)} (filter size inside the box), {\color{ForestGreen}max-pooling (green)}, {\color{red}fully connected (red)}, and {\color{YellowOrange}deconvolutional (orange)} layers, with the number of feature maps on top of each layer, are shown. {\color{Brown}$\rightarrow$ } indicates which feature maps are combined by summation. %ROI-pooling+concat layer extracts fixed size feature map (25x25x480) for each box by concatenating max-pooled features from four sets of feature maps. Proposed bounding boxes and segmentation masks after non-maxima suppression (NMS) are shown for a selected area from \textit{Fluo-N2DL-HeLa} dataset. Box color indicates proposal score, with bright green representing high score and bright red representing low score.}
\caption{a) Ground Truth (GT) graph ($g^{GT}$) showing 4 cells and one event of each type: {\color{red}mitosis (red)}, {\color{blue}death (blue)}, enter (black), exit (black) and {\color{ForestGreen}move (green)}. b) A proposal graph ($G$) constructed from 5 proposals, which contains the $g^{GT}$ as one of its sub-graphs.}
\caption{Datasets: A small region (200x200) from each dataset is shown along with the ground truth cell markers ({\color{red}$\bullet$}) and the boundaries of tracked cells.}
\caption{Assigning (\textcolor{red}{red} box) NOAA weather data, reported in terms of a geographic grid, to health care facilities (\textcolor{red}{red} dots), where the \textcolor{cyan}{blue} color gradient might represent temperature. \label{fig:weather_assign}}
\caption{Feature specific $h_j$ terms, where \textcolor{red}{red} highlights features with a negative association and \textcolor{blue}{blue} highlights those with a positive association. \label{tab:dn_all_res_coeff}}
\caption{The marginal effects of several select covariates, where \textcolor{blue}{blue} shows the kernel density of the original data and the \textcolor{red}{red lines} show the estimation. Rate (y-axis) vs. feature (x-axis). Note that in \ref{fig:me91} and \ref{fig:me101} no kernel density estimate is provided, as these plots are for binary features.\label{fig:margef}}
\caption{The difference in means and paired t-test p-value results, obtained by comparing temperature/humidity values among the bottom 10\% and top 10\% of hand-hygiene compliance rates, by facility (\textcolor{blue}{blue} indicates that either temperature, humidity, or both have a positive difference in means and a p-value $\leq .05$).\label{tab:ttest}}
\caption{Feature specific $h_j$ terms for the Facility 91 model, where \textcolor{red}{red} highlights features with a negative association and \textcolor{blue}{blue} highlights those with a positive association. \label{tab:facility_91_coeff}}
\caption{ Examples of questions from the CLEVR-Humans dataset, along with predicted programs and answers from our model. Question words that do not appear in CLEVR questions are \underline{underlined}. Some predicted programs exactly match the semantics of the question (\textcolor{OliveGreen}{green}); some programs closely match the question semantics (\textcolor{BurntOrange}{yellow}), and some programs appear unrelated to the question (\textcolor{BrickRed}{red}). }
\caption{Marginal orbits' {\color{red} Floquet exponent} $\nu$, defined by the elliptic linear stability coefficient $\lambda_i$ as $\lambda = \exp(\pm 2 \pi i \nu)$, and the Lyapunov exponent $\mu$, defined by $\lambda = \exp(\pm \mu)$, of stable-marginal orbits, or possibly hyperbolic (unstable) orbits. We display orbits with $\mu \leq 0.05$, which we fix as the upper limit on the numerical error.}
\caption{The {\color{red} Floquet exponents} $\nu_j$, where $\lambda_j = \exp(\pm 2 \pi i \nu_j)$ are the linear stability coefficients of four linearly stable periodic three-body orbits.}
\caption{\label{tab:Baseline-performance:-top}Perf: top-1 accuracy (\%, higher is better) on various datasets and parameter cost (\#par., lower is better) for a few baselines and several variants of our method. Rows 1,2: independent baseline performance.\emph{ VGG-B}: VGG \cite{simonyan2014very} architecture B. (S) - trained from scratch. (P) - pre-trained on ImageNet. Rows 3-7: (ours) controller network performance\emph{; $DAN_{sketch}$} as a base network outperforms \textbf{$DAN_{caltech-256}$} on most datasets. A controller network based on random weights ($DAN_{noise}$) works quite well given that its number of learned parameters is a fifth of the other methods. $DAN_{imagenet}$: controller networks initialized from VGG-B model pretrained on ImageNet. $DAN_{imagenet+sketch}$: selective control network based on both VGG-B(P) \& Sketch. We color code the\textcolor{red}{first}, \textcolor{green}{second} and \textcolor{blue}{third} highest values in each column (lowest for \#par).\#par: amortized number of weights learned to achieve said performance for all tasks divided by number of tasks addressed (lower is better).}
\caption{\label{fig:Controller-initialization-scheme} (\emph{a}) Controller initialization schemes. Mean loss averaged over 5 experiments for different ways of initializing controller modules, overlaid with minimal and maximal values. Random initialization performs the worst (\textcolor{red}{random}). Approximating the behavior of a fine-tuned network is slightly better (\textcolor{green}{linear\_approx}) and initializing by mimicking the base network (\textcolor{blue}{diagonal}) performs the best (\emph{b}) Predictability of a control network's overall accuracy average over all datasets, given its transferability measure. }
\caption{Top: Volume-integrated emission of the ULX bubble in the fiducial run (model L40v-2, $L_{\rm w} = 10^{40} \, {\rm erg\,s^{-1}}$, $v_{\rm w} = 10^{-2}c$) as a function of time. Red, blue, black and green lines show bolometric, optical ($5\times10^{14}$\,Hz), radio (5\,GHz) and X-ray (2\,keV) light curves. Bottom: Spectra corresponding to the early adiabatic phase of the forward shock (blue), the transition from adiabatic to radiative (red dashed), and fully radiative (green line).}
\caption{Number of nonzero elements per row {\color{royalblue}in factor $S$} for \HT~matrices from Examples~\ref{ex:holdor1}~and~\ref{ex:holdor2} in 2D and 3D.}
\caption{The three current trajectories on $D$- (\textcolor{blue}{---}) and $Q$-axis (\textcolor{green!50!black}{---}) used in the experiment.}
\caption{Experimental current-flux curves $\phisD(\isD,0)-\PhiM$ (\textcolor{blue}{---}) and $\phisQ(0,\isQ)$ (\textcolor{green!50!black}{---}); solid lines: uncompensated voltage drops; dashed lines: compensated voltage drops; dotted line: averaged curve.}
\caption{Global evolution and illustration of the temporal convergence of KE for the structured grid (left) and skewed grid (right) cases. \textcolor{blue}{$\medtriangleup$} fractional timestep method in the KEcons solver and a timestep $\Delta t = 10$ ms; \textcolor{black}{$\medtriangleup$} fractional timestep method in the KEcons solver and a timestep $\Delta t = 20$ ms; \textcolor{blue}{$\medcircle$} PISO in the KEcons solver and a timestep $\Delta t = 10$ ms; \textcolor{black}{$\medcircle$} PISO in the KEcons solver and a timestep $\Delta t = 20$ ms; \textcolor{blue}{$\medstar$} icoFoam with $\Delta t = 10$ ms; \textcolor{black}{$\medstar$} icoFoam with $\Delta t = 20$ ms; \mythickline{blue} Ham et al. \cite{hamiac-kecons} $\Delta t = 10$ ms; \mythickline{black} Ham et al. \cite{hamiac-kecons} $\Delta t = 20$ ms }
\caption{Contribution of the convective term to the total energy dissipation with $\Delta t = 20$ ms, $\Delta x = 0.03125$ m. The convection term is defined as $\sum_{P} V_P (\hat{u}_i \cdot \frac{\delta u_i u_j}{\delta x_j})$ for the fractional timestep. \mythickline{blue} structured mesh with fractional timestep; \mythickline{red} skewed mesh with fractional timestep; \textcolor{blue}{$\medcircle$} structured mesh with PISO; \textcolor{red}{$\medcircle$} skewed mesh with PISO.}
\caption{Illustration of the unstructured mesh case $\Delta x = 0.0625$ m (left) and global evolution with spatial and temporal convergence of KE (right). \protect\fbcrcl PISO with KEcons solver $\Delta t = 10$ ms and $\Delta x = 0.0625$ m; \mytriangle{black} icoFoam $\Delta t = 10$ ms and $\Delta x = 0.0625$ m; \textcolor{black}{$\medcircle$} PISO with KEcons solver $\Delta t = 20$ ms and $\Delta x = 0.0625$ m; \textcolor{black}{$\medtriangleup$} icoFoam $\Delta t = 20$ ms and $\Delta x = 0.0625$ m; \protect\frcrcl PISO with KEcons solver $\Delta t = 10$ ms and $\Delta x = 0.03125$ m; \mytriangle{red} icoFoam $\Delta t = 10$ ms and $\Delta x = 0.03125$ m; \textcolor{red}{$\medcircle$} PISO with KEcons solver $\Delta t = 20$ ms and $\Delta x = 0.03125$ m; \textcolor{red}{$\medtriangleup$} icoFoam $\Delta t = 20$ ms and $\Delta x = 0.03125$ m.}
\caption{Temporal convergence of KE. \textcolor{blue}{$\medsquare$} icoFoam with pressure correction $\Delta t = 10$ ms; \textcolor{black}{$\medsquare$} icoFoam with pressure correction and $\Delta t = 20$ ms; \textcolor{blue}{$\medstar$} icoFoam without pressure correction $\Delta t = 10$ ms; \textcolor{black}{$\medstar$} icoFoam without pressure correction $\Delta t = 20$ ms; \mythickline{blue} Ham et al. \cite{hamiac-kecons} $\Delta t = 10$ ms; \mythickline{black} Ham et al. \cite{hamiac-kecons} $\Delta t = 20$ ms.}
\caption{Spatial convergence of the error between the numerical and the analytical MMS solution for the structured mesh case, compared with the results of Shunn et al. \cite{shunn-MMS}. \mybarredcircle{black} velocity convergence; \mybarredcircle{blue} velocity convergence in Shunn et al. \cite{shunn-MMS}; \mybarredsquare{black} scalar convergence, \mybarredsquare{blue} scalar convergence in Shunn et al. \cite{shunn-MMS}; \mybarredtriangle{black} density convergence, \mybarredtriangle{blue} density convergence in Shunn et al. \cite{shunn-MMS}.}
\caption{Spatial convergence of the error between the numerical and the analytical MMS solution for the unstructured mesh case, compared with the results of Shunn et al. \cite{shunn-MMS}. \mybarredcircle{black} velocity convergence; \mybarredcircle{blue} velocity convergence in Shunn et al. \cite{shunn-MMS}; \mybarredsquare{black} scalar convergence, \mybarredsquare{blue} scalar convergence in Shunn et al. \cite{shunn-MMS}; \mybarredtriangle{black} density convergence, \mybarredtriangle{blue} density convergence in Shunn et al. \cite{shunn-MMS}.}
\caption{Left: global KE of the 2D analytical MMS solution (\mythickline{black}) as a function of time plotted along with percentage error in KE for the square mesh case (\mythickline{blue}) and the triangular mesh case (\mythickdashedline{blue}). Right: skew-symmetric error contribution to the total KE error for the structured (\mythickline{black}) and the triangular mesh (\mythickdashedline{black}).}
\caption{a) $\Delta\omega_{24}$ variation: \color{blue}- - without WAPSS, \color{red}$-\cdot-$with full order WAPSS,\color{black}--- with WAPSS reduced order.(b) $\Delta\omega_{12}$ variation: \color{blue}- - without WAPSS, \color{red}$-\cdot-$with full order WAPSS,\color{black}--- with WAPSS reduced order.(c) $\Delta\omega_{24}$ variation for different delay ($T_d$) in wide-area loop.(d) $\Delta\omega_{24}$ variation for different time-delay ($T_d$) in remote signal.}
\caption{(a) $\Delta\omega_{24}$ variation:\color{blue}- - without WAPSS,\color{red}$-\cdot-$ with synchronized WAPSS,\color{black}---with non-synchronized WAPSS.(b) $\Delta\omega_{12}$ variation:\color{blue}- - without WAPSS,\color{red}$-\cdot-$ with synchronized WAPSS,\color{black}---with non-synchronized WAPSS. (c) $\Delta\omega_{34}$ variation with synchronized WAPSS for different delays ($T_d$). (d) $\Delta\omega_{24}$ variation with synchronized WAPSS for different delays ($T_d$).}
\caption{(a) $\Delta\omega_{5,4}$ variation:\color{blue}- - without WAPSS,\color{red}$-\cdot-$ with synchronized WAPSS,\color{black}---with non-synchronized WAPSS. (b) $\Delta\omega_{10,4}$ variation:\color{blue}- - without WAPSS,\color{red}$-\cdot-$ with synchronized WAPSS,\color{black}---with non-synchronized WAPSS. (c) $\Delta\omega_{5,4}$ variation for synchronized WAPSS for different delays ($T_d$). (d) $\Delta\omega_{10,4}$ variation for synchronized WAPSS for different delays ($T_d$).}
\caption{Scatter plot of $N = 25$ complex-valued exponent pairs $X \sim N(130,12^2) $ (\protect\unfilledsquare) and potential candidate for common exponent (\protect\bluediamond) }
\caption{The runtime used by \kira{} to reduce topology \texttt{topo7} as defined in \Eq{eq:Beispiel1} and \Eq{eq:topo7}. The parameter $s$ describes the total power of propagators occurring in the numerator. $r_{\max}$ is set to 7. In addition, we also give the time $\bm{T_\pyred}$ used by the \pyred{} module within \kira{} to identify the linearly dependent equations. For comparison the runtime for the same reduction using \reduzetwo{} and \firefive{} is shown. \label{tab:topo7m}}
\caption{Same as \Tab{tab:topo7m} but for topology \texttt{topo4}. ($r_{\max}$ is set to 7.) In all reductions one mass scale is removed using the ratio $m_{2}^{2}=\frac{3}{14}m_{1}^{2}$. \reduzetwo{}, \firefive{} and \kira{} were initialized with 11 cores. \label{tab:topo4} }
\caption{The run time $\bm{T_\pyred}$ for \pyred{} which is called by \kira{} is shown and compared to the total time $\bm{T_\kira}$, which \kira{} needed for a complete reduction of the topology topo4 and topo5 keeping the full mass dependence. \kira{} was initialized with the options \texttt{--algebra} and \texttt{--parallel=13}.}
\caption{ Overview of the proposed models for both directions. \autoref{fig:m2l} depicts the model that learns a mapping from a motion to language by first \emph{encoding} the motion sequence $\vec{M} = (\vec{m}^{(1)} \ldots \vec{m}^{(N)})$ into a \emph{context vector} $\vec{c}$~(in \textcolor{purple}{purple}) using a \emph{stack of bidirectional RNNs}~(BRNNs, in \textcolor{yellow}{yellow}). The context vector is then \emph{decoded} by \emph{another stack of unidirectional RNNs}~(in \textcolor{green}{green}), which also takes the \emph{embedded word}~(in \textcolor{blue}{blue}) generated in the previous timestep as input. A fully-connected layer~(FC, in \textcolor{lightgreen}{light green}) produces the parameters of the output probability distribution, denoted as $\vec{\hat{y}}^{(t)}$. The \emph{decoder}~(in \textcolor{red}{red}) finally takes $\vec{\hat{y}}^{(t)}$ and transforms it into a concrete word $\hat{w}^{(t)}$. Combined, the model thus generates the corresponding description word by word until the special \texttt{EOS} token is emitted and the description $\vec{\hat{w}} = (\hat{w}^{(1)}, \hat{w}^{(2)}, \ldots)$ is obtained. The other direction from language to motion is depicted in \autoref{fig:l2m}. The model works in similar fashion but uses a description in natural language $\vec{w} = (w^{(1)} \ldots w^{(M)})$ to generate the corresponding whole-body motion $\vec{\hat{M}} = (\vec{\hat{m}}^{(1)}, \vec{\hat{m}}^{(2)}, \ldots)$. The models for both directions are trained individually and share no weights. }
\caption{{\color{red}{Number of image sets in each cluster of the ballet dataset.}}}
\caption{Experiment 1: experimental setup and results from shift-invariant deblurring of ``Lena'' image. The observed image (c) is obtained by blurring the reference image (a) with shift-invariant PSF (b), and then corrupting with white Gaussian noise of variance $\sigma^2 = 400$. Image blocks (d) are obtained by splitting observed image into $3 \times 3$ blocks with overlaps of $100 \times 100$ pixels among them. The 2D first-order interpolation weights (e) are of the same size as observed blocks. Plots (i\textendash j) show the image quality of deblurred images obtained for different strengths of regularization. The legends ``\textcolor{red}{Central}'', ``\textcolor{green}{Proposed}'', and ``\textcolor{blue}{Indpndt}'' represent results from the \emph{centralized}, the \emph{proposed} and the \emph{independent} deblurring methods, respectively. Plots (k\textendash l) show impact of extent of overlap on the image quality of deblurred images. The legends ``\textcolor{green}{Proposed-2x2}'' and ``\textcolor{blue}{Indpndt-2x2}'' represent the \emph{proposed} and \emph{independent} deblurring methods for the case when the image is split into $2\times 2$ blocks. Similarly, the other two legends represent the case when the image is split into $3 \times 3$ blocks.}
\caption{Experiment 2: results from shift-variant image deblurring comparing the image quality (in terms of SNR and SSIM) obtained by the three different deblurring methods for different strength of regularization. The legends ``\textcolor{red}{Central3x3}'', ``\textcolor{green}{Proposed-3x3}'', ``\textcolor{blue}{Indpndt-3x3}'' denotes the results from the \emph{centralized}, the \emph{proposed}, and \emph{independent} deblurring methods, respectively, when using only $3\times 3$ grid of PSFs. Similarly, other legends denotes for the results obtained when using $5\times 5$, $6 \times 6$ and $8 \times 8$ grid of PSFs sampled in the field-of-view.}
\caption{ Evaluation on VOT-2015 by EAO, the weighted means of accuracy, robustness and speed. %We run our method and SiamFc-3s \cite{bertinetto2016fully}, and report their speed (FPS). (*) values from the VOT-2015 results \cite{Kristan2015a} in EFO units, which roughly correspond to FPS. %(e.g. the speed of the NCC tracker is 140 FPS with 160 EFO). The \textcolor{red}{first} and \textcolor{green}{second} best scores are highlighted in color.}
\caption{The merged light curve of {\cblue the} $V$-band and 5100~{\AA} continuum and the light curve data of \cite{Miller1979} for Ark 120. % {\cblue The red dashed line represents} a sinusoid with a period of 20.5 years. Note that it is impossible to % intercalibrate the dataset of \cite{Miller1979}. We simply adjust the fluxes to make the mean and standard deviation to be at the same scales as those of the other datasets (see Section~\ref{sec_miller}). {\it We do not use the dataset of \cite{Miller1979} for analysis.}}
\caption{(Top) Power spectral density $P(f)$ and (bottom) $f\times P(f)$ of Ark 120 in the optical and X-ray band (0.3-10 keV). The normalization of the X-ray PSD is adjusted to match the optical PSD. Blue points represent the binned PSD and {\cblue the} grey line represents the raw PSD. {\cblue Slate blue} points represent the X-ray PSD from \cite{Lobban2018}. % Vertical dashed lines correspond to the period of 20.5 years.}
\caption{The best inferred PSD for different PSD models (blue dashed lines). Shaded areas represent the 1$\sigma$ and 2$\sigma$ error bands. Solid lines represent the PSD of the observed data and {\cblue the} grey horizontal dashed line represents the Poisson noise.}
\caption{Dependence of the estimated false alarm probability on the monitoring time length. {\cblue The} vertical dashed line % represents the baseline of the observed data. Beyond the data baseline, the false alarm probability is calculated using {\cblue artificial} light curves for Ark 120. Note that the false alarm probabilities depend on the % configurations of {\cblue artificial} data, which are described in Section 4.4.}
\caption{\cblue Two examples of the extraction of the broad H$\beta$ profile. The top panel shows the mean spectrum of our Lijiang %% observations between 2016 and 2017. The bottom panel shows the historical spectrum from November 22, 1976 digitalized from \cite{Carpriotti1982}, which has narrow wavelength coverage.}
\caption{(Left) H$\beta$ profiles from 1976 to 2017. (Right) Folded H$\beta$ profile series with phases for a complete cycle using a period {\cblue of} 20.5 years. Black % lines represent the overall mean profiles of each phase. Blue and red lines represent profiles from the first and second cycles of the series respectively. % Vertical lines show the wavelength center of {\cblue the} H$\beta$ line.}
\caption{\cblue (Left) The LS periodogram in 12 wavelength bins of the H$\beta$ profile. (Right) The ``multiband'' LS periodogram that combines the information from all the wavelength bins, which is calculated using the Python package \texttt{gatspy} (\citealt{VanderPlas2015s, VanderPlas2015}).}
\caption{Three examples of {\cblue artificial} light curves for Ark 120 used to calculate the false alarm probabilities in Section 4.4. Black points are the observed % data and blue points are {\cblue artificial} data.}
\caption{a) The fabrication starts with the deposition of niobium electrodes on bare SiC (left). After high temperature annealing, the metallic electrodes have been transformed to NbC and the entire surface is covered with FGL (center). A ribbon is then etched in the FGL film using O$_2$ plasma (right). b) Optical microscope image of the chip with 32 four-terminal devices. The scale bar is 1 mm. c) Scanning electron microscope image of a 5.2$\times$4.4~$\mu$m$^2$ ribbon. The scale bar is 1 $\mu$m. d) Transmission electron microscope image of the junction between NbC and SiC showing the FGL flowing continuously over the two materials. The dashed lines materialize continuous graphene layers at the junction between SiC and NbC. The scale bar is 10 nm. e) Resistance times width of the devices as function of the length. For each length 4 to 5 devices were used to determine the plotted mean value with standard deviation as an error bar. The resistance was measured at room temperature (\textcolor{blue}{$\bigcirc$}) and at 4~K (\textcolor{blue}{$\blacksquare$}). The expression $R.W=1400\times L+364$ $\Omega.\mu$m fits the experimental data (\textcolor{blue}{- - -}).\label{Fig2}}
\caption{We show examples of how atomic actions change over time in AVA. The text shows pairs of atomic actions for the people in \textcolor{red}{red} bounding boxes. Temporal information is key for recognizing many of the actions and appearance can substantially vary within an action category, such as opening a door or bottle.}
\caption{The bounding box and action annotations in sample frames of the AVA dataset. Each bounding box is associated with 1 pose action (in \textcolor{orange}{orange}), 0--3 interactions with objects (in \textcolor{red}{red}), and 0--3 interactions with other people (in \textcolor{blue}{blue}). Note that some of these actions require temporal context to accurately label. \label{fig:teaser} }
\caption{We show more examples of how atomic actions change over time in AVA. The text shows pairs of atomic actions for the people in \textcolor{red}{red} bounding boxes.}
\caption{\label{pascal} Pascal's triangle. The encircled coloured numbers correspond to the coefficients of the six polynomials up to multiplicative factors of the size $(n-1)!$ and alternating signs. \textcolor{red}{Red color} indicates a positive sign, \textcolor{blue}{blue color} a negative one.}
\caption{\label{fig:1} \textbf{(a)} Laser cooling transitions for the $A = 133$ isotope of barium II with hyperfine structure of the underlying states. \textbf{(b)} A single \ba\ion and an isotopically pure$^{132}\text{Ba}^+$ ion chain loaded from an enriched microgram source of barium atoms. \textbf{(c)} Laser loading scheme of \ba\for the\ssymbol\$\leftrightarrow$ \psymbol\transition. To Doppler cool\ba, the laser carrier $\nu_{0}^{b}$ is stabilized 4.218(10) GHz above the $^{138}\text{Ba}^+$ resonance. The frequency $\nu_{1}^{b}$, resulting from a second-order sideband at $\nu_{0}^b - 11.744$~GHz, depopulates the \ssymbol, $F$ = 0 state. The frequency $\nu_{h}^{b}$, resulting from a first-order sideband at $\nu_{0}^b -4.300$~GHz, Doppler cools any co-trapped barium II even isotopes and sympathetically cools \ba. This first-order sideband is scanned across the blue shaded region (to $\nu_{0}^b -3.800$~GHz) using a high bandwidth fiber EOM to Doppler heat any other barium II isotopes out of the ion trap.}
\caption{\label{fig:2} Measured hyperfine splittings of the \ssymbol, \psymbol, and \dsymbol\states in$^{133}$Ba$^+$. Solid red lines are fitted Lorenztian profiles. \textbf{(a)} Fluorescence from a single \ba atomic ion with the application of laser frequencies $\nu_{0}^{b}$ and $\nu_{0,1}^{r}$ while scanning laser frequency $\nu_{1}^{b}$. The peak of the fluorescence spectrum yields the \ssymbol\hyperfine qubit splitting$\Delta_1 = 9931(2)_\text{stat}$ MHz. \textbf{(b)} Fluorescence as a function of applied modulation frequency to a laser tuned slightly red of the \ssymbol, $F = 1$ $\leftrightarrow$ \psymbol, $F = 0$ transition. When the applied modulation frequency is near $\Delta_2$, the ion can spontaneously decay to the $F = 2$ states in the \dsymbol\manifold. The resulting decrease in fluorescence gives a\psymbol\hyperfine splitting of$\Delta_2 = 1840 (2)_\text{stat}$ MHz. \textbf{(c)} After applying laser frequencies $\nu_{0,1}^b$, $\nu_{0}^{b} - \Delta_2$, and $\nu_{0}^{r}$, an applied frequency near $\nu_{1}^{r}$ is scanned to repump \ba\out of the$F = 2$ states in the \dsymbol\manifold. The resulting increase in fluorescence rate yields a\dsymbol\hyperfine splitting of$\Delta_3 = 937 (3)_\text{stat}$ MHz. These measurements all have a $\pm$20 MHz systematic uncertainty primarily due to drift of the wavemeter used to stabilize the lasers.}
\caption{Isotope shifts of the \psymbol\$\leftrightarrow$ \ssymbol\and\psymbol\$\leftrightarrow$ \dsymbol\electronic transitions of barium atomic ions and hyperfine$\mathcal{A}$ and $\mathcal{B}$ constants. The isotope shift of the $i$-th electronic transition is defined relative to $^{138}\text{Ba}^+$ and is $\delta \nu^{i} \equiv \nu_{A}^i - \nu_{138}^i$. The isotope shifts of all barium atomic ions are positive with the exception of the isotope shift of the \psymbol\$\leftrightarrow$ \dsymbol\transition in$^{137}\text{Ba}^+$. The bolded values are spectroscopic measurements from this work and have a systematic uncertainty of $\pm$20 MHz. All other isotope shifts are reported from references \citep{hohle:1976, blatt:1982, wendt:1984, vanhove:1985, knab:1987, villemoes:1993}. Columns 3-8 are in MHz.}
\caption{\label{fig:3} King plot for the \psymbol\$\leftrightarrow$ \dsymbol\electronic transition as a function of the\ssymbol\$\leftrightarrow$ \psymbol\electronic transition. Each point represents a pair of barium isotopes ($^A$Ba$^+$,$^{138}$Ba$^+$), labeled by $A$, where the frequency shift is normalized by the mass difference $(\nu_A-\nu_{138})/(A-138)$. Red triangle: $^{133}$Ba$^+$, yellow square: $^{132}$Ba$^+$, and cyan diamond: $^{130}$Ba$^+$ include spectroscopic measurements from this work. Blue circles are reported isotope shifts taken from references \cite{wendt:1984, villemoes:1993}. Yellow and cyan circles are derived from reported isotope shifts of the \ssymbol\$\leftrightarrow$ \psymbol\transition\citep{wendt:1984}, and calculated isotope shifts of the \psymbol\$\leftrightarrow$ \dsymbol\transition\cite{wendt:1984, vanhove:1982}.}
\caption{\textbf{Interactive language learning example.} (a) During training, teacher interacts in natural language with learner about objects. The interactions are in the form of (1)~question-answer-feedback, (2)~statement-repeat-feedback, and (3)~statement from learner and then feedback from teacher. Certain forms of interactions may be excluded for certain set of object-direction combinations or objects (referred to as \emph{inactive combinations/objects}) during training. For example, the combination of \{\emph{avocado}, \emph{east}\} does not appear in question-answer sessions; the object \emph{orange} never appears in question-answer sessions but only in statement-repeat sessions. Teacher provides both sentence feedback as well as reward signal (denoted as {\color{blue}[$+$]} and {\color{red}[$-$]} in the figure). (b) During testing, teacher can ask question about objects around, including questions involving \emph{inactive combinations/objects} that have never been asked before, \emph{e.g.}, questions about the combination of \{\emph{avocado}, \emph{east}\} and questions about \emph{orange}. This testing setup involves \emph{compositional generalization} and \emph{knowledge transferring} settings and is used for evaluating the proposed approach (\emph{c.f. Section~\ref{sec:Exp}}). }
\caption{Comparison of different area calculations for symmetric contours with $p=2$. {\color{red}{$\times$}} indicates the areas calculated by finding the reparametrization, and {\color{blue}{$\square$}} indicates the areas calculated through Shanks transformation of the perturbative expansion. The results agree.}
\caption{The solid line corresponds to the SM value of $C_9^{\rm eff}(q^2)$. The dotted line corresponds to {\color{red}$C_9^{\rm eff, SM}(q^2)+\delta C_9^{\rm NP}$}, where $\delta C_9^{\rm NP}=-0.97-i\times2.13\;$ is the best fit value of the NP fit for $C_9$. The hadronic power corrections fit results are demonstrated by dashed lines via \textcolor{myblue}{$C_9^{\rm eff, SM}+\Delta C_9^{+,PC}$}, \textcolor{myorange}{$C_9^{\rm eff, SM}+\Delta C_9^{-,PC}$} and \textcolor{mygreen}{$C_9^{\rm eff, SM}+\Delta C_9^{0,PC}$}. \label{fig:DeltaC9}}
\caption{{\small Predicting the \textcolor{color0}{cart's position} and \textcolor{color1}{pendulum's angle} behaviour from the cart-pole dataset by applying the control signal of the testing episode to sampled future trajectories from the proposed GPSSM. Learning of the dynamics is demonstrated with \emph{observed} (upper row) and \emph{hidden} (lower row) velocities and with increasing number of training episodes. Ground truth is denoted with the marked lines.}}
\caption{Predicting the \textcolor{color0}{inner} and \textcolor{color1}{outer} pendulum's angle from the double pendulum dataset by applying the control signals of the testing episode to sampled future trajectories from the proposed GPSSM. Learning of the dynamics is demonstrated with \emph{observed} (upper row) and \emph{hidden} (lower row) angular velocities and with increasing number of training episodes. Ground truth is denoted with the marked lines.}
\caption{{\small Demonstration of the identified model that controls the non-linear dynamics of the actuator dataset. The model's fitting on the \textcolor{color0}{train data} and sampled \textcolor{color1}{future predictions}, after applying the control signal to the system. Ground truth is denoted with the marked lines.}}
\caption{% Spectra for~$\chi_{16,16,4}^{\text{r}}$ ({\blue{*}}) and for~$\chi_{16,16,4}^-$ ({\red{$\diamond$}}) as eigenvalue vs index for one broken vertex at~$(1,1,1)$ with the full spectrum shown and the inset showing only the interesting comparative region near $E\approx1.4635$. In the interactive online version, the figure zooms in from showing the entire spectrum to the restricted spectrum from $E\approx1.4570$ to~$E\approx1.4850$, which shows clearly the degeneracy of $E\approx1.4635$ being 2 for~$\chi_{16,16,4}^{\text{r}}$ and being nondegenerate for~$\chi_{16,16,4}^-$ at $E\approx1.4620$ and the other at~$E\approx1.4640$.% }
\caption{Vertical and horizontal strips. \textcolor{blue}{$H_1 \cup H_2 := L(S) \cap S$} and \textcolor{red}{$V_1 \cup V_2 := L^{-1}(S) \cap S$}.}
\caption{\small The structures of $G_{enc}$, $G_{dec}$ and $D$ networks in single-image and multi-image DR-GAN. {\color{blue}{Blue}} texts represent extra elements to learn the coefficient $\omega$ in the $G_{enc}$ of multi-image DR-GAN.}
\caption{Stacked number density of \redmap clusters in each richness bin. ({\it left panel}): Without any cut on $\pmem$. ({\it middle panel}): With $\pmem$>0.5. ({\it right panel}): With $\pmem$>0.8. The number of member galaxies in the outermost region decreases significantly as we apply a stricter $\pmem$ cut.}
\caption{(a): Simulated light curve of an occultation event for a \textcolor{red}{$r = 1.5$} km sized TNO located at $R_h = 40$ au, and traversed at impact parameter $b=0$. The occultation is assumed to occur at $\phi = 30\degree$, and is observed at a wavelength of $500 \ {\rm nm}$. The background object is assumed to be a $V= 13$ mag F6V star. The finite size of the star is taken into consideration. (b) and (c): The same light curve as (a), but sampled at a cadence of (b) 5\Hz and (c) 15\Hz. Solid and dotted lines in panels (b) and (c) show light curves of the occultation sampled with no time offset and with a time offset of$1/10$ and $1/30\ {\rm seconds}$, respectively.}
\caption{ \textcolor{red}{ Example of a scatter plot showing the measured fluxes of a bright star (TYC 6857-238-1, $m_{\rm V} = 10.7$) scaled by their average value against the centroid positions in the x-axis direction of the CMOS sensor. In this example, the star moves in the hour angle direction (parallel to the x-axis direction of the CMOS sensor) in a two-minute observation due to tracking imperfections of the mount. A solid line represents moving average of the fluxes with a window of 0.1 pixels. The standard deviation of the fluxes and of their moving average values are 0.058 and 0.014, respectively. For the data points used in the plot, the relative differences of the positions in the y-axis direction are smaller than 0.2 pixels. } }
\caption{Each {\color{tangerine} tangerine} dot represents the time period between two consecutive donations by the same donor to different campaigns. Each {\color{ocean} ocean} dot represents the life span of a campaign.}
\caption{The behaviour of \car schemes with \lag's contributions. The {\color{tangerine} tangerine} colored bars indicate funds deducted from a campaign. The {\color{ocean} ocean} colored bars indicate funds allocated to a campaign after redistribution.}
\caption{The behaviour of \cpr schemes with \lag's contributions. The {\color{tangerine} tangerine} colored bars indicate funds deducted from a campaign. The {\color{ocean} ocean} colored bars indicate funds allocated to a campaign after redistribution.}
\caption{Comparison of top responses generated for some input dialogue acts between different models. Errors are marked in color (\textcolor{red}{missing}, \textcolor{blue}{misplaced}, \textcolor{orange}{repeated}, \textcolor{green}{grammar} information). \textbf{$^{\dag}$} and \textbf{$^{\natural}$} denotes the baselines and the proposed models, respectively.}
\caption{Comparison of top responses generated for some input dialogue acts between different models. Errors are marked in color (\textcolor{red}{missing}, \textcolor{blue}{misplaced} information). \textit{All2*} are general models.}
\caption[\textcolor{}Evolution of significance of the pulsed detection for \psr ]{Evolution of corresponding $H$-statistics detected for the pulsation of \psr. The black dots and red squares denote the detection significance in the $H$-values and source counts in each 70-day time bin. A comparison of the detection significance between the pre-glitch/post-relaxation stages and the low-flux stage can clearly be seen as well.}
\caption{State reconstruction in a one-dimensional \textcolor{MyRED1}{register}. a)~Schematic representation the Bayesian method with shifting patch in one dimension, assuming nearest~neighbor light contamination. The grid represents the pixels of the sensor, the dots represent empty lattice sites, and the circles represent trapped atoms, where the red (black) filling depicts the estimated probability to be bright (dark). A patch is defined around three sites and the middle set of pixels is used to update the combined-state probabilities only for the atoms inside the patch. Then the patch is shifted, the left atom is excluded from the patch by marginalizing its probability, and a new site at the right is included. The shaded regions correspond to pixels that contain either no information or have already been used. b) State dependent imaging (SDI) of a one-dimensional register of neutral atoms (middle box) initialized in random states. The image is integrated along the vertical direction, and the integrated CCD counts are used to calculate the probability that a lattice site contains an atom in the bright state ($P(B)$) using the Bayesian update algorithm. Position detection imaging (PDI) is used before (top box) and after (bottom box) state detection to verify that the atoms remain trapped in the same lattice site.}{\label{fig:fig4}}
\caption{ Capacity factors $\text{CF}_n^W$ and $\text{CF}_n^S$ for onshore wind and solar PV for the European countries, derived from the EuroStat data~\cite{eurostat1,eurostat2,barometer}. The countries are sorted by their respective mean load $\langle L \rangle$ (in units of GW) over the \highlight{2000-2007} time series. \highlight{*: estimated values, see text for details.} }
\caption{ Scatter clouds for (blue) wind-only and (yellow) solar-only capacity layouts. The diagram plots the overall capacity factor (\ref{eq:CFEU}) vs.\the standard deviation of the overall mismatch (\ref{eq:DeltaEU}). The distribution (\ref{eq:Beta}) and the constraint (\ref{eq:k-factor}) with $K=2$ have been used for the Monte Carlo simulations. The white point in the upper left cloud corners indicate Pareto optimal layouts; see also Figure \ref{fig:examples}c. The line connecting the wind- and solar-only Pareto optimal layouts results from the interpolation \highlight{between these layouts in} (\ref{eq:9a}). The black point marked on this line represents the OPT layout with minimum LCOE. For comparison, the three triangle points mark the (orange) optimal CFprop, (green) optimal CFmax and (blue) optimised GAS layouts for $K=2$. }
\caption{\textcolor{red}{\label{tab:7}} Similar to Table~\ref{tab:2-2} but for the cold accretion scenario.}
\caption{[\textbf{Suspended seat}] The result of the parametric grid search optimization of the suspended seat attached with (a) TMD ($c_a=0$) and (b) NES ($k_a=0$). Optimization has been performed with respect to the stiffness (linear/nonlinear) and damping coefficients of the attachment, and the optimal solutions are marked by a red cross (\textcolor{red}{$\times$}) along with the numeric value of the optimal measure $\gamma$. Optimization of the response displacement (left subplots) and velocity (right subplots) are presented. Parameters without attachment are shown in~\cref{tab:sdof_opt}.}
\caption{[\textbf{Suspended deck-seat}] The result of parametric grid search optimization of the suspended deck-seat attached with (a) TMD ($c_a=0$) and (b) NES ($k_a=0$). Optimization has been performed with respect to the stiffness (linear/nonlinear) and damping coefficients of the attachment and the optimal solutions are marked by a red cross (\textcolor{red}{$\times$}) along with the numeric value of the optimal measure $\gamma$. Optimization of the response displacement (left figures) and velocity (right figures) are presented. Parameters without attachment are shown in~\cref{tab:tdof_opt}.}
\caption{ {\orange Statistical comparisons between the FROC curves shown in \figurename~\ref{fig:Polyp_FROC} for polyp detection (level of significance is $\alpha=0.05$). The curves are compared at 0.01 and .001 false positives per frame, because they coincide with the elbows of the performance curves where they yield relatively higher sensitivity. A red cell indicates that a pair of curves are statistically different in neither of the chosen operating point whereas a green cell indicates at which operating points a statistically significant difference is observed. }}
\caption{{\orange Statistical comparisons between the FROC curves shown in \figurename~\ref{fig:PE_FROC} for pulmonary embolism detection (level of significance is $\alpha$=0.05). Each cell presents a statistical comparison between a pair of FROC curves at 1, 2, 3, 4, and 5 false positives per volume. A red cell indicates that the two curves are not statistically different at any of the five operating points, but a green cell contains the operating points at which the two curves are statistically different. }}
\caption{{\orange Statistical comparisons between the ROC curves shown in \figurename~\ref{fig:QA_ROC} for frame classification (level of significance is $\alpha$=0.05). Each cell presents a statistical comparison between a pair of ROC curves at false positive rate of 10\%, 15\%, and 20\% (0.1, 0.15, and 0.2 on the horizontal axis). A red cell indicates that the two curves are not statistically different at any of the two operating points, but a green cell contains the operating points at which the two curves are statistically different. }}
\caption{{\orange Statistical comparisons between the boxplots shown in \figurename~\ref{fig:CIMT_boxplot}. The p-values larger than 0.05 are highlighted in red.}}
\caption{The test stage of lumen-intima and media-adventitia interface segmentation. (a) A test region of interest. (b) The corresponding confidence map generated by the convolutional neural network. The green and red colors indicate the likelihood of a lumen-intima interface and media-adventitia interface, respectively. (c) The thick probability band around each interface is thinned by selecting the largest probability for each interface in each column. (d) The step-like boundaries are smoothed using 2 open snakes. {\orange (e) Interface segmentation from the ground truth.}}
\caption{{\orange Convergence speed for a deeply fine-tuned CNN and CNNs trained from scratch with three different initialization techniques.}}
\caption{\red{Comparison of \textsc{sparc} and model BTFR statistics. $\rho$ is the Spearman's rank coefficient of the $\Delta R_\text{eff}-\Delta V_\text{f}$ correlation, $s_\text{BTFR}$ is the intrinsic BTFR scatter, $N_\text{f}$ is the number of galaxies with non-flat RCs, and $q$ is the quadratic BTFR curvature. The 3$^\text{rd}$ row shows $s_\text{BTFR}$ for $M_\text{b} > 10^{9.5} M_\odot$ galaxies only.}}
\caption{ (color online) Phase diagram of overlap with one fixed MT and one movable MT in $\tilde v_0-\tilde{f}_d$ plane. We used parameter values $\tilde k_m = 450$, $\tilde\omega = 20$, $\tilde{f}_s = 60$, $N = 5$ and $n_p = 15$. The points denote non-equilibriuam phases obtained from numerical solutions of full non-linear dynamical equations, Eq.(\ref{eq:dyn1}). We find the following phases: (i)~stable {\color{blue} $\diamond$}, (ii)~stable spiral \tikzcircle[black, fill=black]{2pt}, (iii)~stable limit cycle {\color{red} $\blacktriangle$}. The predictions of linear stability analysis are indicated by symbols $s$: linearly stable, $u$: linearly unstable, $us$: unstable spiral, $ss$: stable spiral. The lines denote phase boundaries calculated from Eq. (\ref{eq:pb_unStable}) -- the (green) dashed line is the boundary between $u$ and $us$, (blue) solid line is the boundary between $us$ and $ss$, and the (pink) beaded-line is the boundary between $s$ and $ss$. }
\caption{ (color online) Phase diagram when both the MTs of overlapping pair are considered movable, in the $\tilde v_0-\tilde{f}_d$ plane. The parameter values used $\tilde k_m = 450$, $\tilde\omega = 20$, $\tilde{f}_s = 60$, $N = 5$ and $n_p = 15$. Linear stability estimate of different phases are denoted by $s$: { linearly stable} phase, $u$: { linearly unstable} phase, $us$: { unstable spiral } phase and $ss$: { stable spiral } phase, while the lines show the corresponding phase boundaries. The points denote non-linear dynamics estimate of the phases obtained from numerical integration of Eq.(\ref{eq:full_sys}): stable {\color{blue} $\diamond$}, unstable {\color{blue} $\blacktriangledown$}, stable spiral \tikzcircle[black, fill=black]{2pt}. Note that the linear instability due to Eq.(\ref{eq:two_by_two}) destabilizes all other possible phase behavior within the dashed black line. }
\caption{ (color online) Phase diagrams in $\tilde v_0-\tilde{f}_d$ plane with one movable MT ($a$), and both MTs movable ($b$). The parameter values used $\tilde k_m = 450$, $\tilde\omega = 20$, $\tilde{f}_s = 60$, $N = 50$ and $n_p = 20$. The region $s$ indicates { linearly stable} phase, region $u$ indicates { linearly unstable} phase, region $us$ indicates { unstable spiral } phase and region $ss$ indicates { stable spiral } phase corresponding to linear stability analysis. The points denote corresponding non-linear dynamics estimate obtained from numerical integration: stable {\color{blue} $\diamond$}, unstable {\color{blue} $\blacktriangledown$}, stable spiral \tikzcircle[black, fill=black]{2pt} and stable limit cycle {\color{red} $\blacktriangle$}. Note the stark contrast between the two phase diagrams, while diagram ($a$) does not have any unstable {\color{blue} $\blacktriangledown$} region, diagram ($b$) is devoid of stable limit cycle {\color{red} $\blacktriangle$}. }
\caption{Additional \aastex\symbols}
\caption{Mean classification accuracy (\%) on the CIFAR-10/100 datasets using different constraints. {\bf\color{red}{Red color}} and \underline{\color{blue}{blue color}} indicate the best and second best performing algorithms, respectively.}
\caption{Error analysis of zero-shot recognition by our method on the aP\&Y dataset. (a) Confusion matrix computed based on the per-class prediction results using our full model. Most errors come from semantically similar classes. (b) The label embeddings of both seen classes from Pascal (red) and unseen classes from Yahoo (green). The object classes with similar semantics are close in the embedding space.}
\caption{(a) An optical image of the L cross section of a \emph{Bombyx mori} silk fiber with laser marked $20\times 20~\mu$m$^2$ regions and laser irradiated spots. The region mapped in (b,c) is shown in a solid rectangle in (a). (b,c) Orientation vector map (marker's length Eqn.~\ref{e1} and orientation Eqn.~\ref{e2}) overlayed with the far-field absorbance (color map) at the C-N (b) and N-H (c) bands; these bonds are known to be perpendicular. \blue{S-polarised incident light was perpendicular to the fiber; in the plane of image.}}
\caption{High resolution $1.9~\mu$m ATR FT-IR maps at $1.9~\mu$m resolution of the longitudinal (L) cross sections of silk presented in auto-scale for better viewing; a background-corrected absorbance is ranging from 0 to approximately 0.2. Lateral step size between pixels was 0.5~$\mu$m; as-measured pixelated absorbance maps are presented. \blue{Polarisation of incident light onto ATR prism was $s$ (in the plane of image; along y-axis).}}
\caption{High resolution $1.9~\mu$m ATR maps of transverse (T) cross sections of silk (\emph{Bombyx mori}) with laser 515~nm/230~fs irradiated spots; laser pulse energies, $E_p = 85$~nJ (a) 170~nJ (b) on the sample; the polarisation was linear. The lateral step size was 0.5~$\mu$m; as-measured pixelated absorbance maps are presented. \blue{Polarisation of incident light onto ATR prism was $s$ (in the plane of image; along y-axis).}}
\caption{The orientation of the C=O, C-N, and N-H bonds in amide structure of the L-section of silk fiber~\cite{Alberts,Cruz} \blue{confirmed in this study by the hyper-spectral imaging} (see, Fig.~\ref{f-map}). Only the in-plane components of $\beta$-sheets are drawn without out-off-plane alkyl moieties; hydrogen bonding responsible for $\beta$-sheet crystallisation is shown by the dotted line O$\cdot\cdot\cdot$H. An arrow marks fiber drawing (strain) direction important to alignment of $\beta$-sheets. \blue{Microtome slices allowed to measure absorbance of the lateral flat cross sections without introduction of a fiber shape related anisotropy. }}
\caption{The $P_2(\theta)$ order parameter distribution along silk fiber for different transition dipole moments of Amide A and Amide II bands\blue{, which are perpendicular (see, the difference in color map).} }
\caption{Majority consensus taxonomic tree of objects in the Jovian system. This tree has a tree length score of 128, with a consistency index of 0.46 and a retention index of 0.85. Numbers indicate frequency of the node in the 10000 most parsimonious tree block. \replaced{Colors are indicative of}{Colors represent terminology used in} traditional classification: \textcolor{Amalthea}{Amalthea inner regular family} ; \textcolor{Galilians}{Galilean family}; \textcolor{Themisto}{Themisto prograde irregular} ; \textcolor{Himalia}{Himalia prograde irregular family}; \textcolor{Carpo}{Carpo prograde irregular} ; \textcolor{Ananke}{Ananke irregular family}; \textcolor{Carme}{Carme irregular family} ; \textcolor{Pasiphae}{Pasiphae irregular group}; Unnamed and unclassified. Proposed groups and families are \replaced{indicated}{shown} on the right.}
\caption{Majority Consensus taxonomic tree of objects in the Saturnian system. The tree has a consistency index of 0.30 and a retention index of 0.81. Numbers indicate frequency of the node in the 10000 most parsimonious tree block. \replaced{Colors are indicative of}{Colors represent terminology used in} classical classification: \textcolor{MainRing}{Main ring group, with associated shepherd satellites} ; \textcolor{IcySats}{Mid-sized Icy satellites and Titan}; \textcolor{Trojans}{Trojan satellites} ; \textcolor{Alkanoids}{Alkanoids and associated rings}; \textcolor{Inuit}{`Inuit' prograde irregular family} ; \textcolor{Gallic}{`Gallic' prograde irregular family}; \textcolor{Norse}{`Norse' retrograde irregular family} ; Unnamed and unclassified. Proposed groups and families are \replaced{indicated}{shown} to the right. }
\caption{The evolution of the SED over all 6 visits reconstructed from the HST photometry. \textcolor{red}{[ST: Figure still missing.]}}
\caption{Decay radiation of four isotopes producing the highest energies when considering electrons/positrons and x-rays up to 20\,keV. The solid lines are the N100 model. The dashed lines are the merger model.}
\caption{The \gray spectrum of PKS 1441+25 above 100 MeV averaged over the {\it Fermi} LAT observations in January (blue) and April (red).}
\caption{ The \gray light curve of PKS 1441+25 from January to December 2015 (a). The bin intervals correspond to 1- day (blue data) and 3-days (green data). The light curve obtained by adaptive binning method assuming 20 \% of uncertainty is presented in red (b). The change of photon index for 3-day binning (green) and with adaptive binning method are shown in (c).}
\caption{The broadband SED of PKS 1441+25 for January (red), April (blue) and for the quiescent state (gray). The blue, red and gray lines are the models fitting the data with the electron spectrum given by Eq. \ref{BPL} for January, April and for the quiescent state, respectively. The model parameters are presented in Table \ref{table_fit}. The UV-X-ray and VHE \gray data observed in January and April are from \citet{abeysekara} and HE \gray data ({\it Fermi} LAT) are from this work.}
\caption{{\label{fig:omTphysCcoeff}} Left: physical predictions for thrust ($\tau=1-T$) at LO, NLO and NNLO accuracy in QCD with bands representing scale uncertainty. Data measured by the ALEPH collaboration \cite{Heister:2003aj} is also shown. Right: the $\tau\, C(\tau)$ NNLO coefficient of the thrust distribution. On both figures the lower panels show the ratio of the predictions of Ref.~\cite{Weinzierl:2009ms} (SW) and \eerad\(GGGH) to \colorfulmethod. In the middle panel of the right figure, results from \amcatnlo\\cite{Alwall:2014hca} (MG5) are also shown above the Born kinematic limit of $\tau > 1/3$. }
\caption{Illustration of the non-recurrent LN architecture with one hidden and one output layer. The encoder and decoder paths are highlighted in \textcolor{enc}{\textbf{green}} and \textcolor{dec}{\textbf{yellow}}, respectively.}
\caption{ Overview of (a) feed-forward noise (FFN) and (b) recurrent noise (RN) injection schemes for the encoders (\textcolor{enc}{\textbf{green}}) introduced in subsection \ref{sec:rln_noise} as well as (c) recurrent decoder (RD) and (d) feed-forward decoder (FFD) layouts (\textcolor{dec}{\textbf{yellow}}) introduced in subsection \ref{sec:rln_decoder}. Combining all encoder and decoder layouts gives a total of six model variants including the two no-decoder (ND) baselines ND-RN and ND-FFN.}
\caption{Magnetic-field noise spectrum for different fields as a function of frequency. The test field is applied for different frequencies of 22\,Hz\color{amber} (amber)\color{black}, 72\,Hz\color{blue} (blue)\color{black}, 92\,Hz\color{green} (green) \color{black} and 132\,Hz\color{red} (red)\color{black}.}
\caption{Measured and calculated RMS noise of cavity transmission. Data show in dots \color{blue} (blue)\color{black}, are fitted with a line \color{red} (red)\,\color{black}. Theoretical calculation for the shot noise is presented with dashed lines and an asterix symbol (*) is placed to signify the noise level under normal operation.}
\caption{Example synthetic documents, raw segmentations and results after optional post-processing (Sec.~\ref{section:implement}). Segmentation label colors are: \colorbox{green}{\strut \bf figure}, \colorbox{blue}{\strut \textcolor{white}{\bf table}}, \colorbox{red}{\strut \textcolor{white}{\bf section heading}}, \colorbox{mymagenta}{\strut \textcolor{white}{\bf caption}}, \colorbox{mycyan}{\strut \bf list} and \colorbox{yellow}{\strut \bf paragraph}.}
\caption{Example real documents and their corresponding segmentation. Top: DSSE-200. Middle: ICDAR2015. Bottom: SectLabel. Since these documents are not in PDF format, the simple post-processing in Sec.~\ref{section:implement} can not be applied. One may consider exploiting a CRF~\cite{chen2014semantic} to refine the segmentation, but that is beyond the main focus of this paper. Segmentation label colors are: \colorbox{green}{\strut \bf figure}, \colorbox{blue}{\strut \textcolor{white}{\bf table}}, \colorbox{red}{\strut \textcolor{white}{\bf section heading}}, \colorbox{mymagenta}{\strut \textcolor{white}{\bf caption}}, \colorbox{mycyan}{\strut \bf list} and \colorbox{yellow}{\strut \bf paragraph}. }
\caption{Synthetic documents and the corresponding segmentations. (1) Input synthetic documents. (2) Candidate bounding boxes obtained by parsing the PDF rendering commands. (3) Raw segmentation outputs. (4) Segmentations after post-processing. Segmentation label colors are: \colorbox{green}{\strut \bf figure}, \colorbox{blue}{\strut \textcolor{white}{\bf table}}, \colorbox{red}{\strut \textcolor{white}{\bf section heading}}, \colorbox{mymagenta}{\strut \textcolor{white}{\bf caption}}, \colorbox{mycyan}{\strut \bf list} and \colorbox{yellow}{\strut \bf paragraph}.}
\caption{Synthetic documents and the corresponding segmentations. (1) Input synthetic documents. (2) Candidate bounding boxes obtained by parsing the PDF rendering commands. (3) Raw segmentation outputs. (4) Segmentations after post-processing. Segmentation label colors are: \colorbox{green}{\strut \bf figure}, \colorbox{blue}{\strut \textcolor{white}{\bf table}}, \colorbox{red}{\strut \textcolor{white}{\bf section heading}}, \colorbox{mymagenta}{\strut \textcolor{white}{\bf caption}}, \colorbox{mycyan}{\strut \bf list} and \colorbox{yellow}{\strut \bf paragraph}.}
\caption{Real documents and the corresponding segmentations. Segmentation label colors are: \colorbox{green}{\strut \bf figure}, \colorbox{blue}{\strut \textcolor{white}{\bf table}}, \colorbox{red}{\strut \textcolor{white}{\bf section heading}}, \colorbox{mymagenta}{\strut \textcolor{white}{\bf caption}}, \colorbox{mycyan}{\strut \bf list} and \colorbox{yellow}{\strut \bf paragraph}.}
\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}{$\spadesuit$} = Benign showing collusion potential.}
\caption{Total time for 20 timesteps on BG/Q Mira: AMG ({\color{blue}blue}), XXT ({\color{red}red}), communication (dashed), computation (solid), total time ($\square$), computational linear scaling ({\color{green}green}).}
\caption{Total mean time for 20 timesteps on Titan: AMG ({\color{blue}blue}), XXT ({\color{red}red}), communication (dashed), computation (solid), total time ($\square$), computational linear scaling ({\color{green}green}).}
\caption{Total mean time for 20 timesteps on Beskow: AMG ({\color{blue}blue}), XXT ({\color{red}red}), communication (dashed), computation (solid), total time ($\square$), computational linear scaling ({\color{green}green}).}
\caption{\label{fig:p-finfets}\textcolor{red}{{} }(a) Layout of 6 bulk p-FinFETs showing the strike location and charge collection map for a 11MeV particle strike. (b) Graph showing a comparison between the number of devices collecting more than a certain charge Q (x-axis) in n-FinFETs and p-FinFETs for 11MeV and 5MeV particle strikes.}
\caption{Convergence example of the SGPA RA algorithm w.r.t.\textcolor{red}{{} }$\{\beta_{k,m}\}_{m\in\mathcal{M}}$ when $\{\alpha_{k,m,n}\}$ and $\{\gamma_{m}\}$ are given ($M=20,M_{k}=3$).\label{fig:iterations}}
\caption{Examples of time-series data from Yahoo! and NAB corpora. The {\color{red} red} shaded regions represent the anomaly windows centered at anomalies. The {\color{blue} blue} region marks the data which the benchmark offers for initial parameter estimation and hyperparameter tuning.}
\caption{Score weighting in NAB: \textbf{all} detections outside the window are {\color{black!10!red} \textbf{false alarms}}, whereas only \textbf{the earliest} detection within the window is a {\color{black!60!green} \textbf{true positive}}, and later detections are ignored. }
\caption{An experimental and numerical comparison of $\sigma_{z}$ robustness for three M{\o}lmer-S{\o}rensen entangling gate protocols: the fast DD scheme, as proposed in this article ({\color[rgb]{0 0.4470 0.7410} blue}); the slow DD scheme ({\color[rgb]{0 .75 0} green}); and the slow DD scheme without DD pulses ({\color{red} red}). The number of pulses is denoted as 'm'. The comparison is shown for differing pulse sequences enumerated by the number of DD arms. Two $^{88}Sr+$ were entangled according to the appropriate protocol using a 674 nm laser and the their final state measured via state-selective fluorescence. $\sigma_z$ noise was implemented by detuning all driving lasers from their resonance frequencies. The fidelity with the specific fully entangled state, calculated from measurement results, is shown. $95\%$ confidence intervals are under $\pm 0.03$ and are not plotted. The numerics were done for the MS Hamiltonian with the appropriate pulse sequence and a detuning term, alternating the dynamics between the MS Hamiltonian (Eq.~\ref{MS}) with detuning $\frac{\Delta}{2}\sum_i \sigma_{\beta,i}$, and the detuned $\pi$ pulse. The figures A-E shows the fidelity for a specific state, figure F shows the fidelity for an entangled state up to an arbitrary phase. F shows that high quality entangled states are achieved even at large detunings, albeit at a phase that differs from the zero detuning case. }
\caption{These figures show a comparison of three schemes, the Flower scheme ({\color[rgb]{0 0.4470 0.7410} blue}), the Echo scheme ({\color[rgb]{0 .75 0} green}), and without pulses ({\color{red} red}) (same detuning as in the Echo scheme). The numerics was done for the MS Hamiltonian with the appropriate pulse sequence and a detuning term, alternating the dynamics between the MS Hamiltonian (Eq.~\ref{MS_I}) with detuning $\frac{\Delta}{2}\sum_i \sigma_{\beta,i}$, and the detuned $\pi$ pulse. }
\caption{(a) Comparison of the evolution as a function of time of the sensor measurement $a_1$ obtained from direct numerical simulation ({\color{blue} \bf -----} $a_1^{\bullet}$) and predicted by the identified low-dimensional model \eqref{eq: lift-based model} ({\color{orange} \bf -- --} $a_1^{\circ}$). (b) Trajectory of the true and identified systems in the phase plane ($a_1$, $a_2$). In both cases, the initial condition is close to the linearly unstable fixed point $\mathbf{u}_b(\mathbf{x})$, given by $\mathbf{a}=\mathbf{0}$.}
\caption{(a) Comparison of the evolution as a function of time of the sensor measurement $a_1$ obtained from direct numerical simulation ({\color{blue} \bf -----} $a_1^{\bullet}$) and predicted by the identified low-dimensional model \eqref{eq: lift-based model} ({\color{orange} \bf -- --} $a_1^{\circ}$). (b) Trajectory of the true and identified systems in the phase plane ($a_1$, $a_2$). The initial condition used in the direct numerical simulation has been chosen to lie outside of the limit cycle, and for physical reasons it is also constrained to start close to the paraboloid manifold structuring the phase space of the system.}
\caption{(a) Comparison of the evolution as a function of time of the sensor measurement $a_3$ obtained from direct numerical simulation ({\color{blue} \bf -----} $a_3^{\bullet}$) and predicted by the identified low-dimensional descriptor system made of \eqref{eq: lift-based model} and \eqref{eq: measurement equation} ({\color{orange} \bf -- --} $a_3^{\circ}$). (b) Trajectory of the true system and the identified one in the phase plane ($a_1$, $a_3$).}
\caption{(a) Comparison of the evolution as a function of time of the sensor measurement $a_3$ obtained from direct numerical simulation ({\color{blue} \bf -----} $a_3^{\bullet}$) and predicted by the identified low-dimensional descriptor system made of \eqref{eq: lift-based model} and \eqref{eq: measurement equation} ({\color{orange} \bf -- --} $a_3^{\circ}$). (b) Trajectory of the true system and the identified one in the phase plane ($a_1$, $a_3$).}
\caption{In both figures, the blue line depicts the trajectory of the testing dataset for which we reconstruct the flow field. (a) Delaunay triangulation of the state plane. In addition to the two transient trajectories started from the fixed point, a third trajectory with an initial condition above the limit cycle has been used to obtain this triangulation. (b) Close-up view in the vicinity of the query point ({\color{red} $\bullet$}). The corresponding flow field can then be estimated as a weighted average of the flow fields associated to each vertex of the triangle highlighted in green.}
\caption{Relative AIC\textsubscript{c} criteria for models found by SINDy. The library of polynomial functions used for the identification includes up to 7\textsuperscript{th} degree polynomials in $a_1$ and $a_2$. Magnification in lower panel shows strong (dark gray) and weakly (light gray) supported AIC\textsubscript{c} range. The constrained cubic model ({\color{orange} $\blacksquare$}) identified in \textsection \ref{subsec: dynamical system} lies in the strong support range while its unconstrained counterpart lies just above the weakly supported range.}
\caption{(a) Comparison of the evolution as a function of time of the sensor measurement $a_1$ obtained from direct numerical simulation ({\color{blue} \bf -----} $a_1^{\bullet}$) and predicted by the discrete-time sparse model ({\color{orange} \bf -- --} $a_1^{\circ}$). (b) Trajectory of the true and identified systems in the phase plane ($a_1$, $a_2$). In both cases, the initial condition is close to the linearly unstable fixed point.}
\caption{Observations (grey points) as in Figure \ref{fig:GramicidinData} together with idealization by $\JULES$ ({\ttlred \solidrule}) and its convolution with the lowpass filter ({\myred \solidrule}).}
\caption{Observations (grey points) as in Figure \ref{fig:GramicidinData} together with idealization by $\JSMURF$ ({\green \solidrule}) and its convolution with the lowpass filter ({\mygreen \solidrule}).}
\caption{Observations (grey points) as in Figure \ref{fig:GramicidinData} together with idealization by $\TRANSIT$ ({\ttlblue \solidrule}) and its convolution with the lowpass filter ({\blue \solidrule}).}
\caption{\footnotesize Simulated observations (grey points), true block signal $f$ ({\black \solidrule}) and its convolution ({\black \solidrule}), $\JULES$ ({\ttlred \solidrule}), $\TRANSIT$ ({\ttlblue \solidrule}) and $\JSMURF$ ({\green \solidrule}) idealizations. Convolution of $\JULES$ ({\myred \solidrule}) and $\TRANSIT$ ({\blue \solidrule}) idealizations with the lowpass 4-pole Bessel filter. $\JULES$ provides very accurate idealization, whereas $\TRANSIT$ shifts to the right and estimates a too small amplitude for smaller lengths and $\JSMURF$ misses such peaks.}
\caption{Histograms of the dwell time in the closed state for $\Delta=2$ for all closing events with amplitude between $10$ and $\SI{30}{\pico\siemens}$ together with the true exponential distribution with parameter $0.8$ ({\black \solidrule}) and exponential fits ({\ttlred \solidrule}). We rescaled all lines such that the area under them are standardized to one to make them comparable to the histograms.}
\caption{Histograms of the distance between two flickering events, i.e., events with amplitude between $10$ and $\SI{30}{\pico\siemens}$ and dwell time below $\SI{2.6}{\milli\second}$, for $\Delta=2$ together with the true exponential distribution with parameter $2.5$ ({\black \solidrule}), exponential fits ({\ttlred \solidrule}) and the exponential fits corrected for missed events ({\blue \solidrule}). We rescaled all lines such that the area under them are standardized to one to make them comparable to the histograms.}
\caption{Kernel density estimates with bandwidth $2$ of the amplitudes of events with dwell time below $\SI{2.6}{\milli\second}$ ({\ttlred \solidrule}) and above $\SI{10}{\milli\second}$ ({\blue \solidrule}).}
\caption{Histograms of the dwell time in the closed state for all closing events with amplitude between $10$ and $\SI{30}{\pico\siemens}$ together with exponential fits ({\ttlred \solidrule}) which are rescaled such that the area under them are standardized to one to make them comparable to the histograms.}
\caption{Histograms of the distance between two flickering events, i.e., events with amplitude between $10$ and $\SI{30}{\pico\siemens}$ and dwell time below $\SI{2.6}{\milli\second}$ together with exponential fits ({\ttlred \solidrule}) and the fits corrected for missed events ({\blue \solidrule}). All lines are rescaled such that the area under them are standardized to one to make them comparable to the histograms.}
\caption{Classification accuracy after denoising noisy image input, averaged over ILSVRC2012 validation dataset. \textcolor{red}{Red} is the best and \textcolor{blue}{blue} is the second best results. }
\caption{Segmentation results (mIoU) after denoising noisy image input, averaged over Pascal VOC 2012 validation dataset. \textcolor{red}{Red} is the best and \textcolor{blue}{blue} is the second best results.}
\caption{Magnetic center evolution measurement in Fig. \ref{Fig6}-a case: ($\bullet$) pulsed wire, ($\circ$) rotating coil, (+) stretched wire measurements, and ({\color{green}$\star$}) for \ref{Fig6}-b case stretched wire measurements. QUAPEVA of a) 26 mm, (b) 40.7 mm, (c) 44.7 mm length.}
\caption{(a) The out-of-phase component component as a function of normalized temperature for all MIXx compacts. The temperature has been normalized to $T^*$, defined as the temperature corresponding to the maximum slope of the onset of dissipation. $\chi^{\prime\prime}$ has been normalized to the maximum close to the onset of dissipation. Fits to a power law, $\tau=\tau_0(T_f/T_g-1)^{-z\nu}$, for (b) MIX0 and (c) MIX50. The inset in (b) shows the out-of-phase component as a function of temperature for several frequencies (0.17 to 510 Hz) for MIX0. The \textcolor{red}{ $\ast$} indicates the determined freezing temperature $T_f$. The inset in (c) shows the result of a memory experiment for MIX50.}
\caption{ % Differentiability of synaptic current dynamics: The synaptic current traces from eq~\eqref{eq:syn_dyn_new} (solid lines, upper panels) are shown with the corresponding membrane voltage traces (lower panels). % Here, the gate function is $g = 1/\Delta $ within the active zone %the gate function's active zone of width $\Delta$ (shaded area, lower panels){\red ;} $g = 0$ otherwise. (A,B) The pre-synaptic membrane voltage depolarizes beyond the active zone. % above Despite the different rates of depolarization, both events incur the same amount of charge in the synaptic activity: $\int s ~ dt = 1$. (C,D,E) Graded synaptic activity due to insufficient depolarization levels that do not exceed the active zone. The threshold-triggered synaptic dynamics in eq~\eqref{eq:syn_dyn_old} is also shown for comparison (red dashed lines, upper panels). % The effect of voltage reset is ignored for the purpose of illustration. $\tau = 10$ ms. }
\caption{ Delayed-memory XOR task: Each panel shows the single-trial input, go-cue, output traces, and spike raster of an optimized QIF neural network. The y-axis of the raster plot is the neuron ID. {\red Note %that the initial portion of spike activity exhibits similar patterns according to the first input pulse the similarity of the initial portion of spike patterns for trials of the same first input pulses (A,B,C vs D,E,F). %whereas the spike pattern after the go-cue signal is determined by the desired output pulse In contrast, the spike patterns after the go-cue signal are similar for trials of the same desired output pulses: (A,D: negative output), (B,E: positive output), and (C,F: null output).} }
\caption{\textcolor[rgb]{0,0,0}{Generalized illustration of a non-collaborative setting with profit ($P_i$) maximizing participants $i$, $i\in (A,...,N)$, with limited capacity $L_i$. $Cap_i$ is the capacity usage.}}
\caption{\textcolor[rgb]{0,0,0}{Generalized illustration of centralized planning with profit ($P_i$) maximizing participants $i$, $i\in (A,...,N)$, with limited capacity $L_i$. $Cap_i$ is the capacity usage.}}
\caption{\textcolor[rgb]{0,0,0}{Generalized illustration of decentralized planning with profit ($P_i$) maximizing participants $i$, $i\in (A,...,N)$, with limited capacity $L_i$. $Cap_i$ is the capacity usage. $\bar{R}$ is the subset of requests that have been offered for exchange.}}
\caption{CIFAR10: change of loss function in the direction of white-box and black-box FGSM and PGD examples with $\epsilon=8$ for the same five natural examples. Each line shows how the loss changes as we move from the natural example to the corresponding adversarial example. Top: simple naturally trained model. Bottom: wide PGD trained model. We plot the loss of the original network in the direction of the FGSM example for the original network ({\color{red} red} lines), 5 PGD examples for the original network obtained from 5 random starting points ({\color{blue} blue} lines), the FGSM example for an independently trained copy network ({\color{green} green} lines) and 5 PGD examples for the copy network obtained from 5 random starting points ({\color{black} black} lines). All PGD attacks use 100 steps with step size $0.3$.}
\caption{Original diamond chain and the DM state. An open circle represents a half spin and a bold or thin line represents an exchange interaction. \blue{The DM state consists of singlet dimers (shaded ovals) and free spins (shaded circles).} }
\caption{All the possible \blue{groupings} of spins in the DM state. The grouping rule is that the spins in a group form a partial eigenstate with \blue{a} total spin magnitude \blue{of} 0 or 1/2 for the DM state, and any two of \blue{the} spins in a group are connected by an exchange interaction in Fig. \ref{unit_extend}.}
\caption{Examples of learned attention. The entities in {\color{blue} \bf{bold blue}} draw more attention; those in {\color{mygray} gray} draw less attention. \label{tab:eg_att} }
\caption{Order issue in natural language generation, in which an incorrect generated sentence has \textcolor{red}{\underline{wrong ordered slots}}.}
\caption{Comparison of top responses generated for some input dialogue acts between different models. Errors are marked in color (\textcolor{blue}{missing}, \textcolor{red}{misplaced} slot-value pair). \textbf{$^{\dag}$} and \textbf{$^{\natural}$} denotes the baselines and the proposed models, respectively.}
\caption{ \label{fig:imposed_flow_expt} Effect of imposed flow on passive and active colloidal particles. (a) Passive silica-Pt particles in a flow \textcolor{blue}{($R = 1\;{\mu m}$)}. The plot shows the dependence of the rotation time on flow velocity. $V^*$ is the translational velocity of the passive particle and is used to characterize the flow rate. \textcolor{black}{The red dashed line is a theoretical scaling derived in Sect. S4 and fitted to the data.} Inset: Time-lapse images of a passive particle rolling in flow. (b) Tracked trajectories of passive particles in flow ($V^* = 14 \, \mu m/s$). (c) Optical microscopy image capturing the distribution of orientations of active particles \textcolor{black}{($V_{p} \approx 6\;{\mu m}/s$)} without an imposed flow. (d) Tracked trajectories of active particles without flow. (e) The same system as (c) with an external flow imposed, $V^* = 14 \, \mu m/s$. (f) Tracked trajectories of active particles in flow ($V^* = 14 \, \mu m/s$). (g) The angular probability distributions of active colloidal particles in the absence of an imposed flow, $V^* = 0 \, \mu m/s$ (green), at $V^* = 14 \, \mu m/s$ (blue) and at $V^* = 24 \, \mu m/s$ (red). Inset shows the angular evolution of two different active colloids at imposed flow rates of $V^* = 14 \, \mu m/s$ (blue) and $V^* = 24 \, \mu m/s$ (red).}
\caption{Results for the colour-experiment (n=3). \textbf{(a)} \emph{Accuracy}. DNNs are shown in shades of blue, human data in red; diamonds correspond to \textcolor{alexnet.100}{AlexNet}, squares to \textcolor{googlenet.100}{GoogLeNet}, triangles to \textcolor{vgg.100}{VGG-16}, and circles to \textcolor{human.100}{human observers}; error bars as described in section~\ref{results_general}. \textbf{(b)} \emph{Response distribution entropy}. Plotting conventions as in (a).}
\caption{Estimated stimuli corresponding to $ 50\% $ classification accuracy. \textbf{(a)} \emph{Noise-experiment}. \textbf{(b)} \emph{Eidolon-experiment}. Coherence parameter = 1.0. Top row: stimuli corresponding to threshold for the \textcolor{human.100}{average human observer}. Bottom three rows: stimuli corresponding to the same accuracy for \textcolor{vgg.100}{VGG-16} (second row), \textcolor{googlenet.100}{GoogLeNet} (third row) and \textcolor{alexnet.100}{AlexNet} (bottom row).}
\caption{Estimated stimuli corresponding to $ 50\% $ accuracy for contrast-experiments and eidolon-experiment (coherence parameter = 0.3 and 0.0). Top row: stimuli corresponding to threshold for the \textcolor{human.100}{average human observer}. Bottom three rows: stimuli corresponding to the same accuracy for \textcolor{vgg.100}{VGG-16} (second row), \textcolor{googlenet.100}{GoogLeNet} (third row) and \textcolor{alexnet.100}{AlexNet} (bottom row). The corresponding stimulus levels were calculated by assuming a linear relationship between the two closest data points measured in the experiments.}
\caption{\textbf{Average 3D test error} of the multi-category experiment, where the numbers are shown as [\;prediction$\to$GT / GT$\to$prediction\;]. The mean is computed across categories. For the single-view case, we outperform all baselines in 8 and 10 out of 13 categories for the two 3D error metrics. (All numbers are scaled by 0.01)}
\caption{A screenshot of \grayii\in its interactive mode, which allows a user to integrate and visualize the photon positions in real time. For this particular calculation, we set up a grid of photons originating at a large distance from a black hole with a spin of 0.999. This grid is deformed as it passes near the black hole event horizon. Some of the photons, shown in pink here, are trapped near the horizon. Some others, shown in green and blue, are deflected at large angles.}
\caption{Results of the convergence study using the integral of motion $u^2$. Here, we plot $\max(u^2)$ for $0 \le \lambda < 64$ for the six test problems we performed, as a function of the step $\Delta\lambda$ in the affine parameter. \grayii\converges at 4th order it used the 4th-order Runge-Kutta scheme. The amplitude of the error decreases as$r$ increases except for Test~E. This is an artifact of the oscillatory behavior of $u^2$, in this test, shown in the inset.}
\caption{(\emph{Left}) Unfolded photon orbits in a cylindrical coordinate projection. The vertical axis is the Cartesian $z$ in the Kerr-Schild coordinates. The horizontal axis is the unfolded azimuthal angle $\phi$ in the Kerr-Schild coordinates. We only plot two representative curves here for Tests~B and D. (\emph{Right}) Zoom-in view of the first peak of Model~B as a function of the affine parameter $\lambda$ with different step sizes. The solid circles are the actual output of the \grayii. For each step size, we fit a quadratic equation to the largest three circles. The quadratic equations are then plotted as the different curves here. The color scheme of the solid circles and the curves matches those of Figure~\ref{fig:evolr}.}
\caption{Vertical epicyclic frequency as a function of orbital radius. We use \grayii\to integrate a nearly circular time-like geodesics around a black hole of spin$a = 1$. We introduce a small initial vertical velocity $v_z = 10^{-12}$ to induce vertical oscillation and precession. We then compare the numerical vertical epicyclic frequencies (orange circles) with its analytical expression (blue solid line). The fractional difference between the two curves at a radius $r<1.5$ is less than $10^{-13}$.}
\caption{Results of MonteRay performance ($\blacktriangle, \blacktriangledown, \blacksquare, \bullet$ ) compared with other researchers implementations of neutron transport on GPUs. The performance of the other researchers' work, except Profugus, has been projected with dotted lines (\protect\dashedrule) for multiple CPU cores for using Eq.~\eqref{eq:gpu_ncpus}\@ and a strong scaling efficiency of 78\%. The Profugus result have been projected with a solid line (\solidrule) and as it was the most full featured code that had direct CPU performance comparisons.}
\caption{Relative probability of friendship for groups of users separated by race and gender. The value declares the relative increase (\textcolor{blue}{blue} color) or decrease (\textcolor{red}{red} color) in comparison to the expected demographic population.}
\caption{Relative probability of interactions for groups of users separated by race and gender. The value declares the relative increase (\textcolor{blue}{blue} color) or decrease (\textcolor{red}{red} color) in comparison to the expected demographic population.}
\caption{\textbf{Chosen degree of memory $h$} in simulations for varying true degrees of memory $h_{\textrm{true}}$ and number of observed trajectories $J$. Rows correspond to model selection under a given degree of memory. Columns correspond to the number of trajectories. Depicted are the percent of simulations in which each degree of memory is selected using the different model evaluation criteria (percents of at least $20$ are labeled). Colors coded based on degree of memory: ({\color{Set1_5_red} 1: red}, {\color{Set1_5_blue} 2: blue}, {\color{Set1_5_green}3: green}, {\color{Set1_5_purple}4: purple}, {\color{Set1_5_orange}5: orange}). \emph{Example:} For $h_{\textrm{true}}=1$ and $J=4$, the WAIC$_1$ criteria selected $h=1$ approximately $65\%$ of the time. }
\caption{\small The time evolution of the inflaton and the Higgs for $v_\phi = 10^{-2}\Mpl$ up to $m_\phi t = 150$. The black line is the inflaton condensation $\langle \varphi \rangle^2$, \textcolor{light-red}{the red line} is the inflaton two point function $\langle \varphi^2 \rangle - \langle \varphi \rangle^2$ and \textcolor{dark-blue}{the blue line} is the Higgs two point function $\langle h^2 \rangle$, where the angle brackets denote the spatial average. They are normalized by the initial inflaton amplitude $\varphi_\text{ini}$. The EW vacuum is stable for $(p, q) = (0.5, 0.5)$, while it is destabilized during the preheating for the other cases. The lower right panel corresponds to the case with an accidental cancellation between $\sigma_{\phi h}$ and $\lambda_{\phi h}$. }
\caption{\small The time evolution of the inflaton and Higgs for $v_\phi = 10^{-3}\Mpl$ up to $m_\phi t = 250$. The black line is the inflaton condensation $\langle \varphi \rangle^2$, \textcolor{light-red}{the red line} is the inflaton two point function $\langle \varphi^2 \rangle - \langle \varphi \rangle^2$ and \textcolor{dark-blue}{the blue line} is the Higgs two point function $\langle h^2 \rangle$, where the angle brackets denote the spatial average. They are normalized by the initial inflaton amplitude $\varphi_\text{ini}$. The EW vacuum is stable for $(p, q) = (0.5, 0.5)$, while it is destabilized during the preheating for the other cases. The lower right panel corresponds to the case with an accidental cancellation between $\sigma_{\phi h}$ and $\lambda_{\phi h}$. }
\caption{For the multiple-seed case, (a) plot of the average cascade time step $\langle n_c \rangle$ ($\bigcirc$) and $3.15\big\langle {{d_0}^{-1/2}}\rangle_+$ (\textcolor{red}{$\blacktriangle$}) versus system size $N$ at $\kappa_c \approx0.11494875096512$. The notation $_+$ in $\big\langle {d_0}^{-1/2}\rangle_+$ indicates that only positive values of $d_0$ are considered in taking the average. $d_0$ is measured at $n^*=80$. The guideline has a slope of $0.252$. (b) Plot of $\langle n_c \rangle / N^{0.25}$ (\textcolor{orange}{$\blacksquare$}), $1.023\langle n_c \rangle / N^{0.252}$ ($\bigcirc$), and $3.3\dsqrt /N^{0.254}$ (\textcolor{blue}{$\blacklozenge$}) versus $N$. Data were obtained from ER networks of different sizes but with the same mean degree, $z=8$. $i_0=0.002$, $\mu=\kappa$, and $\eta=0.5$ were used. Average is taken over more than $10^5$ realizations for each data point for $N < 10^8$.}
\caption{For $k$-core percolation with $k=3$, (a) plot of the golden time $\langle n_c \rangle$ ($\bigcirc$) and $4.05\big\langle {d_0(n^*)}^{-1/2}\big\rangle_+$ (\textcolor{red}{$\blacktriangle$}), and $2.4\big\langle {d_1(0)}^{-1/2}\big\rangle_+$ (\textcolor{blue}{$\blacktriangledown$}) versus $N$ for $k$-core percolation with $k=3$ starting from multiples nodes of $O(N)$. $d_0$ is measured at $n^*=60$ and $d_0(0)$ denotes the value of $d_0$ measured at $n=0$. A solid (dashed) guideline has slope of 0.2655 (0.25). (b) Plot of $1.23\langle n_c\rangle/N^{0.2655}$ (\textcolor{orange}{$\blacksquare$}), $\langle n_c\rangle/N^{0.25}$ ({$\bigcirc$}), $4\big\langle {d_0(n^*)}^{-1/2}\big\rangle_+/N^{0.25}$ (\textcolor{blue}{$\blacklozenge$}), and $4.65\big\langle {d_0(n^*)}^{-1/2}\big\rangle_+/N^{0.262}$ (\textcolor{red}{$\blacktriangle$}) versus $N$. Data were obtained from ER networks of different sizes $N$ but with the same mean degree, $z=3.723243$, and $i_0= 0.0567377$. Average is taken over more than $10^5$ realizations for each data point.}
\caption{(a) Plot of the average cascade time step $\langle n_c \rangle$ for the threshold model starting from initial multiple active nodes. Guideline has a slope of $0.263$. (b) Plot of $\langle n_c\rangle/N^{0.25}$ ($\blacksquare$) and $\langle n_c\rangle/N^{0.263}$ (\textcolor{orange}{$\bigcirc$}) versus $N$. Data were obtained from ER networks of different sizes $N$ but with the same $(z,\rho_0,\phi)=(9.191,0.01,0.18)$. Average is taken over more than $4\times 10^4$ realizations for each data point.}
\caption{The \kms{} algorithm. Additions to \kmu{} shown \textcolor{red}{red}}
\caption{Set~1: Mean velocity profile ${U}/U_0$ for different distance $x^*$: \textcolor{red}{$+$ 0.25}, \textcolor{green}{$\times$ 0.5}, \textcolor{blue}{$\ast$ 0.75}, \textcolor{magenta}{$\square$ 1}, \textcolor{blue}{$\blacksquare$ 1.25}, \textcolor{yellow}{$\opencircle$ 1.5}, $\fullcircle$ 1.75, \textcolor{orange}{$\opentriangle$ 2}, \textcolor{gray}{$\blacktriangle$ 2.25}, \textcolor{red}{$\opentriangledown$ 2.5}, \textcolor{green}{$\blacktriangledown$ 2.75} and \textcolor{blue}{$\diamond$ 3}.}
\caption{Set~1: normalized velocity profiles (${U^*}$, $y*$), for different distance $x^*$: \textcolor{red}{$+$ 0.25}, \textcolor{green}{$\times$ 0.5}, \textcolor{blue}{$\ast$ 0.75}, \textcolor{magenta}{$\square$ 1}, \textcolor{blue}{$\blacksquare$ 1.25}, \textcolor{yellow}{$\opencircle$ 1.5}, $\fullcircle$ 1.75, \textcolor{orange}{$\opentriangle$ 2}, \textcolor{gray}{$\blacktriangle$ 2.25}, \textcolor{red}{$\opentriangledown$ 2.5}, \textcolor{green}{$\blacktriangledown$ 2.75} and \textcolor{blue}{$\diamond$ 3}.}
\caption{KITTI image with the \textcolor{blue}{ground-truth} and \textcolor{red}{predicted} bounding boxes and an \textcolor{darkgreen}{attention glimpse}. The lower row corresponds to the hierarchical attention of our model: $1^{st}$ layer extracts an attention glimpse (a), the $2^{nd}$ layer uses appearance attention to build a location map (b). The $3^{rd}$ layer uses the location map to suppress distractors, visualised in (c).}
\caption{ \red Full line: \black the flux \red at the Earth \black of solar axions or ALPs emitted due to the Sikivie process~\cite{CAST2007} for $\g=4\times 10^{-11}$~GeV$^{-1}$ and $m_{a}\lesssim 0.01$~eV. The right-hand-side scale gives the number of converted photons in TASTE per second, assuming $B=3.5$~T, $L=12$~m and the tube diameter of 60~cm. Note that the flux scales as $\g^{2}$ but the number of converted photons scales as $\g^{4}$. \red The dashed line assumes SODART energy-dependent efficiency, see Sec.~\ref{sec:X:opt}, and gives the number of photons collected by the telescope.\black \label{fig:flux} }
\caption{\red The energy dependence of the X-ray optics efficiency for SODART (full line, Ref.~\cite{SODART-preprint}) and CAST (dashed line, Ref.~\cite{CASTeff}).}
\caption{Total number of candidate tracks derived for a single, dense field as a function of the \vtree, \pred and \platew kd-tree linking parameters.}
\caption{Total number of CLEAN NEO tracks derived for a single, dense field as a function of the \vtree, \pred and \platew kd-tree linking parameters.}
\caption{Ratio total number of tracks to number of CLEAN tracks derived for a single, dense field as a function of the \vtree, \pred and \platew kd-tree linking parameters.}
\caption{Ratio number of BAD tracks to number of CLEAN tracks derived for a single, dense field as a function of the \vtree, \pred and \platew kd-tree linking parameters.}
\caption{Ratio number of NONSYNTH tracks to number of CLEAN tracks derived for a single, dense field as a function of the \vtree, \pred and \platew kd-tree linking parameters.}
\caption{Ratio number of MIXED tracks to number of CLEAN tracks derived for a single, dense field as a function of the \vtree, \pred and \platew kd-tree linking parameters.}
\caption{CPU time for running \linkt on a single, dense field as a function of the \vtree, \pred and \platew kd-tree linking parameters.}
\caption{Memory usage of \linkt for a single, dense field as a function of the \vtree, \pred and \platew kd-tree linking parameters.}
\caption{\label{fig:acs_enc} The ACS Enclosure contains a laser (TOPTICA PicoFYb) that delivers a series of 1064\,nm 10~ps pulses into an optical fiber. A Laser Repetition-rate Controller (LRC) locks the fiber laser to a Cs frequency standard (Microsemi 5071A). A Universal Counter (UC: Agilent 53132A) enables a comparison between the Cs clock and a separate GPS-disciplined clock (TrueTime XL-DC). The dashed line represents occasional and brief reconfigurations to monitor the accumulating phase difference between the two clocks, as discussed in the text.}
\caption{\label{fig:lsb} Schematic diagram of the pulse processing system that selects, conditions, and delivers specific ACS laser pulses to the APD injection optics. Thick red and green lines represent optical fibers, while the thin black lines indicate electrical connections. The ACS pulse train is modulated by an Electro-Optic Modulator (EOM), which is normally held in the blocked (``off'') state. The Laser Slicer Board (LSB) generates RF gate pulses to make the EOM transmissive for one or more pulses. Details about the LSB and gate pulse generation are provided in the text. A Second Harmonic Generator (SHG, thermally regulated by a Thermo-Electric Controller, TEC) frequency-doubles the light to green (532\,nm). A variable attenuator (0--60\,dB) allows remote control of the pulse amplitude, generally tuned to deliver$\sim 1$ ACS photon per pulse to the APD array. Three optical taps enable the LSB to monitor the average laser power along the signal path, and also to generate an 80~MHz clock slaved to the pulse train. Signals labeled with \includegraphics[height=10pt]{img/rpi} indicate connections to a Raspberry Pi (GPIO or \iic{} or SPI).}
\caption{Left: raw data from the first LLR+ACS run obtained over a 500\,s period on 2016-09-12. Each dot represents a photon detection during the lunar gate. TDC time measurements are 25\,ps per bin; later photons appear lower in the plot. Yellow dots have been identified as ACS pulses based on nearly-static phase relative to the XL-DC clock. Red dots coincide with predicted lunar return times. Blue dots are the remainder---ones that cannot be confidently identified as either lunar or ACS---and largely represent background, slow avalanches due to carrier diffusion, or delayed crosstalk events in the APD (becoming more pronounced later/lower in the figure). The solid black line is constructed from the independent UC measurement of the APOLLO clock frequency referenced to the Cs clock. Horizontal dashed lines help to emphasize the degree to which the two clocks drift, and vertical dotted lines delimit the shot range during which the XL-DC clock steering reversed direction for 100\,s. Right: histogram of the lunar-prediction-corrected TDC values, smearing the ACS spikes into a triangular distribution while pulling lunar returns into a high-visibility signal---even though weaker than the aggregate ACS signal and spread over approximately the same TDC region. Color assignments match those in the left panel. The masking by ACS photons in the left panel looks five times worse here than it really is, because only one in five stripes is ``live,'' or possible, for a given shot.\label{fig:lun_raw}}
\caption{Using graphical models, we can visualize different ways to define the ``split'' constraint on manifold $\mathcal{S}$ in \eqref{complexity}. Here we consider a two-node network $X=(X_1, X_2)$ and its spatio-temporal stochastic interactions. \textbf{(A)}~$I(X;X^\prime)$ uses constraint \eqref{I_constraint}. \textbf{(B)}~$IF(X \rightarrow X^\prime)$ uses constraint \eqref{ay_constraint}. \textbf{(C)}~$\Phi_G(X \rightarrow X^\prime)$ uses constraint \eqref{amari_constraint}. \textcolor{red}{Dashed lines represent correlations that either may or may not be present in the input distribution $p$. We do not represent these correlations with solid lines in order to highlight (with solid lines) the structure imposed on the stochastic matrices.} Adapted and modified from \cite{amari}.}
\caption{\textcolor{red}{The mutual information $I$} upper bounds $IF$ and $\Phi_G$ when using random weight initializations sampled uniformly between $0$ and $1$ (averaged over 100 trials, with error bars).}
\caption{\textcolor{red}{A full model (left) can have both intrinsic (blue) and extrinsic (red) causal interactions contributing to its overall dynamics. Split models (\textbf{A,B}) formulated with an undirected output edge (purple) attempt to exclusively quantify extrinsic causal interactions (so as to strictly preserve intrinsic causal interactions after the ``split''-projection). However, the output edge can end up explaining away interactions from \textit{both} external factors $Y$ \textit{and} (some) internal factors $X$ (red + blue = purple). As a result, using such a family of split models does not properly capture the \textit{total} intrinsic causal interactions present in a system.}}
\caption{\textcolor{black}{Comparison of the results of the silicon oxide cluster and the ULK-fragment for a) the repair of the OH-damage and b) the repair of the H-damage. The a) blue and b) pink highlighted area of the ULK-fragment shows the OH- and H-damage, respectively. The dotted lines are only a guide for the eyes and show no correlations.}\textcolor{green}{{} \label{fig:ULK-fragment}}}
\caption{\label{fig:results1}\textbf{Actively choosing informative interventions speeds up learning}. The proposed methods ({\color{color2}\textbf{Algorithm 1}}, {\color{color3}\textbf{Algorithm 2}}) outperform {\color{color0}\textbf{only observing}} and {\color{color1}\textbf{randomly intervening}} by each metric of evaluation. Top and bottom row show results of experiments on learning $\mathcal{M}_1$ and $\mathcal{M}_2$ respectively. Experimental details described in Section~\ref{section:experiments}. Each experiment was performed many times; faded lines represent results from single trials, bold lines represent averages of these single trials.}
\caption{Illustration of the BLSTM network for mask prediction. \pfmarker{The numbers in brackets indicate the number of nodes per layer.}}
\caption{\pfmarker{Illustration of the r1MWF-$\mu$ filter for an example sentence (M06\_440C0201\_BUS) from the real test set.} (a) Spectral gain along frequency for different values of $\mu$. (b) The corresponding log-magnitude of one frame of the filtered signals. (c) and (d) Log-magnitude spectrograms of the filtered signals with $\mu=1$ and $\mu=\mu_G$, respectively. }
\caption{ \pfmarker{Relation between FV and WER for the real (solid black markers) and simulated (hollow red markers) data. The dashed lines show the linear regression results separately. A line with positive slope means positive correlation.}}
\caption{$(s,a,x,A)$ is the unique \cyannew{angle-monotone} %$\b$-monotone path to $A$, and $(s,b,x,B)$ is the unique \bluenew{angle-monotone} path to $B$, \bluenew{forming the cycle $(s,a,x,b)$}.}
\caption{Computational time comparison.~All experiments were performed on a computer with 2.7 GHz Intel Core i7 processor and 16 GB memory.~The best and the $2^{\rm nd}$ best results are colored with \textcolor{red}{red} and \textcolor{blue}{blue}, respectively. For frame numbers, frame size, and $p$ for inWLR see Section 3 and 4.}
\caption{Experimental Results on UCF11 Dataset. We report i) the accuracy score, ii) the number of parameters involved in the input-to-hidden mapping in respective models and iii) the average runtime of each training epoch. The models were trained on a Quad core Intel\textregistered Xeon\textregistered E7-4850 v2 2.30GHz Processor to a maximum of 100 epochs}
\caption{\emph{Top}: energy per site for $D=1,2,6$ simple update simulations using \textcolor{funnyBlue}{2x2} and \textcolor{funnyRed}{3x3} unit cells (color online); for each value of $\theta$ only the lowest (in energy) of the two is shown. The black dots correspond to a 1x1 unit cell state (energies for 2x2 and 3x3 simulations coincide). \emph{Bottom}: magnetization per site $m$ and the $II_Q$ and $III_Q$ tensor invariants of the $Q$-matrix per site for the $D=6$ simulations. $I_Q$ is identically zero (not shown).}
\caption{The phase transition between the \textcolor{darkRed}{AFQ3} and \textcolor{darkGreen}{$m=1/2$} phases (color online). \emph{Left:} energy per site for bond dimensions $D=5,6,\ldots,10$ as a function of truncation error, which we extrapolate to zero. \emph{Right:} extrapolated energy per site as a function of $\theta$. We estimate the phase transition to occur at the point where the energies per site intersect, i.e. at $\theta = 0.4886(7)\pi$.}
\caption{Vertical component {\color{revision1}$B_T$} of the magnetic field measured in the Van de Graaff setup at different positions along the axis of the 1~m long, 2-layer HTS shield at room temperature (`no SC shielding`) and with liquid nitrogen cooling (`with SC shielding`) at a nominal steering dipole field of 30~mT. The vertical lines indicate the extension of the HTS shield. The ordinate uses logarithmic scale. In addition, an offset of 0.1~mT (indicated by the horizontal grey line) is added to each measurement.}
\caption{Magnetic field component {\color{revision1}$B_T$} measured at various positions along the axis of the 4.5~inch long, 4-layer HTS and 4.5~inch long, 45-layer HTS shields in the Helmholtz coils setup at a nominal field of 40~mT. The vertical lines indicate the extension of the HTS shields. The ordinate uses logarithmic scale.}
\caption{Magnetic field component {\color{revision1}$B_T$} measured in the center of the 1~m long, 2-layer HTS shield in the Van de Graaff setup as a function of measurement time $t$ for nominal steering dipole magnet fields of {\color{revision1}$B_{a}$~=~15~mT and $B_{a}$~=~45~mT.}}
\caption{Top panel: Displacement $\Delta x_{Li7}$ of the $^{7}_{3}\mbox{Li}^{3+}$ beam passing through the 1~m long, 2-layer superconductor shield in the Van de Graaff setup (at room temperature and in its superconducting state) as a function of the nominal steering dipole field {\color{revision1}$B_{a}$}. Bottom panel: Mean field {\color{revision1}$B_T$} measured in the center of the shield as a function of the steering dipole field. Open symbols indicate time dependence of the measured field.}
\caption{Magnetic field component {\color{revision1}$B_T$} measured in the center of the 4.5~inch long, 45-layer HTS shield prototype inside the MRI magnet as a function of the nominal magnetic field $B_{a}$ in linear scale (top panel) and logarithmic scale (bottom panel). The open markers indicate field measurements showing an increase over time. The vertical lines mark $B_a = 0.5$~T and $B_a = 1.0$~T. {\color{revision1}A dashed line indicating $B_{T} = B_{a} \cdot 0.01$} is shown as well.}
\caption{Relative magnetic permeability $\mu_r$ of 4.5~inch {\color{revision1}long} epoxy / steel powder cylinders (with different fractional masses $f_M$ of steel powder in the mixture) as a function of the applied field $B_{a}$ measured in the MRI magnet.}
\caption{Magnetic flux {\color{revision1}$B_{T}$} in the direction of the Helmholtz coils axis at positions along measurement lines at 1~cm distance from the cloak (4.5~inch long, 4 layer HTS and ferromagnet with {\color{revision1}$\mu_r$~=~2.43(4)}, top panel) and across its center (bottom panel). The plots also show the field for the coils with only the superconducting cylinder (4-layer SC) and the coils itself (reference) measured at the same positions. Vertical lines indicate the cloak dimensions.}
\caption{Magnetic flux {\color{revision1}$B_{T}$} in the direction of the Helmholtz coils axis at positions along measurement lines at 1~cm distance in front of (top figure) and next to (bottom figure) the cloak (4.5~inch long, 4 layer HTS and ferromagnet with {\color{revision1}$\mu_r$~=~2.43(4)}). The plots also show the field for the coils with only the superconducting cylinder (4-layer SC) and the coils itself (reference) measured at the same positions. Vertical lines indicate the cloak dimensions.}
\caption{Magnetic flux {\color{revision1}$B_{T}$} in the direction of the Helmholtz coils axis at positions along a measurement line (the same positions as the top of Fig. \ref{fig:cloak_bvx_1}) at 1~cm distance from the cloak with the 4.5~inch long, four-layer and 4.5~inch long, 45-layer superconductor combined with the same ferromagnet shell ({\color{revision1}$\mu_r$~=~2.43(4)}). Vertical lines indicate the cloak dimensions.}
\caption{Magnetic flux {\color{revision1}$B_{T}$} in the direction of the Helmholtz coils axis at positions along a measurement line (the same positions as the top of Fig. \ref{fig:cloak_bvx_1}) at 1~cm distance from the cloak (4.5~inch long, 45 layer HTS and ferromagnet with {\color{revision1}$\mu_r$~=~2.43(4)}) for different angles between the gap separating the two halves of the HTS shield and the magnetic field. Vertical lines indicate the cloak dimensions.}
\caption{Magnetic flux {\color{revision1}$B_{T}$} in the direction of the MRI axis at positions along measurement lines at various distances from the 4.5~inch long, 45-layer HTS shield (top panel) and the cloak, i.e. the same superconductor shield surrounded by a ferromagnetic shell with {\color{revision1}$\mu_r$~=~2.43(4)} (bottom panel). The nominal field of the MRI magnet is 0.45~T.}
\caption{ A $2$-split $\Sigma$ in the octahedron $\Delta(2,4)$, with the common cell \textcolor{blue}{$H^\Sigma$} and the induced $2$-split in the vertex figure $\fig(e_I)$.}
\caption{Some single-case and multiple-case results from AVA \cite{Elawady2016}, PSUm \cite{Rauschert2011,Liu2013}, and NYm \cite{Cicconet2016} datasets with groundtruth ({\color{blue} blue}) in 1st row (a-e). our methods in 2nd and 3rd rows (f-o) produces better results among El2016 \cite{Elawady2016} in 4rd row (p-t) and Loy2006 \cite{Loy2006} in 5th row (u-y). For each algorithm, the top five symmetry results is presented in such order: {\color{red} red}, {\color{yellow} yellow}, {\color{green} green}, {\color{blue} blue}, and {\color{magenta} magenta}. Each symmetry axis is shown in a straight line with squared endpoints. Best seen on screen.}
\caption{(Color online) TDSE computations of the short and long trajectory contributions to HHG for a two-cycle laser pulse of $2.55\times10^{14}$ W.cm$^{-2}$ peak intensity for model molecules of different internuclear distances indicated in a.u. in the legend. We report the harmonic intensity (a) [(c)] and phase (b) [(d)] calibrated by the atomic reference for the short [long] trajectories {\color{creferee1}(solid lines), and the phase of the transition dipole matrix element computed numerically for the different internuclear distances [dashed lines in (b) and (d)].}}
\caption{ (Color online) Detail of the spin \textcolor{blue}{up} current along the interface in an n-TI junction, showing a ``snake-like" pattern in the ``n side'' (left) of the junction. In the ``TI side" (right), {the spin up electrons flow parallel to the interface to the bottom edge state, where the left-right transmission across the junction takes place.} }
\caption{Height function filtration of a ``clean'' (\emph{left}, \textcolor{darkgreen}{green} points) and a ``noisy'' (\emph{right}, \textcolor{babyblue}{blue} points) shape along direction $\mathbf{d} = (0, -1)^\top$. This example demonstrates the insensitivity of homology towards noise, as the added noise only (1) slightly shifts the dominant points (upper left corner) and (2) produces additional points close to the diagonal, which have little impact on the Wasserstein distance and the output of our layer. \label{fig:npht}}
\caption{ \label{tbl:vectorization} \emph{Left}: Classification accuracies for a linear SVM trained on vectorized (in $\mathbb{R}^N$) persistence diagrams (see Sec.~\ref{subsection:vectorization}). \emph{Right}: Exemplary visualization of the learned structure elements (in $0$-th dimension) for the \texttt{Animal} dataset and filtration direction $\mathbf{d} = (-1,0)^\top$. Centers of the learned elements are marked in \textcolor{blue}{blue}.}
\caption{\textcolor{red}{Most probably will have to changed.} %Automatic Recognition of Facial Expressions of Felt and Unfelt Emotions = how about Automatic recognition of FEFUE? or automatic recognition of unfelt emotions (ARUE)? ...formerly ARDFES. Overview of the propose method. (a) Diagram of the proposed automatic recognition of unfelt facial expressions of emotion method. Feature maps are computed with a VGG-Face CNN and accumulated along facial geometries. Clustering and encoding in the form of Fisher Vectors are used to produce a compact feature vector per sequence. (b) An example of detected facial geometries. (c) Example of activation maps of the CNN.}
\caption{(Color online) Left panel: Maximal value of an effective width of the spatial density profile $\mathrm{max}\left[\sigma(t)\right]$ as a function of $\omega_0$ and $g_W$ for vanishing coupling $g_J = 0$. A sharp crossover between non-spreading excitations (blue) and spreading excitation (dark red) is clearly visible. Different regions of the phase-diagram (bordered with white lines) correspond to a distinct nature of the exciton-vibration dynamics. Right panel: Evolution of the excitation width $\sigma(t)$ for different points on the phase diagram (marked as white squares on the left panel).} \label{fig1} \end{figure} First, we focus on the case of a completely localized initial state $|\boldsymbol{\Psi}_0\rangle$ for $g_J=0$ (Fig.~\ref{fig1}). We qualitatively indicate five different regions on the phase diagram (left panel): {\it (I)} and {\it (II)} where the excitation is dressed by a cloud of vibrations and the excitation does spread; {\it (III)} where due to an exponential reduction of the hopping amplitude the excitation is localized in its initial position \cite{Fesser1982}; {\it (IV)} where the sum of vibration energy and exciton-vibration coupling is larger than the hopping energy, giving rise to Davydov-like soliton behavior; {\it (V)} where $g_W\gtrsim\omega_0$ corresponding to the Discrete Breathers-like behavior \cite{Flach1998,Juanico2007}. Distinct behavior of the system is also visible in these selected areas in the time evolution of $\sigma(t)$ (right panel of Fig.~\ref{fig1}). This picture can be generalized to non-vanishing coupling $g_J$, which we investigate for the the second initial state $|\tilde{\boldsymbol{\Psi}}_0\rangle$ (Fig.~\ref{fig2}). As can be seen, a slight delocalization of the initial state together with non-local coupling $g_J$ dramatically enhance the non-spreading behavior of the wave packet. It is a direct consequence of the non-local terms in \eqref{DavidovEq}. \begin{figure} \includegraphics[scale=0.45]{Figure2.eps} \caption{(Color scale) Maximal value of the wave packet width $\mathrm{max}\left[\sigma(t)\right]$ for different initial states $|\boldsymbol{\Psi}_0\rangle$ and $|\tilde{\boldsymbol{\Psi}}_0\rangle$ (top and bottom row, respectively) and different non-local interactions $g_J=\{0,3,5\}$ (appropriate columns from left to right). Note that strong enhancement of the non-spreading behavior takes place for stronger $g_J$ and for a smeared out initial state.}
\caption{ Charge susceptibility for the square lattice Hubbard-Coulomb model for three different interaction strengths. The color scale corresponds to $V(q) \mbox{Im} \chi_\rho (\omega, q)$. For the spectral function $\mbox{Im} \chi_\rho (\omega, q)$ we use the estimated value after analytic continuation (see (\ref{ConvolutionDeltaFunction})). The same lattice ensembles were used as in Fig.~\ref{fig:DOS1}. Standard Tikhonov regularization with $\lambda=5 \times 10^{-6}$ and additional time averaging according to Fig.~\textcolor{red}{4a} is used in all cases. }
\caption{The (most)-massive-stellar-mass--embedded-cluster-mass ($m_{\mathrm{str, max}}$--$M_{\mathrm{ecl}}$) relation. Optimally-sampled result for the most, second and third massive star mass as a function of embedded cluster mass is shown as solid curve, \textcolor{blue}{blue dashed curve} and \textcolor{green}{green dotted curve}, respectively. Observational data come from: \cite{Kirk2012} (\textcolor{orange}{orange dots}), \cite{Stephens2017} (\textcolor{red}{red dots}), \cite{Weidner2013a} (\textcolor{gray}{gray dots}) and \cite{RamirezAlegria2016} (\textcolor{blue}{blue dots}), where \cite{Weidner2013a} is an inhomogeneous set of data culled from the literature for very young clusters without supernova remnants. The average $M_{\mathrm{ecl}}$ uncertainty is 0.34 dex and the average $m_{\mathrm{str,max}}$ uncertainty is 0.16 dex for the gray dots and the gray color is lighter for data points with larger uncertainties. The thin solid line indicates the $M_{\mathrm{ecl}}=m_{\mathrm{str,max}}$ limit and the horizontal thin dashed line indicates the 150 M$_{\odot}$ limit in our Eq. \ref{eq:1intMstar}.}
\caption{The most-massive young-cluster-mass--galaxy-wide-SFR ($M_{\mathrm{ecl, max}}$--SFR) relation. Optimally-sampled results for different $\beta$ are shown as \textcolor{red}{red solid curve} for $\beta$ following Eq. \ref{eq:beta-SFR}, \textcolor{blue}{blue dotted curve} for $\beta=2$, \textcolor{green}{green dashed curve} for $\beta=2.4$. Observational data (gray dots) adopted from \cite{Weidner2004} have a typical uncertainty of 0.3 dex. The $\beta=2$ and 2.4 curves are almost identical with \cite{Weidner2004}'s figure 6 (middle dotted and middle dashed curves, respectively) that also adopt the $\delta t =10$ Myr assumption. See also \cite{Randriamanakoto2013}.}
\caption{Logarithmic integrated galaxy-wide IMFs, $\xi_{\mathrm{L,IGIMF}}$ (Eq.~\ref{eq:xi_L_IGIMF}), for different SFRs and formed over a 10 Myr epoch. The unit of $\xi_{\mathrm{L,IGIMF}}$ is number of stars per log-mass interval. Each line is normalized to the same values at $m<1$ M$_{\odot}$. Solid curves are IGIMFs for galactic SFR=\textcolor{blue}{$10^{-5}$}, \textcolor{green}{$10^{-4}$} ... $10^{5}$ M$_{\odot}$/yr from bottom left to top right. The dashed line is the canonical IMF (Eq.~\ref{eq:xi_star}) with $\alpha_3=2.3$. The lines lower than the dashed line deviate from the canonical IMF above 1.1 M$_{\odot}$ which is the most massive stellar mass in the least-massive embedded clusters $M \approx M_{\mathrm{min}} = 5$ M$_{\odot}$. The lines on top of the dashed line deviate from the canonical IMF above 1 M$_{\odot}$ which is the lower mass limit for applying $\alpha_3$ as defined in Eq. \ref{eq:xi_star} and \ref{eq:alpha_3}. This plot is comparable with earlier IGIMF works from \cite{Weidner2013b} and \cite{Fontanot2017}.}
\caption{Optimally-sampled galaxy-wide IMFs for different SFRs. The unit of $\xi_{\mathrm{OSGIMF}}$ is number-of-stars-per-linear-mass-interval. The black histogram is the entire stellar sample in the galaxy formed in a 10 Myr epoch. The large scatter in the upper left panel and the serrated feature around $\log_{10}(m/\mathrm{M}_{\odot})=0$ for every panel is explained in the text. The \textcolor{red}{red}, the \textcolor{green}{green} and the \textcolor{blue}{blue} histograms, comprising our decomposition plot, are the OSGIMFs counting only the first, the second and the third most-massive stars of embedded clusters. Note the spoon feature at $m>10^{1.6}$ M$_\odot$ in the red, green and blue curves. The stellar mass distribution is saturated at the high mass end because of the 150 M$_{\odot}$ limit. The thin smooth curves are the IGIMFs as in Fig. \ref{fig:IGIMF_SFR_}, calculated by Eq. \ref{eq:xi_IGIMF} and normalized to give the mass in stars when integrated over the relevant stellar mass ranges.}
\caption{The observed high mass end power-law index of the galaxy-wide IMF as resulting from the here calculated IGIMF, $\alpha_3^{\mathrm{gal}}$ (i.e. $\alpha^{\mathrm{gal}}$ in Eq.~\ref{eq:alpha_gal} for $m>1$ M$_{\odot}$), for a constant SFR over $\delta t=10$ Myr in dependence of the galaxy-wide SFR. In Fig. \ref{fig:OSalphaMstar}, $\alpha_3^{\mathrm{gal}}$ values diverge for different SFRs and also vary for different $m$ at the high mass end. As at each $m$ value there exists a different $\alpha_3^{\mathrm{gal}}$--SFR relation, we plot solid lines for $\log_{10}(m/$M$_{\odot})=$0.2, 0.4, ..., 2, i.e., 1.58, 2.51, ..., 100 M$_{\odot}$ from black to gray (top to bottom) for the fiducial model and dotted lines for the corresponding SolarMetal model defined in Sec. \ref{sec:Model-IGIMF}. \textcolor{blue}{Blue squares} are data from the GAMA galaxy survey \citep{Gunawardhana2011}. \textcolor{red}{Red triangles} and the \textcolor{red}{red dash-dotted line} are data from \cite{Weidner2013b} where the left triangle is for the MW field, middle three triangles are galaxy studies, the right triangle is for the bulges of the MW and M31 and the dash-dotted line is their IGIMF model assuming $\beta=2$. A recent study has suggested that the 2 M$_{\odot}$/yr SFR for MW is overestimated \citep{Chomiuk2011} but we leave this data point the same as in \cite{Weidner2013b}. \cite{Gargiulo2015} report consistency between their IGIMF model assuming $\beta=2$ (\textcolor{orange}{thick yellow dashed line}) and the [$\alpha$/Fe] abundance ratios of elliptical galaxies. The \textcolor{purple}{purple diamond} is an individual analysis for the dwarf galaxy NGC 2915 \citep{Bruzzese2015}. \textcolor{green}{Green stars} are based on the \cite{Lee2009} 11HUGS observations of dwarf galaxies. The black circle is an observation for the solar neighborhood from \cite{Rybizki2015} with adopted MW SFR from \cite{Robitaille2010} as an upper limit of the solar neighborhood SFR because the Sun is located in an inter-arm region where the relevant SFR is significantly smaller (towards the direction indicated by the arrow, see Sec.~\ref{secsub:alpha3-SFR} for further details). The thin horizontal dashed line represents the canonical IMF index $\alpha_2=\alpha_3=2.3$.}
\caption{The optimally-sampled most-, second- and third-most massive star to be found in a galaxy as a function of the galaxy-wide SFR are shown as solid curve, \textcolor{blue}{blue dashed curve} and \textcolor{green}{green dotted curve}, respectively, resulting from a combination of Fig.~\ref{fig:MmaxMecl} and Fig.~\ref{fig:SFRMecl}. The thin horizontal dashed line indicates the mass-limit below which SNII explosions are not likely ($8\,$M$_\odot$), and the vertical thin dashed line shows the SFR below which galaxies are not expected to host SNII events, subject to the axioms adopted in the present study. See Sec.~\ref{secsub:mmax-SFR} for more details.}
\caption{Same as Fig.~\ref{fig:IGIMF_SFR_} but using Eq.~\ref{eq:alpha_3_metal}, with \textcolor{green}{green}, \textcolor{cyan}{cyan}, \textcolor{yellow}{yellow}, and \textcolor{blue}{blue} lines for SFR = $10^{-4}$, $10^{-2}$, $10^{0}$, $10^{2}$ M$_{\odot}$/yr, respectively. There are three cases for each color with [Fe/H] = -3, 0, and 2 from upper to lower lines labeled in the plot, however, the three green lines overlap and the lower two cyan lines also overlap.}
\caption[OTB-2013 dataset]{\textbf{OTB-2013: } Our \textcolor{blue}{p-tracker} (solid line) compared to \textcolor{green}{FCNT}~\cite{Wang2015}, \textcolor{red}{MUSTer}~\cite{hong2015multi}, \textcolor{yellow}{LTCT}~\cite{ma2015long}, \textcolor{cyan}{TLD}~\cite{kalal2012tracking} and \textcolor{magenta}{SPL}~\cite{supancic2013self} (dashed lines). In general, many videos are easy for modern trackers, implying that a method's rank is determined by a few challenging videos (such as the confetti celebration and fireworks on the bottom left). Our tracker learns an UPDATE and REINIT policy that does well on such videos.}
\caption[New Internet dataset]{\textbf{Internet videos} contain new challenges, such as cuts, strange and interesting behaviors, fast motion and complex illumination. Our \textcolor{blue}{p-tracker} (solid line) learns a policy that outperforms \textcolor{green}{FCNT}~\cite{Wang2015}, \textcolor{red}{MUSTer}~\cite{hong2015multi}, \textcolor{yellow}{LTCT}~\cite{ma2015long}, \textcolor{cyan}{TLD}~\cite{kalal2012tracking} and \textcolor{magenta}{SPL}~\cite{supancic2013self} (dashed lines).}
\caption{Relevancy Clustering with Top-10 Regions. \\ \textcolor{red}{Red}: relevant objects (point labels indicate generated captions), \textcolor{blue}{Blue}: irrelevant objects.}
\caption{Cyclical sources of HE \gray{} emission. The values shown in bold type are obtained from a more sensitive analysis of the 3FGL source belonging to the pixel.}
\caption{Figure shows periodic-like \gray{} sources in the \textit{Fermi}-LAT sky. These include three binaries (the corresponding pixels are shown in blue) and seven blazars (shown in yellow). Modulated \gray{} signals from the pixels shown in red are due to instrumental effects, see the text for details. This figure can be seen in colour in the online version.}
\caption{\Structured\methods that accelerate CNNs: (a) a network with 3 conv layers. (b)\sparseconnect\deactivates some connections between channels. (c)\tensordecom\factorizes a\conv\into several pieces. (d)\channelpruning\reduces number of channels in each layer (focus of this paper).}
\caption{Comparisons for \xceptionfifty, under \x{2} acceleration ratio. The baseline network's top-5 accuracy is \xceptionorig\%. Our approach outperforms previous approaches. Most \structured\methods are not effective on Xception architecture (\textit{smaller is better}).}
\caption{Illustration of multi-branch enhancement for residual block. \textbf{Left}: original residual block. \textbf{Right}: pruned residual block with enhancement, $\mathrm{c_x}$ denotes the \featch. Input channels of the first \conv\are sampled, so that the large input\featch\could be reduced. As for the last layer, rather than approximate$\reY_2$, we try to approximate $\reY_1+\reY_2$ directly (Sec.~\ref{sec:multi} Last layer of residual branch).}
\caption{\textbf{Top row}: Single maze, multiple goals, {\color{red} P} represents the position of the player, and {\color{blue}G} is the goal. The first three configurations were from training, and the last tested the agent's ability to generalize. The imagination steps are the shaded blue areas, for different imagination budgets 5, 9, 13 and 15 from left to right. The imagination depth corresponds to the color saturation of the grid cells. The actual action the agent took is shown as a solid black arrow. \textbf{Bottom row:} Multiple mazes and one example run. Left two: two example mazes, one with more candidate goal locations than the other. Middle and right four: one example run of a learned agent on one maze. In this run, the agent performed four imagination steps, then took its first action, then performed one more imagination step. Both the imagination steps and the rollout updates to $c$ are shown. In the rollout updates, the original $Q+c$ values before the rollout are colored gray, and the updates are blue.}
\caption{ % % % % % % [top panel:] Parameters common to all models: BH spin $a$, the radii of the torus inner edge $r_{\rm in}$ and pressure maximum $r_{\rm max}$, the initial tilt angle $\mathcal{T}_{\rm init}$, and simulation duration $t_{\rm sim}$. [bottom panel:] Short and full model name (i.e., S25A93 stands for size and BH spin, W for weakly magnetised, LR or HR for low or high resolution), the strength of the magnetic flux, the resolution $N_{r,\theta,\phi}$, and the quality factor $Q_{r,\theta,\phi}$ (the number of cells per MRI fastest growing wavelength in $r-$, $\theta-$, and $\phi-$directions) at $t = 5\times 10^4 \: t_g$. \highlight{Short names of the models link to 3D animations in a \href{https://www.youtube.com/playlist?list=PL39mDr1uU6a5RYZdXLAjKE1C_GAJkQJNv}{YouTube playlist}.}}
\caption{ An example from the SQuAD dataset. The BiDAF Ensemble model originally gets the answer correct, but is fooled by the addition of an adversarial distracting sentence (\textcolor{blue}{in blue}). }
\caption{Outline response functions for Hubble-type spherical outflow \textcolor{blue}{\textbf{(left)}}, a rotating Keplerian disc viewed at a $20^{\circ}$ angle \textcolor{magenta}{\textbf{(centre)}}, and Hubble-type spherical inflow \textcolor{red}{\textbf{(right)}}. Winds extend from $r_{min}=20r_{g}$ to $r_{max}=200r_{g}$ for an AGN of mass $10^{7} M_{\odot}$. Hubble out/inflows have $V(r_{min})=\pm 3\times 10^{3}$ km s$^{-1}$. Solid lines denote the response from the inner and outer edges of the winds, dotted lines from evenly-spaced shells within the wind. Pale lines describe the edge of the velocity-delay shape of the response function.}
\caption{ Model figure. (a) Two dimensional square lattice with occupied ($n=1$) or unoccupied ($n=0$) sites. Filled circles indicate the occupied sites. Inclinations of the rods towards the horizontal direction show respective particle orientations $\theta\in\left[0,\pi\right]$. (b) Equilibrium move: particle can move to any of the four neighbouring sites with equal probability $1/4$. (c, d) Active move: particle can move to either of its two neighbouring sites with probability $1/2$, if unoccupied, in the direction it is more inclined to, i.e., along $BD$ in (c), and $AC$ in (d). } \label{fig:model} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section*{Model}\label{secmodel} \noindent We consider a collection of apolar particles on a two dimensional square lattice, as shown in schematic diagram Fig. \ref{fig:model}(a). Occupation number `$n_i$' of the $i^{th}$ lattice site can take values $1$ (occupied) or $0$ (unoccupied). Orientation $\theta_i$ of apolar particle at the $i^{th}$ site can take any value between $0$ and $\pi$. The model follows two sequential processes at every step; first, a particle moves to a nearest neighbouring site with {\it some probability}, and then orientation of the particle is updated based on its nematic interaction with its nearest neighbours. We define two kinds of models on the basis of particle movement: (i) `Equilibrium model' (EM) - particle moves with equal probability $1/4$ to any of the four neighbouring sites (Fig. \ref{fig:model}(b)), (ii) `Active model' (AM) - in this model particle movement occurs in two steps. First, it chooses a direction along which it is more inclined. As shown in Fig. \ref{fig:model}(c,d), it chooses the direction of movement along $BD$ if $ \pi /4 < \theta \le 3\pi/4$ and along $AC$ otherwise. In the second step, it moves to a randomly selected site between the two nearest neighbouring sites along the chosen direction. For example, if $BD$ is selected as the direction of movement, then the particle moves to randomly selected site $B$ or $D$ in the second step. In both the models, we consider volume exclusion, i.e., particle movement is allowed only if the selected site is unoccupied. In both the models, the particles also interact with their nearest neighbours. The interaction depends on the relative orientation of the particles and is represented by a modified Lebwohl-Lasher Hamiltonian \cite{llasher} \begin{equation} \mathcal{H} = -\epsilon \sum_{<ij>}n_i n_j \cos2(\theta_i-\theta_j) \label{eqll} \end{equation} where $\epsilon$ is the interaction strength between two neighbouring particles. The interaction in equation (\ref{eqll}) governs the orientation update of the particle. We employ Metropolis Monte-Carlo (MC) algorithm \cite{mcbinder} for orientation update of the particle after the movement trial. In both the models, an order parameter defining the global alignment of the system does not remain conserved during the MC orientation update described above. In actual granular or biological systems where mutual alignment emerges because of steric repulsion, orientation of particles need not to follow a conservation law. An order parameter defined by coarse-graining the orientation in our present model is a class of non-conserved order parameter: {\it Model A} as described by Hohenberg and Halperin \cite{HohenbergHalperin}. Both the models EM and AM comprise of two different physical aspects - motion of the particles and nematic interaction amongst the nearest neighbours. If the particles are not allowed to move, the models reduce to an apolar analogue of the diluted XY-model with nonmagnetic impurities \cite{dilutedxymodel}, where impurities and spins are analogous to vacancies and particles, respectively. However, unlike the diluted XY-model, particles in these models are dynamic. In the EM, the particle diffuses to neighbouring sites, whereas it moves anisotropically in the AM. The anisotropic movement of the active particles arises in general because of the self-propelled nature of the particles in many biological \cite{kemkemer} and granular systems \cite{vnarayanjstat, vnarayanscince}. This move produces an active curvature coupling current in the coarse-grained hydrodynamic equations of motion \cite{shradhanjop, sradititoner}. The AM does not satisfy the detailed balance principle \cite{mcbinder}, because of the orientation update after the anisotropic movement. The coupling of the particle movement with the orientation update in our active model is analogous to the active Ising spin model introduced by Solon and Tailleur \cite{solonprl, solonpre}, where the probabilistic flip of the spins is an equilibrium process, whereas the out-of-equilibrium aspect of the model is attributed to the anisotropic movement probability of the spins. However, their orientation update algorithm \cite{solonprl, solonpre} is similar to kinetic Monte-Carlo, whereas we use Metropolis Monte-Carlo algorithm to update particle orientation. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section*{Numerical Study}\label{secnumerical} \noindent We consider a collection of $N$ particles with random orientation $\theta \in[0,\pi]$ homogeneously distributed on a $L \times L$ lattice ($L=256, 400, 512$) with periodic boundary. Packing density of the system is defined as $C=N/(L \times L)$. We choose a particle randomly, move it to a neighbouring site obeying exclusion, and then update its orientation using Metropolis algorithm. In each iteration, we repeat the same process for $N$ number of times, and we use $1.5\times10^6$ iterations to achieve the steady state of the system. We obtain the steady state results by averaging the observables over next $1.5 \times 10^6$ iterations and use more than twenty realisations for better statistics. The ordering in the system is characterised by a scalar order parameter defined as \begin{equation} S=\sqrt{\left(\frac{1}{N}\sum_i n_i \cos(2 \theta_i)\right)^2+\left(\frac{1}{N}\sum_i n_i \sin(2 \theta_i)\right)^2}. \label{eqops} \end{equation} It is proportional to the positive eigenvalue of the nematic order parameter $\Q$ \cite{pgdgenne}. It takes the minimum value $0$ in the disordered state and the maximum value $1$ in the complete ordered state. First we study the EM as a function of inverse temperature $\beta= 1 / k_BT$ for different packing densities. As shown in Fig. \ref{figtemp}, the system shows disordered isotropic to nematic state (I-N) transition with decreasing temperature. In contrast to the first order I-N transition in the equilibrium Lebwohl-Lasher model in three dimensions \cite{llasher,chaiklub}, we find continuous transition for the EM defined in two dimensions. The observed nature of transition supports the study by Mondal and Roy \cite{mondalroy}. Similar to the diluted XY-model \cite{dilutedxymodel}, the critical inverse temperature $\beta_c(C)$ increases with density in the EM. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection*{Phase diagram} \noindent We construct phase diagram for both the equilibrium model and the active model on the density-temperature plane. As shown in Fig. \ref{fig:phase_snap}(a), two distinct states appear in the EM - (i) an equilibrium isotropic (EI) state on the left side of the red boundary and (ii) an equilibrium nematic (EN) state on the right side of the red boundary. In the EI state, particles remain disordered and homogeneously distributed throughout the system. Consequently, the scalar order parameter $S\simeq0$ in this state. With increasing density or decreasing temperature the particles get mutually ordered and form the EN state ($S>0$). As shown in Fig. \ref{fig:S_C_diag}(a), for a fixed temperature the scalar order parameter increases continuously with increasing density, and the system enters into the nematic state. Both the particle orientation and the coarse-grained density remain homogeneous in the EN state, as shown in the real space snapshot Fig. \ref{fig:phase_snap}(b). Similar to the EM, the active system remains in a homogeneous disordered isotropic (I) state in the high temperature and/or low packing density regime (cyan coloured regime in the phase diagram Fig. \ref{fig:phase_snap}(a)). With increasing density or decreasing temperature, beyond the I state, the active system enters into an inhomogeneous mixed (IM) state (golden regime in the phase diagram Fig. \ref{fig:phase_snap}(a)), where locally ordered high-density domains coexist with disordered low-density regions. In the low temperature regime ($\beta\epsilon \in[1.9, 2.2]$), the I to IM state transition with increasing $C$ occurs with a jump in the scalar order parameter $S$, as shown in Fig. \ref{fig:S_C_diag}(a). In the very beginning of the IM state, as indicated by cross symbols in Fig. \ref{fig:phase_snap}(a), we find a banded state (BS) in the low temperature regime, where particles cluster and align themselves within a strip to form band. However, out of the strip the system remains disordered with low local density, as shown in the real space snapshot Fig. \ref{fig:phase_snap}(b). On further increment of the packing density $C$, bands formed in different directions start mixing leaving the system with many locally ordered high density patches separated by low density disordered regions. Typical real space snapshots for the orientation and the coarse grained density in the IM state are shown in Fig. \ref{fig:phase_snap}(b). The jump in the $S - C$ curve reduces with increasing temperature, and no bands appear in the high temperature ($\beta\epsilon < 1.9$) regime. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[b] \centering \includegraphics[width=0.4\linewidth]{Figure4.pdf} \caption{ Density fluctuation $\Delta N = \sqrt{<N^2> - <N>^2}$. All the active ordered states show large density fluctuation obeying the relation $\Delta N \sim \left<N\right>^{\zeta}$ with $\zeta > 1/2$. The active disordered isotropic state shows normal density fluctuation with $\zeta = 1/2$.} \label{fig:gnf} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Figure \ref{fig:S_C_diag}(a) shows that the I to BS transition occurs in the low temperature regime with a jump in $S$ at a density lower than the corresponding equilibrium I-N transition density $C_{IN}$. These bands appear because of the large activity strength. A linear stability analysis, as detailed later in this paper, shows that the large activity strength induces an instability in the disordered isotropic state. This instability goes away for small activity strength or at high temperature. We also do a renormalised mean field calculation of an effective free energy written for the active nematic. The calculation predicts a jump in the scalar order parameter and shows a shift in the disordered $(S=0)$ to ordered $(S\ne 0)$ state transition density. Both the jump in $S$ and the shift in the transition density reduce with the activity strength or increasing temperature. The I to BS transition is a first order transition. The shift in the disorder-order transition point is a common feature of the active systems. For large activity and low temperature, if the system density is above a certain value but less than $C_{IN}$, the large density fluctuation present in these systems causes local alignment with local density higher than $C_{IN}$. Large density fluctuation is an intrinsic feature of the active systems, and as shown in Fig. \ref{fig:gnf}, we also observe the same in the ordered active states in our model. Due to activity these locally ordered regions move anisotropically and combine with nearby region with similar local ordering. So larger ordered region forms at mean density lower than the equilibrium I-N transition density. Therefore, we find a disordered to ordered state transition at a lower density. For large activity strength, I-BS transition occurs with the jump in scalar order parameter. In our numerical study we calculate the probability $P(S)$ of the scalar order parameter averaging over many iterations and realisations near the I-BS transition point. Figure \ref{fig:S_C_diag}(b) shows $P(S)$ has two peaks, which further supports the first order I-BS transition for large activity strength. In the high density regime (red coloured regime in the phase diagram Fig. \ref{fig:phase_snap}(a)), the AM shows bistability, i.e., it can be either in the locally ordered IM state or in a homogeneous globally ordered (HO) state. As shown in Fig. \ref{fig:S_C_diag}(a), the $S - C$ curve for fixed temperature bifurcates in the high density regime; the lower branch corresponds to the earlier discussed IM state, whereas the higher branch indicates the existence of the globally ordered state. Figure \ref{fig:phase_snap}(b) shows that the system possesses less density inhomogeneity in the HO state compared to the IM state. A finite size scaling of both the HO and the IM state, as shown in the Fig. \ref{fig:S_C_diag}(c), shows that the active nematic possesses non-zero finite order in both these states. Order parameter time series shown in Fig. \ref{fig:S_C_diag}(d) confirms the bistability of the system in the high density regime. Bistability is not generally seen in other agent based numerical simulations of point particles \cite{ngo}; it appears because of finite filling constraint of the model. This feature can be suppressed if we allow more than one particle to sit together. In the complete filling limit $C =1.0$, the AM is equivalent to the EM, and it shows the globally ordered HO state only. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[b] \includegraphics[width=0.6\linewidth]{Figure5.pdf} \caption{ Two-point orientation correlation shown for $\beta\epsilon=2.0$ on log-log scale. (a) Active model: $g_2(r)$ decays exponentially at low density ($\bigcirc$, $\square$) and algebraically at high density ($\diamond$, $\bigtriangleup$). In the bistable regime at high density ($\bigtriangleup$), $g_2(r)$ decays algebraically in the HO state and abruptly in the IM state. (b) Equilibrium model: $g_2(r)$ decays exponentially at low density ($\bigcirc$, $\square$) and algebraically at high density ($\diamond$, $+$, $\bigtriangleup$). Continuous lines are the respective fits, fitted for more than one decade.} \label{fig:correlation} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection*{Two-point orientation correlation} \noindent We further characterise various states on the basis of the two-point orientation correlation in the different states of the equilibrium and the active nematic. It is defined as $ g_2(r) = <\sum_i n_i n_{i+r} $ $\cos\left[2\left(\theta_i-\theta_{i+r}\right)\right]/\sum_i n_i^2 > $ where $r$ represents interparticle distance, and $< . >$ signifies an average over many realisations. Figure \ref{fig:correlation}(a, b) show $g_2(r)$ versus $r$ plots on log-log scale for the AM and the EM, respectively, for a fixed inverse temperature $\beta\epsilon=2.0$. In the AM, $g_2(r)$ decays exponentially at low packing density $C<0.37$, i.e., in the isotropic state. Therefore, the active isotropic is a short-range-ordered (SRO) state. In the BS at $C=0.38$, $g_2(r)$ decays following a power law. Therefore, the system is in a quasi-long-range-ordered (QLRO) state. Ordering increases with density. At high packing density, correlation function confirms the bistability in the active system. At $C=0.82$, $g_2(r)$ shows power law decay in the HO state, whereas in the IM state $g_2(r)$ decays abruptly after a distance $r$. The abrupt change in $g_2(r)$ at a certain distance indicates the presence of locally ordered clusters in the IM state. In contrast, the equilibrium system shows a transition from SRO (exponential decay) isotropic state at low density $C\lsim0.48$ to QLRO (power law decay) nematic state at high density $C\gsim0.50$. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{figure}[b] \includegraphics[width=0.5\linewidth]{Figure6.pdf} \caption{Steady state characteristics of high density states. (a) Orientation distribution $P(\theta)$ of particles calculated from one snapshot in the steady state. $P(\theta)$ fits with Gaussian distribution (continuous lines) for both the HO and the EN states. The IM state shows broad distribution of $\theta$. (b) Distribution of the mean orientation $P(\bar \theta)$ calculated from $\bar \theta$ of each snapshot in the steady state. $P(\bar \theta)$ is broad for the EN state in comparison to the HO state. (c) Steady state autocorrelation $C_{\bar \theta}(t)$ of the mean orientation of the system. All plots are shown for $(\beta\epsilon, C)=(2.0,0.80)$. } \label{fig:distribution} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \subsection*{Orientation distribution and autocorrelation of the mean orientation} \noindent We compare the steady state properties of the active and the equilibrium models in the high density limit. First we calculate the steady state (static) orientation distribution $P(\theta)$ from a snapshot of particle orientation $\theta$. As shown in Fig. \ref{fig:distribution}(a), both the active HO and the equilibrium nematic show Gaussian distribution of orientation. Peak position of $P(\theta)$ for both the EN and the HO state can appear at any point between $0$ and $\pi$ because of the continuous broken rotational symmetry of the Hamiltonian shown in equation (\ref{eqll}). Data shown in Fig. \ref{fig:distribution}(a) is for one realisation only, and for other realisations also the distribution $P(\theta)$ remains Gaussian with peak at other $\theta$ values. Therefore, orientation fluctuation of the particles in the active HO state is same as in the equilibrium nematic state. The distribution $P(\theta)$ in the IM state is very broad and spans over the whole range of orientation. Therefore, the system possess no global ordering in the IM state. We also calculate the time averaged distribution $P(\bar \theta)$ of mean orientation of all the particles in the active HO and the equilibrium nematic states. The mean orientation $\bar \theta(t)$ of all particles is calculated for each iteration time $t$ in the steady state. The distribution $P(\bar \theta)$ of the mean orientation is obtained from these $\bar \theta(t)$ data. This distribution is a measure of the fluctuation in the global orientation of the particles in the steady state. As shown in Fig. \ref{fig:distribution}(b), $P(\bar \theta)$ in the active HO state is narrow in comparison to the broad distribution in the EN state. We also calculate the autocorrelation of the mean orientation $C_{\bar\theta}(t)=<\frac{1}{t} \sum_{\tau=1}^t cos\left[2\left\{\bar\theta\left(t_0\right)-\bar\theta\left(t_0+\tau\right)\right\}\right]>$ in the steady state. As shown in Fig. \ref{fig:distribution}(c), $C_{\bar \theta}(t)$ decreases with time in the EN state, but remains unchanged in the active HO state. Both these results imply that the fluctuation in the global orientation direction $\bar \theta$ in the active HO state is small compared to the EN state. We do not calculate the mean orientation $\bar \theta$ in the active IM state, because the system possesses no global ordering in this state. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section*{Phenomenological approach to understand low density states of the active model}\label{sechydro} \noindent In this section we write the hydrodynamic equations of motion for the active model and characterise the low density states of the system. The equations of motion for the slow variables of the system, i.e., the number density $\wp({\bf r}, t) = \sum_{l}\delta({\bf r}-{\bf R}_l(t))$ and the order parameter $ {\Pi}_{i j}({\bf r}, t) = \wp({\bf r}, t) \Q_{ij}({\bf r}, t) = \sum_{l} ({\bf m}_{li} {\bf m}_{lj} - \frac{1}{2}\delta_{i j}) \delta({\bf r}-{\bf R}_l(t))$ are as follows \cite{shradhanjop, sradititoner}: \begin{equation} \partial_{t}\wp=a_{0}\nabla_{i}\nabla_{j}{\Pi}_{ij}+D_{\wp}\nabla^{2}\wp \label{eqdensity} \end{equation} and \begin{align} \partial_{t}{\Pi}_{ij} =\left\{ \alpha_{1}\left(\wp\right) - \alpha_{2}\left(\Pi:\Pi\right)\right\} \Pi_{ij} + \beta\left(\nabla_{i}\nabla_{j}-\frac{1}{2}\delta_{ij}\nabla^{2}\right)\wp+D_{\Pi}\nabla^{2}{\Pi}_{ij} \label{eqop} \end{align} Here ${\bf R}_l(t)$ represents position of the particle $l$, and ${\bf m}_l = (\cos \theta_l, \sin \theta_l)$ is the unit vector along the orientation $\theta_l$. The total number of particles being a conserved quantity of the system, equation (\ref{eqdensity}) represents a continuity equation $\partial_{t}\wp = -\nabla \cdot{\bf J}$ where the current $J_i = -a_0 \nabla_j {\Pi}_{ij} - D_{\wp} \nabla_i \wp$. The first term of $J_i$ consists of two parts: an anisotropic diffusion current ${\bf J}_{p1} \propto \Q_{ij}\nabla_i \wp$ and an active curvature coupling current ${\bf J}_a \propto a_0 {\wp \nabla_j \Q_{ij}}$ where $a_0$ is the activity strength of the system. The second term represents an isotropic diffusion ${\bf J}_{p2} \propto \nabla \wp$. The $\alpha$ terms in equation (\ref{eqop}) represent mean field alignment in the system. We choose $\alpha_1(\wp) = (\frac{\wp}{\wp_{IN}}-1)$ as a function of density that changes sign at some critical density $\wp_{IN}$. The $\beta$ term represents coupling with density. The last term represents diffusion in order parameter that is written under equal elastic constant approximation for two-dimensional nematic. The steady state solution $\wp({\bf r}, t) = \wp_0$ and ${\Pi}({\bf r}, t) = {\Pi}_{0}$, where ${\Pi}_0 = \sqrt{\frac{\alpha_1\left(\wp_0\right)}{\alpha_2}}$, of equations (\ref{eqdensity}) and (\ref{eqop}) represents a homogeneous ordered state for $\alpha_1(\wp_0) >0 $ at $\wp_0 > \wp_{IN}$, and a disordered isotropic state for $\alpha_1(\wp_0) < 0$ at $\wp_0 < \wp_{IN}$. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %\subsection{Linear stability of disordered isotropic state}\label{linearised} We study the linear stability of the disordered isotropic state ({\bf $\Pi_0 = 0$}) by examining the dynamics of spatially inhomogeneous fluctuations $\delta \wp({\bf r}, t) = \wp({\bf r}, t) - \wp_0$, $\delta{\Pi}_{11} = {\Pi}_{11}({\bf r}, t)$, and $\delta{\Pi}_{12} = {\Pi}_{12}({\bf r}, t)$. We obtain the linearised coupled equations of motion for small fluctuations as \begin{eqnarray} \partial_t \delta \wp &=& a_0 \left( \partial_x^2 - \partial_y^2 \right) \delta {\Pi}_{11} + 2a_0\partial_x\partial_y\delta{\Pi}_{12} + D_{\wp}\nabla^2\delta\wp \label{eq_del_rho} \\ \partial_t \delta {\Pi}_{11} &=& \alpha_1\left(\wp_0\right) \delta{\Pi}_{11} + D_{\Pi}\nabla^2\delta{\Pi}_{11} + \frac{\beta}{2}\left( \partial_x^2 - \partial_y^2 \right) \delta \wp \label{eq_del_w_11} \\ \partial_t \delta {\Pi}_{12} &=& \alpha_1\left(\wp_0\right) \delta{\Pi}_{12} + D_{\Pi}\nabla^2\delta{\Pi}_{12} + \beta\partial_x\partial_y\delta\wp \label{eq_del_w_12} \end{eqnarray} Using Fourier transformation \begin{equation} Y\left({\bf q},\lambda\right) = \int e^{i{\bf q}.{\bf r}} e^{\lambda t} Y\left({\bf r},t\right) d{\bf r} dt \end{equation} we get linear set of equations in the Fourier space as \begin{eqnarray} \lambda \left( \begin{array} {ccc} \delta\wp \\ \delta{\Pi}_{11} \\ \delta{\Pi}_{12} \end{array} \right) = M \left( \begin{array} {ccc} \delta\wp \\ \delta{\Pi}_{11} \\ \delta{\Pi}_{12} \end{array} \right) \label{eq_linear} \end{eqnarray} where $M$ is the coefficient matrix as obtained from equations (\ref{eq_del_rho}), (\ref{eq_del_w_11}), and (\ref{eq_del_w_12}) after the transformation. We solve equation (\ref{eq_linear}) for the hydrodynamic modes $\lambda$. We choose $q_x=q_y=\frac{q}{\sqrt{2}}$ since both the directions are equivalent. Therefore, we obtain \begin{equation} \left( \lambda - \alpha_1 \left( \wp_0 \right) + D_{\Pi} q^2 \right) \left\{ \left( \lambda + D_{\wp} q^2 \right) \left( \lambda - \alpha_1 \left( \wp_0 \right) + D_{\Pi} q^2 \right) - \frac{1}{2} a_0 \beta q^4 \right\} = 0 \end{equation} For small wave-vector $q$, we can find an unstable mode \begin{equation} \lambda_{+} = -2 D_{\wp} q^2 + \frac{a_0 \beta q^4}{2 |\alpha_1(\wp_0)|} - \frac{a_0 \beta q^6(D_{\Pi}-D_{\wp})}{\alpha_1^2(\wp_0)} \label{modeeq} \end{equation} For small $D_{\wp}$ and large actvitity $a_0$ this mode becomes unstable for $q < q_c$, where \begin{equation} q_c^2 = \frac{|\alpha_1(\wp_0)|}{2\Delta D} + \frac{1}{2}\sqrt{\left(\frac{|\alpha_1(\wp_0)|}{\Delta D}\right)^2 - \frac{8 D_{\wp} \alpha_1^2}{\Delta D a_0 \beta}} \label{eqq} \end{equation} provided $\Delta D = D_{\Pi}-D_{\wp}$ is positive and $a_0 \beta > 8 D_{\wp} \Delta D$. Therefore, the unstable mode $\lambda_+$ causes the I - BS transition for small diffusivity, i.e., at low temperature, and for large activity strength $a_0$. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %\subsection{Renormalised mean field study for small $S$}\label{rmf} We also calculate the jump in the scalar order parameter $S$ and the shift in the transition density from equations (\ref{eqdensity}) and (\ref{eqop}). A homogeneous steady state solution of these equations gives a mean field transition from the isotropic to the nematic state at density $\wp_{IN}$ where $\alpha_1(\wp)$ changes sign. Using renormalised mean field (RMF) method we calculate an effective free energy $\mathcal{F}_{eff}(S)$ close to the order-disorder transition where $S$ is small. We consider density fluctuations $\delta \wp$ and neglect order parameter fluctuations. The effective free energy is \begin{equation} \mathcal{F}_{eff}\left(S\right)=-\frac{b_{2}}{2}S^{2}-\frac{b_{3}}{3}S^{3}+\frac{b_{4}}{4}S^{4} \label{eqfenergy} \end{equation} where $b_{2}=\alpha_{1}(\wp)+\alpha_{1}^{\prime}c$, where $c$ is a constant. $\alpha_1^{\prime} = {\partial \alpha_1/\partial \wp|}_{\wp_0}$, $b_{3}=\frac{a_{0}\wp_0\alpha_{1}^{\prime}}{2D_{\wp}}$, and $b_{4}=\frac{1}{2}\wp_0^2\alpha_{2}$. Both $b_3$ and $b_4$ are positive. A detailed calculation for $\mathcal{F}_{eff}$ is shown in Appendix \ref{appRMF}. The density fluctuations introduce a new cubic order term in the free energy $\mathcal{F}_{eff} (S)$ that is proportional to the activity strength $a_0$. The presence of such term produces a jump $\Delta S = S_{c}=\frac{2b_{3}}{3b_{4}} $ at a density $\wp_c = \wp_{IN}(1-\frac{2b_3^2}{9b_4} ) < \wp_{IN}$. Fluctuation in density produces a jump in order parameter and shifts the critical density. Such type of fluctuation induced transitions are called fluctuation dominated first order phase transitions in statistical mechanics \cite{coleman} and are widely studied for many systems \cite{fdfopt,fdfopt2}. The jump in $S$ and the shift in the transition density are proportional to the activity strength $a_0$, and for $a_0=0$ we recover the equilibrium transition. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section*{Discussion}\label{secsummary} \noindent In our present work we have introduced a minimal lattice model for the active nematic and study different ordering states in the density-temperature plane. A brief summary of the results is as follows. In the low density regime, the system is in the disordered isotropic (I) state with short range orientation correlation amongst the particles. In the low temperature regime, large density fluctuation in the active system induces a first order transition from the isotropic to the banded state with a jump in the scalar order parameter at a density lower than the equilibrium isotropic-nematic (I-N) transition density. The linear stability analysis of the isotropic state shows an instability for large activity strength in the low temperature regime. Such instability governs the band formation at density below the equilibrium I-N transition density. As we further increase density, bands vanish and locally ordered patches appear in the inhomogeneous mixed (IM) state. Renormalised mean field calculation confirms the jump in the scalar order parameter and the shift in the transition density. With increasing temperature the shift in the transition density and the jump in scalar order parameter decreases, and no bands appear in the system. The IM state is a state with coexisting aligned and disordered domains, similar to the coexisting or defect-ordered states found in Ref. \cite{ngo, aparnaredner, shimanatcomm, yeomans, juliaarxiv, ozaarxiv, decamp}. In the high density regime, the active nematic shows switching between the IM (low $S$) and the homogeneous ordered (HO, high $S$) states, i.e., the system shows bistability. In the complete filling limit and with excluded volume assumption the active model reduces to the equilibrium model. Therefore, the active model tends to show a homogeneous nematic state in the high density regime. However, large activity strength makes the HO state unstable and leads the system to the IM state. This instability in the HO state is similar to the earlier studies in Ref. \cite{aparnamarchetti, shimanatcomm}. Ngo et al. \cite{ngo} considered a two dimensional off-lattice model for the active nematic without the exclusion constraint. In the low and moderate density regime, they show a homogeneous disordered phase and an inhomogeneous chaotic phase, which are similar to the isotropic and the IM states, respectively. Similar to their study, the spanning area of the IM state (golden regime in the phase diagram Fig. \ref{fig:phase_snap}(a)) along the density axis decreases with the increasing temperature. In the high density limit, they note a homogeneous quasi-ordered phase only, which similar to the HO state in our study. However, we show the bistability between the HO and the IM state in this density limit. In conclusion, our lattice model for the active nematic is a simple one to design and execute numerically, and easy to compare with the corresponding equilibrium model. It shows new features like the BS in the low temperature regime and the bistability in the high density regime, as well as some of the early characterised states, e.g., the IM state. It also shows many basic features of the active nematic like large number fluctuation, long-time decay of orientation correlation, transition from SRO isotropic to QLRO nematic state. The shift in the transition density due to activity strength compared to the equilibrium model can be tested in experimental systems where activity can be tuned. We expect the emergence of the bistability in the high density regime in a two dimensional experimental system composed of apolar particles with finite dimension and high activity strength. It would be interesting to study the model without volume exclusion. In this study, particle orientation has continuous symmetry of $O(2)$. Therefore, the equilibrium limit of our model is an apolar analogue of the two-dimensional XY-model. One can also study the model with discrete orientation symmetry as in Ref. \cite{solonprl, solonpre, prlperuani2012} and compare the results with the corresponding equilibrium model. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % %\section*{Additional information}\label{addinform} %\noindent {\bf Supplementary information} accompanies this paper. % %\noindent {\bf Competing financial interests:} The authors declare no competing financial interests. % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %\newpage % %\begin{titlepage}\centering %{\bf \Large Supplementary Information} %\end{titlepage} %\nopagebreak[4] % %\section*{Supplementary figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \appendix %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \renewcommand\thefigure{\thesection\arabic{figure}} \setcounter{figure}{0} \section{Order-disorder transition in the EM}\label{appEM} \begin{figure}[h] \includegraphics[width=0.4\linewidth]{FigureSI.pdf} \caption{ Order-disorder transition in the equilibrium model. Main - scalar order parameter $S$ versus inverse temperature $\beta\epsilon$ plot for different density $C$. With increasing $\beta\epsilon$ the system goes from the isotropic (small $S$) to the nematic (large $S$) state. Inset - the critical inverse temperature decreases with increasing density. } \label{figtemp} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Renormalised mean field (RMF) study of active nematic for small scalar order parameter}\label{appRMF} \noindent In this section we write an effective renormalised mean field free energy for the scalar order parameter $S$ under the small $S$ approximation. We consider the fluctuations in the density and ignore the order parameter fluctuations in the coupled hydrodynamic equations of motion for the active nematic. Density fluctuation introduces a cubic order term in $S$ in the effective free energy. Such term produces a jump in $S$ at a new transition density $\wp_c$ lower than the equilibrium I-N transition point $\wp_{IN}$. Shift in the transition density and the jump $\Delta S$ are directly proportional to the activity strength $a_0$. We recover the equilibrium limit for zero $a_0$. In the main text we write the coupled hydrodynamic equations of motion for the density $\wp$ and the order parameter ${\Pi} = \wp \Q$ where nematic order parameter \cite{pgdgenne} is defined as \begin{eqnarray} \Q({\bf r}, t)=\frac{S}{2}\left(\begin{array} {c c} \cos 2\theta ({\bf r}, t) & \sin 2\theta ({\bf r}, t) \\ \sin 2\theta ({\bf r}, t) & -\cos 2\theta ({\bf r}, t) \end{array} \right) \label{Q_tensor} \end{eqnarray} Here $\theta$ is the coarse-grained orientation at position ${\bf r}$ and time $t$. These hydrodynamic equations are previously derived in Ref. \cite{shradhanjop}, but with specific coefficients. Here we retain general coefficients. The density equation is a continuity equation $\partial \wp/\partial t = - \nabla \cdot{\bf J}$, where the current ${\bf J}$ has two parts - active and diffusive. Details of these two currents are given in the main text. The activity strength $a_0$ represents the self-propelled nature of the particles, $\beta$ is the coupling coefficient of the density in the order parameter equation, $D_{\wp}$ and $D_{\Pi}$ are the diffusion coefficients in the density and the order parameter equations, respectively. $\alpha_1(\wp)$ and $\alpha_2$ represent alignment in the system, and depend on the model parameters. For metric distance interacting model \cite{shradhanjop}, $\alpha_1(\wp)$ is a function of density and changes sign at the critical density. We choose $\alpha_1(\wp)= \frac{\wp}{\wp_{IN}}-1$ and $\alpha_2 = 1$. Let us consider a small perturbation $\delta\wp$ over the homogeneous steady state solution of the density equation so that $\wp= \wp_0 + \delta \wp$. Now from the density equation, we obtain \begin{align} & a_{0}\nabla_{i}\nabla_{j}{\Pi}_{ij}+D_{\wp}\nabla^{2}\delta \wp = 0 \notag \\ \Rightarrow \; & a_{0}\nabla_{j}{\Pi}_{ij}+D_{\wp}\nabla_{i}\delta \wp =\mathbf{c} \equiv constant \label{eqa3} \end{align} where ${\Pi}_{11} = -{\Pi}_{22} = \frac{S}{2} \cos(2 \theta)$ and ${\Pi}_{12} = {\Pi}_{21} = \frac{S}{2} \sin(2 \theta)$. Considering only the lowest order terms in $S$ and $\theta$, we obtain \begin{equation} \partial_x \delta \wp = -\frac{a_0 \wp_0}{2 D_{\wp}} \partial_x S \Rightarrow \delta \wp(x)=-\frac{a_0 \wp_0}{2 D_{\wp}} S + c_1 \label{eqa4} \end{equation} and \begin{equation} \partial_y \delta \wp = \frac{a_0 \wp_0}{2 D_{\wp}} \partial_y S \Rightarrow \delta \wp(y)=\frac{a_0 \wp_0}{2 D_{\wp}} S + c_2 \label{eqa5} \end{equation} Here we assume the system is aligned along one direction, and the variation in orientation is only along the perpendicular direction. Therefore, we can choose either of equations (\ref{eqa4}) or (\ref{eqa5}). Two constants $c_1$ and $c_2$ are the fluctuations in density when the nematic order parameter is zero. Now from the equation for ${\Pi}_{ij}$, we obtain an effective equation for $S$ as \begin{equation} \partial_{t}S=\left\{ \alpha_{1}\left(\wp\right)-\frac{\wp^2}{2}\alpha_{2}S^{2}\right\} S + \mathcal{O}(\nabla^2 S) + \mathcal{O}(\nabla^2 \wp) \label{eqa6} \end{equation} We neglect all the derivative terms and retain only the polynomials in $S$, i.e., we neglect higher order fluctuations. The Taylor expansion of $\alpha_1(\wp)$ about the mean density $\wp_0$ gives $\alpha_1(\wp)=\alpha_1(\wp_{0}+\delta \wp)=\alpha_1(\wp_{0}) + \alpha_1^{\prime} \delta \wp$ where $\alpha_1^{\prime} = \frac{\partial \alpha_1}{\partial \wp}|_{\wp_{0}}$. This gives \begin{equation} \partial_{t}S=\left\{ \alpha_{1}\left(\wp_{0}\right)+\alpha_1^{\prime} \delta \wp-\frac{\wp_0^2}{2}\alpha_{2}S^{2}\right\} S \label{eqa7} \end{equation} We can write an effective free energy $\mathcal{F}_{eff}\left(S\right)$ so that \begin{equation} \partial_t {S} = -\frac{\delta \mathcal{F}_{eff}(S)}{\delta S} \label{eqa8} \end{equation} Substituting the expression for $\delta \wp$ from equation (\ref{eqa5}), we obtain \begin{equation} -\frac{\delta \mathcal{F}_{eff}}{\delta S} =S\left\{\alpha_{1}\left(\wp_{0}\right)+\alpha_1^{\prime}\left(\frac{a_{0} \wp_0}{2D_{\wp}}S+c_{2}\right)-\frac{\wp_0^2}{2}\alpha_{2}S^{2}\right\} \end{equation} Therefore, \begin{equation} \mathcal{F}_{eff}\left(S\right)=-\frac{b_{2}}{2}S^{2}-\frac{b_{3}}{3}S^{3}+\frac{b_{4}}{4}S^{4} \label{eqa9} \end{equation} where $b_{2}=\alpha_{1}\left(\wp_{0}\right)+\alpha_1^{\prime} c_2$, $b_{3}=\frac{a_{0}\wp_0\alpha_1^{\prime} }{2D_{\wp}}$ and $b_{4}=\frac{1}{2}\wp_0^2\alpha_{2}$. Since the free energy is a state function, we have assumed the integration constant to be zero. Therefore, the fluctuation in the density introduces a cubic order term in the effective free energy $\mathcal{F}_{eff}(S)$. Effective free energy in equation (\ref{eqa9}) is similar to the Landau free energy with a new cubic order term \cite{chaiklub}. Now we calculate the jump $\Delta S$ and the new critical density from the coexistence condition for free energy. Steady state solutions of order parameter ($S=0$ for isotropic and $S \neq 0$ for ordered state) are given by \begin{equation} \frac{\delta \mathcal{F}_{eff}}{\delta S}=\left(-b_{2}-b_{3}S+b_{4}S^{2}\right)S=0 \end{equation} Non-zero $S$ is given by $-b_{2}-b_{3}S_c+b_{4}S_c^{2}=0$. Coexistence condition implies \begin{equation} \mathcal{F}_{eff}(S_c)=\left(-\frac{b_{2}}{2}-\frac{b_{3}}{3}S_c+\frac{b_{4}}{4}S_c^{2}\right)S_c^{2}=\mathcal{F}_{eff}(S=0)=0 \end{equation} Hence we get the solution \begin{equation} S_{c}=-\frac{3b_{2}}{b_{3}} \end{equation} and \begin{equation} b_{2}^{c}=-\frac{2b_{3}^{2}}{9b_{4}} \end{equation} Therefore, the jump at the new critical point is $\Delta S = \frac{2 b_3}{3 b_4}$. Since $b_4 >0$ and hence $b_2^c <0$, the new critical density \begin{equation} \wp_{c}=\wp_{IN}\left(1-\frac{2b_{3}^{2}}{9b_{4}}\right)< \wp_{IN} \label{eqrhoc} \end{equation} is shifted to a lower density in comparison to the equilibrium transition density $\wp_{IN}$. Equation (\ref{eqrhoc}) gives the expression for new transition density as given in the main text. Therefore, using renormalised mean field theory we find a jump $\Delta S$ at a lower density as compared to the equilibrium I-N transition density. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{acknowledgements} \noindent We thank Zoltan G. Soos for helpful discussions. S. M. acknowledges Thomas Niedermayer for useful discussions. S. M. and M. K. acknowledge financial support from the Department of Science and Technology, India, under INSPIRE award 2012 and Ramanujan Fellowship, respectively. \end{acknowledgements} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section*{Author contributions}\label{authorcontri} \noindent R.D., M.K. and S.M. designed the research, discussed the results and prepared the manuscript. R.D. performed simulations. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % BIBLIOGRAPHY %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % Create the reference section using BibTeX: \begin{thebibliography}{10} % \bibitem{sriramrmp} Marchetti, M. C. {\it et al.} Hydrodynamics of soft active matter. {\it Rev. Mod. Phys.} {\bf 85,} 1143-1189 (2013). % \bibitem{tonertusr} Toner, J., Tu, Y. \& Ramaswamy, S. Hydrodynamics and phases of flocks.{\it Ann. Phys.}{\bf 318,} 170-244 (2005). % \bibitem{rev} Ramaswamy, S. The mechanics and statistics of active Matter. {\it Annu. Rev. Cond. Matt. Phys.} {\bf 1,} 323-345 (2010). % \bibitem{revvicsek} Vicsek, T. \& Zafeiris, A. Collective motion.{\it Phys. Rep.}{\bf 517,} 71-140 (2012). % \bibitem{revcates} Cates, M. E. Diffusive transport without detailed balance in motile bacteria: does microbiology need statistical physics? {\it Reports on Progress in Physics} {\bf 75,} 042601 (2012). % \bibitem{harada} Harada, Y., Nogushi, A., Kishino, A. \& Yanagida, T. Sliding movement of single actin filaments on one-headed myosin filaments.{\it Nature (London)}{\bf 326,} 805-808 (1987). % \bibitem{badoual} Badoual, M., J{\"u}licher F \& Prost, J. Bidirectional cooperative motion of molecular motors.{\it Proc. Natl. Acad. Sci. U.S.A.}{\bf 99,} 6696-6701 (2002). % \bibitem{nedelec} N{\'e}d{\'e}lec, F. J., Surrey, T., Maggs, A. C. \& Leibler, S. Self-organization of microtubules and motors.{\it Nature (London)}{\bf 389,} 305-308 (1997). % \bibitem{rauch} Rauch, E. M., Millonas, M. M. \& Chialvo, D. R. Pattern formation and functionality in swarm models.{\it Phys. Lett. A}{\bf 207,} 185-193 (1995). % \bibitem{benjacob} Ben-Jacob, E. {\it et al.} Cooperative formation of chiral patterns during growth of bacterial colonies. {\it Phys. Rev. Lett.} {\bf 75,} 2899-2902 (1995). % \bibitem{animalgroups} Appleby, M. C. {\it Animal Groups in Three Dimensions} (ed. Parrish, J. K. \& Hamner, W. M.) (Cambridge: Cambridge University Press, 1997).% \bibitem{helbingnat} Helbing, D., Farkas, I. \& Vicsek, T. Simulating dynamical features of escape panic.{\it Nature}{\bf 407,} 487-490 (2000). % \bibitem{helbingprl} Helbing D., Farkas, I. J. \& Vicsek, T. Freezing by heating in a driven mesoscopic system.{\it Phys. Rev. Lett.}{\bf 84,} 1240-1243 (2000). % \bibitem{feder} Feder, T. Statistical physics is for the birds. {\it Phys. Today} {\bf 60,} 28-33 (2007). % \bibitem{feare} Feare, C. {\it The Starlings} (Oxford: Oxford University Press, 1984). % \bibitem{kuusela31} Kuusela, E., Lahtinen, J. M. \& Ala-Nissila, T. Collective effects in settling of spheroids under steady-state sedimentation.{\it Phys. Rev. Lett.}{\bf 90,} 094502 (2003). % \bibitem{hubbard} Hubbard, S., Babak, P., Sigurdsson, S. \& Magnusson, K. A model of the formation of fish schools and migrations of fish.{\it Ecological Modeling}{\bf 174,} 359-374 (2004). % \bibitem{naturevschaller} Schaller, V., Weber, C., Semmrich, C., Frey, E. \& Bausch, A. R. Polar patterns of driven filaments.{\it Nature}{\bf 467,} 73-77 (2010). % \bibitem{natureysumino} Sumino, Y. {\it et al.} Large-scale vortex lattice emerging from collectively moving microtubules. {\it Nature} {\bf 483,} 448-452 (2012). % \bibitem{prlperuani2012} Peruani, F. {\it et al.} Collective motion and nonequilibrium cluster formation in colonies of gliding bacteria. {\it Phys. Rev. Lett.} {\bf 108,} 098102 (2012). % \bibitem{vnarayanjstat} Narayan, V., Menon, N. \& Ramaswamy, S. Nonequilibrium steady states in a vibrated-rod monolayer: tetratic, nematic, and smectic correlations.{\it J. Stat. Mech.} P01005 (2006). % \bibitem{vnarayanscince} Narayan, V., Ramaswamy, S. \& Menon, N. Long-lived giant number fluctuations in a swarming granular nematic.{\it Science}{\bf 317,} 105-108 (2007). % \bibitem{kudrolli} Blair, D. L., Neicu, T. \& Kudrolli, A. Vortices in vibrated granular rods.{\it Phys. Rev. E}{\bf67,} 031303 (2003). % \bibitem{ncommsriram} Kumar, N., Soni. H., Ramaswamy, S. \& Sood, A. K. Flocking at a distance in active granular matter.{\it Nat. Comm.}{\bf 5,} 4688 (2014). % \bibitem{polardisks} Deseigne, J., Dauchot, O. \& Chat\'{e}, H. Collective motion of vibrated polar disks. {\it Phys. Rev. Lett.} {\bf 105,} 098001 (2010). % \bibitem{activecolloid1} Klopper, A. Active colloids: Made to order. {\it Nat. Phys.} {\bf 11,} 703 (2015). % \bibitem{activecolloid2} Cates, M. E. \& Tailleur, J. Motility-induced phase separation.{\it Ann. Rev. Cond. Matt. Phys.}{\bf 6,} 219-244 (2015). % \bibitem{activecolloidantoine} Bricard, A., Caussin, J. B., Desreumaux, N., Dauchot, O. \& Bartolo, D. Emergence of macroscopic directed motion in populations of motile colloids.{\it Nature}{\bf 503,} 95-98 (2013). % \bibitem{activecolloidTheurkauff} Theurkauff, I., Cottin-Bizonne, C., Palacci, J., Ybert, C. \& Bocquet, L. Dynamic clustering in active colloidal suspensions with chemical signaling.{\it Phys. Rev. Lett.}{\bf 108,} 268303 (2012). % \bibitem{sradititoner} Ramaswamy, S., Simha, R. A. \& Toner, J. Active nematics on a substrate: Giant number fluctuations and long-time tails.{\it Europhys. Lett.}{\bf 62,} 196-202 (2003). % \bibitem{pgdgenne} de Gennes, P. G. \& Prost, J.{\it The Physics of Liquid Crystals} (Oxford: Clarendon Press, 1995). % \bibitem{shradhanjop} Bertin, E. {\it et al.} Mesoscopic theory for fluctuating active nematics. {\it New J. of Phys.} {\bf 15,} 085032 (2013). % \bibitem{ngo} Ngo, S. {\it et al.} Large-scale chaos and fluctuations in active nematics. {\it Phys. Rev. Lett.} {\bf 113,} 038302 (2014). % \bibitem{shimanatcomm} Shi Xia-qing \& Ma Yu-qiang, Topological structure dynamics revealing collective evolution in active nematics.{\it Nat. Comm.}{\bf 4,} 3013 (2013). % \bibitem{solonprl} Solon, A. P. \& Tailleur, J. Revisiting the flocking transition using active spins.{\it Phys. Rev. Lett.}{\bf 111,} 078101 (2013). % \bibitem{solonpre}Solon, A. P. \& Tailleur, J. Flocking with discrete symmetry: The two-dimensional active Ising model.{\it Phys. Rev. E}{\bf 92,} 042119 (2015). % \bibitem{prlperuani2011} Peruani, F., Klauss, T., Deutsch, A. and Voss-Boechme, A. Traffic jams, gliders, and bands in the quest for collective motion of self-propelled Particle. {\it Phys. Rev. Lett.} {\bf 106,} 128101 (2011). % \bibitem{dimprl} Farrell, F. D. C., Marchetti, M. C., Marenduzzo, D. \& Tailleur, J. Pattern formation in self-propelled particles with density-dependent motility.{\it Phys. Rev. Lett.}{\bf 108,} 248101 (2012). % \bibitem{llasher} Lebwohl, A. \& Lasher, G. Nematic-liquid-crystal order - a Monte Carlo calculation.{\it Phys. Rev. A}{\bf 6,} 426-429 (1972). % \bibitem{mcbinder} Landau, D. P. \& Binder, K.{\it A Guide to Monte Carlo Simulations in Statistical Physics} (Cambridge: Cambridge University Press, 2005). % \bibitem{HohenbergHalperin} Hohenberg, P. C. \& Halperin, B. I. Theory of dynamic critical phenomena.{\it Rev. Mod. Phys.}{\bf 49,} 435-479 (1977). % \bibitem{dilutedxymodel} Leonel, S. A., Coura, P. Z., Pereira, A. R., M\'{o}l, L. A. S. \& Costa, B. V. Monte Carlo study of the critical temperature for the planar rotator model with nonmagnetic impurities.{\it Phys. Rev. B}{\bf 67,} 104426 (2003). % \bibitem{kemkemer} Kemkemer, R., Kling, D., Kaufmann, D. \& Gruler, H. Elastic properties of nematoid arrangements formed by amoeboid cells.{\it Eur. Phys. J. E}{\bf 1,} 215-225 (1999). % \bibitem{chaiklub} Chaikin, P. M. \& Lubensky, T. C.{\it Principles of Condensed Matter Physics} (Cambridge: Cambridge University Press, 2000). % \bibitem{mondalroy} Mondal, E. \& Roy, S. K. Finite size scaling in the planar Lebwohl-Lasher model.{\it Phys. Lett. A}{\bf 312,} 397-410 (2003). % \bibitem{coleman} Coleman, S. \& Weinberg, E. Radiative corrections as the origin of spontaneous symmetry breaking.{\it Phys. Rev. D}{\bf 7,} 1888-1910 (1973). % \bibitem{fdfopt} Halperin, B. I., Lubensky, T. C. \& Ma, S. K. First-order phase transitions in superconductors and smectic-A liquid crystals.{\it Phys. Rev. Lett.}{\bf 32,} 292-295 (1974). % \bibitem{fdfopt2}Chen, J. H., Lubensky, T. C. \& Nelson, D. R. Crossover near fluctuation-induced first-order phase transitions in superconductors.{\it Phys. Rev. B}{\bf 17,} 4274-4286 (1978). % \bibitem{yeomans} Thampi, S. P., Golestanian, R. \& Yeomans, J. M. Instabilities and topological defects in active nematics.{\it Europhys. Lett.}{\bf 105,} 18001 (2014). % \bibitem{juliaarxiv} Doostmohammadi, A., Adamer, M., Thampi, S. P. \& Yeomans, J. M. Stabilization of active matter by flow-vortex lattices and defect ordering. arXiv:1505.04199 (2015).% \bibitem{ozaarxiv} Oza, A. U. \& Dunkel. J. Antipolar ordering of topological defects in active liquid crystals. arXiv:1507.01055 (2015).% \bibitem{decamp} DeCamp, S. J., Redner, G. S., Baskaran, A., Hagan, M. F. \& Dogic, Z. Orientational order of motile defects in active nematics.{\it Nat. Mat.}{\bf 14,} 1110-1115 (2015). % \bibitem{aparnaredner} Putzig, E., Redner, G. S., Baskaran, A. \& Baskaran, A. Instabilities, defects, and defect ordering in an overdamped active nematic. arXiv:1506.03501 (2015).% \bibitem{aparnamarchetti} Baskaran, A. \& Marchetti, M. C. Hydrodynamics of self-propelled hard rods.{\it Phys. Rev. E}{\bf 77,} 011920 (2008). % \end{thebibliography} % %\begin{thebibliography}{10} %% %\bibitem{pgdgenne} de Gennes, P. G. \& Prost, J. {\it The Physics of Liquid Crystals} (Oxford: Clarendon Press, 1995). %% %\bibitem{shradhanjop} Bertin, E. {\it et al.} Mesoscopic theory for fluctuating active nematics. {\it New J. of Phys.} {\bf 15,} 085032 (2013). %% %\bibitem{chaiklub} Chaikin, P. M. \& Lubensky, T. C. {\it Principles of Condensed Matter Physics} (Cambridge: Cambridge University Press, 2000). %% %\end{thebibliography} % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \end{document} % % ****** End of file apstemplate.tex ****** }}
\caption{Example document generated by the Conditional Copy system with a beam of size 5. Text that accurately reflects a record in the associated box- or line-score is highlighted in \textcolor{blue}{blue}, and erroneous text is highlighted in \textcolor{red}{red}.}
\caption{Hammer-Aitoff projection, in Galactic coordinates, of the {\agile} \gray exposure in [$\mathrm{cm^2~s~sr}$] (bin size of 0.5$^\circ$) after one complete rotation in spinning mode, time-centered at the ICECUBE-160731 event time $T_0$. The neutrino event error circle is shown in black. The magenta and yellow contours show, respectively, the Sun/anti-Sun exclusion regions and the average Earth occultation during the considered integration time: ($T_0$-210; $T_0$+210)~s.}
\caption{Spectral energy distribution (SED) of the possible HBL candidate, the faint SDSS J141658.90-001442.5 galaxy, found within the ICECUBE-160731 error circle. The galaxy appears within the 25" error circle of the RASS source 1RXS J141658.0-001449 ($\nu~F_{\nu}$ value shown as black point in the SED), along with a FIRST 2~mJy radio source (red point). Optical and IR data of the SDSS J141658.90 galaxy are from: Sloan Digital Sky Survey (SDSS) -- Release \#7 and\#13 (blue points,\cite{2009ApJS..182..543A, 2016arXiv160802013S}); Catalina Real-Time Transient Survey (CRTS) (magenta points, \cite{2009ApJ...696..870D}); VIKING survey (green points, \cite{2013Msngr.154...32E}); AllWISE Data Release (purple points, \cite{2014yCat.2328....0C}).}
\caption{\bf{\agile} and FERMI-LAT exposures on the ICECUBE-160731 error circle during the period of the detection of the possible \gray precursor AGL J1418+0008. For both instruments, a maximum off-axis angle of 50$^\circ$ between source and FoV center has been assumed.}
\caption{\small Discovering a cutset in AEBP. $Q$ requests a bound on $Q$ from $\phi_1$ while telling it that no factors have been observed (in \textcolor{red}{red}) in other components (empty set \textcolor{red}{$\{\}$}) (a). The bound is computing by only assuming that the message from $A$ is the simplex $\mP(A)$. It then requests a bound from $\phi_2$, this time telling it about $\phi_1$ having already been observed in a different component (b). In (c), it requests a better bound from $\phi_1$, this time telling it about having $\phi_2$, which triggers a request from $\phi_3$ to $A$ with information of having previously observed $\{\phi_1,\phi_2\}$. In (d), a request goes all the way to $B$ with information having observed, among others, $\phi_3$. When $B$ requests a bound from $\phi_4$ (e), it is detected that $\phi_4$'s argument $C$ is also an argument of previously observed factor $\phi_3$, which leads to making $C$ a cutset variable. Bounds are then computed as a \emph{function} of $C$, as opposed of summing it out, all the way back to $Q$, where $C$ will eventually be summed out. }
\caption{Setup for DFG-based OAM down-conversion. LD: diode laser (Toptica, predesigned ,1520 nm-1590 nm); EDFA: Er-doped fiber amplifier (1540 nm-1560 nm); Ti: sapphire laser (Coherent, MBR110); HWP: half-wave plate; QWP: quarter-wave plate; PBS\_B(C): polarizing beam splitter; VPP: vortex phase plate; Filter: long pass filter; SLM: infrared spatial light modulator; P\_M: power meter; F\_M: fiber power meter; PPKTP: (Raycol QPM crystals ; period of 9.375$\mu m$); PPLN: {\color{Revise2}{(HCP, periodically poled lithium niobate (PPLN) chip SFVIS-MA; period of 7.3 $\mu m$)}}.}
\caption{{\color{Revise2}{Single pass conversion efficiency (SPCE) and quantum conversion efficiency (QCE)}} of the OAM frequency bridge. (a) SPCE of the OAM state. The green bar and the red error bars represent the experimental and error values, respectively. The input fields are the Gaussian field ($\ket{0}$), a pure LG state ($\ket{1},\ket{2}$), and the supposition state ($\ket{1}+\ket{-1}$, $\ket{2}+\ket{-2}$). (b) QCE for conversion from visible to infrared for L=0, 1, and 2 in the ring cavity. The x-axis represents the measured intra-cavity pump power. }
\caption{Density matrix and intensity profiles. (a) Input state and density matrix of the output state. The input row describes the intensity and the phase of the input fields; the lists of $Re(\rho)$ and $Im(\rho)$ are the real and imaginary parts of the density matrix for the output fields determined using quantum state tomography. (b) Intensity profiles of the input {\color{Revise2}{visible laser (VL) and the output infrared laser (IL)}} beams determined using visible and infrared CCDs.}
\caption{\label{tab:explanation} Example causal chains for explaining the rise ($\uparrow$) and fall ($\downarrow$) of companies' stock price. The temporally causal {\color{Sepia}$feature$} and {\color{BlueViolet}$target$} are linked through a sequence of predicted cause-effect tuples by different reasoning algorithms: a symbolic graph traverse algorithm \textit{SYMB} and a neural causality reasoning model \textit{NEUR}.}
\caption{$e_{1}\left(n\right)$ \eqref{eq:e1n} for $\boldsymbol{\Sigma}_{AR}$ \eqref{eq:AR} of AR(1) process with $p=50$ and $r=0.5$. } \par\end{centering} \begin{centering} \label{icml-historical-2-1} \par\end{centering} \vskip -0.2in \end{figure} \begin{figure}[h] \vskip 0.2in \begin{centering} \centerline{\includegraphics[scale=0.04]{lance7new}} \protect\protect\protect\protect\caption{$e_{2}\left(n\right)$ \eqref{eq:e2n} for $\boldsymbol{\Sigma}_{AR}$ \eqref{eq:AR} of AR(1) process with $p=50$ and $r=0.5$. }
\caption{Schematic illustration of the proposed algorithm for maximizing a convex quadratic objective over a convex polytope, by successively maximizing a linear sub-estimate of it. The \textcolor{red}{\textit{hot}} color map encodes the function values. The \textcolor{blue}{\textit{j}}\textcolor{green}{\textit{e}}\textcolor{red}{\textit{t}} color map encodes the values of the linear sub-estimate. The point around which the objective is linearized is depicted in red. The global maximum is depicted in blue. The maximum of the linear sub-estimate is depicted in green. Notice that the algorithm travels between extreme but not necessarily adjacent points of the polytope, until it converges to a local maximum. \label{fig:FrankWolfe}}
\caption{The rotational diffusion and arm-disentanglement relaxation times for our star model, with $12$ arms and $6$ monomers per arm. We define $\wi_{rot}=\dot{\gamma}\tau_{rot}$ and $\wi=\dot{\gamma}\tau_{rel}$ with $\tau_{rel}=\max\left[\tau_{rot},\tau_{dis}\right]$.} \begin{tabular}{ccc} \centering System & $\tau_{rot}$ & $\tau_{dis}$ \\ \hline closed melt: $\gamma _{\parallel} = 1.0$, $\gamma _{\perp} = 1.0$ & $710 \pm 40$ & $390 \pm 10$\\ open melt: $\gamma _{\parallel} = 1.0$, $\gamma _{\perp} = 1.0$ & $700 \pm 40$ & $390 \pm 10$\\ solution harmonic bonds & $270 \pm 20$ & $180 \pm 20$\\ solution FENE bonds & $370 \pm 30$ & $950 \pm 90$\\ \end{tabular} \label{tab:relax. time} \end{table} \subsection{Star polymer models} The star polymer model is taken from Ref.~\cite{Hijon_2010}. We use the standard Lennard-Jones units, taking the monomer mass $m_{0}$, unit length $\sigma_0$ and energy $\epsilon_0$ as reference. We consider stars with $f=12$ arms an $m=6$ beads per arm, with a total of $73$ monomers (including the central one). Excluded volume interactions of monomers are modeled by the repulsive Weeks-Chandler-Anderson interaction ($\sigma = 2.415$ and $\epsilon = 1$). The bonds between adjacent monomers $i$ and $j$ are modelled by either harmonic springs or FENE bonds. In the case of harmonic bonds, with a recovery force $-K(r_{ij}-r_{ij}^{eq})$, the spring constant is $K = 20$ and the equilibrium distance $r_{ij}^{eq} = 2.77$ (the equilibrium distance between the central monomer and the first monomer of an arm is larger $r_{ij}^{eq} = 3.9$). Finitely extensible bonds are modeled by the FENE potential \cite{Veldhorst:2015}, with a spring constant $K=20$ and maximum length of the bond $r_{max} = 1.5 \, r_{ij}^{eq}$. \subsection{Melt simulations} For the melt case, we use stars made of harmonic bonds. Simulations are carried out at fixed temperature $T=4$ using molecular dynamics with a dissipative particle dynamic (DPD) thermostat \cite{DPD_Espanol,DPD_Soddemann}. We solve systems with constant volume (closed setup) and also open systems under constant normal load (see Ref. for details \cite{Sablic:2016}). The simulation box is of size $390 \times 117 \times 117$ and the density of the melt in equilibrium corresponds to the occupational factor $\mathit{\Phi} = 0.2$, with about $2000$ molecules. In the closed periodic setup, the shear flow is imposed by the SLLOD algorithm implemented with the Lees-Edwards boundary conditions~\cite{LeesEdwards, Sllod1, Sllod2}. Constant load simulations in an open system, are performed using OBMD~\cite{hmd_prl06,Tools,Flekbus,DelgadoBuscalioni:2015}, which permits to impose an external shear stress at the open ends of the system. We shall use the following coordinates: $x_{1}$ refers to the flow direction, $x_{2}$ to the direction of the velocity gradient and $x_{3}$ to the direction of flow vorticity (sometimes called neutral direction). The DPD thermostat used here introduces friction along the {\em normal} and {\em tangential} directions of any pair of monomers \cite{Sablic:2016,Hijon_2010,DPD_Espanol,DPD_Soddemann} which come closer than the DPD-cutoff radius $R_{DPD} = 2 \times 2^{1/6} \sigma$ (we use a Heaviside kernel for the DPD interaction). The friction coefficients in normal and tangential directions equal $\gamma _{\parallel} = 1.0$ and $\gamma _{\perp} = 1.0$. The equations of motion are integrated by the Velocity-Verlet algorithm~\cite{Tuckerman:2010} with the integration step $0.01 \tau$ for small and moderate shear, and $0.005 \tau$ for high shear rates. A sketch of the star-polymer melt under shear flow from the perspective of one of its constituent polymers is depicted in Fig.~\ref{fig:melt_polymer_view}. \begin{figure} \includegraphics[scale=1.1]{Figure2} \caption{Snapshot of the star-polymer melt under shear flow, drawn from the perspective of one polymer. The latter is depicted in purple and its surrounding polymers are colored in gray. The blue arrows correspond to the direction of the imposed shear, while the black arrows indicate the tank-treading rotation of the polymer. The coordinate unit vectors $\mathbf{e}_{1}$, $\mathbf{e}_{2}$, and $\mathbf{e}_{3}$ define the flow ($x_{1}$), the gradient ($x_{2}$), and the neutral ($x_{3}$) direction, respectively.} \label{fig:melt_polymer_view} \end{figure} \subsection{Star in solution} We simulate a single star polymer in solution using Brownian hydrodynamics \cite{Jendrejack:2000, Jendrejack:2002}. The monomers (representing a coarse description of the molecule) interact via conservative forces (bonds and excluded volume interactions) and also via hydrodynamic interactions. The displacement of monomer $\alpha$ in direction $i$ over time $dt$ has the form $d r_i^{\alpha} = \dot{\gamma} x_2^\alpha \delta^{i,1} dt + \mu_{ij}^{\alpha\beta}(r_{\alpha\beta}) F_j^{\beta} dt + d\tilde{r}_i^{\alpha}$ where the first term indicates the shear flow (acting in $1$-direction) and the mutual drag arises from the mobility tensor $\mu_{ij}^{\alpha\beta}$ which in present calculations consists on the Rotne-Prager-Yamakawa (RPY) approximation~\cite{Jendrejack:2000, Jendrejack:2002}. The Brownian displacement $d\tilde{\bf{r}}$ satisfies a fluctuating dissipation (FD) relation for its covariance $\langle d \tilde{r}_i^{\alpha} d \tilde{r}_j^{\beta}\rangle = 2 k_BT \mu_{ij}^{\alpha\beta} dt$ and to solve $d\tilde{\bf{r}}$ we use the Fixman's method~\cite{Jendrejack:2002}. The integration scheme is an explicit Euler scheme with the time step $dt=0.01 \tau$. In the present study, all simulations are run for $10000 \tau$. \subsection{Monomer rotation dynamics} To provide direct connection with the monomer dynamics, we calculate the angular velocity of rotation of molecules from the autocorrelation function of the gradient-direction coordinate of the last monomer of every arm of the star, relative to the CoM, $({\bf r}_{\alpha}-{\bf r}_{cm})\cdot \hat{\bf x}_2$. This signal is similar to an underdamped oscillator (Fig.~\ref{fig:omega_d_fit}) which can be fitted with the following function~\cite{Usabiaga:2011}: \begin{equation} C\left(t\right) = A^{2} \cos\left(\omegar t + \psi\right)\exp\left(-\Gamma t\right), \label{eqn:tumbling_fit_function} \end{equation} where the damping rate $\Gamma$ represents the decorrelation rate, $\omegar$ the rotation frequency, and $\psi$ a phase constant. Two issues are noticeable from this graph: first, the decorrelation rate $\Gamma$ only starts to significantly increase above $\wi_{rot}>50$. Second, as $\wi$ increases, the quality factor $q=\omegar/\Gamma$ becomes quite large, in particular, compared with what happens in linear polymers under shear \cite{Usabiaga:2011} (which tumble by compressing, like in a tube). Figure \ref{fig:omega_d_fit} (bottom panel) compares the quality factor $q$ for star polymers in solution (S) and melt (M) (with either FENE or harmonic bonds) and that measured in Ref.\cite{Usabiaga:2011} for FENE linear chains with $N=60$ and dumbbells. In the case of star molecules, the arms rotate almost like in a ``wheel'' and a monomer turns around several times ($q$) before decorrelating its initial ``rigid-body'' position. At large shear rates, the differences in values of $q$ are significant [see Fig. \ref{fig:omega_d_fit} (bottom)]. The quality factor is significantly smaller in melts, indicating the hindrance arising from steric interaction amongst close-by molecules. In what follows, we will compare $\omegar$ with $\omegal$ and $\omegae$ and discuss the origin of the decorrelation $\Gamma$, according to the Eckart analysis. \begin{figure} \centering \includegraphics[scale=0.44]{Figure3a}\\ \includegraphics[scale=0.5]{Figure3b} \caption{(Top) Autocorrelation function of position of the final monomers of each polymer's arm in the gradient direction, fitted by Eq.~\ref{eqn:tumbling_fit_function} with parameters: $A = 0.93$, $\omegar = 0.35$, $\Gamma = 0.0084$, and $\Psi = 0.0025$. (Middle panels) The tank-treading frequency $\omegar$ and decorrelation rate $\Gamma$ obtained from the fits (stars in solution and in melt). (Bottom) The quality factor $q=\omegar/\Gamma$ for the dynamics of monomer rotations, comparing our 12-6 stars with linear FENE chains ($N=60$) (with excluded volume interactions) and dumbbells, from Ref.\cite{Usabiaga:2011}.} \label{fig:omega_d_fit} \end{figure} \subsection{Kinetic energies} The kinetic energy balance is illustrated in Table~\ref{tab:kin} for star molecules in solution (having harmonic or FENE bonds) and some values of the shear rate. Displacements describing pure rotations have kinetic energy $T_{\Omega}$ but, coherent (collective) vibrations without angular momentum contribute with the largest energy $T_{\tilde v}$. These are related to overall shape deformations (and in particular, compression/expansion does not introduce angular momentum). Other type of molecular deformations (affine or not) are collected in the velocity ${\bf u}_{\alpha}$ which does provide angular momentum (see Eq. \ref{eqn:diff velocity Eckart}) and feeds the (negative) kinetic energy contribution $T_u=T_{vib-non-ang}+T_{Cori}$ (see Table \ref{tab:kin} and Eq. \ref{eq:kin2}). Equation \ref{eq:tu} confirms that this energy can only be negative because of the Coriolis term. So, in average, ${\bf u}\cdot \left(\Omega\times \delta{\bf r}\right)<0$; in other words ${\bf u}$ contributes in opposite direction to the pure rotation velocity $\Omega\times \delta{\bf r}$. Note that $|T_{u}|$ is subtracted to the pure rotation energy $T_{\Omega}$ to yield the total rotation kinetic energy in the lab frame $T_{rot}^{lab}$ (see Eq. \ref{eq:kin3}). To clarify matters, a sketch illustrating the different types of displacements is drawn in Fig. \ref{fig:coordinate_system_sketch} (bottom panel). In what follows, we analyze these kinetic energies separately. \begin{table*} \small \caption{\ Kinetic energy balance for solution of star polymers with harmonic and FENE bonds. The error bar of the reported values is approximately $5 \%$.} \label{tab:kin} \begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lllll|lllll} \hline \multicolumn{5}{c|}{solution harmonic bonds} & \multicolumn{5}{c}{solution FENE bonds} \\ \hline $W_{i}$ & $T$ & $T_{\Omega}$ & $T_{\tilde{v}}$ & $T_{u}$ & $W_{i}$ & $T$ & $T_{\Omega}$ & $T_{\tilde{v}}$ & $T_{u}$\\ \hline 13.25 & 1102 & 412 & 1028 & -338 & 9.5 & 875 & 303 & 794 & -222\\ 53 & 1117 & 517 & 1042 & -442 & 95 & 876 & 365 & 793 & -282\\ 106 & 1135 & 464 & 1058 & -387 & 570 & 912 & 498 & 808 & -394\\ 424 & 1363 & 1189 & 1275 & -1101 & 1520 & 1086 & 763 & 919 & -596\\ \hline \end{tabular*} \end{table*} \subsection{Pure rotation: tank-treading} Figure~\ref{fig:rotation_Eckart_vs_standard} compares the results for the apparent angular velocity $\omegal$, the Eckart angular velocity $\omegae$ and the frequency of monomers rotation about the CoM $\omegar$. In all considered cases (polymers with either harmonic or FENE bonds in solution, and the melt case), we find that $\omegae=\omegar$ within error bars while $\omegae>\omegal$. Whenever vibrational angular momentum is present, the apparent angular velocity $\omegal$ does not correctly represent molecular rotation \cite{Rhee:1997}. The difference between $\omegal$ and $\omegae$ is larger for stars with Hookean-bonds in solution (see Fig. \ref{fig:rotation_Eckart_vs_standard}). From Eq. \ref{eqn:diff velocity Eckart}, this simply indicates that vibrational angular momentum ($\mathbf{u}_{\alpha}$) has a larger contribution if the molecule is softer (harmonic versus FENE bonds) or has more free space to deform (as in the case of solution compared to melt). Stars with harmonic bonds in solution seems to reach the scaling $\omegal/\dot{\gamma} \sim \wi^{-1}$ (i.e. $\omegal \rightarrow \mathrm{cte}$) as the shear rate is increased (although, in fact, at very large $\dot{\gamma}$, $\omega$ decreases). This apparent scaling was attributed in Ref. \cite{Ripoll:2006} (and subsequent citations) to an universal limiting trend for tank-treading rotation of star polymers. However, although the apparent angular velocity $\omega$ reaches a maximum value, the tank-treading frequency $\omegar$, keeps increasing with $\dot{\gamma}$, like $\omegar\sim \wi^{\alpha}$ with $\alpha=0.5\pm 0.02$. This is shown in Fig.~\ref{fig:rotation_Eckart_vs_standard}) where one can see that $\omega$ and $\omegar$ differ significantly. Finally, in melts, (bottom panel of Fig.~\ref{fig:rotation_Eckart_vs_standard}) we observe that the molecular rotational frequencies are similar in the open and closed environments. This is in agreement with our previous studies (Ref.~\cite{Sablic:2016, Sablic:2016:1}) and indicates that the rheological differences measured in open and closed environments are of thermodynamic origin (density decreases when an open polymer enclosure is sheared). \begin{figure} \includegraphics[width=0.43\columnwidth]{Figure4} %\includegraphics[width=0.5\columnwidth]{Figure4b} %\includegraphics[width=0.5\columnwidth]{Figure4c} \caption{Comparison of the angular velocity computation by the Eckart frame formalism ($\omegae/\dot{\gamma}$ - colored in orange), by the standard approach ($\omegal/\dot{\gamma}$ - colored in magenta), and by the autocorrelation function of the position of the final monomers in every polymer's arm in the gradient direction ($\omegar/\dot{\gamma}$ - colored in green). We study rotations in the solution of star polymers with Hookean (Top) and FENE bonds (Middle) and in the melt of star polymers with Hookean springs (Bottom). In all three systems, the angular velocity obtained by the Eckart frame formalism is higher than the one calculated by the standard approach. In all cases, $\omegar$ matches well with $\omegae$ while the difference $\omegal-\omegar$ is larger in the Hookean spring solution case, followed by the solution of molecules with the FENE bonds and the melt. The reasons for these facts are explained in the text.} \label{fig:rotation_Eckart_vs_standard} \end{figure} \subsection{Vibrational angular momentum and decoherence of rotational motion} Following this line, Eq. \ref{eq:kin3} indicates that the total kinetic energy coming from displacements with angular momentum can be decomposed in a pure rotational part $T_{\Omega}$ and contributions from vibrational angular momentum. It is noted that $T_{\Omega}$ contains contributions from collective displacements and also from fluctuations. Equation \ref{eq:trot} indicates that \begin{equation} T_{\Omega} = \frac{N}{2} \omegae_3^2 \left(G_{11}+G_{22}\right) + \tilde{T}_{rot}, \end{equation} where $\tilde{T}_{rot}$ introduces a significant contribution from the covariances involving zero-average components of the rotational frequency $\boldsymbol{\Omega}$, like $\tilde{T}_{rot} =\langle N \omegae_1^{2} G_{22}\rangle + ...$. Here, $G_{ii}$ represents the diagonal gyration tensor component in the $i$-th direction. The energy of vibrations with angular momentum corresponds to deformations of the arms away from pure rigid body rotation (see Eq. \ref{eqn:diff velocity Eckart}). In solution, these motions arise from Brownian diffusion so we expect that the kinetic energy $|T_u|$ is proportional to $\Gamma D_{arm}$ where $D_{arm}$ is the diffusion coefficient of the center of mass of one star's arm (which is independent on the shear rate). The scaling this hypothesis predicts is validated in Fig. \ref{fig:dis} where $|T_u|$ (normalized with its value at zero shear rate) is compared with $\Gamma\tau_{rot}$ for increasing Weissenberg number. Results for different types of star polymers (harmonic and FENE bonds) confirm that both magnitudes are proportional and indicate that our intuition contains physical insight. In melts, however, both quantities differ significantly (see Fig. \ref{fig:dis} bottom panel) indicating that, in this case, molecular deformations are also determined by other (non-Brownian) mechanisms, like inter-molecular collisions. \subsection{Vibrations without angular momentum and breathing mode} As stated (see Table \ref{tab:kin}), vibrations without angular momentum $T_{\tilde v}$ have the largest contribution to the kinetic energy of the star molecule. This kinetic energy has also a thermal and a coherent contribution. The thermal energy includes the fluctuations in bond length, whose average kinetic energy scales like $N_{sp} K \langle \delta^2\rangle $, with $\delta=r_{\alpha,\beta}-r_{eq}$ the bond length, $N_{sp}=72$ the number of springs in our star molecules and $K$ their spring constant. Excluded volume forces are also central forces ${\bf F}_{\alpha,\beta}\propto{\hat{\bf r}}_{\alpha,\beta}$ (${\hat{\bf r}}_{\alpha,\beta}$ being the unit distance vector between monomers $\alpha$ and $\beta$) so in absence of hydrodynamic interactions they strictly do not contribute to the total angular momentum. It is noted that hydrodynamics spreads over internal forces, and contributes to the angular momentum, with monomer displacements $d{\bf r}_{\alpha}= \boldsymbol{\mu}_{\alpha\beta}{\bf F}_{\beta}\,dt$, where $\boldsymbol{\mu_{\alpha\beta}}$ is the mobility tensor. However, as shown in Ref. \cite{Sablic:2016:1}, the major source of angular momentum comes out from the mean flow. We assume that thermal contribution to $T_{\tilde v}$ is independent on the shear rate. The remaining contribution to $T_{\tilde v}$ is assumed to be associated to overall deformations of the molecular shape and should increase with $\dot{\gamma}$. This separation between thermal and coherent vibrations is clearly revealed in the fit $T_{\tilde v}(\wi)= T_{\tilde v}(0) + \Delta T_{\tilde v}(\wi)$, which is shown in Fig. \ref{fig:vib}, with $T_{\tilde v}(0) = 1029\pm 5$ and $\Delta T_{\tilde{v}}(\wi) = 0.021\,\wi^{1.54}$ for harmonic springs and while $T_{\tilde v}(0) = 792\pm 2$ and $\Delta T_{\tilde v}(\wi) = 2.32\times 10^{-5} \wi^{2.12}$ for FENE bonds (both in solution). In the case of melts we find $T_{\tilde v}(0)=426$ and $\Delta T_{\tilde v}(\wi) =4.60\times10^{-4} \wi^{2.00}$. We expect that the coherent part of the vibrational energy $\Delta T_{\tilde v}(\wi)$ comes out from a collective ``oscillation'' of the molecule shape. Such type of collective vibration was discussed in a previous work on star polymers \cite{Sablic:2016:1}, and was referred to as ``breathing mode''. The dynamics of the breathing mode is revealed in the time correlation of the components of the gyration tensor ($G_{ij}$), given by~\cite{tumbling, Chen:2013, Sablic:2016:1}, \begin{equation} C_{ij}\left(t\right) = \frac{\langle \delta G_{ii}\left(t_{0}\right) \delta G_{jj}\left(t_{0} + t\right)\rangle}{\sqrt{\langle \delta G_{ii}^{j}\left(t_{0}\right)\rangle \langle \delta G_{jj}^{2}\left(t_{0}\right)\rangle}}. \label{eqn:tumbling cross-corr.} \end{equation} where $\delta G_{ii} = G_{ii} - \langle G_{ii} \rangle$. These are damped oscillatory signals with a characteristic frequency $\omegat$. In previous works \cite{Chen:2013, Sablic:2016:1}, the cross-correlation $C_{12}$ has been used to extract the ``tumbling'' time $\tau _{t}$ (as twice the difference between first maximum and first minimum). We define $\omegat= 2\pi/\tau_t$. As explained in Ref. \cite{Sablic:2016:1}, these type of dynamics have been called ``tumbling'' in linear and ring chains, while the word ``breathing'' is more appropriate to describe the star overall shape oscillation, while they perform tank-treading. The energy of ``breathing'' can be estimated from the largest fluctuation in the gyration tensor, taken from the standard deviation of the principal eigenvalue of the gyration tensor ${\bf G}$, i.e. $\mathrm{Std}[G_{1}]= \langle(G_{1}-\langle G_{1}\rangle)^2 \rangle^{1/2}$. A rough estimation of the breathing kinetic energy is then, $T_{B} \equiv \frac{N}{2} \omegat^2 \mathrm{Std}[G_{1}]$, and it is compared with $\Delta T_{\tilde v}$ in Fig. \ref{fig:vib}. In passing, we note that a quite similar outcome is obtained by $T_B\propto \omegat^2 \mathrm{Std}[V^{2/3}]$ which is based on fluctuations (expansion/contraction) of the overall molecular volume $V=\prod_{\alpha} G_{\alpha\alpha}^{1/2}$. Interestingly, we find an excellent agreement (even quantitative) in all cases involving stars with harmonic bonds (solution and melt). However, in the FENE case, the values of $T_B$ and $\Delta T_{\tilde v}$ differ at small and moderate shear rate, and become similar as $\wi$ increases. For moderate and small $\wi$ we find $\Delta T_{\tilde v} <T _B$, indicating that the stronger excluded volume forces in FENE bonds (arm elongations are confined to a fixed value) tend to reduce collective vibrations (breathing) of the star molecules. \begin{figure} \includegraphics[width=0.5\columnwidth]{Figure5} \caption{The absolute value of kinetic energy related to vibrational non-angular momentum $|T_{u}|$ compared with the rate of decorrelation ($\Gamma$) of the monomer pure rotation around the molecule center (see Fig. \ref{fig:omega_d_fit}). Both quantities are normalized with their values at zero shear rate. Top and middle panel, results for solution and bottom panel, for the melt.} \label{fig:dis} \end{figure} \begin{figure*} \includegraphics[scale=0.55, angle=-90]{Figure6} \caption{(a) The angular momentum free vibrational kinetic $T_{\tilde v}$ (symbols) and the fit $T_{\tilde v}(\wi) =T_{\tilde v}(0)+ a \wi^{\beta}$ with $\Delta T_{\tilde v} = a \wi^{\beta}$ the coherent part and $T_{\tilde v}(0)$ the thermal contribution (results for star in solution). (b) The coherent contribution $\Delta T_{\tilde v}$ is compared with the breathing mode energy estimated as $T_B=\frac{N}{2} \omegat^2 \mathrm{Std}[G_{1}]$, where $G_{1}$ is the principal eigenvalue of the gyration tensor and $\omegat$ is the breathing frequency, reported in Ref. \cite{Sablic:2016:1} (results for solution). (c) The same as (b) but for the melt case, and $T_B=0.45\,\frac{N}{2} \omegat^2 \mathrm{Std}[G_{1}]$. } \label{fig:vib} \end{figure*} \subsection{Intrinsic viscosity} One of the major tasks of polymer physics is to relate individual chain dynamics with macroscopic rheological properties. We make such an exercise in this section, taking the shear viscosity as our target macroscopic quantity. In a previous work, we analyzed in some detail the rheology of these stars in melt \cite{Sablic:2016:1} and reported in particular its shear viscosity under shear. Here, we calculate the contribution to the shear viscosity of star polymers in solution from their contribution to the virial part of the shear stress tensor \cite{Doyle:2005}, \begin{equation} {\bm \sigma} =\rho_P \Bigg \langle \sum_{\alpha = 1}^{N} \left(\mathbf{F}^{nb}_{\alpha} + \mathbf{F}^{b}_{\alpha}\right) \otimes \left(\mathbf{r}_{\alpha} - \mathbf{r}_{cm}\right) \Bigg\rangle. \end{equation} Here, $\mathbf{F}^{nb}_{\alpha}$ represents the force on the $\alpha$-th monomer, originating from the non-bonded interactions (i.e. the Weeks-Chandler-Anderson interaction), and $\mathbf{F}^{b}_{\alpha}$ are the forces of the bonds (i.e. either harmonic or FENE). The polymer contribution to the stress tensor is proportional to $\rho_P$, the number density of polymer molecules, and the polymer contribution to the shear viscosity is ~\cite{Doyle:2005} \begin{equation} \eta = - \frac{\sigma_{12}}{\dot{\gamma}}. \end{equation} Using the Carreau fit~\cite{Yasuda:2006, Aho:2011}, we estimate the zero-shear rate viscosity $\eta_0$ and present the normalized viscosity $\eta/\eta_0$. We note that $\eta_0$ is about 1.8 times larger in the case of the harmonic-bond model compared with the FENE bonds. As the shear rate is increased, we find shear thinning $\eta \sim \wi^{-\beta}$ with shear thinning exponents $\beta=0.25$ for FENE bonds and $\beta=0.32$ for harmonic bonds. These values are somewhat smaller than those found in melt, $\beta=0.49$ (see Fig.\ref{fig:eta_vs_Wi}). Viscous dissipation is related to decorrelation times and in fact, the intrinsic viscosity can be expressed as an sum of relaxation times \cite{Doi1994}. For an isolated star in dilute solution, one expects that the main mechanism for dissipation comes from the decorrelation in arm lengths, which takes place at an average rate $\Gamma$ (see Fig. \ref{fig:omega_d_fit}). Thus, as a first estimate, we seek a relation of the form $\eta \propto \Gamma^{-1}$. Figure \ref{fig:eta_vs_Wi} shows that such relation holds relatively well, both in solution and melts. For instance, in solution we see that the softer harmonic bonds leads to faster decorrelation rates and smaller intrinsic viscosity, compared with the more rigid FENE chains. As we indicated in Fig. \ref{fig:dis}, we found that, in solution, $\Gamma$ scales like the kinetic energy $|T_u|$ and consistently, $|T_u|$ is larger in the case of harmonic bonds compared with FENE-stars. In melts, however, one expects that the departure from rigid-body rotation (measured by the velocity $u$ and its kinetic energy $T_u$) arises also from inter-molecular collisions (and not only from Brownian diffusion). This is revealed in the different trends followed by $T_u$ and $\Gamma$ in melts: unlike what it is observed in solution, $T_u$ and $\Gamma$ do not correlate (see Fig. \ref{fig:dis} botom). \begin{figure*} \includegraphics[width=\columnwidth]{Figure7} \caption{The intrinsic shear viscosity $\eta$ is compared with the normalized decorrelation rate of the arms (center-to-end) distance $\Gamma \tau_{rot}$. The left panel corresponds to stars in dilute solution (here, we normalize with the viscosity at zero shear rate $\eta_0$) and the right panel to stars in melt (polymer volume fraction $0.2$). In solution, the shear stress scales like $\sigma_{12} = c \dot{\gamma}/\Gamma$ with $c=15$ for harmonic and $c=7$ for FENE bonds.} \label{fig:eta_vs_Wi} \end{figure*} \section{Conclusions} The main purpose of this work is to show that the Eckart formalism can be used to unveil the complex dynamics of soft molecules in flow. The application of the Eckart formalism to the dynamics of star molecules in shear flow permitted us to warn about the incorrect interpretation of the rotation dynamics of soft molecules (polymers) based on a standard (lab frame) analysis. In particular, the {\em apparent} angular velocity $\omegal$ resulting from such analysis has not a clear dynamical interpretation (it is not the rotation frequency of the molecule). We have shown that the Eckart co-rotating frame correctly extracts the different types of motions in the rotating and vibrating molecule: pure rotation, vibration with no-angular momentum and vibrational angular momentum. Star molecules in shear flow perform a tank-treading motion \cite{Ripoll:2006} whereby monomers rotate around the center of the molecule, but for a given fixed shear rate, the molecule keeps a {\em roughly fixed} ellipsoidal overall shape (more precisely, they do not tumble). At large shear rates, the molecule performs another collective motion, which we called ``breathing mode'' \cite{Sablic:2016:1}, whereby the gyration tensor of the molecule oscillates in time with a characteristic frequency $\omegat$. We have shown that each of these dynamics is associated with a different type of displacement in the Eckart frame. The pure rotational component of the Eckart frame, with a frequency $\omegae$, describes the tank-treading frequency of the star $\omegar$. By extracting the thermal (incoherent) part of the kinetic energy of vibrations without angular momentum component, we find that the kinetic energy of the breathing mode coincides with the energy of ``breathing'' vibrations. Finally, in solution, we find that the decorrelation of the end-to-end arm distance, driven by Brownian diffusion at a rate $\Gamma$, correlates with the kinetic energy associated to vibrations (or more properly, fluctuations) with angular momentum, $T_u$. In melt, such correlation is not observed, and it seems that the energy $|T_u|$ of molecular deformations is mainly determined by intermolecular collisions (and thus density dominated). In this work, we just consider star polymers with $f = 12$ arms and $m = 6$ monomers per arm. According to a recent analysis \cite{Chremos:2015}, star molecules become chain-alike for $f<6$ so our stars are within the ``colloidal-alike'' regime. But, what would be the dynamics of more massive stars? While this question is open to future works, we have good reasons to believe that they will be quite similar to that found for $f=12, m=6$. In fact, several computational works for star polymers in dilute \cite{Ripoll:2006} and semidilute \cite{Chen:2013} conditions, covered a relative large range of values of $f\leq 50 $ and $m< 50$ and (by defining the proper Weissenberg number) they found that all data for $\omega$ collapse in a master curve, indicating that the length of the arms or the functionality was not essentially changing the polymer dynamics. The dynamics would surely change in case of a semidilute solution (or melt) if the stars have {\em very} long arms ($m>100$), because entanglements should play a mayor role in distorting their rotation dynamics. However, we emphasize that the Eckart framework would be still applicable in such regime and provide valuable dynamic information. It also has to be noted that the present analysis can be complementary to the more detailed normal mode analysis of vibrations, within the framework of the theory of molecular vibrations~\cite{Eckart:1935}. In the latter, each internally rotating part of the molecule would require the introduction of additional internal coordinate systems inside the translating and rotating Eckart frame~\cite{Praprotnik:2005, Praprotnik:2005:1, Praprotnik:2005:2, Praprotnik:2005:3, Howard:1937:1, Howard:1937:2, Kirtman:1962, Kirtman:1964, Kirtman:1968}. Presently, we leave this discussion for the future work, since the main aim of this paper is the separation of rotations from vibrations, or the consequences such decomposition brings up in the interpretation of molecular rotations. The objective of this work is to show that the Eckart frame, successfully and routinely used to describe Raman spectra of small molecules, is also a robust and useful tool to investigate the complex dynamics of soft, semiflexible macromolecules. % If in two-column mode, this environment will change to single-column % format so that long equations can be displayed. Use % sparingly. %\begin{widetext} % put long equation here %\end{widetext} % figures should be put into the text as floats. % Use the graphics or graphicx packages (distributed with LaTeX2e) % and the \includegraphics macro defined in those packages. % See the LaTeX Graphics Companion by Michel Goosens, Sebastian Rahtz, % and Frank Mittelbach for instance. % % Here is an example of the general form of a figure: % Fill in the caption in the braces of the \caption{} command. Put the label % that you will use with \ref{} command in the braces of the \label{} command. % Use the figure* environment if the figure should span across the % entire page. There is no need to do explicit centering. % \begin{figure} % \includegraphics{}% % \caption{\label{}} % \end{figure} % Surround figure environment with turnpage environment for landscape % figure % \begin{turnpage} % \begin{figure} % \includegraphics{}% % \caption{\label{}} % \end{figure} % \end{turnpage} % tables should appear as floats within the text % % Here is an example of the general form of a table: % Fill in the caption in the braces of the \caption{} command. Put the label % that you will use with \ref{} command in the braces of the \label{} command. % Insert the column specifiers (l, r, c, d, etc.) in the empty braces of the % \begin{tabular}{} command. % The ruledtabular enviroment adds doubled rules to table and sets a % reasonable default table settings. % Use the table* environment to get a full-width table in two-column % Add \usepackage{longtable} and the longtable (or longtable*} % environment for nicely formatted long tables. Or use the the [H] % placement option to break a long table (with less control than % in longtable). % \begin{table}%[H] add [H] placement to break table across pages % \caption{\label{}} % \begin{ruledtabular} % \begin{tabular}{} % Lines of table here ending with \\ % \end{tabular} % \end{ruledtabular} % \end{table} % Surround table environment with turnpage environment for landscape % table % \begin{turnpage} % \begin{table} % \caption{\label{}} % \begin{ruledtabular} % \begin{tabular}{} % \end{tabular} % \end{ruledtabular} % \end{table} % \end{turnpage} % Specify following sections are appendices. Use \appendix* if there % only one appendix. \appendix %\section{} \section*{Appendix: The Eckart reference configuration} In the Eckart frame formalism of Eqs.~\ref{eqn:Eckart omega} and~\ref{eqn:Eckart inertia}, one needs to define a reference configuration which fixes $c_{i}^{\alpha}$ over time. These are the components of the monomers positions of the reference configuration in the initial internal coordinate system. We choose $c_{i}^{\alpha}$ in three different ways: \textbf{(i)} From an equilibrium configuration of a star polymer at temperature $0$~K. \textbf{(ii)} The reference configuration is obtained by Metropolis Monte Carlo (MC) simulation at the desired temperature $T=4$, which enforces by additional terms in the Hamiltonian that the configuration matches the average gyration tensor components at every shear rate. \textbf{(iii)} The $c_{i}^{\alpha}$s are not constant. Instead, they are changed after a certain number of sampled configurations in the trajectory. An instantaneous configuration is taken as a reference configuration for the following $\tau_w$ in time, i.e. this configuration is used to evaluate the angular velocity of rotation (using the Eckart frame formalism) from all the following trajectory snapshots within the time window $\tau_{w}$. Next, the first configuration following in the trajectory is taken as the new reference configuration. This procedure is thus repeated from the start until the end of the sampled trajectory. We analyze the rotation of molecules for different lengths of the time window and thus give the result for this third characterization of rotation by the Eckart frame formalism in the form of $3$-dimensional plots (Fig.~\ref{fig:Eckart_solution}). In all three described definitions of $c_{i}^{\alpha}$s, the unit base vectors of the internal coordinate system $\mathbf{f_{1}}$, $\mathbf{f_{2}}$, and $\mathbf{f_{3}}$ and the origin of the Eckart frame, defined by $\mathbf{r}_{cm}$, are different in every snapshot of the sampled trajectory. Only the reference components $c_{i}^{\alpha}$s remain constant throughout the whole trajectory in \textbf{(i)} and \textbf{(ii)}, while in \textbf{(iii)} also $c_{i}^{\alpha}$s change in time, as described above. Molecules rotate in a flow-gradient plane. Therefore, the only component of the molecules' angular velocity with non zero-average is in the neutral direction and we denote, ${\bm \omegal} = \left(\omegal_1, \omegal_2, \omegal_3 \right)$ and ${\bm \omegae} =\left(\omegae_1, \omegae_2, \omegae_3 \right)$, where indices $1$, $2$, and $3$ denote the flow, gradient, and neutral direction, respectively. To determine the optimal way to define $c_{i}^{\alpha}$s, we plot, in Fig.~\ref{fig:Eckart_solution}, angular velocities obtained by the Eckart formalism using the definitions \textbf{(i)}, \textbf{(ii)}, and \textbf{(iii)} for solution of star polymers with $12$ arms of $6$ monomers (connected by Hookean springs). Plots for the melt are qualitatively similar and are not shown here. We observe that the approach \textbf{(iii)}, in which $c_{i}^{\alpha}$s change every $\tau _{w}$, gives the angular velocity surface that at the shortest $\tau _{w}$ corresponds to the standard approach (i.e. using Eqs.~\ref{eqn:standard eqn. omega} and~\ref{eqn:standard inertia tensor}). With increasing $\tau _{w}$, it approaches the values obtained by the approaches \textbf{(i)} and \textbf{(ii)}. At a certain value of $\tau _{w}$, we observe a sharp crossover in angular velocity of polymers at very high shear rates, which results in qualitatively different dependencies $\omegae/\dot{\gamma}\left(W_{i}\right)$ emerging only due to the different reference frames. A similar crossover is also observed in melts, but is more prominent in solutions. Furthermore, we observe that this crossover occurs at higher $\tau _{w}$ for the star polymers with longer arms. \begin{figure} \includegraphics[scale=0.7]{Figure8} \caption{Three definitions of reference configuration to calculate the angular velocity by Eckart frame formalism for star polymers in solution: \textbf{(i)} The reference configuration is the equilibrium configuration at temperature $0$~K (green line). \textbf{(ii)} We obtain the reference configuration at every shear rate separately by the Monte Carlo simulation so that it matches the average steady state shape of a polymer (i.e. gyration tensor) at that particular shear rate (blue line). \textbf{(iii)} The reference configuration is taken to be an instantaneous one, but in this way defined $c_{i}^{\alpha}$s are used in computation of the angular velocity only for the following $\tau _{w}$ in time. Afterwards, the reference configuration is replaced with the next instantaneous configuration, from which we define new $c_{i}^{\alpha}$s.} \label{fig:Eckart_solution} \end{figure} Importantly, we find that the definitions \textbf{(i)} and \textbf{(ii)} yield basically the same results, which are also similar to the results obtained by the definition \textbf{(iii)} after the crossover. Therefore, in the manuscript, we present only results obtained by the definition \textbf{(ii)}. % If you have acknowledgments, this puts in the proper section head. \begin{acknowledgments} %\section*{Acknowledgements} J. S. and M. P. acknowledge financial support through grants P1-0002 and J1-7435 from the Slovenian Research Agency. J. S. acknowledges financial support from Slovene Human Resources Development and Scholarship Fund (186. JR). R. D.-B. acknowledges support from the Spanish government under national MINECO project FIS2013-47350-C5-1-R. Partial support from COST Action MP1305 is kindly acknowledged. \end{acknowledgments} % Create the reference section using BibTeX: \bibliography{bibliography} \end{document} }
\caption{Black arrows: Top--down generative model; Program$\to$Spec$\to$Image. {\color{red}Red} arrows: Bottom--up inference procedure. \textbf{Bold:} Random variables (image/spec/program)}
\caption{Neural architecture for inferring specs from images. \textcolor{blue}{Blue}: network inputs. Black: network operations. \textcolor{red}{Red}: draws from a multinomial. \texttt{Typewriter font}: network outputs. Renders on a $16\times 16$ grid, shown in \textcolor{gray}{gray}. STN: differentiable attention mechanism~\citep{jaderberg2015spatial}.}
\caption{Parsing~\LaTeX~output after training on diagrams with $\leq 12$ objects. Out-of-sample generalization: Model generalizes to scenes with many more objects ($\approx$ at ceiling when tested on twice as many objects as were in the training data). Neither SMC nor the neural network are sufficient on their own. \# particles varies by model: we compare the models\emph{with equal runtime} ($\approx 1$ sec/object). Average number of errors is (\# incorrect drawing commands predicted by model)$+$(\# correct commands that were not predicted by model).}
\caption{Verifying tracking results on a typical sequence. Verifier validates tracking results every 10 frames. Most of the time the tracking results are reliable (showing in \textcolor{blue}{blue}). Occasionally, \eg frame \#080, the verifier finds the original tracking result (showing in\textcolor{blue}{blue}) unreliable and the tracker is corrected and resumes tracking based on detection result (showing in \textcolor{red}{red}).}
\caption{Detection based on verification. When finding an unreliable tracking result (showing in \textcolor{blue}{blue} in (a)), the verifier $\VF$ detects the target in a local region ( shown in (b)). The dashed \textcolor{red}{red} rectangles in (b) represent object candidates generated by sliding window. The \textcolor{red}{red} rectangle in (c) is the detection result.}
\caption{Architecture of the proposed model, \textbf{Switch-CNN} is shown. A patch from the crowd scene is highlighted in \textcolor{red}{red}. This patch is relayed to one of the three CNN regressor networks based on the CNN label inferred from \textbf{Switch}. The highlighted patch is relayed to regressor $R_3$ which predicts the corresponding crowd density map. The element-wise sum over the entire density map gives the crowd count of the crowd scene patch.}
\caption{Switch-CNN training algorithm is shown. The training algorithm is divided into stages coded by color. \textbf{Color code index}: \textcolor{BlueViolet}{Differential Training}, \textcolor{OliveGreen}{Coupled Training}, \textcolor{red}{Switch Training}}
\caption{Comparison of classification accuracy for different switch architectures on Part A of the ShanghaiTech dataset~\cite{zhang2016single}. The final switch-classifier selected for all Switch-CNN experiments is highlighted in \textcolor{red}{red}. }
\caption{\label{tab:silver_data_analysis} \textcolor{blue}{We sampled 1866 system generated paraphrase pairs and label them by expert: True is paraphrase and False is non-paraphrase. We show precision at different probability thresholds. If we set cut off value at 0.9, we can get 114025*(846/1866)*(1/3)=17232 paraphrase pairs per month with high precision 82.98\%.}}
\caption{NV center with its axis oriented in the [111] direction. The angle between the NV axis and $\mathbf{B}_0$ is denoted by the angle $\theta$. NV centers can also have their axes oriented in the $\left[ {1\,\bar 1\,\bar 1} \right]$, $\left[{\bar 1\,1\,\bar 1} \right]$, and $\left[ {\bar 1\,\bar 1\, 1} \right]$ directions.} \label{fig:NVorientations} \end{figure} We can explain the main features of these spectra with a simple model. According to Eq.~\ref{eq:nupm}, the Zeeman shift of the ground state energy levels of the NV center depends on $B_{||}$, the component of the applied magnetic field parallel to the axis of the NV center. Figure~\ref{fig:NVorientations} shows an NV center with its axis oriented in the [111] direction. The angle between the NV axis and $\mathbf{B}_0$ is denoted by the angle $\theta$. NV centers can also have their axes oriented in the $\left[{1\,\bar 1\,\bar 1} \right]$, $\left[ {\bar 1\,1\,\bar 1} \right]$, and $\left[{\bar 1\,\bar 1\, 1} \right]$ directions. If we assume that these four orientations are equiprobable, then for $\mathbf{B}_0$ oriented perpendicular to [111] surface, we expect that 1/4 of the NV centers will have $B_{||}=B_0 \cos 0^\circ = B_0$ and 3/4 of the NV centers will have $B_{||}=B_0 \cos 109.5^\circ = -\frac {B_0}{3} $. Model spectra are given in the gray curves included with the $B_0\ne0$ spectra in Fig.~\ref{fig:singlecrystal}~(b)--\ref{fig:singlecrystal}(e). Here we assume that each spectrum is a 3:1 weighted superposition of two functions of the form $I\left( \nu \right)$, one calculated using $B_{||}=B_0$ and the other using $B_{||}= -\frac{B_0}{3}$, where the value of $B_0$ is calculated from the coil geometry and the measured current flowing in the coils. \textit{Thus no new adjustable parameters are needed to fit the $B \ne 0$ spectra.} This is the basis of using ODMR spectroscopy with NV centers to measure magnetic fields.\cite{maze2008nanoscale, balasubramanian2008nanoscale, acosta2011optical, glenn2015single} In fact the small discrepancies between the dip locations in the experimental data relative to the model spectra are likely due to the fact that the model relies on a value of $B_0$ determined indirectly. A value of $B_0$ obtained by fitting the model spectra to the experimental spectra would be a more accurate measure of the magnetic field strength in the region being sampled. Our analysis suggests that we can identify the four fluorescence dips as corresponding to the following values of $m_s$, $\theta$, and $B_{||}$: \vspace{10 pt} \hspace{20 pt} $1 \to{{m_s} = + 1,\;\theta = 0^\circ },\;B_{||}=B_0$ \hspace{20 pt} $2 \to{{m_s} = -1,\;\theta = 109.5^\circ },\;B_{||}=-B_{0}/3$ \hspace{20 pt} $3 \to{{m_s} = + 1,\;\theta = 109.5^\circ },\;B_{||}=-B_{0}/3$ \hspace{20 pt} $4 \to{{m_s} = - 1,\;\theta = 0^\circ },\;B_{||}=B_0$ \vspace{10 pt} \noindent A comparison between the predicted and the experimentally observed dip locations is shown in Fig.~\ref{fig:singlecrystal2}, where we plot the frequency of each of the four dips from Fig. \ref{fig:singlecrystal} as a function of the magnitude of $\mathbf{B}_0$. At low fields the dips are hard to discern -- see, for example, the spectrum in Fig.~\ref{fig:singlecrystal}(b) -- so we plot only the range where distinct dips are observed. The dotted lines, which show the location of these four features predicted by Eq. ~\ref{eq:nupm}, are in good agreement with the experimental observations. Dips 1 and 4 are associated with those NV centers whose axes are along the direction of applied magnetic field, while dips 2 and 3 are associated with centers whose axes make an angle of $109.5^\circ$ with the applied field. Note that for the latter case $B_{||}$ is negative, so Eq.~\ref{eq:dnu} implies that the frequency of the $m_s=+1$ ($m_s=-1$) transition decreases (increases) as the magnetic field strength increases. Since we expect only one out of every four centers to have $\theta = 109.5^\circ,$ we anticipate that that the amplitudes of dips 1 and 4 should be roughly one third the amplitude of dips 2 and 3. The amplitude ratio in the measured spectra is clearly more than the expected 1:3, for reasons that we cannot explain. \begin{figure}[!htbp] \ifeps \includegraphics[width=3.25in]{singlecrystal2.eps} \else \includegraphics[width=3.25in]{singlecrystal2.pdf} \fi \caption{Zeeman splitting of the NV ground state as a function of magnetic field strength. The magnetic field is directed along the [111] direction. We use different symbols to mark the locations of the different fluorescence dips: $\circ \to \left({{m_s} = + 1,\theta = {0^\circ }} \right)$, \mbox{\footnotesize $ \times$ } $\to \left( {{m_s} = -1,\theta = 109.5^\circ } \right)$, \mbox{\tiny $ \square$} $\to \left({{m_s} = +1,\theta = 109.5^\circ } \right)$, \mbox{\footnotesize $ \lozenge$ } $\to \left( {{m_s} = -1,\theta = 0^\circ } \right)$. The dotted lines show the location of these four features predicted by Equation ~\ref{eq:nupm}. } \label{fig:singlecrystal2} \end{figure} %Strictly speaking, we should recalculate the model curves here, though the difference will be minute. \subsection{ODMR in diamond nanocrystals} \begin{figure}[!htbp] \ifeps \includegraphics[width=3.25in]{nanocrystal.eps} \else \includegraphics[width=3.25in]{nanocrystal.pdf} \fi \caption{Zeeman broadening of ODMR in diamond nanocrystals for different magnetic field strengths. Due to the random orientation of the nanocrystals we no longer observed distinct Zeeman levels as the field strength increased, but rather a steady broadening of the $\mathbf{B}_0 = 0$ resonance.} \label{fig:nanozeeman} \end{figure} Optically detected magnetic resonance spectra recorded from diamond nanocrystals are shown in Fig.~\ref{fig:nanozeeman}. The $\mathbf{B}_0 = 0$ spectrum (Fig.~\ref{fig:nanozeeman}(a)) again exhibits a pair of closely spaced dips in fluorescence intensity in the vicinity of 2.87~GHz. This is similar to what we observed in our single crystal measurements, although the resonance in this case is somewhat broader and has a smaller ($\sim 4\;\%$) contrast with the off-resonance fluorescence. This is likely due to inhomogeneous broadening from sampling an ensemble of nanocrystals with slightly different resonant frequencies.\cite{Gruber-Science1997} The data shown here were collected using relatively long scan times (about 1 minute). During this time there is a small upward drift in the laser power, which manifests in the spectrum shown in Fig.~\ref{fig:nanozeeman}(a). We can model these spectra using an approach similar to the one used to analyze the single crystal spectra. As before, we start with the $B_0=0$ spectrum in Fig.~\ref{fig:nanozeeman}(a). Since there is no magnetic field, the fact that our measurements are being made on an ensemble of nanocrystals with different orientations has no real effect and we can proceed just as in the single crystal case. A best fit of the spectrum to a function of the form $I\left(\nu \right)$ gives $C=0.04$, $E=0.0050\;\rm{GHz}$, $D=2.687\;\rm{GHz}$, and $\Gamma=0.012\;\rm{GHz}$. Figure~\ref{fig:nanozeeman}(b)--\ref{fig:nanozeeman}(d) shows ODMR spectra recorded in the presence of a static magnetic field $\mathbf{B}_0$, ranging in magnitude from 0.46 mT through 1.18 mT. The results are noticeably different from the single crystal case; we no longer observe distinct Zeeman levels as the field strength increases, but rather a steady broadening of the $\mathbf{B}_0 = 0$ resonance. This is because the crystal axes of the ensemble of nanocrystals are oriented in random directions with respect to $\mathbf{B}_0$. The gray curves are produced by extending the model used in the single crystal case by summing over all the different (assumed equally likely) orientations of the nanocrystals with respect to $\mathbf{B}_0$. There is qualitative agreement between the model and the measured spectra, but the model does not account for the fact that the experimental spectra are not symmetrical about the center frequency $D$. The absence of well-defined troughs makes it harder to extract accurate values for the magnetic field, but the width of the resonance can serve as an indicator of the strength of the applied magnetic field. Figure~\ref{fig:nanozeeman2} shows the full width at half maximum of the resonance as a function of the magnetic field strength. The open circles represent values for the width derived from the measured spectra while the black curve represent widths obtained from the model spectra. This gives a simple means to translate the measured width of a resonance curve into a magnetic field strength, this time using ODMR spectra recorded from small quantities of nanocrystals. Having a simple method to optically measure magnetic field strengths with a spatial resolution of several microns has significance for biological applications, since diamond nanocrystals can be inserted into living cells.\cite{kucsko2013nanometre} \begin{figure}[!htbp] \ifeps \includegraphics[width=3.25in]{nanodiamonds2.eps} \else \includegraphics[width=3.25in]{nanodiamonds2.pdf} \fi \caption{ Full width at half maximum of the ODMR resonance as a function of the magnetic field strength.The open circles represent values for the width derived from the experimentally measured spectra while the black curve is generated from the model spectra.} \label{fig:nanozeeman2} \end{figure} \subsection{ODMR in an instructional setting} The Wellesley Physics Department's ``junior lab'' course is organized in a fairly standard format: We have a suite of advanced experiments that students rotate through. Working in pairs over the course of about one and one-half weeks (three 3-hour sessions) students can complete a typical experiment. Each experiment comes with a write-up that summarizes the theoretical background, suggests in general terms an appropriate experimental procedure, and then gives some guidance as to how to analyze the data. We try to give enough guidance so that our students can be successful, but not so much that it feels like they are simply following a recipe. The ODMR experiments described here fit well within this framework. Starting from a blank breadboard they construct the fluorescence microscope and then record ODMR spectra from a large single crystal. We then ask them to analyze the data in the framework of the models presented above. Switching to a microcrystal sample presents additional challenges. Exciting and collecting fluorescence from a single microcrystal requires a more careful optical alignment. Also, since the orientation of the microcrystal is not known in advance, the ODMR spectra recorded in the presence of an external magnetic field are typically more complicated and harder to interpret. (A challenging question for students is to see what they can infer about the orientation of the microcrystal from their measurements.) Overall the experience is a good mix of experimental technique, careful data analysis, and modeling based on physics that connects to what they have seen in their quantum mechanics class. A copy of the lab write-up for this experiment is included with the supplementary material. % [AJP to include URL here] In the Advanced Placement Physics course at Dover Sherborn High School, students cycle through a sequence of four modern experiments, each lasting two hours. For the diamond magnetometer experiment, students use the simple set up described above to record ODMR spectra from an NV-rich large single crystal diamond. Using a strong permanent magnet to generate a magnetic field, they observe how the spectrum changes as they vary (by hand) the distance from the magnet to the diamond. They make a video recording of this process, with a field of view that includes the spectrum and the permanent magnet, as well as a ruler taped to the table. By analyzing this video they can determine how the magnet's field varies with distance. \section{Conclusion} What makes a lab experience for advanced undergraduate physics majors both compelling and educational? Physicists have a long and valuable tradition of building their own instruments and many of the most important advances in scientific history were based on a combination of science, engineering, and design. But this tradition may be waning. Both the power and the problem with much modern scientific instrumentation is reflected in the term ``black box'' that is commonly used to describe the equipment. Today's black-box instruments are highly effective in making measurements and collecting data, enabling even novices to perform advanced scientific experiments. But, at the same time, these black boxes are ``opaque'' -- in that their inner workings are often hidden and thus poorly understood by their users. In contrast, the fluorescence microscope used in this experiment can be largely set up from scratch by students. There is some irony in the fact that, in order to eliminate interference from room light, the instrument that students construct is literally located inside a black box. But unlike a metaphorical black box, this is one they can reach inside, to build and to explore. \vspace{20 pt} \section{Appendix: Parts list} Here, for convenience, we give a parts list for the experiments described above. We do not include in these lists commonly available parts such as oscilloscopes, function generators, and dc power supplies. \subsection{Samples} Sources for the samples used in these experiments are listed in Table ~\ref{samples}. For large NV-rich single crystals, suitable starting material is 2 point (that is 4 milligram) high-pressure high-temperature synthetic diamond single crystal from Element Six. It must then be electron irradiated and annealed, as described above. \begin{table}[h!] \centering \caption{Samples} \label{samples} \begin{tabular}{|c|c|c|c|} \hline \textbf{Part} & \textbf{Supplier} & \textbf{Part} & \textbf{Cost(\$)} \\ \hline microcrystals & Ad\'{a}mas & MDNV15umHi50mg & 300 \\ \hline nanocrystals & Ad\'{a}mas & ND-NV-100nm & 300 \\ \hline singlecystals & Element Six & Monocrystal 2pt & 125-400 \\ \hline \end{tabular} \end{table} \subsection{Microscope parts} The optical components and mounts for the fluorescent microscope are listed in Table ~\ref{microscopeparts}. This is generally research grade equipment, but there are a number of lower cost alternatives that can work reasonably well. For example, the \textit{xyz} stage listed below is nice, but it is expensive and the precision it provides is greater than what is required here. An adequate replacement at 1/4 the cost can be obtained from banggood.com, part number 1105874. The same goes for the microscope objective: just about any 10x objective will do. The photodiode with adjustable gain amplifier could easily be replaced by a 50 cent raw photodiode and a home made current to voltage converter. \begin{table}[h!] \centering \caption{Microscope parts} \label{microscopeparts} \begin{tabular}{|c|c|c|c|} \hline \textbf{Part} & \textbf{Supplier} & \textbf{Part} & \textbf{Cost(\$)} \\ \hline 10$times$ microscope objective & Thorlabs & RMS10X & 340 \\ \hline thread adapter & Thorlabs & SM1A3 & 13 \\ \hline dichroic mirror & Thorlabs & DMLP550 & 170 \\ \hline long pass filter & Thorlabs & FEL0600 & 75 \\ \hline short pass filter & Thorlabs & FES0750 & 75 \\ \hline ND filter wheel & Thorlabs & FW1AND & 300 \\ \hline camera & Thorlabs & DCC1645C & 355 \\ \hline 100 mm camera lens & Thorlabs & MVL100M23 & 185 \\ \hline photodiode detector & Thorlabs & PDA36A & 320 \\ \hline mirror & Thorlabs & ME1-G01 & 13 \\ \hline 1'' optics holders (3) & Thorlabs & LRM1 & @15 \\ \hline 1" pedestal posts (5) & Thorlabs & RS1P8E & @22 \\ \hline 0.5" pedestal post & Thorlabs & RS0.5P8E & 22 \\ \hline clamping forks (6) & Thorlabs & CF125C & @11 \\ \hline flip mount & Thorlabs & TRF90 & 82 \\ \hline \textit{xyz} stage & Newport & MS-125-XYZ & 580 \\ \hline construction rails (4) & Thorlabs & XE25L09 & @15 \\ \hline \end{tabular} \end{table} \subsection{Laser parts} The laser system we used was expensive (\$2600) but the Coherent Compass 215M is just as good at 1/5 the price. (Downsides: The Compass 215M is intended as an OEM system and is not available directly from the manufacturer. It is however readily available online -- eBay, for example. Also, you will need to provide a 5V, 4A power supply.) The power stability these systems provide (1\%) improves the signal to noise ratio, but it is not absolutely essential provided you do fast scans. The Laserglow system is stable to only about 10\% but it is even less expensive and is plug and play. The lasers used in these experiments are Class IIIB and eye safety must be taken very seriously. There are a number things we do to make things as safe as possible for novice student users. The first line of defense is the requirement that everyone wear appropriate laser safety glasses (optical density at least 2.0 at 532 nm) while the laser is on. We also place an absorptive neutral density filter with an optical density of at least 2.0 immediately in front of the laser during the alignment phase of the experiment, when laser beams are most likely to go astray. Finally, the black box enclosure for the entire system keeps the beams confined to a small area in the room. \begin{table}[h!] \centering \caption{Laser Parts} \label{laserparts} \begin{tabular}{|c|c|c|c|} \hline \textbf{Part} & \textbf{Supplier} & \textbf{Part} & \textbf{Cost(\$)} \\ \hline laser safety glasses & DiOptika & LG-005L & 50 \\ \hline 40 mW laser module & Thorlabs & DJ532- 40 & 180 \\ \hline laser mount & Thorlabs & LDM21 & 340 \\ \hline current source & Thorlabs & LDC210C & 1100 \\ \hline temperature controller & Thorlabs & TED200C & 1000 \\ \hhline{|=|=|=|=|} 50 mW laser system & Coherent & Compass 215M & 500 \\ \hhline{|=|=|=|=|} 20 mW laser system & Laserglow & LCS- 532 & 300 \\ \hline \end{tabular} %\caption*{} \end{table} \subsection{Microwave parts} External dc power supplies are required for both the voltage controlled oscillator (5V) and the microwave amplifier (12V). \begin{table}[h!] \centering \caption{Microwave parts} \label{microwaveparts} \begin{tabular}{|c|c|c|c|} \hline \textbf{Part} & \textbf{Supplier} & \textbf{Part} & \textbf{Cost(\$)} \\ \hline voltage controlled osc. & Mini-Circuits & ZX95-3150+ & 40 \\ \hline microwave amplifier & Mini-Circuits & ZRL-3500 & 140 \\ \hline 2 dB attenuators (4) & Omni Spectra & 2082-6171-02 & @15 \\ \hline \end{tabular} \end{table} \section{Acknowledgments} This work relied heavily on the support of our colleagues at the \textit{Center for Integrated Quantum Materials}. In particular, we wish to thank Nathalie de Leon for first suggesting this project and for many subsequent discussions that were essential to our progress. The development of the simplified setup greatly benefited from help from Michael Walsh and Hannah Clevenson as well as Marko Lon\v{c}ar, Anna Schneidman, Robert Hart, John Free, and Danielle Braje. Daniel Twitchen and Mathew Markham at Element 6 graciously provided the large single crystal samples. Jim MacArthur from the the Harvard Physics Electronics Shop built the high resolution microwave source, based on a design by Sasha Zibrov, while Paul Horowitz made some useful suggestions on the design of the low resolution microwave source. Wellesley students Catherine Matulis, Hanae Yasakawa, Phyllis Ju and Hannah Peltz Smalley were instrumental in getting our diamond studies at Wellesley off the ground. This work is supported by the Center for Integrated Quantum Materials under NSF grant DMR-1231319. We are also grateful for support from Wellesley College, including the Sally Etherton Cummins Summer Science Research Endowed Fund and the T.T. and W.F. Chao Summer Scholars Program in Natural Sciences Endowed Fund. \ifbibtex \bibliography{ODMR} % need .bib file containing all cites to generate the references \else \begin{thebibliography}{99} \bibitem{donnally1963some}B. L. Donnally and E. Bernal G, ``Some Experiments on Nuclear Magnetic Resonance'', Am. J. Phys. \textbf{31}, 779--784 (1963) \bibitem{biscegli1982advanced} C. Biscegli, H. Panepucci, H. A. Farach, and C. P. Poole Jr, ``Advanced laboratory NMR spectrometer with applications'', Am. J. Phys. \textbf{50}, 48--50 (1982). \bibitem{footenote1} This is the specification for the Bruker 500 MHz NMR spectrometer that is the workhouse used in organic chemistry labs at Wellesley College. \bibitem{Gruber-Science1997} A. Gruber, A. Drabenstedt, C. Tietz, L. Fleury, J. Wrachtrup, and C. von Borczyskowski, ``Scanning Confocal Optical Microscopy and Magnetic Resonance on Single Defect Centers'', Science \textbf{276}, 2012--2014 (1997). \bibitem{hanson2007spins} R. Hanson, L. Kouwenhoven, J. Petta, S. Tarucha, and L. Vandersypen, ``Spins in few-electron quantum dots'', Rev. Mod. Phys. \textbf{79}, 1217--2065 (2007). \bibitem{childress2007coherent} L. I. Childress, \textit{Coherent manipulation of single quantum systems in the solid state}, Ph.D. thesis, Harvard University Cambridge, Massachusetts (2007). \bibitem{ChildressPhysicsToday2014} L. Childress, R. Walsworth, and M. Lukin, ``Atom-like crystal defects: From quantum computers to biological sensors'', Phys. Today \textbf{67}(10), 38--43 (2014). \bibitem{ladd2010quantum} T. D. Ladd, F. Jelezko, R. Laflamme, Y. Nakamura, C. Monroe, and J. L. O'Brien, ``Quantum computers'', Nature \textbf{464}, 45--53 (2010). \bibitem{doi2014deterministic} Y. Doi, T. Makino, H. Kato, D. Takeuchi, M. Ogura, H. Okushi, H. Morishita, T. Tashima, S. Miwa, S. Yamasaki, et al., ``Deterministic electrical charge-state initialization of single nitrogen-vacancy center in diamond'', Phys. Rev. X \textbf{4}, 011057 (2014) \bibitem{maze2008nanoscale} J. Maze, P. Stanwix, J. Hodges, S. Hong, J. Taylor, P. Cappellaro, L. Jiang, M. G. Dutt, E. Togan, A. Zibrov, A. Yacoby, R. L. Walsworth, and M. D. Lukin, ``Nanoscale magnetic sensing with an individual electronic spin in diamond'', Nature \textbf{455}, 644--647 (2008). \bibitem{balasubramanian2008nanoscale} G. Balasubramanian, I. Chan, R. Kolesov, M. Al-Hmoud, J. Tisler, C. Shin, C. Kim, A. Wojcik, P. R. Hemmer, A. Krueger, \textit{et al.}, ``Nanoscale imaging magnetometry with diamond spins under ambient conditions'', Nature \textbf{455}, 648--651 (2008). \bibitem{schirhagl2013nitrogen} R. Schirhagl, K. Chang, M. Loretz, and C. L. Degen, ``Nitrogen-Vacancy Centers in Diamond: Nanoscale Sensors for Physics and Biology'', Annu. Rev. Phys. Chem. \textbf{65}, 83--105 (2014). \bibitem{patange2013instrument} O. Patange, \textit{On an Instrument for the Coherent Investigation of Nitrogen-Vacancy Centres in Diamond}, Master's thesis, University of Waterloo (2013). \bibitem{markov2014limits} I. L. Markov, ``Limits on fundamental limits to computation'', Nature \textbf{512}, 147--154 (2014). \bibitem{footnote2} Other examples of quantum materials include single atomic layer materials such as graphene,\cite{geim2013van} which could enable ultrafast signal processing, and topological insulators,\cite{moore2010birth} which may lead to error-free channels for transporting spin information. Developing next generation electronic devices based on quantum materials is the mission of the National Science Foundation supported \textit{Center for Integrated Quantum Materials} (CIQM), which is supporting the work described here. See \texttt{ciqm.harvard.edu} for more information. \bibitem{resnick2000beyond} M. Resnick, R. Berg, and M. Eisenberg, ``Beyond black boxes: Bringing transparency and aesthetics back to scientific investigation'', J. Learn. Sci. \textbf{9}(1), 7--30 (2000). \bibitem{footnote3} A number of factors are responsible for the weakness of the interaction of the spin with its surroundings. For example, because the $^{12}\rm{C}$ nucleus has a spin of zero, the ``spin-spin'' interactions are small. Also, the low mass and stiff interatomic bonds of the carbon lattice lead to a high Einstein temperature ($T_E = 2300 \rm{K}$), which makes the interaction of the NV center) with the vibrational modes of the surrounding lattice unusually weak at room temperature. \bibitem{wrachtrup2006processing} J. Wrachtrup and F. Jelezko, ``Processing quantum information in diamond'', J. Phys. Condens. Matter \textbf{18}, S807--S824 (2006). \bibitem{footnote4} Equation \ref{eq:gyro} is only valid in the limit that the magnetic field strength is much smaller than the zero-field splitting between the $m_s=0$ state and the $m_s=\pm1$ states; i.e when $B_{||} << 2.87 \rm{GHz}/\gamma$. This condition is satisfied for all the experiments described here. \bibitem{doherty2013nitrogen} M. W. Doherty, N. B. Manson, P. Delaney, F. Jelezko, J. Wrachtrup, and L. C. Hollenberg, ``The nitrogen-vacancy colour centre in diamond'', Phys. Reports \textbf{528}, 1 (2013). \bibitem{footnote5} Element 6 sells HPHT diamond with suitably high nitrogen concentrations. See reference \onlinecite{2017arXiv170205332F} for more detailed guidance on creating high NV concentrations in such samples via electron irradiation and annealing. \bibitem{acosta2010temperature} V. Acosta, E. Bauch, M. Ledbetter, A. Waxman, L.-S. Bouchard, and D. Budker, ``Temperature dependence of the nitrogen-vacancy magnetic resonance in diamond'', Phys. Rev. Lett. \textbf{104}, 070801 (2010). \bibitem{acosta2011optical} V. M. Acosta, \textit{Optical magnetometry with nitrogen-vacancy centers in diamond}, Ph.D. thesis, Berkeley (2011). \bibitem{glenn2015single} D. R. Glenn, K. Lee, H. Park, R. Weissleder, A. Yacoby, M. D. Lukin, H. Lee, R. L. Walsworth, and C. B. Connolly, ``Single cell magnetic imaging using a quantum diamond microscope'', Nature methods \textbf{12}, 736 (2015). \bibitem{kucsko2013nanometre} G. Kucsko, P. Maurer, N. Yao, M. Kubo, H. Noh, P. Lo, H. Park, and M. Lukin, ``Nanometre-scale thermometry in a living cell'', Nature \textbf{500}, 54 (2013). \bibitem{le2013optical}D. Le Sage, K. Arai, D. Glenn, S. DeVience, L. Pham, L. Rahn-Lee, M. Lukin, A. Yacoby, A. Komeili, and R. Walsworth, ``Optical magnetic imaging of living cells'', Nature \textbf{496}, 486 (2013). \bibitem{footnote6} Excellent lower cost alternative laser systems are the Coherent Compass 215M or the Laserglow Technolgies LCS-532 series. There are even less expensive 532 nm lasers, but these are typically not temperature stabilized. This results in significant fluctuations in output power that may obscure the relatively small variations in fluorescence intensity that we observe here. To reduce the detrimental effects of power fluctuations it helps to record the spectra quickly. Also, for the laser intensities used here, the fluorescence is linearly proportional to the laser power and for longer scans we have had good success in minimizing the effect of laser power fluctuations by monitoring the laser power and normalizing. \bibitem{tisler2009fluorescence} J. Tisler, G. Balasubramanian, B. Naydenov, R. Kolesov, B. Grotz, R. Reuter, J.-P. Boudou, P. A. Curmi, M. Sennour, A. Thorel, \textit{et al.}, ``Fluorescence and spin properties of defects in single digit nanodiamonds'', ACS Nano \textbf{3}, 1959 (2009). \bibitem{geim2013van} A. Geim and I. Grigorieva, Nature \textbf{499}, ``Van der Waals heterostructures'', 419 (2013). \bibitem{moore2010birth} J. E. Moore, ``The birth of topological insulators'', Nature \textbf{464}, 194 (2010). \bibitem{2017arXiv170205332F} D. Farfurnik, N. Alfasi, S. Masis, Y. Kauffmann, E. Farchi, Y. Romach, Y. Hovav, E. Buks, and N. Bar-Gill, ``Enhanced concentrations of nitrogen-vacancy centers in diamond through TEM irradiation'', ArXiv e-prints (2017), arXiv:1702.05332 [quant-ph]. \end{thebibliography} \fi \end{document}}
\caption{Critical threshold $p_c$ as a function of $\gamma=k_{sB}^*/k_{sB}$ for different values of $z_A$, the internal connectivity of network $A$, where its internal degree distribution is RR. The curves represent different values of $z_A$: $z_A=3$ (\protect\bckline), $z_A=5$ (\protect\rddotted) and $z_A=10$ (\protect\bludash). Panel (a) corresponds to the Giant Component rule. Panel (b) corresponds to the ``mass rule'', with $q(h)=1$ for $h=1,2,3$, and panel (c) to the k-core rule with $k^*_X=2$. Note that in panel (b) $p_c \sim \gamma^{1/4}$ when $\gamma\to 0$, and thus corresponding curves appear finite even for very small $\gamma>0$.}
\caption{\textbf{Visualization of the Theorem \ref{mainthm2}}. Consider the set of selected points {\color{myblue} $\mathbf{s}$} and the points in the remainder of the dataset {\color{myred} $[n] \setminus \mathbf{s}$}, our results shows that if $\mathbf{s}$ is the $\delta_{\mathbf{s}}$ cover of the dataset, $\left| {\color{myred} \frac{1}{n}\sum_{i \in [n]} l(\mathbf{x}_i,y_i,A_{\mathbf{s}})} -{\color{myblue} \frac{1}{|\mathbf{s}|}\sum_{j \in \mathbf{s}} l(\mathbf{x}_j,y_j;A_{\mathbf{s}})} \right| \leq \mathcal{O}\left(\delta_\mathbf{s}\right) + \mathcal{O}\left(\sqrt{\frac{1}{n}}\right)$} \label{fig:thm} \end{figure} In order to show that this bound applies to CNNs, we prove the Lipschitz-continuity of the loss function of a CNN with respect to input image for a fixed true label with the following lemma where max-pool and restricted linear units are the non-linearities and the loss is defined as the $l_2$ distance between the desired class probabilities and the soft-max outputs. CNNs are typically used with cross-entropy loss for classification problems in the literature. Indeed, we also perform our experiments using the cross-entropy loss although we use $l_2$ loss in our theoretical study. Although our theoretical study does not extend to cross-entropy loss, our experiments suggest that the resulting algorithm is very effective for cross-entropy loss. \begin{lemma} Loss function defined as the 2-norm between the class probabilities and the softmax output of a convolutional neural network with $n_c$ convolutional (with max-pool and ReLU) and $n_{fc}$ fully connected layers defined over C classes is $\left(\frac{\sqrt{C-1}}{C} \alpha^{n_c+n_{fc}}\right)$-Lipschitz function of input for fixed class probabilities and network parameters. \end{lemma} Here, $\alpha$ is the maximum sum of input weights per neuron (see appendix for formal definition). Although it is in general unbounded, it can be made arbitrarily small without changing the loss function behavior (\ie keeping the label of any data point $\mathbf{s}$ unchanged). We defer the proof to the appendix and conclude that CNNs enjoy the bound we presented in Theorem~\ref{mainthm2}. In order to computationally perform active learning, we use this upper bound. In other words, the practical problem of interest becomes $\min_{\mathbf{s}^1:|\mathbf{s}^1 \leq b|} \delta_{\mathbf{s}^0\cup \mathbf{s}^1}$. This problem is equivalent to the k-Center problem (also called min-max facility location problem) \citep{facility}. In the next section, we explain how we solve the k-Center problem in practice using a greedy approximation.{\parfillskip0pt\par} \begin{wrapfigure}{R}{0pt} \begin{minipage}{0.44\textwidth} \vspace{-8mm} \begin{algorithm}[H] \caption{k-Center-Greedy} \label{alg:greedy} \begin{algorithmic} \STATE{\bfseries Input:} data $\mathbf{x}_i$, existing pool $\mathbf{s}^0$ and a budget $b$ \STATE Initialize $\mathbf{s}=\mathbf{s}^0$ \REPEAT \STATE $u=\arg\max_{i \in [n] \setminus \mathbf{s}} \min_{j \in \mathbf{s}} \Delta(\mathbf{x}_i, \mathbf{x}_j)$ \STATE $\mathbf{s} = \mathbf{s} \cup \{u\}$ \UNTIL{$|\mathbf{s}|=b+|\mathbf{s}^0|$} \STATE{\bfseries return} $\mathbf{s} \setminus \mathbf{s}^0$ \end{algorithmic} \end{algorithm} \vspace{-10mm} \end{minipage} \end{wrapfigure} \subsection{Solving the k-Center Problem} \label{sec:alg} We have so far provided an upper bound for the loss function of the core-set selection problem and showed that minimizing it is equivalent to the \emph{k-Center} problem (minimax facility location \citep{facility}) which can intuitively be defined as follows; choose $b$ center points such that the largest distance between a data point and its nearest center is minimized. Formally, we are trying to solve: \begin{equation} \min_{\mathbf{s}^1:|\mathbf{s}^1| \leq b} \max_i \min_{j \in \mathbf{s}^1 \cup \mathbf{s}^0} \Delta(\mathbf{x}_i,\mathbf{x}_j) \end{equation} Unfortunately this problem is NP-Hard \citep{cook}. However, it is possible to obtain a $2-OPT$ solution efficiently using a greedy approach shown in Algorithm~\ref{alg:greedy}. If $OPT=\min_{\mathbf{s}^1} \max_i \min_{j \in \mathbf{s}^1 \cup \mathbf{s}^0} \Delta(\mathbf{x}_i,\mathbf{x}_j)$, the greedy algorithm shown in Algorithm~\ref{alg:greedy} is proven to have a solution ($\mathbf{s}^1$) such that; $ \max_i \min_{j \in \mathbf{s}^1 \cup \mathbf{s}^0} \Delta(\mathbf{x}_i,\mathbf{x}_j) \leq 2 \times OPT$. Although the greedy algorithm gives a good initialization, in practice we can improve the $2-OPT$ solution by iteratively querying upper bounds on the optimal value. In other words, we can design an algorithm which decides if $OPT \leq \delta$. In order to do so, we define a mixed integer program (MIP) parametrized by $\delta$ such that its feasibility indicates $\min_{\mathbf{s}^1} \max_i \min_{j \in \mathbf{s}^1 \cup \mathbf{s}^0} \Delta(\mathbf{x}_i,\mathbf{x}_j) \leq \delta$. A straight-forward algorithm would be to use this MIP as a sub-routine and performing a binary search between the result of the greedy algorithm and its half since the optimal solution is guaranteed to be included in that range. While constructing this MIP, we also try to handle one of the weaknesses of k-Center algorithm, namely robustness. To make the k-Center problem robust, we assume an upper limit on the number of outliers $\Xi$ such that our algorithm can choose not to cover at most $\Xi$ unsupervised data points. This mixed integer program can be written as: \begin{equation} \begin{aligned} Feasible(b,\mathbf{s}^0,\delta, \Xi): &\sum_j u_j, = |\mathbf{s}^0|+ b, \quad && \sum_{i,j} \xi_{i,j} \leq \Xi \\ &\sum_j \omega_{i,j} = 1\quad \forall i, \quad && \omega_{i,j} \leq u_j \quad \forall i,j \\ & u_i =1 \quad \forall i\in \mathbf{s}^0, \quad &&u_i \in \{0, 1\} \quad \forall i \\ &\omega_{i,j} = \xi_{i,j} \quad \forall i,j \quad \mid&& \Delta(\mathbf{x}_i,\mathbf{x}_j) > \delta . \end{aligned} \label{eqmip} \end{equation} In this formulation, $u_i$ is 1 if the $i^{th}$ data point is chosen as center, $\omega_{i,j}$ is $1$ if the $i^{th}$ point is covered by the $j^{th}$, point and $\xi_{i,j}$ is 1 if the $i^{th}$ point is an outlier and covered by the $j^{th}$ point without the $\delta$ constraint, and $0$ otherwise. And, variables are binary as $u_i, \omega_{i,j}, \xi_{i,j} \in \{0,1\}$. We further visualize these variables in a diagram in Figure~\ref{mip}, and give the details of the method in Algorithm~\ref{alg:bin}. \begin{figure*}[h] \begin{minipage}[t]{0.5\textwidth} \vspace{-3mm} \begin{algorithm}[H] \caption{Robust k-Center} \label{alg:bin} \begin{algorithmic} \STATE{\bfseries Input:} data $\mathbf{x}_i$, existing pool $\mathbf{s}^0$, budget $b$ and outlier bound $\Xi$ \STATE{\bfseries Initialize} $\mathbf{s}_g =$ k-Center-Greedy($\mathbf{x}_i, \mathbf{s}^0, b$) \STATE $\delta_{2-OPT} = \max_j \min_{i \in \mathbf{s}_g} \Delta(\mathbf{x}_i,\mathbf{x}_j)$ \STATE $lb=\frac{\delta_{2-OPT}}{2}$, $ub=\delta_{2-OPT}$ \REPEAT \IF{$Feasible(b, \mathbf{s}^0,\frac{lb+ub}{2},\Xi)$} \STATE $ub=\max_{i,j \mid \Delta(\mathbf{x}_i,\mathbf{x}_j) \leq \frac{lb+ub}{2}} \Delta(\mathbf{x}_i,\mathbf{x}_j) $ \ELSE \STATE $lb=\min_{i,j \mid \Delta(\mathbf{x}_i,\mathbf{x}_j) \geq \frac{lb+ub}{2}} \Delta(\mathbf{x}_i,\mathbf{x}_j) $ \ENDIF \UNTIL{$ub = lb$} \STATE{\bfseries return} $\{i\ s.t.\ u_i=1\}$ \end{algorithmic} \end{algorithm} \end{minipage} \quad \begin{minipage}[t]{0.45\textwidth} \vspace{-1mm} \includegraphics[width=\textwidth]{mip.pdf} \vspace{-5mm} \caption{Visualizations of the variables. In this solution, the $4^{th}$ node is chosen as a center and nodes $0,1,3$ are in a $\delta$ ball around it. The $2^{nd}$ node is marked as an outlier.} \label{mip} \end{minipage} \end{figure*} \subsection{Implementation Details} \label{sec:imp} One of the critical design choices is the distance metric $\Delta(\cdot,\cdot)$. We use the $l_2$ distance between activations of the final fully-connected layer as the distance. For weakly-supervised learning, we used Ladder networks \citep{ladder} and for all experiments we used VGG-16 \citep{vgg} as the CNN architecture. We initialized all convolutional filters according to \citet{he_et_al}. We optimized all models using RMSProp with a learning rate of $1\mathrm{e}{-3}$ using Tensorflow~\citep{tensorflow}. We train CNNs from scratch after each iteration. % of the active learning. We used the Gurobi~\citep{gurobi} framework for checking feasibility of the MIP defined in (\ref{eqmip}). As an upper bound on outliers, we used $\Xi=1\mathrm{e}{-4} \times n$ where $n$ is the number of unlabelled points. \section{Experimental Results} \label{sec:exp} We tested our algorithm on the problem of classification using three different datasets. We performed experiments on CIFAR~\citep{cifar} dataset for image classification and on SVHN\citep{svhn} dataset for digit classification. CIFAR~\citep{cifar} dataset has two tasks; one coarse-grained over 10 classes and one fine-grained over 100 classes. We performed experiments on both. We compare our method with the following baselines: $i)$\textbf{Random:} Choosing the points to be labelled uniformly at random from the unlabelled pool. $ii)$\textbf{Best Empirical Uncertainty:} Following the empirical setup in \citep{gal_active}, we perform active learning using max-entropy, BALD and Variation Ratios treating soft-max outputs as probabilities. We only report the best performing one for each dataset since they perform similar to each other. $iii)$ \textbf{Deep Bayesian Active Learning (DBAL)\citep{gal_active}:} We perform Monte Carlo dropout to obtain improved uncertainty measures and report only the best performing acquisition function among max-entropy, BALD and Variation Ratios for each dataset. $iv)$ \textbf{Best Oracle Uncertainty:} We also report a best performing oracle algorithm which uses the label information for entire dataset. We replace the uncertainty with $l(\mathbf{x}_i,y_i,A_{\mathbf{s}^0})$ for all unlabelled examples. We sample the queries from the normalized form of this function by setting the probability of choosing the $i^{th}$ point to be queried as $p_i=\frac{l(\mathbf{x}_i,y_i,A_{\mathbf{s}^0})}{\sum_j l(\mathbf{x}_j,y_j,A_{\mathbf{s}^0})}$. $v)$\textbf{k-Median:} Choosing the points to be labelled as the cluster centers of k-Median (k is equal to the budget) algorithm. $vi)$\textbf{Batch Mode Discriminative-Representative Active Learning(BMDR)\citep{kdd13}:} ERM based approach which uses uncertainty and minimizes MMD between iid. samples from the dataset and the actively chosen points. $vii)$\textbf{CEAL \citep{wang2016cost}:} CEAL \citep{wang2016cost} is a weakly-supervised active learning method proposed specifically for CNNs. we include it in the weakly-supervised analysis. We conducted experiments on active learning for fully-supervised models as well as active learning for weakly-supervised models. In our experiments, we start with small set of images sampled uniformly at random from the dataset as an initial pool. The weakly-supervised model has access to labeled examples as well as unlabelled examples. The fully-supervised model only has access to the labeled data points. We run all experiments with five random initializations of the initial pool of labeled points and use the average classification accuracy as a metric. We plot the accuracy vs the number of labeled points. We also plot error bars as standard deviations. We run the query algorithm iteratively; in other words, we solve the discrete optimization problem $\min_{\mathbf{s}^{k+1} : |\mathbf{s}^{k+1}| \leq b} E_{\mathbf{x},y \sim p_\mathcal{Z}} [l(\mathbf{x},y; A_{\mathbf{s}^{0} \cup \ldots, \mathbf{s}^{k+1}})]$ for each point on the accuracy vs number of labelled examples graph. We present the results in Figures~\ref{fig:ressemi} and \ref{fig:resnosemi}. \begin{figure}[tb] \centering \begin{subfigure}[b]{\textwidth} \includegraphics[width=\textwidth]{fig_ws5.pdf} \end{subfigure} \vspace{-5mm} \caption{Results on Active Learning for Weakly-Supervised Model (error bars are std-dev)}\label{fig:ressemi} \vspace{-3mm} \label{fig:resns} %\end{figure*} %\begin{figure*}[ht] % \centering \vspace{5mm} \begin{subfigure}[b]{\textwidth} \includegraphics[width=\textwidth]{fs_fig5.pdf} \end{subfigure} \vspace{-5mm} \caption{Results on Active Learning for Fully-Supervised Model (error bars are std-dev)}\label{fig:resnosemi} \vspace{-5mm} \label{fig:ress} \end{figure} Figures~\ref{fig:resns} and \ref{fig:ress} suggests that our algorithm outperforms all other baselines in all experiments; for the case of weakly-supervised models, by a large margin. We believe the effectiveness of our approach in the weakly-supervised case is due to the better feature learning. Weakly-supervised models provide better feature spaces resulting in accurate geometries. Since our method is geometric, it performs significantly better with better feature spaces. We also observed that our algorithm is less effective in CIFAR-100 when compared with CIFAR-10 and SVHN. This can easily be explained using our theoretical analysis. Our bound over the core-set loss scales with the number of classes, hence it is better to have fewer classes. One interesting observation is the fact that a state-of-the-art batch mode active learning baseline (BMDR \citep{kdd13}) does not necessarily perform better than greedy ones. We believe this is due to the fact that it still uses an uncertainty information and soft-max probabilities are not a good proxy for uncertainty. Our method does not use any uncertainty. And, incorporating uncertainty to our method in a principled way is an open problem and a fruitful future research direction. On the other hand, a pure clustering based batch active learning baseline (k-Medoids) is also not effective. We believe this is rather intuitive since cluster sentences are likely the points which are well covered with initial iid. samples. Hence, this clustering based method fails to sample the tails of the data distribution. Our results suggest that both oracle uncertainty information and Bayesian estimation of uncertainty is helpful since they improve over empirical uncertainty baseline; however, they are still not effective in the batch setting since random sampling outperforms them. We believe this is due to the correlation in the queried labels as a consequence of active learning in batch setting. We further investigate this with a qualitative analysis via tSNE \citep{tsne} embeddings. We compute embeddings for all points using the features which are learned using the labelled examples and visualize the points sampled by our method as well as the oracle uncertainty. This visualization suggests that due to the correlation among samples, uncertainty based methods fail to cover the large portion of the space confirming our hypothesis. \noindent\textbf{Optimality of the k-Center Solution:} Our proposed method uses the greedy 2-OPT solution for the k-Center problem as an initialization and checks the feasibility of a mixed integer program (MIP). We use LP-relaxation of the defined MIP and use branch-and-bound to obtain integer solutions. The utility obtained by solving this expensive MIP should be investigated. We compare the average run-time of MIP\footnote{On Intel Core i7-5930K@3.50GHz and 64GB memory} with the run-time of 2-OPT solution in Table~\ref{tab:runtime}. We also compare the accuracy obtained with optimal k-Center solution and the 2-OPT solution in Figure~\ref{fig:twoopt} on CIFAR-100 dataset. \begin{figure*}[tb] \begin{minipage}[t]{0.49\textwidth} \begin{center} \begin{subfigure}[b]{0.49\textwidth} \includegraphics[width=\columnwidth]{fig1_a_1.pdf} \caption{Uncertainty Oracle} \end{subfigure} \begin{subfigure}[b]{0.49\textwidth} \includegraphics[width=\columnwidth]{fig1_b_1.pdf} \caption{Our Method} \end{subfigure} \end{center} \caption{tSNE embeddings of the CIFAR dataset and behavior of uncertainty oracle as well as our method. For both methods, the initial labeled pool of 1000 images are shown in blue, 1000 images chosen to be labeled in green and remaining ones in red. Our algorithm results in queries evenly covering the space. On the other hand, samples chosen by uncertainty oracle fails to cover the large portion of the space.} \end{minipage} \quad \begin{minipage}[t]{0.49\textwidth} \vspace{-50mm} \captionof{table}{Average run-time of our algorithm for $b=5k$ and $|\mathbf{s}^0|=10k$ in seconds.} \vspace{-2mm} \setlength{\tabcolsep}{1mm} \resizebox{\columnwidth}{!}{% \begin{tabular}{@{}ccccc@{}} \toprule Distance& Greedy & MIP & MIP & \\ Matrix &(2-OPT) & (iteration) & (total) & Total \\ \midrule 104.2 & 2 & 7.5 & 244.03 & 360.23 \\ \bottomrule \end{tabular}} \label{tab:runtime} \vspace{4mm} \includegraphics[width=\textwidth]{mip_100_3.pdf} \vspace{-2mm} \captionof{figure}{We compare our method with k-Center-Greedy. Our algorithm results in a small but important accuracy improvement. } \label{fig:twoopt} \end{minipage} \end{figure*} As shown in the Table~\ref{tab:runtime}; although the run-time of MIP is not polynomial in worst-case, in practice it converges in a tractable amount of time for a dataset of 50k images. Hence, our algorithm can easily be applied in practice. Figure~\ref{fig:twoopt} suggests a small but significant drop in the accuracy when the 2-OPT solution is used. Hence, we conclude that unless the scale of the dataset is too restrictive, using our proposed optimal solver is desired. Even with the accuracy drop, our active learning strategy using 2-OPT solution still outperforms the other baselines. Hence, we can conclude that our algorithm can scale to any dataset size with small accuracy drop even if solving MIP is not feasible. \section{Conclusion} We study the active learning problem for CNNs. Our empirical analysis showed that classical uncertainty based methods have limited applicability to the CNNs due to the correlations caused by batch sampling. We re-formulate the active learning problem as core-set selection and study the core-set problem for CNNs. We further validated our algorithm using an extensive empirical study. Empirical results on three datasets showed state-of-the-art performance by a large margin. \begin{thebibliography}{51} \providecommand{\natexlab}[1]{#1} \providecommand{\url}[1]{\texttt{#1}} \expandafter\ifx\csname urlstyle\endcsname\relax \providecommand{\doi}[1]{doi: #1}\else \providecommand{\doi}{doi: \begingroup \urlstyle{rm}\Url}\fi \bibitem[Abadi et~al.(2016)Abadi, Agarwal, Barham, Brevdo, Chen, Citro, Corrado, Davis, Dean, Devin, et~al.]{tensorflow} Martin Abadi, Ashish Agarwal, Paul Barham, Eugene Brevdo, Zhifeng Chen, Craig Citro, Greg~S Corrado, Andy Davis, Jeffrey Dean, Matthieu Devin, et~al. \newblock Tensorflow: Large-scale machine learning on heterogeneous distributed systems. \newblock \emph{arXiv:1603.04467}, 2016. \bibitem[Berlind \& Urner(2015)Berlind and Urner]{BerlindU15} C.~Berlind and R.~Urner. \newblock Active nearest neighbors in changing environments. \newblock In \emph{ICML}, 2015. \bibitem[Brinker(2003)]{brinker2003incorporating} Klaus Brinker. \newblock Incorporating diversity in active learning with support vector machines. \newblock In \emph{ICML}, volume~3, pp.\59--66, 2003.\bibitem[Cook et~al.(1998)Cook, Cunningham, Pulleyblank, and Schrijver]{cook} William~J Cook, William~H Cunningham, William~R Pulleyblank, and Alexander Schrijver. \newblock \emph{Combinatorial optimization}, volume 605. \newblock Springer, 1998. \bibitem[Dasgupta(2004)]{dasgupta2004analysis} Sanjoy Dasgupta. \newblock Analysis of a greedy active learning strategy. \newblock In \emph{NIPS}, 2004. \bibitem[Dasgupta(2005)]{NIPS2004_2636} Sanjoy Dasgupta. \newblock Analysis of a greedy active learning strategy. \newblock In L.~K. Saul, Y.~Weiss, and L.~Bottou (eds.), \emph{Advances in Neural Information Processing Systems 17}, pp.\337--344. MIT Press, 2005.\newblock URL \url{http://papers.nips.cc/paper/2636-analysis-of-a-greedy-active-learning-strategy.pdf}. \bibitem[Demir et~al.(2011)Demir, Persello, and Bruzzone]{demir2011batch} Beg{\"u}m Demir, Claudio Persello, and Lorenzo Bruzzone. \newblock Batch-mode active-learning methods for the interactive classification of remote sensing images. \newblock \emph{IEEE Transactions on Geoscience and Remote Sensing}, 49\penalty0 (3):\penalty0 1014--1031, 2011. \bibitem[Donahue et~al.(2016)Donahue, Kr{\"a}henb{\"u}hl, and Darrell]{bigan} Jeff Donahue, Philipp Kr{\"a}henb{\"u}hl, and Trevor Darrell. \newblock Adversarial feature learning. \newblock \emph{arXiv:1605.09782}, 2016. \bibitem[Dumoulin et~al.(2016)Dumoulin, Belghazi, Poole, Lamb, Arjovsky, Mastropietro, and Courville]{ali} Vincent Dumoulin, Ishmael Belghazi, Ben Poole, Alex Lamb, Martin Arjovsky, Olivier Mastropietro, and Aaron Courville. \newblock Adversarially learned inference. \newblock \emph{arXiv:1606.00704}, 2016. \bibitem[Elhamifar et~al.(2013)Elhamifar, Sapiro, Yang, and Shankar~Sasrty]{elhamifar2013convex} Ehsan Elhamifar, Guillermo Sapiro, Allen Yang, and S~Shankar~Sasrty. \newblock A convex optimization framework for active learning. \newblock In \emph{ICCV}, 2013. \bibitem[Freund et~al.(1997)Freund, Seung, Shamir, and Tishby]{freund1997selective} Yoav Freund, H~Sebastian Seung, Eli Shamir, and Naftali Tishby. \newblock Selective sampling using the query by committee algorithm. \newblock \emph{Machine learning}, 28\penalty0 (2-3), 1997. \bibitem[Gal \& Ghahramani(2016)Gal and Ghahramani]{gal_bayes} Yarin Gal and Zoubin Ghahramani. \newblock Dropout as a bayesian approximation: Representing model uncertainty in deep learning. \newblock In \emph{International Conference on Machine Learning}, 2016. \bibitem[Gal et~al.(2017)Gal, Islam, and Ghahramani]{gal_active} Yarin Gal, Riashat Islam, and Zoubin Ghahramani. \newblock Deep bayesian active learning with image data. \newblock \emph{arXiv preprint arXiv:1703.02910}, 2017. \bibitem[Ganti \& Gray(2012)Ganti and Gray]{ganti2012upal} Ravi Ganti and Alexander Gray. \newblock Upal: Unbiased pool based active learning. \newblock In \emph{Artificial Intelligence and Statistics}, pp.\422--431, 2012.\bibitem[Golovin \& Krause(2011)Golovin and Krause]{golovin2011adaptive} Daniel Golovin and Andreas Krause. \newblock Adaptive submodularity: Theory and applications in active learning and stochastic optimization. \newblock \emph{Journal of Artificial Intelligence Research}, 42:\penalty0 427--486, 2011. \bibitem[Gonen et~al.(2013)Gonen, Sabato, and Shalev-Shwartz]{gonen2013efficient} Alon Gonen, Sivan Sabato, and Shai Shalev-Shwartz. \newblock Efficient active learning of halfspaces: an aggressive approach. \newblock \emph{The Journal of Machine Learning Research}, 14\penalty0 (1):\penalty0 2583--2615, 2013. \bibitem[Goodfellow et~al.(2014)Goodfellow, Pouget-Abadie, Mirza, Xu, Warde-Farley, Ozair, Courville, and Bengio]{gan_original} Ian Goodfellow, Jean Pouget-Abadie, Mehdi Mirza, Bing Xu, David Warde-Farley, Sherjil Ozair, Aaron Courville, and Yoshua Bengio. \newblock Generative adversarial nets. \newblock In \emph{NIPS}, 2014. \bibitem[Guillory \& Bilmes(2010)Guillory and Bilmes]{guillory2010interactive} Andrew Guillory and Jeff Bilmes. \newblock Interactive submodular set cover. \newblock \emph{arXiv:1002.3345}, 2010. \bibitem[Guo(2010)]{guo2010} Yuhong Guo. \newblock Active instance sampling via matrix partition. \newblock In \emph{Advances in Neural Information Processing Systems}, pp.\802--810, 2010.\bibitem[Guo \& Schuurmans(2008)Guo and Schuurmans]{guo_et_al} Yuhong Guo and Dale Schuurmans. \newblock Discriminative batch mode active learning. \newblock In \emph{Advances in neural information processing systems}, pp.\593--600, 2008.\bibitem[Hanneke(2007)]{hanneke2007bound} Steve Hanneke. \newblock A bound on the label complexity of agnostic active learning. \newblock In \emph{Proceedings of the 24th international conference on Machine learning}, pp.\353--360. ACM, 2007.\bibitem[Har-Peled \& Kushal(2005)Har-Peled and Kushal]{har2005smaller} Sariel Har-Peled and Akash Kushal. \newblock Smaller coresets for k-median and k-means clustering. \newblock In \emph{Annual Symposium on Computational geometry}. ACM, 2005. \bibitem[He et~al.(2016)He, Zhang, Ren, and Sun]{he_et_al} Kaiming He, Xiangyu Zhang, Shaoqing Ren, and Jian Sun. \newblock Deep residual learning for image recognition. \newblock In \emph{Proceedings of the IEEE conference on computer vision and pattern recognition}, pp.\770--778, 2016.\bibitem[Hoi et~al.(2006)Hoi, Jin, Zhu, and Lyu]{hoi_et_al} Steven~CH Hoi, Rong Jin, Jianke Zhu, and Michael~R Lyu. \newblock Batch mode active learning and its application to medical image classification. \newblock In \emph{Proceedings of the 23rd international conference on Machine learning}, pp.\417--424. ACM, 2006.\bibitem[Inc.(2016)]{gurobi} Gurobi~Optimization Inc. \newblock Gurobi optimizer reference manual, 2016. \newblock URL \url{http://www.gurobi.com}. \bibitem[Joshi et~al.(2009)Joshi, Porikli, and Papanikolopoulos]{joshi2009multi} Ajay~J Joshi, Fatih Porikli, and Nikolaos Papanikolopoulos. \newblock Multi-class active learning for image classification. \newblock In \emph{CVPR}, 2009. \bibitem[Joshiy et~al.(2010)Joshiy, Porikli, and Papanikolopoulos]{porikli} A.~J. Joshiy, F.~Porikli, and N.~Papanikolopoulos. \newblock Multi-class batch-mode active learning for image classification. \newblock In \emph{2010 IEEE International Conference on Robotics and Automation}, pp.\1873--1878, May 2010.\newblock \doi{10.1109/ROBOT.2010.5509293}. \bibitem[Kapoor et~al.(2007)Kapoor, Grauman, Urtasun, and Darrell]{kapoor2007active} Ashish Kapoor, Kristen Grauman, Raquel Urtasun, and Trevor Darrell. \newblock Active learning with gaussian processes for object categorization. \newblock In \emph{ICCV}, 2007. \bibitem[Krizhevsky \& Hinton(2009)Krizhevsky and Hinton]{cifar} Alex Krizhevsky and Geoffrey Hinton. \newblock Learning multiple layers of features from tiny images. \newblock 2009. \bibitem[Li \& Guo(2013)Li and Guo]{li2013adaptive} Xin Li and Yuhong Guo. \newblock Adaptive active learning for image classification. \newblock In \emph{CVPR}, 2013. \bibitem[Maaten \& Hinton(2008)Maaten and Hinton]{tsne} Laurens van~der Maaten and Geoffrey Hinton. \newblock Visualizing data using t-sne. \newblock \emph{Journal of Machine Learning Research}, 9\penalty0 (Nov):\penalty0 2579--2605, 2008. \bibitem[MacKay(1992)]{mackay1992information} David~JC MacKay. \newblock Information-based objective functions for active data selection. \newblock \emph{Neural computation}, 4\penalty0 (4):\penalty0 590--604, 1992. \bibitem[McCallumzy \& Nigamy(1998)McCallumzy and Nigamy]{mccallumzy1998employing} Andrew~Kachites McCallumzy and Kamal Nigamy. \newblock Employing em and pool-based active learning for text classification. \newblock In \emph{ICML}, 1998. \bibitem[Netzer et~al.(2011)Netzer, Wang, Coates, Bissacco, Wu, and Ng]{svhn} Yuval Netzer, Tao Wang, Adam Coates, Alessandro Bissacco, Bo~Wu, and Andrew~Y Ng. \newblock Reading digits in natural images with unsupervised feature learning. \newblock In \emph{NIPS workshop on deep learning and unsupervised feature learning}, volume 2011, pp.\~5, 2011. \bibitem[Radford et~al.(2015)Radford, Metz, and Chintala]{dcgan} Alec Radford, Luke Metz, and Soumith Chintala. \newblock Unsupervised representation learning with deep convolutional generative adversarial networks. \newblock \emph{arXiv:1511.06434}, 2015. \bibitem[Rasmus et~al.(2015)Rasmus, Berglund, Honkala, Valpola, and Raiko]{ladder} Antti Rasmus, Mathias Berglund, Mikko Honkala, Harri Valpola, and Tapani Raiko. \newblock Semi-supervised learning with ladder networks. \newblock In \emph{NIPS}, 2015. \bibitem[Roy \& McCallum(2001)Roy and McCallum]{roy2001toward} Nicholas Roy and Andrew McCallum. \newblock Toward optimal active learning through monte carlo estimation of error reduction. \newblock \emph{ICML}, 2001. \bibitem[Salimans et~al.(2016)Salimans, Goodfellow, Zaremba, Cheung, Radford, and Chen]{salimans2016improved} Tim Salimans, Ian Goodfellow, Wojciech Zaremba, Vicki Cheung, Alec Radford, and Xi~Chen. \newblock Improved techniques for training gans. \newblock In \emph{NIPS}, 2016. \bibitem[Settles(2010)]{settles2010active} Burr Settles. \newblock Active learning literature survey. \newblock \emph{University of Wisconsin, Madison}, 52\penalty0 (55-66):\penalty0 11, 2010. \bibitem[Simonyan \& Zisserman(2014)Simonyan and Zisserman]{vgg} Karen Simonyan and Andrew Zisserman. \newblock Very deep convolutional networks for large-scale image recognition. \newblock \emph{arXiv:1409.1556}, 2014. \bibitem[Stark et~al.(2015)Stark, Haz{\i}rbas, Triebel, and Cremers]{captcha} Fabian Stark, Caner Haz{\i}rbas, Rudolph Triebel, and Daniel Cremers. \newblock Captcha recognition with active deep learning. \newblock In \emph{GCPR Workshop on New Challenges in Neural Computation}, 2015. \bibitem[Tong \& Koller(2001)Tong and Koller]{tong2001support} Simon Tong and Daphne Koller. \newblock Support vector machine active learning with applications to text classification. \newblock \emph{JMLR}, 2\penalty0 (Nov):\penalty0 45--66, 2001. \bibitem[Tsang et~al.(2005)Tsang, Kwok, and Cheung]{tsang2005core} Ivor~W Tsang, James~T Kwok, and Pak-Ming Cheung. \newblock Core vector machines: Fast svm training on very large data sets. \newblock \emph{JMLR}, 6\penalty0 (Apr):\penalty0 363--392, 2005. \bibitem[Wang et~al.(2016)Wang, Zhang, Li, Zhang, and Lin]{wang2016cost} Keze Wang, Dongyu Zhang, Ya~Li, Ruimao Zhang, and Liang Lin. \newblock Cost-effective active learning for deep image classification. \newblock \emph{Transactions on Circuits and Systems for Video Technology}, 2016. \bibitem[Wang \& Ye(2015)Wang and Ye]{kdd13} Zheng Wang and Jieping Ye. \newblock Querying discriminative and representative samples for batch mode active learning. \newblock \emph{ACM Transactions on Knowledge Discovery from Data (TKDD)}, 9\penalty0 (3):\penalty0 17, 2015. \bibitem[Wei et~al.(2013)Wei, Liu, Kirchhoff, and Bilmes]{wei2013using} Kai Wei, Yuzong Liu, Katrin Kirchhoff, and Jeff~A Bilmes. \newblock Using document summarization techniques for speech data subset selection. \newblock In \emph{HLT-NAACL}, 2013. \bibitem[Wei et~al.(2015)Wei, Iyer, and Bilmes]{wei2015submodularity} Kai Wei, Rishabh~K Iyer, and Jeff~A Bilmes. \newblock Submodularity in data subset selection and active learning. \newblock In \emph{ICML}, 2015. \bibitem[Wolf(2011)]{facility} Gert~W Wolf. \newblock Facility location: concepts, models, algorithms and case studies., 2011. \bibitem[Xu \& Mannor(2012)Xu and Mannor]{robust} Huan Xu and Shie Mannor. \newblock Robustness and generalization. \newblock \emph{Machine learning}, 86\penalty0 (3):\penalty0 391--423, 2012. \bibitem[Yang et~al.(2015)Yang, Ma, Nie, Chang, and Hauptmann]{yang2015multi} Yi~Yang, Zhigang Ma, Feiping Nie, Xiaojun Chang, and Alexander~G Hauptmann. \newblock Multi-class active learning by uncertainty sampling with diversity maximization. \newblock \emph{International Journal of Computer Vision}, 113\penalty0 (2):\penalty0 113--127, 2015. \bibitem[Yu et~al.(2006)Yu, Bi, and Tresp]{yu2006active} Kai Yu, Jinbo Bi, and Volker Tresp. \newblock Active learning via transductive experimental design. \newblock In \emph{Proceedings of the 23rd international conference on Machine learning}, pp.\1081--1088. ACM, 2006.\end{thebibliography} \appendix \section{Proof for Lemma 1} \begin{proof} We will start with showing that softmax function defined over $C$ class is $\frac{\sqrt{C-1}}{C}$-Lipschitz continuous. It is easy to show that for any differentiable function \mbox{$f:\mathbb{R}^n\rightarrow\mathbb{R}^m$}, \[ \left \| f(\mathbf{x})-f(\mathbf{y})\right \|_2 \leq \left \|J\right \|^*_F \left\| \mathbf{x}-\mathbf{y}\right\|_2 \, \, \forall \mathbf{x},\mathbf{y}\in\mathbb{R}^n \] where $\left \|J\right \|^*_F = \max\limits_{\mathbf{x}} \left \|J\right \|_F$ and $J$ is the Jacobian matrix of $f$. Softmax function is defined as \[ f(x)_i = \frac{\exp(x_i)}{\sum\limits_{j=1}^{C}\exp(x_j)}, \, i={1,2,...,C} \] For brevity, we will denote $f_i(x)$ as $f_i$. The Jacobian matrix will be, \[ J = \begin{bmatrix} f_1(1-f_1) & -f_1f_2 & ... & -f_1f_C \\ -f_2f_1 & f_2(1-f_2) & ... & -f_2f_C \\ ... & ... & ... & ... \\ -f_{C}f_{1} & -f_{C}f_{2} & ... & -f_{C}(1-f_{C}) \end{bmatrix} \] Now, Frobenius norm of above matrix will be, \[ \left \| J \right \|_F = \sqrt{\sum\limits_{i=1}^{C}\sum\limits_{j=1, i\neq j}^{C}f_{i}^{2}f_{j}^{2} + \sum\limits_{i=1}^{C} f_i^2(1-f_i)^2} \] It is straightforward to show that $f_i = \frac{1}{C}$ is the optimal solution for $\left \| J \right \|^{*}_F = \max\limits_{x}\left \| J \right \|_F $ Hence, putting $f_i = \frac{1}{C}$ in the above equation , we get \mbox{$\left \| J \right \|^{*}_F = \frac{\sqrt{C-1}}{C}$}. Now, consider two inputs $\mathbf{x}$ and $\mathbf{\tilde{x}}$, such that their representation at layer $d$ is $\mathbf{x}^d$ and $\mathbf{\tilde{x}}^d$. Let's consider any convolution or fully-connected layer as $\mathbf{x}^d_j = \sum_i w_{i,j}^d \mathbf{x}^{d-1}_i$. If we assume, \mbox{$\sum_i |w_{i,j}| \leq \alpha \quad \forall i,j,d$}, for any convolutional or fully connected layer, we can state: \[ \|\mathbf{x}^d - \mathbf{\tilde{x}}^d\|_2 \leq \alpha \|\mathbf{x}^{d-1} - \mathbf{\tilde{x}}^{d-1}\|_2 \] On the other hand, using $|a-b| \leq |\max(0, a) - \max(0,a)|$ and the fact that max pool layer can be written as a convolutional layer such that only one weight is $1$ and others are $0$, we can state for ReLU and max-pool layers, \[ \|\mathbf{x}^d - \mathbf{\tilde{x}}^d\|_2 \leq \|\mathbf{x}^{d-1} - \mathbf{\tilde{x}}^{d-1}\|_2 \] Combining with the Lipschitz constant of soft-max layer, \[ \|CNN(\mathbf{x};\mathbf{w}) - CNN(\mathbf{\tilde{x}};\mathbf{w})\|_2 \leq \frac{\sqrt{C-1}}{C} \alpha^{n_c+n_{fc}} \|\mathbf{x}-\mathbf{\tilde{x}}\|_2 \] Using the reverse triangle inequality as \[ |l(\mathbf{x},y;\mathbf{w})-l(\mathbf{\tilde{x}},y;\mathbf{w})| = |\| CNN(\mathbf{x};\mathbf{w}) -y\|_2 -\|CNN(\mathbf{\tilde{x}};\mathbf{w})-y\|_2 | \leq \|CNN(\mathbf{x};\mathbf{w}) - CNN(\mathbf{\tilde{x}};\mathbf{w})\|_2, \] we can conclude that the loss function is $\frac{\sqrt{C-1}}{C} \alpha^{n_c+n_{fc}}$-Lipschitz for any fixed $y$ and $\mathbf{w}$. \end{proof} \section{Proof for Theorem 1} Before starting our proof, we state the Claim 1 from \cite{BerlindU15}. Fix some $p,p^\prime \in [0,1]$ and $y^\prime \in \{0,1\}$. Then, \[ p_{y \sim p}(y \neq y^\prime) \leq p_{y \sim p^\prime}(y \neq y^\prime) + |p - p^\prime| \] \begin{proof} We will start our proof with bounding $E_{y_i \sim \eta(\mathbf{x}_i)}[l(\mathbf{x}_i,y_i; A_{\mathbf{s}})]$. We have a condition which states that there exists and $\mathbf{x}_j$ in $\delta$ ball around $\mathbf{x}_i$ such that $\mathbf{x}_j$ has $0$ loss. \[ \begin{aligned} E_{y_i \sim \eta(\mathbf{x}_i)}[l(\mathbf{x}_i,y_i; A_{\mathbf{s}})] &= \sum_{k\in [C]} p_{y_i \sim \eta_k(\mathbf{x}_i)}(y_i = k) l(\mathbf{x}_i,k; A_{\mathbf{s}}) \\ &\overset{(d)}{\leq} \sum_{k\in [C]} p_{y_i \sim \eta_k(\mathbf{x}_j)}(y_i = k) l(\mathbf{x}_i, k; A_{\mathbf{s}}) + \sum_{k\in [C]} |\eta_k(\mathbf{x}_i)-\eta_k(\mathbf{x}_j)| l(\mathbf{x}_i, k; A_{\mathbf{s}}) \\ &\overset{(e)}{\leq} \sum_{k\in [C]} p_{y_i \sim \eta_k(\mathbf{x}_j)} (y_i = k) l(\mathbf{x}_i,k; A_{\mathbf{s}}) + \delta \lambda^\eta L C\\ \end{aligned} \] With abuse of notation, we represent \mbox{$\{y_i=k\} \sim \eta_k(\mathbf{x}_i)$} with \mbox{$y_i \sim \eta_k(\mathbf{x}_i)$}. We use Claim 1 in $(d)$, and Lipschitz property of regression function and bound of loss in $(d)$. Then, we can further bound the remaining term as; \[ \begin{aligned} \sum_{k\in [C]} p_{y_i \sim \eta_k(\mathbf{x}_j)} (y_i = k) l(\mathbf{x}_i,k; A_{\mathbf{s}}) =& \sum_{k\in [C]} p_{y_i \sim \eta_k(\mathbf{x}_j)} (y_i = k) [l(\mathbf{x}_i,k; A_{\mathbf{s}}) - l(\mathbf{x}_j,k; A_{\mathbf{s}}) ] \\ &\quad+ \sum_{k\in [C]} p_{y_i \sim \eta_k(\mathbf{x}_j)} (y_i = k) l(\mathbf{x}_j,k; A_{\mathbf{s}}) \\ &\leq \delta \lambda^l \end{aligned} \] where last step is coming from the fact that the trained classifier assumed to have $0$ loss over training points. If we combine them, \[ E_{y_i \sim \eta(\mathbf{x}_i)}[l(\mathbf{x}_i,y_i,A_{\mathbf{s}})] \leq \delta( \lambda^l+\lambda^\mu LC) \] We further use the Hoeffding's Bound and conclude that with probability at least $1 - \gamma$, \[ \left| \frac{1}{n}\sum_{i \in [n]} l(\mathbf{x}_i,y_i; A_{\mathbf{s}}) - \frac{1}{|\mathbf{s}|}\sum_{j \in \mathbf{s}} l(\mathbf{x}_j,y_j;A_{\mathbf{s}}) \right| \leq \delta (\lambda^l + \lambda^\mu LC)+ \sqrt{\frac{L^2 \log(1/\gamma)}{2n}} \] \end{proof} \end{document} }
\caption{A typical search and reconnaissance scenario with an internal semi-autonomous agent (robot) and an external supervisor (human) -- a video demonstration can be accessed at \textcolor{blue}{\url{https://www.youtube.com/watch?v=u_t1TQotzo4}}. }
\caption{Comparison of the PR/SR score with state-of-the-art trackers in the OPE based on the 11 sequence attributes: illumination variation (IV), scale variation (SV), occlusion (OCC), deformation (DEF), motion blur (MB), fast motion (FM), in-plane rotation (IPR), out-of-plane rotation (OPR), out-of-view (OV), background cluttered (BC) and low resolution (LR). The best and the second best results are shown in {\color{red} \textbf{red}} and {\color{blue}\textbf{blue}} colours respectively.}
\caption{The results of VOT2014 baseline and region-noise experiments with and without-re-initialization. The best and the second best results are shown in {\color{red} \textbf{red}} and {\color{blue}\textbf{blue}} colours respectively.}
\caption{(Colour online) The maximum power (in units of $\kappa_b\hbar\omega_a$) with respect to average number of excitations $\bar{n}_{\text{h}}$ in the effective hot bath. The black solid, red dotted, and blue dashed curves indicate the results of the semiclassical, classical, and quantum calculations{\color{red}.} All the other parameters are as explained in the text. }
\caption{The stacked, weighted 2D $\kappa$ profile recovered from the analysis of CMB temperature data around 3697 \redmap{} clusters. Each pixel in the cutout is 0.5 arcminutes on a side.}
\caption{The azimuthally averaged $\kappa$ profile recovered from measurement of CMB lensing around 3697 \redmap{} clusters detected in DES Y1 data (blue points with errorbars). The errorbars shown are the diagonal elements of the full covariance determined from a jackknife resampling of the cluster sample; there is significant covariance between adjacent data points. The orange region indicates the allowed range of model predictions (68\% confidence region) given the results of the fit for parameters of the cluster mass-richness relation.}
\caption[Caption for multi posterior figure]{Constraints on the amplitude, $A$, and richness scaling, $\alpha$, of the \redmap{} mass-richness relation obtained from fits to CMB lensing measurements. Contours represent $1\sigma$ and $2\sigma$ levels. The blue curve in the panel at right shows the prior on $\alpha$, which dominates our constraint. Numerical results are summarized in Table~\ref{tab:mr_parameters}.}
\caption[Caption for mass-richness comparison figure]{Comparison of constraints obtained on the \redmap{} mass-richness relation from CMB cluster lensing (this work) and cluster-galaxy lensing (\citealt{Melchior:2017} and \citealt{Simet:2017}). The solid band and lines illustrate the $1\sigma$ ranges allowed by the different constraints. Note that unlike the \citet{Melchior:2017} and \citet{Simet:2017} analyses, this work imposes an informative prior on the slope of the mass-richness relation.}
\caption{Constraints on the shifts $\Delta z^i$ applied to the \metacal\$n_{\rm PZ}(z)$ distributions for the weak lensing source galaxies are plotted for three different validation techniques. Shifts derived from resampling the COSMOS 30-band redshifts are described in this paper, and agree well with those derived (for bins 1--3 only) using angular correlations between the source population and redMagic galaxies (WZ) by \citet{xcorr} (COSMOS constraints plotted here have been expanded as per Appendix~\ref{correlationappendix} to include the effects of poorly known correlation between bins). These are also consistent with the weak lensing shear ratio tests conducted by \citet{gglpaper}. The final validation constraints on $\Delta z^i$ are taken as the combination of the COSMOS and WZ results for each redshift bin (where available), and yield the 68\% confidence intervals denoted by the \magenta{black points and error bars} in the 1-d marginal plots. The dashed lines at $\Delta z^i=0$ indicate no mean shift from the BPZ posteriors---the validation processes yield shifts that are non-zero at $\approx 1\sigma$ level. }
\caption{The point $\widetilde{p}(\widetilde{t}_0) \approx (0.670125, 0.571734, 0.473343)$ in $\mathcal{A}$ maximally far from the curves corresponding to the isosceles and right triangles. The common distance to the boundary of the region $\mathcal{A}$ is $\arcsin \left(\frac{\widetilde{\alpha}}{1+\widetilde{\alpha}^2}\right)\approx 0.069629$. Permuting the coordinates of $\widetilde{p}(\widetilde{t}_0)$ yields a ring of six points equidistant from the equilateral triangle point $\left(\frac{1}{\sqrt{3}},\frac{1}{\sqrt{3}},\frac{1}{\sqrt{3}}\right)$. The difference is too small to see, but the radius of the circle around the equilateral point ({\tiny $\blacksquare$}) in the right figure is smaller than that of the inscribed circle around $\widetilde{p}(\widetilde{t}_0)$ by $\approx 0.000333$.} \label{fig:inscribed5} \end{figure} The distance from $\widetilde{p}(t)$ to the boundary great circle $x-y=0$ is $\arcsin\left(\frac{\sin t}{2}\right)$, and the distance to the point $q(x)$ on the curve of right triangles is \[ \widetilde{d}(t,x) = \arccos\left(\frac{\sqrt{\frac{1-x^2}{x^2+1}} \cos t}{\sqrt{3}}+x \sqrt{\frac{1-x^2}{x^2+1}} \left(\frac{\cos t}{\sqrt{3}}-\frac{\sin t}{\sqrt{2}}\right)+x \left(\frac{\sin t}{\sqrt{2}}+\frac{\cos t}{\sqrt{3}}\right)\right). \] Hence, the challenge is to determine the smallest value of $t$ for which $\widetilde{d}(t,x) = \arcsin\left(\frac{\sin t}{2}\right)$. Again, we solve this equation for $x$, and then find the smallest $t$ which makes $x$ real, which is \[ \widetilde{t}_0 = 2 \arctan \widetilde{\alpha} \] where $\widetilde{\alpha} \approx 0.069912$ is the smallest positive root of the even, palindromic polynomial \begin{align*} & 131072 z^{52}-30081024 z^{50}+715784192 z^{48}-10181738496 z^{46}+83609604096 z^{44}\\ & -443259328512 z^{42}+1410471953408 z^{40}-1858643071488 z^{38}+18137673285920 z^{36}\\ & -14367112128688 z^{34}+56162265469488 z^{32}-73041229883512 z^{30}+73382345772378 z^{28}\\ & -122601623733111 z^{26}+73382345772378 z^{24}-73041229883512 z^{22}+56162265469488 z^{20}\\ & -14367112128688 z^{18}+18137673285920 z^{16}-1858643071488 z^{14}+1410471953408 z^{12}\\ & -443259328512 z^{10}+83609604096 z^8-10181738496 z^6+715784192 z^4-30081024 z^2+131072. \end{align*} Again, more details are in the supplementary materials~\cite{supplementary-materials}. Therefore, the incenter of $\mathcal{A}$ is \[ \widetilde{p}(\widetilde{t}_0) \approx (0.670125, 0.571734, 0.473343) \] and we can convert to side lengths to conclude: \begin{proposition}\label{prop:acute} The least symmetric acute triangle has side lengths \begin{multline*} \frac{2}{3(1+\widetilde{\alpha}^2)^2}\left(1-\sqrt{6}\widetilde{\alpha}+\widetilde{\alpha}^2+\sqrt{6}\widetilde{\alpha}^3+\widetilde{\alpha}^4,1+4\widetilde{\alpha}^2+\widetilde{\alpha}^4,1+\sqrt{6}\widetilde{\alpha}+\widetilde{\alpha}^2-\sqrt{6}\widetilde{\alpha}^3+\widetilde{\alpha}^4\right)\\ \approx (0.550933, 0.673120, 0.775946). \end{multline*} \end{proposition} \begin{figure}[htbp] \centering \includegraphics[height=1.4in]{mostscalene4.pdf} \caption{The least symmetric acute triangle, with side lengths $\approx (0.550933, 0.67312, 0.775946)$.} \label{fig:mostacute} \end{figure} % subsection acute (end) \section{Conclusion and open questions} The emphasis in the chemistry literature is on finding \emph{chirality measures} whose maxima are considered the most chiral triangles. Of course, our least symmetric triangle from Proposition~\ref{prop:nondegenerate} can also be thought of in this way: it maximizes the minimum distance to the 9 great circles representing isosceles and degenerate triangles. Since this function is continuous but only piecewise smooth, a natural question to ask is: does there exist a reasonable smooth function on the sphere which vanishes precisely at the isosceles and degenerate triangles and which is maximized on the hyperoctahedral group orbit of the point $\frac{1}{\sqrt{13+6 \sqrt{2}}} \left(1+2 \sqrt{2},1+\sqrt{2},1\right)$?\footnote{Rassat and Fowler~\cite{Rassat:2003jn} showed that \emph{any} non-isosceles triangle is the most chiral triangle according to some chirality measure in a particular family, but their chirality measures are slightly unnatural to define on the sphere.} And similarly for the other special points we have found? At the very least, we expect, based on numerical experiments and O'Hara's results on planar convex bodies~\cite{OHara:2012fm}, that these points are the limits of maxima of suitably renormalized Riesz-type potentials as the exponent goes to $-\infty$. The problem of extending this analysis to $n$-gons is substantial and interesting. As alluded to in the introduction, the spherical tiling induced by the hyperoctahedral group action on the sphere (Figure~\ref{fig:fundamental domains}) generalizes to a decomposition of $n$-gon space for any $n$, namely the images of the fundamental domain of the standard hyperoctahedral group action on $G_2(\R^n)$. In Section~\ref{sec:non-degenerate} we found the incenter of $\mathcal{T}$ and hence, by taking this point's hyperoctahedral group orbit, of all the other triangles in the tiling. What are the incenters of the cells in the decomposition of $n$-gon space, which we can interpret as the least symmetric $n$-gon? Similarly, what is the least symmetric $n$-gon in $\R^3$? Just as $n$-gons in the plane are modeled by $G_2(\R^n)$, $n$-gons in space correspond to points in the complex Grassmannian $G_2(\C^n)$. The natural generalization of the hyperoctahedral group is given by replacing $(\Z/2)^n = (O(1))^n$ with $(U(1))^n$, yielding the group $(U(1))^n \rtimes S_n$, which acts on $G_2(\C^n)$. The fundamental domain (in the sense of Hermann~\cite{Hermann:1962eu}) of this action gives the space of unordered $n$-gons in $\R^3$, and it seems challenging to describe this domain and then to find the point furthest from the boundary. Finding the least symmetric triangle can be interpreted as finding the optimal shape satisfying certain constraints. As in Section~\ref{sec:obtuse}, where we added the constraints that triangles should be obtuse or acute, other constraints are also interesting; for example, what is the most knotted\footnote{Meaning furthest from the subset of unknotted polygons.} trefoil knot in the space of $23$-gons? Since our polygon model has generalizations to continuous curves in the plane~\cite{Younes:2008gy} and to framed curves in space~\cite{Needham:2017vn}, we are not restricted to asking such questions only about polygons. Finally, while we have presented the triangle from Proposition~\ref{prop:nondegenerate} as the least symmetric triangle, we could also think of it as the \emph{median} triangle since it is equidistant from the three distinct pieces of the boundary of the region $\mathcal{T}$ parametrizing unordered triangles. This framing suggests the obvious question: what is the \emph{mean} triangle? Or, indeed, the mean $n$-gon? We intend to address this question in future work. \section*{Acknowledgments} We would like to thank Vance Blankers, Jason Cantarella, Renzo Cavalieri, Andy Fry, Tom Needham, Eric Rawdon, and Gavin Stewart for stimulating conversations about the geometry of polygon space. We are especially grateful to Noah Otterstetter for his participation in our early conversations about this project. This work was supported by a grant from the Simons Foundation (\#354225, CS).\bibliography{scalene-special,papers-export} \end{document} }
\caption{The F-measure and MAE of different saliency detection methods on six large-scale saliency detection datasets. The best three results are shown in \textcolor[rgb]{1,0,0}{red},~\textcolor[rgb]{0,1,0}{green} and \textcolor[rgb]{0,0,1}{blue}. The proposed methods rank first or second on these datasets.}
\caption{The F-measure and MAE of different saliency detection methods on five frequently used datasets. The best three results are shown in \textcolor[rgb]{1,0,0}{red},~\textcolor[rgb]{0,1,0}{green} and \textcolor[rgb]{0,0,1}{blue}, respectively. The proposed methods rank first and second on these datasets.}
\caption{{\color{blue}CBC-GWRM $u(t,x)$ (\textit{left}) and exact solution $u_e(t,x)$ (\textit{right})} of the forced wave equation in the range $t=[0,80]$. CBC-GWRM parameters are $N_s=5$, $K=12$, and $L=5$.}
\caption{Pose estimation results in Leeds Sports Pose dataset. First images are from test set with the superimposed ground truth skeleton depicted in {\color{red} red} and the predicted skeleton in {\color{green} green}. Second images are corresponding nearest neighbors.}
\caption{\small Prior network structures (a,b) and our memory block (c). The \textcolor{blue}{blue circles} denote a recursive unit with an unfolded structure which generates the short-term memory. The \textcolor{green}{green arrow} denotes the long-term memory from the previous memory blocks that is directly passed to the gate unit.}
\caption{\small Basic MemNet architecture. The \textcolor{Red}{red dashed box} represents multiple stacked memory blocks.}
\caption{\small Multi-supervised MemNet architecture. The outputs with \textcolor{Purple}{purple} color are supervised. }
\caption{\small Ablation study on effects of long-term and short-term connections. Average PSNR/SSIMs for the scale factor $\times2$, $\times3$ and $\times4$ on dataset Set$5$. \textcolor{red}{Red} indicates the best performance.}
\caption{\small Benchmark image denoising results. Average PSNR/SSIMs for noise level $30$, $50$ and $70$ on $14$ images and BSD$200$. \textcolor{red}{Red} color indicates the best performance and \textcolor{blue}{blue} color indicates the second best performance.}
\caption{\label{Figure:OmegaResultsEnergyBins}(Color online) Measurement of the helicity asymmetry~$E$ in the reaction $\gamma p\to p\omega$ using a circularly-polarized photon beam and a longitudinally-polarized target. The data are shown in 100-MeV-wide bins for the photon energy range $E_\gamma\in [1.1,\,2.3]$~GeV. The CLAS-FROST results (red circles {\color{red} $\bullet$}) are compared with results from the CBELSA/TAPS Collaboration~\cite{Eberhardt:2015lwa}, which used the radiative decay mode, $\omega\to \pi^0\gamma$~(blue squares {\tiny\color{blue} $\blacksquare$}). The black solid line represents the BnGa PWA solution. The data points include statistical uncertainties only; the total systematic uncertainty is given as bands at the bottom of each distribution.}
\caption{\label{Figure:OmegaResultsAngleBins}(Color online) Measurement of the helicity asymmetry~$E$ in the reaction $\gamma p\to p\omega$ using a circularly-polarized photon beam and a longitudinally-polarized target. The data are shown in 0.25-wide-bins in cos\,$\theta^{\,\omega}_{\rm c.m.}$ and in 100-MeV-wide bins for the photon energy range $E_\gamma\in [1.1,\,2.3]$~GeV. The CLAS-FROST results (red circles {\color{red} $\bullet$}) are compared with results from the CBELSA/TAPS Collaboration~\cite{Eberhardt:2015lwa}, which used the radiative decay mode, $\omega\to \pi^0\gamma$~(blue squares {\tiny\color{blue} $\blacksquare$}). The black solid line represents the BnGa PWA solution. The data points include statistical uncertainties only; the total systematic uncertainty is given as bands at the bottom of each distribution.}
\caption[]{The plot {\tikz[baseline=-0.5ex]\draw[fill=black,radius=1.5pt] (0,0) circle ;} for ${\rm Prob}(\chi(\mu)>0 \mid \chi(\mu)\neq 0)$ and the plot {\tikz[baseline=-0.5ex]\draw[fill=white,radius=1.5pt] (0,0) circle ;} for ${\rm Prob}(\chi(\mu)<0 \mid \chi(\mu)\neq 0)$ where $1\leq n\leq 38$.% }
\caption{\textbf{Effect of inter-group interaction rate on community collapse.} \tiny There is no individualization tendency in this experiment ($s = 0$). In \ref{fig:2dClusterUppEffectNoUniqNormal} the dashed line shows that at $u_b = 0.01$ in almost half of experiments identity loss occurred. In \ref{fig:2dTimeUppEffectNoUniqNormal} for those values of $u_b$ which we have red {\color{red}{x}}, communities did not loose their identities and did not collapse over the experiment run time (30,000 pairwise interactions).}
\caption{a) Stark shift canceling Raman scheme. The hyperfine ground states are coupled through two virtual levels (dotted lines) located symmetrically around the excited state manifold. The coupling fields are created by phase-modulation of a {\color{col3}carrier field}. By choosing the carrier detuning, $\Delta\approx\Dhfs/2$ and the ratio of the sideband to carrier powers $A= (J_1(z)/J_0(z))^2\approx 1.5$, the ac-Stark shifts are canceled. A $\pi$ phaseshift of the {\color{col1}red-detuned sideband} compared to the {\color{col0}blue-detuned sideband} ensures constructive interference of the Raman coupling amplitudes. b) Intensity of \SI{1}{\milli \watt} of light ($\lambda=\SI{852}{nm}$) guided by a nanofiber ($\diameter=\SI{470}{nm}$) as a function of distance to the fiber axis, together with an exponential fit (dotted line). c) Trapping potential for Cs atoms in the electronic ground state. See main text for details.}
\caption{An example of how intra-library collusion (ILC) could happen in practice. Libraries are able to use permissions in \textcolor{blue}{blue} because they have been granted to the app, while permissions in \textcolor{red}{red} are unavailable for libraries within that app. Overall, \texttt{library-2} is able to access a total of four permissions on the device.}
\caption{\textcolor{red}{Schematic illustration of crater relaxation caused by meteorite-impact-induced seismic shaking on small astronomical objects.}}
\caption{Comparison between profiles taken by a laser profiler and those computed by the NDT model. The initial shape for the model computation is approximated by \textcolor{red}{$h-h_0\sim({r_0}^{1.6}-r^{1.6})$} with $r_0=40$~mm. \textcolor{red}{The time step and spatial resolution of the numerical calculation are $10^{-5}$ s and $0.5$ mm, respectively.} Experimental and computed curves agree well in relatively steep regimes. When the shape of the pile get flattened to some extent, granular media reach a jammed state, where the relaxation almost halts.}
\caption{Screen captures of oscilloscope displaying electric signal from homodyne detector and quadrature distribution of cat state of phases corresponded to anti-squeezing and squeezing. The number of the overlaid events is 10,000 and electric signal is recorded 200 ns around the timing of photon detection. For anti-squeezing: (a) Screen capture of post processing measurement. (b) Screen capture of real-time measurement. White triangle marks represent the timing of the generation of cat states ($\xi=0.25$). The coloring represents frequency of the distribution at each time and change from blue to green, yellow and red as the frequency increase. (c) Correlation plot between quadrature of post processing and real-time measurement. (d) Quadrature distribution of post processing. (e) Quadrature distribution of real-time measurement (This corresponds to the histogram of electric signal at the white triangle mark in (b)) . The correspoding figures for squeezing are (f-j). (See \textcolor{blue}{Visualization 1} for the screen captures of real-time measurement for phases in between.)}
\caption{\textcolor{blue}{(Color online)} The calculated enthalpy of $P4/nmm$ HBr relative to the $C2/m$ HBr under pressure, given by the PBE-D2 and PBEsol-D2 functionals, (a) without and (b) with considering the ZPE, respectively. The inset shows the difference of Gibbs free energy of $P4/nmm$ HBr relative to $C2/m$ HBr at different temperature and pressure, with PBE-D2 functional. (c) From $P4/nmm$ HBr to $C2/m$ phase, with bond breaking and interlayer sliding. Once the hydrogen atom moves away from the center of the HBr$_4$ tetrahedron, half of the H-Br bonds in the $P4/nmm$ phase will be broken, which leads to forming a 1D chain-like structure. }
\caption{\textcolor{blue}{(Color online)} (a) Calculated band structures, projected density of states (DOS), and (b) Fermi surface of $P4/nmm$ HBr at 170 GPa. (c) The contour plot of Fermi surface at the plane of $k_z$ = 0. }
\caption{\textcolor{blue}{(Color online)} Phonon dispersion curves, Eliashberg spectral functions $\alpha^2F$($\omega$) together with the electron-phonon integral $\lambda$($\omega$) and phonon density of states (PHDOS) for $P4/nmm$-HBr, $P4mm$-HCl$_{0.5}$Br$_{0.5}$ at 170 GPa, and $P4/nmm$-HI at 150 GPa. The phonon linewidth $\gamma_{q,j}$($\omega$) of each mode (q,j) is illustrated by the size of pink circles along the phonon dispersions. }
\caption{\textcolor{blue}{(Color online)} The calculated band structures and density of states (DOS) for the $P4/nmm$ structure of (a) HBr at 170 GPa, (b) HI at 150 GPa and (d) HF at 150 GPa. (c) DOS for the $P4mm$ structure of HCl$_{0.5}$Br$_{0.5}$ at 170 GPa. }
\caption{\textcolor{blue}{(Color online)} Contour plots of electron localization function (ELF) on the (100) plane along the (a) H-Br and (b) H-Cl bonds, respectively, in the $P4mm$ HCl$_{0.5}$Br$_{0.5}$ at 170 GPa. (c) Crystal structure of $P4mm$ HCl$_{0.5}$Br$_{0.5}$. The red, green, and blue balls represent the hydrogen, chlorine, and bromine atoms, respectively. }
\caption{\textcolor{blue}{(Color online)} The vibrational modes at the $M$ point for the lowest twofold degenerate frequency of (a), (b) $P4/nmm$ HBr and (c), (d) $P4mm$ HCl$_{0.5}$Br$_{0.5}$. They belong to a 2D $M_3$ irreducible representation. }
\caption{\textcolor{blue}{(Color online)} $T_c$ (red solid line), the integral EPC parameter $\lambda$ (blue solid line), and the logarithmically averaged phonon frequency $\omega_{log}$ of $P4mm$-HCl$_{0.5}$Br$_{0.5}$, $P4/nmm$-HBr, and $P4/nmm$-HI versus pressure.}
\caption{Conceptual model of particle dynamics, on which the numerical model is based. Nanophytoplankton and diatoms (in {\color{Green}green}) take up nutrients and \chem{CO_2}, which are released again from respiration and remineralisation of \chem{bSiO_2}, POC, DOM and lithogenic particles. The nutrients are not represented in the figure, because only particles impact \chem{^{231}Pa} and \chem{^{230}Th}. Zooplankton are denoted by the {\color{black}black} box. All sinking particles are denoted by {\color{Blue}blue} boxes. Effectively, this figure comprises the internal cycling of \textsc{Pisces}, minus details that are not of interest here, plus the lithogenic dust model. DOM stands for Dissolved Organic Carbon, sPOM and bPOM stand for small and big Particulate Organic Matter which are both subject to the differential lability scheme \citep{aumont2017}, and \chem{bSiO_2} stands for biogenic silica. % Sinking is denoted by the {\color{Red}red} arrows; tripple arrows means fast, normal arrows slow. }
\caption{The conceptual reversible scavenging model for the radionuclides. The radioisotopes, \chem{^{230}Th} and \chem{^{231}Pa}, are depicted in {\color{Orange}orange} when in the dissolved phase.}
\caption{Hologram printed on a glass substrate and an optical reconstruction of the hologram: (a) printed hologram and (b) optical reconstruction (\textcolor{blue}{see Visualization 1}).}
\caption{ Avalanche size distributions for three runs of the simulation, supercritical (\textcolor{red}{red}), subcritical (\textcolor{green}{green}) and critical (\textcolor{blue}{blue}). Black line: reference $\alpha=-2$ power law as reported by \cite[Fig.~3A]{Beggs2003} for \SI{1}{\milli \second}-binned LFP data. Inset: Kolmogorov-Smirnov statistic $D$ comparing the data against a theoretical power law with the estimated parameters. Filled circles in the inset correspond to the runs shown in the main plot. }
\caption{Active information storage (\textcolor{blue}{blue}, left axis) and nonparametric causal density (\textcolor{red}{red}, right axis) for different rewiring probabilities $p$. Each measure peaks at one side of the critical region around $p=0.3$ where the system shows power law-like statistics (see Fig.~\ref{fig:ksInset}).}
\caption{MI between pairs of modules using linear-Gaussian (\textcolor{red}{red}) and nonparametric (\textcolor{blue}{blue}) estimators.}
\caption{Streamline Slimming: The GoogLeNet example and the running time is measured using bvlc\_googlenet model in Caffe on a Samsung Galaxy S5. Left panel: convolution (in green), LRN (in red), pooling (in red). Right Panel: single convolution layer. The three layers in the left panel are merged and regenerated as a convolution layer (i.e., slim layer) in the right panel.}
\caption{Branch Slimming: The GoogLeNet example and the running time is measured using bvlc\_googlenet model in Caffe on a Samsung Galaxy S5. Left panel: four branches in parallel, convolution layer, convolution + convolution, convolution + convolution, convolution + pooling. Right panel: two branches in parallel, convolution + convolution, convolution + convolution. Two branches are reduced.}
\caption{\label{tabel:Market-state} Comparison of our method with state-of-the-art methods on the Market-1501 dataset. We use ResNet-50 as backbone. The best, second and third highest results are in \textcolor{red}{red}, \textcolor{blue}{blue} and \textcolor{green}{green}, respectively.}
\caption{Annihilation spectra for (a) 1,2-trans-dichloroethylene (C$_2$H$_2$Cl$_2$) and (b) tetrachloroethylene (C$_2$Cl$_4$): ($\bullet$) measured data; (\textcolor{magenta}{$\textbf{---}$}) total $\ZeffTot$ from Eq.~\eqref{Eq:ZeffTot}; (\textcolor{blue}{$\cdots$}) $\ZeffDir$; (\textcolor{blue}{$\textbf{-- --}$}) $\ZeffRes$, described by Eq.~\eqref{Eq:ZeffRes} and including only IA modes, using $\Gamma^e_\nu /\Gamma_\nu =1$; and ($-\cdot-$) $\ZeffMRA$, scaled by the factor $\eta$. Vertical bars show resonant energies $\varepsilon_\nu$; tall black bars and short red bars denote IA and infrared-inactive modes, respectively. The fit parameters from Eq.~(\ref{Eq:ZeffTot}) are (a) $\varepsilon_b = 15$~meV, $\eta = 0.76$ and (b) $\varepsilon_b = 57$~meV, $\eta = 0.13$. Insets show unidentified resonances in detail, where (\textcolor{magenta}{$\cdots$}) show $\ZeffTot$ with $\ZeffRes$ obtained from Eq.~\eqref{Eq:ZeffRes} \emph{including} the contribution of infrared-inactive C-C stretch modes with $\Gamma^e_\nu /\Gamma_\nu $ chosen to give the best fit of the data, yielding $\Gamma^e_\nu /\Gamma_\nu =0.39$ and 0.22 in (a) and (b), respectively.}
\caption{Schematics of different algorithms to return a loop-free prediction. Nodes such as {\protect\tikz[baseline=(X.base)]{\protect\node[specialState,inner sep=2pt] (X) {\small$1$}}} are selected by the algorithm, with thick edges such as {\protect\tikz{\protect\coordinate (X) at (0,0) {}; \protect\coordinate (Y) at (0.5,0) {}; \protect\draw[->,ultra thick] (X) to (Y); \protect\node[draw=none] at (0.25,-0.00625) {}}} denoting the sequence ordering. {\sc LoopElim} removes the loop from the Viterbi solution (here the POI sequence $(1,2,3,1)$), possibly returning a path of shorter length than requested; {\sc Greedy} incrementally selects POIs which have not been selected before, and locally maximises the sub-path score; {\sc ILP} solves an integer linear program to find the optimal length $l$ path in a complete graph over POIs; {\sc ListViterbi} finds where the second-best sequence diverges from the standard Viterbi sequence ($(1,2,3,1)$ as before); if not loop-free, it finds where the third-best diverges from the second-best, \emph{etc}. }
\caption{Remnants of the two most recent known supernovae in the Milky Way. Left: {\colorbox{yellow}{G1.9+0.3}}\index{G1.9+0.3} (about 1900 CE) (X-rays) (K.~Borkowski). Center: \cb{Cassiopeia A}\index{Cassiopeia A} (5 GHz, VLA; DeLaney et al.~2014). Right: Cassiopeia A (about 1680 CE). Green: Si band. Red: Fe K$\alpha$ band. (both with {\sl Chandra}; U.~Hwang). Blue: $^{44}$Ti emission (68 keV) with NuSTAR (Grefenstette et al.~2014).}
\caption{OO code metrics used for all studies in this paper. Last line, shown in \colorbox{lightgray}{gray}, denotes the dependent variable. For more details, see~\cite{krishna17a}.}
\caption{The number of changes recommended by each of the planners. The values on each row represents the percentage score indicating the number of times the metric has recommended for change. Note that XTREE (highlighted in \colorbox{lightgray}{gray}) recommends changes to far fewer metrics than the other methods.}
\caption{A simple example of computing overlap. Here a `$+$' represents an \textit{increase}, a `$-$' represents a \textit{decrease}, and a `$\cdot$' represents \textit{no-change}. Columns shaded in \colorbox{lightgray}{gray} indicate a match between developer's change and the recommendation made by a planner.}
\caption{The figure lists defect datasets used in this paper. The bellwether dataset is highlighted in \colorbox{lavenderpink}{light gray}.}
\caption{Area Under Planner Effectiveness Curve (AUPEC) obtained with the \ktest~for all planners. $\bigtriangledown$ indicates AUPEC for defects reduced and $\bigtriangleup$ indicates AUPEC for defects increased. \textit{Larger} values for $\bigtriangledown$ are preferable and \textit{smaller} values for $\bigtriangleup$ are preferable. For each row, cells with the largest AUPEC values shaded in \colorbox{lightgray}{gray}. Note that in 14 out of 18 cases, XTREE/BELLTREE reduces far more defects than it increases. Cells labeled with ${\dagger}$ indicates the best planner for \textit{reducing} defects. Note that in all cases, XTREE/BELLTREE outperform other planners. To compare XTREE with BELLTREE, cells are labeled with ${\ddagger}$. In 5 cases XTREE is better than BELLTREE, in 7 cases BELLTREE is better than XTREE, and in 6 cases they are comparable.}
\caption{(a) schematic setup and laser printing of the ultrathin lens on optical metasurfaces. (b) Amplitude contrast presented by the transmission difference between the printed and non-printed zones. {\color{black}The transmitted signals are normalized and taken under a white-light illumination.} (c) Experimental transmittance spectra of the printed zones which are printed with different laser intensities. {\color{black}For clarity, the spectra have been stacked with a +0.1 displacement.} (d) Corresponding absolute transmittance at the wavelength of 532\,nm under different printing power which are read from (c) (indicated by the dashed red line). (e) Simulated transmittance spectra for laser modulated metasurfaces.{\color{black}For clarity, the spectra have been stacked with a +0.1 displacement.} (f) Normalized electric field distribution of the original plasmonic nanostructure at a 532\,nm excitation.}
\caption{(a) Microscopic image of a fabricated FZP. Scale bar: 20\,$\rm\mu$m. (b) SEM image of a selected region (red box) in (a) which shows both the printed and unprinted areas. Scale bar: 2\,$\rm\mu$m. (c) Experimentally obtained image of the focused plane under a 532\,nm laser illumination. (d) Fitted experimental focal field intensity for laser illumination with wavelength of 532\,nm by integrating the intensity signals in a radial manner, which results in (a) . (e) Measured beam intensity profile of the FZP in the axial direction around the focal point. The intensity (I) along the center of focal beam is plotted along the$z$ axis.}
\caption{ Top panel: Temporal evolution of the energy spectra of particles from an NSM. \red{ For model NSM-RS (blue solid lines) and NSM-P (green dotted lines), we plot spectra at $t=0, 10^6, 5\times10^6, 10^7,$ and $2\times10^7$yr, from top to bottom. For models NSM-MB (red dashed lines), spectra are plotted for a shorter time span, i.e., $t=0, 10^5, 5\times10^5, 10^6,$ and $2\times10^6$yr. } Horizontal bars at the top indicate the detectable energy range of an instrument on the UHCRE satellite and stony iron meteorites. The detectors of UHCRE can detect UHCRs in the energy range of the dashed bar while the geomagnetic cutoff energy for the satellites is at 1.5 GeV/$A$. Bottom panel: The diffusion distance, $\lambda^{1/2}$, as a function of energy $E(t)$. The values at $t=10^5, 10^6, 10^7,$ and $3\times10^7$ yr are plotted from bottom to top. }
\caption{ The predicted energy spectra of the cosmic ray protons (red) and iron (blue) at the solar system. \red{ We overplot 15 lines at different time points at $2 \times 10^5$ yr intervals. } Crosses and filled circles show the observational results of protons in AMS2 \citep{Aguilar15} and CREAM \citep{Yoon11}, respectively. Squares, triangles, and open circles are the observed flux of iron by HEAO-3 \citep{Engelmann90}, TRACER \citep{Obermeier11} and CREAM \citep{Ahn09}, respectively. }
\caption{\red{The predicted UHCR flux in the detectable energy range by the UHCRE satellite as a function of time. The solid blue and dotted green lines denote results of model NSM-RS and NSM-P, respectively. The dash-dotted magenta line denotes SNR-UHCRs. The black line shows the predicted flux of iron in the SNR model times $(X_r/X_{\rm Fe})_\odot$. In this energy range, the result of NSM-MB is far smaller and not plotted. } }
\caption{\red{The predicted UHCR flux in the detectable energy range by meteorites. The red dashed line show the result of model NSM-MB. Line types and colors of other lines are the same as Figure~\ref{meteo}. } }
\caption{ UHCR flux averaged over the period from $t=0$ to $t$ for models \red{NSM-RS (blue solid), NSM-MB (red dashed),} NSM-P (green dotted), and SNR (magenta dash-dotted) in the meteorite energy range. Thin lines show the instantaneous flux (the same as Figure~\ref{meteo} but a different scale) for models \red{NSM-RS (light-blue), NSM-MB (pink), and} NSM-P (light-green). Four panels show results with different positions of $\theta = 0, \pi/2, \pi, 3\pi/2$ on a ring with $r = 8$ kpc. }
\caption{ Spallation fractions $ \langle 1-e^{-\nu} \rangle$ in the energy range of the UHCRE satellite in models NSM-RS (blue \red{solid}), NSM-P (green dotted), and SNR (magenta dash-dotted). The results of NSM-RS and NSM-P are almost overlapped. }
\caption{ Time averaged spallation fractions $\overline{ \langle 1-e^{-\nu} \rangle }$ in models \red{NSM-RS (blue solid), NSM-MB (red dashed)}, NSM-P (green dotted), and SNR (magenta dash-dotted) in the meteorite energy range. Thin lines show the instantaneous value $ \langle 1-e^{-\nu} \rangle$ in models NSM-RS (light-blue), NSM-MB (pink), and NSM-P (light-green). \red{The lines of NSM-RS and NSM-P are almost overlapped.} }
\caption{A schematic of a first-order \ac{IIR} implementation (specific to the DE2) with fixed point coefficients ($B_0$, $B_1$ and $A_1$) with 10-bit fractional resolution is shown. The \ac{FPGA} implementation is constructed from simple operations including multiplication ($\times$), addition ($\sum$), and division ($\div$) by the power of 2 implemented as a shift operation, saturation logic (\protect\includegraphics[height=0.35cm]{saturation_icon}) and clock delays (\protect\includegraphics[height=0.35cm]{clock_icon}.). The placement of the multiplication and division by $2^{10}$ operations is critical to correctly implement this \ac{IIR} filter with proper over-sampling. The bit width of each data path is noted in grey, and we note that the 14 bit width of the input and output signals are limited by the \ac{ADC} and \ac{DAC} conversion stages.}
\caption{Evolution of the normalised specific activity of \thtTtE\(dashed blue), \thtTtL\(dashed red), and the evolution of the ratio of the two (thick black) for a material where all radium has been removed. It can be seen that the ratio evolves rapidly over the first 5--10 years after processing (zero time corresponds to the point at which all radium is removed) reaching equilibrium after about 50 years. Additionally, it is only on this timescale that \gray\spectroscopy will give a true measurement of the specific activity of\thtTt\in a material such as this.\label{fig:bateman}}
\caption{A comparison between simulated $^{214}$Bi spectra for a sample in the well of Lumpsey. The red spectrum is produced by a simulation which uses simple \gray\branching ratios and the blue spectrum is produced using the\geant\radioactive decay libraries which includes the summing effects described in the main text. For ease of comparison, the blue spectrum is offset by\SI{1}{\keV}. The resolution observed in these spectra is as measured for Lumpsey in Figure~\ref{fig:DetectorResn}. \label{fig:Bi214CCF}}
\caption{Comparison between normalised spectra from resistor samples run on Lunehead (red) and Chaloner (blue). It is clear that there is a significant peak \pbtoz\seen at\SI{46}{\kilo\electronvolt} in Chaloner that is not reproduced in Lunehead. For comparison, the full photopeak efficiency for the sample on each detector is plotted using dashed lines in the same colour as the spectra. It is clear that Chaloner maintains high efficiency at the \pbtoz\\gray\energy whereas the efficiency has fallen off precipitously in Lunehead. The measured specific activity for\pbtoz, shown in Table~\ref{tab:resistors}, is significantly out of equilibrium with the measured value for \utTeL. \label{fig:resistors}}
\caption{Calculated contaminations for a sample of resistors screened using both Chaloner and Lunehead. A high level of agreement is seen across the \utTe\chain with the exception of\pbtoz\which has a measured specific activity 58$\times$ higher than those isotopes above it in the chain (determined using the combined Lunehead/Chaloner measurement of \utTeL). A measurement by Lunehead alone would not observe this.}{\footnotesize \begin{tabular}{C{1.1cm}|C{1.1cm}|C{1.1cm}|C{1.1cm}|C{1.1cm}|} % \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \cline{2-5} \rule{0pt}{2.4ex}& \multicolumn{4}{ c| }{{all values in \SI{}{\micro\becquerel\per resistor}}} \\ \hline \multicolumn{1}{ |c| }{{Detector} } & \rule{0pt}{2.4ex}{\utTeE\} & \rule{0pt}{2.4ex}{\utTeL\} & \rule{0pt}{2.4ex}{\pbtoz\} & \rule{0pt}{2.4ex}{\utTF\}\\ \hline \multicolumn{1}{ |c| }{ Chaloner} & 5.8$\pm$1.7 & 3.7$\pm$0.9 & 267$\pm$9 & $<$0.3 \\ \multicolumn{1}{ |c| }{ Lunehead} & $<$29 & 5.6$\pm$1.1 & - & $<$0.6 \\ \multicolumn{1}{ |c| }{ \textbf{Combined}} & \textbf{5.8$\pm$1.7} & \textbf{4.6$\pm$1.0} & \textbf{267$\pm$9} & \textbf{$<$0.3} \\ \hline\hline \multicolumn{1}{ |c| }{{Detector} } & \rule{0pt}{2.4ex}{{\thtTtE\}} & \rule{0pt}{2.4ex}{{\thtTtL\}} & \rule{0pt}{2.4ex}{{\kfz\}} & \rule{0pt}{2.4ex}{{\cosz\}} \\ \hline \multicolumn{1}{ |c| }{ Chaloner} & 1.9$\pm$1.0 & 1.4$\pm$0.3 & 29$\pm$6 & $<$0.3\\ \multicolumn{1}{ |c| }{ Lunehead} & 2.7$\pm$1.1 & 1.8$\pm$0.4 & 28$\pm$6 & $<$0.4 \\ \multicolumn{1}{ |c| }{ \textbf{Combined}} & \textbf{2.3$\pm$1.1} & \textbf{1.5$\pm$0.3} & \textbf{29$\pm$6} & \textbf{$<$0.3} \\ \hline \end{tabular}% }
\caption{Comparison of different methods on 6 datasets under different settings. \blu{Blue} indicates the best performance under each setting while \red{red} indicates the best performance under all settings. ``-C'', ``-R'', and ``-RC'' means using the CRF postprocessing, the ResNet50 backbone, and both of them, respectively.}
\caption{Benchmark of the eigenvalue solver MGK1d with particle code AWECS. {\color{\mycolor}Left and right frames show the imaginary ($\omega_i$) and real ($\omega_r$) parts of the eigenvalue divided by $v_{ti}/R$ respectively for different values of $\beta_e$. The blue diamonds and green squares indicate the ITG and KBM branches by solving Eqs. (\ref{eq:gkem_e_a})-(\ref{eq:vort_e_a}). The magenta pentagram indicates the electrostatic ITG branch with the constraint $\Psi=0$, which is tested in \cite{Xie2017}. }}
\caption{{\color{\mycolor}Left and right frames show the imaginary ($\omega_i$) and real ($\omega_r$) parts of the eigenvalue divided by $v_{ti}/R$ respectively for different values of $k_\theta\rho_i$. Different eigenstates of KBM co-exist, labeled with $l=0,1,2$.} The $l=1$ KBM dominates for $3\lesssim k_\theta\rho_i \lesssim7$, and the $l=2$ KBM dominates for $k_\theta\rho_i \gtrsim 7$.}
\caption{{\color{\mycolor}Mode structures of $\delta\phi$ (left), $\delta\psi$ (center) and $\delta A_\parallel$ (right) for the most unstable mode with $l=1$ and $\omega/(v_{ti}/R)=-17.2+7.04\imath$ (first row), the ground state mode (third most unstable) with $l=0$ and $\omega/(v_{ti}/R)=-14.5+5.31\imath$ (second row) and second most unstable mode with $l=2$ and $\omega/(v_{ti}/R)=-35.4+5.8\imath$ (third row) KBMs with $k_\theta\rho_i=5.0$. The blue solid line and the red dashed line indicate the real and imaginary parts of $\delta\phi$, $\delta\psi$ and $\delta A_\parallel$.}}
\caption{Values of the initial radii of the embedded star clusters, $R_{cl}$, \citep{Marks2012}, $\alpha_3$ values (see Eq.~(\ref{eq:MKDP})), and the number of massive stars $N(m_{\star}>8 M_{\odot})$ computed as a function of initial stellar mass, $M_{UCD}=M_{ecl}$, and metallicity. For comparison, we also show $N(m_{\star}>8M_{\odot})$ for the \textcolor{green}{CAN IMF}. The predicted values of the embedded cluster/UCD initial radii, $R_{cl}$, are small. However, due to stellar and dynamical evolution, UCDs expand by approximately a factor of ten \citep{Dabringhausen2010}. We use the $R_{cl}$ values as an empirical extrapolation from GCs and therefore at UCD scales departures are possible.}
\caption{Schematic representation of the optimization-based DDM analysis. The procedure is based on an optimization cycle (yellow arrows), fed by the experimental image structure function $D(q,t)$ and by an initial set of parameters $\left( A_0(q), B_0(q) \right)$. The object function is the dispersion $\sigma^2$ of the reconstructed mean square displacements (see Eq. \ref{appsigma2}). New values of $\left(A(q), B(q) \right)$ are iteratively generated in order to minimize the object function. The output of the procedure is the optimal set of parameters $\left( A(q), B(q) \right)$ leading to the best estimate of MSD($t$).} \label{fig:diagrams} \end{figure*} The general idea is sketched in Fig.\ref{fig:diagrams}, where we show a block-diagram that depicts our fitting-free %optimization-based procedure. This procedure is based on an optimization cycle initially fed by a tentative amplitude-baseline parameter pair $\left(A_0(q),B_0(q)\right)$, for $q$-values within a given interval $[q_1,q_2]$. These parameters are used to invert the corresponding image structure functions (Eq. \ref{eq:msdddm2}), leading to a "bundle" of MSDs. If the considered pair $\left(A_0(q),B_0(q)\right)$ is the correct one for all $q$s, than the estimates for the MSD given by Eq. \ref{eq:msdddm2} are completely $q$-independent, resulting in an almost perfect collapse of all the curves. Any deviation of the parameters from the correct values introduces a $q$-dependent dispersion. In our optimization scheme, the dispersion $\sigma^2$ of the curves (see Appendix A for details) plays the role of an objective function: new values of $\left(A(q),B(q)\right)$ are iteratively generated until a minimum of $\sigma^2$ is found. This algorithm, implemented in a custom code developed in MATLAB\textregistered, was found to rapidly and robustly converge to a minimum for a wide range wave-vectors. Results obtained for the tracer MSD with this optimization-based procedure (Fig. \ref{fig:newton}b) are in excellent agreement with those obtained with the fitting-based analysis (Fig. \ref{fig:newton}a) over the whole investigated range of delay times $1.3 \times 10^{-3}$ $s< t<10^2$ $s$, which validates the procedure. Also, we note that the $q$-averaged MSD shown in Fig. \ref{fig:newton}b were obtained by averaging the MSD in the range $1.36$ $\mu m^{-1}<q<9.06$ $\mu m^{-1}$; this range is wider than that probed with the fitting-based procedure. The usable $q$-range is larger in the optimization-based procedure, because the full relaxation of the image structure functions is here not a requirement for the determination of the MSD, since $A(q)$ and $B(q)$ can be obtained self-consistently. Let us stress that the optimization-based procedure is largely model- and operator-independent. The only required external parameters are the relevant $q$-range $[q_1,q_2]$ over which the optimization is performed and the initial values of the parameters $(A_0(q),B_0(q))$. The key importance of all these properties when studying arbitrary samples is described in detail in the next subsection. \subsection{Viscoelastic fluid} \begin{figure}[!t] \centering \includegraphics[width=.5\textwidth]{MSD_PEO_new3} \caption{(a) Two-dimensional MSD of LB1 (orange triangles) and LB5 (blue circles) tracers in a viscoelastic polymer solution (2\% PEO2 in water) obtained from DDM analysis. Black dots: same quantity obtained from PT analysis for the sample with LB5 tracers. The small insets show representative images of the two samples: the one loaded with subdiffraction LB1 particles (left upper corner) and the one loaded with LB5 tracers (right lower corner), respectively. (b) Comparison of the storage moduli $G^{\prime}$ estimated from the DDM-reconstructed MSD of LB1 tracers (orange triangles) and LB5 tracers (blue circles), respectively. (c) same as in panel (b) for the loss moduli $G^{\prime \prime}$.} \label{fig:PEO2_MSD} \end{figure} \begin{figure}[!t] \centering \includegraphics[width=.5\textwidth]{PEORHEO_new} \caption{Comparison of the viscoelastic moduli $G^{\prime}$ and $G^{\prime \prime}$ of a 2\% PEO polymer solution in water, obtained with different methods. Gray circles (triangles): $G^{\prime}$ ($G^{\prime \prime}$) obtained with traditional rheology; continuous blue (orange) line: $G^{\prime}$ ($G^{\prime \prime}$) obtained with DWS using LB5 tracers; black continuous (dashed) line: $G^{\prime}$ ($G^{\prime \prime}$) obtained with DDM microrheology (weighted average of the results of LB1 and LB5 tracers, shown individually in Fig. \ref{fig:PEO2_MSD}).} \label{fig:PE2_rheo} \end{figure} In this subsection, we apply the optimization-based procedure to the data obtained with our model viscoelastic fluid, an aqueous solution of PEO, that exhibits elastic behavior at short-times, high frequencies. The expected short time elastic plateau in the MSD would contribute to the baseline $B(q)$, requiring an independent determination of the camera noise. Such requirement would be similarly involved as the calibration procedure needed in PT-$\mu$r experiments to account for the tracer localization uncertainty. While such calibration is technically feasible, the optimization-based DDM analysis permits a calibration-free implementation of DDM-$\mu$r. Application of the optimization-based procedure to the PEO solutions with small ($\sim 100$ nm) and large ($\sim 500$ nm) tracers provides the results shown in Fig. \ref{fig:PEO2_MSD}a. The accessible range of probed timescales is very similar for the two tracer sizes. For comparison we also show the results obtained with PT for the sample containing the larger tracers as black points in Fig. \ref{fig:PEO2_MSD}a; for the smaller particles tracking is not feasible, as easily appreciated from the images shown as insets. \begin{figure*}[!t] \centering \includegraphics[width=.99\textwidth ]{MSD_PEO_new33} \caption{Effect of the model-dependent determination of the noise baseline $ B(q)$ on the reconstructed MSD. (a) Symbols: image structure function $D(q,t)$ (for $q=6.04$ $\mu m^{-1}$) obtained from DDM on LB5 tracers in a viscoelastic solution of PEO in water. The continuous red line is an exponential fit to the data obtained at large delay times ($t>20$ s); this fit allows for the estimation of the plateau height $A^{\prime}(q)$. Inset: close-up of the short time behavior of $D(q,t)$ (symbols). The data is fitted with different functions, leading to different estimates of the baseline $B(q)$: linear fit over the first $10$ data points (continuous blue line), linear fit over the first $5$ data points (dashed orange line), fit over the first 20 data points with a function of the form $y=ax^{0.5} +b$ (dashed-dotted yellow line) and fit over the same interval with a function of the form $y=ax^{0.25} +b$ (dotted green line). (b) Mean square displacement obtained from Eq. \ref{eq:msdddm} using the amplitude $A(q)=A^{\prime}(q)-B(q)$ and the noise baseline $B(q)$ obtained from the different fitting models shown in the inset of panel (a). Curves are color-coded according to the fits, the black continuous line is the result of the model-free procedure shown in Fig. \ref{fig:PEO2_MSD}(a) as blue circles.} \label{fig:PEO2rightandwrong} \end{figure*} For each tracer size, we extracted from the MSD the mechanical moduli $G^{\prime}$ and $G^{\prime \prime}$, as shown in Fig. \ref{fig:PEO2_MSD}(b) and (c). Results obtained for $G^{\prime}$ with the two tracer sizes are off by about 10-20\% at small frequencies but the two datasets are compatible within the experimental errors. To obtain a statistically significant estimate for the moduli, we combine the data obtained with the two tracers and show the results as black lines in Fig. \ref{fig:PE2_rheo}. These data are in good agreement with the results obtained with traditional rheology, shown as open symbols, and also with the results obtained with DWS, shown as closed symbols. DDM-$\mu$r extends traditional rheology by one decade at high frequency, whereas at low frequency similar performances are obtained, at least as far as the storage modulus is concerned. However, improvements in the low-frequency region may be expected by increasing the mechanical stability of the microscope setup. Let us underline that, without additional calibration steps, it would be very difficult to extract meaningful MSD and thus mechanical moduli with a fitting-based analysis of the DDM-data. In the limit of short times the MSD displays a non-trivial scaling, compatible with a power-law $MSD \simeq t^\gamma$ with an exponent $\gamma$ close to $0.5$. The counterpart of this behavior in the Fourier space is an image structure function taking the form of $D(q,t) \simeq C(q) t^{\gamma}+B(q)$ at short times. Clearly, an exponential or a polynomial fit of $D(q, t)$ are inadequate to describe this behavior and any estimate of the baseline $B(q)$ based on an exponential or a polynomial fit provides a biased, incorrect result, as shown in Fig. \ref{fig:PEO2rightandwrong}. Choosing other model functions, such as for instance a power-law with different exponents, also fails, even though the data may seem deceivingly well described at short times. By contrast, the optimization-based procedure self-consistently determines the MSD without need of fitting the experimentally determined image structure functions. \section{Conclusions} Microrheology is a very powerful complement to traditional, mechanical rheology \cite{Waigh:2005jk,Cicuta:2007vn,Squires:2010hb,Waigh:2016sf}. For the high-frequency range, rheology is usefully complemented by DWS-$\mu$r \cite{Mason:1995fq}, whereas in the low-frequency limit both DLS-$\mu$r \cite{mason1996rheology} and PT-$\mu$r \cite{Mason:1997yu} have been usefully employed in the past. PT-$\mu$r is technically the less demanding technique, not requiring any laser source or digital correlation board and is also very flexible for biophysical applications, owing to the possibility of employing different sample contrast mechanisms. However, in its practical realization one encounters some challenges. Accurate tracking algorithms require several input parameters, such as a typical value for the particle radius, a score cut-off to discriminate signals that are not due to presence of a particle, an intensity threshold to consider bright pixels as particles, etc. The results of the tracking depends severely on the choice of these parameters that, even for experienced users, may be sometimes more difficult than expected \cite{chenouard2014objective}. Also, the extraction of the tracer MSD from PT trajectories requires the knowledge of the intrinsic particle localization uncertainty, which is usually determined by calibration with particles that are kept fixed in space or that freely diffuse in a Newtonian fluid with similar optical properties \cite{Waigh:2005jk}. We have shown here that DDM \citep{Cerbino:2008if}, a technique that retains the simplicity and flexibility of PT in terms of experimental setup and applications, can be also used for accurate microrheology experiments. We also show that DDM-$\mu$r outperforms PT with small particles in bright field microscopy. Finally, if an optimization-based algorithm is used instead of the standard fitting-based approach, DDM-$\mu$r does not require any calibration or user input, which limits dramatically the degree of arbitrariness on the determination of the mechanical moduli of the sample. However, particle tracking is expected to be superior to DDM in the presence of unwanted and moving scatterers that, being potentially discarded by an accurate particle tracking, would affect DDM-$\mu$r experiments. It is likely that these and other DDM features, such as its capability to handle optically dense samples, for which tracking becomes extremely challenging if not impossible, will make DDM-$\mu$r a useful addition to the portfolio of rheo-scientists, both in academic and in industrial research laboratories. \appendix \section{Optimization-based determination of MSD} In this appendix we describe the fitting-free optimization procedure used to extract from the experimental image structure function $D(q,t)$ the best estimate for the tracers' mean square displacement. The main steps of the procedure are the following: \begin{enumerate} \item Choice of the interval $[q_1,q_2]$ of wave-vectors over which the optimization is performed. The interval should be a subset of the accessible $q$-range with a fair signal to noise ratio. This condition can be also checked retrospectively at the end of the procedure, when a $q$-resolved estimate of the amplitude $A(q)$ and the noise background $B(q)$ is obtained. \item Choice of the initial set of parameters $\left(A_0(q),B_0(q)\right)$. This can be done, for example, by fitting, for each $q \in[q_1,q_2]$, $D(q,t)$ with a linear function near the origin and with a exponential function for large delays (as done for example in Fig.\ref{fig:PEO2rightandwrong}). \item Calculation of the mean square displacement $MSD(t|q)$ using Eq. \ref{eq:msdddm} for each $q$ in the selected interval. \item Determination, for each delay time $t$, of the subset $J(t)$ of $q$-values such that $MSD(t|q)<q^{-2}$. This choice ensures that, if $q \in J(t)$, then $D(q,t)$ has not completely lost track of the signal correlation for that value of $q$ and can be thus meaningfully inverted. Let $N(t)$ be the number of elements in $J(t)$. \item Calculation of the average mean square displacement \begin{equation} MSD(t)=\frac{1}{N(t)}\sum_{q\in J(t)}MSD(t|q). \label{appMSD} \end{equation} \item Calculation of the $t$-dependent dispersion $\sigma_t^2(t)$ as \begin{equation} \sigma_t^2(t)=\frac{1}{N(t)-1}\sum_{q\in J(t)} \log^2 \frac{MSD(t|q)}{MSD(t)}. \label{appsigma1} \end{equation} and of the total dispersion $\sigma^2$ as \begin{equation} \sigma^2=\sum_{t}\sigma^2(t). \label{appsigma2} \end{equation} \item Generation of a new set of parameters and repetition of the procedure from step 3 unless a local minimum in $\sigma^2$ is reached (or the prescribed maximum number of iterations is exceeded). \item If the procedure converges to a minimum of $\sigma^2$, the optimal set of parameters $\left(A(q),B(q) \right)$ represents the best estimate for the $q$-dependent amplitude and noise baseline, respectively, and the corresponding average mean square displacement (Eq. \ref{appMSD}) is the best estimate for the tracer's MSD. \end{enumerate} Many algorithms are available to search the minimum of $\sigma^2$ and to guide the generation of new sets of parameters in step 7. In our implementation, the optimization cycle 3-7 was realized using the MATLAB function \textit{fminsearch}, which is based on the simplex search method of Lagarias \textit{et al.} \cite{fminsearch}. More refined implementation could possibly include suitable weights when computing the averages in the right-hand sides of Eqs. \ref{appMSD}-\ref{appsigma2}, accounting for the different statistical errors affecting each term. Also, an effective weighting scheme could provide an efficient way to reject the contribution of the most noisy wave-vectors making unnecessary the explicit selection of a predetermined optimization interval (step 1). \begin{acknowledgments} We thank Stefano Buzzaccaro, Marco Caggioni and Giuliano Zanchetta for stimulating discussions. We thank Gora Conley for careful reading of the manuscript. We acknowledge funding from the Italian Ministry of University and Scientific Research (MIUR) - Project RBFR125H0M, from Regione Lombardia and CARIPLO foundation - Project 2016-0998 and from the Swiss National Science Foundation - Project 200021-157214. \end{acknowledgments} \bibliographystyle{apsrev4-1} %merlin.mbs apsrev4-1.bst 2010-07-25 4.21a (PWD, AO, DPC) hacked %Control: key (0) %Control: author (72) initials jnrlst %Control: editor formatted (1) identically to author %Control: production of article title (-1) disabled %Control: page (0) single %Control: year (1) truncated %Control: production of eprint (0) enabled \begin{thebibliography}{36}% \makeatletter \providecommand \@ifxundefined[1]{% \@ifx{#1\undefined} }% \providecommand \@ifnum[1]{% \ifnum #1\expandafter \@firstoftwo\else \expandafter \@secondoftwo\fi }% \providecommand \@ifx[1]{% \ifx #1\expandafter \@firstoftwo\else \expandafter \@secondoftwo\fi }% \providecommand \natexlab[1]{#1}% \providecommand \enquote[1]{``#1''}% \providecommand \bibnamefont[1]{#1}% \providecommand \bibfnamefont[1]{#1}% \providecommand \citenamefont[1]{#1}% \providecommand \href@noop [0]{\@secondoftwo}% \providecommand \href[0]{\begingroup \@sanitize@url\@href}% \providecommand \@href[1]{\@@startlink{#1}\@@href}% \providecommand \@@href[1]{\endgroup#1\@@endlink}% \providecommand \@sanitize@url[0]{\catcode `\\12\catcode `\$12\catcode `\&12\catcode `\#12\catcode `\^12\catcode `\_12\catcode `\%12\relax}% \providecommand \@@startlink[1]{}% \providecommand \@@endlink[0]{}% \providecommand \url[0]{\begingroup\@sanitize@url\@url}% \providecommand \@url[1]{\endgroup\@href{#1}{\urlprefix }}% \providecommand \urlprefix[0]{URL }% \providecommand \Eprint[0]{\href }% \providecommand \doibase[0]{http://dx.doi.org/}% \providecommand \selectlanguage[0]{\@gobble}% \providecommand \bibinfo[0]{\@secondoftwo}% \providecommand \bibfield[0]{\@secondoftwo}% \providecommand \translation[1]{[#1]}% \providecommand \BibitemOpen[0]{}% \providecommand \bibitemStop[0]{}% \providecommand \bibitemNoStop[0]{.\EOS\space}% \providecommand \EOS[0]{\spacefactor3000\relax}% \providecommand \BibitemShut[1]{\csname bibitem#1\endcsname}% \let\auto@bib@innerbib\@empty%</preamble> \bibitem[{\citenamefont{Coussot}(2014)}]{coussot2014rheophysics}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Coussot}},\}\href@noop {} {\emph{\bibinfo{title}{Rheophysics: Matter in all its States}}}\(\bibinfo{publisher}{Springer},\\bibinfo{year}{2014})\BibitemShut{NoStop}% \bibitem[{\citenamefont{Wyss}(2016)}]{Wyss:2016eu}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{H.~M.}\\bibnamefont{Wyss}},\}\enquote{\bibinfo{title}{Rheology of soft materials},}\in\\href{\doibase 10.1002/9781119220510.ch9}{\emph{\bibinfo{booktitle}{Fluids, Colloids and Soft Materials}}}\(\bibinfo{publisher}{John Wiley and Sons},\\bibinfo{year}{2016})\pp.\\bibinfo{pages}{149--163}\BibitemShut{NoStop}% \bibitem[{Note1()}]{Note1}% \BibitemOpen \bibinfo{note}{As long as the stress or strain is small enough not to change materials' mechanical properties.}\BibitemShut{Stop}% \bibitem[{\citenamefont{Kavanagh}\and\\citenamefont{Ross-Murphy}(1998)}]{Kavanagh:1998fp}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{G.~M.}\\bibnamefont{Kavanagh}}\and\\bibinfo{author}{\bibfnamefont{S.~B.}\\bibnamefont{Ross-Murphy}},\}\href{\doibase http://dx.doi.org/10.1016/S0079-6700(97)00047-6}{\bibfield{journal}{\bibinfo{journal}{Prog. Polym. Sci.}\}\textbf{\bibinfo{volume}{23}},\\bibinfo{pages}{533 } (\bibinfo{year}{1998})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Ewoldt}\\emph{et~al.}(2015)\citenamefont{Ewoldt}, \citenamefont{Johnston},\and\\citenamefont{Caretta}}]{Ewoldt:2015qr}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{R.~H.}\\bibnamefont{Ewoldt}}, \bibinfo{author}{\bibfnamefont{M.~T.}\\bibnamefont{Johnston}}, \and\\bibinfo{author}{\bibfnamefont{L.~M.}\\bibnamefont{Caretta}},\}\enquote{\bibinfo{title}{Experimental challenges of shear rheology: how to avoid bad data},}\in\\href@noop {} {\emph{\bibinfo{booktitle}{Complex Fluids in Biological Systems}}}\(\bibinfo{publisher}{Springer},\\bibinfo{year}{2015})\pp.\\bibinfo{pages}{207--241}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Cicuta}\and\\citenamefont{Donald}(2007)}]{Cicuta:2007vn}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Cicuta}}\and\\bibinfo{author}{\bibfnamefont{A.~M.}\\bibnamefont{Donald}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Soft Matter}\}\textbf{\bibinfo{volume}{3}},\\bibinfo{pages}{1449} (\bibinfo{year}{2007})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Mason}\\emph{et~al.}(1997)\citenamefont{Mason}, \citenamefont{Ganesan}, \citenamefont{Van~Zanten}, \citenamefont{Wirtz},\and\\citenamefont{Kuo}}]{Mason:1997yu}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{T.~G.}\\bibnamefont{Mason}}, \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Ganesan}}, \bibinfo{author}{\bibfnamefont{J.~H.}\\bibnamefont{Van~Zanten}}, \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Wirtz}}, \and\\bibinfo{author}{\bibfnamefont{S.~C.}\\bibnamefont{Kuo}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{79}},\\bibinfo{pages}{3282} (\bibinfo{year}{1997})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Squires}\and\\citenamefont{Mason}(2010)}]{Squires:2010hb}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{T.~M.}\\bibnamefont{Squires}}\and\\bibinfo{author}{\bibfnamefont{T.~G.}\\bibnamefont{Mason}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Annu. Rev. Fluid Mech.}\}\textbf{\bibinfo{volume}{42}} (\bibinfo{year}{2010})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Waigh}(2005)}]{Waigh:2005jk}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{T.~A.}\\bibnamefont{Waigh}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Rep. Prog. Phys}\}\textbf{\bibinfo{volume}{68}},\\bibinfo{pages}{685} (\bibinfo{year}{2005})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Waigh}(2016)}]{Waigh:2016sf}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{T.~A.}\\bibnamefont{Waigh}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Rep. Prog. Phys.}\}\textbf{\bibinfo{volume}{79}},\\bibinfo{pages}{074601} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Mason}\and\\citenamefont{Weitz}(1995)}]{Mason:1995fq}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mason}}\and\\bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Weitz}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{74}},\\bibinfo{pages}{1250} (\bibinfo{year}{1995})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Mason}\\emph{et~al.}(1996)\citenamefont{Mason}, \citenamefont{Gang},\and\\citenamefont{Weitz}}]{mason1996rheology}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mason}}, \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Gang}}, \and\\bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Weitz}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{J. Mol. Struct.}\}\textbf{\bibinfo{volume}{383}},\\bibinfo{pages}{81} (\bibinfo{year}{1996})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Dasgupta}\\emph{et~al.}(2002)\citenamefont{Dasgupta}, \citenamefont{Tee}, \citenamefont{Crocker}, \citenamefont{Frisken},\and\\citenamefont{Weitz}}]{Dasgupta:2002jt}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{B.~R.}\\bibnamefont{Dasgupta}}, \bibinfo{author}{\bibfnamefont{S.~Y.}\\bibnamefont{Tee}}, \bibinfo{author}{\bibfnamefont{J.~C.}\\bibnamefont{Crocker}}, \bibinfo{author}{\bibfnamefont{B.~J.}\\bibnamefont{Frisken}}, \and\\bibinfo{author}{\bibfnamefont{D.~A.}\\bibnamefont{Weitz}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Physical Review E}\}\textbf{\bibinfo{volume}{65}},\\bibinfo{pages}{051505} (\bibinfo{year}{2002})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Rathgeber}\\emph{et~al.}(2009)\citenamefont{Rathgeber}, \citenamefont{Beauvisage}, \citenamefont{Chevreau}, \citenamefont{Willenbacher},\and\\citenamefont{Oelschlaeger}}]{Rathgeber:2009mb}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Rathgeber}}, \bibinfo{author}{\bibfnamefont{H.-J.}\\bibnamefont{Beauvisage}}, \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Chevreau}}, \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Willenbacher}}, \and\\bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Oelschlaeger}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Langmuir}\}\textbf{\bibinfo{volume}{25}},\\bibinfo{pages}{6368} (\bibinfo{year}{2009})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Cerbino}\and\\citenamefont{Trappe}(2008)}]{Cerbino:2008if}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Cerbino}}\and\\bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Trappe}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{100}},\\bibinfo{pages}{188102} (\bibinfo{year}{2008})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Croccolo}\\emph{et~al.}(2006)\citenamefont{Croccolo}, \citenamefont{Brogioli}, \citenamefont{Vailati}, \citenamefont{Giglio}, \citenamefont{Cannell},\and\\citenamefont{Sadhal}}]{Croccolo:2006et}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Croccolo}}, \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Brogioli}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Vailati}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Giglio}}, \bibinfo{author}{\bibfnamefont{D.~S.}\\bibnamefont{Cannell}}, \and\\bibinfo{author}{\bibfnamefont{S.~S.}\\bibnamefont{Sadhal}},\}\href{\doibase 10.1196/annals.1362.030}{\bibfield{journal}{\bibinfo{journal}{Ann.NY Acad.Sci.}\}\textbf{\bibinfo{volume}{1077}},\\bibinfo{pages}{365} (\bibinfo{year}{2006})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Giavazzi}\and\\citenamefont{Cerbino}(2014)}]{Giavazzi:2014sj}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Giavazzi}}\and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Cerbino}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{J. Opt.}\}\textbf{\bibinfo{volume}{16}},\\bibinfo{pages}{083001} (\bibinfo{year}{2014})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Giavazzi}\\emph{et~al.}(2009)\citenamefont{Giavazzi}, \citenamefont{Brogioli}, \citenamefont{Trappe}, \citenamefont{Bellini},\and\\citenamefont{Cerbino}}]{Giavazzi:2009xd}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Giavazzi}}, \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Brogioli}}, \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Trappe}}, \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Bellini}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Cerbino}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Phys. Rev. E}\}\textbf{\bibinfo{volume}{80}},\\bibinfo{pages}{031403} (\bibinfo{year}{2009})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Wilson}\\emph{et~al.}(2011)\citenamefont{Wilson}, \citenamefont{Martinez}, \citenamefont{Schwarz-Linek}, \citenamefont{Tailleur}, \citenamefont{Bryant}, \citenamefont{Pusey},\and\\citenamefont{Poon}}]{Wilson:2011wa}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{L.~G.}\\bibnamefont{Wilson}}, \bibinfo{author}{\bibfnamefont{V.~A.}\\bibnamefont{Martinez}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Schwarz-Linek}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Tailleur}}, \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Bryant}}, \bibinfo{author}{\bibfnamefont{P.~N.}\\bibnamefont{Pusey}}, \and\\bibinfo{author}{\bibfnamefont{W.~C.~K.}\\bibnamefont{Poon}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Phys. Rev. Lett.}\}\textbf{\bibinfo{volume}{106}},\\bibinfo{pages}{018101} (\bibinfo{year}{2011})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Reufer}\\emph{et~al.}(2012)\citenamefont{Reufer}, \citenamefont{Martinez}, \citenamefont{Schurtenberger},\and\\citenamefont{Poon}}]{Reufer:2012ri}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Reufer}}, \bibinfo{author}{\bibfnamefont{V.~A.}\\bibnamefont{Martinez}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Schurtenberger}}, \and\\bibinfo{author}{\bibfnamefont{W.~C.~K.}\\bibnamefont{Poon}},\}\href{\doibase 10.1021/la204904a}{\bibfield{journal}{\bibinfo{journal}{Langmuir}\}\textbf{\bibinfo{volume}{28}},\\bibinfo{pages}{4618} (\bibinfo{year}{2012})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{He}\\emph{et~al.}(2012)\citenamefont{He}, \citenamefont{Spannuth}, \citenamefont{Conrad},\and\\citenamefont{Krishnamoorti}}]{He:2012bx}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{He}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Spannuth}}, \bibinfo{author}{\bibfnamefont{J.~C.}\\bibnamefont{Conrad}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Krishnamoorti}},\}\href{\doibase 10.1039/c2sm26392k}{\bibfield{journal}{\bibinfo{journal}{Soft Matter}\}\textbf{\bibinfo{volume}{8}},\\bibinfo{pages}{11933} (\bibinfo{year}{2012})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Gao}\\emph{et~al.}(2015)\citenamefont{Gao}, \citenamefont{Kim},\and\\citenamefont{Helgeson}}]{Gao:2015zh}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Gao}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kim}}, \and\\bibinfo{author}{\bibfnamefont{M.~E.}\\bibnamefont{Helgeson}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Soft Matter}\}\textbf{\bibinfo{volume}{11}},\\bibinfo{pages}{6360} (\bibinfo{year}{2015})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Germain}\\emph{et~al.}(2016)\citenamefont{Germain}, \citenamefont{Leocmach},\and\\citenamefont{Gibaud}}]{Germain:2015gf}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Germain}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Leocmach}}, \and\\bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Gibaud}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Am. J. Phys}\}\textbf{\bibinfo{volume}{84}},\\bibinfo{pages}{202} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Dehaoui}\\emph{et~al.}(2015)\citenamefont{Dehaoui}, \citenamefont{Issenmann},\and\\citenamefont{Caupin}}]{Dehaoui:2015mz}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Dehaoui}}, \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Issenmann}}, \and\\bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Caupin}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Proc. Natl. Acad. Sci. U.S.A.}\}\textbf{\bibinfo{volume}{112}},\\bibinfo{pages}{12020} (\bibinfo{year}{2015})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Wittmeier}\\emph{et~al.}(2015)\citenamefont{Wittmeier}, \citenamefont{Leeth~Holterhoff}, \citenamefont{Johnson},\and\\citenamefont{Gibbs}}]{Wittmeier:2015qy}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Wittmeier}}, \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Leeth~Holterhoff}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Johnson}}, \and\\bibinfo{author}{\bibfnamefont{J.~G.}\\bibnamefont{Gibbs}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Langmuir}\}\textbf{\bibinfo{volume}{31}},\\bibinfo{pages}{10402} (\bibinfo{year}{2015})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Sentjabrskaja}\\emph{et~al.}(2016)\citenamefont{Sentjabrskaja}, \citenamefont{Zaccarelli}, \citenamefont{De~Michele}, \citenamefont{Sciortino}, \citenamefont{Tartaglia}, \citenamefont{Voigtmann}, \citenamefont{Egelhaaf},\and\\citenamefont{Laurati}}]{Sentjabrskaja:2016fq}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Sentjabrskaja}}, \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Zaccarelli}}, \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{De~Michele}}, \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sciortino}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tartaglia}}, \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Voigtmann}}, \bibinfo{author}{\bibfnamefont{S.~U.}\\bibnamefont{Egelhaaf}}, \and\\bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Laurati}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Nat. Commun.}\}\textbf{\bibinfo{volume}{7}} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Wulstein}\\emph{et~al.}(2016)\citenamefont{Wulstein}, \citenamefont{Regan}, \citenamefont{Robertson-Anderson},\and\\citenamefont{McGorty}}]{Wulstein:2016tt}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{D.~M.}\\bibnamefont{Wulstein}}, \bibinfo{author}{\bibfnamefont{K.~E.}\\bibnamefont{Regan}}, \bibinfo{author}{\bibfnamefont{R.~M.}\\bibnamefont{Robertson-Anderson}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{McGorty}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Opt. Express}\}\textbf{\bibinfo{volume}{24}},\\bibinfo{pages}{20881} (\bibinfo{year}{2016})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Kodger}\\emph{et~al.}(2017)\citenamefont{Kodger}, \citenamefont{Lu}, \citenamefont{Wiseman},\and\\citenamefont{Weitz}}]{Kodger:2017mw}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{T.~E.}\\bibnamefont{Kodger}}, \bibinfo{author}{\bibfnamefont{P.~J.}\\bibnamefont{Lu}}, \bibinfo{author}{\bibfnamefont{G.~R.}\\bibnamefont{Wiseman}}, \and\\bibinfo{author}{\bibfnamefont{D.~A.}\\bibnamefont{Weitz}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Langmuir}\} (\bibinfo{year}{2017})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Drechsler}\\emph{et~al.}(2017)\citenamefont{Drechsler}, \citenamefont{Giavazzi}, \citenamefont{Cerbino},\and\\citenamefont{Palacios}}]{drechsler2017active}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Drechsler}}, \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Giavazzi}}, \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Cerbino}}, \and\\bibinfo{author}{\bibfnamefont{I.~M.}\\bibnamefont{Palacios}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{bioRxiv}\,\\bibinfo{pages}{098590}} (\bibinfo{year}{2017})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Crocker}\and\\citenamefont{Grier}(1996)}]{crocker1996methods}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{J.~C.}\\bibnamefont{Crocker}}\and\\bibinfo{author}{\bibfnamefont{D.~G.}\\bibnamefont{Grier}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{J. Colloid Interface Sci.}\}\textbf{\bibinfo{volume}{179}},\\bibinfo{pages}{298} (\bibinfo{year}{1996})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Weitz}\and\\citenamefont{Pine}(1992)}]{pine_book}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Weitz}}\and\\bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Pine}},\}\href@noop {} {\emph{\bibinfo{title}{Dynamic Light Scattering}}}\(\bibinfo{publisher}{Oxford University Press, Oxford},\\bibinfo{year}{1992})\BibitemShut{NoStop}% \bibitem[{\citenamefont{{Giavazzi}}\\emph{et~al.}(2017)\citenamefont{{Giavazzi}}, \citenamefont{{Edera}}, \citenamefont{{Lu}},\and\\citenamefont{{Cerbino}}}]{2017arXiv_apodization}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{{Giavazzi}}}, \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{{Edera}}}, \bibinfo{author}{\bibfnamefont{P.~J.}\\bibnamefont{{Lu}}}, \and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{{Cerbino}}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{ArXiv e-prints}\} (\bibinfo{year}{2017})},\\Eprint{http://arxiv.org/abs/1707.07501}{arXiv:1707.07501 [cond-mat.soft]} \BibitemShut{NoStop}% \bibitem[{\citenamefont{Berne}\and\\citenamefont{Pecora}(2000)}]{Berne:2000ye}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{B.~J.}\\bibnamefont{Berne}}\and\\bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Pecora}},\}\href@noop {} {\emph{\bibinfo{title}{Dynamic light scattering: with applications to chemistry, biology, and physics}}}\(\bibinfo{publisher}{Dover Publications},\\bibinfo{year}{2000})\BibitemShut{NoStop}% \bibitem[{\citenamefont{Cheng}(2008)}]{Glycerolwater}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{N.-S.}\\bibnamefont{Cheng}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Ind. Eng. Chem. Res.}\}\textbf{\bibinfo{volume}{47}},\\bibinfo{pages}{3285} (\bibinfo{year}{2008})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Chenouard}\\emph{et~al.}(2014)\citenamefont{Chenouard}, \citenamefont{Smal}, \citenamefont{De~Chaumont}, \citenamefont{Ma{\v{s}}ka}, \citenamefont{Sbalzarini}, \citenamefont{Gong}, \citenamefont{Cardinale}, \citenamefont{Carthel}, \citenamefont{Coraluppi}, \citenamefont{Winter} \emph{et~al.}}]{chenouard2014objective}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Chenouard}}, \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Smal}}, \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{De~Chaumont}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Ma{\v{s}}ka}}, \bibinfo{author}{\bibfnamefont{I.~F.}\\bibnamefont{Sbalzarini}}, \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Gong}}, \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Cardinale}}, \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Carthel}}, \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Coraluppi}}, \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Winter}}, \emph{et~al.},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{Nat. Methods}\}\textbf{\bibinfo{volume}{11}},\\bibinfo{pages}{281} (\bibinfo{year}{2014})}\BibitemShut{NoStop}% \bibitem[{\citenamefont{Lagarias}\\emph{et~al.}(1998)\citenamefont{Lagarias}, \citenamefont{Reeds}, \citenamefont{Wright},\and\\citenamefont{Wright}}]{fminsearch}% \BibitemOpen \bibfield{author}{\bibinfo{author}{\bibfnamefont{J.~C.}\\bibnamefont{Lagarias}}, \bibinfo{author}{\bibfnamefont{J.~A.}\\bibnamefont{Reeds}}, \bibinfo{author}{\bibfnamefont{M.~H.}\\bibnamefont{Wright}}, \and\\bibinfo{author}{\bibfnamefont{P.~E.}\\bibnamefont{Wright}},\}\href@noop {} {\bibfield{journal}{\bibinfo{journal}{SIAM J. Optim.}\}\textbf{\bibinfo{volume}{9}},\\bibinfo{pages}{112} (\bibinfo{year}{1998})}\BibitemShut{NoStop}% \end{thebibliography}% \end{document} }
\caption{\red{Plot of $T_c/\omega_D$ vs. electron density $n$ for various values of coupling strength, $V/t = 2$, $3$, and $4$. We have use a value of $t_2 = -0.2 t$; the EDOS is plotted in the insert, and resembles very closely the result obtained using DFT.\cite{quan16} As an example, with $V = 2t$ ($V/(16t) = 0.125$), and $\omega_D = 100$ meV, then $T_c \approx 200$ K (at $n\approx 1$). In this range of $\omega_D$ the results for $T_c$ scale with $\omega_D$.}}
\caption[Figure]{Statistics of two Poisson spike trains with a fixed total rate of 1000Hz divided among the two spike trains. (A) Dependence of the Victor-Purpura distance on the rate ratio for three different values of the time scale parameter $q$. (B) \highlight{The operational range for spike time coding is marked in grey.} (C) Same dependence for the ISI-distance, the SPIKE distance (compare \citeauthor{Mulansky15}, \citeyear{Mulansky15}) and the RI-SPIKE-distance (see \citeauthor{Satuvuori17}, \citeyear{Satuvuori17}). (D) Operational ranges for spike time coding. For all distances and rate ratios values can go down to almost zero which is indicated by increasingly darker shades of grey. \label{Fig:V-P_2D}}
\caption[Figure]{The floor effect for spike trains with a very low rate. Mean overall rates (dashed lines) and overall numbers of spikes in the two spike trains at 5\% confidence (solid lines) over 10,000 realizations for a spike train pair with a rate ratio of one (red curves) and a pair with a very high rate ratio (green curves). While the mean values which perfectly match the expectation values (curves not visible) are growing linearly with the rate of the process with higher rate (note the logarithmic x-scale), the actual spike counts can only attain discrete values. Moreover, due to the floor effect it takes \highlight{1.5 spikes} for spike train pairs with rate ratio one and \highlight{3 spikes} for spike train pairs with very high rate ratio to get at least a single spike in either of the two spike trains at 95\% probability. % ##### Old version: Moreover, due to the floor effect it takes a rate of 1.5Hz for spike train pairs with rate ratio one and a rate of 3Hz for spike train pairs with very high rate ratio to get at least a single spike in either of the two spike trains at 95\% probability. ##### \label{Fig:Floor}}
\caption[Figure]{Simple examples illustrating some of the most important differences between the \highlight{Victor-Purpura distance $D_V$ and the SPIKE-distance $D_S$}. (A) Floor effect: SPIKE-distance looks for timing in a single spike. (B) Spike trains with different rates: Victor-Purpura distance ignores timing information of extra spikes. (C) The Victor-Purpura distance is insensitive to exactly matching spikes. (D) Simple clustering example: In contrast to the SPIKE-distance, for spike trains with different rates the Victor-Purpura distance can never really focus on the timing information, even for large q-values. \label{Fig:Examples}}
\caption{2D projection (left) of 5K popular last names' \textit{embeddings}. Same-ethnicity names stand close indicating similar embeddings. Insets (left to right) highlight \textit{White} \fcolorbox{red}{white}{1}, \textit{Black} \fcolorbox{red}{white}{2} and \textit{Hispanic} \fcolorbox{red}{white}{3} names. \textit{API} ( \fcolorbox{blue}{white}{4} and \fcolorbox{blue}{white}{5}) in Fig. \ref{fig:chinese_and_indian}. }
\caption{Two distinct Asian clusters. Left: Chinese/ Vietnamese names (\fcolorbox{blue}{white}{4}). Right: Indian names (\fcolorbox{blue}{white}{5}). It shows name embeddings capture nationality signals.}
\caption{Overview of several common default Bayes Factors (from the $\texttt{R}$-package $\texttt{BayesFactor}$ \citep{BayesFactor}, and their robustness against different kinds of optional stopping. Between parentheses is the type of prior used, in the taxonomy introduced in this paper. The {\color{red} \bf but..} indicates that, formally, prior calibration works for the priors, yet, because we are in the default setting, the Bayes factor is not fully subjective, so prior calibration is not too meaningful --- which is just the main point of this paper. A second caveat is that sampling from the prior on $\mu,\sigma$ is not possible in these situations since the prior is improper; yet the results of \cite{OptionalStoppingTechnical} show that prior calibration can still be mathematically defined, and does hold. }
\caption{Three traps create three rings of magnetic nanoparticles. The rings interact with one another (see \textcolor{blue}{Visualization 3}). [From Masajada et al., Opt. Lett. \textbf{38}, 3910 (2013).]}
\caption{Three traps create three rings of magnetic nanoparticles. The rings interact with one another (see \textcolor{blue}{Visualization 3}). [From Masajada et al., Opt. Lett. \textbf{38}, 3910 (2013).]}
\caption{Frequency spectra for a range of $\alpha$ for NACA 0000, 0006, 0012, and 0018 airfoils with Gurney flap. Dominant vortex shedding frequency - {\large$\bullet$}, harmonic vortex shedding frequency - $\square$. Regions with dominant $St \in [0.12,0.18]$ are classified as 2S and P regimes by the \textcolor{OliveGreen}{green} and \textcolor{Dandelion}{yellow} regions respectively, and $St \in [0.06,0.10]$ as 2P regime by the \textcolor{Red}{red} region.}
\caption{Wake classification diagram for NACA 0000, 0006, 0012, and 0018 airfoils are illustrated. All cases simulated in the current study are categorized into different characteristic wake regimes respect to $h/c$ and $\alpha$ values. Different regimes are described by: steady - \textcolor{NavyBlue}{blue}, 2S - \textcolor{OliveGreen}{green}, P - \textcolor{Dandelion}{yellow}, 2P - \textcolor{Red}{red} and transition between two regimes - \textcolor{Gray}{gray}. The boundaries for all the regimes are obtained by polynomial curve fitting.}
\caption{Illustration of near field flow of three cases classified under the 2S, (a) - (c), P, (d) - (f) and 2P, (g) - (i) regimes. The figure depicts instantaneous vorticity contour plots of flow over NACA 0000 airfoil at $\alpha = 6^{\circ}, 12^{\circ}$ and $15^{\circ}$ with a Gurney flap of $h/c = 0.06$. Corresponding instantaneous lift data, represented by the \textcolor{RedOrange}{\large$\bullet$}, is also displayed on the bottom.}
\caption{\footnotesize Comparison of the type I errors and powers of the SW tests for the NRPPs, NMPPs, and deviance and Pearson residuals. Response variable is simulated from the true model at varying sample sizes, and nonlinear covariate effects of $\beta_1=0.5$ (\protect\solidline \protect\markcircle), 1 (\protect\dashedline \protect\marktriangle) and 2 (\protect\dottedline $+$). True model: NB model with $\mu_{i}=\exp(\beta_1 x^2_{i})$. Wrong model: NB model with $\mu_{i}=\exp(\beta_1 x_{i})$.}
\caption{\footnotesize Comparison of the type I errors and powers of the SW tests for the NRPPs, NMPPs, and deviance and Pearson residuals. Response variable is simulated from the true model at varying sample sizes and the over-dispersion parameters of $k=1$ (\protect\solidline \protect\markcircle), 2 (\protect\dashedline \protect\marktriangle) and 10 (\protect\dottedline $+$). True model: NB model with mean $\exp(\beta_0+\beta_1 x)$. Wrong model: Poisson model with mean $\exp(\beta_0+\beta_1 x)$.}
\caption{\footnotesize Comparison of the type I errors and powers of the SW tests for the NRPPs, NMPPs, and deviance and Pearson residuals. Response variable is simulated from the true model at varying sample sizes and percentages of excessive zeroes of $p=10\% $ (\protect\solidline \protect\markcircle), $30\%$ (\protect\dashedline \protect\marktriangle) and $50\%$ (\protect\dottedline $+$). True model: ZIP model with mean $\exp(\beta_0+\beta_1 x)$. Wrong model: Poisson model with mean $\exp(\beta_0+\beta_1 x)$. }
\caption[TeV Sample distribution in L-L planes]{\label{Fig_corrs} Luminosity in the VHE $\gamma$-ray band vs. luminosity in radio (a), mid-infrared (b), optical (c), X-ray (d), and $\gamma$-ray (e) bands in logarithmic scale. The different symbols represent the data of different groups (see text): low state group A (\textcolor{blue}{circle}), high state group A (\textcolor{red}{triangle}) and group B (\textcolor{green}{cross}). The correlation functions of the various groups are shown with low state group A (\textcolor{blue}{blue dashed line}), the combined dataset of high state group A and group B (\textcolor{red}{red line}), and the combined dataset of low state group A and group B (\textcolor{blue}{blue line}).}
\caption{Wavelength depended atmospheric opacity at different locations in the atmosphere of HD 189733 b {\color{\clr} for the M profile and under nominal eddy conditions}. The background coloured contours present the contribution of gaseous components, while the foreground black line contours correspond to the aerosol {\color{\clr}extinction} contribution.}
\caption{\label{fig3} (a) Echo-detected field sweep. $A_e$ (open circles) is shown as a function of $B_0$ (parallel to the wire). (b) COMSOL\textsuperscript \textregistered simulation of the $\epsilon_{100}$ component of the strain field in the silicon around the wire. (c) Spin coherence time measurement at $B_0=3.74\,$mT. $A_e$ plotted as a function of the delay $2\tau$ between $\pi/2$ pulse and echo (red triangles). An exponential fit (black solid line) yields $T_2=1.65\pm0.03\,$ms. (d) $T_1$ and $T_2$ as a function of $B_0$. Error bars are within the marker size.}
\caption{Additional \aastex\symbols}
\caption{Low-dimensional visualization of \textit{Abalone} and \textit{ANN-Thyroid-1v3} using t-SNE. Plus signs are anomalies and circles are nominals. A \textcolor{red}{red} coloring indicates that a true anomaly point was queried. A \textcolor{green}{green} indicates a nominal point was queried. Grey circles correspond to unqueried nominals. To make unqueried anomalies stand out visually, we indicate them with \textcolor{blue}{blue} plus signs.}
\caption{Random trees in Isolation Forest (IF) for synthetic data. The points in \textcolor{red}{red} are true anomalies; points in gray are true nominals. Figure~\ref{fig:ifor_1_trees} shows the leaf node regions for a single tree generated by random IF splits. Figure~\ref{fig:ifor_contours_1_trees} shows the contours of anomaly scores assigned to the nodes of this tree. Deeper \textcolor{red}{red} means more anomalous; deeper \textcolor{blue}{blue} means more nominal. The \textcolor{red}{red} circles are the true anomalies among the top ranked 35 instances. The \textcolor{britishracinggreen}{green} circles are the true nominals among the top ranked 35 instances. The left sidebar in Figure~\ref{fig:ifor_contours_1_trees} shows the ranking of true anomalies (\textcolor{red}{red} dots). Ideally, true anomalies should be near the top on this bar.}
\caption{Incorporating feedback in Isolation Forest (IF) for synthetic data (Figure~\ref{fig:synthetic_data}). Figures~\ref{fig:ifor_iter_00} -- \ref{fig:ifor_iter_32} show anomaly score contours in the same way as explained in Figure~\ref{fig:ifor_contours_trees}. The \textcolor{red}{red} and \textcolor{britishracinggreen}{green} circles are the instances that have been presented for labeling. The x-axis in Figure~\ref{fig:ifor_synth_num_seen} represents the number of instances presented to the analyst, and the y-axis represents the number of true anomalies discovered. The \textcolor{red}{red} curve in Figure~\ref{fig:ifor_synth_num_seen} shows the number of true anomalies discovered when we incorporate feedback; the \textcolor{blue}{blue} curve in Figure~\ref{fig:ifor_synth_num_seen} shows the number of true anomalies discovered when no feedback was incorporated.}
\caption{ a) Illustration of the rules used to compute the fractal dimension of the complete and accessible perimeters with the yardstick method. Suppose the sticks start to follow the coast from the bottom. The green circle shows the area of coast covered by a particular stick. The \textbf{X}'s represent the next possible starting points of that particular stick. If the closest point along the coast (red \textcolor{red}{\textbf{X}}) is always chosen as the next starting point, we obtain the complete perimeter. If, on the other hand, the most distant point (blue \textcolor{blue}{\textbf{X}}) is chosen, then we obtain the accessible perimeter. b) Paths made by sticks of equal sizes of the complete (blue sticks) and accessible (red sticks) perimeters.}
\caption{\label{Nokia} Four networks of investor group trading relationships in Nokia security. Investor group positions are fixed in all four plots. Node sizes depend on node degrees in each network. The first network (a) is inferred using the C3NET algorithm on the original data set. The second network (b) is inferred by bagging C3NET. For the third (c) and fourth (d) networks, the whole six-year period is divided into 12 six-month sub-periods. For each of those 12 sub-periods, a C3NET and bagged C3NET networks are inferred. Then those 12 networks are aggregated into a final network that covers the whole 6-year period. \protect\tikz{\protect\draw[black, thin, fill=cc686e9] (4pt,4pt) circle [radius=4pt];} - Households, \protect\tikz{\protect\draw[black, thin, fill=cea8615] (4pt,4pt) circle [radius=4pt];} - Non-profit organizations, \protect\tikz{\protect\draw[black, thin, fill=cRestWorld] (4pt,4pt) circle [radius=4pt];} - Other companies \protect\tikz{\protect\draw[black, thin, fill=c00caff] (4pt,4pt) circle [radius=4pt];} - Financial and insurance companies, \protect\tikz{\protect\draw[black, thin, fill=cFinancial] (4pt,4pt) circle [radius=4pt];} - Government institutions.}
\caption{\label{Nokia6monthSignLinks}54 most re-occurring links in 12 Nokia networks estimated over non-overlapping 6-month periods. \protect\tikz{\protect\draw[black, thin, fill=cgreeen] (0pt,0pt) rectangle (8pt, 8pt);} - inferred relationship, \protect\tikz{\protect\draw[black, thin, fill=white] (0pt,0pt) rectangle (8pt, 8pt);} - no relationship.}
\caption{\label{MultiNetworks}Networks summarizing investor group trading similarities in 100 securities over 6 years. The starting point for both networks is a set of $1\,200$ networks inferred for each security over 12 six-month, non-overlapping periods. The difference between networks comes from the aggregation order. The first network is first aggregated security-wise and then time-wise, while the third network is aggregated in reverse order. Network (3) in Figure \ref{BaggingLayers} represents network (a), while network (5) in the same figure represents network (b). \protect\tikz{\protect\draw[black, thin, fill=cc686e9] (4pt,4pt) circle [radius=4pt];} - Households, \protect\tikz{\protect\draw[black, thin, fill=cea8615] (4pt,4pt) circle [radius=4pt];} - Non-profit organizations, \protect\tikz{\protect\draw[black, thin, fill=cRestWorld] (4pt,4pt) circle [radius=4pt];} - Other companies \protect\tikz{\protect\draw[black, thin, fill=c00caff] (4pt,4pt) circle [radius=4pt];} - Financial and insurance companies, \protect\tikz{\protect\draw[black, thin, fill=cFinancial] (4pt,4pt) circle [radius=4pt];} - Government institutions.}
\caption{\textcolor[rgb]{0.98,0.00,0.00}{(color online)} \textbf{ Order parameters in colloidal glasses.} {\bf a,} Particles in the polar coordinate system with reference particle i as pole O and with x-axis as polar axis. We use $g_{i}(r,\theta)$ to characterize the probability of finding particles at the point $(r,\theta)$ with radial coordinate $r$ and angular coordinate $\theta$. Spatial distribution of {\bf b,} orientational fluctuation function $O_{i}$, {\bf c,} translational fluctuation function $T_{i}$, and {\bf d,} local Debye-Waller factor $\alpha_i$. All three parameters are plotted by the ranks of individual particles. Spearman's rank correlation between $\alpha_i$ and {\bf e,} $O_{i}$, {\bf f,} between $\alpha_i$ and $T_{i}$.}
\caption{\textcolor[rgb]{0.98,0.00,0.00}{(color online)} \textbf{ Emergence of orientational order during glass transition.} {\bf a,} Self intermediate scattering functions for liquids in different packing fractions $\phi$. {\bf b,} The dependence of $\chi_4$ on duration $\Delta t$. $\chi_4$ reaches its maximum value with $\Delta t$ close to $\alpha$-relaxation time as indicated by the arrows. {\bf c,} $\phi$ dependence of the rank correlation between $\alpha_i$ and distribution fluctuations $O_{i}$ ($T_{i}$). There is a crossover at $\phi_c\sim0.7$ where orientational order becomes better correlated with dynamical heterogeneity than translational order. Spatial distribution of polar vector $g_i^{\rm max}$ pointing to the direction of maximum $g_i(\theta)$ at {\bf d,} $\phi=0.83$ and {\bf e,} $\phi=0.54$. {\bf f,} Average dot production between particle $i$'s polar vector $g_i^{\rm max}$ and particle $j$'s polar vector $g_j^{\rm max}$ at the distance $r$ to the particle $i$. The polar vectors are correlated within three shells. }
\caption{Comparison of performance between grasp deemed {\color{blue}best} and {\color{red}worst} with respect to the metric $m_{ag}$ measured at the ``average'' state.}
\caption{ Visualization of labeled trajectories and motion models of three layers. The first figure is the original image covered with labeled trajectories: the trajectories are clustered into three layers: \textcolor{red}{background layer(Red)}, \textcolor{blue}{car1 layer(Blue)} and \textcolor{green}{car2 layer(Green)}. The following three figures show the estimated motion models for the three layers, where the arrow and length indicate the direction and magnitude of motion. }
\caption{ Visualization of labeled trajectories and motion models of three layers. The first figure is the original image covered with labeled trajectories: the trajectories are clustered into three layers: \textcolor{red}{background layer(Red)}, \textcolor{blue}{car1 layer(Blue)} and \textcolor{green}{car2 layer(Green)}. The following three figures show the estimated motion models for the three layers, where the arrow and length indicate the direction and magnitude of motion. }
\caption{Legend: {\color{red}$\times$} is the true parameter $p^{\star}$; {\color{blue}$+$} are the sampled parameters; {\color{green} $+$} is the final selected parameter in the interval $t\in[t_k,t_{k+1})$; the potentially optimal hyperectangles at each iteration is highlighted with thicker lines. \label{fig:res} }
\caption{\textbf{A Critical Density of Bookmarks is Required for Stable Domain Formation.} (\textbf{A}) Using the clustered pattern of bookmarks at different densities $\phi$, we quantify the deviation from a ``perfect'' block-like epigenetic pattern. To do this we define the ``fidelity'', $\chi$, as $1-\Delta^2$ where $\Delta^2=Var \left[P_{\rm red}(i),\Pi(i)\right]$, i.e. the variance of the probability $P_{\rm red}(i)$ of observing a red bead at position $i$ with respect to the perfect square wave $\Pi(i)=0.5\left[\sgn\left(\sin{\left(\pi i/n_d\right)}\right)+1\right]$, where $n_d$ is the number of beads in a domain (here $n_d = 100$). The fidelity $\chi$ jumps abruptly from a value near its lower bound of $1/2$ towards unity, at the critical $\phi_c\simeq 0.04$. (\textbf{B,C}) Kymographs representing the behaviour of the system at the points circled in red and grey in (\textbf{A}). } \label{fig:accuracy} \end{figure} %fig 4: heterochrom +bookmarks replication and excision \begin{figure*}[t!] \centering \includegraphics[width=0.8\textwidth]{Panel_heterVSbookmarks.png} \caption{\textbf{Asymmetric Interactions and Bookmark Excision but not DNA Replication Affect the Epigenetic Landscape.} (\textbf{A-B}) Here we consider the case in which blue-blue interactions are stronger than red-red ones. We set $\epsilon_{\rm blue}=1 k_BT$ and $\epsilon_{\rm red}=0.65 k_BT$ with $f=2$. The central region of a chromatin segment $L=2000$ beads long is initially patterned with bookmarks at density $\phi=0.1>\phi_c$ (this region is indicated in the kymograph by an orange arrowhead). Blue beads invade non-bookmarked regions thanks to the thermodynamic bias whereas the local red state is protected by the bookmarks. (\textbf{C-D}) The chromatin fibre undergoes replication cycles which extensively perturb the pattern of PTM of histones on chromatin. A semi-conservative replication event (R) occurs every $10^5$ $\tau_{Br}$ and half of the (non-bookmarked) beads become grey. The epigenetic pattern is robustly inherited. (\textbf{E-F}) The chromatin fibre undergoes semi-conservative replication followed by excision of bookmarks (R+E). At each time, 1/4 of the initial bookmarks are removed and turned into grey (recolourable) beads. The epigenetic pattern is inherited until $\phi<\phi_c$. At this point, the central red domain is either immediately lost (not shown) or it can be sustained through some replication cycles (\textbf{F}) by local memory (LM). \red{See also Suppl. Movie 5 for a direct comparison of the behaviour with and without bookmarks.}}% This entails that bookmarks are crucial element that allow epigenetic memory. In addition, this figure suggests that activated bookmarks may be able to disrupt an homogeneous landscape and establish \emph{de novo} transient epigenetic domains~\cite{Ciabrelli2017}.} \label{fig:excision} \end{figure*} %%fig 5 Real drosophila chromosome \begin{figure*}[t!] \centering \includegraphics[width=0.83\textwidth]{Panel_drosophila.png} \caption{\textbf{GBM Alone is Able to Recapitulate the Distribution of Polycomb Marks in \emph{Drosophila} S2 cells.} Here we perform chromosome-wide simulations of Ch3R of Drosophila S2 cells at $3$ kbp resolution ($L=9302$) with GBM. (\textbf{A}) The location of PSC/PRE (bookmarks) are mapped onto beads using ChIP-Seq data from Ref.~\cite{Follmer2012}. Using the ``9-states'' HMM data~\cite{Kharchenko2011}, gene deserts (regions lacking any mark in ChIP-seq data, state 9), promoter/enhancers (state 1) and transcriptionally active regions (states 2-4) are permanently coloured grey, red and green, respectively. The remaining beads ($\sim$20\%) are initially unmarked (white) and may become either heterochromatin (blue) or polycomb (purple). (\textbf{B}) \emph{In silico} ChIP-seq data for H3K27me3 (top half, purple lines) is compared with \emph{in vivo} ChIP-seq~\cite{Kharchenko2011} (bottom half, grey line). Small orange arrows at the top of the profile indicate the location of the bookmarks. The excellent quantitative agreement between the datasets is captured by the Pearson correlation coefficient $\rho=0.46$ -- to be compared with $\rho=0.006$ obtained between a random and the experimental datasets. We highlight that not all the bookmarked beads foster the nucleation of H3K27me3 domains (see big purple/orange arrowheads in the insets, corresponding to the HOX cluster). The reason can be found by analysing the 3D conformations of the chromosome (\textbf{C}). The non-nucleating bookmarks (orange arrowheads), although near in 1D, are found far from potential target beads in 3D space (purple arrowheads) and so fail to yield large H3K27me3 domains. \red{See also Suppl. Movie 6 for a direct comparison of the results with and without bookmarks.} } %This suggests an intriguing relationship between 1D information and 3D organisation in preserving epigenetic memory~\cite{Ciabrelli2017}. We finally speculate that the these bookmarks may be functioning as nucleation points in other cell lines which display different 3D organisations. %The p-value testing independence between the simulated and experimental data-sets is $p=3$ $10^{-13}$, to be compared with $p=0.75$ obtained using a random data-set against the experimental one. \label{fig:drosophila} \end{figure*} \subsection{Biasing Epigenetic Landscapes with Asymmetric Interactions} Thus far, we have considered symmetric interactions between like-coloured beads. In other words, red-red and blue-blue interaction strengths were equal. However, such binding energies may differ if mediated by distinct proteins. Consider the case where red and blue marks encode Polycomb repression and constitutive heterochromatin, respectively. If the blue-blue interaction is larger than the red-red one, the thermodynamic symmetry of the system is broken and the blue mark eventually takes over all non-bookmarked regions (Fig.~\ref{fig:excision}A). However, if there are bookmarks for the red mark, they locally favour the red state, whereas the stronger attraction globally favours the blue mark. This competition creates an additional route to form stable domains as exemplified in Figure~\ref{fig:excision}A,B. Here, red bookmarks (identified by orange beads) are concentrated in the central segment of a chromatin fibre. Starting from a swollen and epigenetically disordered fibre, where red, blue and grey beads are equal in number, we observe that blue marks quickly invade non-bookmarked regions and convert red beads into blue ones (a process mimicking heterochromatic spreading \emph{in vivo}~\cite{Hathaway2012}). However, the central segment containing the bookmarks displays a stable red domain (Fig.~\ref{fig:excision}A,B). %Interestingly, bookmarking is necessary but not sufficient to yield stable coexistence between domains, as we further require that the attraction between red bookmarks and red beads is large enough to locally overcome the thermodynamic preference of the blue marks. \subsection{Bookmark Excision but not DNA Replication Destabilises the Epigenetic Landscape} We next asked whether the epigenetic pattern established through GBM is also stable against extensive perturbations such as DNA replication. In order to investigate this scenario we simulated semi-conservative replication of the chromatin fibre by replacing half of the (non-bookmarked) beads with new randomly coloured beads~\cite{Berry2017}. In Figure~\ref{fig:excision}C-D we show that our model can ``remember'' the established epigenetic pattern through multiple rounds of cell division. Importantly, the combination of ``memory'' and local epigenetic order (via bookmarks) may allow cells to display ``epialleles'', i.e., alleles with different transcriptional behaviours thus explaining local (or ``cis-'') memory~\cite{Berry2015,Berry2017}. We next considered a set-up relevant in light of recent experiments in {\it Drosophila}~\cite{Laprell2017,Coleman2017}, where the role of Polycomb-Response-Elements (PREs) in epigenetic memory was investigated. In these works, polycomb-mediated gene repression was perturbed as a consequence of artificial insertion or deletion of PREs. In Figure~\ref{fig:excision} we thus performed a simulated dynamic experiment where replication was accompanied by random excision of bookmarks~\cite{Laprell2017} (Fig.~\ref{fig:excision}E,F); in practice, we remove $1/4$ of the initial number of bookmarks at each replication event. Then each ``cell cycle'' successively dilutes the bookmarks which at some point can no longer sustain the local red state and the region is consequently flooded with blue marks. Importantly, the system does not display immediate loss of the red domain as soon as $\phi<\phi_c$; on the contrary, this domain is temporarily retained through local memory (see Fig.~\ref{fig:excision}F, LM)~\cite{Angel2011,Berry2015,Berry2017}. This originates from an enhanced local density of marks together with the positive read/write feedback (see SM). [The persistence of the local memory can be tuned via the parameters of our polymer model.] These results are again consistent with experiments, as regions of the \emph{Drosophila} genome marked with H3K27me3 are only gradually lost after PRE excision~\cite{Laprell2017}. Similarly, epialleles have been observed to be temporarily remembered across cell division~\cite{Berry2015}. \red{We finally highlight that the results presented in Fig.~\ref{fig:excision} are independent on the chosen initial configuration. In SM (Figs.S4-S5) we show that starting from a collapsed and epigenetically disordered chromatin (CD phase), resembling heavily condensed and sparsely marked mitotic structures, leads to the same behaviour and strongly supports the robustness of our findings.} \subsection{Chromosome-Wide Simulations Predict the Epigenetic Landscape in \emph{Drosophila}} Simplified models considered thus far are useful to identify generic mechanisms; we now aim to test our model in a realistic scenario. To do so, we perform polymer simulations of the whole right arm of chromosome 3 in {\it Drosophila} S2 cells. Bookmarks (orange, in Fig.~\ref{fig:drosophila}) are located on the chromosome using PSC ChIP-Seq data~\cite{Follmer2012}, as PSC binds to PREs during inter-phase and mitosis~\cite{Follmer2012} as well as recruiting PRC2 (via molecular bridging). Some other beads are permanently coloured according to the ``9-state'' Hidden Markov Model (HMM,~\cite{Kharchenko2011}). If they correspond to gene deserts (state 9), promoter/enhancers (state 1) or transcriptionally active regions (states 2-4) they are coloured grey, red and green, respectively. We further introduce an interaction between promoter and enhancer beads to favour looping, plus, an attractive interaction between gene desert (grey) beads mimicking their compaction by H1 linker histone~\cite{Sexton2012} (see SM for full list of parameters). The remaining 20\% of the polymer is left blank and these ``unmarked'' beads are allowed to dynamically change their chromatin state into heterochromatin (blue) or polycomb (purple) according to our recolouring scheme. We evolve the system to steady state and we evaluate the probability of finding a Polycomb mark at a certain genomic position. [To determine these probability, a bookmarked bead is counted as bearing the H3K27me3 mark when it is near beads with polycomb marks, or within large stretches of bookmarked beads]. This provides us with an \emph{in silico} ChIP-seq track for Polycomb marks which can be compared with \emph{in vivo} ChIP-Seq data~\cite{Kharchenko2011} (see Fig.~\ref{fig:drosophila}B). The two are in excellent agreement (Pearson correlation coefficient $\rho=0.46$, against $\rho=0.006$ for a random dataset). Remarkably, not all bookmarked segments (orange) are populated by Polycomb marks; instead we observe that H3K27me3 spreading requires appropriate 3D folding (Fig.~\ref{fig:drosophila}B-C, insets). Bookmarks which do not contact other bookmarks due to the local epigenetic landscape do not nucleate H3K27me3 spreading. Again, this is consistent with 3D chromatin conformation being crucial for the spreading and establishment of epigenetic patterns~\cite{Ciabrelli2017,Engreitz2013,Deng2014}. \section{Discussion} We proposed and investigated a new biophysical mechanism for the \emph{de novo} establishment of epigenetic domains and their maintenance through interphase and mitosis. Our simplest model requires only one element: a positive feedback between readers (e.g., binding proteins HP1, PRC2, etc.) and writers (e.g., methyltransferases SUV39, EzH2, etc.). We performed large-scale simulations in which chromatin is modelled as a semi-flexible bead-and-spring polymer chain overlaid with a further degree of freedom representing a dynamic epigenetic patterning. Specifically, each bead is assigned a colour corresponding to the local instantaneous epigenetic state. Readers are implicitly included by setting an attraction between like-coloured beads~\cite{Brackley2016nar,Barbieri2012}, whereas writers are modelled by performing re-colouring moves according to realistic and out-of-equilibrium rules~\cite{Dodd2007,Dodd2011} (see Fig.~\ref{fig:model}). %Intriguingly, we note that the main qualitative behaviours are retained when detailed balance is restored~\cite{Michieletto2016prx} (see SM). %and (ii) the presence of ``bookmarks'', i.e. DNA-sequence specific transcription factors which can recruit read/write machineries. We find that, if read-write positive feedback is sufficiently strong, a single histone mark can spread over the whole fibre and drives a discontinuous transition to a collapsed-ordered state (see Fig.~\ref{fig:phasediagr}). This state is stable and robust against extensive perturbations such as those occurring during replication~\cite{Talbert2006,Zentner2013,Cavalli2013}, when most histones are removed or displaced~\cite{Alberts2014,Zentner2013,Festuccia2017}. In other words, our model displays ``epigenetic memory''. The main limitation of this simple model is that epigenetic order in real chromosomes is local, rather than global. Distinct epigenetic domains coexist on a chromosome, thereby forming an ``heterogeneous'' epigenetic pattern. Our main result is that this feature of real chromosomes can be reproduced by our model when we include genomic bookmarking (GBM). %here the is a lot of peter -- %i add ``dynamically associated'' instead of bound %Bookmarks are usually considered to be transcription factors that remain (dynamically) associated through mitosis so gene activity can be inherited from one interphase to the next~\cite{Kadauke2013a,Deluz2016,Sarge2005}. %There are many classical examples, including PcG~\cite{Lanzuolo2007,Laprell2017}, GATA~\cite{Kadauke2012}, Esrbb~\cite{Festuccia2016}, PSC~\cite{Follmer2012} and Sox2~\cite{Teves2016}. %This list will probably soon expand as many TFs are now know to remain associated to DNA through mitosis – contrary to what had been thought~\cite{Teves2016}. Here, we envisage bookmarks which can perform functions typical of many TFs: they recruit read/write machineries, and hence nucleate the spreading of epigenetic marks and the establishment of epigenetic domains. We also assumed that bookmarking TFs are permanently bound to DNA, however our conclusions should hold even for dynamic bookmarks that switch between bound and unbound state~\cite{Brackley2016ephe,Teves2016}. %fig 6 a model for differentiation w bookmarks+pve feedback \begin{figure*}[t!] \centering \includegraphics[width=0.80\textwidth]{Panel_differentiationWbookmarks_model.png} \caption{\textbf{Model for Cellular Differentiation}. We speculate that cellular differentiation may be driven by a two-step process. First, sequence-specific factors (bookmarks) are expressed as a consequence of environmental and positional cues. Second, the positive feedback set up by read/write machineries drives the establishment and maintenance of tissue-specific epigenetic patterns. As a consequence, genomic bookmarks are key targets to understand cellular differentiation and reprogramming.} \label{fig:differentiation} \end{figure*} %The inclusion of GBM allows us to identify some basic rules required for the establishment of stable epigenetic domains. For instance, GBM must be locally clustered and above a critical density (about one in 400 nucleosomes, Fig.~\ref{fig:accuracy}). This scenario may be relevant to cases where regions bookmarked with ``repressive'' GBM coexist alongside other regions bookmarked with ``active'' GBM. We find that stable domains can be formed with only one type of bookmark when the competing epigenetic mark is thermodynamically favoured (Fig.~\ref{fig:excision}). This result rationalises the common understanding that heterochromatin can spread at lengths (blue mark in Fig.~\ref{fig:excision}A,B) and it is stopped by actively transcribed (bookmarked) regions. Further, it is in agreement with recent genome editing experiments in {\it Drosophila}: when PRE is inserted into the genome, it provides a bookmark for H3K27me3 which leads to spreading of that mark~\cite{Laprell2017}, whereas PRE excision leads to (gradual) loss of the mark~\cite{Laprell2017} (Fig.~\ref{fig:excision}). Additionally, the expression of HOX and other Polycomb-regulated genes (which contain multiple PREs) is predicted by our model to be less sensitive to deletion of single PREs~\cite{De2016}. We suggest that this is because domains remain stable if bookmark density is kept above the critical threshold (Fig.~\ref{fig:accuracy}). %These findings can be readily explained within our model if the 3D attraction between polycomb marks is slightly weaker than that between, e.g., H3K9me3-modified regions (Fig.~\ref{fig:excision}). Our results strongly suggest that bookmarks can establish specific epigenetic domains by exploiting the local diffusion of chromatin and thereby ``infecting'' 3D-proximal chromatin segments. The local increase in the density of a mark is then stopped either by thermodynamics (Fig.~\ref{fig:excision}A) or competition with other bookmarks (Fig.~\ref{fig:bookmarks1}B). Crucially, our model does not require any boundary element to stop the spreading of marks, which is instead self-regulated. Losing bookmarks (via artificial excision or DNA mutation) will thus impair the ability of cells to inherit the cell-line-specific epigenetic patterns. In addition, we argue that newly activated bookmarks (for instance subsequently to inflammation response or external stimuli~\cite{Kirmes2015,Feuerborn2015,Wood2014}) may drive the \emph{de novo} formation of transient epigenetic domains which allow the plastic epigenetic response to environmental changes. We show that our model can recreate the pattern of H3K27me3 in \emph{Drosophila} S2 cells starting solely from the position of PSC proteins acting as Polycomb bookmarks % transcriptionally-determined pattern of TFs mix of static and dynamic epigenetic marks, with ~20\% of the genome initially being unmarked. Then, our recolouring scheme completes the epigenetic landscape, and quantitatively recapitulates distributions of H3K27me3 seen \emph{in vivo}. Intriguingly, our simulations show that not all bookmarks end up in H3K27me3 domains; whether or not they do, depends on their network of chromatin contacts in 3D. This is agreement with recent experiments~\cite{Ciabrelli2017,Engreitz2013,Deng2014} and it is also reminiscent of the well-known position effect according to which the activity of a gene depends on its local environment~\cite{Feuerborn2015}. \red{While our framework can be directly applied to model competition between repressive epigenetic marks, the deposition of active marks may be better modelled as resulting from a co-transcriptional positive feedback loop. In light of this, in the SM we show that a model with thermodynamically favoured heterochromatin competing with local recolouring due to transcription leads to results that are qualitatively similar to those presented in the previous sections, as long as promoters are seen as bookmarks for active marks (see SM for more details).} Our results also prompt several further questions. First, starting from a stem cell, how might different cell lineages be established? We suggest that environmental and morphological cues trigger production of lineage-specific bookmarks such as GATA~\cite{Kadauke2013a} and PSC~\cite{Follmer2012}, which nucleate the positive feedback between readers and writers to generate and sustain new cell-line specific epigenetic patterns (Fig.~\ref{fig:differentiation}). Thus, bookmarks are here envisaged as key elements that should be targeted in order to understand, and manipulate, cellular differentiation. Second, how might reprogramming factors like Nanog work? We argue that their binding can ``mask'' the action of pre-existing bookmarks, thereby allowing the establishment of new epigenetic patterns~\cite{Festuccia2016} (see also BioRxiv: https://doi.org/10.1101/127522). In conclusion, we have extended the existing notion of GBM to include the ability of nucleating the spreading of epigenetic marks by triggering \emph{local} read/write feedback loops. This model predicts the \emph{de novo} establishment of heterogeneous epigenetic patterns which can be remembered across replication and can adapt in response to GBM-targeted perturbations. Within our framework, architectural elements such as CTCF~\cite{Alberts2014}, Cohesins~\cite{Fudenberg2016} and SAF-A~\cite{Nozawa2017} may provide the initial 3D chromatin conformation upon which the GBM-driven establishment of epigenetic landscape takes place. \section{ACKNOWLEDGEMENTS} We acknowledge the European Research Council for funding (Consolidator Grant THREEDCELLPHYSICS, Ref. 648050). Work in the Papantonis lab is supported by CMMC core funding. The authors thank C. A. Brackley, A. Buckle, N. Gilbert, J. Allan and G. Cavalli for insightful remarks on the manuscript. %\bibliographystyle{apsrev4-1} \bibliographystyle{nar} \bibliography{Epigenetics} \end{document} }
\caption{\textbf{Asymmetric Interactions and Bookmark Excision but not DNA Replication Affect the Epigenetic Landscape.} (\textbf{A-B}) Here we consider the case in which blue-blue interactions are stronger than red-red ones. We set $\epsilon_{\rm blue}=1 k_BT$ and $\epsilon_{\rm red}=0.65 k_BT$ with $f=2$. The central region of a chromatin segment $L=2000$ beads long is initially patterned with bookmarks at density $\phi=0.1>\phi_c$ (this region is indicated in the kymograph by an orange arrowhead). Blue beads invade non-bookmarked regions thanks to the thermodynamic bias whereas the local red state is protected by the bookmarks. (\textbf{C-D}) The chromatin fibre undergoes replication cycles which extensively perturb the pattern of PTM of histones on chromatin. A semi-conservative replication event (R) occurs every $10^5$ $\tau_{Br}$ and half of the (non-bookmarked) beads become grey. The epigenetic pattern is robustly inherited. (\textbf{E-F}) The chromatin fibre undergoes semi-conservative replication followed by excision of bookmarks (R+E). At each time, 1/4 of the initial bookmarks are removed and turned into grey (recolourable) beads. The epigenetic pattern is inherited until $\phi<\phi_c$. At this point, the central red domain is either immediately lost (not shown) or it can be sustained through some replication cycles (\textbf{F}) by local memory (LM). \red{See also Suppl. Movie 5 for a direct comparison of the behaviour with and without bookmarks.}}
\caption{\textbf{GBM Alone is Able to Recapitulate the Distribution of Polycomb Marks in \emph{Drosophila} S2 cells.} Here we perform chromosome-wide simulations of Ch3R of Drosophila S2 cells at $3$ kbp resolution ($L=9302$) with GBM. (\textbf{A}) The location of PSC/PRE (bookmarks) are mapped onto beads using ChIP-Seq data from Ref.~\cite{Follmer2012}. Using the ``9-states'' HMM data~\cite{Kharchenko2011}, gene deserts (regions lacking any mark in ChIP-seq data, state 9), promoter/enhancers (state 1) and transcriptionally active regions (states 2-4) are permanently coloured grey, red and green, respectively. The remaining beads ($\sim$20\%) are initially unmarked (white) and may become either heterochromatin (blue) or polycomb (purple). (\textbf{B}) \emph{In silico} ChIP-seq data for H3K27me3 (top half, purple lines) is compared with \emph{in vivo} ChIP-seq~\cite{Kharchenko2011} (bottom half, grey line). Small orange arrows at the top of the profile indicate the location of the bookmarks. The excellent quantitative agreement between the datasets is captured by the Pearson correlation coefficient $\rho=0.46$ -- to be compared with $\rho=0.006$ obtained between a random and the experimental datasets. We highlight that not all the bookmarked beads foster the nucleation of H3K27me3 domains (see big purple/orange arrowheads in the insets, corresponding to the HOX cluster). The reason can be found by analysing the 3D conformations of the chromosome (\textbf{C}). The non-nucleating bookmarks (orange arrowheads), although near in 1D, are found far from potential target beads in 3D space (purple arrowheads) and so fail to yield large H3K27me3 domains. \red{See also Suppl. Movie 6 for a direct comparison of the results with and without bookmarks.} }
\caption{End-to-End Training. mAP when fine-tunning different parts of the pipeline. In \colorbox{lightgray}{gray}, the modules that are fine-tunned in every experiment.}
\caption{{\bf Top row}: Positions in physical kpc of gas (black dots) and star (orange dots) particles for the galaxy of the simulated sample with the highest SFR. This object has total (i.e. as determined by {\small{SUBFIND}}) M$_{\star}=4.30\times10^{10}$ M$_\odot$, M$_{\rm gas}=1.94\times10^{11}$ M$_\odot$, M$_{\rm tot}=1.89\times10^{12}$ M$_\odot$, SFR $=6.77$ M$_\odot$ yr$^{-1}$ and $\log\big($sSFR$/$yr$^{-1}\big)=-9.80$. The stellar and gas mass inside the ($30$ kpc)$^3$ cube are, respectively, M$_{\rm\star,c}=3.51\times10^{10}$ M$_\odot$ and M$_{\rm gas,c}=1.19\times10^{10}$ M$_\odot$. Left panel: xz edge-on projection. Middle panel: xy face-on projection. Right panel: yz edge-on projection. {\bf Bottom left panel}: SFR$-$M$_{\star}$ relation for our simulated galaxies. The grey solid (+ triple dot-dashed) line is the best fit SFR$-$M$_{\star}$ relation for star forming local galaxies of \citet{renzinipeng2015}. The synthetic sample represents disc galaxies on the main sequence. {\bf Bottom right panel}: simulated specific SFR$-$M$_{\star}$ relation. The vertical and horizontal dashed lines mark, respectively, the median M$_{\star}$ and sSFR of the final sample: $\log(\tilde{{\rm M}}_{\star}/$M$_\odot)=9.78$, $\log(\tilde{\rm sSFR}/$yr$^{-1})=-10.04$. To facilitate the subsequent analysis, we divided the plot in four quadrants: Q$1$ = low M$_{\star}-$ low sSFR, \color{red} Q$2$ \color{black}= low M$_{\star}-$ high sSFR, \color{blue} Q$3$ \color{black}= high M$_{\star}-$ low sSFR and \color{orange} Q$4$ \color{black}= high M$_{\star}-$ high sSFR (see text).} \label{fig_gal11_proj} \end{figure*} This configuration comes in two different setups: {\textit {Ref-L025N0752}} and {\textit {Recal-L025N0752}}. The first one is the initial reference setup, while in the second the subgrid stellar and AGN feedback parameter values were re-calibrated to better match the observed low redshift GSMF \citep{schaye2015}. In practice, the main difference is that feedback is slightly more effective in {\textit {Recal}}. This prevents overcooling problems and leads to more realistic gas and stellar distributions in galaxies. For this reason, we used {\textit {Recal-L025N0752}} for the analysis presented in this paper. \subsection{Galaxy Sample} \label{gsample} For this project, we have developed a pipeline to optimise the analysis of EAGLE simulations. The pipeline first (and only once) loads the original EAGLE snapshot and reads the output of {\small SUBFIND} (i.e. the catalogue of substructures within DM haloes). Then it extracts and saves only the information needed by the user (e.g. only galaxies with stellar mass, M$_{\star}$, and$/$or specific star formation rate, sSFR $\equiv$ SFR/M$_{\star}$, in a given range, et cetera). Since the original snapshot can be several gigabytes in size, this procedure drastically reduces both memory and time access requirements, speeding up considerably the subsequent analysis. We used the pipeline to extract cubes of size $30$ physical kpc around all the galaxies with M$_{\star} \ge 10^{9}$ M$_\odot$ from {\textit {Recal-L025N0752}} at $z=0$ (taking into account periodic boundary conditions and a buffer of extra $70$ kpc per side for SPH interpolation, see Section \ref{binn}). Each cube is a {\small GADGET} format file containing a new header and the following information for gas ($g$) and star ($s$) particles: position ($g$+$s$), velocity ($g$+$s$), mass ($g$+$s$), temperature ($g$), density ($g$), smoothing length ($g$), SFR ($g$), metallicity ($g$+$s$) and age ($s$). The initial sample included $266$ galaxies. \begin{figure*} \centering \vspace{0.3cm} \includegraphics[width=7.85cm, height=7.1cm]{color_rotational_mapxz_200} \includegraphics[width=7.85cm, height=7.1cm]{color_rotational_mapyz_200} \vspace{0.25cm} \includegraphics[width=7.85cm, height=7.1cm]{color_mapxz_200} \includegraphics[width=7.85cm, height=7.1cm]{color_mapxz_Ha_200} \vspace{0.25cm} \includegraphics[width=7.85cm, height=7.1cm]{color_mapxy_200} \includegraphics[width=7.85cm, height=7.1cm]{color_mapxy_Ha_200} \caption{{\bf Top row}: examples of mean (rotation) velocity maps created using all the gas in the cube. Left panel: edge-on xz projection $-$ mean $v_{\rm y}$. Right panel: edge-on yz projection $-$ mean $v_{\rm x}$. {\bf Middle row}: gas velocity dispersion maps in the edge-on xz projection $-$ $\sigma_{\rm y}$. Left panel: {\it all gas}. Right panel: {\it warm gas}, that is only pixels where $3.8\le\log\big(\langle$T$_{\rm gas}\rangle/$K$\big)\le4.2$ (see Section \ref{binn}). Warm gas is clearly associated with the galactic disc and virtually absent when moving away from the galaxy plane, where the all gas $\sigma_{\rm y}$-map peaks. {\bf Bottom row}: gas velocity dispersion maps in the face-on xy projection $-$ $\sigma_{\rm z}$. Left panel: {\it all gas}. Right panel: {\it warm gas}. Now, all the lines-of-sight pierce the disc (where warm gas dominates) and the two maps are almost identical.} \label{fig_maps} \end{figure*} We then rotated the particle distribution ($g$+$s$) around the gravitational potential minimum (using the angular momentum per unit mass of star particles), in order to place each galaxy face-on in the xy plane and edge-on in the xz and yz planes. Accordingly, we rotated the velocity vector of each particle in the cube and used the velocity of the (rotated) stellar centre of mass as the velocity reference frame. At this point, we visually inspected the sample and only selected unperturbed galaxies with a prominent disc structure in both gas and stars, a number of gas particles (N$_{\rm gas}$) inside the $30$+$70$ kpc cube sufficient to make our analysis robust, and fairly regular$/$symmetric gas rotation velocity maps (constructed as explained in Section \ref{binn}). The final sample includes $43$ galaxies with $9.02\le\log($M$_{\star}/$M$_\odot)\le10.76$ and $7.8\times10^{3}\lesssim$ N$_{\rm gas}\lesssim9.4\times10^{4}$ ($\langle$N$_{\rm gas}\rangle\approx3.8\times10^{4}$)\footnote{Note that a galaxy with M$_{\star}=10^{9}$ M$_\odot$ in {\textit {Recal-L025N0752}} contains more than $4.4\times10^{3}$ star particles.}. In the top row of Fig. \ref{fig_gal11_proj} we plot positions of gas (black dots) and star (orange dots) particles for the galaxy with the highest SFR. The left and right panels show the edge-on xz and yz projections: the galactic disc is clearly visible in both star and gas components. The middle panel shows the face-on xy projection. The gas distribution is arranged in a complex clumpy structure as a result of the interplay between star formation and associated feedback processes. The bottom left panel of Fig. \ref{fig_gal11_proj} shows the distribution in the SFR$-$M$_{\star}$ plane of our final sample of $43$ galaxies. The minimum and maximum star formation rates are $0.065$ and $6.77$ M$_\odot$ yr$^{-1}$. The grey solid (+ triple dot-dashed) line is the best fit SFR$-$M$_{\star}$ relation for star forming local galaxies of \citet{renzinipeng2015}: $\log\big($SFR$/[$M$_\odot$ yr$^{-1}]\big)=(0.76\pm0.01)\log($M$_{\star}/$M$_\odot)- 7.64\pm0.02$. The final sample represents disc galaxies on the main sequence. We plot the simulated specific SFR$-$M$_{\star}$ relation in the bottom right panel of Fig. \ref{fig_gal11_proj}. The vertical and horizontal dashed lines mark, respectively, the median stellar mass, $\log(\tilde{{\rm M}}_{\star}/$M$_\odot)=9.78$, and median specific star formation rate, $\log(\tilde{\rm sSFR}/$yr$^{-1})=-10.04$. In the following sections we will study how the gas velocity dispersion distribution varies as a function of M$_{\star}$ and sSFR. For this reason, we split the plot in four quadrants: Q$1$ = low M$_{\star}-$ low sSFR, \color{red} Q$2$ \color{black}= low M$_{\star}-$ high sSFR, \color{blue} Q$3$ \color{black}= high M$_{\star}-$ low sSFR and \color{orange} Q$4$ \color{black}= high M$_{\star}-$ high sSFR. Q$1$ and Q$4$ contain, respectively, $7$ and $8$ galaxies, while both Q$2$ and Q$3$ contain $14$ objects. \subsection{Binning and warm gas} \label{binn} We binned the gas particles in each galactic cube on a $2$D spatial (+ $1$D depth) grid of pixels with linear size $2$ kpc ($15$ pixels per cubic side), which is comparable to the effective resolution of SAMI after accounting for a typical AAT seeing of $2.1$ arcsec\footnote{In Appendix \ref{appendix_b} we will explore the impact of cube size ($30$ and $60$ kpc) and grid resolution ($2$ and $3$ kpc) on our results.}. To obtain SPH quantities on the grid, we followed the procedure described in Section \color{blue} 4 \color{black} of \citet{altay2013}. We started by extracting a buffer of additional $70$ kpc per side around the central cube of volume $(30$ kpc$)^3$, to ensure that all the gas particles in the simulation whose $3$D SPH kernels intercept one or more pixels of the grid in the $2$ spatial directions and the cube margins in the line-of-sight direction were taken into account. Then, we assigned a truncated Gaussian kernel to each gas particle (Eqs. \color{blue} 8 \color{black} and \color{blue} 9 \color{black} of \citealt{altay2013}) and integrated it over the square pixels. Using this procedure, we calculated the density weighted (mean) velocity, velocity dispersion and temperature along the line-of-sight in all the pixels. We considered different projections. In the edge-on xz (yz) projection, the line-of-sight direction is the direction y (x) perpendicular to the xz (yz) plane. The corresponding mean velocity and velocity dispersion are, respectively, $v_{\rm y}$ and $\sigma_{\rm y}$ for the xz projection and $v_{\rm x}$ and $\sigma_{\rm x}$ for the yz projection. The velocity dispersion for the face-on projection xy is $\sigma_{\rm z}$. The two top panels of Fig. \ref{fig_maps} show examples of mean (rotation) velocity maps in the two edge-on projections created using all the gas in the cube. SAMI observations of low-redshift galaxies with outflows are largely based on the detection of H$\alpha$ emitting gas at T $\sim 10^4$ K. From now on, to better compare with these observations we will distinguish between {\it all gas} and {\it warm gas}. In an SPH simulation, the fluid conditions at any point are defined by integrating over {\it all} particles, weighted by their kernel. Selecting only a subset of them (e.g. only star forming or cold$/$hot gas) would break mass$/$momentum$/$energy conservation laws. Therefore, we define \\ \noindent{\it Warm gas}: pixels in a velocity$/$velocity dispersion$/$temperature map where the density weighted gas temperature (calculated using all the particles) is in the range $3.8\le\log\big(\langle$T$_{\rm gas}\rangle/$K$\big)\le4.2$. \\ \noindent Gas with temperature around $10^4$ K is usually referred to as warm to distinguish it from cold gas in molecular clouds (T $<100$ K) and hot gas in the halo or in supernova bubbles (T $>10^5$ K), based on the model of a three-phase ISM medium \citep{mckeeost1977}. H$\alpha$ emission in real galaxies is mostly from \text{H\,\textsc{\lowercase{II}}} regions, and hence correlates strongly with star formation rate \citep{kennicutt1998}. To a good approximation, in our analysis warm pixels trace pixels with density weighted SFR greater than zero. We introduced this temperature cut to qualitatively compare with the kinematic signatures seen in SAMI observations, without having to model complicated and uncertain radiative transfer effects. In the middle row of Fig. \ref{fig_maps} we plot edge-on velocity dispersion maps (xz $-$ $\sigma_{\rm y}$) for $a$) all gas (left panel) and $b$) warm gas (right panel). In $a$) the velocity dispersion increases when moving away from the disc of the galaxy (in both vertical directions) and peaks at abs(z$_{\rm gas}$) $\sim9$ kpc. On the other hand, in $b$) the high-$\sigma$ part is completely suppressed and warm gas is mainly associated with the galactic disc\footnote{We stress again that the map in $b$) is the map in $a$) with only pixels fulfilling the condition $3.8\le\log\big(\langle$T$_{\rm gas}\rangle/$K$\big)\le4.2$ included.}. This has important consequences for our analysis. We will explore them in Sections \ref{total_vs_ha} and \ref{veldistr3}. The situation is different in the bottom two panels of Fig. \ref{fig_maps}, which show face-on velocity dispersion maps (xy $-$ $\sigma_{\rm z}$) for all gas (left panel) and warm gas (right panel). This time the two maps are almost identical (except for two pixels). This is due to the fact that now all the lines-of-sight pierce the disc, where warm gas dominates. The $\sigma$-maps in Fig. \ref{fig_maps} are the base of all our analyses and we will discuss them more in the next sections. \section{Signatures of outflows: the velocity dispersion distribution} \label{gkin} \begin{figure} \centering \includegraphics[width=8.7cm]{multi_hist_face_edge_gas_Ha} \caption{Pixelated velocity dispersion probability distributions (i.e. the fraction of pixels $-$ with pixel size $= 2$ kpc $-$ per velocity dispersion bin) calculated using all gas and warm gas in different projections. The bin size is $20$ km s$^{-1}$. Black diamonds + solid line and red triangles + short dashed line: all gas and warm gas, respectively, in the edge-on xz projection $-$ $\sigma_{\rm y}$. Blue squares + dot-dashed line and orange crosses + long dashed line: all gas and warm gas, respectively, in the face-on xy projection $-$ $\sigma_{\rm z}$. Each line was created by stacking the histograms of all the $43$ galaxies in the simulated sample. Errors are Poissonian. Since in the face-on projection the lines-of-sight in different pixels always pierce the galactic disc, where warm gas dominate, the all gas and warm gas $\sigma_{\rm z}$-distributions are very similar (see the bottom panels of Fig. \ref{fig_maps}). In the edge-on projection, the $\sigma_{\rm y}$-distribution for all gas is shifted to higher velocity dispersions than the warm gas distribution, which is associated with the galactic disc and therefore has a more prominent peak at low $\sigma_{\rm y}=30$ km s$^{-1}$ (cf. the two middle panels of Fig. \ref{fig_maps}).} \label{fig_total_vs_ha} \end{figure} The aim of this work is to determine whether or not current and upcoming IFS surveys can succesfully identify large-scale outflows and provide meaningful constraints on the physical processes driving gas out of galaxies. For this reason, we apply observationally-based analysis techniques to the simulations. In particular, in the next sections we compare with the observational investigations of SAMI galaxies with outflows presented by \citet{ho2014} and \citet{ho2016}. We stress that observational and numerical analyses estimate the kinematic state of the gas in two different ways. We track the motion of particles as sampled by the SPH scheme in the simulation, without including any radiative transfer effects. Instead, \citet{ho2014,ho2016} extract kinematic information from emission line spectra of galaxies. Thus, while observations only probe ionised gas, simulations take into account all the gas in the galactic halo. To facilitate the comparison, we therefore introduced in the previous section the definition of warm gas that we will use throughout the paper. \citet{ho2014} studied the nature of a prototypical low redshift isolated disc galaxy with outflows: SDSS J$090005.05$+$000446.7$ (SDSS J$0900$, $z = 0.05386$). Its emission line spectrum was decomposed using the spectral fitting pipeline {\small{LZIFU}} \citep{LZIFU2016}, a likelihood ratio test and visual inspection. Emission lines were modelled as Gaussians composed of up to three kinematic components with small, intermediate and high velocity dispersion relative to each other (i.e from narrow to broad features). Fig. \color{blue} 6 \color{black} of \citet{ho2014} shows the velocity dispersion distribution of SDSS J$0900$. The statistically prominent narrow component peaks at $\sim40$ km s$^{-1}$ and is associated with the (rotationally supported) disc of the galaxy. More interestingly, the $\sigma$-distribution extends to very high values ($450$ km s$^{-1}$) with the broad kinematic component peaking at $\sim300$ km s$^{-1}$. The authors argue that this high-$\sigma$ component traces shock excited emission in biconical outflows, likely driven by starburst activity. \subsection{All gas vs warm gas} \label{total_vs_ha} Following \citet{ho2014}, we begin our analysis by investigating the velocity dispersion distribution of the simulated galaxies. Throughout the paper, we adopt the following procedure. Using $\sigma$-maps like those shown in the bottom four panels of Fig. \ref{fig_maps}, we first calculate the histogram of the pixelated velocity dispersion for each galaxy (i.e. the fraction of pixels per velocity dispersion bin), where the pixel size is $2$ kpc, the upper limit is the maximum $\sigma$ of the entire sample (that changes depending on the projection and if all gas or warm gas is considered) and the bin size is always $20$ km s$^{-1}$. Then, we stack these histograms and normalize by the number of galaxies to obtain the final {\it pixelated velocity dispersion (probability) distribution} and its associated Poissonian errors. In Fig. \ref{fig_total_vs_ha} we examine the differences between the pixelated velocity dispersions calculated using all gas and warm gas in different projections for all the $43$ simulated galaxies. In the face-on (xy view) projection, the $\sigma_{\rm z}$-distributions of all gas (blue squares and dot-dashed line) and warm gas (orange crosses and long dashed line) are very similar, with the highest fractions at low $\sigma_{\rm z}$ and a declining trend at larger velocity dispersions, and share the same $\max(\sigma_{\rm z})=201.31$ km s$^{-1}$. The only difference is that the all gas distribution shows slightly larger statistics at high $\sigma_{\rm z}$. According to the bottom panels of Fig. \ref{fig_maps}, this is not surpising. When the lines-of-sight in different pixels pass through the galactic disc, particles in the warm gas regime are the majority and dominate the $\sigma_{\rm z}$-maps. On the other hand, the edge-on (xz view) $\sigma_{\rm y}$-distribution for all gas (black diamonds and solid line, $\sigma_{\rm y,\max}=211.08$ km s$^{-1}$) is shifted to higher velocity dispersions than the warm gas distribution (red triangles and short dashed line, $\sigma_{\rm y,\max}=169.29$ km s$^{-1}$), which has a more prominent peak at $30$ km s$^{-1}$ and then rapidly drops to lower fractions at high $\sigma_{\rm y}$. The reasons for this are mentioned in Section \ref{sform_feedb} and highlighted by the two middle panels of Fig. \ref{fig_maps}. In our simulated disc galaxies, warm gas traces the disc and is virtually absent when moving away from the galaxy plane. In Section \ref{gal_winds}, we will show how the extraplanar velocity dispersion is dominated by outflowing gas. Therefore, the lack of warm gas outside the disc is a direct consequence of the thermal stellar feedback implemented in EAGLE that heats outflowing particles up to a temperature higher than the warm range (T $\sim10^{4}$ K). Galactic winds in our simulations are mostly hot (T $>10^5$ K) and, compared to observed galaxies, may entrain insufficient gas with T $<10^5$ K. Note that such high-temperature gas would not be visible in SAMI observations. This issue with \mbox{EAGLE} was already noted by \citet{turner2016} in a study of the $z\approx3.5$ intergalactic medium and has important implications for our work too. A direct comparison with H$\alpha$-based SAMI observations should be done using warm gas (since H$\alpha$ emitting gas has a temperature T $\sim10^4$ K). However, the paucity of such gas in the outflows of our simulated galaxies could lead to misleading results. Specifically, in edge-on projections: underestimation of the outflowing gas mass$/$incidence of galactic winds (that may be there but just too hot to be detected in H$\alpha$) and poor sampling of the extraplanar gas. In the next section we will show an example of this problem. % We also note that \citet{tanner2016b}, using synthetic absorption % lines generated from hydrodynamic simulations, found that the % outflow velocity of neutral gas, $v_{\rm o}$, is four to five times % lower than the average velocity of the hottest gas, the difference % increasing with gas ionisation. As a consequence, absorption lines % of neutral or low ionised gas could severely underestimate hot gas % $v_{\rm o}$ and outflow energetics. \subsection{Disc component and off-plane gas} \label{veldistr3} \begin{figure*} \centering \includegraphics[width=8.7cm]{multi_hist_disk_out} \includegraphics[width=8.7cm]{multi_hist_disk_out_Ha} \caption{Pixelated velocity dispersion probability distributions (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) calculated by assuming different cuts in the z direction of the edge-on (xz view) $\sigma_{\rm y}$-maps (with pixel size $=2$ kpc) to separate the galactic disc component from the extraplanar one. {\bf Left panel}: all gas. {\bf Right panel}: warm gas. Black filled triangles and solid line, (internal) pixels with abs(z) $<$ $3$ kpc. Red filled inverted triangles and dashed line, (external) pixels with abs(z) $>$ $3$ kpc. Orange diamonds and triple dot-dashed line, (external) pixels with abs(z) $>$ $5$ kpc. Errors are Poissonian. In the all gas case, excluding the galactic disc shifts the $\sigma_{\rm y}$-distribution to larger velocity dispersions. In the warm gas case, the off-plane distributions are poorly sampled and only marginally different from the disc one due to the scarcity of extraplanar gas at T $\sim10^4$ K.} \label{fig_hist_disc} \end{figure*} We focus on the difference between gas in the disc and extraplanar gas. To do so, we take the edge-on $\sigma_{\rm y}$-maps (xz view) and restrict the number of pixels in the direction perpendicular to the disc plane by applying various vertical cuts. The resulting pixelated velocity dispersion distributions (including Poissonian errors) are shown in Fig. \ref{fig_hist_disc}. The two panels refer to all gas (left) and warm gas (right). In the left panel, when all gas and only pixels with abs(z) $<$ $3$ kpc (black filled triangles and solid line) are included\footnote{ Here and throughout the paper, the vertical cuts are considered from the edge of the pixels, not the centre.}, the distribution shows a sizeable peak at $\sigma_{\rm y}=30$ km s$^{-1}$ and then quickly declines to low fractions. With this pixel selection, we are targeting only a thin layer of gas in the edge-on galactic discs. Red filled inverted triangles and the dashed line represent the complementary distribution calculated using only pixels with abs(z) $>$ $3$ kpc. In this case, the distribution is shifted to larger $\sigma$-values than before, while the low-$\sigma$ part is greatly reduced. This demonstrates that the low-$\sigma$ peak is indeed associated mainly with the galactic disc, in qualitative agreement with \citet{ho2014}. In the left panel of Fig. \ref{fig_hist_disc} we also show the velocity dispersion distribution (for all gas) calculated using only pixels with abs(z) $>$ $5$ kpc (orange diamonds and triple dot-dashed lines). The low-$\sigma$ peak and high-$\sigma$ tail become, respectively, slightly less and more important when moving further away from the galactic disc, supporting our previous conclusion. The right panel of Fig. \ref{fig_hist_disc} illustrates how using only warm gas can be misleading, in the framework of EAGLE simulations. All three distributions are very similar (despite the fact that they probe rather different environments) and more noisy at $\sigma_{\rm y}>50$ km s$^{-1}$ than the corresponding distributions for all gas. In the edge-on warm $\sigma_{\rm y}$-maps there are only a few pixels with abs(z) $>$ $3$ and (especially) $5$ kpc, therefore poor sampling affects the results in these two cases. Considering only warm gas in EAGLE would lead to the wrong conclusion that planar and extraplanar gas components are kinematically similar. This is a consequence of the thermal implementation of stellar feedback that produces hot outflows. When all gas (warm \& hot) is considered, the extraplanar (mainly hot) gas is kinematically clearly distinct from the (mainly warm) disc (left panel of Fig.\ref{fig_hist_disc}). In the rest of the paper, whenever possible and appropriate (e.g. to study face-on velocity dispersion distributions or the impact of general galactic properties like M$_{\star}$ and sSFR) we will show results obtained using warm gas. However, including all gas will be necessary to ensure a robust description of galactic winds and outflow signatures (see e.g. Section \ref{asym}). \subsection{Outflows} \label{gal_winds} In the previous section we have demonstrated how the low-$\sigma$ part of the edge-on velocity dispersion distribution is associated with the galactic disc. Now, we study the origin of the high-$\sigma$ tail. In the xz edge-on projection, we consider a pixel of the $\sigma$-map as {\it outflow dominated} if the density weighted vertical velocity of its gas particles, $\langle v_{\rm z}\rangle$, is positive in the semi-plane with z$_{\rm gas}>0$ kpc or negative in the negative z$_{\rm gas}$ semi-plane (i.e. if the gas particles contributing to the pixel are predominantly moving away from the galactic plane). Otherwise, a pixel is flagged as {\it non-outflow dominated}\footnote{We mask pixels in the galactic plane (i.e. those with $-1 < $ z$_{\rm gas}/$kpc $<1$), since their outflowing/non-outflowing status is undefined.}. The corresponding $\sigma_{\rm y}$-distributions are shown in Fig. \ref{fig_v_esc}. Errors are Poissonian. The non-outflow dominated $\sigma$-distribution (black diamonds and solid line) resembles the abs(z) $<$ $3$ kpc distribution (i.e. associated with the galactic disc) visible in the left panel of Fig. \ref{fig_hist_disc} (black filled triangles and solid line). With respect to these two diagrams, the outflow dominated $\sigma$-distribution of Fig. \ref{fig_v_esc} (red triangles and dashed line) is shifted to higher velocity dispersions and has a more statistically prominent high-$\sigma$ tail (in agreement with the abs(z) $>$ $3$ and $5$ kpc distributions of Fig. \ref{fig_hist_disc}). When only extraplanar gas is considered (i.e. pixels with abs(z) $>$ $3$ kpc), we find that in $42$ out of $43$ simulated galaxies, the number of outflow dominated pixels is greater than the number of non-outflow dominated pixels. On average, this excess is a factor of $\sim3.2$ times and can be up to $>15$ times. \begin{figure} \centering \includegraphics[width=8.7cm]{multi_hist_outflows_rr} \caption{Pixelated velocity dispersion probability distributions (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) for outflow dominated (red triangles and dashed line) and non-outflow dominated (black diamonds and solid line) pixels. Errors are Poissonian and we consider all gas in the edge-on xz projection $-$ $\sigma_{\rm y}$. The velocity dispersion distribution based on outflow dominated pixels is more prominent at high-$\sigma$.} \label{fig_v_esc} \end{figure} \begin{figure*} \centering \includegraphics[width=8.7cm]{multi_hist_select_poiss_bin2_tot_Ha} \includegraphics[width=8.7cm]{multi_hist_select_poiss_bin2_tot_xy_Ha} \caption{Pixelated velocity dispersion probability distributions (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) for galaxies in the four M$_{\star}-$sSFR quadrants defined in the bottom right panel of Fig. \ref{fig_gal11_proj}: Q$1$ = low M$_{\star}-$ low sSFR, \color{red} Q$2$ \color{black}= low M$_{\star}-$ high sSFR, \color{blue} Q$3$ \color{black}= high M$_{\star}-$ low sSFR and \color{orange} Q$4$ \color{black}= high M$_{\star}-$ high sSFR. {\bf Left panel}: edge-on xz projection $-$ $\sigma_{\rm y}$. {\bf Right panel}: face-on xy projection $-$ $\sigma_{\rm z}$. Errors are Poissonian and we consider only warm gas. In general, galaxies with higher M$_{\star}$ present a more extended $\sigma$-distribution, while the sSFR has a secondary effect with respect to stellar mass.} \label{fig_hist_tot} \end{figure*} \begin{figure*} \centering \includegraphics[width=8.4cm]{ms-sigma_sfr} \includegraphics[width=8.4cm]{multi_hist_sfr_dens_mstar_Ha} \includegraphics[width=8.4cm]{multi_hist_sfr_dens_mstar} \caption{{\bf Top panel}: relation between star formation rate surface density, $\Sigma_{\rm SFR}$, and M$_{\star}$ for our simulated galaxies. $\Sigma_{\rm SFR}={\rm SFR}/(2\pi r_{50}^2)$, where SFR is the total star formation rate and $r_{50}$ is the radius within which half of the galaxy stellar mass is included. The plot is divided in four sectors: S$1$ = low M$_{\star}-$ low $\Sigma_{\rm SFR}$, \color{red} S$2$ \color{black}= low M$_{\star}-$ high $\Sigma_{\rm SFR}$, \color{blue} S$3$ \color{black}= high M$_{\star}-$ low $\Sigma_{\rm SFR}$ and \color{orange} S$4$ \color{black}= high M$_{\star}-$ high $\Sigma_{\rm SFR}$ (see Section \ref{sigma_SFR}). {\bf Bottom panels}: pixelated velocity dispersion probability distributions (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) for galaxies in the four sectors defined in the top panel. Errors are Poissonian. We only consider the edge-on xz projection $-$ $\sigma_{\rm y}$. Left panel: warm gas. As in the previous section, a trend with stellar mass is visible. Right panel: all gas. At fixed M$_{\star}$, galaxies with high $\Sigma_{\rm SFR}$ give rise to a $\sigma$-distribution shifted to larger velocity dispersions compared to galaxies with low $\Sigma_{\rm SFR}$.} \label{fig_hist_sfrd} \end{figure*} Our results are qualitatively consistent with the observations of \citet{ho2014}: in low redshift disc galaxies, the low-$\sigma$ component of the velocity dispersion distribution is associated with the (rotationally supported) disc, while the high-$\sigma$ component mainly traces extraplanar, outflowing gas. In the case of strong disc-halo interactions through galactic winds, the velocity dispersion of the extraplanar gas is broadened (up to $300$ km s$^{-1}$ in the extreme case of M82) by both the turbulent motion of the outflowing gas and line splitting caused by emissions from the approaching and receding sides of the outflow cones \citep[][and references therein]{ho2016}. There is an important caveat to consider here. In this work, outflowing material includes both particles that are actually leaving their host galaxy and particles that will eventually stop and fall back to the disc. At this stage, our numerical analysis is not able to differentiate between these two components (but this also applies to observations). Theoretical predictions on how much gas actually escapes from galaxies are crucial, since the escaping mass is very hard to measure observationally \citep{joss2003, joss2007}. Recent simulations run by different groups indicate that wind recycling becomes particularly important at $z<1$ and galaxies of all masses reaccrete more than $50$\% of the expelled gas \citep[e.g.][]{oppe2010, nelson2015, christensen2016, alcazar2016}. This point will be addressed in an upcoming paper (Crain et al. in prep). Note that the thermal$/$buoyant winds in our EAGLE discs will allow particles without enough velocity$/$thermal energy to escape to float up to the top of the galactic halo \citep{bower2016}. \section{Impact of stellar mass, specific SFR, SFR surface density and gas temperature} \label{veldistr1} In this section, we study the impact of different galactic properties on the overall shape of the velocity dispersion distribution. Since we do not focus primarily on the high-$\sigma$ tail associated with outflows, results are presented for warm gas to better compare with the observational analysis of \citet{ho2014, ho2016}. We begin with stellar mass and the specific star formation rate. Fig. \ref{fig_hist_tot} shows the result: the edge-on xz projection $-$ $\sigma_{\rm y}$ in the left panel and the face-on xy projection $-$ $\sigma_{\rm z}$ in the right panel (errors are Poissonian). We divided our galaxies according to the four M$_{\star}-$sSFR quadrants in the bottom right panel of Fig. \ref{fig_gal11_proj} (the same colour code applies). A clear trend with stellar mass is visible in both panels, while the sSFR appears to have a secondary effect. It is interesting to note how, especially in the edge-on projection (left panel), the velocity dispersion distribution of low mass galaxies (black diamonds + solid line and red triangles + dashed line, respectively associated with Q$1$ and Q$2$) declines shortly after the peak at $30$ km s$^{-1}$ (associated with the disc) and drops to zero already at $70$ km s$^{-1}$. The predominance of the disc component in Q$1$ and Q$2$ is present also when all gas distributions (not shown here) are used. The $\sigma$-distributions of galaxies with high-M$_{\star}$ (blue squares + dot-dashed line and orange crosses + triple dot-dashed line, respectively associated with Q$3$ and Q$4$) are more extended, and prominent at high-$\sigma$, than those of low mass galaxies (regardless of the range in sSFR). At $\sigma<100$ km s$^{-1}$, this is due to the fact that in the synthetic sample warm gas mainly traces the galactic disc and objects in Q$3$ and Q$4$ are generally bigger. Due to the SFR$-$M$_{\star}$ relation and the fact that in EAGLE there is a direct connection between SFR and stellar feedback, these objects also have higher outflowing activities than galaxies in Q$1$ and Q$2$. Despite the lack of warm gas in EAGLE's galactic winds, this causes the broadening of the $\sigma$-distributions to higher velocity dispersion. % Both at low- and high-M$_{\star}$, the difference between the mean % sSFR in the low- and high-sSFR subsamples is $\sim0.33$ dex, while % the difference in the mean SFR (stellar mass) between the low- and % high-M$_{\star}$ subsamples is $0.74$ ($0.82$) dex, more than double % (while the mean sSFR decreases by only $0.13$ dex). This explains % why stellar mass has a larger impact than the specific star % formation rate. Our simulated face-on distributions, with $\max(\sigma_{\rm z})=201.31$ km s$^{-1}$, do not extend as far as the velocity dispersion diagram of the SDSS J$0900$ galaxy in \citet{ho2014}, with $\max(\sigma_{\rm z})\sim 450$ km s$^{-1}$. This might be in part due to the fact that SDSS J$0900$ is more massive, $\log($M$_{\star}/$M$_\odot)=10.8$, and has a higher SFR ($\sim5-15$ M$_\odot$ yr$^{-1}$, depending on the adopted SFR indicator) than objects in our synthetic sample, but could also indicate that the effect of EAGLE stellar feedback on gas kinematics is too weak. Simulated galaxies in Q$2$, Q$3$ and Q$4$ have distributions that reach $\sigma_{\rm z}>100$ km s$^{-1}$\footnote{In the edge-on projection, only galaxies with high-M$_{\star}$ (Q$3$ and Q$4$) have distributions with $\max(\sigma_{\rm y})>100$ km s$^{-1}$.}, which is the starting point of the broad kinematic component associated with outflowing gas in SDSS J$0900$. \subsection{SFR surface density} \label{sigma_SFR} \begin{figure*} \centering \includegraphics[width=8.7cm]{hist_temp_gal} \includegraphics[width=8.7cm]{multi_hist_temp} \caption{{\bf Left panel}: average temperature distribution of gas particles inside cubes of linear size $30$ kpc centred on each simulated galaxy. Two distinct regions are visible in the gas temperature histogram: the bulk of particles with T $<10^5$ K, and a tail of particles with T $\ge 10^5$ K. {\bf Right panel}: effect of gas temperature on the pixelated velocity dispersion probability distribution (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) of galaxies. Black diamonds and solid line: pixels where the density weighted gas temperature is T $<10^5$ K. Red triangles and dashed line: pixels with T $\ge10^5$ K. Errors are Poissonian and we only consider the edge-on xz projection $-$ $\sigma_{\rm y}$. Pixels with temperature T $<10^5$ K trace the galactic disc (low-$\sigma$), while those with T $\ge10^5$ K are associated with higher-$\sigma$ (i.e. extraplanar, outflowing gas).} \label{fig_hist_temp_cut} \end{figure*} \citet{ho2016} found that, on average, wind galaxies have higher star formation rate surface densities than those without strong wind signatures. We checked this result with our simulated sample in Fig. \ref{fig_hist_sfrd}. Following \citet{ho2016}, the star formation rate surface density is defined as $\Sigma_{\rm SFR}=$ SFR$/(2\pi r_{50}^2)$, where SFR is the total star formation rate of the object as determined by {\small SUBFIND} and $r_{50}$ is the radius within which half of the galaxy stellar mass is included. The top panel of Fig. \ref{fig_hist_sfrd} shows the $\Sigma_{\rm SFR}-$M$_{\star}$ relation of EAGLE galaxies. As for SFR and M$_{\star}$, the two quantities are positively correlated. The vertical and horizontal dashed lines mark, respectively, the median stellar mass, $\log(\tilde{{\rm M}}_{\star}/$M$_\odot)=9.78$, and median star formation rate surface density, $\log\big(\tilde{\rm \Sigma}_{\rm SFR}/[$M$_\odot$ yr$^{-1}$ kpc$^{-2}]\big)=-2.77$. We divided the plot in four sectors: S$1$ = low M$_{\star}-$ low $\Sigma_{\rm SFR}$ (16 objects), \color{red} S$2$ \color{black}= low M$_{\star}-$ high $\Sigma_{\rm SFR}$ (5 objects), \color{blue} S$3$ \color{black}= high M$_{\star}-$ low $\Sigma_{\rm SFR}$ (6 objects) and \color{orange} S$4$ \color{black}= high M$_{\star}-$ high $\Sigma_{\rm SFR}$ (16 objects). The corresponding velocity dispersion distributions are shown in the bottom panels of Fig. \ref{fig_hist_sfrd}: warm gas on the left and all gas on the right (we only consider the edge-on xz projection $-$ $\sigma_{\rm y}$). We start by considering the warm gas case (left panel of Fig. \ref{fig_hist_sfrd}). Trends are similar to those of the left panel of Fig. \ref{fig_hist_tot}. At low masses (S$1$ and S$2$), the velocity dispersion distributions of galaxies with low and high $\Sigma_{\rm SFR}$ are almost identical (black diamonds + solid line and red triangles + dashed line), with a narrow peak at $30$ km s$^{-1}$. These galaxies have relatively low SFRs and weak outflowing activities, therefore warm gas mainly traces their galactic discs (of similar size). The probability distributions of high-mass galaxies (blue squares + dot-dashed line and orange crosses + triple dot-dashed line, respectively associated with S$3$ and S$4$) are shifted to larger values than those of low-mass galaxies. As discussed in the previous section, part of the shift is driven by the increase in stellar mass, but a correlation with the SFR surface density is now visible (a more extended high-$\sigma$ tail for objects in S$4$ with high $\Sigma_{\rm SFR}$). Patterns are different in the all gas case (right panel of Fig. \ref{fig_hist_sfrd}). A trend with stellar mass is still present (black \& red vs blue\& orange points and lines), but, at fixed M$_{\star}$, galaxies with high $\Sigma_{\rm SFR}$ give rise to a $\sigma$-distribution shifted to larger velocity dispersions compared to galaxies with low $\Sigma_{\rm SFR}$ (red \& orange vs black\& blue points and lines). As in the observations of\citet{ho2016}, this result, only partially visible before due to the lack of warm gas in the outflows of EAGLE galaxies, indicates that the star formation rate surface density correlates with the outflowing activity even when $\Sigma_{\rm SFR}$ is rather low, as it is the case of our simulated galaxies (we will explore the correlation in more detail in Section \ref{asym}). According to \citet{heckman2002}, starburst-driven winds are observed to be ubiquitous in galaxies with $\log\big(\Sigma_{\rm SFR}/[$M$_\odot$ yr$^{-1}$ kpc$^{-2}]\big) > -1$ \citep[see also the results of][]{sharma2017}. In our sample, $\log\big(\Sigma_{\rm SFR,\max}/[$M$_\odot$ yr$^{-1}$ kpc$^{-2}]\big) = -1.95$, almost a dex lower\footnote{In \citet{ho2016}, winds are seen at $-3\lesssim\log\big(\Sigma_{\rm SFR}/[$M$_\odot$ yr$^{-1}$ kpc$^{-2}]\big) \lesssim -1.5$.}. \subsection{The role of gas temperature} \label{veldistr_temp} Since the predictive power of realistic numerical simulations allows us to study the effect of additional galactic properties, which are usually hard to measure observationally, we looked for a way to better relate the shape of the velocity dispersion distribution to feedback processes. In EAGLE, stellar feedback is implemented thermally, and we found that temperature is a good proxy to distinguish low- and high-$\sigma$ parts in our simulated galaxies. The average temperature histogram of gas particles inside cubes of volume ($30$ kpc)$^3$ centred on each simulated galaxy splits into two regions: the bulk of particles with T $<10^5$ K, and a tail of particles with T $\ge 10^5$ K (see the left panel of Fig. \ref{fig_hist_temp_cut}). For this reason, we divided pixels in the velocity dispersion distribution where the density weighted gas temperature is above and below $10^5$ K. The result is visible in the right panel of Fig. \ref{fig_hist_temp_cut} (we consider all gas in the edge-on xz projection $-$ $\sigma_{\rm y}$, errors are Poissonian). The condition on temperature produces two very different probability distributions. When pixels with T $<10^5$ K are selected (black diamonds and solid line), the distribution peaks at $\sigma_{\rm y}=30$ km s$^{-1}$, then quickly drops to low fractions. On the other hand, pixels with T $\ge 10^5$ K give rise to a distribution shifted to larger velocity dispersions and where the low-$\sigma$ section is less prominent (red triangles and dashed line). These trends, which we find are also visible in the face-on xy projection $-$ $\sigma_{\rm z}$, support the results of the previous sections (see in particular the left panel of Fig. \ref{fig_hist_disc} and Fig. \ref{fig_v_esc}). Thus, EAGLE simulations of low redshift disc galaxies indicate a direct correlation between the thermal state of the gas and its state of motion as described by the velocity dispersion distribution: \begin{itemize} \renewcommand{\labelitemi}{$\bullet$} \item Low-$\sigma$ peak $\,\,\Leftrightarrow\,$ galactic disc$/$gas with T $<10^5$ K; \item High-$\sigma$ tail $\,\Leftrightarrow\,$ outflows$/$gas with T $\ge 10^5$ K. \end{itemize} The real picture is certainly more complicated than this. For example, blobs of cold gas at relatively high density could be entrained in hot, diffuse winds \citep{veilleux2005, cooper2008, cooper2009}. Despite the simplifications made in our analysis, the predicted correlation between EAGLE's thermal$/$buoyant outflows and high temperature gas can be very useful to guide and interpret real observations. \section{SIGNATURES OF WINDS: GAS KINEMATICS IN EAGLE AND SAMI GALAXIES} \label{asym} \citet{ho2016} proposed an empirical identification of wind-dominated SAMI galaxies. In this section, we apply the same methodology to our simulated sample. The authors defined two dimensionless quantities to measure the (ionised gas) extraplanar velocity dispersion and asymmetry of the velocity field. The first quantity is the velocity dispersion to rotation ratio parameter: \begin{equation} \eta_{50}=\sigma_{50}/v_{\rm rot}, \end{equation} where $\sigma_{50}$ is the median velocity dispersion of all pixels outside $\tilde{r}_{\rm e}$ (the $r$-band effective radius increased by approximately 1 arcsec to reduce the effect of beam smearing) with signal-to-noise in H$\alpha$ $-$ S$/$N(H$\alpha$) $-$ greater than $5$. $v_{\rm rot}$ is the maximum rotation velocity measured from the pixels along the optical major axis (for galaxies without sufficient spatial coverage, they used the stellar mass Tully-Fisher relation to infer $v_{\rm rot}$). The second quantity is the asymmetry parameter: \begin{equation} \label{eq_xi} \xi={\rm std}\,\bigg(\frac{v_{\rm gas}-v_{\rm gas,flipped}}{\sqrt{{\rm Err}(v_{\rm gas})^2+{\rm Err}(v_{\rm gas,flipped})^2}} \bigg), \end{equation} where ${\rm std}=$ standard deviation. To obtain $\xi$, the authors first flipped the line-of-sight velocity map over the galaxy major axis, $v_{\rm gas,flipped}$, and then subtracted the flipped map from the original one, $v_{\rm gas}$. ${\rm Err}(v_{\rm gas})$ and ${\rm Err}(v_{\rm gas,flipped})$ are the corresponding $1\sigma$ error maps from {\small{LZIFU}}. The standard deviation is again calculated taking into account only pixels outside $\tilde{r}_{\rm e}$ with S$/$N(H$\alpha$) $>5$. \citet{ho2016} used $\eta_{50}$ and $\xi$ to quantify the strength of disc-halo interactions and to distinguish galactic winds from extended diffuse ionised gas (eDIG) in a sample of $40$ low redshift disc galaxies. Since galactic winds both perturb the symmetry of the extraplanar gas velocity and increase the extraplanar emission line widths, they should show high $\xi$ and high $\eta_{50}$. On the other hand, eDIG is more closely tied to the velocity field of the galaxy and therefore should result in low $\xi$ and low $\eta_{50}$. \begin{figure*} \centering \vspace{0.6cm} \includegraphics[width=8.7cm]{pasymmetry_total_linear_sfrdens_fit_noise} \includegraphics[width=8.7cm]{plot_asim_sfrd_corr_SFR_Ha_log} \caption{{\bf Left panel}: asymmetry parameter $\xi$ vs velocity dispersion to rotation ratio parameter $\eta_{50}$. Dots: simulated edge-on xz projection. Squares: simulated edge-on yz projection. Synthetic data are colour coded according to the SFR surface density of the parent galaxy, $\Sigma_{\rm SFR}={\rm SFR}/(2\pi r_{50}^2)$. The vertical and horizontal dotted lines mark the observational limits above which galaxies show strong disc-halo interactions according to \citet{ho2016}, $\eta_{50}>0.3$ and $\xi>1.8$. Crosses: observational SAMI data, where red \color{red} \textsf{x} \color{black} symbols represent wind-dominated galaxies and black \textsf{x} symbols represent galaxies without strong wind signatures. {\bf Right panel}: correlation between the asymmetry parameter $\xi$ and $\Sigma_{\rm SFR}$. Grey dots and squares: simulated galaxies in the edge-on xz and yz projections, respectively. Crosses: observational data from \citet{ho2016}. In both panels, the range of the simulated parameters (determined using all gas) is broadly consistent with the observations. A clear positive $\xi-\Sigma_{\rm SFR}$ correlation is visible in both EAGLE and SAMI data.} \label{fig_asym} \end{figure*} We calculated $\eta_{50}$ and $\xi$ for our simulated galaxies and plot them (dots and squares) along with the observational data (black$/$red crosses) in Fig. \ref{fig_asym}. We remind the reader that there are some differences between the two analyses. The underlying assumptions are the same (i.e. regular, not warped discs and rotation maps, exclusion of mergers and systems undergoing major interactions) but, for example, ${\rm Err}(v_{\rm gas})$ and ${\rm Err}(v_{\rm gas,flipped})$ in Eq. \ref{eq_xi} are undefined in our analysis of EAGLE galaxies, since we rely on the SPH scheme to directly determine the kinematic state of the gas, without performing any emission line fitting. Instead, to obtain an estimate of the noise consistent with SAMI data, we fit a polynomial to \citet{ho2016}'s ${\rm Err_{\rm obs}}(v_{\rm gas})-$z [kpc] scatter plot and then add the observational noise to the simulated velocity maps. Note that the error on $v_{\rm gas}$ increases with the distance from the galaxy plane because the S$/$N of extraplanar H$\alpha$ emission is lower than that of H$\alpha$ gas in the disc. Furthermore, observations only consider ionised gas, while we take into account all (warm + hot) gas in the galactic cubelet to better sample any outflowing activity (see Sections \ref{total_vs_ha} and \ref{veldistr3}). Despite these differences, the range in $\eta_{50}$ and $\xi$ is similar in simulations and observations. \citet{ho2016} empirically defined as {\it wind-dominated} those galaxies with $\eta_{50}>0.3$ and $\xi>1.8$. These limits are visible in the left panel of Fig. \ref{fig_asym} as the vertical and horizontal black dotted lines. Accordingly, red \color{red} \textsf{x} \color{black} symbols in the figure represent wind-dominated SAMI galaxies ($15$ out of $40$), while black \textsf{x} symbols are associated with ($25$) observed objects without strong wind signatures. Among the EAGLE galaxies, just one object has $\eta_{50}<0.3$ (and only in the yz projection value). Although this would be consistent with a wind-dominated galaxy sample, the range in $\xi$ does not support this conclusion. The mean asymmetry parameter is $1.63$ (below the threshold of \citealt{ho2016}) and only five discs have $\xi>1.8$ in both projections. We are interested in studying the interdependency between $\eta_{50}$, $\xi$ and relevant galactic properties. As for the observational sample, the Spearman rank correlation test indicates no significant correlation between the two parameters: $\rho=0.12$ with a significance of $0.27$. \citet{ho2016} argue that if winds are the only mechanism disturbing the extraplanar gas, then a trend between $\eta_{50}$ and $\xi$ should be expected. The fact that both works fail to find a significant correlation suggests that, when applied to current data and simulations, the $\xi-\eta_{50}$ plot might not be accurate enough for the identification of wind-dominated galaxies. \citet{ho2016} speculate that gas accreted on to galaxies through satellite accretion would cause a large velocity asymmetry of the extraplanar gas without affecting much the off-plane velocity dispersion (i.e. large $\xi$ and small $\eta_{50}$), and therefore complicate the interpretation of the $\xi-\eta_{50}$ plot in terms of galactic winds and eDIG. We saw in Section \ref{gal_winds} how the extraplanar gas distribution of EAGLE objects is dominated by outflows (the ratio of outflow to non-outflow dominated pixels is on average $\sim3.2$). Unfortunately, this does not lead to a significant $\xi-\eta_{50}$ correlation in the simulated sample. However, Fig. \ref{fig_asym} highlights a different trend within EAGLE and SAMI data. In the left panel, simulations are colour coded according to the SFR surface density of the parent galaxy, $\Sigma_{\rm SFR}={\rm SFR}/(2\pi r_{50}^2)$. In general, objects with low$/$high $\xi$ have low$/$high $\Sigma_{\rm SFR}$ (i.e. bluish points are below reddish points). This positive correlation is even more visible in the right panel of Fig. \ref{fig_asym}, where we plot $\xi$ as a function of $\Sigma_{\rm SFR}$. The Spearman rank correlation test indicates a significant correlation between the two parameters both in simulations\footnote{For each simulated galaxy, there are two values of $\xi$, corresponding to the edge-on projections xz and yz. We averaged the two values into a single one to calculate the asymmetry$-\Sigma_{\rm SFR}$ correlation.} (grey dots and squares, $\rho=0.67$ with a significance of $\sim10^{-6}$) and observations\footnote{\citet{ho2016} quoted both spectral energy distribution (SED) and H$\alpha$ based SFRs for their sample. Here we use SFR$_{\rm H\alpha}$ to calculate the observed $\Sigma_{\rm SFR}$. We checked that our conclusions do not change when using SFR$_{\rm SED}$.} (black$/$red crosses, $\rho=0.56$ with a significance of $1.6\times10^{-4}$). Since the asymmetry parameter marks the incidence of galactic winds in disc galaxies, this result is in qualitative agreement with Section \ref{sigma_SFR} and the conclusions of \citet{ho2016}: the star formation rate surface density correlates with outflowing activity. \\ At this point, it is important to remark that numerical results could be affected by statistical noise in various ways. For example, in low mass, small galaxies high asymmetry could be due to poor sampling of the extraplanar gas distribution, rather than galactic winds. Moreover, since our discs are fairly regular, the two sets of $\eta_{50}$ and $\xi$ parameters associated with the edge-on xz and yz projections should contain the same information, and therefore be statistically similar. However, the analysis is limited by the small size of the synthetic sample ($43$ galaxies), which is due to the small box size of the cosmological simulation used ($25$ cMpc). In principle, increasing resolution and box size of the simulations will alleviate the impact of statistical noise. Note that the position of SAMI galaxies in the $\xi-\eta_{50}$ plot can also be biased in different ways. For example, see the discussion in \citet{ho2016} on how, for galaxies without a direct measurement of $v_{\rm rot}$, the lack of correlation between $\xi$ and $\eta_{50}$ is partially due to the scatter in the Tully-Fisher relation. Furthermore, the high asymmetry seen in some of the observed galaxies might be accentuated by inclination effects (while all the simulated galaxies are perfectly edge-on). In summary, according to our analysis, the $\xi-\eta_{50}$ plot does not seem to provide a clear, unambiguous tool to identify wind-dominated galaxies. Although using simulations with higher resolution (and better quality observational data) could improve the results, different observables (e.g. the velocity dispersion distribution) appear to estimate the outflowing activity of star forming galaxies more accurately than $\xi$ and $\eta_{50}$. \section{Conclusions} \label{concl} In this paper we have presented an analysis of stellar feedback-driven galactic outflows based on the comparison between hydrodynamic simulations and IFS observations from the SAMI survey \citep{bryant2015}. We have extracted cubes of $30$ physical kpc around unperturbed disc galaxies from the highest resolution cosmological simulation of the EAGLE set \citep{schaye2015}. Our final sample includes $43$ main sequence objects with $9.02\le\log($M$_{\star}/$M$_\odot)\le10.76$ and $\log(\tilde{\rm sSFR}/$yr$^{-1})=-10.04$ (Fig. \ref{fig_gal11_proj}). We have divided each cubelet into a grid of pixel size $2$ kpc (which is comparable to the effective resolution of SAMI) and created gas rotation velocity and velocity dispersion maps \mbox{(Fig. \ref{fig_maps})}. In our study the terms {\it galactic winds} and {\it outflows} are synonyms that we have used to identify both gas in the process of leaving a galaxy and gas that will eventually stop and fall back to the disc. This work is the theoretical counterpart of the observational analyses on SAMI galaxies with outflows presented in \citet{ho2014} and \citet{ho2016}. In the first part of the paper, we have focussed on the pixelated velocity dispersion (probability) distribution as a tracer of galactic wind signatures. To better compare with SAMI observations that mainly target H$\alpha$ emitting gas at T $\sim10^4$ K, we have distinguished between {\it all gas} and {\it warm gas}. The latter identifies pixels in a velocity$/$velocity dispersion$/$temperature map where the density weighted gas temperature (calculated using all the particles) is in the range $3.8\le\log\big(\langle$T$_{\rm gas}\rangle/$K$\big)\le4.2$. We have found that, in EAGLE galaxies, warm gas traces the disc and is virtually absent when moving away from the galaxy plane (middle two panels of Fig. \ref{fig_maps}). For this reason, the edge-on velocity dispersion for warm gas has a less prominent high-$\sigma$ tail (that is mostly associated with outflows) and a lower $\sigma_{\max}$ than the all gas distribution, while the two face-on distributions are very similar (Fig. \ref{fig_total_vs_ha}). The lack of warm gas outside the disc is a direct consequence of the thermal stellar feedback implemented in EAGLE. Galactic winds in our simulations are mostly hot (with a \mbox{T $>10^5$ K} they would not be seen in SAMI observations), buoyant \citep[rather than ballistic, cf.][]{bower2016} and, compared to observed galaxies, may entrain insufficient gas with T $<10^5$ K \citep[as pointed out also by][]{turner2016}. Throughout the paper, whenever possible and appropriate we have performed our analysis using warm gas. However, when studying outflowing material and galactic wind signatures we have included all gas to obtain more reliable results. We have targeted the galactic disc by taking into account only pixels with abs(z) $<$ $3$ kpc in the edge-on xz projection (with the disc lying in the xy plane). The $\sigma$-distribution peaks at $30$ km s$^{-1}$ and then quickly declines to low fractions. The complementary distributions (pixels with abs(z) $>$ $3$, $5$ kpc) extend to progressively larger values and have a less prominent low-$\sigma$ peak (Fig. \ref{fig_hist_disc}). This demonstrates that the low-$\sigma$ part of the velocity dispersion distribution is associated mainly with the galactic disc. On the other hand, (extraplanar) outflowing gas dominates the high-$\sigma$ tail (Fig. \ref{fig_v_esc}). Both these results are in qualitative agreement with the observations of \citet{ho2014}. In general, galaxies with higher stellar mass present a more extended $\sigma$-distribution (in both the edge-on and face-on projections), while the specific star formation rate has a secondary effect with respect to M$_{\star}$ (Fig. \ref{fig_hist_tot}). At fixed stellar mass and when all gas is used, the edge-on probability distribution of galaxies with high $\Sigma_{\rm SFR}=$ SFR$/(2\pi r_{50}^2)$ (where $r_{50}$ is the radius within which half of the galaxy stellar mass is included) is shifted towards larger velocity dispersions (and more extended) than the low-$\Sigma_{\rm SFR}$ one (Fig. \ref{fig_hist_sfrd}). As in the SAMI observations of \citet{ho2016}, this indicates that the star formation rate surface density correlates with the outflowing activity even at the low $\Sigma_{\rm SFR}$ seen in our simulated galaxies: $\log\big(\Sigma_{\rm SFR}/[$M$_\odot$ yr$^{-1}$ kpc$^{-2}]\big) \le -1.95$. We have studied the impact of temperature on the velocity dispersion distribution and found that there is a direct correlation between the thermal state of the gas and its state of motion. Gas with temperature T $<10^5$ K traces the low-$\sigma$ galactic disc, while gas with T $\ge10^5$ K is mainly associated with higher-$\sigma$ (i.e. extraplanar, outflowing gas, Fig. \ref{fig_hist_temp_cut}). Our results imply the following relations for low redshift disc galaxies in EAGLE: \renewcommand{\labelitemi}{$\bullet$} \begin{itemize} \item Low-$\sigma$ peak $\,\,\Leftrightarrow\,\,$ galactic disc $\,\,\Leftrightarrow\,\,$ gas with \mbox{T $<10^5$ K}; \item High-$\sigma$ tail $\,\,\Leftrightarrow\,\,$ galactic winds $\,\,\Leftrightarrow\,\,$ gas with \mbox{T $\ge 10^5$ K}. \end{itemize} We have applied the empirical identification of wind-dominated SAMI galaxies proposed by \citet{ho2016} to the simulated sample. The ranges of the two parameters used to measure the strength of disc-halo interactions, $\eta_{50}$ (velocity dispersion to rotation ratio) and $\xi$ (asymmetry of the extraplanar gas), are similar in simulations and observations, and most of EAGLE galaxies have $\eta_{50}>0.3$ (one of the thresholds introduced by \citealt{ho2016} to define when outflows become important, the other being $\xi>1.8$). Although this would be consistent with a wind-dominated synthetic sample, the range in $\xi$ does not support this conclusion: only five discs have $\xi>1.8$. Moreover, the lack of a significant correlation between $\eta_{50}$ and $\xi$ suggests that the $\xi-\eta_{50}$ plot is currently not accurate enough to provide a clear, unambiguous way to identify wind-dominated galaxies (left panel of Fig. \ref{fig_asym}). However, we have found a significant correlation between the asymmetry parameter $\xi$ and the SFR surface density of the parent galaxy (right panel of Fig. \ref{fig_asym}) that qualitatively confirms our result that $\Sigma_{\rm SFR}$ correlates with the outflowing activity. % Since $\xi$ marks the incidence of galactic winds in disc galaxies, % this result is in agreement with the analysis presented in the first % part of the paper and the conclusions of \citet{ho2016}: the star % formation rate surface density correlates with outflowing activity. In Appendix \ref{appendix_b} we have tested the impact of different cube sizes and grid resolutions. Increasing the pixel size from $2$ to $3$ kpc has a minimum impact on the $\sigma$-distribution, while changing the cube size from $30$ to $60$ kpc reduces the statistical relevance of the high-$\sigma$ tail (Fig. \ref{fig_hist_res_test}). Finally, in Appendix \ref{constrsim} we have tested our analysis on idealized simulations of isolated disc galaxies (Fig. \ref{fig_cdiscc_maps}). Results support the picture emerging from the analysis of EAGLE galaxies, where outflows are responsible for the high-$\sigma$ tail of the velocity dispersion distribution (Fig. \ref{fig_hist_discs}). \\ In summary, the velocity dispersion distribution has the potential to provide valuable information on the outflowing activity in galaxies. Notably, shape and extension of the high-$\sigma$ tail correlate with the strength of galactic winds. The comparison with SAMI observations has highlighted a limitation of EAGLE's stellar feedback: there is a dearth of cold and warm gas in the (mostly hot) simulated galactic outflows. Our results emphasise the double benefit of comparing simulations and observations: simulations are a valuable tool to interpret IFS data, and detailed observations guide the development of more realistic simulations. The next steps are to calculate how much gas actually escapes from galaxies and probe star formation activity and galactic winds in the time domain, following thermodynamic and kinematic changes as they happen. This is where the predictive power of simulations can play a pivotal role. As pointed out by \citet{guidi2016SDSS}, it will be important moving towards an unbiased, consistent comparison between simulated and real galaxies. The work presented in this paper extends the investigation of \citet{ho2014,ho2016} and contributes to placing SAMI observations of wind-dominated galaxies in a physical context. Together, these studies provide important constraints for ongoing IFS surveys and, in the longer term, will help with planning for the next generation multi-object integral field units including HECTOR, the successor of SAMI \citep{lawrence2012,joss2015,bryant2016}. \section*{Acknowledgements} ET would like to thank Lemmy Kilmister for constant inspiration during the writing of this paper and Jeremy Mould \& Paul Geil for many insightful discussions. RAC is a Royal Society University Research Fellow. SMC acknowledges the support of an Australian Research Council Future Fellowship (FT100100457). Support for AMM is provided by NASA through Hubble Fellowship grant\#HST-HF2-51377 awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS5-26555. The SAMI Galaxy Survey is based on observations made at the Anglo-Australian Telescope. The Sydney-AAO Multi-object Integral field spectrograph (SAMI) was developed jointly by the University of Sydney and the Australian Astronomical Observatory. The SAMI input catalogue is based on data taken from the Sloan Digital Sky Survey, the GAMA Survey and the VST ATLAS Survey. The SAMI Galaxy Survey is funded by the Australian Research Council Centre of Excellence for All-sky Astrophysics (CAASTRO), through project number CE110001020, and other participating institutions. The SAMI Galaxy Survey website is\url{http://sami-survey.org/}. Part of this research was done using the National Computational Infrastructure (NCI) {\it Raijin} distributed-memory cluster and supported by the Flagship Allocation Scheme of the NCI National Facility at the Australian National University (ANU). For the post-processing we used the {\it Edward} and {\it Spartan} High Performance Computing (HPC) clusters at the University of Melbourne. This work also used the DiRAC Data Centric system at Durham University, operated by the Institute for Computational Cosmology on behalf of the STFC DiRAC HPC Facility (\url{http://dirac.ac.uk}). The DiRAC system was funded by BIS National E-Infrastructure capital grant ST/K00042X/1, STFC capital grants ST/H008519/1 and ST/K00087X/1, STFC DiRAC Operations grant ST/K003267/1 and Durham University. DiRAC is part of the National E-Infrastructure. This work was supported by STFC grant ST/L00075X/1. \bibliographystyle{mnras} \bibliography{tescari_sami_eagle_winds_resub2} \appendix \section{variations of cube size and grid resolution} \label{appendix_b} Our galactic cubelets have a linear dimension of $30$ kpc and the pixel size of the velocity dispersion maps is $2$ kpc. In this appendix, we investigate the effect of different cube sizes and grid resolutions on our analysis. The result is shown in Fig. \ref{fig_hist_res_test}. Black circles and solid line represent the pixelated velocity dispersion distribution for the fiducial choice of parameters metioned above (we consider all gas in the edge-on xz projection $-$ $\sigma_{\rm y}$). \begin{figure} \centering \includegraphics[width=8.7cm]{multi_hist_select_poiss_res_test} \caption{Pixelated velocity dispersion probability distribution (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) of galaxies: resolution tests. Black circles and solid line: fiducial model with cube size = $30$ kpc and pixel size = $2$ kpc. Orange squares and triple dot-dashed line: cube size = $30$ kpc, \color{orange} pixel size = $3$ kpc\color{black}. Blue inverted triangles and dot-dashed line: \color{blue} cube size = $60$ kpc\color{black}, pixel size = $2$ kpc. Errors are Poissonian. We consider all gas in the edge-on xz projection $-$ $\sigma_{\rm y}$. Decreasing the resolution of the grid does not affect the $\sigma$-distribution, while changing the cube size from $30$ to $60$ kpc reduces the statistical relevance of the high-$\sigma$ tail.} \label{fig_hist_res_test} \end{figure} To obtain the distribution described by the orange squares and triple dot-dashed line, we kept fixed the cube size to $30$ kpc and increased the pixel size to $3$ kpc. From Fig. \ref{fig_hist_res_test}, it is clear that the decrement in grid resolution does not affect much the new $\sigma$-distribution, which is essentially a smoothed version of the fiducial model. On the other hand, changing the cube size from $30$ to $60$ kpc (while keeping the grid resolution fixed to $2$ kpc) has a large impact (blue inverted triangles and dot-dashed line). The new velocity dispersion distribution has a similar low-$\sigma$ part as the previous two, but declines more rapidly at $\sigma_{\rm y}>70$ km s$^{-1}$. This is due to the fact that the larger cubes include gas in the outskirts of galaxies, where the velocity dispersion is generally small. The additional pixels contribute mostly to the low-$\sigma$ part, while the increased total number of bins reduces the statistical relevance of the high-$\sigma$ tail. \section{testing the methodology with idealized simulations of disc galaxies} \label{constrsim} \begin{figure*} \centering \includegraphics[width=4.37cm, height=3.77cm]{gas_nofb_xz.png} \includegraphics[width=4.37cm, height=3.77cm]{gas_weak_xz.png} \includegraphics[width=4.37cm, height=3.77cm]{gas_fb1_xz.png} \includegraphics[width=4.37cm, height=3.77cm]{gas_strong_xz.png} \vspace{0.25cm} \includegraphics[width=4.37cm, height=3.97cm]{color_mapxz_nofb} \includegraphics[width=4.37cm, height=3.97cm]{color_mapxz_fb0p75} \includegraphics[width=4.37cm, height=3.97cm]{color_mapxz_fb1} \includegraphics[width=4.37cm, height=3.97cm]{color_mapxz_strong} \caption{{\bf First row}: positions in physical kpc of gas particles for our constrained disc galaxies. {\bf Second row}: corresponding gas velocity dispersion maps. From left to right: the first column shows the no feedback run; second column $-$ weak feedback run; third column $-$ fiducial model; fourth column $-$ strong feedback run. In each panel, we consider all gas in the edge-on xz projection $-$ $\sigma_{\rm y}$.} \label{fig_cdiscc_maps} \end{figure*} Throughout this work, we have analysed galaxies extracted from one particular run of the EAGLE simulations: {\textit{Recal-L025N0752}}. As discussed in Section \ref{conf}, the L025N0752 configuration has the highest resolution of the set, and therefore comes in only two slightly different setups: \textit{Ref} and \textit{Recal}. Fully exploring the parameter space around the fiducial model would have been numerically too demanding, and was done instead using configurations with lower resolution \citep[L050N0752 and L025N0376, see][]{crain2015}. Consequently, we have not been able to study the impact of variations in the feedback strength, which are important for interpreting real data and also validating our previous results. For this reason, we ran four simulations of an isolated disc galaxy. We used a different version of the {\small{GADGET-3}} code that includes smoothed particle hydrodynamics with a higher order dissipation switch (SPHS) than classic SPH. The main advantage of SPHS is that it suppresses spurious numerical errors in the calculation of fluid quantities before they can propagate \citep{hobbs2013, power2014}. At temperatures T$_{\rm gas}\ge10^4$ K, the gas is assumed to be of primordial composition and cools radiatively following \citet{KWH}. At T$_{\rm floor}=100$ K $\le$ T$_{\rm gas}<10^4$ K, the gas cools following the prescription of \citet{mash2008} for gas of solar abundance. Gas is also prevented from cooling to the point at which the Jeans mass for gravitational collapse becomes unresolved. Gas above a fixed density threshold forms stars with a star formation efficiency of $10$\% \citep{power2016}. The subgrid model for star formation is calibrated to follow the Kennicutt-Schmidt law \citep{kennicutt1998}. \begin{figure} \centering \includegraphics[width=8.37cm]{multi_hist_discs_combined_cumulative} \caption{Cumulative fraction of the pixelated velocity dispersion probability distribution (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) for our constrained disc galaxies. Black diamonds and solid line: fiducial model. Red triangles and dashed line: strong feedback. Orange crosses and triple dot-dashed line: weak feedback. Blue squares and dot-dashed line: no feedback. The thin grey horizontal line marks the saturation point of each distribution (cumulative fraction $=1$). We consider all gas in the edge-on xz projection $-$ $\sigma_{\rm y}$. As soon as feedback is introduced, the distributions shift to larger velocity dispersions.} \label{fig_hist_discs} \end{figure} The four {\it constrained} discs are identical in all but the strength of stellar feedback. The latter is quantified by the feedback factor $f_{\rm f}$, which simply (linearly) scales the amount of thermal energy deposited by supernovae into neighbouring gas particles. To calculate the energy released by SNe at any given time (where only SNe II are considered and $E_{\rm SN} = 10^{51}$ erg), the code integrates over a \citet{salpeter55} IMF in the range $8-100$ M$_{\odot}$ (to determine the number of SNe II) and adopts an approximate main-sequence time of $t_{\rm MS}\approx11.8$ Myr. At $t_{\rm MS}$ after the formation of the star particle, the energy injection is implemented as a delta function in time \citep{hobbs2013}. We varied the feedback factor from zero to strong feedback ($f_{\rm f}=1.50$), the fiducial model being the one with $f_{\rm f}=1.00$ (see Table \ref{tab_asym_discs}). Each galaxy is composed of a (gas + stellar) disc of radius $10$ kpc and a stellar bulge embedded in a DM halo (with a NFW $-$ \citealt{NFW1996} $-$ concentration parameter $c=10$), sampled with $308,012$ total particles (of which N$_{\rm DM} = 100,000$). The mass and spatial resolutions are $1.6\times10^5$ M$_{\odot}$ and $0.01$ kpc, respectively. All systems rotate (the DM halo spin parameter is $0.04$) and have the same total mass M$_{\rm tot}=2.57\times10^{12}$ M$_{\odot}$. Gas and stellar masses change according to $f_{\rm f}$ (as well as N$_{\rm gas}$ and N$_{\star}$), with $\langle$M$_{\star}\rangle=7.42\times10^{10}$ M$_{\odot}$. We stress that these discs are not meant to faithfully reproduce physical and morphological properties of the EAGLE galaxies previously used. Our main goal here is testing the methodology. \begin{table} \centering \begin{tabular}{llcc} \\ \hline & Run & Feedback & $\max(\sigma_{\rm y})$ \\ & & factor ($f_{\rm f}$) & [km s$^{-1}$] \\ \hline & strong feedback & 1.50 & 738.74 \\ & fiducial model & 1.00 & 786.67 \\ & weak feedback & 0.75 & 1117.06 \\ & no feedback & 0.00 & 157.30 \\ \hline \\ \end{tabular} \caption{Run name, feedback factor $f_{\rm f}$ (that regulates the amount of thermal energy injected by SNe into nearby gas particles) and gas maximum velocity dispersion $\sigma_{\rm y}$ for the edge-on projection of our idealized disc galaxies.} \label{tab_asym_discs} \end{table} In the first row of Fig. \ref{fig_cdiscc_maps} we plot positions of gas particles for each single disc. From left to right, the following runs are shown: no feedback, weak feedback, fiducial model and strong feedback. We consider all gas in the edge-on xz projection. All systems are plotted at the same evolutionary stage of $t\sim50$ Myr. As expected, the gas disc becomes more and more perturbed and the amount of outflowing material increases when moving from the no feedback to the strong feedback case. This trend is partially visible also in the second row of Fig. \ref{fig_cdiscc_maps}, where we show the corresponding gas velocity dispersion maps. It is interesting to note how the weak feedback run produces the largest velocity dispersions, $\max(\sigma_{\rm y})\approx1120$ km s$^{-1}$ (see Table \ref{tab_asym_discs}). This somewhat counterintuitive result is due to the fact that, compared to the fiducial and strong feedback simulations, only a few gas particles left the disc at the evolutionary stage considered and therefore the weak feedback $\sigma$-map is subjected to large statistical fluctuations. In Fig. \ref{fig_hist_discs} the cumulative pixelated velocity dispersion probability distributions of the constrained discs are visible. We consider all gas in the edge-on xz projection $-$ $\sigma_{\rm y}$. The no feedback run quickly saturates to one at $\sigma_{\rm y}=150$ km s$^{-1}$. In this case, only the disc component is present (according to the leftmost panel in the second row of Fig. \ref{fig_cdiscc_maps}). This $\sigma$-distribution is statistically different from the other three. As soon as feedback is introduced, the distributions shift to progressively larger velocity dispersions following the increase in feedback strength. This is particularly visible up to $\sigma_{\rm y}\sim 400$ km s$^{-1}$, where there is a crossover between the weak feedback (orange crosses and triple dot-dashed line) and the strong feedback (red triangles and dashed line) runs for the reason explained in the previous paragraph. These trends fully support the low-$\sigma$ $\Leftrightarrow$ disc + high-$\sigma$ $\Leftrightarrow$ outflows correlations emerging from the analysis of EAGLE galaxies presented in the paper. We refrained from calculating the $\xi$ and $\eta_{50}$ parameters for the constrained discs because the results would be completely dominated by statistical noise$/$fluctuations. \label{lastpage} \end{document} }
\caption{Pixelated velocity dispersion probability distributions (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) for galaxies in the four M$_{\star}-$sSFR quadrants defined in the bottom right panel of Fig. \ref{fig_gal11_proj}: Q$1$ = low M$_{\star}-$ low sSFR, \color{red} Q$2$ \color{black}= low M$_{\star}-$ high sSFR, \color{blue} Q$3$ \color{black}= high M$_{\star}-$ low sSFR and \color{orange} Q$4$ \color{black}= high M$_{\star}-$ high sSFR. {\bf Left panel}: edge-on xz projection $-$ $\sigma_{\rm y}$. {\bf Right panel}: face-on xy projection $-$ $\sigma_{\rm z}$. Errors are Poissonian and we consider only warm gas. In general, galaxies with higher M$_{\star}$ present a more extended $\sigma$-distribution, while the sSFR has a secondary effect with respect to stellar mass.}
\caption{{\bf Top panel}: relation between star formation rate surface density, $\Sigma_{\rm SFR}$, and M$_{\star}$ for our simulated galaxies. $\Sigma_{\rm SFR}={\rm SFR}/(2\pi r_{50}^2)$, where SFR is the total star formation rate and $r_{50}$ is the radius within which half of the galaxy stellar mass is included. The plot is divided in four sectors: S$1$ = low M$_{\star}-$ low $\Sigma_{\rm SFR}$, \color{red} S$2$ \color{black}= low M$_{\star}-$ high $\Sigma_{\rm SFR}$, \color{blue} S$3$ \color{black}= high M$_{\star}-$ low $\Sigma_{\rm SFR}$ and \color{orange} S$4$ \color{black}= high M$_{\star}-$ high $\Sigma_{\rm SFR}$ (see Section \ref{sigma_SFR}). {\bf Bottom panels}: pixelated velocity dispersion probability distributions (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) for galaxies in the four sectors defined in the top panel. Errors are Poissonian. We only consider the edge-on xz projection $-$ $\sigma_{\rm y}$. Left panel: warm gas. As in the previous section, a trend with stellar mass is visible. Right panel: all gas. At fixed M$_{\star}$, galaxies with high $\Sigma_{\rm SFR}$ give rise to a $\sigma$-distribution shifted to larger velocity dispersions compared to galaxies with low $\Sigma_{\rm SFR}$.}
\caption{{\bf Left panel}: asymmetry parameter $\xi$ vs velocity dispersion to rotation ratio parameter $\eta_{50}$. Dots: simulated edge-on xz projection. Squares: simulated edge-on yz projection. Synthetic data are colour coded according to the SFR surface density of the parent galaxy, $\Sigma_{\rm SFR}={\rm SFR}/(2\pi r_{50}^2)$. The vertical and horizontal dotted lines mark the observational limits above which galaxies show strong disc-halo interactions according to \citet{ho2016}, $\eta_{50}>0.3$ and $\xi>1.8$. Crosses: observational SAMI data, where red \color{red} \textsf{x} \color{black} symbols represent wind-dominated galaxies and black \textsf{x} symbols represent galaxies without strong wind signatures. {\bf Right panel}: correlation between the asymmetry parameter $\xi$ and $\Sigma_{\rm SFR}$. Grey dots and squares: simulated galaxies in the edge-on xz and yz projections, respectively. Crosses: observational data from \citet{ho2016}. In both panels, the range of the simulated parameters (determined using all gas) is broadly consistent with the observations. A clear positive $\xi-\Sigma_{\rm SFR}$ correlation is visible in both EAGLE and SAMI data.}
\caption{Pixelated velocity dispersion probability distribution (cf. Fig. \ref{fig_total_vs_ha} and the beginning of Section \ref{total_vs_ha}) of galaxies: resolution tests. Black circles and solid line: fiducial model with cube size = $30$ kpc and pixel size = $2$ kpc. Orange squares and triple dot-dashed line: cube size = $30$ kpc, \color{orange} pixel size = $3$ kpc\color{black}. Blue inverted triangles and dot-dashed line: \color{blue} cube size = $60$ kpc\color{black}, pixel size = $2$ kpc. Errors are Poissonian. We consider all gas in the edge-on xz projection $-$ $\sigma_{\rm y}$. Decreasing the resolution of the grid does not affect the $\sigma$-distribution, while changing the cube size from $30$ to $60$ kpc reduces the statistical relevance of the high-$\sigma$ tail.}
\caption{Ablation study on the Cityscapes validation set using the VGG based Deeplab-LargeFOV model. Black numbers for the convolutional layers indicate fixed dilation parameters, {\color{red}red} numbers or ranges in our learnable dilated convolution layers indicate the initial values or distributions before training. }
\caption{Actual ground-state solutions \textcolor{magenta}{(dotted)} to our three-dimensional bound-state equation \cite{WL05:LS}, obtained straightforwardly by application of variational techniques [$\varphi_2(r)$, left] or a conversion to an equivalent matrix eigenvalue problem [$\varphi_2(p)$, right], vs.\the initial Salpeter component\textcolor{blue}{(dashed)}, this inversion's \emph{starting point\/}.}
\caption{Additional \aastex\symbols}
\caption{Sketch of the optical path of the laser system and visualization as arranged on an optical table with a size of $2.4\,\mathrm{m}$ x $1.5\,\mathrm{m}$ (shown not to scale). The symbols on top of the light path indicate out-of-plane (\protect\includegraphics{Symbol_OutPol.pdf}) and in-plane polarization (\protect\includegraphics{Symbol_PlanePol.pdf}). Optical mirrors operated at an angle of incident (AOI) of $45^\circ$ are represented by black solid lines. The mirror labeled M1 serves as reference point in \Fig\ref{fig:BeamPath}, where an advanced laser-beam path is shown to scale. The drop defines the origin of our coordinate system, where $\vec{e}_z$ is aligned to the laser-beam propagation and $\vec{e}_x$ to gravity. Abbreviations that are used in the figure are explained in Nomenclature.\label{fig:Optics}}
\caption{Clustering performance comparisons on various data sets. The leftmost shows our \textbf{rank1count} by setting a threshold automatically. For the rest of the columns, we show f-scores using optimal (oracle-supplied) thresholds. For BBT and Buffy, we show average scores over six episodes. The full table with individual episode results is given in Appendix~\ref{sec:performance_comparisons}. Best viewed in color (\textcolor{red}{\textbf{1st place}}, \textcolor{orange}{\textbf{2nd place}}, \textcolor{ForestGreen}{\textbf{3rd place}}).}
\caption{{\bf Clustering results from {\em Buffy the Vampire Slayer}}. A failure example can be seen in frame (e), in which the main character Buffy (otherwise in a \textcolor{RoyalPurple}{purple box}) in shown in a \textcolor{RubineRed}{pink box}.}
\caption{{\bf Clustering results from {\em the Big Bang Theory}}. A failure example can be seen in frame (d), in which the main character Howard (otherwise in a \textcolor{Magenta}{magenta box}) in shown in a \textcolor{Gray}{gray box}.}
\caption{Clustering performance comparisons evaluated on traditional measures. (\textcolor{red}{\textbf{1st place}},\textcolor{orange}{\textbf{2nd place}},\textcolor{ForestGreen}{\textbf{3rd place}}).}
\caption{\textcolor{rev}{ With each (partial) edge on the boundary, only one vertex is associated.}}
\caption{Dust mass derived from parameter set ($A_V$, $T_\mathrm{d}$) for Himiko in Model A. {The dust mass corresponding to ($A_V$, $T_\mathrm{d}$) is shown by the grey scale,} and the level of dust mass is shown in the bar on the top. {The dust mass is only shown in the area where the SED fitting is successful. The yellow shaded area is the region where the mm flux exceeds the ALMA upper limit, while the red shaded area is the region where we do not obtain a satisfactory fit to the rest UV--optical data} (i.e.\reduced$\chi^2\geq 3$).}
\caption{ a) Sheet resistance versus temperature for five different samples. b-f) ARRES measurements for conducting (\SampleCond{}, \SampleTwente{}) and non-conducting (\SampleIns{}, \SampleSrO{}, \SampleSputtered{}) samples. Sample~\SampleCond{} (b,~\ElectricARRESCond{}) 8~u.c.\@\LAO{} grown in the LEEM, sample~\SampleTwente{} (c,~\ElectricTwente{}) 4~u.c.\@\LAO{} grown in a conventional PLD setup, sample~\SampleIns{} (d,~\ElectricOutOfFocus{}) 8~u.c.\@\LAO{} grown with out-of-focus PLD laser, sample~\SampleSrO{} (e,~\ElectricSrO{}) 5~u.c.\@\LAO{} grown on \SrO{}-terminated \STO{} and sample~\SampleSputtered{} (f,~\ElectricSputtered) 5~nm \LAO{} grown with sputter deposition. }
\caption{Phase diagram of the purely repulsive Yukawa system as a function of the Coulomb coupling parameter $\Gamma$ and the screening parameter $\kappa$. The experimentally investigated samples in the different single-phase regions are represented by {\textcolor{dgreen}{$+$}} (fcc), {\textcolor{red}{$\ocircle$}} (bcc) and {\textcolor{blue}{$\times$}} (fluid) and the three coexisting phases in the sample under triple conditions by {\textcolor{dgreen}{\ding{72}}} (fcc), {\textcolor{red}{\ding{72}}} (bcc) and {\textcolor{blue}{\ding{72}}} (fluid). The theoretically predicted fluid-solid coexistence line is indicated by {\textcolor{red}{$\blacktriangle$}}~\cite{Hamaguchi1997}, {\textcolor{blue}{$\medbullet$}}~\cite{Meijer1991}, {\textcolor{green}{$\blacksquare$}}~\cite{Stevens1993} and the corresponding analytical expression $\Gamma = 106\,{\rm e}^{\kappa}/(1+\kappa+0.5\kappa^2)$~\cite{vaulina2000} by a solid line. The theoretically predicted fcc-bcc coexistence is indicated by {\textcolor{blue}{$\blacktriangle$}} connected by a line. The inset shows the data close to the triple point.}
\caption{Evolution of Ad-Blockers on Google Play: \textcolor{red}{(a)} Number of Ad-Blockers available on Google Play over time. \textcolor{red}{(b)} Distribution of app rating vs. installs per Ad-Blocker. }
\caption{Screenshots showing inefficiency of DashVPN, a system-level Ad-Blocker: \textcolor{red}{(a)} DashVPN does not block click-baits from Taboola (\texttt{taboola.com}) and \textcolor{red}{(b)} shows full-screen ads from Google.}
\caption{\blue{Group sizes for simulation scenarios for the Melon dataset. We vary the numbers of observations per group in order to simulate highly imbalanced group sizes.}}
\caption{\blue{The performance of the evaluated classification methods for highly imbalanced datasets is evaluated. The six scenarios refer to the scenarios described in Table \ref{tab:ImbalancedMelon}. Especially setup 1 and 6 cause a problem for most approaches while LP presents itself as mostly robust towards imbalanced group sizes.}}
\caption{{\it Top panel:} The light curve of $0.1<\:E_\gamma\:<500$ GeV \grays from NGC 1275 from August 4, 2008 to March 5, 2017, with 3-day (blue) binning. {\it Middle panels:} Sub intervals covering F1 (left) and F2 (right). F1 is shown with 8-hour (blue) and 12-hour (red) time intervals and F2 with 3-hour (blue) and 6-hour (red) bins. The red dashed lines show the fit of F1 and F2 with Eq. \ref{func}. {\it Lower panels:} The plot of Npred/$\sqrt{\rm Npred}$ vs Flux/$\Delta {\rm Flux}$ for 8-hour (left) and 3-hour (right) bins.}
\caption{Qualitative comparison between our proposed joint deep set network (JDS) and the deep set networks with Dirichlet-Categorical (DS (DC)) and Negative Binomial (DS (NB)) as the cardinality loss. For each image, the ground truth tags and the predictions for our JDS and the two baselines are denoted below. {\textcolor{red}{False positives}} are highlighted in red. Our JDS approach reduces both cardinality and classification error.}
\caption{More results generated by the proposed framework compared with the original descriptions provided by the Flickr users. \textcolor{blue1}{Blue} indicates a precise description of the image content that does not even appear in the original descriptions. \textcolor{green1}{Green} shows a successful recovery of the landmarks. Figure is best viewed in color.}
\caption{Results generated by the proposed framework that could be further improved. \textcolor{purple1}{Purple} indicates wrong or unrelated phrases. \textcolor{red1}{Red} shows a wrong location or activity based on the original descriptions provided by the users. However, for some cases, these locations or activities cannot be recovered solely based on the image content. Figure is best viewed in color.}
\caption{The \gray light curve of 3C 120 from August 4, 2008, to August 4, 2016. (a) The bin intervals correspond to 90 days. (b) The light curve obtained by adaptive binning method assuming 20 \% of uncertainty. The change of photon index is shown in the insert.}
\caption{\label{R1}(a) and (b) Plot of mean translocation time and standard deviation, obtained from 2000 successful translocation events, for a polymer of length $N=32$ translocating through pore of width $W=3$. The polymer entering the pore from the stiff end ($S_nF_n$) is shown by symbol \textcolor{goldenrod}{$\RIGHTcircle$} while the polymer entering from the flexible end ($F_nS_n$) is shown by \textcolor{goldenrod}{$\LEFTcircle$}. The accuracy of detecting heteropolymers (c) $F_{2}S_{2}$ (\textcolor{magenta}{$\blacktriangledown$}) and (d) $S_{8}F_{8}$ (\textcolor{blue}{$\blacktriangle$}) through Pores $\beta$ and $\gamma$. }
\caption[]{\la{fig:repeatability_small} Turbulence statistics with different post arrangements with $\phi_s=1/9$ at $Re_\tau\approx200$. (\aaa) Mean streamwise velocity profile; (\bbb) $w^+\approx4$, rms velocity and pressure fluctuations; (\ccc) $w^+\approx26$, rms velocity and pressure fluctuations. \dashed: smooth wall. \linesolidsquar, R04$_{1}$; \linesquar, R04$_{2}$; \linesoliddtri, R26$_{1}$; \linetri, R26$_{2}$; \linesolidrtri, R26$_{3}$; \lineltri, R26$_{4}$. }
\caption[]{ (\aaa) Slip length, $b^+$, versus texture size, $L^+_n=w^+/\sqrt{\phi_s}$, with $\phi_s=1/9$. Symbols are DNS solutions; \solidcircle: randomly distributed posts, \smalltimes: aligned posts \citep{Seo2016}. Lines are Stokes flow solutions; \solid: Equation (\ref{eq:sbrpros}) by \cite{Sbragaglia2007a}; \dasheddot: Ybert \etal \cite{Ybert2007} (\bbb) Slip length, $b^+$, normalized by texture width, $w^+$, versus solid fraction, ${\phi_s}$, at $w^+\approx 4$. \solidcircle: R04$_{1}$, \soliddtrian: R04$\phi_{s16}$, \solidsqua: R04$\phi_{s25}$, \solid: \cite{Sbragaglia2007a} (\ccc) Mean velocity profiles subtracted by slip velocity for $\phi_s=1/9$, \linecirc, Patterned slip with randomly distributed posts, R04$_{1}$; \linesolidcirc, Patterned slip, R26$_{1}$; \linediam, Homogenized slip using the slip length from R04$_{1}$, \linesoliddiam, Homogenized slip using the slip length from R26$_{1}$. \dashed: smooth-wall. }
\caption[]{ (\aaa) Wall rms pressure fluctuation due to stagnation, $\tilde{p}^+_{rms_{0}}$ against nominal texture size $L^+_n$; (\bbb) $\tilde{p}^+_{rms_{0}}$ versus slip velocity $U_s^+$. \linesolidcirc: randomly distributed posts; \linecross: aligned posts. (\ccc) Self-similar profiles of wall pressure distribution due to stagnation on the gas-liquid interface. The plots show pressure profiles on the centerline going through the middle of post width. $L^+_n\approx13$; \linesolidtri: $L^+_n\approx26$; \linesolidrtri: $L^+_n\approx38$; \linesoliddtri: $L^+_n\approx77$. }
\caption[]{ Interface deformation angle, $(d\eta/dx|_L)$, due to stagnation pressure for randomly distributed posts versus number of posts, $N_p$, as compared to aligned posts with Weber number $We^+={5\times 10^{-3}}$ for texture size (\aaa) $L_n^+\approx77$; (\bbb) $L_n^+\approx26$. \circle: Randomly distributed posts; Insets: histogram of interface deformation angle due to stagnation pressure for randomly distributed posts versus number of posts. \solid: mean of $(d\eta/dx|_L)$ of randomly distributed posts; \dashed: $(d\eta/dx|_L)$ of aligned posts. }
\caption[]{ (\aaa) Comparison of maximum interface deformation slope, max $(d\eta/dx)|_{L}$, due to stagnation pressure for randomly distributed posts and aligned posts versus effective texture size, $L^+_n=w^+/\sqrt{\phi_s}$, at Weber number $We^+= {5\times 10^{-3}}$. \linesolidcirc: randomly distributed posts; \linecross: aligned posts \citep{Seo2015}. For $L^+_n\approx13$ and $L^+_n\approx 77$, error bars are plotted by taking account for different texture arrangements. (\bbb) The critical texture size, $L^+_c$, is plotted against $We^+$ when $\theta_{adv}=120^{\circ}$. \solid: current study, \dashed: aligned posts. }
\caption{Summary of experimental results: ``average accuracy $\pm$ SD ({\color{blue} rank})". The ``ave. rank" column shows the average rank of each method. The lower the average rank, the better the overall performance of the method.}
\caption{\redHL{Accuracy loss distribution for BS. \label{fig:hist_bs}}}
\caption{\redHL{Accuracy loss under SPS+ for different {\em targetAccLoss} values} }
\caption{\redHL{Accuracy loss distribution for PR with gnu04 input set. \label{fig:hist_pr_gnu04}}}
\caption{\redHL{Accuracy loss distribution for PR with gnu05 input set. \label{fig:hist_pr_gnu05}}}
\caption{\redHL{Accuracy loss distribution for PR with pk input set. \label{fig:hist_pr_pk}}}
\caption{ \redHL{Statistical fault injection outcome for temporal noise tolerance. Bit 23 on the x-axis demarcates the most significant; bit 0, the least significant.} \label{fig:profile}}
\caption{\redHL{Energy consumption under SPS+ for different {\em targetAccLoss} values.} \label{fig:energyMinDPS}}
\caption{\redHL{Input Sensitivity Analysis for Pagerank (PR) with different datasets from the Snap Database\protect\cite{snapnets}. Only \textbf{gnu04} is deployed for profiling.}}
\caption{Pearson correlation coefficient {\it vs.} distance correlation coefficient for 528 pairs of variables, calculated for galaxies with redshift $0 \leq z < 0.5$. The subplots for each galaxy type are shown in four middle frames together with the superposition of the four subplots in the large left frame. The locations of potential outlier pairs in the scatter plot for {\bf\color{OliveGreen}Type 3 (Spiral: Sbc)} galaxies is also shown in the large right frame. }
\caption{Comparison of the classification accuracies ($\%$) on the Parliament dataset \cite{VrigkasM_SETN14}. Results highlighted with \colorbox{blue!15}{light purple} indicate statistically significant improvement over the second best method using paired t-test.}
\caption{Comparison of the classification accuracies ($\%$) on TVHI the dataset \cite{Perez12}. Results highlighted with \colorbox{blue!15}{light purple} indicate statistically significant improvement using paired t-test.}
\caption{Comparison of the classification accuracies ($\%$) on the SBU dataset \cite{Ykiwon12}. Results highlighted with \colorbox{blue!15}{light purple} indicate statistically significant improvement using paired t-test.}
\caption{Comparison of the classification accuracies ($\%$) on the USAA dataset \cite{FuHXG12}. Results highlighted with \colorbox{blue!15}{light purple} indicate statistically significant improvement using paired t-test.}
\caption{Comparison of the classification accuracies ($\%$) on Parliamment \cite{VrigkasM_SETN14}, TVHI \cite{Perez12}, SBU \cite{Ykiwon12}, and USAA \cite{FuHXG12} datasets using CNN features. Results highlighted with \colorbox{blue!15}{light purple} indicate statistically significant improvement using paired t-test.}
\caption{ \highlight{$uv$-coverage of our interferometric observations.} }
\caption{\blue{Values of $e(k,r)$ defined by \eqref{e:log-error} for different sinc quadrature spacing $k$ and $r=0$ (left), $r=\beta$ (middle) and $r=1$ (right). Here $M$ and $N$ are chosen to balance the three error terms coming from the sinc quadrature $\HH^r(\Omega)$ for a given sinc quadrature spacing $k$ (see, Remark~\ref{rem:MN})}. }
\caption{\blue{Values of $e(k,r)$ defined by \eqref{e:log-error} for different sinc quadrature spacing $k$ and $r=0$ (left), $r=\beta$ (middle) and $r=1$ (right). In contrast with the experiments provided in Figure~\ref{f:sinc}, here $M=N=1/k^2$. }}
\caption{Classification accuracy of different learning paradigms on the SoBiR dataset. In individual learning, each attribute is learned separately. In multi-task learning, the average loss of all attributes is backpropagated in the network. In CILICIA, four clusters were formed and attributes are in descending order based on their intra cross-correlation. Results highlighted with \colorbox{blue!15}{light purple} indicate statistically significant improvement using the paired-sample t-test.}
\caption{Performance comparison on the VIPeR dataset. Five clusters were formed and attributes are in descending order based on their intra cross-correlation. Results highlighted with \colorbox{blue!15}{light purple} indicate statistically significant improvement using the z-test.}
\caption{The wall-normal profile of the optimal forcing with three different eddy viscosities in (\ref{eq:2.2}): (\textit{a}) $|\hat{f}_{w,v}|$, (\textit{b}) $|\hat{f}_{w,w}|$. Here, \protect \solid, the Cess eddy viscosity ($\nu_t=\nu_t(y)$ in (\ref{eq:2.3})); \protect \dashdot, a constant eddy viscosity ($\nu_t=\max_y \nu_t(y)$); \protect \dashed, no eddy viscosity ($\nu_t=0$).}
\caption{ Results of applying GRC to the symmetric matrices of Figure \textcolor{green!50!black}{1} and the matrix (1) in the Sec. 3.2 (which is also turned out to be symmetric). 'GRC2' denotes GRC(L=2) and so on. GRC with $L \geq 3$ generates almost identical residual norms at the beginning in both graphs, as predicted by theory. The difference at later iterations in the left seems due to the accumulation of floating point errors. }
\caption{ Results of applying GRC using eq. (\ref{eq.psiresid}) instead of eq. (\ref{eq.cprcpsi}) to the symmetric matrices of Figure \textcolor{green!50!black}{1} and the matrix (1) in the Sec. 3.2. In this case, $L \geq 2$ is sufficient in theory. In the right, all the contours overlap completely. }
\caption{Qualitative results of dual-trained VQA~(top) and VQG~(bottom) models. Labeled questions/answers are given to infer the counterpart. \textcolor{red}{Red} denotes the failure cases. \textcolor{green}{Green} indicates the correct ones. The question is viewed as \emph{correct} if it corresponds to the given answer. }
\caption{Comparing effect of code length (in bits) on recall (in \%). Our model can convert raw real valued features (\textcolor{green}{green}) to binary codes (\textcolor{red}{red}) which boosts the performance. Binary codes obtained by popular unsupervised hashing method (LSH) (\textcolor{blue}{blue}) have lower performance than the corresponding raw real-valued features.}
\caption{The probability of an intrusion subsequence. Process events ($S \setminus I$) are \textbf{black}, intrusion events ($I$) are \textcolor{red}{\textbf{red}}, extra events at the beginning and at the end are \textcolor{gray}{\textbf{gray}} and dashed.}
\caption{\label{quali}Qualitative assessment of model R9. The three columns on the left show simulated images, while the right columns use photographs. In some cases, details are magnified in insets with increased contrast. Top row: input image, middle: {\color{red} hook-pull}, {\color{OliveGreen} sit}, {\color{MidnightBlue} break}, bottom: {\color{red} place-on}, {\color{OliveGreen} walk}, {\color{MidnightBlue} grasp}. Note, in case of simultaneous presence of two affordances the colors blend. }
\caption{Relative interest over time in web searches according to Google Trends (\url{www.google.es/trends/}). Terms legend: CRISP-DM in \textcolor[rgb]{0,0,1}{blue}, KDD in \textcolor[rgb]{1,0,0}{red}, SEMMA in \textcolor[rgb]{0,0.58,0}{green} (the latter having a relative interest close to zero).}
\caption{ \label{fig1} \blue{A schematic representation of a CSTF's single helical column, with the structural period $2 \Omega$ and inclination angle $\chi$ indicated.} }
\caption{ \label{fig2} Scaled value of $\ln\lec{\sf det}[{\underline{\underline Y}}(q)]\ric$ mapped against Re$\les q \ris/\ko$ and Im$\les q \ris/\ko$ for $\psi = 30^\circ$. The darkest regions contain the zeros of ${\sf det}[{\underline{\underline Y}}(q)]$. \blue{Solutions listed in Table~\ref{tab1} are indicated.} }
\caption{Results on category prediction \textcolor{blue}{TODO: (1) compute top-3 for visual nets; (2) compute chance level; (3)}}
\caption{System setup for the robotic scrub nurse. The surgeon is conducting a surgery (black) while the robotic nurse (\textcolor{orange}{orange}) picks up the requested instrument from mayo stand (\textcolor{brown}{brown}) and delivers to surgeon. The surgeon is monitored by Myo armband (\textcolor{green}{green}), Epoc headset (\textcolor{red}{red}) and Kinect (\textcolor{purple}{purple}) for turn-taking prediction.}
\caption{Cluster of stable and unstable phase in information space. Each point is a row of the emission matrix for the reduced Markov model with $2$ states. The plot shows the change in the Markov model as the process moves from stable and unstable. \color{red} Red diamonds \color{black} represent the unstable phase while \color{green} green diamonds \color{black} represent the stable phase.}
\caption{\magenta{Fig.~5\,(a) in Ref.~\cite{Cui:2017nnn}. The upper side of each curve is excluded with $1.6\,\sigma$ C.L. We superimpose the purple line representing Eq.~\eqref{val:sigma} and the light purple region denoting the error of $\sigma_\text{SI}$ coming from the uncertainty of the overall coupling $f_N=0.30\pm0.03$; see Ref.~\cite{Cline:2013gha}.}}
\caption{Upper: Allowed regions for $\lambda_S=0$. For each $m_t$, the region above the line with corresponding color is allowed. Each vertical line denotes the lower bound on $m_\tx{DM}$ from the positivity of potential: $V_{\varphi\leq\Lambda}>0$. The envelope of the rainbow-colored lines, indicated by the black line, gives the lower bound on $r$ for each $m_\tx{DM}$ when one varies $m_t$. %We have shaded the region $m_\tx{DM}\geq1585\GeV$ where perturbativity is violated. (See Fig.~\ref{fig:mtfix-main} for the corresponding plot with right-handed neutrinos.) Lower: Excluded regions for $\lambda_S=0$ (lower left) and $\lambda_S=0.6$ (lower right). %The vertical black (gray) shade in $m_\tx{DM}\geq1585\GeV$ ($m_\tx{DM}\geq1223\GeV$) is excluded by the perturbativity for $\lambda_S=0$ (0.6). \magenta{ The blue and red regions are excluded by the upper bound on the tensor-to-scalar ratio $r<0.09$~\cite{Ade:2015xua} and by lower bound on the DM mass $m_\tx{DM}>0.7\TeV$ from PandaX-II~\cite{Cui:2017nnn}, respectively.}%. \magenta{The red region is excluded by the lower bound on the DM mass %$m_\tx{DM}=630\GeV$ from XENON1T~\cite{Aprile:2017iyp} (solid) and % $m_\tx{DM}>0.7\TeV$ from PandaX-II~\cite{Cui:2017nnn}.} }
\caption{The lower bound on $r$ for each fixed $m_t$ (colored) with $\lambda_S=0$. The black line is their envelope, which is identical to the ones in Figs.~\ref{fig:houraku2sigma} and \ref{fig:houraku4sigma} except for their right-most boundary where they follow the $m_t=176\GeV$ (blue) and 178\,GeV (purple) lines in this figure, respectively. See the left of Fig.~\ref{fig:N0} for the corresponding plot without right-handed neutrinos and for the explanation of the shaded region. }
\caption{Matrix that shows the median average \# refactored API classes per library. For each library, we summarized the median values across all library versions. Table includes median ($\bar{x}$) of matched refactored classes. 0 represents a value less than 0.01. (--) reports no matched classes. }
\caption{Classification of \RO~for API classes with \textit{presver} ratio. Note that one class may be classified under several refactoring types. Note (--) represents no matches. We also show the total of all breakages (cu. + ncu.) and use to colors to highlight when \colorbox{green!25}{prsv} = low and \colorbox{red!25}{prsv} = high.}
\caption{Sketch of reduction from \mincoloring~to \textsc{Frog/Br/Opt}. Agent $x$'s paths correspond to deciding a color $\gamma$ for each vertex $i$. Depicted in the gray rectangle, the first time $x$ chooses color $\gamma$ is on vertex $i$ (first time on the line). Then (1) $x$ waits one step on edge $(v_{i,\gamma},v_{i,\gamma}')$ (because of an agent $r_1$). Then on the same line, (2) neighbors in $\CG$ are Backfired (can't put the same color on a neighbor) and non-neighbors are discounted to delay zero (by heavily delaying agents $r_1$ and disarming their eventual backfires). Transit edges (dotted) have delay one to allow for backfires to work. Hence, a valid coloring of size $k$ would correspond to a path $\pi_x$ of length $\eta+k$ which does not enable to find $\beta$ for an $L$-reduction. Therefore, to bring the correspondence back from affine to linear, we technically multiply all the costs by $M$, with $M$ times more agents and vertices, but not on the dotted edges. }
\caption{DGP1: Empirical coverage rates for $T=100$. `\textbf{OLS}': (unrestricted) VAR in levels estimated by OLS; `\textbf{True Rank}': VECM estimated with knowledge of the true rank; `\textbf{AIC}' and`\textbf{BIC}': rank estimation using AIC and BIC, respectively; `\textbf{BERS/AIC}' and `\textbf{BERS/BIC}': Bootstrap Endogenous Rank Selection with respectively AIC and BIC used for rank selection; `\textbf{MA}' : Model Averaging with weights as in \eqref{eq:weights}; `\textbf{FDBb/AIC}' and `\textbf{FDBb/BIC}': FDB bagging with respectively AIC and BIC used for rank selection; `\textbf{WIMP}': WIMP method with weights as in \eqref{eq:weights}. The {\bf\color{Pink2}{pink}} lines show CPs for all nine impulse responses; the {\bf\color{red}{red}} line is the median of these per horizon. For ease of comparison, the median and minimum coverage of the OLS intervals is always reported in \textbf{black}.}
\caption{68\% confidence intervals of impulse responses to a government spending shock identified as in \citet{BlanchardPerotti02}. \textbf{Dashed} lines are OLS intervals, {\bf\color{HotPink2}{dotted}} lines FDBb/AIC intervals, {\bf\color{red}{solid}} lines WIMP intervals.}
\caption{68\% confidence intervals of peak multipliers implied by government spending and tax-cut shocks based on \citet{BlanchardPerotti02} \textbf{[B\&P]}, \citet{Ramey11} \textbf{[Ramey]}, \citet{MountfordUhlig09} \textbf{[M\&U]} and \citet{MertensRavn12,MertensRavn14} \textbf{[M\&R]}. \textbf{Dashed} lines are OLS intervals, {\bf\color{HotPink2}{dotted}} lines FDBb/AIC intervals, {\bf\color{red}{solid}} lines WIMP intervals.}
\caption{Sequence diagram of the trading workflow. Solid lines represent RIAPS messages and Ethereum transactions, while dashed lines represent smart-contract events. Messages and transaction in \textcolor{red}{red} stop the trading workflow.}
\caption{A schematic bifurcation diagram that summarises the results obtained for various parameters considered in the present work. Equilibria are presented by lines ({\color{black} \rule[0.6 pt]{10 pt}{1.2 pt}}, stable equilibria; {\color{black} \rule[0.6 pt]{5 pt}{1.2 pt}}\;{\color{black} \rule[0.6 pt]{5 pt}{1.2 pt}}, unstable equilibria; \protect\tikz \protect\fill[black] (0.5ex,0.5ex) circle (0.3ex); {\color{black} \rule[0.6 pt]{10 pt}{1.2 pt}}, stability depends on parameters). The lines with the double arrows indicate regimes where flapping occurs, with the top and bottom lines representing the peak-to-peak flapping amplitude. The diagram shows that with decreasing $K_B$ (moving left to right), the system transitions from the undeformed equilibrium to a stable deformed equilibrium. Following this, the system bifurcates to small-deflection deformed flapping, then large-amplitude flapping (chaotic flapping can also occur in this regime depending on parameters), and finally to a deflected-mode regime whose dynamics depend on Reynolds number: for $Re = 20$, no flapping occurs and the large-deflection state is an equilibrium of the fully-coupled system; for $Re = 200$, the large-deflection state is characterised by small-amplitude flapping. The above diagram only corresponds to cases when flapping occurs. For $Re = 20$ and $M_\rho < O(1)$, no flapping occurs and the only two regimes are the undeformed and deformed equilibria.}
\caption{Markers: peak flapping frequency at $Re = 200$ for the parameters corresponding to the bifurcation diagrams shown in figure \ref{fig:bif_Re200}; (\dashedrule): bluff-body shedding frequency ($=0.2/L_p$, where $L_p$ is the projected length of the flag to the flow defined using the maximum tip deflection at a given stiffness).}
\caption{Comparison among models over the UCI datasets. \textcolor{red}{Red} indicates the best result, followed by the second best in \textcolor{blue}{blue}. All models were evaluated based on NMSE values, the lower is the better. }
\caption{Comparison among models for prediction of \textbf{arousal} using different modality features: audio (Audio), video appearance (V-A), and video geometry (V-G). \textcolor{red}{Red} indicates the best result, followed by the second best in \textcolor{blue}{blue} based on CCC.}
\caption{Comparison among models for prediction of \textbf{valence} using different modality features: audio (Audio), video appearance (V-A), and video geometry (V-G). \textcolor{red}{Red} indicates the best result, followed by the second best in \textcolor{blue}{blue} based on CCC.}
\caption{Initial mode assignment (\S\ref{sec:cluster}) for bouncing ball trajectories (\S\ref{sec:ball}). The ball falls~($m_1$), reaches the \textcolor{resetfrom}{guard region}, jumps to \textcolor{resetto}{post-guard region}, and bounces up~($m_2$).}
\caption{ \textcolor{blue}{ Additional experimental results for LANDMARKS dataset: \emph{mid} (first row), \emph{low} (second row), and \emph{night} (third row) illumination conditions. Each row contains the same four plots as in the main Figure \ref{fig:LANDMARKS_main}. Namely, the first column shows the test overall accuracy (OA), the second one the CPU time needed to train each method, the third column shows the trade-off between test OA and training CPU time, and the last one provides the CPU time needed at the test step (production time). The legend, which is common to all plots and is included only in the top left one, is also the same as in Figure \ref{fig:LANDMARKS_main}. } }
\caption{ \textcolor{blue}{ Experimental results for LANDMARKS dataset, \emph{high} illumination condition. From left to right and top to bottom, the first plot shows the test overall accuracy (OA) of RFF-GPC, VFF-GPC, and GPC for the different values of $n$ (number of training examples) and $D$ (number of Fourier frequencies) considered. The second one is analogous, but displays the CPU time needed to train each method (instead of the test OA). The third plot summarizes the two previous ones, providing a trade-off between test OA and training CPU time. The last plot is analogous to the first and second ones, but showing the CPU time used at the test step (production time). The legend for second and fourth plots is the same as the one in the first plot. However, notice that in the third plot the GPC lines degenerate into single points (since GPC does not depend on $D$). In both legends, the numbers indicate the amount $n$ of training examples used, which determines the width/size of the lines/points too. As further explained in the main text, shown results are the mean over five independent runs. } }
\caption{\label{fig:Application} (a) Eight SDSS observations of asteroid (5129) Groom plotted in the scheme proposed. The error bars represent the errors transformed to PC space. The spectra of the observations are plotted together in the lower corner of the graph. The specific SDSS observation IDs are s14307, s0e2e7, sca267, sc8e1d, scf0dc, scdde4, scc719, and scbc65. (b) Three asteroids with multiple observations presenting spectral variability. The SDSSMOC observation IDs for each asteroid: (2252) CERGA (\textcolor{blue}{blue}): s3bafc, s3f6f3, s3eea4, s41189, s424bf, sbbb68, sf0125, sef985, sef5d2; (3631) Sigyn (\textcolor{purple}{purple}): s16290, s16001, sf0e38, s042fc, s132c2; and (55926) 1998 FE60 ({\bf black}): s17150, s18825, s17c18, s3924d, s2cd9b.}
\caption{Attribute-based PR scores on OTB benchmark compared with recent trackers, where the best results of deep and non-deep trackers divided by dash line are in \textcolor{red}{red} and \textcolor{green}{green} colors, respectively. }
\caption{Attribute-based SR scores on OTB benchmark compared with recent trackers, where the best results of deep and non-deep trackers divided by dash line are in \textcolor{red}{red} and \textcolor{green}{green} colors, respectively. }
\caption{Variation of $\varphi$ as a function of (a) Reynolds number when Stokes number and heat mixing parameter are kept constant. \red{Subscripts $\eta$ and $l$ refer to normalizing by Kolmogorov and large-eddy turnover times, respectively: $St_l = \tau_p/\tau_l, St_{\eta} = \tau_p / \tau_{\eta}, \sigma_l = \tau_{th}/\tau_l, \sigma_{\eta}=\tau_{th}/\tau_{\eta}$;} (b) heat mixing parameter $\sigma_l$ and Stokes number $\St_l$.}
\caption{ (a) and (b) show the trajectories in a delay-time coordinate. The solid line denotes the ground truth and the (red) circles are (a) the noisy observations and (b) the LSTM predictions. (c) shows the next-step prediction by LSTM. The solid line ($\frac{~~~}{~~~}$) denotes the ground truth and the hollow ($\circ$) and the solid ({\color{red}$\bullet$}) circles are the noisy observation and the LSTM prediction, respectively. (d) shows the predictive probability distributions from ($\circ$) the standard CE and ({\color{red}$\bullet$}) the regularized CE.}
\caption{ (a) A 500-step forecast of the Mackey-Glass equation. The contours represent the probability distribution conditioned on the last observation; $p(\yhat_{s+n}|\yhat_s)$. The dashed lines indicate the 95\% confidence interval and the solid circles ({\color{red}$\bullet$}) are the ground truth. (b) shows the predicted STD in time. The dashed line indicates the noise level.}
\caption{ (a) and (b) show the trajectories in a delay-time coordinate. The solid line is the ground truth and the circles ({\color{red}$\bullet$}) in (a) are the noisy observations and in (b) are the LSTM predictions. }
\caption{ (a) A 500-step forecast of the Ikeda equation. The contours represent the probability distribution conditioned on the last observation; $p(\yhat_{s+n}|\yhat_s)$. The dashed lines indicate the 95\% confidence interval and the solid circles ({\color{red}$\bullet$}) are the ground truth. (b) The predicted standard deviation in time. The dashed line indicates the noise level.}
\caption{\textit{Source graph}. Configuration of the Alice source. {\red to be completed }}
\caption{The distribution of far-UV escape fractions and stellar masses predicted by the \bluetides\simulation at$z=8$. The shading denotes the logarithmic density of galaxies. The solid line shows the median escape fraction in bins of $\log_{10}(M_{*}/M_{\odot})$. The inner and outer shaded regions show the range of containing the central $68.3\%$ and $95.4\%$ of objects respectively. Observational constraints at $z>6$ from \citet{Willott2015}, \citet{Schaerer2015}, \citet{Watson2015}, and \citet{Laporte2017} are also shown alongside observational constraints at $z=1.2-4.0$ from \citet{Pannella2015}. The object HFLS3 \citep{Riechers2013, Cooray2014}, which has $\log_{10}(f_{\rm esc, FUV})\approx -4$, is omitted as it falls far below other observations. Observed stellar masses are converted to assume a \citet{Chabrier2003} initial mass function. The right-hand axes shows the corresponding far-UV attenuation in magnitudes. Tabulated values of $f_{\rm esc, FUV}$ in bins of stellar mass are presented in Table \ref{tab:L_fesc}.}
\caption{Intrinsic/total (solid dark line), obscured (solid light line), and unobscured (dashed line) star formation rate distribution functions (bottom panels) and cumulative star formation rate densities (top panels) predicted by \bluetides\at$z\in\{8,9,10\}$.}
\caption{\textcolor{yellow}{Expected performance outcomes for different system configurations}}
\caption{To be written \red{To be written} \red{To be written}To be written \red{To be written} \red{To be written}To be written \red{To be written} \red{To be written}To be written \red{To be written} \red{To be written}}
\caption{\red{To be written. QCD-plot: remove $a_1$, $n$, and $a_0$; add labels; replace QCD with PDG}}
\caption{Phase shift $\Delta \phi \, (t)-\langle\Delta \phi \, (t)\rangle$ as a function of time, where $\langle \cdot \rangle$ denotes time average, for $f=21.6$ Hz. \textcolor{blue}{- -}: $I=2A$, \textcolor{black}{-}: $I=2.12$ A, \textcolor{red}{-.}: $I=2.5$ A. \label{fig:time_signal}}
\caption{ The function {\color{green4} ${-}\log\sig({-}X)$} used in\linebreak inference and the $\bie$ operator, versus {\color{blue} $X$}. }
\caption{On the left a) \grays counts map above 10~GeV in the $8^{\circ} \times 8^{\circ}$ region around Westerlund 2. The 3FGL catalog sources are labeled with a black cross and blue circle for identified and unassociated sources, respectively. The white dashed line shows the extended emission related to FGES~J1036.4--5834. On the right b) Residual map after subtracting all the identified catalog source and the diffuse background, centred on the best-fit position of the extended source to the south of Westerlund 2. Also shown are the best-fit 2D disk template (white circle), the position and extension of FGES~J1023--5746 (black circle), and the known SNRs and bright pulsars in the field of view. The extension of the H.E.S.S. source HESS~J1026--582 is also marked with a black circle. The white contours are obtained from the TS-map, for levels 3, 4 and 5. }
\caption{Left panel a): Gas column densities derived from CO data, integrated in the velocity range of [-11, 21] km/s. Middle panel b): \ion{H}{ii} column density derived from the Planck free-free maps. Right panel c): Total gas column density by summing up the molecular gas and \ion{H}{ii} gas. The \gray emission is overlaid with white contours together with the positions of the energetic pulsars in the field and the position of Westerlund 2 (black circle). The white disk indicate the \gray template, in which the total gas mass is calculated.}
\caption{{\it Left panel}: The white circle and black arcs label the regions to subtract the \gray emissivities. The position of Westerlund 2 is marked as cyan diamond. {\it Right panel}: The profile of \gray emissivities per H atom above 1~GeV with respect to the distance to Westerlund 2. For comparison, we show $1/r$ (solid curve), and constant (dotted curve) profiles, which are predicted by the continuous injection, and the impulsive injection, respectively. }
\caption{Three traps create three rings of magnetic nanoparticles. The rings interact with one another (see \textcolor{blue}{Visualization 3}). [From Masajada \textit{et al.}, Opt. Lett. \textbf{38}, 3910 (2013)].}
\caption{Three traps create three rings of magnetic nanoparticles. The rings interact with one another (see \textcolor{blue}{Visualization 3}). [From Masajada \textit{et al.}, Opt. Lett. \textbf{38}, 3910 (2013)].}
\caption{ (Color online) X-band (9.4~GHz) magnetic resonance spectra at various temperatures, mostly in the magnetically ordered region ($T_{\mathrm N}=107$~K), for the external field along the particular in-plane direction (100). Open squares and circles indicate the resonance fields of additional lines at fields below the main line, see also corresponding symbols in \textcolor{blue}{Fig. \ref{FigOrdXQ}}. Stars indicate the spin-flop field as determined from magnetization data \cite{kliemt15a,kliemt17a}. }
\caption{ (Color online) Temperature dependence of resonance field $H_{\rm res}$ and linewidth $\Delta H$ for the external field along two different in-plane directions and two microwave frequencies as indicated. Solid lines guide the eyes. Open squares and circles indicate $H_{\rm res}$ of additional lines as shown in \textcolor{blue}{Fig. \ref{FigSpecs100pH}}. }
\caption{Determining the neighbors of the \textcolor{Lavender}{upper left point} by looking at its \textcolor{Apricot}{neighbors} in the grid.}
\caption{Target distance from Horn 1 \cite{DUMMY:4}. The proton beam enters the target from the left in this diagram (\textcolor{red}{red arrow}).}
\caption{({\textbf{a}}) $m=\widetilde{m}(t)$ for $d_1 = 500\mu$m; ({\textbf{b}}) $h=\widetilde{h}(t)$ for $d_1 = 500\mu$m; ({\textbf{c}}) Experimental curves for $d_2 = 1$mm of $m=\widetilde{m}(t, H_R)$ for $H_R = 20.0 \pm 2.0\%$ and $H_R = 40.0 \pm 2.0\%$. ({\textbf{d}}) Experimental curves of $h=\widetilde{h}(t, H_R)$ for two different $H_R$ at $d_2=1$mm. ({\textbf{e}}) $h=\widetilde{h}(t)$ for $d_1$ and $d_2$ for different $\theta_t$. ({\textbf{f}}) $h=\widetilde{h}(\alpha)$ for $d_1$ and $d_2$ where $h$ increases with $\alpha$. ({\textbf{g}}) $h=\widetilde{h}(B_o)$ for $d_1$ and $d_2$. ({\textbf{h}}) Experimentally measured PSZ saturation, $\Phi_{Expt}=\widetilde{\Phi}(B_o)$, for $d_1$ and $d_2$. ({\textbf{Legend}}): All performed at $H_R = 20\%$, $\theta_t=90^{\circ}$, $d_1$ (hollow points): ({\color{blue}{$\star$}}) $\alpha=0$; ({\color{yellow}{$\bigtriangleup$}}) $\alpha=0.25$; ({\color{red}{$\Box$}}) $\alpha=0.50$; ({\color{black}{$\diamondsuit$}}) $\alpha=0.75$; ({\color{green}{$\circ$}}) $\alpha=1$. All performed at $H_R = 20\%$, $\theta_t=90^{\circ}$, $d_2$ (filled points): ($\openbigstar$) $\alpha=0$; ($\protect\markersix$) $\alpha=0.25$; ($\protect\markerseven$) $\alpha=0.50$; ($\protect\markereight$) $\alpha=0.75$; ($\protect\markerone$) $\alpha=1$. All performed at $\alpha=1$ and $H_R = 20\%$: ({\color{magenta}{$\davidsstar$}}) $d_1$, $\theta_t=75^{\circ}$; ({\color{cyan}{$\bigtriangledown$}}) $d_1$, $\theta_t=45^{\circ}$; ($\protect\markertwo$) $d_2$, $\theta_t=75^{\circ}$; ($\protect\markerthree$) $d_2$, $\theta_t=45^{\circ}$. All performed at $H_R = 40\%$ and $d_2$: ($\protect\markerfive$) $\alpha=0$; ({\color{green}{$\varhexstar$}}) $\alpha=1$. }
\caption{The PSNR(dB) results of different methods on Set12 dataset with noise levels 15, 25 35, 50 and 75. The best two results are highlighted in \textcolor[rgb]{1.00,0.00,0.00}{red} and \textcolor[rgb]{0.00,0.00,1.00}{blue} colors, respectively}
\caption{\wm\task: (a) CT-GRU (blue) and GRU (orange) response to probe on sequences like$\{\sym{l}/0, \symx/0, \ldots, \symx/t \}$ for a range of $t$. %storage-retrieval lags, $t$. (b) Storage timescales, $\log_{10}(\btau_k^\sto )$, and event-detection weights, $\bW^\evt$. The CT-GRU modulates storage time scale of symbol based on the context. \label{fig:denis}}
\caption{Event sequences for (a) \cluster, (b) \hawkes, and (c) \reddit. Time is on horizontal axis. Color denotes event label; in (a), irrelevant labels are rendered as dashed black lines. }
\caption{In column a) a \ac{YSO} model at \SI{140}{\pc}, b) a star-forming region at \SI{10}{\kilo\pc} and c) the center of a galaxy at \SI{500}{\kilo\pc}. Top: Ideal single band synthetic observation at about \SI{8}{\microns} (a, b) and at about \SI{500}{\microns} (c) extracted directly from the radiative transfer output. Bottom: a,b) Realistic synthetic three-color images including field stars in \ac{GLIMPSE} colors and c) \emph{Herschel} three-color image. The maximum value of the stretch for the fluxes in \ac{IRAC} \SI{8}{\microns}, \ac{IRAC} \SI{4.5}{\microns}, \ac{IRAC} \SI{3.6}{\microns} (\acs{RGB}) are \SI[mode=text]{3.0e-14}{ergs.s^{-1}.cm^{-2}} (red), \SI[mode=text]{4.2e-14}{ergs.s^{-1}.cm^{-2}} (green) and \SI[mode=text]{1.3e-14}{ergs.s^{-1}.cm^{-2}} (blue) for Figure~\ref{C2:Fig:caseplot}\textcolor{lmugreen}{(a)} and \SI[mode=text]{1.2e-13}{ergs.s^{-1}.cm^{-2}} (red), \SI[mode=text]{2.0e-13}{ergs.s^{-1}.cm^{-2}} (green) and \SI[mode=text]{8.0e-14}{ergs.s^{-1}.cm^{-2}} (blue) for Figure~\ref{C2:Fig:caseplot}\textcolor{lmugreen}{(b)}. The maximum value of the stretch for the fluxes in \ac{SPIRE} \SI{500}{\microns}, \ac{PACS} \SI{160}{\microns}, \ac{PACS} \SI{70}{\microns} (\acs{RGB}) in for Figure~\ref{C2:Fig:caseplot}\textcolor{lmugreen}{(c)} are\SI[mode=text]{3.0e-14}{ergs.s^{-1}.cm^{-2}} (red), \SI[mode=text]{2.8e-13}{ergs.s^{-1}.cm^{-2}} (green) and \SI[mode=text]{9.3e-15}{ergs.s^{-1}.cm^{-2}} (blue).}
\caption{The stellar-mass evolution of MW-mass (solid blue line boxes) and M31-mass (dashed red line boxes) galaxy progenitors, including counts for each redshift bin. The data points show the stellar masses of all sources in ZFOURGE over $z=0.2-2.5$. The red curve shows the 80 per cent stellar mass completeness limit for star forming and passive galaxies in ZFOURGE ({\color{blue}Spitler et al.} {\color{blue}2017, in prep}).}
\caption{Additional \aastex\symbols}
\caption{Contour map of the QCE and P3 for the SRC and DRC. {\color{Revise}{The simulation parameters were $\delta_1=0.03$, $\delta_2=0.02$, and $\alpha=0.04/W$ for the results presented in Fig. (a)-(d).}} Panel (a) shows the QCE versus R1 and R2. Panel (b) shows a plot of P3 versus P1 and P2, {\color{Revise}{where $R_1$=93\% and $R_2$=85\%.}} {\color{Revise2}{Panels (c) and (d) show the optimized reflectivity of the input mirror versus the input powers for the SRC and DRC, respectively.}}}
\caption{The results for SRC-SFG and DRC-SFG. (a) P3 and QCE were varied by changing P1 (P2 was fixed at 90 mW) for the SRC, {\color{Revise2}{where $\delta_1$=0.039; $\delta_2$= 0.02; and $\alpha=0.05$.}} (b) P3 versus $P_1^{cav}$ and $P_2^{cav}$ (P2 was fixed at 90 mW). (c): P3 and QCE were varied with P1 (P2 was fixed at 161 mW) for the DRC, where $\delta_1=0.0347$, $\delta_2=0.0295$, and $\alpha=0.05$. (d) The corresponding plot for P3 versus $P_1^{cav}$ and $P_2^{cav}$ for the DRC (P2 was fixed at 161 mW). }
\caption{\textcolor{red}{This is basically an illustration of the method, similar as figure \ref{fig:VGHVG}, so we need to produce a good figure (and only one)}}
\caption{Snapshots of simulations at $\Pr = \tau = 0.1$, and $R_0^{-1} = 2.25$ (top row) and $R_0^{-1} = 4.5$ (bottom row), in simulations with $L_x = L_y = L_z= 100d$. Note that $R_L^{-1} \simeq 1.5$ and $R_c^{-1} = 5.5$ at these parameters. Left: Volume rendering of {\color{red} $|\nabla \tilde{C}|$}, the magnitude of the gradient of the compositional {\color{red} perturbations}. Right: Normalized horizontally averaged {\color{red} total} temperature, {\color{red} total} composition, {\color{red} total} density, and rms vertical velocity profiles taken at the same times as the snapshots. }
\caption{Dependence of the emitted power from the seeding power and the injection current. Note that the maximum output power of 125\,mW can be reached already for 7\,mW seed power. The color scale of the emitted power is in units of milli watt.}
\caption{Plot of the measured output power vs the current and the temperature. The data was taken for an injection beam power of approximately 3\,mW. The color scale is in units of milli watt.}
\caption{Ramping the current in constant current (CC) mode. a) The output power of the AR-coated slave diode for 3\,mW of seeding power at 20$^{\circ}$C. The power is ramped up (blue) and down (red). Note that amplification occurs at any current and power. b) The output power of a slave based on an FP-diode (Sharp GH0781JA2C) for \textcolor{blue}{1-3}\,mW of seeding power. Note that the amplification works only for narrow ranges of currents and temperatures.}
\caption{Power of the amplified light in constant power mode (a) Ramping the temperature by 0.6\,K/min. (b) Applying temperature shocks of up to 1.5\,K/min. The left axis of the graph (blue color) is the emitted power, and the right axis (purple color) is the corresponding case temperature. The stability is better than$50\,\mu \text{W}_\text{RMS}$. In the absence of large temperature fluctuations the laser is stable to $20\,\mu \text{W}_\text{RMS}$ or $2*10^{-4}$}
\caption{Lane detection score on No-VP set (\textcolor{red}{\textbf{Red}}: Best).}
\caption{Simultaneous detection and classification F1 score for lane classes (\textcolor{red}{\textbf{Red}}: Best).}
\caption{Simultaneous detection and classification recall score for road marking classes (\textcolor{red}{\textbf{Red}}: Best).}
\caption{Schematic illustration of the mechanism for exposing monitoring data to users of the \gui. In this example, two users (Operators~1 and~2) request data from a combination of sources. The data transmitted to the users are first stored in a \redis database, which is integrated with the \python Web server. \redis is filled by directly acceding lower-level elements of the system. In this example, these are two telescopes (\lst-1 and \mst-5), and the monitoring database, an \oes component.}
\caption{Pairs of images are compared to each other in feature space and similarity is measured using cosine similarity \textsc{top} and L2 distance \textsc{bottom}. Pairs of images with same identity shown in {\color{blue}blue}, and pairs with different identities shown in {\color{red}red} (10 each). Columns show left to right: sweeping of azimuth, elevation, lighting azimuth, and lighting elevation with all other parameters held. Dashed vertical lines show range of transformation values seen at training time. Ideally cosine similarity would be 1 everywhere for the blue lines, indicating perfect transformation invariance. For dissimilar faces, the red curves would be less than 1. We see that invariance to lighting is easy, even beyond the range of training examples (see central box in Figure \ref{fig:faces_panel}). Elevation is particularly hard, so two features of the same person begin to differ at large elevations.}
\caption{\small Typical Examples. In each row, we show an example of our motion-triggered method selecting visual observations. The gray block represents non-triggered frames, and red block represents triggered frames. Each block stands for half second. A few triggered (red boxes) and non-triggered (regular boxes) frames are visualized. At the end of each example, We show the trigger ratio and the correctly predicted intention. More results are shown in the \highlight{Sec.1 of the} supplementary material. }
\caption{Comparison between the measured LNA impedance and lumped circuit model (simulated in FEKO). We have verified that the small residual difference between the measured and model impedance has an insignificant \rednew{(< 1\%)} impact on the resulting beam model \rednew{at 216\,MHz}.}
\caption{Stokes I, Q, U and V images (a,b,c,d respectively) obtained from 2 minutes observation started at 13:14:48 UTC on 2014-03-06. \red{The images were beam corrected in 1.28\,MHz coarse channels and averaged in the 200-212\,MHz band. Only part of the band was used to avoid radio-frequency interference that affected the upper part of the band (most likely due to digitial TV) which caused subtle artefacts in the Q, U and V images. The images obtained with\textsc{wsclean} were beam corrected using the FEE model. The false Stokes leakages are within $\pm$5\% in image centres and get a bit higher closer to the edges. The $2^{nd}$ order polynomial surfaces fitted to false Stokes Q, U and V leakages are shown in Fig.~\ref{fig:2d_leakages} and leakages averaged in declination bins are show in Figures~\ref{fig:stokes_q_leakage_vs_dec_1078140304},~\ref{fig:stokes_u_leakage_vs_dec_1078140304}~and~\ref{fig:stokes_v_leakage_vs_dec_1078140304}.}}
\caption{False Stokes Q, U and V leakages (a,b,c respectively) surfaces obtained from fit of $2^{nd}$ order polynomial to leakages of the brightest sources from the Q, U and V images in Fig.~\ref{fig:iquv_images} \rednew{(the colour scale is the same for all three images)}.}
\caption{\red{Comparison of Stokes Q leakage measured in images at 200--231\,MHz calibrated with the three different models. The leakage data from individual sources were averaged in 5$\degree$ bins (100-200 sources per bin). The analytic model performs the worst of all three models. The FEE models performs better at positive declinations and converges to AEE model at negative declinations (both are within measurements errors on this relatively small data sample).}}
\caption{\red{Comparison of Stokes U leakage measured in images at 200--231\,MHz calibrated with the three different models. The leakage data from individual sources were averaged in 5$\degree$ bins (100-200 sources per bin). All three models have similar values of the U leakage $\approx 5$\%. Note that the errors of false Stokes U for the analytical model are significantly larger than for the FEE and AEE models (Fig.~\ref{fig:stokes_u_leakage_vs_dec_1078140304}) because the hour angle dependence of false Stokes U was reduced significantly for the non-analytic models.}}
\caption{\red{Comparison of Stokes V leakage measured in images at 200--231\,MHz calibrated with the three different models. The leakage data from individual sources were averaged in 5$\degree$ bins (100-200 sources per bin). All the models have values within the errors bars, but AEE model has V leakage $\approx$1-2\% consistent across the band, which is of similar magnitude to that reported by \citet{offringa-2016} below 200\,MHz.}}
\caption{\red{False Stokes Q leakages calculated from GLEAM images corrected with the originally used AEE model (upper image) and the new FEE model (lower image). The small scattered data points represent false Stokes Q calculated for all the individual sources (around 85000 in total in both cases) identified in the images of the three fields at declinations $\delta= -13\degree, 1.6\degree$ and $+18.6\degree$ (with approximately 40100, 29000 and 16000 sources respectively). The large data points with error bars were calculate as the mean and standard deviation in 3$\degree$ bins.}}
\caption{\red{The data points are the ratio $\widetilde{B}_{2,2} / \widetilde{B}_{1,1}$ based on the GLEAM data from bottom row (frequency ranges 200-208\,MHz, 208-216\,MHz, 216-223\,MHz and 223-231\,MHz) of Fig.4 in\citep{gleam_nhw}. The data were collected with the beam pointing at local meridian at declination $\approx 2 \degree$. The GLEAM data were binned in $3 \degree$ bins in declination (with about 1000-1200 sources averaged in the central bins down to 100-200 in the bins near the image edges). The solid lines represent the same ratio predicted when the AEE beam model is used to correct unpolarised sky brightness propagated through the FEE beam model (assumed to represent the ``true'' MWA tile beam). The simulated curves were normalised by values in the image centre in order to replicate the normalisation of flux to The Molonglo Reference Catalogue (MRC) catalogue \citep{1981MNRAS.194..693L} performed on the GLEAM data.}}
\caption{\label{fig:AC} Fourier spectrum of the time evolution of the limit cycle velocity in the doublet-singlet mode for $W=\frac{8}{3}D$ as functions of $F^{\rm D}$. (a) Amplitude of the first three essential Fourier modes ($f/f_1=1,2,3$) normalized by $A_0\equiv \tilde{v}\left( 0\right)$. (b) The whole cycle, $T_{\rm whole}$, the residence time in the triplet state, $T_{\rm tri}$, and that in the doublet-singlet state, $T_{\rm ds}$, in the unit of $t_0$. The dashed lines stand for $F^{\rm D}_{\rm C1}=24F^{\rm D}_0$, the critical value of $F^{\rm D}$ where the crossover between $A_1$ and $A_3$ is observed and $T_{\rm whole}$ shows the minimum as well. The dotted line denotes $F^{\rm D}_{\rm C3}=27.32F^{\rm D}_0$, the critical value of the transition of the dynamical modes. The plots in the right panel show the magnified images at approximately $F^{\rm D}_{\rm C3}$ in the $F^{\rm D}$-axis. The dash-dotted lines in the magnified plots denote $F^{\rm D}_{\rm C2}=27.2F^{\rm D}_0$ above which $A_3$ starts to decrease. } \end{figure} To clarify the behavior around the dynamical mode transitions observed in our system, we performed a spectrum analysis on the time evolution of the velocity of a tagged particle by using the following equation: \begin{align} \tilde{v} \left( f \right) = \int_{-\infty}^{\infty}e^{i2\pi f t}v\left( t\right)dt,\label{eq:fourier} \end{align} where $v\left(t\right)$ is the velocity of the tagged particle and $f$ is the frequency. Because we confirmed that the altering of $W$ does not yield any essential change in the qualitative behavior of the system, in the following, we analyze the dynamics of the system with $W=8D/3$ as a typical case. In Fig.~\ref{fig:fourier}, we show the absolute value of the power spectra obtained by eq.(\ref{eq:fourier}) for $F^{\rm D}=20F_0^{\rm D}$. The spectra of essential frequencies can be recognized as the fundamental frequency $f_1$ and its harmonics. Hereafter, we refer to the spectrum at $f=0$ as the zeroth spectrum, and to the other essential spectra as the first, second, third $\cdots$ in increasing order of frequency. The zeroth spectrum reflects the average velocity. The first essential spectrum comes from the whole cycle with three peaks, while the third spectrum reflects the cycle between neighboring triplet peaks, such as (b) and (d) in Fig.~\ref{fig:sche_p3}. Because the characteristic three-peaked shape of the velocity time evolution can be almost reproduced by using the three spectra of $f_1$, $f_2$ and $f_3$, these spectra can be regarded as the essential spectra. It is difficult to give a clear physical interpretation for each peak since the three-peaked pattern depends on all three spectra, but we can say that the third spectrum at least reflects the difference in the magnitude of the spectrum between the average velocity and the fastest velocity of the triplet cluster. We show the amplitudes of the first three spectra, $A_i\equiv \left|\tilde v(i \cdot f_1)\right| \left( i \in {1, 2, 3}\right)$, as functions of $F^{\rm D}$ in Fig.~\ref{fig:AC}(a). In Fig.~\ref{fig:AC}(b) we also show $T_{\rm tri}$, the total residence time in the triplet state during one whole cycle~((a)-($\text{a}^\prime$) in Fig.~\ref{fig:sche_p3}), and $T_{\rm ds}$, the one in the doublet-singlet state. The period of one whole cycle $T_{\rm whole}=T_{\rm ds} +T_{\rm tri}$ is also plotted. In Fig.~\ref{fig:AC}(a), the amplitudes of $A_i$ are normalized by $A_0\equiv \tilde v(0)$. As $F^{\rm D}$ increases, $A_1$ decreases and $A_3$ increases. Although $A_3$ is smaller than $A_1$ when $F^{\rm D}$ is smaller than a threshold value $24F^{\rm D}_0$, above the threshold value, $A_3$ exceeds $A_1$. That the third spectrum is greater than the first reflects the fact that the average velocity is higher than the minimum value of the doublet velocity ((c) or (e) in Fig.~\ref{fig:sche_p3}). This crossover occurs because the contribution of the triplet state becomes more dominant as $T_{\rm tri}$ becomes larger. Interestingly, at around $F^{\rm D} = 24F^{\rm D}_0$, $T_{\rm whole}$ also exhibits a shallow minimum. The non-monotonicity in $T_{\rm whole}$ can be understood by considering the contributions from $T_{\rm ds}$ and $T_{\rm tri}$ separately. If the average velocity of the particles becomes faster, the time needed by the doublet to catch up to the singlet in front becomes shorter, which results in a smaller $T_{\rm ds}$~(see Fig.~\ref{fig:AC}~(b)). At the same time, as $F^{\rm D}$ becomes larger, the triplet cluster becomes more stable, and $T_{\rm tri}$ becomes longer. These two effects are in a trade-off relation, and the latter dominates the system when $F^{\rm D}\ge 24F^{\rm D}_0$. This value $F^{\rm D}_{\rm C1} = 24F^{\rm D}_0$ is referred to as the first critical value, and the system has two more critical values. The second critical value is $F^{\rm D}_{\rm C2} = 27.2F^{\rm D}_0$, where $A_3$ shows a maximum, and above $F^{\rm D}_{\rm C2}$, the amplitude of the third spectrum, $A_3$, decreases abruptly. The existence of this second critical force can be explained as follows. As $F^{\rm D}$ approaches $F^{\rm D}_{\rm C2}$, the residence time in the triplet state becomes longer very sensitively to the change in $F^{\rm D}$, although the velocity does not change greatly. This leads to a dominant contribution of $A_3$ to the zeroth peak amplitude $A_0$ which reflects the average velocity. In other words, the average velocity approaches the triplet velocity. Because $A_3$ reflects the difference in the average velocity and the fastest velocity of the triplet state, the magnitude of the spectra decreases upon increasing $F^{\rm D}$ above $F_{\rm C2}^{\rm D}$. The third critical force is $F^{\rm D}_{\rm C3} = 27.32F^{\rm D}_0$, where the dynamical mode transition between the doublet-singlet mode and the triplet mode occurs. As shown in Fig.~\ref{fig:AC}(b), the dynamical mode transition can be understood as a continuous divergence of the residence time in the triplet state $T_{\rm tri}$ ($T_{\rm ds}$ changes monotonously and almost linearly decreases with $F^{\rm D}-F^{\rm D}_{\rm C3}$). Therefore, we characterize the divergence in analogy with the second order phase transition by using the following equation: \begin{align} T_{\rm tri} = T_{\rm tri}^{\rm C}\left( \frac{F^{\rm D} - F^{\rm D}_{\rm C3}}{ F^{\rm D}_{\rm C3}}\right)^{-\beta},\label{eq:fit} \end{align} where $T_{\rm tri}^{\rm C}$ and $\beta$ are constants and are found to be $T_{\rm tri}^{\rm C}\approx 17.9$, $\beta \approx 0.15$ by a fitting of Eq.~(\ref{eq:fit}) using the least squares method to the residence time $T_{\rm tri}$. The transition of the dynamical modes can also be understood from the view point of the stability of the triplet state. When the triplet state breaks up and the doublet leaves the singlet behind, the doublet has a positive relative acceleration to the singlet. Therefore, we consider the relative acceleration of the doublet to the singlet. The equations of motion for three individual particles are as follows: \begin{align} M_{\rm p} {\boldsymbol{a}_1} &= \bF^{\rm D}+\bF_1^{\rm H}+\bF_{12}^{\rm P},\label{eq:a1}\\ M_{\rm p} {\boldsymbol{a}_2} &= \bF^{\rm D}+\bF_2^{\rm H}+\bF_{21}^{\rm P}+\bF_{23}^{\rm P},\label{eq:a2}\\ M_{\rm p} {\boldsymbol{a}_3} &= \bF^{\rm D}+\bF_3^{\rm H}+\bF_{32}^{\rm P}\label{eq:a3}, \end{align} where the particle indexes $i$=1,2,3 are assigned to the particles from the front to the rear in order, $\boldsymbol{a}_i$ stands for the acceleration of particle $i$, and $\bF_{ij}^{\rm P}$ represents the particle-particle direct interaction force exerted on particle $i$ by particle $j$. Now, we would like to consider the motion of the front doublet (composed of particles 1 and 2) relative to the trailing singlet (particle 3). As the motion of the front doublet, we can simply consider the motion of the center of mass of the two particles in the front (particles 1 and 2). Since the action reaction relation $\bF_{12}^{\rm P}+\bF_{21}^{\rm P}=\boldsymbol{0}$ holds, the summation of both sides of Eqs.~(\ref{eq:a1}) and (\ref{eq:a2}) leads to the following equation: \begin{align} M_{\rm p}{\boldsymbol{a}_{\rm d}} &= \bF^{\rm D}+\frac{\bF_{23}^{\rm P}}{2}+\frac{\bF_1^{\rm H}+\bF_2^{\rm H}}{2}\label{eq:a4}, \end{align} where $\boldsymbol{a}_{\rm d}=\left(\boldsymbol{a}_1+\boldsymbol{a}_2\right)/2$ stands for the acceleration of the doublet. Then, the acceleration of the doublet relative to the singlet can be expressed by using Eqs.~(\ref{eq:a3}) and (\ref{eq:a4}), as follows: \begin{align} M_{\rm p}\left({\boldsymbol{a}_{\rm d}-\boldsymbol{a}_{\rm s}} \right)&= \frac{\bF_{23}^{\rm P}}{2} - \bF_{32}^{\rm P} +\frac{\bF_1^{\rm H}+\bF_2^{\rm H}}{2} -\bF_3^{\rm H},\label{eq:a5} \end{align} where we rename the acceleration of the singlet $\boldsymbol{a}_3$ as $\boldsymbol{a}_{\rm s}$. As mentioned before, even when the triplet state is formed, the three particles do not always have the exactly same velocity. However, there is a time when the velocity of the leading doublet (or the average velocity of the front two particles) and that of the rear end particle coincide. We will refer to this moment as the marginal point. The stability of the triplet state can be quantified by the relative acceleration at this marginal point. For a measure for the relative acceleration, now we introduce the following quantity at the marginal point: \begin{align} {\cal F} = \left( \frac{\bF_{23}^{\rm P}}{2} - \bF_{32}^{\rm P} +\frac{\bF_1^{\rm H}+\bF_2^{\rm H}}{2} -\bF_3^{\rm H} \right)\cdot \boldsymbol{e}_{\rm y}. \end{align} This is the $y$-component of the right-hand side of Eq.~(\ref{eq:a5}) at the marginal point. In Fig.~\ref{fig:EF}, ${\cal F}$ is shown as a function of $F_{\rm D}$. When ${\cal F}$ is positive, the doublet has a positive acceleration relative to the singlet, which leads to the breakage of the triplet state, {\it i. e.}, the doublet-singlet mode. As ${\cal F}$ is positive but becomes decreases, the residence time in the triplet state becomes longer. On the other hand, when ${\cal F}$ is negative, the singlet will further approach the doublet, and eventually, all three particles start to move at the exactly same velocity, which results in the stable triplet state. In agreement with the result of the spectrum analysis, the change in the sign of ${\cal F}$ occurs at $F_{\rm C3}^{\rm D}=27.32F^{\rm D}$. \begin{figure} \begin{center} \includegraphics[width=\linewidth]{/EF} \caption{\label{fig:EF} The evaluation function ${\cal F}$ as a function of $F^{\rm D}$. The values of ${\cal F}$ are normalized by ${\cal F}_0={\cal F}\left(27F_0^{\rm D}\right)$. }
\caption{Frequency dependence of (a) Re$(\sigma_{eff})$, and (b) Re$(\epsilon_{eff})$, obtained from THz-TDS measurements of $l$-CNT film at 10 K, 323 K, 423 K and 533 K, along with their associated Drude fits (\full). (c) Temperature dependence of the scattering rate $\gamma$ (\opentriangledown) obtained from Drude fits of measured spectra Re$(\sigma_{eff}(\omega))$ and Re$(\epsilon_{eff}(\omega))$ in the range 0.3--2 THz for $l$-CNT film. The data has been fitted in the range 300--500 K with a linear function $\gamma_{e-ph}=\alpha T/d$ (\dashed), where d = 1.6 nm is the mean diameter of our CNTs, and $\alpha=49$ m/Ks is a fitting constant. (d) Temperature dependence of the value $\gamma-\gamma_{e-ph}$ calculated from the data in (c).}
\caption{ (a) Scanning electron microscope image of the fabricated device, with a close-up of the beamsplitter region. (b) Simulation of the \blue{z-component of the electric field (absolute value) inside the multi-mode interferometer when injecting the fundamental TM mode into waveguide H.} }
\caption{ (a) Calculated and (b) measured joint spectral intensity of the SPDC source, evidencing the emission of nearly frequency-uncorrelated photons. \blue{The value of the pump incidence angle $\theta \simeq 1.4^{\circ}$ can be deduced from the measured signal and idler central wavelengths \cite{Caillet09,Orieux13}.} }
\caption{ROC curves for two random forest classifiers for the image data set. Each random forest was trained on 10K images and tested on 10K images. The shaded regions are approximate 95\% confidence bands \gray{with} and \green{without} Poisson bootstrapping the test data set. The inset shows the full ROC curve. Throughout this paper the x-axis of the ROC curves is on a log-scale and the logs are taken base-10.}
\caption{Confidence bands for random forests trained on \red{1M images} and \green{200K images}. Also shown are two standard deviation confidence intervals (black) for log FPR and TPR at several points along the curve. The inset shows the full ROC curves. }
\caption{Confidence bands for random forests with \red{256 trees} and \green{128 trees}. Also shown are two standard deviation confidence intervals (black) for log FPR and TPR at several points along the curve. The inset shows the full ROC curves.}
\caption{Confidence bands for random forests with \red{256 trees} and \green{512 trees}. Also shown are two standard deviation confidence intervals (black) for log FPR and TPR at several points along the curve. The inset shows the full ROC curves. }
\caption{Confidence bands for random forests whose trees have \red{maximum depth 20} and \green{maximum depth 10}. Also shown are two standard deviation confidence intervals (black) for log FPR and TPR at several points along the curve. The inset shows the full ROC curves. }
\caption{Confidence bands for random forests trained \red{with bootstrapping} and \green{without bootstrapping} of the training data set. Also shown are two standard deviation confidence intervals (black) for log FPR and TPR at several points along the curve. The inset shows the full ROC curves. }
\caption{Confidence bands for random forests trained by \red{considering 64 features} and \green{considering 256 features} when splitting each node. Also shown are two standard deviation confidence intervals (black) for log FPR and TPR at several points along the curve. The inset shows the full ROC curves. }
\caption{Trajectory 1. Comparison of the initial ({\color{blue}---}) and final (- - -) error signals on the two joint angles $\theta_1(t)$ and $\theta_2(t)$.}
\caption{Trajectory 2. Comparison of the initial ({\color{red}---}) and final (- - -) error signals on the two joint angles $\theta_1(t)$ and $\theta_2(t)$.}
\caption{Trajectory 1. Comparison of the initial input $u_0(t) = y_d/G_0$ (- - -) and the final learned input $u_{20}(t)$ ({\color{blue}---}).}
\caption{Trajectory 2. Comparison of the initial input $u_0(t) = y_d/G_0$ (- - -) and the final learned input $u_{20}(t)$ ({\color{red}---}).}
\caption{{\bf The faithful embedding of a causal set.} An unlabeled causal set (left) with $N=200$ spacetime elements, indicated by the green points, and $5373$ causal relations, indicated by the gray lines, is faithfully embedded into the blue region (right). The causal set is bounded below by a null boundary and above by a constant-time hypersurface, with a timelike boundary of constant radius separating the two. \blue{This particular region demonstrates how in practice we can encounter causal sets with a non-trivial combination of boundaries.} A general embedding algorithm for a given causal set is unknown, but it may be possible to extract information from the causal set structure about the types of hypersurfaces which form the bounding region.}
\caption{{\bf The spacetime representation for the causal interval.} The representation of the causal interval in the $(1+1)$-dimensional Minkowski spacetime is shown in the left panel, where the orange vertical curves correspond to the expected paths of chains and the blue horizontal curves correspond to the expected paths of antichains. The orange and blue straight lines crossing the center, \blue{denoted the representation-induced origin,} respectively represent the maximum chain and antichain. In the center panel, the empirical chain lengths (orange) fall nearly perfectly across the expected values (green), given by~(\ref{eq:geo_to_chain}). The radial coordinates are inferred by averaging over values sampled from the marginal distribution~(\ref{eq:marginal_x}). Fluctuations increase with radial distance due to finite-size effects. The right panel shows the antichain widths, i.e., cardinalities, for the same causal sets, ranked by a time coordinate inferred from the intersection of each antichain $a_i$ with the maximum chain. All data is averaged over ten graphs with height $L=1$ and size $N=2^{14}$, and the shaded regions indicate the standard deviation of the mean.}
\caption{{\bf Measurement of timelike boundary volume.} The causal set is partitioned into antichains, each of whose elements lie at a constant graph distance to the minimal elements $\mathcal{P}$. In each antichain, elements near the timelike boundary have the fewest number of relations (left). Those with a number of relations in the range $k\in[k_{min},k_{min}+\epsilon)$ are selected as candidates (red). The center panel shows the resulting set of candidates $\mathcal{T}$ on top of the antichain partitions, where each partition's elements are the same shade of green. Maximal chains $\{\mathcal{B}_i\}$, which are proxies for timelike geodesics, are then constructed by maximizing the number of elements $\mathcal{T}$ in each chain, shown by the bold black lines (right). \blue{The origin is always taken to be at the center of the region to enable a natural extension to higher dimensions.}}
\caption{{\bfseries Visualizing ECIS Data:} Resistance measurements as recorded by an ECIS\textsuperscript\textregistered device. (a) A single response for each of three different cell lines, each exposed to two different AC frequencies. (b) Response of H1299 cell line to all nine AC frequencies.}
\caption{ \textbf{Experimental setup and spatial properties of the solid-state magnet assembly.} \textbf{(a)} Cross-sectional view of a ultra-high vacuum chamber housing a surface-electrode trap (indicated at the centre), used for spatial manipulation of single atoms. Two sets of rare-earth, ring magnets generate a magnetic (quantisation) field $\mathbf{B}_0$ along their symmetry axis $\hat{z}$ (indicated by \textbf{\color{myred} - -}) . In addition, three individual pairs of magnetic field (shim) coils are mounted on corresponding mechanical support structure, marked with (\textbf{\color{myorange}--}). The shim coils enable fine tuning of $\mathbf{B}_0$ along longitudinal and orthogonal (vertical and horizontal) directions. Preparation and detection laser beams ({\color{myblue} $\mathbf{\rightarrow}$}) enter the chamber along $\mathbf{B}_0$. A home-build biquad antenna (sketched in the top left corner) is used to apply microwaves around $2\pi\,\times\,1,600\,$MHz for internal state manipulation of the atom. \textbf{(b)} The magnetic-field variation of the solid-state magnets close to their geometrical centre (inset shows larger region) along $\hat{z}$, calculated using Eq.\,\ref{eq:mag}. From numerical calculations, considering all directions, we infer a diameter of spherical volume $d_{\text{dsv}} \simeq 150\,$\textmu m, where $\Delta B_0/B_0 \leq 1\,\times\,10^{-6}$. }
\caption{Disagreement between polygon-shaped (in {\color{blue}blue} colour) and rectangle-shaped (in {\color{red}red} colour) groundtruth regions. It is found out that evaluation results using the polygon-shaped groundtruth provides a more accurate representation of algorithm's performance.}
\caption{\label{tab:plane_synth_data_training} Instance Detection performance (mAP) of Faster-RCNN~\cite{faster_rcnn} re-trained on synthetic and evaluated on real data for two different feature extractor networks (\blue{InceptionResnet}~\cite{inception_resnet} and \red{Resnet101}~\cite{resnet101}). Training on synthetic data and evaluating on real data results in low detection performance in case of the \ptgrey camera and the total absence of any reasonable detection result in case of the \asus camera.}
\caption{\label{fig:incremental} Finetuning the feature extractor after 400K, 900K and 1200K steps where the pre-trained feature extractor was frozen for the \ptgrey and the \asus cameras. We show results for the InceptionResnet~\cite{inception_resnet} and Resnet101~\cite{resnet101} architectures.}
\caption{\label{tab:main_exp} Outcomes of all our experiments. We give numbers for \blue{InceptionResnet}~\cite{inception_resnet} / \red{Resnet101}~\cite{resnet101}. Except for the experiments with real data (last column), all experiments were performed on synthetic data only. We emphasized the best results trained on synthetic data.}
\caption{\label{fig:influence} Influences of the different building blocks for synthetic rendering for the \ptgrey and the \asus cameras. Results were obtained with InceptionResnet~\cite{inception_resnet} as the feature extractor. Blurring is clearly a useful yet simple operation to apply to the synthetic images to improve the results.}
\caption{\label{fig:rfcn} Results using RFCN~\cite{rfcn} on the \ptgrey dataset. Freezing the feature extractor boosts performance significantly on this method as well. We observe the same results if we train Mask-RCNN on the \asus dataset.}
\caption{(Color online) Characteristics of NVs in nanodiamond. \textbf{(a)} Atomic force microscope image of nanodiamonds deposited on a mica plate. \textbf{(b)} Histogram of nanodiamond sizes that are predominantly within 25-61nm ($43\pm18$ nm). \textbf{(c)} Typical ODMR measurement of a nanodiamond NV center with an applied magnetic field $B_{\parallel}=$ 37.8G (\protect\scalebox{0.6}{\color{Red}$\square$}) and 66G ({\color{Cyan}$\circ$}). \textbf{(d)} The time evolution of the spin state population $\mbox{P}_{\ket{0}}$ in spin echo experiment for three typical NVs. By fitting the data with a function of $(1/2)[1+\exp(-(T/T_{SE})^{\alpha})]$, we estimate the parameters $[T_{SE},\alpha]$ as follows: $[ 2.142\pm 0.018\mu s, 1.448 \pm 0.026]$ (NV1, {\color{Orange}$\circ$}), $[ 4.292\pm 0.133\mu s, 1.47\pm 0.10]$ (NV2, \protect\scalebox{0.6}{\color{Cyan}$\square$}), $[ 2.990 \pm 0.083\mu s, 1.576 \pm 0.101]$ (NV3, \protect\scalebox{0.8}{\color{Red}$\triangledown$}). }
\caption{(Color online) \textbf{(a)} The extended coherence times as achieved by CCDD scheme (\protect\scalebox{0.6}{\color{Red}$\square$}) and XY8 pulse dynamical decoupling ({\color{Blue}$\circ$}) as a function of the effective Rabi frequency. The values in panel are the number of XY8 cycles (up to 12) that achieve the longest possible coherence time. \textbf{(b)} The estimated sensitivity for the measurement of an oscillating field using CCDD (\protect\scalebox{0.6}{\color{Red}$\square$}) and XY8 ({\color{Blue}$\circ$}) scheme. The achievable Rabi frequency is $\Omega=8.5$MHz, the coherence time is assumed as $T_2=15.22 \mu s$ (XY8) and $T_c=31.22 \mu s$ (CCDD). The oscillating field strength is $\gamma b=(2\pi)100$kHz. The other experiment parameters are the same as Fig.\ref{fig:XY8_CCD}.}
\caption{\label{fig.scheme} Schematic representation of the numerical setup. (a) A chain of $J$ $3$-dimensional oscillators observed through the measurement functions $h_j(\mb{x})$, which can either return $y_j$ (gray disk), $x_j$ (blue square) or $z_j$ (red triangle). (b) For the benchmark of each network $h_i=h_j,~\forall i,\,j$, and the recorded variable is the one that provides best observability of the dynamics. For the cases of mixed measurements the experimental design explores both the position and number of lower-ranked variables in terms of observability. Chain networks of $J=5$ R\"ossler and Cord systems are considered.}
\caption{\label{fig.rossler-coherent-1z}(a-e) Impact of a single $z$ measurement of an oscillator -- the other four being measurements of the $y$ variable of the other four oscillators -- in the synchronization analysis of a chain of $J=5$ R\"ossler oscillators in the spiral chaotic regime. The vertical gray lines are as in Fig.~\ref{fig.rossler-coherent-5y}. The red star indicates missing features. Roughly, the oscillator observed through $z$ becomes ``invisible" to the svM-SSA. (f-j) A similar situation is found when the measurement of $z$ is removed from the analysis, but with some improvement in resolution enhancement [see vertical arrows, only showed in (a) and (f)]. (k-o) The use of $\log_{10} z$, instead of $z$, fully recovers the overall features of the benchmark (see Fig.~\ref{fig.rossler-coherent-5y}b).}
\caption{\label{fig.rossler-coherent-1z}(a-e) Impact of a single $z$ measurement of an oscillator -- the other four being measurements of the $y$ variable of the other four oscillators -- in the synchronization analysis of a chain of $J=5$ R\"ossler oscillators in the spiral chaotic regime. The vertical gray lines are as in Fig.~\ref{fig.rossler-coherent-5y}. The red star indicates missing features. Roughly, the oscillator observed through $z$ becomes ``invisible" to the svM-SSA. (f-j) A similar situation is observed when the measurement of $z$ is removed from the analysis. (k-o) The use of $\log_{10} z$, instead of $z$, fully recovers the overall features of the benchmark (see Fig.~\ref{fig.rossler-coherent-5y}b).}
\caption{\label{fig.cord-image-plot}Phase synchronization in a chain of $J=5$ chaotic cord oscillators. (a)~Mean observed frequencies computed through phase estimates taken with the Poincar\'e section$xxx$ \textcolor{red}{????????}. (b)~Singular values from the svM-SSA, computed with the $x$ time series (the benchmark). The analysis through the leading $\lambda_1^*$ shows that mixed measurements with (c)~$z$ and (d)~$y$ have similar detrimental effects. Dropping the $z$-measured oscillator from the analysis provides results that are similar to those computed with mixed measurements with (e)~one and (f)~two $z$ variables [the selected cases correspond to the left-shifted symbolic sets in (c)].}
\caption{Measured current (\textcolor[rgb]{0.5,0,0.5}{+}) finite element current (\textcolor{red}{$\circ$}) for $V=0.3$\,V. The solid lines are the heavy hole, light hole, and total current for the 1d theory fitted to the currents calculated by the finite element method.}
\caption{Measured current (\textcolor[rgb]{0.5,0,0.5}{+})and the second iteration finite element current (\textcolor{red}{$\circ$}) for $V=0.3$\,V. The solid lines are the heavy hole, light hole, and total current for the 1d theory fitted to the currents calculated by the finite element method.}
\caption{Additional \aastex\symbols}
\caption{\Fermi\\gray\SED of Kes\,73,%\Fermi\ \gray\ spectral energy distribution of source~A, fitted with various models (see \S\ref{subsec:origin}). Systematic errors (see \S\ref{subsec:spec}) are indicated by black bars, and the statistical errors are indicated by red bars. In the upper panel, the long-dashed and short-dashed lines indicate the IC and p-p interaction components of Model I, respectively. In the bottom panel, the long-dashed line represents the magnetar emission in both Model III and Model IV; the short-dashed blue and black lines indicate the p-p interaction component in Model III and the IC component in Model IV, respectively.}
\caption{Figures (a) and (b) show the performance of the \textit{original} (\protect\purple), \textit{optimized original} (\protect\green), \textit{lazy} (\protect\blue) \textit{packed lazy} (\protect\orange), \textit{implicit} (\protect\yellow) and \textit{packed implicit} (\protect\dblue) layouts.}
\caption{Figures (a), (b) and (c) show the performance of the \textit{implicit} (\protect\purple) and the \textit{optimized original} tiered vector (\protect\green) for different tree widths.}
\caption{Figures (a),(b) and (c) show the performance of the \textit{implicit} (\protect\purple) and the \textit{optimized original} tiered vector (\protect\green) for different tree heights.}
\caption{Figures (a) and (b) show the performance of the \textit{base} (\protect\purple), \textit{rotated} (\protect\green), \textit{non-aligned sizes} (\protect\blue), \textit{non-templated} (\protect\orange) layouts.}
\caption{ Functions $\Psi_\eta(x)$ and $\Psi_\kappa(x)/x$ for neutrons (\full) and protons (\dashed) from the exact solutions of the system (\ref{eq:psi_eq}) for the Av18+TBF microscopic model and $n_{\rm B}=0.3$~fm$^{-3}$. Thick red and thin blue lines correspond to thermal conductivity and shear viscosity problems, respectively. % For the thermal conductivity the function $\Psi_\kappa (x)/x$ is plotted. }
\caption{\label{fig3} (color online) (a), (b), (c), (d): Time-averaged mean-squared displacements (TAMSD), $ \delta^2(\Delta,T)$ and their ensemble average, $ \langle \delta^2(\Delta,T) \rangle $ ({\color{red} $ \filledmedtriangleup $}) shown for two sets of five-angle arches unclogged with vibration amplitudes $ \Gamma = 8.4 $ (left) and $ \Gamma = 2.8 $ (right). The ensembles are contrasted for an intermediate and a long total averaging time, $ T = \num{5e3} $ (a), (b) and $ T = \num{4.5e4} $ (c), (d). The ensembles include all arches that have survived until T or longer. The black lines show a linear slope $ \sim \Delta $, the expected slope for a subdiffusive CTRW with a power law waiting time distribution. (e), (f): Scaling of $ \langle \delta^2(\Delta,T) \rangle $ for both sets of arches. The different symbols show $ \langle \delta^2(\Delta,T) \rangle $ evaluated for fixed values of the lag time $ \Delta = 24 $ ({\color{color1} $ \ast $}), $ \Delta = 44 $ ({\color{color2} $ \bullet $}), $ \Delta = 81 $ ({\color{color3} $ \plus $}), $ \Delta = 149 $ ({\color{color4} $ \filledmedtriangleup $}), $ \Delta = 272 $ ({\color{color5} $ \medtriangleright $}), $ \Delta = 496 $ ({\color{color6} $ \filledmedtriangleright $}), $ \Delta = 904 $ ({\color{color7} $ \medtriangledown $}), $ \Delta = 1649 $ ({\color{color8} $ \filledmedtriangledown $}), $ \Delta = 3007 $ ({\color{color9} $ \medtriangleleft $}), $ \Delta = 5,484 $ ({\color{color10} $ \filledmedtriangleleft $}). The black line indicates a T dependence $ \sim T^{-0.7 = -1 + 0.3} $.}
\caption{\label{fig4}(color online) Distributions $ \phi(\xi) $ of the normalized TAMSD $ \xi = \frac{\delta^2}{\langle\delta^2\rangle} $ for $ \Gamma = 8.4 $ $ (a) $ and for $ \Gamma = 2.8 $ $(b)$. The shapes of the distributions were found to depend on the ratio $ \Delta/T $, and are shown here for $ \Delta/T = 0.1 $ and different $ T $ values given in the legend, which are the same for both plots. The averaged $ \phi(\xi) $, black solid line, is used to compute the EB at $\Delta/T=0.1$. Inset to (b): EB parameter, $ \langle \xi^2 \rangle - \langle \xi \rangle^2 $ for amplitudes $ \Gamma = 2.8 $ ({\color{red} $ \bullet $}) and $ \Gamma = 8.4 $ ({\color{blue} $ \filledmedtriangleup $}) as a function of $ \Delta/T $, computed using averaged $\phi(\xi)$ over a range of $\Delta/T$~\cite{SI}. The EB parameter is roughly constant in the range $ \Delta/T = 0.1\textendash0.4 $. The dashed black line indicates the expected EB value for a power law $\psi(t)$ with $ \alpha = 0.3 $. }
\caption{Offline (\orange{orange}) and online (\blue{blue}) workflows.}
\caption{\label{fig:Tails_pdf_log_log} The PDF of tracer displacements in log-scale. Symbols represent the data from \cite{Leptos2009}. Solid lines correspond to the simulations. The dashed lines correspond to the power laws: \symbol{\dashed}{}{10}{0}{black}{black}: $x^{-4}$. \symbol{\dashed}{}{10}{0}{magenta}{black}: $\textcolor{magenta}{x^{-3}}$. }
\caption{\label{fig:Deff_beta_0_5_kbts} Effective diffusion coefficient of tracers for the squirmer model with $\beta = 0.5$. \symbol{\solid}{\bigcircle}{22}{0}{black}{black}: simulations. \symbol{}{\ssquareb}{0}{0}{black}{blue}: data from \cite{Leptos2009}. }
\caption{\label{fig:Tracers_disp_time_dep} PDF of tracer displacements at $\Delta t = 0.12s$ for the time-dependent and steady models with $\phi_v = 2.2\%$. \symbol{}{\losange}{0}{0}{black}{black}: data from \cite{Leptos2009}. \symbol{\solid}{}{0}{0}{blue}{black}: $\bar{\beta} = 0.5$, time-dependent. \symbol{\dashed}{}{0}{0}{blue}{black}: $\beta = 0.5$, steady. \symbol{\dashdot}{}{0}{0}{magenta}{black}: $\bar{\beta} = 0.12$, time-dependent. \color{magenta}{\tdot \tdot \tdot \tdot \tdot \tdot} \color{black}{: $\beta = 0.12$, steady.} }
\caption{Map of the 6.7\,GHz methanol masers toward NIRS3. Triangles and circles represent maser spots before and after the burst, respectively. The velocity-integrated emission of the 6.7\,GHz masers (gray scale) and the JVLA 5\,GHz radio continuum emission(red contours) are also shown.}
\caption{NEOWISE multi-color ({\color{blue} W1},{\color{green} W2}, {\color{yellow} W3}, {\color{red} W4}) light curve of G107, folded by the 34.6\,d period, and normalized to minimum brightness. Zero represents the date of the maser flare peak. Large crosses mark IRAC flux ratios, normalized to the low state. The Gaussian fit to the methanol maser light curve (\cite{fujisawa_periodic_2014}) is shown for comparison. The apparent jumps at phase of $\sim$4\,d are due to long-term brightening.}
\caption{\emph{HST} COS and STIS spectra of SN\,2017egm after normalization of the continuum using Savitsky-Golay convolution. Absorption lines originating from the Milky Way Galaxy are marked as green, while lines from the host system are marked as red (host component) or blue (companion component) and jointly labeled with purple text. Dotted lines and lighter text colours indicate notable nondetections or marginal detections. The resolution in the mid-UV (bottom-most panel) is not sufficient to resolve the two velocity components in the host system.\label{fig:fullspec}}
\caption{Profiles of prominent absorption lines from various species associated with the host-galaxy system of SN 2017egm (see also Figure~\ref{fig:fullspec}). All lines except for the sub-DLA (top plot) show a bimodal profile, with the red component at a redshift matching that of NGC\,3191 and the bluer component (-235\,km\,s$^{-1}$) at the exact redshift of the companion galaxy SDSSJ101857.98+462714.6. Red and blue curves show modeled intrinsic profiles and the green curves are the combined results after convolution with the COS instrumental line-spread-function (LSF). %In reality, most of the lines are likely to contain narrower saturated sub-components that prevent an exact abundance determination. \label{fig:linefit}}
\caption{\red{The monopole of the galaxy bispectrum in the moderately squeezed configuration \eqref{msc}, at redshifts $z=0.5,1.0,1.5$. The dashed curve is the Newtonian approximation, the solid curve is the result when the {\em only} the relativistic dynamical corrections from \eqref{e15}--\eqref{e19} are included, i.e, only the $\alpha_I$ terms in \eqref{e26}.} }
\caption{\red{{\em Upper:} The monopole of the galaxy bispectrum in the moderately squeezed configuration \eqref{msc}, at redshifts $z=0.5,1.0,1.5$. The dashed curve is the Newtonian approximation, the solid blue curve is the monopole of Paper II~\cite{Jolicoeur:2017nyt} [i.e. only the $\Gamma_I$ terms in \eqref{e26}], and the solid red curve is the full result from \eqref{e26r}, i.e. including relativistic dynamical corrections. \\ {\em Lower:} Percentage difference, at the redshifts $z=0.5,1.0,1.5$, from dynamical corrections to the Paper II monopole.}}
\caption{ Redshift evolution at $k = 0.01\;\mathrm{Mpc}^{-1}$ ($\approx$ the equality scale): the galaxy bispectrum monopole for the Newtonian approximation (dashed), Paper II (blue) and this paper (red) {\em (left)}; the percentage difference relative to the Newtonian approximation for Paper II and this paper {\em (middle)}; \red{the percentage difference between this paper and Paper II {\em (right)}}.}
\caption{Evolution of an obscured AGN in the high-resolution simulation of the A1A0 merger. {\em Left panels:} SDSS $ugz$ images, generated from {SUNRISE} simulations, show the disturbed merger morphology during the late stages of the merger. In the first snapshot, the BH separation has just fallen below 10 kpc, and the final snapshot occurs just after the BH merger. The $ugz$ filter combination is chosen to enhance contrast between the dust-obscured nuclei and star-forming regions. {\em Right panels:} from top to bottom, the evolution of the bolometric AGN luminosity $L_{\rm AGN}$, line-of-sight gas column density $N_{\rm H}$, and \wise\\wonetwo\color throughout the merger is shown. The vertical line denotes the time of BH merger. In the$L_{\rm bol}$ and $N_{\rm H}$ plots, the blue and magenta curves correspond to each BH prior to merger, and the blue curve shows the post-merger evolution. In the $N_{\rm H}$ and \wonetwo\plots, the error bars show the range of values over all viewing angles.$N_{\rm H}$ is calculated along the line of sight to each BH, in an aperture 64 pc in size (consistent with the effective spatial resolution of the high-resolution simulations). \wise\colors are calculated for the entire galaxy. For clarity, error bars are plotted for only a subset of snapshots. The dashed and dotted lines in the\wonetwo\plot denote single-color cuts of 0.5 and 0.8, used in the literature and in this work. Note that higher time resolution is used for the{SUNRISE} calculations in the late phases of the merger. The AGN luminosity and column density peak during the galaxies' coalescence, when the galaxies are morphologically disturbed; this luminous, obscured AGN phase is closely traced by red \wonetwo\colors.\label{fig:evol_with_images}}
\caption{Optical and IR SEDs are shown for the high-resolution A1A0 simulation snapshots corresponding to the optical images in Figure \ref{fig:evol_with_images}. Green solid lines give the fiducial simulation SED, and black dotted lines give corresponding SED in the AGNx0 simulation. The gray shaded regions denote the four \wise\bands (from left to right:$W1$, $W2$, $W3$, \&$W4$). In the first two snapshots, the system is still dominated by star formation, and prominent PAH and 9.7 $\mu$m silicate absorption features are apparent. In snapshots ``c" and ``d" we see the reddening of the mid-IR colors with increasing AGN contribution, particularly in the last snapshot. The \wonetwo\colors in each snapshot are 0.1 (``a"), 0.2 (``b"), 0.8 (``c"), and 1.3 (``d").\label{fig:seds}}
\caption{In the top panel, the simulated \wise\\wonetwo\color is shown versus the AGN luminosity for all eight mergers in the simulation suite. In the bottom panel, the AGN luminosity is shown as a fraction of the total bolometric luminosity in the host ($f_{\rm AGN} = L_{\rm AGN}/L_{\rm tot}$). Each point represents a single snapshot, where the \wise\color is averaged over all sight lines. The green upward triangles, red squares, and orange downward triangles denote snapshots in the Early, Late, and Post-Merger stages, respectively. The gray dotted and dot-dashed lines denote the single-color cuts considered in this work (\wonetwo\$> 0.5$ \& 0.8, respectively). The global\wise\color is strongly correlated with the AGN contribution to the total host SED, with the reddest\wonetwo\colors produced predominantly by luminous AGN in the Late and Post-Merger stages.\label{fig:w1w2_flbol_scatter}}
\caption{For each simulation, and each merger phase, the ``contamination" fraction ($t_{\rm WISE,AGNx0}/t_{\rm WISE}$) of mid-IR colors by star formation in the AGNx0 simulations is shown. Specifically, this is the amount of time in each AGNx0 simulation for which \wise\colors\wonetwo\$>0.5$ are produced solely by star-formation heating, relative to the total \wise\\wonetwo\$>0.5$ lifetime in the corresponding fiducial simulation. All quantities are averaged over 7 viewing angles. The left panel shows results by simulation number as defined in Table \ref{table:merger_models}, and the middle and right panels show the contamination versus the maximum sSFR and SFR, respectively. 3/8 simulations have a nonzero lifetime with red \wise\colors from star formation, occurring only during intense starbursts and constituting at most 15-25\% of the Late merger phase. When these starburst-induced red \wise\phases occur in the AGNx0 simulation, the corresponding fiducial simulation{\em always} has a simultaneously-active AGN, so this is not ``true" contamination of the \wise\AGN selection.\label{fig:sf_contam}}
\caption{The evolution of merging galaxies in \wise\color-color space (\wonetwo\versus\wtwothree) is shown for the gas-rich merger simulations (\# 0-5). The thick color curves show the evolution for the fiducial simulations, where the color scale denotes the bolometric AGN luminosity as shown in the color bar. The thin black curves show the evolution for the corresponding AGNx0 simulations (in which the AGN luminosity is artificially set to zero for the radiative transfer calculation). The cyan circles denote the initial snapshot, the orange triangles denote the time of BH merger, and the red squares denote the final snapshot at the end of the simulation. The gray dotted and dot-dashed lines show single-color\wonetwo\selection cuts used in the literature and in this work. The gray dashed lines denote the two-color ``wedge" selection criteria of J11, and the blue solid lines denote the two-color cut motivated by this work (Equation\ref{eqn:mycut}, \S\\ref{ssec:reliability}). The \wonetwo\color typically peaks near the time of BH merger, and in the pure starburst (AGNx0) simulations it rarely exceeds$0.5$. The red \wtwothree\colors produced in purely starburst systems are also clearly distinguishable from the AGN in color-color space.\label{fig:color-color}}
\caption{The median mid-IR AGN selection completeness -- that is, the fraction of AGN that would be selected with a given mid-IR color cut -- is shown for all eight mergers in the simulation suite. The left side of the plots show selection completeness versus minimum \wonetwo\for a single-color cut. The right side of the plots show the same quantities for the two-color cuts (\wonetwo\and\wtwothree) of J11 and in this work (see \S\\ref{ssec:reliability} for details). The blue and cyan circles denote the median \wise\AGN fraction for the whole simulation, and the red and orange stars denote the median fraction for the Late+Post merger phases only. Error bars denote the inter-quartile range. For moderate-luminosity AGN ($L_{\rm bol} > 10^{44}$ erg s\inv), the completeness decreases steadily with stricter \wonetwo\color cuts. The two-color cut proposed in this work (\S\\ref{ssec:reliability}) has a median completeness of 100\% for moderate-to-high luminosity, merger-triggered AGN. \label{fig:wiseagn_frac_allsim}}
\caption{For the six gas-rich merger simulations, the mid-IR dual AGN lifetime is shown versus projected separation, as a fraction of the total time in each separation bin (i.e., the dual AGN duty cycle). The leftmost ``negative" separation bin denotes the post-merger phase, capped at 100 Myr. Here, a ``dual AGN" is defined when each BH has a bolometric luminosity $> 10^{44}$ erg s\inv, and the system is considered a ``Dual MIR AGN" if the \wonetwo\$>0.5$ criterion is simultaneously met. The thick red histogram denotes the \wise-selected dual AGN lifetime, while the dotted black histogram denotes \wise-selected AGN where only a {\em single} BH is active ($L_{\rm bol} > 10^{44}$ erg s\inv). The thin blue histogram denotes the phases where a dual AGN is present that would {\em not} be selected via \wise\\wonetwo\$>0.5$ colors; such a phase appears in only a single bin in the top-left plot. Thus, \wise\color selection is extremely efficient at identifying dual AGN hosts, particularly at small separations where many such duals may still be unresolved.\label{fig:wise_dual_projsep_w05}}
\caption{The mid-IR selection completeness for {dual} AGN (pairs of simultaneously-active AGN) is shown in a similar manner as the single-AGN completeness in Figure \ref{fig:wiseagn_frac_allsim}. Specifically, the \wise-selected dual AGN lifetime is plotted as a fraction of the total dual AGN lifetime. Blue circles and red stars denote the median fraction for the total merger and the Late phase, respectively, versus the \wonetwo\single-color cut assumed. The cyan circles and orange stars similarly show$t_{\rm WISE,Dual}/t_{\rm Dual}$ for the two-color cuts defined in J11 and in this work. Dual AGN are defined as systems where each BH exceeds the minimum luminosity threshold $L_{\rm AGN} > 10^{44}$ erg s\inv. Mid-IR selection is even more effective at identifying dual AGN than single AGN, with a median completeness of $\sim$ 100\% for all the selection criteria considered in this work. \label{fig:wise_dual_frac}}
\caption{The fraction of the {\em total} mid-IR lifetime for which a dual AGN is present ($t_{\rm WISE,Dual}/t_{\rm WISE}$) is shown for various mid-IR selection criteria. Blue (cyan) circles denote this fraction for the total merger (for single- \& two-color\wise\cuts, respectively), and red (orange) stars similarly denote the fraction for the Late+Post merger phase. Note that the dual AGN lifetime is zero by definition in the Post-merger phase. Although dual mid-IR AGN are efficiently triggered advanced mergers (Figure\ref{fig:wise_dual_projsep_w05}), the Late merger phase is intrinsically short. The fraction of \wise\AGN containing duals varies greatly between mergers, particularly in the Late phase, and the short lifetime of the Late phase yields a somewhat higher probability of finding dual AGN at larger separations. Nonetheless, we find that most merger-triggered mid-IR AGN, and$\sim$ 30-40\% of mid-IR AGN in advanced mergers, are actually dual AGN with $L_{\rm AGN} > 10^{44}$ erg s\inv. \label{fig:wise_dual_to_wise_allsim}}
\caption{The total gas column density (\nh) along the AGN line of sight is shown versus merger stage; this provides an upper limit on the amount of galactic-scale obscuration. Prior to the BH merger, the merger evolution is binned by projected BH separation, decreasing from right to left (following the convention used in the literature and throughout this work). After the BH merger, the evolution is binned by the time elapsed since the BH merger, increasing from right to left (such that time evolution can be followed from right to left across both pre- and post-merger panels). \nh\is shown for AGN with bolometric$L_{\rm AGN} > 10^{44}$ erg s\inv\(cyan), for mid-IR AGN selected using the \wise\two-color cut proposed in this work (black), and for dual mid-IR AGN selected with the same color cut (red). The column density in each bin is calculated as the time-weighted median over all viewing angles, for all eight fiducial simulations; the error bars denote the median absolute deviation. The AGN column density peaks at$\ga 10^{23}$ - $10^{24}$ cm s$^{-2}$ in the late stages of the merger, at BH separations $< 3$ kpc and in the first 40 Myr post-merger. This demonstrates that AGN are preferentially obscured in advanced mergers, and that environmentally-driven obscuration can be significant during mergers even on galactic scales. \label{fig:column_dens}}
\caption{\colorbox{gray!30}{Optimized} cleanup.}
\caption{(Color online) (a) Size distribution of the active components under different dynamic processes: bond percolation, site percolation and the Watts threshold model. Symbols are the results of $10^{7}$ Monte-Carlo simulations, and curves show the exact probability given by our approach~\redcopy{ Eq.\,\eqref{Eq:sizedistribution}}. (b) Exact mean size, \redcopy{Eq.\,\eqref{Eq:averageSize}}, of the final active component for bond and site percolations as a function of the occupation probability $p$. The graph has $N=20$ and 54 directed edges.}
\caption{(Color online) Individual response function and activation probabilities. The colors of the nodes are used to show their activation probability, summed over all possible outcomes. \redcopy{Square nodes can make a spontaneous transition [with response function $F_s(p,q)$, see Eq.~\eqref{eq:examplersponse_spont}], while round nodes need active precursors to make a transition [they have the response function $F_w(p,\tau)$, see Eq.~\eqref{eq:examplersponse_thresh}].} The parameters of the spontaneously activating nodes are: $p=0.4$ and $q=0.3$ for nodes 1, 5 and 11; and $p=0.6$ and $q=0.1$ for nodes 0, 6 and 9. The parameters of the threshold nodes are: $p=0.6$ and $\tau=2$ for node 2, 4, 8 and 12; and $p=0.7$ and $\tau=1$, for nodes 3, 7 and 10.}
\caption{\textit{Main panel}: RGB color image of NGC\,232:\ha~intensity in red, V-band intensity in green and \oiii$\lambda5007$ intensity in blue. White contours trace the \ha~velocity map in steps of 50 \kms, with the solid contours displaying receding velocities and the dashed countours the approaching velocities. The black solid line indicates the location of zero velocity. The bottom-left inset shows the co-added spectra (within a resolution element of $\sim 0$\farcs67) for three different positions located at (i) the nucleus (red), (ii) the disk (blue), and (iii) the elongated emission-line structure (orange). Each location is marked with an arrow in the main panel. These spectra cover the wavelength range of \hb~to \oiii$\lambda5007$~ with the flux shown in units of 10$^{-16}$ \flux. The upper-right inset shows the \oiii~intensity map in logarithmic scale using colors and zoomed in at the location of the jet structure. On-top of that image is superimposed the contour map of the 1.4\,GHz radio emission obtained by the VLA, with the contour levels being 0.35, 2.1, 6.7 and 15.4 mJy/beam intensities, respectively. The white straight lines indicate the location of the semi-major (North-East to South-West) and semi-minor (North-West to South-East) axis of the galaxy. The white dashed-line shows the mean position angle$\mathrm{PA_{[OIII]}}$ of the \oiii\structure.\textit{Upper right panel} shows the velocity field of \oiii$\lambda$5007 color-coded in \kms. The spatial extension of the jet-like structure is shown with a black solid contour line. Its location is defined by the position where the \oiii~ intensity reaches $\sim 0.04 \times 10^{-16}$ \flux. The white and black crosses show the photometric (V-band) and kinematic centers, respectively. \textit{Bottom right panel} shows the { 2D} position-velocity diagram of NGC\,232 relative to its kinematic center. The semi-transparent grey squares represent the\ha~velocity, with the blue (red) squares tracing the maximum (minimum) envelope of these velocities. The blue squares trace the rotation curve of the galaxy. Orange squares represent the \oiii\position-velocity diagram for the jet-like structure. It comprises those spaxels located within the black contour shown in the upper-right panel.}
\caption{Color-maps showing the spatial distribution of the line ratios \nii/\ha~(left panel) and \oiii/\hb~(following panel) for all spaxels with a S/N $>2$ in the intensity of each involved emission line. In both panels, the white contour refers to the line ratio threshold of log(\oiii/\hb) = 0.3, which is used to define our jet-like structure. Boxes of sizes $1 \farcs 8 \times 0 \farcs 6$ are drawn in the leftmost-panel along the jet axis. The position of the box centers (black dots) are defined by their respective peak in \oiii/\hb\ratio (see next panel) within each box. The photometric center is shown with a black star. The black contours in the\oiii/\hb\map represent the same radio contours as in the top-right inset of Fig.\ref{rgb}. The rightmost panel shows the classical BPT diagram together with the \citet{kauffmann} (black dashed-line) and \citet{Kewley2001} (continuous black line) demarcation curves. The dotted line represents the LINER/Seyfert transition according to \citet{Kewley2006} (with LINERs typically below this curve). In this BPT panel, the red contours represent the location of all line ratios that lie outside the white contour on both the \nii/\ha\and\oiii/\hb\panels. The outer (inner) red contour encloses 68 (34) percent of the spaxels, respectively. The red filled-circles in the BPT diagram represent the location for the co-added fluxes within each black box (see leftmost panel). The black filled-star represents the values corresponding to the box at the nucleus location. The size of each filled marker represents its projected radial distance with respect to the photometric center, being the largest ones the farthest from the nucleus. The closest (farthest) is at 1\farcs3 (6\farcs2).}
\caption{Additional \aastex\symbols}
\caption{ Comparison of AP and mAP in \% of our model and state-of-the-art methods on the PASCAL VOC 2007 dataset. The best results and second best results are highlighted in {\color{red}{red}} and {\color{blue}{blue}}, respectively. Best viewed in color.}
\caption{Comparison of our model and state-of-the-art methods on the MS-COCO dataset. The best results and second best results are highlighted in {\color{red}{red}} and {\color{blue}{blue}}, respectively. Best viewed in color.}
\caption{ Cycle-consistent adversarial adaptation of pixel-space inputs. %We train our model in two stages. By directly remapping source training data into the target domain, we remove the low-level differences between the domains, ensuring that our task model is well-conditioned on target data. We depict here the image-level GAN loss (\textcolor{green}{green}), the feature level GAN loss (\textcolor{orange}{orange}), the source and target semantic consistency losses (black), the source cycle loss (\textcolor{red}{red}), and the source task loss (\textcolor{purple}{purple}). For clarity the target cycle is omitted. }
\caption{Sound modeling using 59,306 training points and 691 test points. The intensity of the time series can be seen in (a). Train time for RBF kernel hyperparameters is in (b) and the time for inference is in (c). The standardized mean absolute error (SMAE) as a function of time for an evaluation of the marginal likelihood and all derivatives is shown in (d). Surrogate is ({\color{blue} ------}), Lanczos is ({\color{red} - - -}), Chebyshev is {(\color{magenta} --- $\diamond$ ---}), scaled eigenvalues is ({\color{green} --- + ---}), and FITC is ({\color{black} --- o ---}).}
\caption{1-dimensional perturbations for the exact RBF and Mat\'ern 1/2 kernel where the data is$1000$ equally spaced points in the interval $[0,4]$. The exact values are ($\bullet$), Lanczos is ({\color{red}-----}), Chebyshev is ({\color{green}-----}). The error bars of Lanczos and Chebyshev are $1$ standard deviation and were computed from $10$ runs with different probe vectors}
\caption{1-dimensional perturbations with the SKI (cubic) approximations of the RBF and Mat\'ern 1/2 kernel where the data is 1000 points drawn from$\mathcal{N}(0,2)$. The exact values are ($\bullet$), Lanczos with diagonal replacement is ({\color{red}-----}), Chebyshev with diagonal replacement is ({\color{green}-----}), Lanczos without diagonal replacement is ({\color{magenta}-----}), Chebyshev without diagonal replacement is ({\color{black}-----}), and scaled eigenvalues is ({\color{blue} $\times$}). Diagonal replacement makes no perceptual difference for the RBF kernel so the lines are overlapping in this case. The error bars of Lanczos and Chebyshev are $1$ standard deviation and were computed from $10$ runs with different probe vectors}
\caption{Results for training with different sizes of the training data (2\%, 5\%, 10\%, 25\%) and evaluated on the dev set. `Rand.\skill weights' is `All skill tasks' model with random weights. (\textcolor{blue}{w/} fine-tuning)}
\caption{The Four Observables Parameter Space (4OPS) plot. Top left: peak luminosity of our literature SLSN~I sample in the 400\,nm band (\bluemag{0}) versus the decline in magnitude over 30 days $\Delta\bluemag{30}$. Top right: \bluemag{0} versus color at $+30$\,d (\thirtycol). Bottom left: peak color (\peakcol) versus $\Delta\bluemag{30}$. Bottom right: \peakcol\versus\thirtycol. 99.72\% confidence bands from the Bayesian linear regression are also shown for each panel. The four plots allow the definition of a main population of SLSN I regardless of the peak luminosity. A SLSN~I will belong to the main population if it falls in the confidence interval for the blue areas in the A and D panels or, alternatively, in the orange areas of the B and C panels. 4OPS can also be used to predict missing observables for a SLSN.}
\caption{Excerpt of the figure eight scenario: radar measurements with indicated Doppler velocity from the front left (\protect\includegraphics{ext_figures/blueDot.pdf}) and front right (\protect\includegraphics{ext_figures/yellowSquare.pdf}) sensor, estimated trajectory (solid) and exemplary vehicle poses (solid rectangles), reference trajectory (dashed) and reference poses (dashed rectangles)}
\caption{Scenario with two oncoming vehicles: Estimated (solid) and ground truth (dashed) trajectories, exemplary vehicle poses (estimates: solid rectangles, ground truth: dashed rectangles), corresponding measurements with Doppler velocity (front left: \protect\includegraphics{ext_figures/blueDot.pdf}, front right: \protect\includegraphics{ext_figures/yellowSquare.pdf}, rear left: \protect\includegraphics{ext_figures/greenCircle.pdf}, rear right: \protect\includegraphics{ext_figures/redSquare.pdf}), and sensor \glspl{FOV}}
\caption{Comparison of the measurement clusters that contributed the most during update: cluster measurements (\protect\includegraphics{ext_figures/blueDot.pdf}), other measurements (\protect\includegraphics{ext_figures/blackDot.pdf}), average vehicle estimate with center of the rear axle (rectangle and cross), reference vehicle (dashed rectangle)}
\caption{Two vehicles driving closely: Estimated (solid) and ground truth (dashed) vehicle poses, corresponding measurements with Doppler velocity (front left: \protect\includegraphics{ext_figures/blueDot.pdf}, front right: \protect\includegraphics{ext_figures/yellowSquare.pdf}, rear left: \protect\includegraphics{ext_figures/greenCircle.pdf}, rear right: \protect\includegraphics{ext_figures/redSquare.pdf}), and sensor \glspl{FOV}}
\caption{Histogram of cardinality estimation errors for the multi-object scenarios: variational model (\protect\includegraphics{ext_figures/blueBar.pdf}) and direct scattering model (\protect\includegraphics{ext_figures/yellowBar.pdf})}
\caption{T-intersection scenario: Two excerpts with estimated (solid) and true (dashed) trajectories (only available in the lidar \gls{FOV}), exemplary vehicle poses (solid rectangles) and true poses (dashed rectangles), corresponding measurements (front left: \protect\includegraphics{ext_figures/blueDot.pdf}, front right: \protect\includegraphics{ext_figures/yellowSquare.pdf}, rear left: \protect\includegraphics{ext_figures/greenCircle.pdf}, rear right: \protect\includegraphics{ext_figures/redSquare.pdf}), and sensor \glspl{FOV}}
\caption{Qualitative results on GoogleRef (top) and DrivingRef (bottom)) with real speech of the text descriptions and the corresponding prediction box shown in \textcolor{red}{Red} and ground truth text and the corresponding bounding box are shown in \textcolor{green}{Green}.}
\caption[Sketch of a domain decomposite]{ An examplary decomposition of the computational domain is displayed with 13 different reference domains that are mapped to the subdomains of the decomposition, where identical colors classify the different reference domains. The different types are locally periodic cells (\tikzsquare{colp}), branching cells (\tikzsquare{colbr1} \tikzsquare{colbr2} \tikzsquare{colbr3}), double branching corner cells (\tikzsquare{colcn1} \tikzsquare{colcn2} \tikzsquare{colcn3}) and coupling cells (\tikzsquare{colcp11} \tikzsquare{colcp12} \tikzsquare{colcp21} \tikzsquare{colcp22} \tikzsquare{colcp31} \tikzsquare{colcp32}).}
\caption{The optimal shape for the compression load case and a subdomain structure with $13$ cell types is depicted (top). We show in addition a color coding of the von Mises stresses using the colorbar \protect\includegraphics[height=1ex,width=5em]{graphics/cb_stress} on the set $[\Pf > 0.5]$ (bottom).}
\caption{Performance of various algorithms for testing the uniqueness of Markov boundary: proposed Alg. 2-AF (blue `\textcolor{blue}{+}'); Alg. 2-KI (green `\textcolor{green}{$\times$}'); Alg. 3 (black `$\diamond$'); Alg. 4 (red `\textcolor{red}{$\circ$}'). The x-axis is in logarithm and starts from $200$.}
\caption{\color{blue}3D-CNN Training Algorithm}
\caption{\color{blue}Deep CRF}
\caption{\color{blue}Segmentation Accuracies on Different datasets. A `*' denotes that the best average accuracy (shown in bold) is significantly better than the accuracy achieved by the corresponding method according to a statistical paired t-test for comparing classifiers.}
\caption{\color{blue}Overall accuracies with different numbers of training samples per class for Indian Pines data.}
\caption{\color{blue}Overall accuracies with different numbers of training samples per class for Pavia University data.}
\caption{\color{blue}Overall accuracies with different numbers of training samples per class for Griffith-USGS data.}
\caption{\color{blue}Effect of Data Augmentation on Overall Accuracies (\%) for all three datasets}
\caption{\color{blue}Training and validation losses for Indian Pines data}
\caption{\color{blue}Training and validation losses for Pavia University data}
\caption{\color{blue}Training and validation losses for Griffith-USGS data}
\caption{\color{blue}Effect of spectral depth in convolution kernels}
\caption{\color[HTML]{3531FF} RUNNING TIME COMPARISON (MEASURED IN MINUTES)}
\caption{Comparison with the state-of-the-art in terms of mAP on the VOC2007 \emph{test} set. Our number is marked in \textcolor{red}{red} if it is the best in the column}
\caption{Comparison with the state-of-the-art in terms of CorLoc ($\%$) on the VOC2007 \emph{trainval} set. Our number is marked in \textcolor{red}{red} if it is the best in the column}
\caption{Comparison with the state-of-the-art in terms of mAP on the VOC2012 \emph{val} set. Our number is marked in \textcolor{red}{red} if it is the best in the column. \underline{Underline} is used if the C-WSL variant outperforms its baselines}
\caption{Comparison with the state-of-the-art in terms of \emph{CorLoc} on the VOC2012 \emph{train} set. Our number is marked in \textcolor{red}{red} if it is the best in the column. \underline{Underline} is used if the C-WSL variant outperforms its baselines}
\caption{An example control application and its precedence graph \predG. \capt{The execution starts with sensor readings -- either $\tau_1$ or $\tau_2$. After both have been received by the controller, actuation values are computed ($\tau_3$), multicast to the actuators ($m_3$), and applied ($\tau_5$ and $\tau_6$). %All tasks must be completed before the application end-to-end deadline $\app.d$. }}
\caption{Performance: cumulative return and drawdown for three typical stocks. Trading strategy based on probability predictions from ordered logit model in purple ({\color{purple}{ $\boldsymbol{ -}$ }}), separate logits in bordeaux ({\color{bordeaux}{ $\boldsymbol{-}$ }}), GARCH in red ({\color{red}{ $\boldsymbol{-}$ }}), FHS in black ({\color{black}{ $\boldsymbol{-}$ }}), and benchmark buy and hold in grey ({\color{grey}{ $\boldsymbol{-}$ }}).}
\caption{Performance: cumulative return and drawdown for three typical stocks. Trading strategy based on probability predictions from ordered logit model in purple ({\color{purple}{ $\boldsymbol{ -}$ }}), separate logits in bordeaux ({\color{bordeaux}{ $\boldsymbol{-}$ }}), GARCH in red ({\color{red}{ $\boldsymbol{-}$ }}), FHS in black ({\color{black}{ $\boldsymbol{-}$ }}), and benchmark buy and hold in grey ({\color{grey}{ $\boldsymbol{-}$ }}) color.}
\caption{$p$-values for all datasets, computed by the signed-rank test. \xhsp{We perform a one-tailed paired Wilcoxon signed-rank test, where the null-hypothesis ($\mathcal{H}_0$) is that the paired differences for the results of our PCA model and of the compared method come from a distribution with zero median, against the alternative ($\mathcal{H}_1$) that the paired differences have a non-zero median (greater than zero for Dice, sensitivity and specificity, and less than zero for surface distances)}. In addition, we use the Benjamini-Hochberg procedure to reduce the false discovery rate (FDR). We highlight, in \textcolor{green}{green}, the results where our PCA model performs statistically significantly better. The results show that our PCA model outperforms other methods on most of the measures.}
\caption{Medians (top), and means with standard deviations (bottom) for validation measures for all the methods and all the datasets. We highlight the best results in \textcolor{green}{green} based on the \textit{median} values. Among all datasets, our PCA model has the best median on Dice overlap scores and generally on surface distances. \xhsp{Exception is BEaST which achieves a lower maximum surface distances on the LPBA40 and the TBI datasets}. In addition, our model also has the best mean and variance for the Dice overlap scores and the surface distances on most of these datasets.}
\caption{\label{fig:fig5} Extensional stress normalized by the Young modulus measured in the linear regime at low strain (fig.~\ref{fig:fig2}a) as a function of the strain, for samples with different amounts of telechelic polymers, $\beta$, as indicated in the legend. For all samples, the strain rates have been adjusted such that $Wi\sim1$. The symbols are the experimental data points and the continuous line is the linear regime expectation $\sigma / E = \varepsilon$. Fracture occurs at $\sigma_c$ and $\varepsilon_c$. Inset: Variation of $\chi=\frac{\sigma_c}{E \varepsilon_c}$ with $\beta$. The dotted line indicates the linear {\color{myc} viscoelastic} regime.}
\caption{\label{fig:fig10} Correlation between the non-linear elasticity quantified by $\chi=\frac{\sigma_c}{E \varepsilon_c}$ and the departure from the parabolic shape quantified by (a) the length $\delta$, (b) the normalized length $\frac{\delta}{\lambda}$ (see text for a definition of $\lambda$). The dotted lines indicate linear {\color{myc} viscoelasticity}. Inset: normalized length as a function of $\chi-1$ in a log-log plot. The best fit yields a powerlaw with an exponent $1/3$. }
\caption{ This is a basic ablation showing the effect of adding each method one at a time. The scores for CUB birds (CUB) and CompCars (CCars) are the single class classification accuracy. PASCAL VOC uses mean average precision (mAP). The mean column is for CUB and CCars, but we show a mean of CUB, CCars and VOC as ``All''. VOC is the \emph{Post Hoc} classification results run after the fact to see how well our CUB/CCars surrogate set matches a core self-supervised benchmark test. The baseline two patch protocol uses color dropping and matches the protocol of \cite{Doersch15}. Gains in CUB/CCars appear to correlate with gains in VOC (but not perfectly). The largest gains for both CUB/CCars and VOC are from rotation with classification, chroma blurring and random aperture. Also notice that the results for CompCars is only one percentage point less than the ImageNet pretrained network. \textcolor{myblue}{\textsuperscript{\S}}The ImageNet pretrain for VOC uses conv1 through conv5. All fully connected (fc) layers are initialized new.}
\caption{These are classification mAP \cite{Krahenbuhl16}, detection mAP \cite{Girshick15} and segmentation mIU \cite{Long15} test results over PASCAL VOC \cite{Everingham10}. Mean scores are shown for classification + detection \emph{(C + D)} as well as for all three if the segmentation score is available \emph{(All)}. The bottom three results are ours and include all methods except for WV. These are: CB, YJ, TP, EPC, UBT, RA and RWC. RWC (four rotations) gives the best results, but adding in RRM yields only slightly better results. \textcolor{myblue}{\textsuperscript{\textdagger}}To conserve space, we have taken the largest of two scores when network weights have been rescaled \cite{Krahenbuhl16}. \textcolor{myblue}{\text{*}}Denotes that this is an estimate for the score based on a very recent result with a different network other than AlexNet. The estimate is computed by adding the gain reported in the work to a mutual baseline method that has an AlexNet result and also appears in our table (namely \cite{Doersch15}). \textcolor{myblue}{\textsuperscript{\textbullet}}Results were published while this paper was under review. \textcolor{myblue}{\textsuperscript{\textdaggerdbl}}Using corrected results from ArXiv paper, not ECCV.}
\caption{Demonstrating conditioned data generation with a sinusoid that has Gaussian noise added in the y-axis direction. The x-axis value is selected from the interval [1,5] and the algorithm is asked to predict the missing y-axis value. \textcolor{blue}{Blue} denotes the ground truth data and \textcolor{cadmiumorange}{orange} is generated data.\label{fig:serpent}}
\caption{Conditioned data generation being demonstrated with a simple data set of three Gaussians. The x-axis value is selected from the interval [1,5] and the algorithm is asked to predict the missing y-axis value. \textcolor{blue}{Blue} denotes the ground truth data and \textcolor{cadmiumorange}{orange} is generated data. Some stray data points remain between the Gaussians but generally the algorithm captures the shape of the distribution well.\label{fig:three_gaussians_conditioned}}
\caption{An illustrative example of how the predictions $\widehat{\by}$ (black squares) will start to get closer to a point of actual data (\textcolor{blue}{blue} dots). \label{fig:evolution}}
\caption{An illustrative example of the idea of the algorithm. The predictions $\widehat{\by}$ (black squares) are matched to the actual data (\textcolor{blue}{blue} dots). \label{fig:learning_illustration}}
\caption{Initial situation. \textcolor{blue}{Blue} denotes the training data and black denotes the generated predictions.}
\caption{The first step. We choose a random data point $\by$ (\textcolor{ao_english}{green} dot) and find the closest prediction $\widehat{\by}$ (\textcolor{red}{red} asterisk). These form a pair drawn as a black line.}
\caption{Resonance slice at the $\mu B_0=20\;keV$ plane. (a): Colorbar value equals to $\Omega/\omega_\theta$ and indicates the strength of resonance. Red lines show resonances condition being satisfied, $\Omega=0$, with $l=9,10,11$ and so on. (b): Broken KAM surface domains are indicated by the red dots, which are detected by phase vector rotation at mode amplitude $A=1.5\times10^{-4}$.\\ Four parallel {\color{blue}blue} lines from bottom to top are initial particle deposit positions Res 1, 2, 3, 4 with $\mathcal{E}'=30\;keV,45.09\;keV,51.93\;keV,60.22\;keV$, respectively. Also, trapped-loss (T-L), trapped (T), and co-passing (C-P) domains are shown. Particle orbits intersecting with the magnetic axis (A), the right wall (R) and the left wall (L) are shown with pink, purple and orange lines, respectively. The dashed black line shows particle orbit passing through the $(\psi_p=0.22,\theta=0)$ point, where the mode structure peaks.}
\caption{ \blue{ The difference in the optimization trajectory between the DLD and the Geman-McClure estimator: the face models before the registration (top), the deformed shapes after one step of the optimization by the Geman-McClure estimator (middle) and the DLD (bottom). The target point sets and the source point sets are colored blue and red, respectively. The columns of the figure show the face models rendered at different camera positions. } }
\caption{ \blue{ The performance comparison between the DLD without the pre-alignment (top) and the GMM-ASM with the pre-alignment (bottom). The number shown in each square represents the number of successes for 71 different body postures, obtained from the SCAPE dataset. Each trial was conducted in a leave-one-out manner. The position of the square represents the rigid transformation added to the 71 body postures. } }
\caption{ \blue{ The performance comparison between the DLD (top) and the Geman-McClure estimator (bottom). The number shown in each square represents the number of successes for 50 different 3D face models, extracted from the FLAME dataset. The position of the square represents the rigid transformation added to the 50 face models. } }
\caption{Model predictions for \textit{generalized} words (\textbf{top}) and triplets (\textbf{bottom}). Model (PIX, REG) and embedding types (\textit{EMB}, \textit{RND}) are as indicated in the legend on the left. The Subject's location is given (\textcolor{blue}{\textbf{blue}} box) and the model predicts the Object (\textcolor{red}{\textbf{red}}). In PIX, the intensity of the red corresponds to the predicted probability. The \textit{generalized} (unseen) words are \underline{underlined}.}
\caption{Performance on PASCAL VOC 2012 validation set. The number in {\setlength{\fboxsep}{1.5pt}\colorbox{mygray2}{gray block}} represents the performance of categories with \emph{pixel-level} annotations.}
\caption{\label{fig:comp}\textcolor{gray}({Color online) Material distribution in the system, in the case of uniform (a) and blurred (b) QD. } }
\caption{\label{fig:mag_el}\textcolor{gray}({Color online) (a) Magnetic field dependence of the $12$ lowest electron states in the uniform InAs QD. Energy $E=0$ refers to the unstrained GaAs valence band edge. Schematic pictures on the left describes spin configuration at magnetic field close to zero, while insets on the right presents the configuration at B=$12$~T. (b) Corresponding axial projection of the envelope angular momentum. The inset in the right bottom corner presents the in-plane probability density of $e_{3}$ and $e_{5}$ states at $B_{z}=0$ and $B_{z}=12$~T. } }
\caption{\label{fig:mag_p_el}\textcolor{gray}({Color online) Magnetic field dependence of the electron $p$-shell energy levels in the uniform InAs QD. } }
\caption{\label{fig:mag_h}\textcolor{gray}({Color online) (a) Magnetic field dependence of the lowest hole energy levels. Energy $E=0$ refers to the unstrained GaAs valence band edge. (b) Corresponding axial projection of the envelope angular momentum.} }
\caption{\label{fig:mag_p_h}\textcolor{gray}({Color online) Magnetic field dependence of the hole $p$-shell energy levels in the uniform InAs QD. Solid lines denote the results obtained from the effective model for a fitting (a) without SO parameter (b) with SO parameter included. } }
\caption{\label{fig:compar}\textcolor{gray}({Color online) The electron (a,b) and hole (c,d) g-factor as a function of material composition in the case of uniform and blurred QD. } }
\caption{The directions of data flow \textit{solid arrows} during the forward pass and gradient flow \textit{dotted arrows} during the backward pass of our iterative update procedure. \textit{Solid} blocks indicate that the block is frozen during that update step while \textit{dotted} block indicate that it is being updated. {\color{red} Red} denoted source information and {\color{blue} Blue} denotes target information.}
\caption{\textcolor{red}{Formula parameters}}
\caption{Sampled two and one-dimensional marginal posterior for HD~219134~b interior parameters: gas mass \menv, gas metallicity \Zenv, intrinsic luminosity \Lenv, mass of water \mice, radius of rocky interior \rsolid, core radius \rc, and mantle's relative abundances $\fesima$ and $\mgsima$. Blue dots explain the data within 1-$\sigma$ uncertainty. \reve{Dashed curves represent the prior distributions assumed.}}
\caption{Sampled two-dimensional (2-D) marginal posterior for HD~219134~c interior parameters: gas mass \menv, gas metallicity \Zenv, intrinsic luminosity \Lenv, mass of water \mice, radius of rocky interior \rsolid, core radius \rc, and mantle's relative abundances $\fesima$ and $\mgsima$. Blue dots explain the data within 1-$\sigma$ uncertainty. \reve{Dashed curves represent the prior distributions assumed.}}
\caption{Sampled one-dimensional marginal posterior for selected parameters of HD~219134~b: (a) gas radius fraction \renv, (b) water mass fraction \mice$/M$, (c) rock radius fraction \rsolid$/R$, and (d) relative core radius \rc/\rsolid. The posterior distributions depend on precision on bulk abundance constraints (light and dark green curves), and mass and radius uncertainties (light and dark purple curves). For comparison, the Earth-like solution is highlighted in red.}
\caption{Sampled one-dimensional marginal posterior for selected parameters of HD~219134~c: (a) gas radius fraction \renv, (b) water mass fraction \mice$/M$, (c) rock radius fraction \rsolid$/R$, and (d) relative core radius \rc/\rsolid. The posterior distributions depend on precision on bulk abundance constraints (light and dark green curves), and mass and radius uncertainties (light and dark purple curves). For comparison, the Earth-like solution is highlighted in red.}
\caption{\label{figure1}Crystal structure analysis of Pr$_2$Ir$_2$O$_7$/YSZ(111). (a) $\theta$--$2\theta$ XRD patterns of as-grown (bottom) and post-annealed (top) Pr-Ir-O thin film on a YSZ(111) substrate. Thin film peaks are highlighted in blue. (b) Schematic illustration of STEM sampling and observation direction. A microwedge sample was cut from the central part of the Hall bar by focused ion beam milling. (c) Cross-sectional crystal structure of the Pr$_2$Ir$_2$O$_7$/YSZ(111) interface \cite{VESTA}. Pr, Ir, and Zr atoms are drawn in blue (\textcolor[rgb]{0,0,1}{\textbullet}), red (\textcolor[rgb]{1,0,0}{\textbullet}), and green (\textcolor[rgb]{0,1,0}{\textbullet}), respectively. Oxygen is omitted. The model is slightly tilted around the $[11\overline{2}]$ direction to show the atomic arrangement in the depth direction. (d) and (e) Cross-sectional HAADF-STEM images of a single grain in the film, taken near the Pr$_2$Ir$_2$O$_7$ surface and the Pr$_2$Ir$_2$O$_7$/YSZ interface, respectively. Pr, Ir, and Zr atom positions are shown in the insets for the regions marked with the orange outlines. The colors correspond to those used in (c). Yellow (\textcolor[rgb]{1,0.75,0}{\textbullet}) represents Pr and Ir atoms that are alternately arranged in the depth direction. Crystal axes are shown in (d). The scale bars correspond to 3 nm.}
\caption{\label{figure3}Magnetotransport properties of a Pr$_2$Ir$_2$O$_7$ thin film. (a) and (b) Magnetic field ($B$) dependence of the Hall resistivity ($\rho_{{\rm xy}}(B)$), measured at 1037 mK and 9 mK, respectively. $\rho_{{\rm xy}}(B)$ is defined as $[\rho_{{\rm xy}}(B)-\rho_{{\rm xy}}(-B)]/2$ to eliminate the $\rho_{{\rm xx}}$ component. The insets show the magnified plot for the low field hysteresis part of $\rho_{{\rm xy}}(B)$. (c) Temperature dependence of the spontaneous Hall resistivity, i.e., the absolute value of $\rho_{{\rm xy}}(B)$ at $B=0$ obtained after a field cycle. Orange open diamond (\textcolor[rgb]{1,0.33,0}{$\diamond$}) indicates the temperature at which $\rho_{{\rm xx}}(0~{\rm T})$ reaches a minimum. (d) Transverse MR curves as a function of magnetic field. Measurement temperature are denoted on the right. Spin configurations labeled in (c) and (d) are indicated for Pr $4f$ moments. The magnetic field was applied along the [111] direction in all measurements. Red and blue lines in (a), (b), and (d) corresponds to up and down sweeps of the magnetic field, respectively. Illustrations of the (e) 2-in--2-out and (f) 3(1)-in--1(3)-out spin configurations for Pr $4f$ moments.}
\caption{\label{figure4}Magnetic phase diagram for a Pr$_2$Ir$_2$O$_7$ thin film as a function of temperature and magnetic field, mainly governed by Pr 4$f$ moments. The Q = (001) order, 2-in--2-out and 3(1)-in--1(3)-out states appear in areas colored in blue, green, and red, respectively. Blue open squares (\textcolor[rgb]{0,0,1}{$\square$}) denote the magnetic field values below which a hysteresis loop is observed in the vicinity of zero magnetic field in the MR curves in Fig.~\ref{figure3}(d). Red solid circles (\textcolor[rgb]{1,0,0}{\textbullet}) represent points where the MR curves in Fig.~\ref{figure3}(d) and Supplemental Fig.~S5 pass through a minimum. Orange open diamond (\textcolor[rgb]{1,0.33,0}{$\diamond$}) indicates the temperature at which the spontaneous Hall effect starts to develop. The magnetic field was applied along the [111] direction.}
\caption{The CCDFs of strength, degree, and link-weight are presented in time-aggregated networks with the activity exponent $\gamma=2.5$, where the memory exponent $\beta=0$ ({\color{red} $\square$}), 0.25 ({\color{green} $\bigtriangledown$}), 0.5 ({\color{blue} $\bigcirc$}), 0.75 ({\color{violet} $\bigtriangleup$}), and 1 ({\color{gray} $\Diamond$}). (a) The strength CCDF scales as $C(s)\sim s^{-\gamma_s+1}$ with $\gamma_s=2.5$. As expected, they are irrespective of $\beta$. (b) The degree CCDF scale as $C(k)\sim k^{-\gamma_k+1}$. The slopes of dashed lines represent $\gamma_k$ (From top to bottom, $\gamma_k=2.5, 3, 4,$ and 7, respectively). For $\beta=1$, $C(k)$ exponentially decays. (c) The link-weight CCDF is presented as $C(w)$, where the slope of the dashed line is $-1.5$. As long as $\beta$ is not so small, $C(w)$ algebraically decays with the same exponent of $C(s)$. However, for $\beta\ll 1$, $C(w)$ follows no longer a simple power- aw. Eventually, for $\beta=0$, $C(w)=\delta(w-1)$. Numerical data are averaged over 100 network configurations for $N=40000$ and $T=N$.}
\caption{Decay-constant ratio $f_{H_q}(m_q)/f_{H_q}(m_{ud}),$ $m_{ud}\equiv\frac{1}{2}\,(m_u+m_d),$ as a function of the light-quark-mass variable $(m_q-m_{ud})/(m_s-m_{ud})$ for three different ans\"atze (constant, linear, and linear plus chiral logarithms) for redefined effective threshold$z_{\rm eff}\equiv\sqrt{s_{\rm eff}}-m_Q-m_q$ \cite{LMSLD}, compared with the findings (\textcolor{blue}{squares}) of Ref.~\cite{LMSIB}.}
\caption{Effect of the multi-scale prediction strategy on a excerpt of the ISPRS Vaihingen dataset. Small objects or surfaces with ambiguous spatial context are regularized by the multiple scales prediction aggregation.\\ (white: roads, {\color{blue!80!black} blue}: buildings, {\color{cyan!80!black} cyan}: low vegetation, {\color{green!80!black} green}:~trees, {\color{yellow!80!black} yellow}: cars)}
\caption{Effect of the fusion strategy on an excerpt of the ISPRS Potsdam dataset. Confusion between impervious surfaces and buildings is significantly reduced thanks to the contribution of the nDSM in the V-FuseNet strategy.\\ (white: roads, {\color{blue!80!black} blue}: buildings, {\color{cyan!80!black} cyan}: low vegetation, {\color{green!80!black} green}:~trees, {\color{yellow!80!black} yellow}: cars)}
\caption{Disputable inconsistencies between our predictions and the ground truth.\\ (white: roads, {\color{blue!80!black} blue}: buildings, {\color{cyan!80!black} cyan}: low vegetation, {\color{green!80!black} green}:~trees, {\color{yellow!80!black} yellow}: cars)}
\caption{Comparison of the relERT (shown on a log-scale) of the single best solver (HCMA, depicted as \textcolor{myred}{$\bullet$}), and the best two algorithm selectors: a kernel-based SVM whose features were selected using a greedy forward-backward strategy (Model 1, \textcolor{mygreen}{$\blacktriangle$}) and an extension of the previous model, which performed a $(10+50)$-GA feature selection on top (Model 2, \textcolor{myblue}{$\blacksquare$}). The performances are shown separately per dimension and selector (top: Model 1, bottom: Model 2). For better visibility, the areas between the curves are highlighted grey and the five BBOB groups are separated from each other by vertical dot-dashed lines.}
\caption{\color{Gray} The depth-diameter (d-D) for modeled craters as given by Bray et al. [2014] and Cox and Bauer [2015], plotted against the observed d-D [Schenk 2002]. Bray et al. [2014] placed the conductive shell thickness at 7 km, while the estimate by Cox and Bauer [2015] was 10 km. In our study, in addition to ice shell over ocean, we also consider the conductive ice over warm convective ice scenario. The three transitions as defined by [Schenk 2002] are annotated on the plot.}
\caption{\color{Gray} The Europan crater profiles (left column) alongside the crater images (right column). The crater rims are shown with arrows. All panels, except (a), have the same vertical scale. The horizontal scale varies across panels. The crater profiles for panels (a-d) were digitized from Figure 6 in Bray et al. [2014], and panel (e) from Figure 9c in Schenk and Turtle [2009]. Crater images are from the Galileo spacecraft. All craters, except (a), are complex craters with anomalous morphologies (Transition II) [Schenk 2002].}
\caption{\color{Gray} The diagram depicting the two setups modeled in our study: (a) a fully conductive shell over ocean, and (b) a conductive ice lid overlying warm convective ice. Schenk [2002] estimated the depth of the ocean at 19-25 km and noted that craters at Transition III would exhibit features indicative of the rheological changes (e.g. multiring structures). While it is assumed that there is ocean at some depth in the conductive-convective scenario (panel (b)), it is not explicitly modeled. This is because at crater sizes investigated in this study (D $<$ 23 km), the presence of the ocean at depths proposed by Schenk [2002] is not expected to make an appreciable contribution.}
\caption{\color{Gray} Examples of (a) strength-depth and (b) temperature-depth profiles used in our simulations. Legend reports conductive lid thickness and temperature of the warm convecting ice (e.g., solid red curve is 6 km thick conductive layer overlying warm ice at 265 K).}
\caption{\color{Gray} Examples of crater profiles used to determine when the crater fully formed for two representative simulations. The end times for each simulation are reported in Tables S1-S3. The simulation is considered over when crater collapse is finished, and no notable change in crater depth, diameter and morphology takes place. The simulation scenario, impactor radius and reported end time ($t_{end}$) are shown in figure panels. Note that the vertical and horizontal scales in the two panels are different. The times prior to ($t_{end}$) are denoted with solid grey and red lines, and the later times with dark green and light green broken lines, and solid blue line (also shown in legend).}
\caption{\color{Gray} Diagram depicting the method for measuring crater depth and diameter, as well as the uncertainty in crater diameter. Since the highest point on the crater rim can be anywhere from one to several cells wide, depending on a simulation, we first found the median (R), which also serves as the measurement of the crater radius (diameter, D = 2R). Then, we obtained the intersecting point (Rwall) of the crater wall and the pre-impact surface. The uncertainty in crater radius is expressed as Rerr = 0.25(R - Rwall), and uncertainty in overall diameter as Derr = 2Rerr. Note that the diagram is meant for visual purposes only, and as such, it does not represent the actual crater proportions.}
\caption{\color{Gray} Time series of the craters produced by a 320 m in radius impactor. Shown are the time steps at (a) 180 s, (b) 240 s, and (c) 440 s. In all panels, the 6 km thick conductive lid over warm convective ice at 265 K (left) and the 8 km ice shell over ocean scenario (right) are plotted side by side for better comparison. In ice over ocean case, the uplifted crater floor is due to the contribution of the ocean underneath. There is a larger uplift of much warmer and more ductile material (see Section 3.3) in the convective ice case as compared to the 8 km ice shell at 180 s and 240 s into the simulation (Figure 4a,b).}
\caption{\color{Gray} Crater depth versus diameter for the (a) ice shell over ocean, and warm convective ice at (b) 255 K and (c) 265 K scenarios, plotted against observed d-D [Schenk, 2002]. Note that the vertical scale in (a) differs from that in (b) and (c). Triangles represent craters for which it was not possible to reliably measure d-D due to warm material overflow, and half-filled points represent the craters for which it was possible to determine the location of the crater rim before it was engulfed by the warm material (see Section 3.3). While these points are not listed separately in legend, their color corresponds to the simulation setups shown in (b) and (c).}
\caption{\color{Gray} The modeled crater profiles for best fit results from our simulations for conductive scenario (8 km shell over ocean (black)) and conductive-convective scenarios (5 km ice lid over warm convective ice at 255 K (blue), and 6 km ice lid over warm convective ice at 265 K (red)). The legend shown in panel (a) applies to all panels.}
\caption{\color{Gray} Time series showing the crater formation for the projectile with Ri = 320 m impacting a 4 km thick ice lid over warm convective ice at 255 K.}
\caption{\color{Gray} (a) The crater profiles at 140 s and 500 s into the simulation. Panels (b) and (c) show the time series at 140 s and 500 s. Note that the vertical scale in (b,c) is different from that in (a).}
\caption{\color{Gray} (a) The crater profiles at 140 s and 500 s into the simulation. Panels (b) and (c) show the time series at 140 s and 500 s. Note that the vertical scale in (b,c) is different from that in (a).}
\caption{\color{Gray} (a) The crater profiles at 140 s and 500 s into the simulation. Panels (b) and (c) show the time series at 140 s and 500 s. Note that the vertical scale in (b,c) is different from that in (a).}
\caption{\color{Gray} (a) The crater profiles at 140 s and 500 s into the simulation. Panels (b) and (c) show the time series at 140 s and 500 s. Note that the vertical scale in (b,c) is different from that in (a).}
\caption{(Left) Comparison of lift ({\Large \color{blue}$\circ$}, {\small \color{red}$\square$}, {\small \color{black}$+$}) and drag coefficients ({\small \color{blue}$\triangle$}, {\small \color{red}$\nabla$}, {\small \color{black}$\times$}). (Right) Comparison of pressure distribution over the airfoil. For both plots, data are shown from present study (blue), Kojima et al. [\citen{Kojima:JA2013}] (red), and Yeh et al.~[\citen{Yeh:AIAA17}] (black). }
\caption{The coefficients of drag and lift versus the coefficient of momentum. The baseline values are indicated by \dashed ~and the controlled cases include pure blowing ($\bigcirc$), co-rotating ($\triangledown$), and counter-rotating ($\triangle$).}
\caption{The correction term $2S C_\mu$ versus the lift and drag forces. The baseline values are indicated by \dashed~ and the controlled cases include pure blowing ($\bigcirc$), co-rotating ($\triangledown$), and counter-rotating ($\triangle$).}
\caption{The correction term $S^2 C_\mu$ versus the lift and drag forces. The baseline values are indicated by \dashed ~and the controlled cases include pure blowing ($\bigcirc$), co-rotating ($\triangledown$), and counter-rotating ($\triangle$).}
\caption{The drag (bottom) and lift (top) forces versus the modified coefficient of momentum $C_\mu^* = (1+S)^2 C_\mu$. The baseline values are indicated by \dashed ~and the controlled cases include pure blowing ($\bigcirc$), co-rotating ($\triangledown$), and counter-rotating ($\triangle$). Inserts in the figure are representative time-averaged flows using $\overline{u}_x = 0$ iso-surface with Reynolds stress $\tau_{xy}$ overlaid in color (follows the visualization set-ups of \figs \ref{fig:baseline} and \ref{fig:A9_sepReyZ}).}
\caption{Time history of lift and drag for baseline ($-$), case D ({\color{red}$-$}), and case E ({\color{blue}$-$}). Representative instantaneous flow fields are visualized with the $Q$-criterion isosurface colored with pressure, following the same setup in \fig \ref{fig:baseline}. Symbols on the lift curves indicate the time of visualization.}
\caption{Comparison on Translation Examples. We italicize some {\em \color{blue} translation errors} and highlight the {\bf \color{red} correct ones} in bold.}
\caption{Frames of 5 videos in Flash-MNIST and the corresponding label. The \textcolor[rgb]{0.44,0.188,0.627}{\textbf{purple bold digits}} in label are digits flashing in the video.}
\caption{\label{tab:8tasks} Performance measured in accuracies achieved by each method on each of the learned tasks at the end of the 8 tasks sequence (Table for bar plot in Figure~\textcolor{red}{5a} of the main paper).}
\caption{\textbf{One-Shot \rgb Camera Localization. Top:} We use our pose estimates to fit a 3D scene model to test images. \textbf{Middle:} We visualize a camera trajectory by plotting camera positions as dots and connecting consecutive frames. We show our estimates in \textcolor{cyan}{cyan} and ground truth in \textcolor{green}{green}. Our results are highly accurate with very few outliers. We trained from \rgb images and ground truth poses only, without using the 3D scene model. \textbf{Bottom:} Results of competing methods which are less accurate and produce many wrong estimates. To improve the visualization, we connect only consecutive frames within 5 meters range. \textcolor{red}{Red:} In the spirit of PoseNet \cite{kendall2015convolutional}, we train a CNN to predict poses directly. \textcolor{orange}{Orange:} Results of DSAC \cite{brachmann2017dsac} trained with a 3D scene model. }
\caption{\textbf{Generalization Capabilities.} We train two CNNs for scene coordinate regression and pose hypothesis scoring following DSAC~\cite{brachmann2017dsac} but using only two training images (\textcolor{gray}{gray}). The system cannot generalize to an unseen test image (\textcolor{green}{green}), and produces an estimate far off (\textcolor{orange}{orange}, 3.7m and 12.6$^\circ$ pose error). Exchanging the scoring CNN for a soft inlier count (see Sec.~\ref{sec:meth:score}), we obtain an accurate estimate (\textcolor{blue}{blue}, 0.1m and 0.3$^\circ$ pose error). Note that, apart from the scoring function, we fix all components of the test run, including pose hypotheses sampling. This experiment illustrates that scene coordinate regression generalizes well, while score regression does not.}
\caption{\textbf{Estimated Camera Trajectories}. We plot estimated camera locations as trajectories within the respective 3D scene model (untextured for 7Scenes). We show ground truth in \textcolor{green}{green}, DSAC \cite{brachmann2017dsac} estimates in \textcolor{orange}{orange}, and our results in \textcolor{blue}{blue} and \textcolor{cyan}{cyan} when trained with and without a 3D model, respectively. Note that DSAC produces many wrong estimates despite being trained with a 3D model.}
\caption{\label{Fig_TSonOB} Optical layout of the test bed. For clarity, the TS is shown next to the OB. In the experiment, it is placed on top of the OB (shown right) at the position marked with the dashed orange square with dedicated mounting feet (MF). The origin of the coordinate system is on the surface of the OB. The positive z axis points upwards (out of the paper plane). The TS can be adjusted in all three lateral degrees of freedom x, y, z and it can be rotated around all three axes. \Rx{} beam:\green,\LO{} beam:\blue,\Tx{} beam:\red.}
\caption{\label{Fig_ModulationBenchElectronics} Schematic of laser preparation, electronics, and interferometer readout. Fibres:\green, laser beams:\red, cables:\grey}
\caption{\textbf{We explore training instance segmentation models with partial supervision}: a subset of classes (\textcolor{green}{green} boxes) have instance mask annotations during training; the remaining classes (\textcolor{red}{red} boxes) have only bounding box annotations. This image shows output from our model trained for 3000 classes from Visual Genome, using mask annotations from only 80 classes in COCO.}
\caption{\textbf{Mask predictions from the class-agnostic baseline (top row) \vs our \methodname approach (bottom row)}. \textcolor{green}{Green} boxes are classes in set $A$ while the \textcolor{red}{red} boxes are classes in set $B$. The left 2 columns are $A=\{\text{voc}\}$ and the right 2 columns are $A=\{\text{non-voc}\}$.}
\caption{\textbf{Example mask predictions from our \methodname on 3000 classes in Visual Genome.} The \textcolor{green}{green} boxes are the 80 classes that overlap with COCO (set $A$ with mask training data) while the \textcolor{red}{red} boxes are the remaining 2920 classes not in COCO (set $B$ without mask training data). It can be seen that our model generates reasonable mask predictions on many classes in set $B$. See \S\ref{sec:exp_vg} for details.}
\caption{Comparisons of three photometric redshift estimators to available spectroscopic redshifts in Stripe 82, reproduced from \catpaper{}. The comparison is limited to $i < 22.5$ and $0.01 < z_{\rm spec} < 0.8$. The left and middle panels are from the \redmapper{} project \citep{rozo15}, while the right panel compares neural-network \photozs{} from \citet{reis12}. The 3$\sigma$-clipped dispersion is listed in each panel along with the fraction of catastrophic outliers defined by $\abs{\Delta z} > 0.1$. Contours are plotted at high data densities with 0.3 dex logarithmic spacing in the left and middle panel and 0.4 dex in the right panel. The 1-to-1 relation is plotted in each panel as a thin light grey line. \label{fig:photoz_specz}}
\caption{\label{fig:design} The design of cosmological parameter sampling for our simulations (100 massive + 1 massless neutrino models in total). The two fiducial models ($M_\nu$=0.0~eV and 0.1~eV, $\Omega_m$=0.3, $A_s$=2.1$\times$10$^9$) are marked in red. All parameter values are listed in Table~\ref{tab: CosmoParsm}. {\red A flat universe ($\Omega_\Lambda+\Omega_m=1$) is assumed.} The other cosmological parameters are fixed at $h$=0.7, $n_s$=0.97, $\Omega_b$=0.046, and $w$=$-1$.}
\caption{\label{fig:pmatter_fidu}{\bf Upper}: the matter power spectra for the fiducial models, where $\Omega_m$=0.3, $A_s$=2.1$\times10^9$, and $M_\nu=0.1$~eV (massive, solid curves) and 0 eV (massless, dashed curves). Other cosmological parameters are fixed at $h$=0.7, $n_s$=0.97, $\Omega_b$=0.046 and $w$=$-1$. The simulations (green curves) have 1024$^3$ DM particles and box size 512~$h$/Mpc. We also show the theory curves from linear theory for comparison (black curves). {\bf Lower}: the fractional difference between the $P_m$ of the massive model and that of the massless model, {\red i.e.}, the suppression due to massive neutrinos, measured in our simulations (green solid curves) and compared with linear theory (black solid curves) and two versions of \texttt{Halofit} (dashed curves). We also show additional higher-resolution runs, with the same number of particles but half the box size (labeled as ``higher res.'', brown solid curve). The dashed vertical lines denote the approximate division between linear and nonlinear scales.}
\caption{\label{fig:pmatter_all} The difference of the matter power spectra of massive models from the massless fiducial model ($z$=0), {\red shown for both simulations (solid curves) and linear theory (dashed curves)}. Parameters are listed in Table~\ref{tab: CosmoParsm}.}
\caption{\label{fig:pmatter_all} (Cont.) The difference of the matter power spectra of massive models from the massless fiducial model ($z$=0), {\red shown for both simulations (solid curves) and linear theory (dashed curves)}. Parameters are listed in Table~\ref{tab: CosmoParsm}.}
\caption[]{\label{tab: CosmoParsm} Cosmological parameters for our simulations. {\red A flat universe ($\Omega_\Lambda+\Omega_m=1$) is assumed.} Other parameters are fixed at $h=0.7$, $n_s=0.97$, $\Omega_b=0.046$, and $w=-1$. All points are visualized in Fig.~\ref{fig:design}.}
\caption{ $(a)$ ABC-GAN architecture for ABC computation. The external dependence on noise ($\epsilon$) is shown in \textcolor{red}{red}. Two distinct network paths (shown in \textcolor{green}{green} and \textcolor{pink}{pink}) correspond to two different optimizations~(resp., \textcolor{green}{\PA} and \textcolor{pink}{\PB}) akin to ratio test in GANs; $(b)$ ABC-GAN implementation $\alpha=10^{-3}$, $m=50$.}
\caption{$(a)$ Posterior samples for ABC-GAN (shown in \textcolor{green}{green}) for one run in the mixture of normals experiment (Section~\ref{sec:mixturenormal}) with $5000$ iterations with a mini-batch size of $10$ and sequence length $10$, using a learning rate $10^{-3}$. The probability density function for the mixture normal and the low-variability component are shown in \textcolor{red}{red} and \textcolor{blue}{blue} respectively. $(b)$ Histogram of posterior samples for first run of BOLFI. Total number of samples is $5000$. We note that the low-probability space is not captured by BOLFI compared to ABC-GAN.}
\caption{ \textcolor{\refrevisions1}{Comparing the Juvinas (purple) spectrum with Albite (black), Anorthite (grey), ferrosilite (brown) and enstatite (blue). Purple dashed vertical lines indicate features spotted in the Juvinas spectrum. Laboratory spectra of Anorthite (grey) and Albite (black) measured by \cite{salisbury91}. Wavelength peak positions of bands in these spectra close to the bands of olivine and pyroxene are indicated in the plot.} }
\caption{\label{Fig-Results} (a) $V(\dot{\gamma})$ measured on control surface (\textcolor{black}{$\blacktriangle$}), HA840-h (\textcolor{ForestGreen}{$\blacksquare$}), HA840-l (\textcolor{orange}{$\blacklozenge$}), and HA58 (\textcolor{magenta}{\Large $\bullet$}) brushes. The solid line is the GCB prediction for non-deformable surfaces separated by $20$ nm. (b) Experimentally measured $h(\dot{\gamma})$ (symbols as in (a)). Solid line indicates the constant value of $z=20$ nm used in GCB theory. Error bars accounting for standard error and uncertainty on $h$ and $V$ are about the size of the symbols.}
\caption{\label{Fig-Model} (a) $h(\dot{\gamma})-H_0$ data on HA brushes (symbols as in Fig. \ref{Fig-Results}, vertical scale according to arrows), and theoretical predictions for $\delta(\dot{\gamma})$ with $M=5$ Pa (dashed line), 57 Pa (solid line), and 15000 Pa (dotted line). The shaded area around the theoretical curves is defined by the predictions obtained when $\xi$ is varied from lower to upper bound for each brush. Inset: sketch showing the location of the no-slip plane at $\xi$ below the brush surface. (b) Theoretical ($V_\text{th}$) {\it vs} experimental ($V_\text{exp}$) velocities (symbols as in (a)). The solid line corresponds to $V_\text{th}=V_\text{exp}$ and the dashed line to $V_\text{th}=1.2V_\text{exp}$. Variations of $V_\text{th}$ due to changes in $\xi$ and error bars on $V_\text{exp}$ are about the symbol size. Inset: measured (symbols) and predicted (lines) deviation from linearity, $\Delta V=V(\dot{\gamma})-S\dot{\gamma}$, with $S$ the slope in the limit of small shear rates. (c) Best fit values (\textcolor{ProcessBlue}{$\blacksquare$}), measured reference (\textcolor{ProcessBlue}{\Large$\bullet$}) and predictions (solid line) for $M$ as a function of $\xi$. Error bars correspond to the uncertainty on $\xi$.}
\caption{\label{Fig-testmodelSI} $\delta$ $vs$ $\dot{\gamma}$ measured on HA840-h brush (\textcolor{red}{$\blacksquare$}), predicted with $M=57$ Pa and $\xi=74$ nm (solid line), $M=57$ Pa and $\xi=0$ (dash-dotted line), $M=120$ Pa and $\xi=0$ (dashed line), $M=57$ Pa and $\xi=74$ nm without inertial force (blue dotted line). }
\caption{\label{Fig-strainSI} Brush strain $\mathcal{S}$ $vs$ $\dot{\gamma}$ computed for the HA840-h (\textcolor{ForestGreen}{$\blacksquare$}), HA840-l (\textcolor{orange}{$\blacklozenge$}), and HA58 (\textcolor{magenta}{\Large $\bullet$}) brushes.}
\caption{Extending the variational autoencoder (VAE) model to the case of hybrid labeled/unlabeled observations. % (a) (\subref{fig:vae-standard}) the standard VAE model extended to the paired observation case: an encoder {\color{blue}$q(\z|\image,\pose)$} and decoder {\color{DarkGreen}$p(\image,\pose|\z)$} mapping to/from a latent code $\z$; % (b) (\subref{fig:vae-semi}) the hybrid VAE model introducing a discriminative variational model {\color{red}$q(\pose|\image)$} % (c) (\subref{fig:discriminative-network}) The architecture of the discriminative network. }
\caption{\red{Permeability ($k$), porosity ($\epsilon$), average pore sizes ($s$), and related dimensionless parameters for tested foams. Permeability and pore size values from \cite{manes2011turbulent} are noted in parenthesis for the 10 and 60 ppi foams. Note that $s^+ = s u_\tau /\nu$ and $Re_k = \sqrt{k} u_\tau/\nu$ are defined using the friction velocity upstream of the porous section. Porosity ($\epsilon$) was estimated from solid volume displacement in water and permeability was estimated from pressure drop experiments.}}
\caption{Contour maps showing variation in premultiplied frequency spectra \red{(normalized by $U_e^2$)} for streamwise velocity as a function of wall-normal distance $y/\delta$ over the smooth wall (a), the 100 ppi foam (b), the 10 ppi foam (c), the thin 20 ppi foam (d), the baseline 20 ppi foam (e), and the thick 20 ppi foam (f). The spectra are plotted against a normalized streamwise length scale, $U/f\delta$, computed using Taylor's hypothesis. The white box in (a) denotes the region typically associated with VLSMs while the markers ($*$) represent the nominal frequency $f_{KH}$ for structures resembling Kelvin-Helmholtz vortices. \red{The spectra refer to the same physical measurement location, corresponding to $x/h=42$ for the foams with $h = 12.7$mm and $x/h=21$ and $84$ for the thick and thin 20 ppi foams, respectively.}}
\caption{(a) Scaled velocity gradient $(y/U_e)\partial U/\partial y$ plotted as a function of $y/\delta$ for the smooth wall profile and porous foam data. Equation (\ref{eq:log-law-fit}) was fitted to these data to estimate the normalized displacement height, $y_d/h$, and the friction velocity weighted by the von Karman constant, $u_\tau/(\kappa U_e)$, shown in (b) and (c), respectively. Using these estimates for $y_d$ and $u_\tau/\kappa$, the roughness height $k_0/h$, shown in (d), was evaluated from the velocity profiles using equation (\ref{eq:log-law}). Dotted lines in (a) represent the upper limits of $y/\delta=0.16,0.20$ and $0.25$ employed in the fitting procedure. The dashed line represents the minimum lower limit, $y/\delta > 0.02$. Larger marker sizes in (b,c,d) denote higher values for the upper limit. The red cross in (c) represents the friction velocity estimated via a linear fit to the near-wall velocity measurements over the smooth wall. \red{Horizontal error bars in panels (b)-(d) represent uncertainty in pore size, $s$.}}
\caption{Average PSNR/SSIM for scale factors $\times 2$, $\times 3$ and $\times 4$ on datasets Set5 \cite{set5}, Set14 \cite{set14}, B100 \cite{b100} and Urban \cite{selfex}. \textcolor{red}{Red} indicates the best performance and \textcolor{blue}{blue} indicates the second-best performance.}
\caption{\red{Schematics of a typical parallel heat flux radial profile $q_{||}(r_u)$ given by Eq. (\ref{eq:qparfit}) (solid line). The dashed line represents the heat flux associated with the far SOL, $q_f\exp(-r_u/\lambda_f)$. The power entering the near and far SOL, $P_n$ and $P_f$ respectively, are given by the integral of the red and green shaded areas. \label{fig:pnpf}}}
\caption{Time traces of a) line-averaged electron density $n_{e,av}$ (blue) and plasma current $I_p$ (red) b) N$_2$ \red{(blue) and D$_2$ (red)} flow measured by the piezoelectric valve c) total radiated power $P_{rad}$ from bolometric measurements \red{(blue), the plasma effective charge $Z_{eff}$ (red) and the loop voltage $V_{loop}$ (magenta), rescaled for plotting} d) electron temperature on axis $T_{e,ax}$ (blue) and in the edge region $T_{e,edge}$ (red) from Thomson scattering measurements. e) TCV cross section together with the magnetic equilibrium reconstruction provided by LIUQE \cite{Hofmann1988}. The IR camera field of view (red dashed lines), the location of the flush mounted LPs (orange dots) \red{and of the TS measurements (blue crosses), the trajectory of the RP (magenta thick line)}, and the position of the valve used for N$_2$ injection (green rectangle) are also shown. \label{fig:discharge}}
\caption{\red{Plasma emissivity $\epsilon$ for discharge \#56142 before (a), during (b) and after (c) N$_2$ injection, computed from the tomographic inversion of 64 gold foil bolometers measurements.} \label{fig:bolo}}
\caption{\red{Power radiated inside each flux surface $P_{rad,\rho}$ for discharge \#56142 before (red), during (blue) and after (green) N$_2$ injection.} \label{fig:prad}}
\caption{\red{Electron temperaure radial profile $T_e(\rho)$ from Thomson scattering measurements for discharge \#56142 before (red circles), during (blue squares) and after (green diamonds) N$_2$ injection. Smooth interpolated profiles are plotted with solid lines.} \label{fig:thomson}}
\caption{\red{Radial profiles at the OMP of parallel heat flux $q_{||}(r_u)$ , before N$_2$ injection (red) and for $f_{rad}>70\%$ (blue). The fit of $q_{||}(r_u)$ with Eq. (\ref{eq:qparfit}) is shown with thick lines, while the heat flux associated with the far SOL $q_{||,f}(r_u) = q_f \exp(−r_u/\lambda_f)$ is shown with a black dashed line for the case before N$_2$ injection. The LCFS position from LIUQE is marked by a black vertical line. The LCFS location according to the method used in Ref. \cite{Tsui2017} is shown with vertical dotted lines.}\label{fig:RP}}
\caption{Floating potential profile along the direction of the limiter, $Z-Z_{ax}$, with $Z_{ax}$ the vertical position of the plasma magnetic axis, before N$_2$ injection \red{(red dots)} and for $f_{rad}>70\%$ (blue squares). The profiles are interpolated with cubic splines. The drop in the floating potential $\Delta V_{fl}$ is shown. \red{Measurements from LPs} shaded by the neighboring tiles \red{are plotted with open symbols}. \label{fig:LP}}
\caption{\textcolor{blue}{The two different possible approaches to defining the energy conversions between} $E_k$, $\Pi_1$ and $\Pi_2$ in a turbulent multi-component compressible stratified fluid \textcolor{blue}{discussed in this paper}.}
\caption{Two-level parallel MPI/OpenMP pseudo-code for one iteration. \textcolor{red}{MPI routines} and \textcolor{blue}{OpenMP compiler directives} are in red and blue colors, respectively. By removing highlighted lines, which represent work on data transfers between spatial subdomains, one obtains pure OpenMP pseudo-code. By discarding blue lines, one obtains pure MPI pseudo-code with spatial domain decomposition. The serial pseudo-code corresponds to un-highlighted black lines.}
\caption{Comparison results on the three benchmark datasets. Performances are measured by the rank1, rank5 and rank10 matching accuracy of the cumulative matching curve, as well as mAP. The best performances are indicated in {\color{red}\textbf{red}} and the second indicated in {\color{blue}\textbf{blue}}.}
\caption{\small (a) Bifurcation diagram for a channel with $\alpha=30$, $\alpha_w=0.285$, $\alpha_h=0.024$ and $s=40$ spanning flow rate and bubble diameter. The path of a pitchfork bifurcation separates regions of unstable and stable steady on-rail propagation. The path of a limit point encloses a region in which there are no steady on-rail states. Schematic bifurcation diagrams of the bubble speed as a function of flow rate are shown for two values of the bubble diameter indicated with dashed lines: (b) $D=1.27$ ($D^*=8.74$ mm) and (c) $D=1.37$ ($D^*=9.44$ mm). Solid (dashed) lines denote stable (unstable) solutions, while blue (black) lines denote off-rail (on-rail) states, respectively. In (b), the upper \textcolor{red}{$\blacksquare$} denotes the pitchfork bifurcation point that stabilises the on-rail state. Upon the increase of the bubble diameter, the stable on-rail solution branch interacts with the unstable RVB solutions in the lower part of diagram (b), which results in its disconnection in (c) so that there is no simple steady on-rail solution in the region delimited by the upper two saddle node points \textcolor{red}{\Large\bf$\bullet$} in (c), although the unstable RVB solution remains in the lower part of the diagram. The disconnection of the stable on-rail solution branch arises through a transcritical bifurcation at $D \approx$ 1.365 ($Q=0.039$), marked as T in (a).}
\caption{Representative errors made by different systems, indicated in \red{red}.}
\caption{\small \textbf{Overview of our framework.} We take as input an image and set of bounding boxes. The scene layout $\HB$ is predicted by the {\bf \textcolor{msgreen}{layout module}}, a skip-connected CNN. Each bounding box is then represented by features from three sources: ROI-pooled features extracted from a {\bf \textcolor{msblue}{fine module}} that uses the original resolution, full image context features from the {\bf \textcolor{msred}{coarse module}}, and the {\bf \textcolor{mspurple}{bounding-box location}}. These features are concatenated and passed through several layers, culminating in the prediction of a shape code $\sB$, scale $\cB$, translation $\tB$, and rotation of the object $\qB$. The shape code is mapped to voxels $\VB$ by the {\bf \textcolor{msorange}{shape decoder}}. }
\caption{Grapheme Wins. Phoneme errors indicated in \red{red}.}
\caption{Phoneme Wins. Grapheme errors indicated in \red{red}.}
\caption{Grapheme Wins for Multi-dialect. Phoneme errors indicated in \red{red}.}
\caption{GTSRB class sample distribution. \textcolor{blue}{Blue}: seen classes (\#22).\textcolor{orange}{Orange}: unseen classes (\#21).}
\caption{TT100K class sample distribution. \textcolor{blue}{Blue}: seen classes (\#24).\textcolor{orange}{Orange}: unseen classes (\#12).}
\caption{ The $N$-dependence of $\delta_\mathrm{tr}$. The Hamiltonian and the lattice are the same as in \cite{Gemmer2017, Iyoda2017}. The parameters are given by $\omega=1,8,-50$, $g=0.1$, and $\beta=0.1$. $N^\ast_{\omega=1}$ and $N^\ast_{\omega=8}$ are the average particle numbers in the canonical ensemble for $\omega=1,8$, respectively. For each data point, 10 energy eigenstates are sampled, and the error bar represents their standard deviation. The red circle indicates the parameters used in \cite{Gemmer2017} ($\omega=-50$ and $N=4$). \textcolor{OliveGreen}{The blue circle indicates the parameters used in \cite{Iyoda2017} ($\omega=1$ and $N=4$).} }
\caption{Wormhole attack experiment results. This table presents the capabilities of the Wormhole Jammer setup for each SF/Packet Size pair. \colorbox{green!30}{S}: successful jamming (>95\%), \colorbox{yellow!30}{M}: mixed success (0-95\%), \colorbox{red!30}{F}: failure to jam (<0\%).}
\caption{{\bf Experimental setup}. {\bf (a)} Top view schematic of the T$^3$C facility (not to scale). Air is captured at the leeward side of the cavitators, as indicated. We attached 2, 3, or 6 cavitators equally distributed around the perimeter of the inner cylinder. The rotation of the cylinders is shown as $\omega_i$ and $-\omega_o.$ (b) Vertical cross-section, showing the position of the torque sensor. \red{The sensor is located in the inner cylinder, so that the torque between the driving shaft and the inner cylinder is measured.} To control the void fraction, we fill the cylinder only partially with water, so that the void fraction $\alpha$ is controlled by measuring the relative height of the water level. \red{Turbulent mixing ensures axial mixing between the two phases, whereas the centrifugal accelerations push the water towards the outer cylinder and, consequently, the air towards the inner cylinder.}}
\caption{\red{ Snapshots of air cavities at $\Re_s = 8\times 10^5$, for 3 different rotation ratios: {\bf (a)} $a=0.14$ (counter-rotation), {\bf (b)} $a=0$ (stationary OC) and {\bf (c)} $a=-0.2$ (co-rotation). The direction of the cylinder rotation is indicated by the curved arrows, in which $\omega_i$ and $\omega_o$ indicate the direction of the inner and outer cylinder, respectively. The global gas volume fraction is $\alpha=2\%$. The vertical bars and horizontal rings are essential structural parts of the setup. (I) The air cavity. (II) The cavitator. (III) Cylinder not covered with an air cavity. (IV) The contact line of the cylinder-water-air interface. (V) In the counter-rotating case, many bubbles are trapped in these `Taylor vortices'. In fig.\(b) and especially in fig.\(c), the flow is radially more stably stratified. Therefore, less bubbles are present and the air cavity is better visible.}}
\caption{Snapshots of air cavities at $\Re_i = 5 \times 10^5$ with a stationary outer cylinder, 3 cavitators, and 2\% of air. We zoomed in on the top of the cylinder. We show 2 photos taken at time = $t_1$ --- when the cavitator is visible, and $t_2$ --- when the closure region is visible. {\bf (a)} Cavitator (vertical white strip in image) with development of the air cavity \red{at $t_1$}. {\bf (b)} The closure region of the air cavity \red{at $t_2$}. The visible white bar here is {\it not} a cavitator, \red{but a blank that is mounted flush with the cylinder surface}. Note the dependence of the cavity length on the height. The white arrow indicates the position of the closure region, which is governed by the re-entry jet mechanism.}
\caption{Streamwise air cavity length on the inner cylinder as a function of $\Re_i$. The outer cylinder is stationary. We used 3 cavitators. The coverage is extracted by visual means from a series of images similar to those of fig.\\ref{Chap_Six_fig:visu2}. We show results for three different axial positions, close to the top ($z/L=3/4$), at mid-height ($z/L=1/2$) and close to the bottom ($z/L=1/4$). The estimated error bar is shown \red{in the bottom left corner of the graph. In dashed black, we added the streamwise length between two cavitators $2 \pi r_i/3$, which is the upper limit of the streamwise cavity length. On the right y-axis, we normalized the streamwise cavity length with the distance between two cavitators.} }
\caption{Percentage of air cavity coverage on the inner cylinder as a function of $\Re_i$. The outer cylinder is stationary. We used 3 cavitators. The coverage is calculated by integrating the streamwise cavity lengths (see fig.\\ref{Chap_Six_fig:length}). The estimated error is shown \red{in the bottom right corner of the graph}.}
\caption{Global dimensionless torque and drag reduction percentage for 0\%, 2\%, and 4\% of air. Here we mounted 3 cavitators, and we kept the outer cylinder stationary. {\bf (a)} Dimensionless torque $G$ as a function of the inner Reynolds number $\Re_i$. {\bf (b)} Drag reduction percentages as a function of inner Reynolds number $\Re_i$. As comparison, we also show the bubbly DR results from \citet{gil13}. \red{A typical error bar is shown in both graphs.}}
\caption{\red{The drag reduction, as shown in fig.\\ref{Chap_Six_fig:DR} as a function of air cavity coverage (fig.\ref{Chap_Six_fig:coverage}). Note that both the DR and the coverage depend on $\Re_i$. A typical error bar is shown for both the drag reduction as the coverage percentage. }}
\caption{Dimensionless torque and DR for the case of counter-rotating cylinders with 3 cavitators as a function of shear Reynolds number $\Re_s$. The rotation ratio equals $a=0.2$. {\bf (a)} Dimensionless torque $G$ as a function of shear Reynolds number $\Re_s$. {\bf (b)} Drag reduction percentages as a function of shear Reynolds number $\Re_s$. The DR is significantly smaller than for the case of only inner cylinder rotation (fig.\\ref{Chap_Six_fig:DR}). \red{A typical error bar is shown in both graphs.} }
\caption{{\bf (a)} Dimensionless torque with 2, 3, or 6 cavitators as a function of $\Re_i$ for stationary outer cylinder. {\bf (b)} The drag reduction for $\alpha=2\%$ for the case with 2, 3 or 6 cavitators. The DR percentages are similar for a constant gas volume fraction $\alpha$, although the global torque is increased by the cavitators, which induce an additional pressure drag \citep{zhu18}. \red{A typical error bar is shown in both graphs.} }
\caption{Dimensionless torque and DR as a function of $\Re_i$. The reference case is without cavitators and with $\alpha=0$. The other cases are measured with 3 cavitators. The outer cylinder is stationary. {\bf (a)} The dimensionless torque as a function of the inner Reynolds number $\Re_i$. {\bf (b)} The net drag reduction as a function of the inner Reynolds number $\Re_i$. The net DR as compared to the reference case is negative, i.e.\instead of drag reduction we observe a drag increase.\red{A typical error bar is shown in both graphs.}}
\caption{{\bf (a)} The dimensionless torque $G$ as a function of gas flow rate $\dot{Q}$. We observe that $G$ does not depend on $\dot{Q}$. Here, the number of cavitators is 3, and we measure at $\Re_i=1\times 10^6$ with a void fraction of $\alpha=2\%$. The OC is kept stationary. On the top x-axis, we non-dimensialized $\dot{Q}$ with the volume of the system $V$, which equals $\dot{\alpha}$, i.e.\the void fraction injected per minute. Injecting more air than shown here is not possible as it leads to an increase in void fraction as water is pushed out of the system, and thus to an unfair comparison.\red{In plot {\bf (b)}, we show the same data but made dimensionless: $(G-G_0)/G_0$, in which $G_0 = G(Q=0)$. In this way, the relative change of $G$ is shown. A typical error bar is shown in both graphs.}}
\caption[The linear bin-cut examples]{Effects of binning and cut-off level on linear mock data. Panels on the left-hand side are for a $DS$ value of 4, while \textcolor[rgb]{1.00,0.00,0.00}{those on the right-hand side} are for a $DS$ value of 3.1. The number of bins and sigma cut-off are (10, 6) and (20, 3) for the upper and lower left-hand side panels, and (10, 6) and (30, 6) for the upper and lower right-hand side panels, respectively. Symbols are the same as those in Fig.~\ref{fig:psulin}.}
\caption{Detection with quantitative results using BER, smaller is better. For our proposed architecture, we use image triplets of ISTD training set. These models are tested on three datasets. The best and second best results are marked in {\color{red}{red}} and {\color{blue}{blue}} colors, respectively. }
\caption{Detection with quantitative results using BER, smaller is better. For our proposed architecture, we use image pairs of SBU training set together with their roughly generated shadow-free images by Guo et al. \cite{guo2013paired} to form image triplets for training. The best and second best results are marked in {\color{red}{red}} and {\color{blue}{blue}} colors, respectively. }
\caption{Removal with quantitative results using RMSE, smaller is better. The original difference between the shadow and shadow-free images is reported in the third column. We perform multi-task training on ISTD and compare it with three state-of-the-art methods. The best and second best results are marked in {\color{red}{red}} and {\color{blue}{blue}} colors, respectively.}
\caption{Component analysis of ST-CGAN on ISTD by using RMSE for removal and BER for detection, smaller is better. The metrics related to shadow and non-shadow part are also provided. The best and second best results are marked in {\color{red}{red}} and {\color{blue}{blue}} colors, respectively.}
\caption{Performance evaluation on WILLOW dataset~(Varying the number of outliers). \textcolor{red}{\textbf{Red}} and \textcolor{blue}{\textbf{blue}} bold numbers indicate the best and the second-best performances.}
\caption{This figure is analogous to Fig.\4, with the difference that the measurements are done at$V_{\rm g}$ = 0.5 V, the point of lowest DC conductivity ($\sigma_{\rm 0} \approx$ 2 e$^2$/h), corresponding to graphene close to the charge neutrality point. The calculations are done for $E_{\rm F}$ = 0.05 eV, although due to puddles the effective Fermi energy in the experiment could be significantly larger. We observe that the value of the THz photoconductivity is positive, while in Fig.\4 it is negative. Panels\textbf{A)} and \textbf{B)} show the THz photoconductivity for three pump photon energies $E_{\rm ph} = $ 0.5, 1.0, and 2.5 eV. Panel \textbf{C)} shows the calculated carrier multiplication factor CM, defined in the main text. To understand the magnitude of the photoconductivity, we note that the added carrier density in the conduction band corresponds to $\Delta n$ = CM$\cdot n_{\rm exc}$. Using CM = 3 (for 2 eV photon energy), $n_{\rm exc}$ = 0.2$\cdot$10$^{12}$/cm$^2$, and a mobility of $\mu$ = 1000 cm$^2$/Vs, we obtain an increased conductivity of $\Delta \sigma = \Delta n e\mu$ = 4 e$^2$/h. This is agrees with the calculated increase in conductivity (see Fig.\6B for 2 eV photon energy). The measured value is significantly lower, most likely due to electron-hole puddles.}
\caption{ \textbf{A)} Illustration of the optical pump -- THz probe measurement technique applied to turbostratic graphene supported by SiC substrate (see Methods). This device consists of multiple graphene monolayers with random relative orientation and very low doping. Through these measurements, we further examine graphene close to the charge neutrality point, as in \textcolor{blue}{\textbf{\ref{Fig4}}} for the gate-tunable sample. \textbf{B)} The THz photoconductivity $\Delta \sigma_{\rm THz}$ as a function of the pump pulse fluence, parametrized by the photoexcited carrier density $n_{\rm exc}$, for several values of the pump photon energy $E_{\rm ph} = $ 1.5, 1.8, 2.0, and 2.5 eV (empty circles). The turbostratic graphene samples are characterized by a very low doping, so this panel should be compared to \textcolor{blue}{\textbf{\ref{Fig4}A}}. The solid lines are linear fits to the data. \textbf{C)} The THz photoconductivity at $n_{\rm exc} = 0.2 \times 10^{12}$/cm$^2$ [vertical dashed line in \textbf{B)}], as a function of the pump photon energy. The dashed line is a linear fit to the data. Consistently with the THz photoconductivity measured on gated samples, we observe larger THz photoconductivity for larger photon energy, as discussed in \textcolor{blue}{\textbf{\ref{Fig4}}}. This trend indicates that efficient interband heating takes place in the, intrinsically undoped, turbostratic graphene samples. }
\caption{Visualizing which angles fool the classifier for 50 random examples. For each dataset and model, we visualize one example per row. \textcolor{red}{Red} corresponds to \emph{misclassification} of the images. We observe that the angles fooling the models form a highly non-convex set.}
\caption{Performance comparison between image denoising algorithms on widely used classical images, in terms of PSNR (in dB). The best results are highlighted with \textcolor{red}{red} color while the \textcolor{blue}{blue} color represents the second best denoising results.}
\caption{\label{Fig. 1} (a) Schematic of the substrate used for the condensation experiments. Transparent interdigitated ITO electrodes (red) are patterned on the glass substrate (grey), which is then coated with a hydrophobic dielectric polymer film (green). A schematic of a condensate droplet under EW is also shown. (b) Comparison between breath figures without EW (control) (C-i to C-iii) and under EW ($U_{rms}=150$ V; $f=1$ kHz) (EW-i to EW-iii) at different time $(t)$ instants. The (e)lectrode-(g)ap geometry underneath the dielectric film is indicated by the solid red and white lines. \textcolor{blue}{Gravity points from top-to-bottom i.e. along the negative y--direction.}}
\caption{Ensemble average global error $\langle\mathcal{E}_k\rangle$, for each material described in Tables~\ref{Tab:Disks} and \ref{Tab:CoarseGrained}, as a function of the $k$-th step along the 3 proposed decimation processes. The average and standard deviations (error bars) were obtained by applying Eqs.~\eqref{Eq:EAGErr}. The statistically sensitive decimation step $Z$ was computed with Eqs.~\ref{Eq:DIS}, applied to the extrema, [\textbf{---}] $\min\left(\ell_{\beta,j}\right)$ and [$\boldsymbol{\cdots}$] $\max\left(\ell_{\beta,j}\right)$, and the average, [\textbf{- -}] $\overline{\ell_{\beta,j}}$, of the correlation lengths of the different descriptors $\beta$ and phases $j$.}
\caption{Placebo tests on Basque Country terrorism data: {\protect\tikz \protect\draw[color={rgb:red,4;green,0;yellow,1}] (0,0) -- plot[mark=o, mark options={scale=2}] (0.25,0) -- (0.5,0);}, DID; {\protect\tikz \protect\draw[color={rgb:red,244;green,226;blue,66}] (0,0) -- plot[mark=triangle*, mark options={scale=2,fill=white}] (0.25,0) -- (0.5,0);}, ED; {\protect\tikz \protect\draw[color={rgb:red,0;green,5;blue,1}] (0,0) -- plot[mark=+, mark options={scale=2}] (0.25,0) -- (0.5,0);}, MC-NNM; {\protect\tikz \protect\draw[color={rgb:red,66;green,200;blue,244}] (0,0) -- plot[mark=x, mark options={scale=2}] (0.25,0) -- (0.5,0);}, RVAE; {\protect\tikz \protect\draw[color={rgb:red,66;green,107;blue,244}] (0,0) -- plot[mark=diamond, mark options={scale=2}] (0.25,0) -- (0.5,0);}, SCM; {\protect\tikz \protect\draw[color={rgb:red,244;pink,66;blue,223}] (0,0) -- plot[mark=triangle, mark options={scale=2, rotate=180}] (0.25,0) -- (0.5,0);}, VT-EN.\label{basque-sim}}
\caption{Placebo tests on West German reunification data: {\protect\tikz \protect\draw[color={rgb:red,4;green,0;yellow,1}] (0,0) -- plot[mark=o, mark options={scale=2}] (0.25,0) -- (0.5,0);}, DID; {\protect\tikz \protect\draw[color={rgb:red,244;green,226;blue,66}] (0,0) -- plot[mark=triangle*, mark options={scale=2,fill=white}] (0.25,0) -- (0.5,0);}, ED; {\protect\tikz \protect\draw[color={rgb:red,0;green,5;blue,1}] (0,0) -- plot[mark=+, mark options={scale=2}] (0.25,0) -- (0.5,0);}, MC-NNM; {\protect\tikz \protect\draw[color={rgb:red,66;green,200;blue,244}] (0,0) -- plot[mark=x, mark options={scale=2}] (0.25,0) -- (0.5,0);}, RVAE; {\protect\tikz \protect\draw[color={rgb:red,66;green,107;blue,244}] (0,0) -- plot[mark=diamond, mark options={scale=2}] (0.25,0) -- (0.5,0);}, SCM; {\protect\tikz \protect\draw[color={rgb:red,244;pink,66;blue,223}] (0,0) -- plot[mark=triangle, mark options={scale=2, rotate=180}] (0.25,0) -- (0.5,0);}, VT-EN.\label{germany-sim}}
\caption{Pre-period densities of log per-capita state government education spending by treatment status:{\protect\tikz \protect\draw[color=black] (0,0) -- plot[mark=square, mark options={scale=2, fill=white}] (0.25,0) -- (0.5,0);}, Control; {\protect\tikz \protect\draw[color={rgb:red,104;green,122;blue,255}] (0,0) -- plot[mark=square*, mark options={scale=2,fill={rgb:red,104;green,122;blue,255}}] (0.25,0) -- (0.5,0);}, Treated \label{educ-dense}}
\caption{Placebo tests on education spending data: {\protect\tikz \protect\draw[color={rgb:red,4;green,0;yellow,1}] (0,0) -- plot[mark=o, mark options={scale=2}] (0.25,0) -- (0.5,0);}, DID; {\protect\tikz \protect\draw[color={rgb:red,244;green,226;blue,66}] (0,0) -- plot[mark=triangle*, mark options={scale=2,fill=white}] (0.25,0) -- (0.5,0);}, ED; {\protect\tikz \protect\draw[color={rgb:red,0;green,5;blue,1}] (0,0) -- plot[mark=+, mark options={scale=2}] (0.25,0) -- (0.5,0);}, MC-NNM; {\protect\tikz \protect\draw[color={rgb:red,66;green,200;blue,244}] (0,0) -- plot[mark=x, mark options={scale=2}] (0.25,0) -- (0.5,0);}, RVAE; {\protect\tikz \protect\draw[color={rgb:red,66;green,107;blue,244}] (0,0) -- plot[mark=diamond, mark options={scale=2}] (0.25,0) -- (0.5,0);}, SCM; {\protect\tikz \protect\draw[color={rgb:red,244;pink,66;blue,223}] (0,0) -- plot[mark=triangle, mark options={scale=2, rotate=180}] (0.25,0) -- (0.5,0);}, VT-EN. \label{educ-sim}}
\caption{\label{fig:curves} Top-1 and top-5 validation accuracy of ST-ResNet-18 on ImageNet as a function of the number of multiplications, the number of additions, and model size, along with the values obtained in related works BWN \cite{rastegari2016xnor}, TWN \cite{li2016ternary}, TTQ \cite{zhu2016trained}, ABC-Net-1/2/3/5 \cite{lin2017towards} (``+'' signs, the suffix reflects the ranking according to accuracy), and the full-precision model (FP). The numbers associated with the marker types correspond to the ratio of the number of hidden SP units and output channels, $r/\cout$. Different colors indicate different combinations of output patch size $p$ and number of convolution groups $g$: {\color{\colorb}Blue: $p=2$, $g=1$}; {\color{\colora}green: $p=1$, $g=1$}; {\color{\colorc} red: $p=1$, $g=4$}. Selected models trained with KD are shown with filled markers.}
\caption{A simplified schematic of the electrical set up. Function Generator 1 (FG1) is used to drive the main electrodes via an amplifier (Amplifier 1) and a helical resonator. The phase at the output of the resonator can be gauged using the monitor coil (mc). The two channels of a dual function generator (DFG) \red{are} used to apply additional rf voltages to the two rf radial electrodes. Only one radial electrode (RE) \red{is shown} for simplicity. Another \red{dual} function generator (FG4) is used to apply \red{a differential} rf signal to the inner electrode\red{s} of the electrode \red{assemblies}. An oscilloscope monitors the phase of the various signals. FG2, FG3, \red{FG4 and FG5} are clock-synchronised to FG1. See the main text for further description. }
\caption{(a) Normalised micromotion amplitude as a function of the phase of FG3. The solid line is a linear fit to the data. Also shown in the right column are the correlated fluorescence counts at the individual points (i), (ii) and (iii). The micromotion amplitude is deduced by the sinusoidal fit at the trap frequency as shown in the red lines. The change of the polarity is clearly visible between (i) and (iii). (b) Normalised micromotion amplitude as a function of the \red{common} phase of FG4 \red{and FG5}, detected with the cavity emission counts at the SPCM. \red{The error bars are the sinusoidal fitting errors of the correlated fluorescence counts.}}
\caption{The cavity emission as the ion is being moved along the standing wave of the cavity mode. \red{The x-axis is the differential voltage applied to the upper and lower inner electrodes as probed by the oscilloscope (see Fig.} \ref{fig:MovingIonWithRFSetup_JMO}\red{). Note that this voltage is not of the same magnitude as the actual voltages on the inner electrodes due to inline electrical losses.} The error bars are standard deviations from \red{10} measurements. The separation of the two anti-nodes correspond to a spatial separation of 433 nm as indicated by the red arrow.}
\caption{Diagram illustrating the training procedure for GibbsNet. The \textbf{unclamped chain} (\textcolor{black!50}{dashed box}) starts with a sample from an isotropic Gaussian distribution $ \mathcal{N}(0, I) $ and runs for $ N $ steps. The last step (iteration $N$) shown as a \textcolor{myred}{solid pink box} is then compared with a single step from the \textbf{clamped chain} (\textcolor{mygreen}{solid blue box}) using joint discriminator $ D $.}
\caption{ Overview of trajectories of various fluid elements in $R=0.3-0.6 \unit{kpc}$. Each line indicates the track of the fluid parcels from $t =399.5\unit{Myr}$ to $404.5\unit{Myr}$. \pink{Note that those lines are different from the configuration of magnetic field lines.} Colors denote vertical velocity; redder colors correspond to upward flows and bluer colors correspond to downward flows. This figure clearly shows ubiquitous vertical flows. The \pink{boxed} region shows the Region X. }
\caption{ Time evolution of vertical velocity of different fluid elements. \pink{ Labels 'a','b' and 'c' correspond to the trajectories shown in Figure \ref{fig_traject}. } \pink{ Since we focus on flows in the upper hemisphere, positive (negative) $v_z$ corresponds to upward (downward) flows. } }
\caption{\pink{ The time evolution of MF1. The red dashed line denotes the original position at $t=401.0\ \unit{Myr}$. The gray line is MF1 at $t=401.15\ \unit{Myr}$ transported along velocity field of the gas from the red dashed line %assumed the magnetic field line is frozen completely. The magenta line is the magnetic field line tracked from the loop top region (arrow) of the gray line. % denotes the magnetic field line drawn from snap data at $t=401.15\ \unit{Myr}$ across the same point (the arrow shown) with gray line. The lines on the $x-y$ plane denote the projection of each line. } }
\caption{$G_{\textrm{ind}}^{-1}$ as a function of {\color{red} the matching scale }$X_{0}$. The values of $X_0$ are in units of GeV$^{-1}$, and the vertical axis is in units of GeV$^2$.}
\caption{ \label{fig:fkmEval500MatlabUCAT} Pseudotransparency plots of performance measures relating to the unimodal category for the family \eqref{eq:fkm} depicted in Figure \ref{fig:fkm} with $n = 500$ and using a Gaussian kernel. From left to right, we show empirical distributions of $\text{ucat}$ {\color{blue}(blue) for TDE}, $\text{ucat}$ {\color{red}(red) for CV}, and the empirical probability that the estimate of $\text{ucat}$ is correct. From top to bottom, we show $k = 1, \dots 3$. Each panel has $m = 1, \dots 10$ along the horizontal axis.}
\caption{\textcolor{red}{Elaborate this.}Agreement of the CNNs composing the ensemble (first row patient 11.).}
\caption{The closed trajectories in the unsteady flow for $Re = 80$ (solid lines) and $250$ (dashed lines). Orbits 1, 2 \& 3 are explicitly labelled for$Re = 250$. Topologically similar three closed trajectories are found for all $Re\in\left[50,300\right]$. The uniform free-stream veloctiy $U_\infty$ is along the positive $x-$axis.\label{fig:unsteady_flow_closed_traj}} \end{center} \end{figure} For every $Re\in[50,300]$, we find three distinct closed fluid particle trajectories whose time period $T$ is the same as that of the flow. In figure~\ref{fig:unsteady_flow_closed_traj}, we show the three trajectories for $Re = 250$ (dashed lines), which are labelled as orbits 1, 2 and 3. Topologically similar orbits for $Re = 80$ are also shown in figure~\ref{fig:unsteady_flow_closed_traj} (solid lines). Orbits 1 \& 2, which are symmetric counterparts of each other but with a time lag of $T/2$, are non-self-intersecting, whereas orbit 3 intersects itself on the centreline. With increasing $Re$, the three orbits move closer to the cylinder and simultaneously increase in their spatial extent. Previous studies by \citet{giannetti2015wkbj} identified the same three closed orbits, but only for $Re = 190$ \& $260$. As mentioned in section~\ref{sec:floq}, for all the three orbits at a given $Re$, only purely transverse wave vectors are periodic with the time period $T$, owing to which we compute growth rates only for $\boldsymbol{k_i} = \boldsymbol{\hat{e}_z}$. Furthermore, orbits 1 \& 2 correspond to the same growth rates, and hence we plot all results only for orbits 1 \& 3. \subsection{Inviscid growth rates}\label{sec:inviscid_growth_rates} %\subsection{Inviscid Growth Rates} \label{sec:unsteady_growth_rates} \begin{figure} \begin{center} \includegraphics[width=1\textwidth,angle=0]{unsteady_growthrates_inv_upd_temp11_cropped.eps} \caption{$(a)$ Inviscid growth rate $\sigma_0^{re}$ and $(b)$ the imaginary part $\sigma_0^{im}$ of the complex growth rate, for orbits 1 \& 2 (solid line) and orbit 3 (dashed line) plotted as a function of $Re$ for purely transverse perturbations. Corresponding values from \citet{giannetti2015wkbj} for $Re=190$ \& $260$ are indicated by {\scriptsize$\bigcirc$} (orbits 1 \& 2) and {\normalsize$\triangle$} (orbit 3). The markers on the curves indicate the actual values of $Re$ at which the growth rate calculations were performed. \label{fig:unsteady_flow_growth_rates}} \end{center} \end{figure} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% In figure~\ref{fig:unsteady_flow_growth_rates}$(a)$, we plot the inviscid growth rate $\sigma_0^{re}$ for orbit 1 (solid line) and orbit 3 (dashed line) as a function of $Re$. For orbit 1, $\sigma_0^{re}$ is zero at $Re = 50$, which is immediately followed by a bifurcation as indicated by the positive $\sigma_0^{re}$ at $Re = 60$. For $Re\lesssim190$, $\sigma_0^{re}$ increases monotonically, attaining a local maximum of $\sigma_0^{re} \approx 0.8$ at $Re \approx 190$, before decreasing to zero at $Re \approx 250$. Remarkably, orbit 1 is stable for a small range of $Re$ around $Re = 250$, above which it again becomes unstable with $\sigma_0^{re}$ increasing to 1 at $Re = 300$. The two bifurcations associated with orbit 1, immediately below and above $Re = 250$, are discussed in more detail later in this section. In contrast, orbit 3 is unstable in the entire range of $\left[50,300\right]$, with its $\sigma_0^{re}$ monotonically increasing with $Re$. The instability associated with orbit 1 slightly dominates over that of orbit 3 in the range $70\lesssim Re\lesssim157$, above which the instability of orbit 3 clearly dominates up to $Re = 300$. Also included in figure~\ref{fig:unsteady_flow_growth_rates}$(a)$ are the growth rate estimates of \citet{giannetti2015wkbj} at $Re = 190$ \&$260$, which are in reasonable agreement with our results. The difference between our computed growth rates and those of \citet{giannetti2015wkbj} may be attributed to differences in the respective numerical schemes used to generate the base flows. We recall here that the imaginary part of the complex growth rate is given by $\sigma_0^{im}=\Im(\log E_m)/T$, where $E_m$ is the eigenvalue that corresponds to the inviscid growth rate. In figure~\ref{fig:unsteady_flow_growth_rates}$(b)$, we plot $\sigma_0^{im}$ as a function of $Re$ for orbits 1 \& 3. For orbit 1,$\sigma_0^{re}>0$ and $\sigma_0^{im}=\pi/T$ ($T$ is the time period of the closed trajectory) in the range $60\le Re\le248$, indicating that the corresponding eigenvalue $E_m$ is real and less than -1. For $251\le Re\le300$, $\sigma_0^{re}>0$ with $\sigma_0^{im} = 0$, i.e. the corresponding eigenvalue $E_m$ is real and greater than 1. The instability on orbit 1 is therefore seen to switch from being asynchronous ($\sigma_0^{im}\ne 0$) for $Re\le248$ to synchronous ($\sigma_0^{im} = 0$) for $Re\ge251$. In physical terms, a synchronous instability has the same time period as that of the base flow. In contrast, at $Re = 260$, \citet{giannetti2015wkbj} report $\sigma_0^{im}=\pi/T$, suggesting that $E_m$ is real and less than -1. We recall, however, from figure~\ref{fig:unsteady_flow_growth_rates}$(a)$ that $\sigma_0^{re}$ for orbit 1 from our calculations is in good quantitative agreement with that of \citet{giannetti2015wkbj} at $Re = 260$. We are unable to identify the source of the discrepancy in $\sigma_0^{im}$ for orbit 1 as \citet{giannetti2015wkbj} report their results only at $Re = 190$ \&$Re = 260$. For orbit 3, the eigenvalue $E_m$ is real and positive at all $Re$, resulting in $\sigma_0^{im}$ being zero at all $Re$. To investigate the bifurcations in the stability on orbit 1 at $Re\approx50$ and $Re\approx250$, we track the evolution of the non-trivial Floquet exponents $\sigma_1$ and $\sigma_2$ with $Re$; the trivial Floquet exponent is $\sigma_3 = 0$. $\sigma_1$ and $\sigma_2$ are defined such that $\Re(\sigma_1)\ge \Re(\sigma_2)$. At $Re =50$, $\sigma_1 = \sigma_2^*$ ($^*$ denotes complex conjugate) with $\Re(\sigma_1)<0$ and $\Re(\sigma_2)<0$, whereas at $Re=60$, $\sigma_1 = -\sigma_2^*$ with $\Im(\sigma_1)=\Im(\sigma_2) = \pi/T$. In other words, the stability characteristics on orbit 1 switch from being stable-focus-like at $Re = 50$ to unstable-saddle-like at $Re=60$, but with non-zero imaginary parts in the Floquet exponents. The stability property on orbit 1 remains unstable-saddle-like, with $\Im(\sigma_1)\ne 0$ and $\Im(\sigma_2)\ne 0$, for $Re<249$. For the two bifurcations at around $Re = 250$, we plot $\sigma_1$ and $\sigma_2$ on the complex plane as $Re$ is varied in the neighbourhood of $Re = 250$ (figure~\ref{fig:eigenvalue_variation_around_250}). For $Re = 245$ and 248, $\Re(\sigma_1)>0$ and $\Re(\sigma_2)<0$, with both the exponents having the same positive imaginary part. The exponents then switch to being complex conjugates with negative real parts at $Re = 249$. This bifurcation from unstable-saddle-like behaviour (with $\Im(\sigma_1)\ne 0$ and $\Im(\sigma_2)\ne 0$) to stable-focus-like behaviour is the reverse of what occurs at $Re\approx50$. The Floquet exponents remain as complex conjugates with $\Re(\sigma_1)<0$ and $\Re(\sigma_2)<0$ for $Re = 249.5$, 250 \& 250.5. We then observe a switch to Floquet exponents with$\sigma_1 = -\sigma_2^*$ at $Re = 251$, with $\Im(\sigma_1)=\Im(\sigma_2)=0$. This bifurcation is therefore from stable-focus-like to unstable-saddle-like properties. For all $Re\ge 251$, the instability property on orbit 1 remains saddle-like with $\sigma_1 = -\sigma_2^*$ and $\Im(\sigma_1)=\Im(\sigma_2)=0$. \begin{figure} \begin{center} \includegraphics[width=0.9\textwidth,angle=0]{complex_growthrates_on_complex_plane_cropped.eps} \caption{Path traced by the non-trivial Floquet exponents $\sigma_1$ (solid line) and $\sigma_2$ (dashed line) on the complex plane as $Re$ is varied in the vicinity of $Re=250$, with $\sigma_1$ and $\sigma_2$ defined such that $\Re(\sigma_1)\ge\Re(\sigma_2)$. Indicated right next to every datapoint is the corresponding value of $Re$.\label{fig:eigenvalue_variation_around_250}} \end{center} \end{figure} \subsection{Viscous growth rates}\label{sec:viscous_growth_rates} To explore the relation between our results and existing knowledge on secondary instabilities in the cylinder wake, we incorporate finite-wavenumber, finite-$Re$ effects in the following manner~\citep{gallaire2007three} \begin{equation}\label{eq:gr_corr} \sigma_{\nu}(\beta,Re) = \sigma_0(Re) - \frac{\beta^2}{Re} - \frac{A(Re)}{\beta}, \end{equation} where $\sigma_{\nu}$ is the corrected growth rate, $\sigma_0$ the inviscid growth rate from the local stability calculations, $\beta$ the transverse wave number, and $A$ a model parameter. The first correction term, $\beta^2/Re$, in equation~\ref{eq:gr_corr} follows from the study of~\citet{landman1987strainedvortices} that showed that weak viscous effects always serve to suppress the inviscid local instabilities. The second correction term, $A/\beta$, in equation~\ref{eq:gr_corr} is based on a previous study~\citep{bayly1988centrifugal} that proposed the construction of localized eigenmodes from local stability calculations. The results of \citet{bayly1988centrifugal} were specific to centrifugal instability on a streamline with locally maximum inviscid growth rate in a steady flow, implying that the validity of equation~\ref{eq:gr_corr} for a closed trajectory in an unsteady flow is unknown. However, equation~\ref{eq:gr_corr} has been employed for centrifugal and non-centrifugal-type instabilities, which don't necessarily satisfy all the assumptions of \citet{bayly1988centrifugal}, with reasonable accuracy~\citep{sipp1999vortices,gallaire2007three}. This suggests that equation~\ref{eq:gr_corr} may represent a reasonably accurate generic model for finite-$\beta$, finite-$Re$ corrections. The model parameter, $A$ in equation~\ref{eq:gr_corr} is assumed to be independent of $\beta$ \citep{gallaire2007three} and a function of the base flow Reynolds number $Re$ only. To estimate $A(Re)$, we use inputs from known secondary instability characteristics. The self-sustained mode-B secondary instability has a characteristic span-wise wavelength of around 1$D$, making it more likely to be captured in a short-wavelength framework than mode-A (characteristic span-wise wavelength $\approx 4D$). Furthermore, global mode Floquet analysis reveals the emergence of an instability at $Re \approx 260$ with a dominant mode of span-wise wavelength around $0.822D$ \citep{barkley1996three}, and is associated with the mode-B instability observed in experiments and three-dimensional numerical simulations. Therefore, we explore the possible connection between our local instability calculations and the mode-B secondary instability, specifically focusing on the relevance of the bifurcation in the instability of orbit 1 at $Re\approx250$. \begin{figure} \begin{center} \includegraphics[width=0.9\textwidth,angle=0]{corr_growth_rates_2x2_cropped.eps} \caption{$(a)$ The maximum growth rate $\sigma_{\nu}^\ast=\sigma_0-3\beta^{\ast 2}/Re$ (with $\beta^\ast = 7.64$) as a function of $Re$ (solid line with markers) for orbits 1 and 2. The inviscid growth rate $\sigma_0$ for orbit 1 is reproduced here for comparison (solid line with no markers). $(b)$ The corrected growth rate $\sigma_\nu$ (equation~\ref{eq:gr_corr}) as a function of $\beta$ using two methodologies M1 and M2 for orbit 1 at $Re=280$; for comparison, the global stability growth rates from~\citet{barkley1996three} are plotted using open circles. $(c)$ and $(d)$: Similar data as in $(a)$ and $(b)$, respectively is plotted for oribt 3. \label{fig:corr_growth_rates}} \end{center} \end{figure} \citet{barkley1996three} report that the most unstable perturbation mode at the birth of the new instability at $Re \approx 260$ corresponds to a span-wise wavenumber of $\beta = \beta^* \approx 7.64$. While the exact values of $\beta^\ast$ are not reported for $Re> 260$ in \citet{barkley1996three}, the most unstable wavenumber is known to be around the same value as that for $Re = 260$. For $\beta = \beta^*$ to be the most unstable transverse wavenumber in the local stability analysis, i.e. for $\sigma_\nu$ (equation \ref{eq:gr_corr}) to attain a maximum at $\beta = \beta^*$, we require \begin{equation}\label{eq:beta_max_law} A=\frac{2\beta^{*3}}{Re}, \end{equation} thus providing us with an estimate for $A$. The resulting maximum growth rate $\sigma_\nu(\beta^*,Re) = \sigma_0(Re) - 3\beta^{\ast 2}/Re$ is denoted as $\sigma_\nu^*$. In figure~\ref{fig:corr_growth_rates}$(a)$, we plot $\sigma_\nu^*$ (solid line with markers) as a function of $Re$ for orbit 1. Strikingly, based on $\sigma_\nu^*$, orbit 1 is stable for all $Re\lesssim262$. This critical $Re$ of 262 above which orbit 1 is unstable is close to the mode-B critical $Re$ of 260 reported by~\citet{barkley1996three}. Additionally, as noted in figure~\ref{fig:unsteady_flow_growth_rates}$(b)$, the instability on orbit 1 is synchronous ($\sigma_0^{im}=0$) for $Re\ge 262$, and hence leading us to suggest that it is indeed closely related to the synchronous mode-B secondary instability. To further explore the relation between the local instability on orbit 1 and the mode-B secondary instability, we evaluate the extent to which equation~\ref{eq:gr_corr} captures the growth rate variation with $\beta$. In figure~\ref{fig:corr_growth_rates}$(b)$, we plot $\sigma_\nu$ for orbit 1 as a function of $\beta$ at $Re = 280$ (solid line), with the value of $A$ chosen based on equation~\ref{eq:beta_max_law}; such a choice for $A$ is denoted as methodology M1. The solid line, representing methodology M1, in figure~\ref{fig:corr_growth_rates}$(b)$ is in reasonable qualitative agreement with the growth rate variation based on the results of \citet{barkley1996three} (solid line with markers). Specifically, the range of unstable $\beta$ based on $\sigma_\nu$ is $[3.72, 13.42]$, whereas the unstable range reported by \citet{barkley1996three} is $[6.23,10.2]$. In figure~\ref{fig:corr_growth_rates}$(b)$, we also plot $\sigma_\nu$ vs. $\beta$ using an alternate method M2 to estimate $A$, where we match $\sigma_\nu(\beta^*)$ with the corresponding value of growth rate at $\beta^\ast$ from \citet{barkley1996three}. While methodology M2 does not guarantee that $\sigma_\nu$ attains a maximum at $\beta = \beta^*$, we find remarkable qualitative and quantitative agreement with \citet{barkley1996three}. Based on methodology M2, the range of unstable $\beta$ is $[5.87, 11.84]$. In summary, the results from figures~\ref{fig:corr_growth_rates}$(a)$ \&$(b)$ indicate that the mode-B secondary instability is possibly a manifestation of the local instability on orbits 1 \& 2. To investigate the effects of finite-$\beta$, finite-$Re$ corrections on the local instability identified on orbit 3, we plot $\sigma_\nu^*$ as a function of $Re$ in figure~\ref{fig:corr_growth_rates}$(c)$, with $A$ again estimated using $\beta^* = 7.64$. Based on the corrected growth rate, orbit 3 is stable for all $Re$ less than the critical value of $Re\approx 190$, a value of Reynolds number that lies in the so-called transition regime \citep{williamson1996vortex}. However, $Re = 190$ represents the critical Reynolds number of the mode-A secondary instability, whose characteristic transverse wavelength is $4D$. Therefore, it is not clear what the implications of the transition at $Re = 190$ in figure~\ref{fig:corr_growth_rates}$(c)$ are. We also note that orbit 3 is of noticeably smaller spatial extent than that of orbit 1, and also contains segments where its radius of curvature is negligibly small, potentially raising questions on whether a span-wise wavelength of $1D$ can be considered ``short-wavelength" for orbit 3. In a manner similar to figure~\ref{fig:corr_growth_rates}$(b)$, we plot $\sigma_\nu$ as a function of $\beta$ (for both methodologies M1 and M2) at $Re=280$ for orbit 3, along with the mode-B instability growth rate variation reported by \citet{barkley1996three}. For both methodologies M1 and M2, a significantly larger range of beta is unstable when compared with the estimate from global analysis. This further suggests that the local instability on orbit 3, probably, has no relation to the mode-B secondary instability. Finally, we performed similar calculations as in figure~\ref{fig:corr_growth_rates} to explore the relation between the local instabilities and the mode-A secondary instability. The span-wise wavelength associated with the mode-A secondary instability is around $4D$ ($\beta \approx 1.59$), and hence not necessarily a ``short-wavelength instability". Assuming a $\beta^*$ of 1.59, finite-$Re$, finite-$\beta$ corrections do not significantly modify the inviscid growth rates on orbits 1 or 3. Therefore, our study is inconclusive about the role of local instabilities in the mode-A secondary instability. The related study by \citet{giannetti2015wkbj} performed similar inviscid local stability calculations on the three closed trajectories for $Re = 190$ \& 260. The inviscid instability on orbit 1 was reported to be asynchronous for both$Re$, leading \citet{giannetti2015wkbj} to suggest that it is related to the mode-C secondary instability. Orbit 3 was found to display a synchronous instability, and was linked with both modes A \& B secondary instabilities. Our study, however, finds that the inviscid instability on orbit 1 undergoes an asynchronous-to-synchronous bifurcation at$Re\approx 250$. Additionally, with finite-wavenumber, finite-$Re$ corrections, the synchronous instability on orbit 1 is shown to occur for $Re\gtrsim262$ only. Furthermore, we also find the mode B span-wise wavelength of around $1D$ to be consistent with the local instability on orbit 1, but not on orbit 3. \section{Conclusions}\label{sec:concl} In this paper, we have presented a local stability analysis in the near-wake region resulting from the uniform flow past a two-dimensional circular cylinder for Reynolds numbers in the range $50\le Re\le300$. The inviscid local stability equations were solved on closed fluid particle trajectories in the unsteady flow for purely transverse perturbations. Three closed trajectories with the time period the same as that of the base flow were identified for all $Re\in\left[50,300\right]$. Two of these closed trajectories, denoted orbits 1 \& 2, are non-self-intersecting and symmetric counterparts of each other. The third closed trajectory, referred to as orbit 3, is self-intersecting. The inviscid growth rate associated with orbits 1 \& 2 undergoes bifurcations at $Re\approx 50$ and $Re\approx 250$, with the instabilities in the ranges $50\lesssim Re \lesssim248$ and $251\lesssim Re\le300$ being asynchronous and synchronous, respectively. In contrast, orbit 3 was found to be inviscidly unstable in the entire range of $\left[50,300\right]$, with the corresponding growth rate increasing monotonically with $Re$. Finite-wavenumber, finite-Reynolds number corrections on the computed inviscid growth rates were then obtained by assuming that the most unstable perturbation mode occurs for the mode-B span-wise wavelength of $0.822D$. Based on this corrected growth rate, the inviscid instability on orbits 1 \& 2 is suppressed for$Re\lesssim262$, and a synchronous instability occurs for $Re>262$. This transition Reynolds number of $Re\approx262$ is remarkably close to the mode B critical Reynolds number from global stability analysis, experiments and numerical simulations. Additionally, the corrected growth rate variation with the span-wise wavenumber for $Re=280$ shows excellent qualitative and quantitative agreement with the corresponding mode B instability growth rates from the global analysis. These results strongly suggest that the three-dimensional, short-wavelength instability on orbit 1 is closely connected to the mode B secondary instability in the wake of a circular cylinder. The physical relevance of the inviscid local instability on orbit 3 is unclear. In the future, it may be worth designing direct numerical simulations with the aim of studying the evolution of localized three-dimensional perturbations on the closed trajectories in the two-dimensional cylinder wake, which may lead to devising flow control strategies. Also, it would be interesting to explore the existence of other closed trajectories in the time-periodic cylinder wake, whose time period may be any integer multiple of the flow time period. We thank S. Ajith Kumar for his help with the numerical simulations of the base flow in the early phase of this study. \bibliography{paperPRF_refs} \end{document} % % ****** End of file apssamp.tex ****** }
\caption{ Quantitative evaluation of state-of-the-art SR algorithms: average PSNR/SSIM for scale factors $2\times$, $4\times$ and $8\times$. \red{\textbf{Red}} text indicates the best SSIM score and \blue{\underline{blue}} text indicates the second best performance of SSIM score. }
\caption{Illustration of topological operations on a heterogeneous table (a) consisting of numerical (\protect\includegraphics[height=0.7em]{figs/hashtag.pdf}) and categorical (\protect\includegraphics[height=0.7em]{figs/category.pdf}) attributes and their results reflected in the aggregation hierarchy: sorting (b), filtering (c), grouping by a single categorical attribute (d), grouping by the Cartesian product of two categorical attributes (e), and aggregating (f).}
\caption{\textbf{Complementary polysaccharide prioritization allows robust coexistence in gut \emph{Bacteroides} species.} \\ \textbf{(A)} The polysaccharide utilization network of \emph{Bacteroides} species in the human gut (data taken from\cite{Chia2017}). The character labels represent 9 different polysaccharides: starch (S), mucin (M), galactan (G), pectin (P), arabinogalactan (A), hemicellulose (HC), cellulose (C), hyaluronan (H), chondroitin sulfate (CS) --- known to be frequently present in human diets (legend in the box on the left), whereas the colored circles represent 7 different \emph{Bacteroides} species routinely found in human gut microbiome: \emph{Bacteriodes fragilis}, \emph{B. ovatus}, \emph{B.vulgatus}, \emph{B.caccae}, \emph{B.cellulosilyticus}, \emph{B. thetaiotaomicron}, \emph{Parabacteroides distasonis}. %(legend in the gray box on the top). Undirected links between microbes and polysaccharides indicate a species' ability to metabolize that polysaccharide. %These particular species are known to regularly co-occur in the human gut \cite{Chia2017} at high abundances, without apparent metabolic conflict, even though the sets of metabolites they utilize is densely overlapping. \textbf{(B)} Examples of microbial nutrient preferences (the most preferred nutrient of each of the microbes) are sorted into three categories: complementary (top) where microbes' top preferred nutrients (\#1) are all distinct from each other; random (middle) preferences where all ranked lists are randomly generated;%which represents partial conflict between these top preferences and even second preferences (\#2) for the next layer and competition; and maximal conflict (bottom) which represents the maximum intersection between the sets of top (\#1) and second (\#2) preferred nutrients of different microbes.%(see Methods for additional details of how these lists were generated). \textbf{(C)} For 1,000 randomly sampled microbial preferences from each category, we simulated the stable marriage model to compute the expected per species microbial abundances (see Methods: Studying complementarity through different ranked interaction tables) for each case as box plots. The box plots quantify the distribution of average microbial abundance assumed to be inversely proportional to the rank of utilized nutrient. The average abundance is the largest in the case of complementary nutrient choices, All differences between distributions of abundances in each category are highly statistically significant according to the Kolmogorov-Smirnov test with a $P$-value threshold of $0.01$. }
\caption{Average PSNR and SSIM results for bicubic degradation on datasets Set5~\cite{bevilacqua2012low}, Set14~\cite{zeyde2010single}, BSD100~\cite{MartinFTM01} and Urban100~\cite{huang2015single}. The best two results are highlighted in \textcolor[rgb]{1.00,0.00,0.00}{red} and \textcolor[rgb]{0.00,0.00,1.00}{blue} colors, respectively.}
\caption{Average PSNR and SSIM results of different methods with different degradations on Set5. The best results are highlighted in \textcolor[rgb]{1.00,0.00,0.00}{red} color. The results highlighted in \textcolor[rgb]{0.45,0.45,0.45}{gray} color indicate unfair comparison due to mismatched degradation assumption.}
\caption{Contribution of extension, shear,and bending deformation mechanisms (see Eq.~\ref{eqn2a}) to the effective compliance ($\rightarrow \xi_{\mathrm{mode}}/\xi_{\mathrm{tot}}$ where $\xi_{\mathrm{tot}} = \sum\xi$) of wavy CNTs that comprise the A-PNC as a function of the waviness ratio ($w$) and ratio of the intrinsic CNT longitudinal and shear moduli ($\rightarrow Y/G$) for $Y/G = 1000$ (\protect\tikz{\protect\filldraw[draw=black,fill=black] (0,0) circle [radius=0.1cm];}), $Y/G = 250$ (\protect\tikz{\protect\filldraw[rounded corners=0.01cm,draw=blue,fill=blue] (0,0) rectangle (0.2cm,0.2cm);}), $Y/G = 100$ (\protect\tikz{\protect\filldraw[rounded corners=0.01cm,draw=red,fill=red,rotate=180](0,0)--(0.2cm,0)--(0.1cm,0.2cm)--cycle;}), and $Y/G = 10$ (\protect\tikz{\protect\filldraw[rounded corners=0.01cm,draw=green,fill=green](0,0)--(0.2cm,0)--(0.1cm,0.2cm)--cycle;}). (a) Extension mode contributions. (b) Shear mode contributions. (c) Bending mode contributions.}
\caption{\label{fig:tref}Simulated fraction of the deposited energy in bulk tungsten for various incident pion energies as a function of \tred.}
\caption{Asymptotic power of the knockoff procedure as compared to the oracle, for two different distributions $\Pi^*$: point mass (left) and exponential (right). $\epsilon = 0.2, \delta = 1, \sigma = 0.5$. {\it Top}: Theoretical prediction. Each of the points marked on the top graphs corresponds to a particular value of $q$, and was generated as follows. (i) Fix $q$; (ii) find $\lambda$ and $t'$ s.t. $\fdpinfty(t) = q$ and $\fdphatinftyprime(t') = q$, where $\fdpinfty$ and $\fdphatinftyprime$ are given in \eqref{eq:fdp-infty} and \eqref{eq:fdp-hat-infty-pi-naught-hat}; (iii) plot the point $\left( \tppinfty(t),\tppinftyprime(t') \right)$, where $\tppinfty$ and $\tppinftyprime$ are given in \eqref{eq:tpp-infty} and \eqref{eq:tpp-infty-prime}. The broken line is the 45 degrees line. The knockoff procedure adapts to an unknown $\Pi^*$: in both situations the asymptotic power (TPP) of the knockoffs procedure is very close to that of the oracle procedure. {\it Bottom}: Simulation results, $\epsilon = 0.2$; $n=p=5000$; $\sigma = 0.5$. Each of the two graphs is meant to be an analogue of the asymptotic limits above. } \label{fig:intro} \end{figure} While the main focus is on the testing problem, we address also the issue of model selection and prediction. Motivated by the work of \citet{abramovich2006special}, which formally connected FDR control and estimation in the sequence model, our aim is to investigate whether favorable properties associated with FDR control carry over to the linear model. As \citet{abramovich2006special} remark at the outset, in the special case of the sequence model, a full decision-theoretic analysis of the predictive risk is possible. For the linear model case, \citet{bogdan2015slope} proposed a penalized least squares estimator, SLOPE, {motivated} from the perspective of multiple testing and FDR control, and \citet{su2016slope} even proved that this estimator is asymptotically minimax under the same design as that considered in the current paper, although in a different sparsity regime, namely $\epsilon\to 0$. In truth, the linear model is substantially more difficult to analyze than the sequence model. Fortunately, however, the specific working assumptions we adopt allow to analyze the (asymptotic) predictive risk, again relying on the elegant results of \citet{bayati2012lasso} for the Lasso estimates. While our results lack the rigor of those in, e.g., \citet{abramovich2006special} and \citet{su2016slope}, we think that the findings reported in Section \ref{sec:prediction} are valuable in investigating a problem for which conclusive answers are currently not available. The rest of the paper is organized as follows. Section \ref{sec:tradeoff} reviews some relevant results from \citet{su2017false} and presents the oracle false discovery rate-power tradeoff diagram for the Lasso. In Section \ref{sec:knockoffs} we present a simple testing procedure which uses knockoffs for FDR calibration, and prove some theoretical guarantees. A power analysis, which compares the knockoff procedure to the oracle procedure, is carried out in Section \ref{sec:power}. Section \ref{sec:prediction} is concerned with using knockoffs for prediction, and Section \ref{sec:discussion} concludes with a short discussion. \section{An oracle tradeoff diagram}\label{sec:tradeoff} Adopting the setup from \citet{su2017false}, we consider the linear model \begin{equation}\label{eq:model} \by = \bX\bbeta + \bz \end{equation} in which $\bX\in \RR^{n\times p}$ has i.i.d.~$\calN(0, 1/n)$ entries and the errors $z_i$ are i.i.d.~$\calN(0,\sigma^2)$ with $\sigma\geq 0$ fixed and arbitrary. We assume that the regression coefficients $\beta_j,\j=1,...,p$ are i.i.d.~copies of a random variable $\Pi$ with $\EE(\Pi^2)<\infty$ and $\PP(\Pi\neq 0) = \epsilon\in (0,1)$ for a fixed constant $\epsilon$. In this section we will be interested in the case where $n,p\to \infty$ with $n/p \to \delta$ for a positive constant $\delta$. Note that $\Pi$ is assumed to not depend on $p$, and so the expected number of nonzero elements $\beta_j$ is equal to $\epsilon \cdot p$. \noindent Let $\widehat{\bbeta}(\lambda)$ be the Lasso solution, \begin{equation}\label{eq:lasso} \widehat{\bbeta}(\lambda) = \displaystyle \argmin_{\bb\in \RR^p} \frac{1}{2}\|\by - \bX\bb\|^2 + \lambda \|\bb\|_1, \end{equation} and denote $V(\lambda) = |\{j: \widehat{\beta}_j(\lambda) \neq 0, \beta_j = 0\}|$, $T(\lambda) = |\{j: \widehat{\beta}_j(\lambda) \neq 0, \beta_j \neq 0\}|$, $R(\lambda) = |\{j: \widehat{\beta}_j(\lambda)\neq 0\}|$ and $k = |\{j: \beta_j \neq 0\}|$. Hence $V(\lambda)$ is regarded as the number of false `discoveries' made by the Lasso; $T(\lambda)$ is the number of true discoveries; $R(\lambda)$ is the total number of discoveries, and $k$ is the number of true signals. We would like to remark here that $\beta_j = 0$ implies that $\bX_j$ (the $j$-th variable) is independent of $\by$ marginally {\it and} conditionally on any subset of $X_{-j}:=\{X_1,...,X_{j-1},X_{j+1},...,X_p\}$; hence the interpretation of rejecting the hypothesis $\beta_j = 0$ as a false discovery is clear and unambiguous. The false discovery proportion (FDP) is defined as usual as \begin{equation*} \tFDP(\lambda) = \frac{V(\lambda)}{1\vee R(\lambda)} \end{equation*} and the true positive proportion (TPP) is defined as \begin{equation*} \tTPP(\lambda) = \frac{T(\lambda)}{1\vee k}. \end{equation*} \citet{su2017false} build on the results in \citet{bayati2012lasso} to devise a tradeoff diagram between TPP and FDP in the linear sparsity regime. Lemma A.1 in \citet{su2017false}, which is adopted from \citet{bogdan2013supplementary} and is a consequence of Theorem 1.5 in \citet{bayati2012lasso}, predicts the limits of FDP and TPP at a fixed value of $\lambda$. Throughout, let $\eta_t(\cdot)$ be the soft-thresholding operator, given by $\eta_{t}(x) = \mathrm{sgn}(x)(|x|-t)_+$. Also, let $\Pi^*$ denote a random variable distributed according to the conditional distribution of $\Pi$ given $\Pi \neq 0$; that is, \begin{equation} \Pi = \begin{cases} \Pi^*, &\text{w.p.} \ \epsilon, \\ 0, &\text{w.p.} \ 1-\epsilon. \end{cases} \end{equation} Finally, denote by $\alpha_0$ the unique root of the equation $(1+t^2)\Phi(-t) - t\phi(t) = \delta/2$. \begin{lemma}[Lemma A.1 in \citealp{su2017false}; Theorem 1 in \citealp{bogdan2013supplementary}; see also Theorem 1.5 in \citealp{bayati2012lasso}]\label{lem:infty} The Lasso solution with a fixed $\lambda>0$ obeys \begin{equation*} \begin{aligned} \frac{V(\lambda)}{p}&\stackrel{\PP}{\to}2(1-\epsilon)\Phi(-\alpha),\\ \frac{T(\lambda)}{p}&\stackrel{\PP}{\to}\PP(|\Pi + \tau W|>\alpha\tau, \Pi \neq 0) = \epsilon \PP(|\Pi^* + \tau W|>\alpha\tau), \end{aligned} \end{equation*} where $W\sim \calN(0,1)$ independently of $\Pi$, and $\tau>0,\\alpha >\max\{\alpha_0, 0\}$ is the unique solution to \begin{equation}\label{eq:system} \begin{aligned} \tau^2 &= \sigma^2 + \frac{1}{\delta}\EE(\eta_{\alpha\tau}(\Pi + \tau W) - \Pi)^2 \\ \lambda &= \left( 1 - \frac{1}{\delta}\PP(|\Pi + \tau W|>\alpha\tau) \right) \alpha\tau. \end{aligned} \end{equation} \end{lemma} As explained in \citealp{su2017false}, Lemma \ref{lem:infty} is a consequence of the fact that, under the working assumptions, $(\bbeta, \widehat{\bbeta}(\lambda))$ is in some (limited) sense asymptotically distributed as $(\bbeta, \eta_{\alpha\tau}(\bbeta + \tau\mathbf{W}))$, where $\mathbf{W} \sim\calN_p(\mathbf{0}, \mathbf{I})$ independently of $\bbeta$; here, the soft-thresholding operation acts on each component of a vector. %(here $\alpha$ and $\tau$ are the solution to \eqref{eq:system}). It follows immediately from Lemma \ref{lem:infty} that for a fixed $\lambda > 0$, the limits of FDP and TPP are \begin{equation}\label{eq:fdp-infty} \tFDP(\lambda) \stackrel{\PP}{\longrightarrow} \frac{2(1-\epsilon)\Phi(-\alpha)}{2(1-\epsilon)\Phi(-\alpha) + \epsilon \PP(|\Pi^* + \tau W|>\alpha\tau)} \equiv \fdpinfty(\lambda) \end{equation} and \begin{equation}\label{eq:tpp-infty} \tTPP(\lambda) \stackrel{\PP}{\longrightarrow} \PP(|\Pi^* + \tau W|>\alpha\tau) \equiv \tppinfty(\lambda). \end{equation} \medskip The parametric curve $(\fdpinfty(\lambda), \tppinfty(\lambda))$ implicity defines the FDP level attainable at a specific TPP level for the Lasso solution with a fixed $\lambda$ as \begin{equation}\label{eq:tradeoff-oracle} q^{\Pi^*}(\tppinfty(\lambda); \epsilon, \delta, \sigma) = \fdpinfty(\lambda). \end{equation} Interpreted in the opposite direction, for known $\Pi^*$ and $\epsilon$, we have an {\it analytical expression} for the asymptotic power of a procedure that, for a given $q$, chooses $\lambda$ in \eqref{eq:lasso} in such a way that $\fdpinfty(\lambda) = q$. We describe how to compute the tradeoff curve $q^{\Pi^*}(\cdot ; \epsilon, \delta, \sigma)$ for a given $\Pi^*$ and $\epsilon$ in the Appendix. Because in practice $\Pi^*$ and $\epsilon$ would usually be unknown, we refer to this procedure as the {\it oracle} procedure. The oracle procedure is not a ``legal'' multiple testing procedure, as it requires knowledge of $\epsilon, \Pi^*$ and $\sigma^2$, which is of course realistically unavailable. Still, it serves as a reference for evaluating the performance of valid testing procedures. Competing with this oracle is the subject of the current paper, and the particular working assumptions above were chosen simply because it is possible under these assumptions to obtain a tractable expression for the empirical distribution of the Lasso solution. Before proceeding, we would like to mention another oracle procedure. Hence, consider the procedure which for $\lambda$ uses the value \begin{equation} \lambda^* = \inf\{\lambda: \tFDP(\lambda)\leq q\}. \end{equation} This can be thought of as an oracle that is allowed to see the actual $\bbeta$ but is still committed to the Lasso path, and stops the last ``time" the FDP is smaller than $q$. From now on, we think of $\lambda$ as decreasing in time, and regard a variable $j$ as entering {\it before} variable $j'$ if $\sup \{\lambda: \widehat{\beta}_{j'}(\lambda)\neq 0\} \leq \sup \{\lambda: \widehat{\beta}_{j}(\lambda)\neq 0\}$. Although this may appear as a stronger oracle in some sense, it is asymptotically equivalent to the ``fixed-$\lambda$" oracle discussed earlier: by Theorem 3 in \citet{su2017false}, the event \begin{equation} \bigcap_{\lambda>0.01} \left\{ \tFDP(\lambda) \geq q^{\Pi^*}(\tTPP(\lambda) - 0.001; \epsilon, \delta, \sigma) \right\} \end{equation} has probability tending to one, hence the lower bound on FDP provided by $q^{\Pi^*}(\cdot; \epsilon, \delta, \sigma)$ holds uniformly in $\lambda$. %provides a lower bound on the attainable TPP for a given FDP level, which applies also to a procedure that uses a {\it data-dependent} value $\tau(\by, \bX)>0$ for $\lambda$. %Since the function $q^{\Pi}(\cdot; \delta, \sigma)$ is monotone, a consequence of Theorem 3 and Lemma A.1 in \citet{su2017false} is as follows. %For fixed $q\in (0,1)$, suppose that $\lambda = \lambda(q)>0$ is such that $\fdpinfty(\lambda) = q$. %Then the Lasso procedure with $\lambda$ has asymptotic FDP equal to $q$ and asymptotic TPP equal to $\tppinfty(\lambda)$. %Furthermore, no procedure that chooses $\lambda$ adaptively as $\tau(\by, \bX)>0$ and achieves limiting TPP equal to $\tppinfty(\lambda(q))$, can have smaller $FDP$ than $q$. %For a fixed $q\in (0,1)$, we will therefore refer to the Lasso procedure with fixed $\lambda = \lambda(q)$ as the {\it oracle} procedure. %Crucially, note that $\lambda(q) = \lambda(q;\Pi)$. \section{Calibration using knockoffs}\label{sec:knockoffs} %The previous section discusses an oracle rule that chooses a fixed value of lambda---which depends on the distribution $\Pi^*$ and on $\epsilon$---to achieve exact asymptotic FDP control, and attains the corresponding power specified through the function $q^{\Pi^*}$. %In practice, we will usually not know $\Pi^*$ and $\epsilon$, and would like to consider procedure that can control $\mathrm{FDR}(\lambda) = \EE(\tFDP(t))$ for finite $n,p$ and without using $\Pi^*, \epsilon$. If we limit ourselves to procedures that use the Lasso with a fixed---or even data-dependent---value of $\lambda$, from the previous section we see that the achievable TPP can never asymptotically exceed that of the oracle procedure. Hence the asymptotic power of the oracle procedure serves as an asymptotic benchmark. In this section we propose a competitor to the oracle, which utilizes knockoffs for FDR calibration. While the procedure presented below is sequential and formally does not belong to the class of Lasso estimates with a data-dependent choice of $\lambda$, it will be almost equivalent to the corresponding procedure belonging to that class. We discuss this further later on. \begin{subsection}{A knockoff filter for the i.i.d.~design}\label{subsec:iid} Let the model be given by \eqref{eq:model} as before. In this section, unless otherwise specified, it suffices to assume that the entries of $\bX$ are i.i.d.~from an arbitrary known continuous distribution $G$ (not necessarily normal). Furthermore, unless otherwise indicated, in this section we treat $n,p$ as fixed and condition throughout on $\bbeta$. {\it The definitions in the current section override the previous definitions, if there is any conflict. } Let $\Xtil \in \RR^{n \times r}$ be a matrix with i.i.d.~entries drawn from $G$ independently of $\bX$ and of all other variables, where $r$ is a fixed positive integer. Next, let \begin{equation*} \mathbb{X} := [\bX \ \Xtil] \in \RR^{n\times (p+r)} \end{equation*} denote the augmented design matrix, and consider the Lasso solution for the augmented design, \begin{equation}\label{eq:lasso-til} \widehat{\bbeta}(\lambda) = \displaystyle \argmin_{\bb\in \RR^{p+r}} \frac{1}{2}\|\by - \mathbb{X} \bb\|^2 + \lambda \|\bb\|_1. \end{equation} For simplicity we keep the notation $\widehat{\bbeta}(\lambda)$ for the solution corresponding to the augmented design, although it is of course different from \eqref{eq:lasso} (for one thing, \eqref{eq:lasso-til} has $p+r$ components versus $p$ for \eqref{eq:lasso}). %From here on, unless otherwise specified, $\widehat{\bbeta}(\lambda)$ will denote the solution to \eqref{eq:lasso-til} and $\widehat{\beta}_j(\lambda),\ j=1,...,p+r$, its components. Let \begin{equation*} \calH = \{1,...,p\},\ \ \ \calH_0 = \{j\in \calH: \beta_j = 0\},\ \ \ \calK_0 = \{p+1,...,p+r\} \end{equation*} be the sets of indices corresponding to the original variables, the null variables and the knockoff variables, respectively. Note that, because we are conditioning on $\bbeta$, the set $\calH_0$ is nonrandom. Now define the statistics \begin{equation}\label{eq:stat} T_j = \sup\{\lambda: \widehat{\beta}_j(\lambda) \neq 0\},\ \ j=1,...,p+r. \end{equation} Informally, $T_j$ measures how early the $j$th variable enters the Lasso path (the larger, the earlier). Let \begin{equation*} V_0(\lambda) = |\{j\in \calH_0: T_j \geq \lambda\}|,\ \ V_1(\lambda) = |\{j\in \calK_0: T_j \geq \lambda\}| \end{equation*} be the number of true null and fake null variables, respectively, which enter the Lasso path ``before" time $\lambda$. Also, let \begin{equation*} R(\lambda) = |\{j\in \calH: T_j \geq \lambda\}| \end{equation*} be the total number of original variables entering the Lasso path before ``time" $\lambda$. A visual representation of the quantities defined above is given in Figure \ref{fig:illustration}. \begin{figure}[H] \centering \includegraphics[width=.75\textwidth]{kf_illustration.pdf} \caption{ Representation of test statistics. Each $T_j$ is represented by a marker: round markers correspond to original variables ($\calH$), triangles represent fakes (knockoffs, $\calK_0$); black corresponds to nonnull ($\calH\setminus\calH_0$), red represents null (true or fake, $\calH_0\cup \calK_0$). %Illustration of the knockoff filter for i.i.d.~design. %Suppose that there are as many knockoffs as original variables ($r=p$), and only the 19 largest statistics are shown (as markers). %Round markers represent original variables, triangles represent fake (knockoffs); black represents nonnull, red represents null (true or fake). %At point $\lambda$, there are $V_0(t) = 3$ (unobservable) true nulls, and $V_1(t) = 2$ (observable) fake nulls, with statistics bigger than $\lambda$. %There is a total of $R(t) = 9$ original variables that are bigger than $\lambda$. %Because the red markers appear in a random order, the FDP can be estimated by $V_1(t)/R(t) = 2/9$. %The actual FDP (unobservable) is $V_0(t)/R(t) = 3/9$. } \label{fig:illustration} \end{figure} The class of procedures we propose (henceforth, knockoff procedures) reject the null hypotheses corresponding to \begin{equation*} \{j \in \calH: T_j\geq \widehat{\lambda}\}, \end{equation*} where \begin{equation}\label{eq:lambda-hat} \widehat{\lambda} = \inf\left\{ \lambda \in \Lambda: \frac{ (1 + V_1(\lambda)) \cdot \frac{|\calH| \widehat{\pi}_0}{1+|\calK_0|} }{R(\lambda)} \leq q\right\} \end{equation} for an estimate $\widehat{\pi}_0$ of $\pi_0 := |\calH_0|/|\calH|$ and for a set $\Lambda$ associated with $\widehat{\pi}_0$. Note that conditionally on $\bbeta$, $\pi_0$ is a (nonrandom) parameter, and in this section we speak of estimating $\pi_0$ rather than $1-\epsilon$. Hence the tests in this class are indexed by the estimate of $\pi_0$ (and by $\Lambda$). %\footnote{while $\pi_0$ is a random variable, our analysis below is conditional on $\calH_0$, in which case $\pi_0$ becomes a parameter.}. Denote by $\mathbb{X}_j, \j=1,...,p+r$, the columns of the augmented matrix $\mathbb{X}$. To see why the procedure makes sense, observe that, conditionally on $\bbeta$, the distribution of $(\mathbb{X}, \by)$ is invariant under permutations of the variables in the set $\{\calH_0\cup \calK_0\}$ (just by symmetry). Therefore, conditionally on $|\{j\in \calH_0\cup \calK_0: T_j \geq \lambda\}| = k$, any subset of size $k$ of $\{\calH_0\cup \calK_0\}$ is equally probable to be the selected set $\{j\in \calH_0\cup \calK_0: T_j \geq \lambda\}$. It follows that \begin{equation*} V_0(\lambda) / |\calH_0| \approx V_1(\lambda) / |\calK_0|. \end{equation*} %With $\widehat{\pi}_0\cdot |\calH|$ being an estimate of $|\calH_0| \approx (1-\epsilon)|\calH|$, Therefore, \begin{equation*} V_0(\lambda) \approx V_1(\lambda) \cdot \frac{|\calH| \pi_0}{|\calK_0|} \approx V_1(\lambda) \cdot \frac{|\calH| \widehat{\pi}_0}{|\calK_0|} \end{equation*} and hence \eqref{eq:lambda-hat} can be interpreted as the smallest $\lambda$ such that an estimate of \begin{equation*} \tFDP(\lambda) \equiv \frac{ V_0(\lambda) }{ 1\vee R(\lambda)} \end{equation*} is below $q$. The fact that the procedure defined above, with $\widehat{\pi}_0 = 1$, controls the FDR, is a consequence of the more general result below. \begin{theorem}[FDR control for knockoff procedure with $\widehat{\pi}_0 = 1$]\label{thm:fdr-control} Let $T_j,\j=1,...,p+r$ be test statistics with a joint distribution that is invariant to permutations in the set $\calH_0\cup \calK_0$ (that is, reordering the statistics with indices in the extended null set, leaves the joint distribution unchanged). Set $V_0(t) = |\{j\in \calH_0: T_j \geq t\}|$, $V_1(t) = |\{j\in \calK_0: T_j \geq t\}|$, and $R(t) = |\{j\in \calH: T_j \geq t\}|$. Then the procedure that rejects $\calH_0^j$ if $T_j\geq \tau$, where \begin{equation}\label{eq:tau} \tau = \inf\left\{ t \in \RR: \frac{ (1 + V_1(t)) \cdot \frac{|\calH| }{1+|\calK_0|} }{1\vee R(t)} \leq q\right\} \end{equation} controls the $\tFDR$ at level $q$. In fact, $$ \EE \left[ \frac{ V_0(\tau) }{ 1\vee R(\tau)} \right] \leq q\frac{|\calH_0|}{|\calH|}. $$ \end{theorem} \begin{proof} We will use the following lemma, which is proved in the Appendix. \begin{lemma}\label{lem:hyper} Let $X\sim \text{Hyper}(n_0, n_1; m)$ be a hypergeometric random variable, i.e., $$ \PP(X = k) = \frac{{n_0 \choose k} {n_1\choose m-k}}{{n_0+n_1 \choose m}} $$ with $m>0$ and set $Y = X/(1+m-X)$. Then\footnote{Here and below we adopt the convention that ${0\choose 0} = 1$, and ${n\choose k} = 0$ if $n<0$ or if either $k<0$ or $k>n$.} $$ \EE Y = \frac{n_0}{1+n_1} \left( 1 - \frac{{n_0-1 \choose m}}{{n_0+n_1 \choose m}} \right). $$ In particular, $$ \EE Y \leq \frac{n_0}{1+n_1} $$ for any value of $m$. \end{lemma} We proceed to prove the theorem. Denote $m_0 = |\calH_0|$ and recall that $|\calH| = p$ and $|\calK_0| = r$. Recall also that we are conditioning throughout on $\bbeta$ (so that, in particular, $\calH_0$ is nonrandom). We will show that FDR control holds conditionally on $\bbeta$ (it therefore also holds unconditionally). With the notation above, $$ \tau = \inf\left\{ t: \frac{(1+V_1(t))\cdot \frac{p}{1+r}}{1\vee R(t)} \leq q \right\}. $$ %Using a manipulation similar to that in \citet{barber2015controlling}, We have \begin{align*} \tFDR = \EE \left[ \frac{V_0(\tau)}{1\vee R(\tau)} \right] = \EE \left[ \frac{(1+V_1(\tau))\cdot \frac{p}{1+r}}{1\vee R(\tau)} \cdot \frac{V_0(\tau)}{(1+V_1(\tau))\cdot \frac{p}{1+r}} \right] &\leq q\EE\left[ \frac{V_0(\tau)}{(1+V_1(\tau)) \cdot \frac{p}{1+r}} \right]\\ &= q \frac{1+r}{p}\EE\left[ \frac{V_0(\tau)}{1+V_1(\tau)} \right]. \end{align*} We will show that \begin{equation} \EE\left[ \frac{V_0(\tau)}{1+V_1(\tau)} \right] \leq \frac{m_0}{1+r}. \end{equation} Consider the filtration $$ \calF_t = \sigma\left( \{V_0(s)\}_{0\leq s\leq t}, \{V_1(s)\}_{0\leq s\leq t}, \{R(s)\}_{0\leq s\leq t} \right). $$ That is, at all times $s\leq t$, we have knowledge about the number of true null, fake null and nonnull variables with statistics larger than $s$. We claim that $$ \frac{V_0(t)}{1+V_1(t)} $$ is a supermartingale w.r.t. $\{\calF_t\}$, and $\tau$ is a stopping time. Indeed, let $t>s$. We need to show that \begin{equation}\label{eq:supermart} \EE \left[\frac{V_0(t)}{1+V_1(t)} | \calF_s\right] \leq \frac{V_0(s)}{1+V_1(s)}. \end{equation} The crucial observation is that, conditional on $V_0(s)$, $V_1(s)$ {\it and} $V_0(t) + V_1(t) = m$, $$ V_0(t )\sim \text{Hyper}(V_0(s), V_1(s); m). $$ This follows from the exchangeability of the test statistics for the true and fake nulls (in other words, conditional on everything else, the order of the red markers in Figure \ref{fig:illustration} is random). %We demonstrate this with reference to Figure \ref{fig:illustration}: at time $s$, we are told that exactly five true nulls and exactly three fake nulls have entered before time $s$. %We are also told that a total of five nulls (true or fake) have entered ``even before" time $t>s$. %Then those five that entered ``even before" time $t>s$ must be a subset of the $5+3=8$ nulls that we know have entered at some point before $s$, and, because the red markers appear in a random order, each subset of five nulls among those eight, is equally likely. %Then $V_0(t)$ is counting the number of ``successes" in selecting at random five items from an urn that has five ``successes" and three ``failures". Now, Lemma \ref{lem:hyper} assures us that \begin{equation} \EE \left[\frac{V_0(t)}{1+V_1(t)} | V_0(s), V_1(s), V_0(t) + V_1(t) = m\right] \leq \frac{V_0(s)}{1+V_1(s)} \end{equation} regardless of the value of $m$. Hence, the supermartingale property \eqref{eq:supermart} holds (and it is also clear that $\tau$ is a stopping time). By the optional stopping theorem for supermartingales, \begin{equation}\label{eq:optional} \EE \left[\frac{V_0(t)}{1+V_1(t)} \right] \leq \EE \left[\frac{V_0(0)}{1+V_1(0)} \right] = \EE \left( \EE \left[\frac{V_0(0)}{1+V_1(0)} | V_0(0)+V_1(0) \right] \right). \end{equation} But conditional on $V_0(0)+V_1(0)=m$ we have that $V_0(0 )\sim \text{Hyper}(m_0, r; m)$, and so, invoking Lemma \ref{lem:hyper} once again, we have $$ \EE \left[\frac{V_0(0)}{1+V_1(0)} | V_0(0)+V_1(0)=m \right] \leq \frac{m_0}{1+r} $$ no matter the value of $m$. \end{proof} \end{subsection} \begin{subsection}{Estimating the proportion of nulls}\label{subsec:pi-naught} From our analysis, we see that FDR control would continue to hold if we replaced the number of hypotheses $p=|\calH|$ with the number of nulls $m_0=|\calH_0|$ in the definition of $\tau$. Of course, this quantity is unobservable, or, equivalently, $\pi_0 = |\calH_0|/|\calH|$ is unobservable. We can, however, estimate $\pi_0$ using a similar approach to that of \citet{storey2002direct}, except that we replace calculations under the theoretical null with the counting of knockoff variables at the ``bottom" of the Lasso path: specifically, for any $\lambda_0$ we have $$ \frac{ |\{j\in \calH_0: T_j\leq \lambda_0\}| }{|\calH_0|} \approx \frac{ |\{j\in \calK_0: T_j\leq \lambda_0\}| }{|\calK_0|}, $$ again due to the exchangeability of null features. It follows that $$ \begin{aligned} |\calH_0| \approx |\calK_0| \cdot \frac{ |\{ j\in \calH_0:T_j\leq \lambda_0 \}| }{ |\{ j\in \calK_0:T_j\leq \lambda_0 \}| } &\leq (|\calK_0| + 1) \cdot \frac{ 1+|\{ j\in \calH_0:T_j\leq \lambda_0 \}| }{ |\{ j\in \calK_0:T_j\leq \lambda_0 \}| } \\ &\leq (|\calK_0| + 1) \cdot \frac{ 1+|\{ j\in \calH:T_j\leq \lambda_0 \}| }{ |\{ j\in \calK_0:T_j\leq \lambda_0 \}| }. \end{aligned} $$ For small $\lambda_0$, we expect only few non-nulls among $\{j\in \calH: T_j\leq \lambda_0\}$ so that $|\{j\in \calH_0:T_j\leq \lambda_0\}| \approx |\{j\in \calH:T_j\leq \lambda_0\}|$. Hence, we propose to estimate $\pi_0$ by \begin{equation}\label{eq:pi-naught-hat-lambda} \widehat{\pi}_0 = \frac{(|\calK_0| + 1)}{|\calH|} \cdot \frac{ 1+|\{ j\in \calH:T_j\leq \lambda_0 \}| }{ |\{ j\in \calK_0:T_j\leq \lambda_0 \}| }. \end{equation} %or, in practice, by the minimum between one and \eqref{eq:pi-naught-hat-lambda}. We again state our result more generally. \begin{theorem}[FDR control for knockoff procedure with an estimate of $\widehat{\pi}_0$]\label{thm:fdr-control-est} Set \begin{equation}\label{eq:pi-naught-hat} \widehat{\pi}_0 = \frac{(|\calK_0| + 1)}{|\calH|} \cdot \frac{ 1+|\{ j\in \calH:T_j\leq t_0 \}| }{ |\{ j\in \calK_0:T_j\leq t_0 \}| } \end{equation} ($\widehat{\pi}_0$ is set to $\infty$ if the denominator vanishes). In the setting of Theorem \ref{thm:fdr-control}, the procedure that rejects $\calH_0^j$ if $T_j\geq \tau$, where \begin{equation}\label{eq:tau-est} \tau = \inf\left\{ t \geq t_0: \frac{ (1 + V_1(t)) \cdot \frac{|\calH| \widehat{\pi}_0}{1+|\calK_0|} }{1\vee R(t)} \leq q \right\} \end{equation} obeys $$ \tFDR \leq q. $$ %Furthermore, if $\widehat{\pi}_0$ is estimated by the minimum of \eqref{eq:pi-naught-hat} and 1, then, under the working assumptions stated at the beginning of Section \ref{sec:tradeoff}, and if $r = \ceil{\rho p}$ for a constant $\rho>0$, we have that %$$ %\lim_{n,p\to \infty} \tFDR \leq q %$$ %where the expectation in the last inequality is taken also with %respect to $\bbeta$, distributed as in Section %\ref{sec:tradeoff}.\footnote{This is the only place in Section % \ref{sec:knockoffs} treating the coefficients as random.} \end{theorem} \begin{proof} We use the same notation as in the proof of Theorem \ref{thm:fdr-control}. As before, we have that $$ \tFDR \leq q\EE\left[ \frac{V_0(\tau)}{(1+V_1(\tau)) \cdot \frac{p \widehat{\pi}_0}{1+r}} \right]. $$ Again, the process $\{V_0(t)/(1+V_1(t))\}_{t\geq t_0}$ starting at $t_0$, is a supermartingale w.r.t. $\{\calF_t\}_{t\geq t_0}$ and $\tau$ is a stopping time. Then the optional stopping theorem gives \begin{equation} \begin{aligned} \EE\left[ \frac{V_0(\tau)}{(1+V_1(\tau)) \cdot \frac{p \widehat{\pi}_0}{1+r}} \right] &\leq \EE\left[ \frac{V_0(t_0)}{(1+V_1(t_0)) \cdot \frac{p \widehat{\pi}_0}{1+r}} \right]\\ &=\EE\left[ \frac{V_0(t_0)}{1+V_1(t_0)} \cdot \frac{r-V_1(t_0)}{1+p-R(t_0)} \right]\\ &\leq \EE \left[ \frac{V_0(t_0)}{1+V_1(t_0)} \cdot \frac{r-V_1(t_0)}{1+m_0-V_0(t_0)} \right]. \end{aligned} \end{equation} As before, $V_0(t_0) \sim \text{Hyper}(m_0, r;m)$ conditional on $V_0(t_0) + V_1(t_0)=m$. In the Appendix we prove that if a random variable $X$ follows this distribution, then \begin{equation}\label{eq:hypergeometric-identity} \EE \left[ \frac{X}{1+m-X} \cdot \frac{r+X-m}{1+m_0-X} \right] = 1 - \frac{ {m_0\choose m_0\wedge m} {r\choose m-m_0\wedge m} }{ {m_0+r\choose m} } \end{equation} In particular, for all $m$ the result is no more than $1$, thereby completing the proof of the theorem. %\medskip % %Now let %\begin{equation*} %\widehat{\pi}_0 = 1 \ \wedge\ \left( \frac{(|\calK_0| + 1)}{|\calH|} \cdot \frac{ 1+|\{ j\in \calH:T_j\leq t_0 \}| }{ |\{ j\in \calK_0:T_j\leq t_0 \}| }\right). %\end{equation*} %As in the first part of the proof, %$$ %\tFDR \leq q\EE\left[ \frac{V_0(\tau)}{(1+V_1(\tau)) \cdot \frac{p \widehat{\pi}_0}{1+r}} \right], %$$ %and by the same reasoning as above, %\begin{equation} %\begin{aligned} %\EE\left[ \frac{V_0(\tau)}{(1+V_1(\tau)) \cdot \frac{p \widehat{\pi}_0}{1+r}} \right] &\leq \EE\left[ \frac{V_0(t_0)}{(1+V_1(t_0)) \cdot \frac{p \widehat{\pi}_0}{1+r}} \right]\\ %&=\EE\left[ \frac{V_0(t_0)}{1+V_1(t_0)} \cdot \max \left( \frac{r-V_1(t_0)}{1+p-R(t_0)}, \frac{1+r}{p} \right) \right]\\ %&\leq \EE \left[ \frac{V_0(t_0)}{1+V_1(t_0)} \cdot \max \left( \frac{r-V_1(t_0)}{1+m_0-V_0(t_0)}, \frac{1+r}{p} \right) \right]. %\end{aligned} %\end{equation} %Now, as in equation \eqref{eq:fdp-prime} below, we have %$$ %\frac{V_0(t_0)}{p} = \frac{|\{j\in \calH_0: T_j \geq t_0\}|}{p} \approx \frac{|\{j\in \calH_0: \widehat{\beta}_j(t_0)\neq 0, \beta_j = 0\}|}{p} \to 2(1-\epsilon')\Phi(-\alpha_1'), %$$ %where $\epsilon' = \epsilon/(1 + \rho)$ and where $\alpha_1'$ is the same as in equation \eqref{eq:fdp-hat-infty-pi-naught-hat}, when replacing $\lambda_0$ with $t_0$. %Also, using equation \eqref{eq:fdp-fake-prime}, we have %$$ %\frac{V_1(t_0)}{r} = \frac{|\{j\in \calK_0: T_j \geq t_0\}|}{r} \approx \frac{|\{j\in \calK_0: \widehat{\beta}_j(t_0)\neq 0\}|}{r} \to 2\Phi(-\alpha_1'). %$$ %Therefore, %$$ %\frac{V_0(t_0)}{1+V_1(t_0)}\bigg/\frac{p}{r} \to \frac{2(1-\epsilon')\Phi(-\alpha_1')}{2\Phi(-\alpha_1')} = 1-\epsilon' %$$ %and %$$ %\frac{r-V_1(t_0)}{1+m_0-V_0(t_0)}\bigg/\frac{r}{p} \to \frac{[1-2\Phi(-\alpha_1')]}{(1-\epsilon')[1-2\Phi(-\alpha_1')]} = \frac{1}{1-\epsilon'},\ \ \ \ \ \ %\frac{1+r}{p}\bigg/\frac{r}{p} \to 1. %$$ %Hence, %$$ %\frac{V_0(t_0)}{1+V_1(t_0)} \cdot \max \left( \frac{r-V_1(t_0)}{1+m_0-V_0(t_0)}, \frac{1+r}{p} \right) \to (1-\epsilon') \cdot \max(\frac{1}{1-\epsilon'}, 1) = 1, %$$ %which completes the proof of the second part of the theorem. \end{proof} In practice, we of course suggest to use the minimum between one and \eqref{eq:pi-naught-hat-lambda} as an estimator for $\pi_0$. While we cannot obtain a result similar to Theorem \ref{thm:fdr-control-est} when the truncated estimate of $\pi_0$ is used instead of \eqref{eq:pi-naught-hat-lambda}, we can appeal to results from approximate message passing, in the spirit of the calculations from Section \ref{sec:power}, to obtain an approximate asymptotic result. Thus, let \begin{equation*} \widehat{\pi}_0 = 1 \ \wedge\ \left( \frac{(|\calK_0| + 1)}{|\calH|} \cdot \frac{ 1+|\{ j\in \calH:T_j\leq t_0 \}| }{ |\{ j\in \calK_0:T_j\leq t_0 \}| }\right). \end{equation*} and adopt the working assumptions stated at the beginning of Section \ref{sec:power}. As in the proof of the theorem above, $$ \tFDR \leq q\EE\left[ \frac{V_0(\tau)}{(1+V_1(\tau)) \cdot \frac{p \widehat{\pi}_0}{1+r}} \right], $$ and by the same reasoning as above, \begin{equation} \begin{aligned} \EE\left[ \frac{V_0(\tau)}{(1+V_1(\tau)) \cdot \frac{p \widehat{\pi}_0}{1+r}} \right] &\leq \EE\left[ \frac{V_0(t_0)}{(1+V_1(t_0)) \cdot \frac{p \widehat{\pi}_0}{1+r}} \right]\\ &=\EE\left[ \frac{V_0(t_0)}{1+V_1(t_0)} \cdot \max \left( \frac{r-V_1(t_0)}{1+p-R(t_0)}, \frac{1+r}{p} \right) \right]\\ &\leq \EE \left[ \frac{V_0(t_0)}{1+V_1(t_0)} \cdot \max \left( \frac{r-V_1(t_0)}{1+m_0-V_0(t_0)}, \frac{1+r}{p} \right) \right]. \end{aligned} \end{equation} Now, as in equation \eqref{eq:fdp-prime} below, we have $$ \frac{V_0(t_0)}{p} = \frac{|\{j\in \calH_0: T_j \geq t_0\}|}{p} \approx \frac{|\{j\in \calH_0: \widehat{\beta}_j(t_0)\neq 0, \beta_j = 0\}|}{p} \to 2(1-\epsilon')\Phi(-\alpha_1'), $$ where $\epsilon' = \epsilon/(1 + \rho)$ and where $\alpha_1'$ is the same as in equation \eqref{eq:fdp-hat-infty-pi-naught-hat}, when replacing $\lambda_0$ with $t_0$. Also, using equation \eqref{eq:fdp-fake-prime}, we have $$ \frac{V_1(t_0)}{r} = \frac{|\{j\in \calK_0: T_j \geq t_0\}|}{r} \approx \frac{|\{j\in \calK_0: \widehat{\beta}_j(t_0)\neq 0\}|}{r} \to 2\Phi(-\alpha_1'). $$ Treating the approximations above as equalities (in Section \ref{sec:power} we argue that these should be good approximations), we have $$ \frac{V_0(t_0)}{1+V_1(t_0)}\bigg/\frac{p}{r} \to \frac{2(1-\epsilon')\Phi(-\alpha_1')}{2\Phi(-\alpha_1')} = 1-\epsilon' $$ and $$ \frac{r-V_1(t_0)}{1+m_0-V_0(t_0)}\bigg/\frac{r}{p} \to \frac{[1-2\Phi(-\alpha_1')]}{(1-\epsilon')[1-2\Phi(-\alpha_1')]} = \frac{1}{1-\epsilon'},\ \ \ \ \ \ \frac{1+r}{p}\bigg/\frac{r}{p} \to 1. $$ Hence, $$ \frac{V_0(t_0)}{1+V_1(t_0)} \cdot \max \left( \frac{r-V_1(t_0)}{1+m_0-V_0(t_0)}, \frac{1+r}{p} \right) \to (1-\epsilon') \cdot \max(\frac{1}{1-\epsilon'}, 1) = 1. $$ We conclude that, up to the approximations above, $$ \limsup_{n,p\to \infty} \tFDR \leq q. $$ \end{subsection} \section{Power Analysis}\label{sec:power} The main observation in this paper is that we can use the consequences of Theorem 1.5 in \citet{bayati2012lasso} to obtain an asymptotic TPP-FDP tradeoff diagram for the Lasso solution associated with the {\it knockoff} setup; that is, the Lasso solution corresponding to the {augmented} design $\mathbb{X}$. Indeed, the working assumptions required for applying Lemma \ref{sec:tradeoff} continue to hold, only with a slightly different set of parameters. Recall that $\mathbb{X}\in \RR^{n\times(p+r)}$ has i.i.d.~$\calN(0,1/n)$ entries. Taking $r = \ceil{\rho p}$ for a constant $\rho>0$, we have that $n/(p+r) \to \delta' := \delta/(1+\rho)$ as $n,p\to \infty$. Furthermore, because $\EE Y$ is related to $\mathbb{X}$ through $$ \EE Y = \mathbb{X} \begin{bmatrix} \bbeta \\ \mathbf{0} \end{bmatrix}, % = \mathbb{X} [\bbeta \ \mathbf{0}]^T $$ where $\mathbf{0} = (0,....,0)^T\in \RR^r$, we have that the {empirical distribution} of the coefficients corresponding to the {\it augmented} design converges exactly to \begin{equation}\label{eq:pi-prime} \Pi' = \begin{cases} \Pi^*, &\text{w.p.} \ \epsilon',\\ 0, &\text{w.p.} \ 1-\epsilon', \end{cases} \end{equation} where $\epsilon' := \epsilon/(1 + \rho) = \lim \{p\epsilon/(p+\ceil{\rho p})\}$. Fortunately, it is only the {\it empirical} distribution of the coefficient vector that needs to have a limit of the form \eqref{eq:pi-prime}, for Lemma \ref{lem:infty} to follow from Theorem 1.5 of \citet{bayati2012lasso}; see also the remarks in Section 2 of \citet{su2017false}. The working hypothesis of Section \ref{sec:tradeoff} therefore holds with $\delta$ and $\epsilon$ replaced by $\delta'$ and $\epsilon'$. Recalling the definitions of $\calH, \calH_0$ and $\calK_0$ from Section \ref{sec:knockoffs}, a counterpart of Lemma \ref{lem:infty} asserts that \begin{align} \frac{|\{j\in \calH: \widehat{\beta}_j(\lambda)\neq 0, \beta_j = 0\}|}{p} &\longrightarrow 2(1-\epsilon)\Phi(-\alpha'), \\ \label{eq:fdp-prime} \frac{|\{j\in \calH: \widehat{\beta}_j(\lambda)\neq 0, \beta_j \neq 0\}|}{p} &\longrightarrow \epsilon \PP (|\Pi^* + \tau'W|>\alpha'\tau') \end{align} and \begin{equation}\label{eq:fdp-fake-prime} \frac{|\{j\in \calK_0: \widehat{\beta}_j(\lambda)\neq 0\}|}{r} \longrightarrow 2\Phi(-\alpha'), \end{equation} where $\widehat{\beta}_j(\lambda)$ is the solution to \eqref{eq:lasso-til}; $W$ is a $\calN(0,1)$ variable independent of $\Pi$; and $\tau'>0,\\alpha' >\max\{\alpha_0', 0\}$ is the unique solution to \eqref{eq:system} when $\delta$ is replaced by $\delta'$ and $\epsilon$ is replaced by $\epsilon'$. Hence, \begin{equation}\label{eq:fdp-infty-prime} \begin{aligned} \tFDP(\lambda) &= \frac{|\{j\in \calH_0: T_j \geq \lambda\}|}{1\vee |\{j\in \calH: T_j \geq \lambda\}|} \\ &= \frac{|\{j\in \calH_0: \widehat{\beta}_j(\lambda')\neq 0 \text{ for some } \lambda' \geq \lambda\}|}{1\vee |\{j\in \calH: \widehat{\beta}_j(\lambda')\neq 0 \text{ for some } \lambda' \geq \lambda\}|}\\ %\end{aligned} %\end{equation} %Now, under the assumption that variables for which $\widehat{\beta}_j(\lambda')\neq 0 \text{ for some } \lambda' > \lambda$ and $\widehat{\beta}_j(\lambda)= 0$ are extremely rare (although we cannot rule this event out completely), we can approximate the last term in \eqref{eq:fdp-infty-prime} by %\begin{equation}\label{eq:fdp-infty-prime-cont} %\begin{aligned} &\approx \frac{|\{j\in \calH_0: \widehat{\beta}_j(\lambda)\neq 0\}|}{1\vee |\{j\in \calH: \widehat{\beta}_j(\lambda)\neq 0\}|} \\ &\longrightarrow \frac{2(1-\epsilon)\Phi(-\alpha')}{2(1-\epsilon)\Phi(-\alpha') + \epsilon \PP (|\Pi^* + \tau'W|>\alpha'\tau')} \equiv \fdpinftyprime(t). \end{aligned} \end{equation} %where $\widehat{\beta}_j(\lambda)$ is the solution to \eqref{eq:lasso-til} with $\lambda = t$; $W$ is a $\calN(0,1)$ variable independent of $\Pi$; and $\tau'>0,\ \alpha' >\max\{\alpha_0', 0\}$ is the unique solution to \eqref{eq:system} when $\lambda = t$ and when $\delta$ is replaced by $\delta'$ and $\epsilon$ is replaced by $\epsilon'$. Above, we allowed ourselves to approximate $|\{j\in \calH: \widehat{\beta}_j(\lambda')\neq 0 \text{ for some } \lambda' \geq \lambda\}|$ with $|\{j\in \calH: \widehat{\beta}_j(\lambda)\neq 0\}|$, and likewise for $\calH_0$. If least-angle regression \citep{efron2004least} %\done{Give ref.~to LAR.} were used instead of the Lasso, the two quantities would have been equal; but because the Lasso solution path is considered here, we cannot completely rule out the event in which a variable that entered the path drops out, and the former quantity is in general larger. Still, we expect the difference to be very small, at least for large enough values of $\lambda$ (corresponding to small FDP). %\ejc{You also have % a denominator.} This is verified by the simulation of Figure \ref{fig:simulations}: the green circles, computed using the statistics $T_j$, are very close to the red circles, computed when rejections correspond to variables with $\widehat{\beta}_j(\lambda)\neq 0$. All three simulation examples were computed for a single realization on the original $n\times p$ design (not the augmented design). Similar results were obtained when we repeated the experiment (i.e., for a different realization of the data). %\done{I would say that that this is not a special simulation. This is what we would see again and again.} \begin{figure}[] \centering \includegraphics[width=.8\textwidth]{simulations.pdf} \caption{ Discrepancy between testing procedures for three different values of $\epsilon$. In all three examples $n=p=5000$, $\Pi^*\sim \exp(1)$, and $\sigma = 0.5$; and TPP is plotted against FDP. Green circles correspond to the procedure that uses the statistics $T_j$ in \eqref{eq:stat}. Red circles correspond to the procedure that rejects when $\widehat{\beta}_j(\lambda)\neq 0$. Black circles are theoretical predictions from Section \ref{sec:power} with $\delta = 1$. Each curve is is computed from a single realization of the data, and uses an $n\times p$ design matrix (i.e., not the augmented design). % \ejc{Here or in the main % text, perhaps comment on the maximum difference between green and % red. AW: I did compute the max differences, and they were small, but I didn't want to include a comment because if we did, we might need to explain what exactly we are taking the difference between: the green and the red points have different x values} } \label{fig:simulations} \end{figure} \noindent Similarly, \begin{equation}\label{eq:tpp-infty-prime} \begin{aligned} \tTPP_{\text{aug}}(\lambda) &\equiv \frac{|\{j\in \calH\setminus \calH_0: T_j \geq \lambda\}|}{1\vee |\{j\in \calH: \beta_j \neq 0\}|} \\ &\approx \frac{|\{j\in \calH \setminus \calH_0: \widehat{\beta}_j(\lambda)\neq 0\}|}{1\vee |\{j\in \calH: \beta_j \neq 0\}|}\\ &\longrightarrow \PP (|\Pi^* + \tau'W|>\alpha'\tau') \equiv \tppinftyprime(\lambda) \end{aligned} \end{equation} and \begin{equation}\label{eq:fdp-hat-infty-gen} \begin{aligned} \widehat{\tFDP}_{\text{aug}}(\lambda) &\equiv \frac{ (1 + |\{j\in \calK_0: T_j \geq \lambda\}|) \cdot \frac{|\calH| \widehat{\pi}_0}{1+|\calK_0|} }{|\{j\in \calH: T_j \geq \lambda\}|} \approx \frac{ (1 + |\{j\in \calK_0: \widehat{\beta}_j(\lambda)\neq 0\}|) \cdot \frac{|\calH| \widehat{\pi}_0}{1+|\calK_0|} }{|\{j\in \calH: \widehat{\beta}_j(\lambda)\neq 0\}|}\\ &\longrightarrow \frac{2\Phi(-\alpha')}{2(1-\epsilon)\Phi(-\alpha') + \epsilon \PP (|\Pi^* + \tau'W|>\alpha'\tau')} \cdot \lim \widehat{\pi}_0, \end{aligned} \end{equation} where $W$, $\tau'>0,\\alpha' >\max\{\alpha_0', 0\}$ are as before, and where $\widehat{\pi}_0$ is any estimate of $\pi_0$ that has a limit. Taking $\widehat{\pi}_0$ to be the minimum between one and \eqref{eq:pi-naught-hat}, we have that the limit in \eqref{eq:fdp-hat-infty-gen} can be approximated by \begin{equation}\label{eq:fdp-hat-infty-pi-naught-hat} \begin{aligned} &\frac{2\Phi(-\alpha')}{2(1-\epsilon)\Phi(-\alpha') + \epsilon \PP (|\Pi^* + \tau'W|>\alpha'\tau')} \\ &\cdot \left(1 \wedge \left\{ 1-\epsilon + \frac{ \epsilon [\PP (|\Pi^* + \tau'_2 W|>\alpha'_2\tau'_2) - \PP (|\Pi^* + \tau'_1 W|>\alpha'_1\tau'_1)] }{2[\Phi(-\alpha'_2) - \Phi(-\alpha'_1)]} \right\}\right) \equiv \fdphatinftyprime(\lambda), \end{aligned} \end{equation} where $\tau'_1>0,\\alpha'_1 >\max\{\alpha_0', 0\}$ is the unique solution to \eqref{eq:system} when $\lambda = \lambda_0$ and when $\delta$ is replaced by $\delta'$ and $\epsilon$ is replaced by $\epsilon'$; and where $\tau'_2>0,\\alpha'_2 >\max\{\alpha_0', 0\}$ is the limit of the unique solution to \eqref{eq:system} when $\lambda \to 0^+$ also with $\delta'$ (resp.~$\epsilon'$) in lieu of $\delta$ (resp.~$\epsilon$). Indeed, \begin{equation}\label{eq:tau-naught-I} \begin{aligned} \frac{|\{j\in \calH: 0<T_j< \lambda_0\}|}{p} &= \frac{|\{j\in \calH: T_j>0\}| - |\{j\in \calH: T_j\geq \lambda_0\}|}{p}\\ &\approx \frac{|\{j\in \calH: \widehat{\beta}_j(0^+)\neq 0\}| - |\{j\in \calH: \widehat{\beta}_j(\lambda_0)\neq 0\}|}{p}\\ &\longrightarrow 2(1-\epsilon)\Phi(-\alpha'_2) + \epsilon \PP (|\Pi^* + \tau'_2 W|>\alpha'_2\tau'_2)\\ &- 2(1-\epsilon)\Phi(-\alpha'_1) + \epsilon \PP (|\Pi^* + \tau'_1 W|>\alpha'_1\tau'_1) \end{aligned} \end{equation} and, by a similar calculation, \begin{equation}\label{eq:tau-naught-II} \begin{aligned} \frac{|\{j\in \calK_0: 0<T_j\leq \lambda_0\}|}{|\calK_0|} \longrightarrow 2\Phi(-\alpha'_2) - 2\Phi(-\alpha'_1); \end{aligned} \end{equation} using \eqref{eq:tau-naught-I} and \eqref{eq:tau-naught-II} to compute the limit of \eqref{eq:pi-naught-hat}, we obtain \eqref{eq:fdp-hat-infty-pi-naught-hat}. \begin{figure}[] \centering \includegraphics[width=.8\textwidth]{tradeoff-exp-multiple.pdf} \caption{ % \ejc{This is our main result. Any chance we can % highlight/emphasize this a bit more?} Tradeoff diagrams for $\Pi^*\sim \exp(1)$. $\sigma = 0.5$; $\rho=1$. Different panels correspond to different combinations of $\delta$ and $\epsilon$, as appears in the title of each of the plots. The black (resp.~light blue) curve corresponds to the oracle (resp.~knockoffs). Markers denote the pair $(\mathrm{tpp},\mathrm{fdp})$ of asymptotic power and type I error, attained for three example values of $q$: black for oracle, light blue for knockoff. The knockoff procedure loses a little bit of power due to the estimate of $1-\epsilon$ (proportion of true nulls). } \label{fig:tradeoff-exp} \end{figure} As $\lambda>0$ varies, a parametric curve $(\fdpinftyprime(\lambda), \tppinftyprime(\lambda))$ is traced which specifies the FDP level attainable at a specific TPP level for the Lasso solution corresponding to the {\it augmented} design, as \begin{equation} q_{\text{aug}}^{\Pi^*}(\tppinftyprime(\lambda); \epsilon, \delta, \sigma) = \fdpinftyprime(\lambda). \end{equation} Figure \ref{fig:tradeoff-exp} shows this curve against the analogous curve $(\fdpinfty(\lambda), \tppinfty(\lambda))$, associated with the original design $\bX$ and the oracle procedure, when $\Pi^*$ is exponential with mean $1$. The different panels correspond to different combinations of $\delta$ and $\epsilon$. The proportion $\rho$ was taken to be $1$ in all scenarios, in other words, the number of knockoff variables $r$ was taken to be equal to the number of original variables $p$; we observed only very slight differences in the curves when $\rho$ was varied between $0.1$ and $1$. Figure \ref{fig:tradeoff-distributions} shows the two curves, $(\fdpinfty(\lambda), \tppinfty(\lambda))$ and $(\fdpinftyprime(\lambda), \tppinftyprime(\lambda))$, for different distributions $\Pi^*$. \begin{figure}[] \centering \includegraphics[width=.8\textwidth]{tradeoff-distributions.pdf} \caption{ Tradeoff diagrams for different $\Pi^*$, with $\epsilon = 0.2$, $\delta = 1$ and $\sigma = 0.5$. } \label{fig:tradeoff-distributions} \end{figure} In all situations depicted in Figures \ref{fig:tradeoff-exp} and \ref{fig:tradeoff-distributions}, the light blue curve, corresponding to the augmented setup, lies very close to the black curve, which corresponds to the original design. Hence, augmenting the design with knockoffs seems to have little effect on the tradeoff diagram, or, the achievable power for a given FDP level, at least for FDP levels small enough to be of interest. For example, if we set $q=0.05$, then there exist values $\lambda$ and $t'$ such that $\fdpinfty(\lambda) = \fdpinftyprime(\lambda') = 0.05$ and $\tppinfty(\lambda) \approx \tppinftyprime(\lambda')$. Of course, the knockoff procedure achieves this without knowledge of $\epsilon$ and $\Pi^*$, while the oracle relies on knowing these quantities in choosing $\lambda$. Note, still, that for a given input $q$, the knockoff procedure essentially (asymptotically) chooses $\lambda'$ such that $\fdphatinftyprime(\lambda')$, not $\fdpinftyprime(\lambda')$, is equal to $q$. Because the former uses a conservative estimate of $1-\epsilon$, $\fdphatinftyprime(\lambda')$ is usually larger than $\fdpinftyprime(\lambda')$ and, consequently, the power attained by the knockoff procedure is a little lower. This gap is depicted by the ``+" markers in the plots, each pair of blue and black markers corresponding to a particular value of $q$. For example, for $\delta = 1$ and $\epsilon = 0.2$, oracle power is $0.187$ and knockoffs power is $0.18$ at $q = 0.1$. The value of $t_0$ used is $0.1$. Figure \ref{fig:intro} gives an alternative representation (see caption). %\done{Perhaps, give numbers? At say $q = 0.1$, oracle power is bla, knockoffs power is bla.} \section{Prediction}\label{sec:prediction} Armed with a procedure that controls the FDR for the model \eqref{eq:model}, we now explore utilizing knockoffs for estimation purposes (the prediction error and the estimation error are equivalent here because we have uncorrelated predictors). Our pursuit is to a large degree inspired by the work of \citet{abramovich2006special}, which for the first time rigorously linked FDR control and estimation error. %Indeed, \citet{abramovich2006special} showed formally that in the Gaussian sequence model, FDR (hard-) thresholding adapts to unknown sparsity and achieves asymptotic minimaxity in estimation, provided that [...]. %Turning to the linear model, it is natural to ask if a procedure that controls the FDR still exhibits competitive performance in terms of the risk. %In fact, the SLOPE estimator \citep{bogdan2015slope} was in part motivated by such prospects, even though strict FDR control and asymptotic minimaxity have so far only been proven for some very special cases of the linear model and in a different sparsity regime than ours \citep{su2016slope}. At the same time, the working assumptions we adopt in the current paper allow us to take further advantage of the results of \citet{bayati2012lasso}, namely, to asymptotically analyze not only the power or the knockoff procedure but also the {\it estimation} error associated with it. As in the preceding analysis of testing errors, the (normalized) risk of the Lasso estimator at a fixed $\lambda$ tends in probability to a simple limit. Specifically, by Corollary 1.6 of \citet{bayati2012lasso}, we have that, for a given pair of $\epsilon$ and $\Pi^*$, and for fixed $\lambda$, \begin{equation} \lim \frac{1}{p} \| \widehat{\bbeta}(\lambda) - \bbeta \|^2 \longrightarrow \EE\left[ \left(\eta_{\alpha\tau}(\Pi + \tau W) - \Pi \right)^2 \right] \equiv R^{\infty}(\lambda; \Pi^*, \epsilon), \end{equation} where $W\sim \calN(0,1)$ independently of $\Pi$, and $\tau>0,\\alpha >\max\{\alpha_0, 0\}$ is the unique solution to \eqref{eq:system}. Because we have a tractable form for the limiting risk, we can minimize it over $\lambda$ to obtain an {\it oracle} choice for $\lambda$, \begin{equation}\label{eq:oracle-lambda} \lambda^{\infty}_{\text{OL}}(\Pi^*, \epsilon) = \argmin_{\lambda} R^{\infty}(\lambda; \Pi^*, \epsilon). \end{equation} The oracle choice of $\lambda$ depends on $\Pi^*$ and on $\epsilon$ and therefore does not produce a legitimate estimator for $\bbeta$, but it sets a benchmark for evaluating the performance of a procedure that plugs in any (possibly data-dependent) value for $\lambda$. Consider now a fixed $q$ and the knockoff procedure from Section \ref{subsec:pi-naught}, which selects the $j$-th variable if $T_j\geq \tau$, where $\tau$ is given by \eqref{eq:lambda-hat} and $\widehat{\pi}_0$ by \eqref{eq:pi-naught-hat} and $\Lambda = [t_0,\infty)$. This is a legal variable selection procedure (it does not depend on $\Pi^*$ or on $\epsilon$) with the FDR control guarantees of Theorem \ref{thm:fdr-control-est}. In the limit as $n,p\to \infty$, it corresponds to a choice of $\lambda$ defined by \begin{equation*} \fdphatinftyprime(\lambda) = q \end{equation*} where $\fdphatinftyprime(\lambda)$ is given in \eqref{eq:fdp-hat-infty-pi-naught-hat}. Call this the ``knockoff" choice for $\lambda$ and denote it by $\lambda^{\infty}_{\text{KO}}(q)$. Then the quantity \begin{equation}\label{eq:inflation} \nu(q) \equiv \sup_{\Pi^*} \left\{ \frac{R^{\infty}(\lambda^{\infty}_{\text{KO}}(q); \Pi^*, \epsilon)}{R^{\infty}(\lambda^{\infty}_{\text{OL}}(\Pi^*, \epsilon); \Pi^*, \epsilon)} \right\} = \sup_{\Pi^*} \left\{ \frac{R^{\infty}(\lambda^{\infty}_{\text{KO}}(q); \Pi^*, \epsilon)}{\inf_{\lambda}R^{\infty}(\lambda; \Pi^*, \epsilon)} \right\} \end{equation} is the worst-case (in $\Pi^*$) inflation of the asymptotic risk due to using $\lambda^{\infty}_{\text{KO}}(q)$ instead of the optimal choice for $\lambda$ \citep[our use of the term {\it risk inflation} alludes to the similarity to the criterion proposed in][]{foster1994risk}. It is not clear that $\nu(q)$ is at all finite; but if it were true that there exists a choice of $q$ such that for all $\epsilon, \delta, \sigma$, $\nu(q)$ is bounded by $1 + \Delta$ for small $\Delta$, it would be rather reassuring. Even for fixed $\epsilon, \delta$ and $\sigma$, computing \eqref{eq:inflation} is not trivial because we need to maximize the ratio over all distributions $\Pi^*$; neither were we able to {\it bound} \eqref{eq:inflation} in general. The problem is much simpler in the case where $\Pi^*$ is a point mass at some nonzero (positive) value $\lambda$, because in this case the risk inflation is a univariate function. In Figure \ref{fig:point} we plot the risk inflation as a function of $\lambda$ for $\epsilon = 0.1$ and $\delta = 1, \sigma = 0.5$. The maximum risk inflation is $1.177$ and attained at $t=1.9$, which is quite interesting as this is a case of neither very low nor very high ``signal to noise" ratio. Changing $\epsilon$ from $0.1$ to $0.05$ yielded maximum risk inflation of only $1.00032$. \begin{figure}[H] \centering \includegraphics[width=.65\textwidth]{risk_inflation_pointmass.pdf} \caption{ Risk inflation for point mass at $\lambda$, as a function of $\lambda$. $\epsilon = 0.1, \\delta = 1,\\sigma = 0.5$. The maximum $1.177$ is attained at $t=1.9$. } \label{fig:point} \end{figure} We continue our investigation with an empirical study. Hence, we chose a large number of members of a flexible parametric family of distributions, and evaluate the risk inflation on each. Specifically, we took $\delta = 1,\\sigma = 1$ and $\epsilon = 0.1$, and consider the family of mixtures of Gamma distributions $$ \left\{ \sum_{i=1}^8 w_i \text{Gam}(\alpha_i): 0\leq w_i\leq 1, \sum_{i=1}^8 w_i = 1 \right\} $$ where $\boldsymbol{\alpha} = (0.1, 0.8, 1.5, 2.2, 2.9, 3.6, 4.3, 5.0)$ and where $\text{Gam}(\alpha)$ is the Gamma distribution with shape parameter $\alpha$ and rate one. We then consider a restriction of the original family, namely, the subset $$ \left\{ \sum_{i=1}^8 (u_i/\sum u_j) \text{Gam}(\alpha_i): u_i \in \{0,1,2,3,4\} \right\} $$ which includes 383,809 unique members. A few of these distributions are plotted in color in Figure \ref{fig:mix} (the black curves correspond to distributions which belong to the original mixture family but not the restricted one). Next, for each of these 383,809 distributions $\Pi^*$ we compute the ratio $$ \frac{R^{\infty}(\lambda^{\infty}_{\text{KO}}(q); \Pi^*, \epsilon)}{\inf_{\lambda}R^{\infty}(\lambda; \Pi^*, \epsilon)} $$ numerically. For $q=0.7$, this ratio was always below $1.13$. Figure \ref{fig:hist-ratio} shows a histogram of the 383,809 values: the distribution with the maximum risk inflation appears in Figure \ref{fig:mix} in red; the ``runner-ups" look similar. We also computed the ratio for some distributions that belong to the original mixture family but not the restricted one, and these are shown in black in Figure \ref{fig:mix}; the ratio for both was less than $1.12$. To summarize, for $q=0.7$ the risk inflation did not exceed $17\%$ in any of our examples. \begin{figure}[H] \centering \includegraphics[width=.5\textwidth]{mixture.pdf} \caption{ Gamma mixture distributions. The colored lines represent distributions in the restricted family of 383,809 distributions. Black lines are distributions that belong to the original mixture family, but not the restricted one. The red curve corresponds to the distribution that attained the maximum ratio among the 383,809 distributions. Numbers in legend are risk inflation values corresponding to the different distributions in the plot. } \label{fig:mix} \end{figure} \begin{figure}[H] \centering \includegraphics[width=.65\textwidth]{hist.pdf} \caption{ Histogram of the 383,809 computed risk ratios. } \label{fig:hist-ratio} \end{figure} %\ejc{Do we have numbers for $\epsilon = 0.05?$ AW: We don't---we'd need to redo this on the cluster if we wanted} \section{Discussion}\label{sec:discussion} The Lasso is nowadays frequently used for selecting variables in a linear model. Here we have considered the controlled variable selection problem, where a valid procedure must satisfy that the expected proportion of incorrectly selected variables (the FDR) is kept below a pre-specified level $q$. This connects our work to that of \citet{tibshirani2016exact} and \citet{fithian2015selective}, who consider hypothesis testing for procedures which sequentially select variables into a model. However, the null hypothesis tested in these papers is different, corresponding to an ``incremental" test versus a ``full-model" test; see the elaborate and illuminating discussion in \citet{fithian2015selective}. %\begin{figure}[H] %\centering % \includegraphics[width=.8\textwidth]{tradeoff-distributions.pdf} %\caption{ %Tradeoff diagrams for different $\Pi^*$, with $\epsilon = 0.2$, $\delta = 1$ and $\sigma = 0.5$. %} %\label{fig:tradeoff-distributions} %\end{figure} We have proposed an adaptation of the Lasso-based knockoff procedure of \citet{barber2015controlling}, which has provable FDR guarantees. Under the working assumptions, which include a Gaussian i.i.d.~design, it is possible to analyze the asymptotic tradeoff between the false discovery proportion and power of the proposed procedure. Also, while the i.i.d.~$\calN(0,1)$ design is perhaps not realistic, our aim is to compare knockoffs against the best possible mechanism for selection over the lasso path, and such a mechanism has been developed explicitly only for the i.i.d.~$\calN(0,1)$ setting. We have demonstrated that the proposed procedure comes close in performance to an oracle procedure, which uses knowledge about the underlying signal to choose the Lasso tuning parameter in such a way that FDP is controlled. Moving away from the testing problem, we have also demonstrated that FDR calibration with knockoffs has potential in achieving good prediction performance, adaptively in the sparsity $\epsilon$ and in $\Pi^*$. In this article the focus was on procedures tied to the Lasso path. There is certainly room for investigating the use of other test statistics to measure feature importance. Using the Lasso statistics $T_j$ in \eqref{eq:stat} for the knockoff procedure might be suboptimal, as can be seen in Figure \ref{fig:tradeoff-distributions} and as was implied in \citet{su2017false}: in the Lasso path false discoveries are interspersed between true ones. As a matter of fact, an optimal oracle procedure exists for the setup considered in the current paper. Namely, the procedure which rejects the hypotheses with $T^*_j \leq \tau$ where $$ T_j^* = \PP(\beta_j = 0|\by, \bX) $$ is the ``local" false discovery rate at $(\by,\bX)$ and where $\tau$ is selected so that the procedure has FDR control at $q$. \citet{candes2016panning} have already proposed to use similar Bayesian statistics. Note that $T_j^*$ defined above depends on $\Pi^*$ and $\epsilon$, hence we associate it with an oracle. It would be interesting to think if there is a way to mimic this oracle with a computationally feasible empirical Bayes procedure. Even if we commit to Lasso statistics, it might be possible to increase power by ordering the variables differently. For example, \citet{candes2016panning} suggested to use the absolute value of the Lasso coefficient, evaluated at the cross-validated $\lambda$, to measure feature importance. We leave for future work the investigation of the potential gain in power in using alternative knockoff statistics. \appendix \section{Proofs} \begin{proof}[Proof of Lemma \ref{lem:hyper}] By definition, $$ \EE Y = \sum_{1\leq k\leq m} \frac{k}{1+m-k} \cdot \frac{{n_0 \choose k} {n_1\choose m-k}}{{n_0+n_1 \choose m}} $$ Assuming $n_0, n_1>0$, for $1\leq k\leq m$ we have $$ k {n_0\choose k} = n_0{n_0-1 \choose {k-1}},\ \ \ \ \ \ \ \ \ \ \frac{1}{1+m-k}{n_1\choose {m-k}} = \frac{1}{n_1 + 1} {n_1+1 \choose {m-k+1}}. $$ Therefore, $$ \EE Y = \frac{n_0}{1+n_1}\cdot \frac{1}{ {n_0+n_1\choose m} } \sum_{1\leq k \leq m} {n_0-1 \choose {k-1}} \cdot {n_1+1 \choose {m-k+1}} = \frac{n_0}{1+n_1} \left( 1 - \frac{ {n_0-1 \choose m} }{ {n_0+n_1\choose m} } \right). $$ If $n_0 = 0$, then $X=Y=0$. If $n_1 = 0$, then $X=Y=m$. In both cases the claimed result still holds. \end{proof} \bigskip \begin{proof}[Proof of identity \eqref{eq:hypergeometric-identity}] Assuming $m_0,r>0$, we have \begin{align*} &\EE \left( \frac{X}{1+m-X} \cdot \frac{r+X-m}{1+m_0-X} \right) = \sum_{ k = 1\vee (m-r+1) }^{m\wedge m_0} \frac{k}{1+m-k} \cdot \frac{r+k-m}{1+m_0-k} \cdot \frac{{m_0\choose k}{r\choose {m-k}}}{{m_0+r\choose m}}\\ &=\sum_{ k = 1\vee (m-r+1) }^{m\wedge m_0} \frac{m_0!}{(k-1)!(m_0-k+1)!} \cdot \frac{r!}{(m-k+1)!(r-m+k-1)!} \cdot \frac{m!(m_0+r-m)!}{(m_0+r)!}\\ &=\sum_{ k = 1\vee (m-r+1) }^{m\wedge m_0} \frac{ {m_0\choose {k-1}} {r\choose m-k+1} }{ {m_0+r\choose m} } = 1 - \frac{ {m_0\choose m_0\wedge m} {r\choose m-m_0\wedge m} }{ {m_0+r\choose m} } \end{align*} If $m_0 = 0$, then $X=0$ and the identity still holds. If $r = 0$, then $X=m$ and the identity again continues to holds. \end{proof} \section{Computing the curve $q^{\Pi^*}$} To compute $q^{\Pi^*}(\tppinfty(\lambda); \epsilon, \delta, \sigma)$ for specific $\Pi^*$ and $\epsilon$, we need to be able to compute the pair $(\fdpinfty(\lambda),\tppinfty(\lambda))$ for each $\lambda>0$. This, in turn, requires to find for each $\lambda$ the pair $(\alpha, \tau)$ given by the system of equations \eqref{eq:system}. Hence, the functionals $$ f(\Pi^*, \epsilon) := \EE(\eta_{\alpha\tau}(\Pi + \tau W) - \Pi)^2 = (1-\epsilon)\EE(\eta_{\alpha\tau}(\tau W) )^2 + \epsilon \EE(\eta_{\alpha\tau}(\Pi^* + \tau W) - \Pi^*)^2 $$ and $$ g(\Pi^*, \epsilon) := \PP(|\Pi + \tau W|>\alpha\tau) = 2(1-\epsilon)\Phi(-\alpha) + \epsilon \PP(|\Pi^* + \tau W|>\alpha\tau), $$ which appear in \eqref{eq:system}, are evaluated numerically for a given CDF $F_{\Pi^*}$ of $\Pi^*$. For a fixed value of $\lambda$, our code takes as an input $\epsilon$ and a CDF $F_{\Pi}$ of $\Pi$ (as well as $\delta$ and $\sigma^2$) and returns a solution for $(\alpha, \tau)$ in \eqref{eq:system}. Thus, for given values of $\alpha$ and $\tau$, we need to be able to compute $ f(\Pi, \epsilon) %:= \EE(\eta_{\alpha\tau}(\Pi + \tau W) - \Pi)^2 = (1-\epsilon)\EE(\eta_{\alpha\tau}(\tau W) )^2 + \epsilon \EE(\eta_{\alpha\tau}(\Pi + \tau W) - \Pi)^2 $ and $ g(\Pi, \epsilon) %:= \PP(|\Pi + \tau W|>\alpha\tau) = 2(1-\epsilon)\Phi(-\alpha) + \epsilon \PP(|\Pi + \tau W|>\alpha\tau),, $ defined in Section \ref{sec:tradeoff}, as functionals of $F_{\Pi}$. We explain how we implement this. Let $R(\mu):= \EE \left(\eta_{\alpha\tau}(\mu + \tau W) - \mu\right)^2$ and $Q(\mu):=\PP\left(|\mu + \tau W|>\alpha\tau \right)$. Writing $f(\Pi, \epsilon)$ as a Riemann-Stieltjes integral and using integration by parts, we have \begin{align*} f(\Pi, \epsilon) &:= \EE(\eta_{\alpha\tau}(\Pi + \tau W) - \Pi)^2 = \int_{-\infty}^\infty R(\mu) \dF_{\Pi} \\ &= R(\infty)F_{\Pi}(\infty) - R(-\infty)F_{\Pi}(-\infty) - \int_{-\infty}^\infty F_{\Pi}(\mu) \dR(\mu)\\ &= R(\infty)F_{\Pi}(\infty) - R(-\infty)F_{\Pi}(-\infty) - \int_{-\infty}^\infty F_{\Pi}(\mu) R'(\mu) d\mu\\ &= 1 + \alpha^2\tau^2 - \int_{-\infty}^\infty F_{\Pi}(\mu) [ 2\mu \left(\Phi\left(\alpha - \mu/\tau \right)\right)- \left(\Phi\left(-\alpha - \mu/\tau \right)\right)] d\mu \end{align*} where we substituted a closed-form expression for the derivative of the risk of soft thresholding \citep[see, e.g., ][Ch. 2.7]{johnstone2011gaussian}. The last expression is evaluated using numerical integration. Similarly, \begin{align*} g(\Pi, \epsilon) &:= \PP(|\Pi + \tau W|>\alpha\tau) = \int_{-\infty}^\infty Q(\mu) \dF_{\Pi} \\ &= Q(\infty)F_{\Pi}(\infty) - Q(-\infty)F_{\Pi}(-\infty) - \int_{-\infty}^\infty F_{\Pi}(\mu) \dQ(\mu)\\ &= Q(\infty)F_{\Pi}(\infty) - Q(-\infty)F_{\Pi}(-\infty) - \int_{-\infty}^\infty F_{\Pi}(\mu) Q'(\mu) \d\mu\\ &= 1- 1/\tau \int_{-\infty}^\infty F_{\Pi}(\mu) [ \phi(\alpha - \mu/\tau) - \phi(-\alpha - \mu/\tau) ] \d\mu, \end{align*} and, again, the last expression is evaluated using numerical integration. \bibliographystyle{plainnat} \bibliography{/Users/assafweinstein/Dropbox/Research/References.bib} \end{document} }
\caption{Graphs showing calculated values for $t_p(t_{\text{NL}})$ and $t_d(t_{\text{NL}})$ from simulations in \secref{sec:lifetime}. Observed values (`{\color{red} $+$}') and mean values (`{\color{green} $\bullet$}') from fit routines.}
\caption{PZT driver voltage noise spectral density. \textcolor{red}{\textbf{(red)}} noise spectrum of the PZT driver described here with no load applied, measured with the output at 5~VDC. The superimposed high frequency tail shown in \textcolor{mygreen}{\textbf{(green)}} shows the noise when a 0.7~$\mu$F capacitor is connected across the output, typical of the capacitance of a PZT. Two additional noise spectra are included for comparison: a common commercial PZT driver (Thorlabs MDT694b) \textcolor{blue}{\textbf{(blue)}}, also at 5~VDC, and the noise model presented for the design by Pisenti et al.\cite{PRR16e} \textcolor{black}{\textbf{(black)}}. The noise floor for the measuremed traces is plotted in \textcolor{mygray}{\textbf{(gray)}}.}
\caption{Absorption spectroscopy of a rubidium vapor cell. The heavy black curve in (A) shows a mode-hop-free scan over essentially the full range of the PZT driver (roughly 15~GHz for the laser used here), showing the familiar D$1$ lines. (B) Voltage \textcolor{blue}{\textbf{(blue)}} applied to the laser PZT, and transmission \textcolor{red}{\textbf{(red)}} through a confocal Fabry-Perot etalon (FSR = 1 GHz). The same scan performed with the feed-forward signal to the laser disconnected is shown on both plots in gray. Mode hops can be seen as discontinuities in the vapor cell transmission in (A), while the corresponding trace of Fabry-Perot transmission in (B) clearly shows a non-linear relationship between PZT voltage and laser frequency as well as regions of multi-mode behavior.}
\caption{ Comparison results of AP and mAP in \% of our model and the previous state of the art methods on the VOC07 dataset. The best results and second best results are highlighted in {\color{red}{red}} and {\color{blue}{blue}}, respectively. Best viewed in color.}
\caption{Comparison results of our model and the previous state of the art methods on the MS-COCO dataset. The best and second best results are highlighted in {\color{red}{red}} and {\color{blue}{blue}}, respectively. Best viewed in color.}
\caption{\label{fig:fig4} $\Delta \sigma(0.2 \text{ T})$ vs $\sigma(0)$ (a) at various fixed $V_\TG$, and (b) at various fixed $V_\BG$. There are duplicate data points in (a) and (b); for example, ``\textcolor{violet}{+}" at $\sigma(0)=637\times10^{-6} e^2/h$ in (a) and ``\textcolor{red}{$\Box$}" at the same $\sigma(0)$ in (b) are the same data point at $V_\TG=0$ and $V_\BG=34$ V. $\alpha_0$ (the prefactor for WL) and $\alpha_1$ (the prefactor for WAL) obtained from the theoretical fits (Eq.~\ref{eq:1}) at various gate voltages used in (a) and (b) are shown in (c) and (d), respectively. Eye-guiding lines of the universal curves are shown as solid smooth lines. }
\caption{$L_2$ and $L_\infty$ errors for each flow field variable predicted by the GAN after a single time step using ground truth inputs as a function of time-step sizes of $\Delta t U_{\infty}/D = 0.096$, $0.192$, $0.288$, $0.384$, and $0.480$. $L_{2}$ errors for $u/U_{\infty}$, $v/U_{\infty}$, and $p/\frac{1}{2}\rho U_{\infty}^{2}$ are denoted by \textcolor{red}{$-\square-$}, \textcolor{green}{$-\square-$}, and \textcolor{blue}{$-\square-$}, respectively; $L_{\infty}$ errors for $u/U_{\infty}$, $v/U_{\infty}$, and $p/\frac{1}{2}\rho U_{\infty}^{2}$ are denoted by \textcolor{red}{$-\diamond-$}, \textcolor{green}{$-\diamond-$}, and \textcolor{blue}{$-\diamond-$}, respectively. \textbf{\textcolor{black}{---}}, \textbf{\textcolor{red}{---}}, \textbf{\textcolor{green}{---}}, and \textbf{\textcolor{blue}{---}} denote the 1st, 2nd, 3rd, and 4th order slopes.}
\caption{$L_2$ and $L_\infty$ errors for each flow field variable predicted by the GAN using recursive inputs with time-step size of $\Delta t U_{\infty}/D = 0.096$. Flow fields are predicted on time $t U_{\infty}/D = 0.096$, $0.192$, $0.288$, $0.384$, and $0.480$. $L_{2}$ errors for $u/U_{\infty}$, $v/U_{\infty}$, and $p/\frac{1}{2}\rho U_{\infty}^{2}$ are denoted by \textcolor{red}{$-\square-$}, \textcolor{green}{$-\square-$}, and \textcolor{blue}{$-\square-$}, respectively; $L_{\infty}$ errors for $u/U_{\infty}$, $v/U_{\infty}$, and $p/\frac{1}{2}\rho U_{\infty}^{2}$ are denoted by \textcolor{red}{$-\diamond-$}, \textcolor{green}{$-\diamond-$}, and \textcolor{blue}{$-\diamond-$}, respectively. \textbf{\textcolor{black}{---}} and \textbf{\textcolor{red}{---}} denote the 1st and 2nd order slopes.}
\caption{{\bf State-recycling techniques.} {\bf A,} Simplified sketch illustrating the experimental technique to detect `time' evolution in a photonic lattice through state-recycling. Both the ``linear" and ``ring" recycling schemes are illustrated. The propagation distance is the analogous time ($z\!\leftrightarrow\!t$). {\bf B,} One-dimensional driven lattice with nearest-neighbour couplings $J_{1, 2}$, which are varying periodically in time. {\bf C,} Photonic implementation of the driven lattice in (B). Here, the state-recycling is performed using a linear cavity, the facets of which are indicated by two parallel blue planes. {\bf D,} Floquet spectrum of the driven lattice in (B, C) consists of two linearly dispersive bands for the following driving protocol: $J_{1, 2}\!=\!0, \pi/T$ for $0\! \le \! t \le \! T/2$ and $J_{1, 2}\!=\! \pi/T, 0$ for $T/2\! \le \! t \! \le \!T$ where $T$ is the driving period. {\bf E-I,} Excitation of a linearly dispersive band. Experimentally observed intensity distributions at $t\!=\!(1/2+N)T$, $N\!=\! 0, 1, 2, 3$. The red square indicates the waveguide which was excited at the input. {\color{Seba}The effective propagation distances are indicated on each image.} \label{fig_data}}
\caption{ {\color{Seba}{\bf Discrete diffraction in the presence of a synthetic electric field.}{\bf A-D,} Quasi-real-time evolution of light intensity in a 1D straight photonic lattice consisting of twenty coupled single-mode waveguides. The small facet angles of the substrate (inset in E) cause a linear phase shift along the lattice at each facet, which effectively produces a time-periodic (pulsed) synthetic electric field. A-D show the output intensity distributions that were experimentally obtained after different effective propagation distances indicated on each image. Light was launched at a single waveguide indicated by the red square. {\bf E-H,} Comparison between the experimental observations in (A-D) and the associated numerical results. The vertical axis is the normalized optical intensity and the horizontal axis is the waveguide number. The measured intensity patterns (red bars) agree with the numerical simulations (blue) upon adding the effects of a pulsed (synthetic) electric field. The dynamics corresponds to approximately half a period of a Bloch oscillation, which is produced by the synthetic electric field. For comparison, we show the numerical results when omitting the pulsed electric field (cyan bars), where the Bloch oscillation is absent. } \label{fig_discrete}}
\caption{{\bf Quasi-real-time propagation of topological edge modes.} {\bf A,} White-light micrograph of the facet of the driven square lattice. {\bf B-E,} One-way (here, counter-clockwise) propagation of the edge modes for effective times, $t\!=\!2T, \ 4T, \ 6T$ and $8T$; {\color{Seba}here $2T\!\leftrightarrow\!L\!=\!70 \ $mm.} The edge modes are excited with $\approx85\%$ efficiency by exciting the $(4, 1)$ site on the bottom-right edge (indicated by the red square). The edge modes are neither back-scattered by a corner nor by the defect [here, a missing waveguide at the $(8, 4)$ site]. {\bf F-H,} Time evolution of the bulk state after times, $t\!=\!2T, \ 4T$ and $6T$. The weakly dispersive bulk bands are equally excited by launching light at the $(4, 5)$ site. The delocalization of the state becomes evident after long detection times. \label{fig_EDGE_BULK}}
\caption[]{ \label{fig:map_SM} (color online) Recorded optical spectral holeburning spectra for site~II of \ybiso{} crystal measured for different magnetic field amplitudes applied in the direction close to $\mathbf{D}_2$-axis. Zero frequency detuning corresponds to the central frequency at which spectral holeburning is performed. Black regions correspond to lower absorption (holes), while white lines correspond to lower transmission (antiholes) regions. White and black dashed lines indicate energy level splittings of ground and excited states, respectively. \red{The deviation from the data is attributed to the imperfect calibration of the laser scan.} }
\caption[]{\label{fig:abs} (color online) \red{Energy level diagram of \ybiso{} showing the zero-field hyperfine level structures of the lowest crystal-field levels of the ground \gstate{}(0) and excited \estate{}(0) states, both for site~I (a) and site~II (c). The ground-state hyperfine level structures were determined from the zero-field ODMR measurements, while the excited-state structure was determined using SHB measurements (see Section~\ref{sec:results}). In (b) and (d) we show high-resolution optical absorption spectra of the \transition{} transition for site~I and site~II, respectively. These were recorded at low temperature (3~K) for light polarized along $\mathbf{D}_2$ (solid line) and $\mathbf{D}_1$ (dashed line) crystal axis. The position of all optical-hyperfine transitions have been calculated using the hyperfine $\mathbf{A}$ tensors measured in this work (dashed vertical lines). Zero optical detuning refers to the central transition wavelength given in Sec. \ref{sec:crystal}.}}
\caption[]{\label{fig:map} (color online) Experimental results. (a) Optically detected magnetic resonance (ODMR) lines at zero magnetic field for site~I of \ybiso{}. The oscillating magnetic field $B_{\text{ac}}$ was applied in $\mathbf{b}$ direction. Three different transitions are shown and their corresponding levels are depicted on the right. The variation of the measured spin linewidths could be attributed to the power broadening effect. (b) Recorded optical spectral holeburning (SHB) spectra of site~I of \ybiso{} crystal measured for different magnetic field amplitudes applied in the direction close to $\mathbf{D}_1$-axis. Zero frequency detuning corresponds to the central frequency at which spectral holeburning is performed. Black regions correspond to lower absorption (holes), while white lines correspond to lower transmission (antiholes) regions. \red{The bending of some of the holes and anti-holes at around 30~mT of magnetic field is a result of avoided crossings of the associated ground-state levels. Many of these also show an unusual a hole-antihole transformation in this field region (see also \figref{fig:hole} and Sec. \ref{sec:holes}).}}
\caption[]{\label{fig:hole} (color online) \red{Example of the hole-antihole transformation. (a) Shown is the transformation of the antihole corresponding to the $\ket{2_g}\longleftrightarrow\ket{3_g}$ ground-state transition (at 2046 MHz for zero field). The data is taken from the \figref{fig:map} for magnetic field amplitudes around 30~mT. We believe this transformation to be due to an increased relaxation rate of this transition, which in turn is caused by the rapid change in wavefunctions in this field region (see (b)). The dashed line represents the predicted spectral position of the antihole. The discrepancy between the measured and calculated values can be attributed to the accuracy of the calibration of the laser scan. (b) The change in wavefunction of states $\ket{2_g}$ and $\ket{3_g}$ can be visualized by plotting their overlap with the separable states $\ket{\uparrow \Downarrow}$ and $\ket{\downarrow \Downarrow}$, where $\ket{\uparrow,\downarrow} \equiv \ket{S_z = \frac{1}{2}, -\frac{1}{2}}$ and $\ket{\Downarrow,\Uparrow} \equiv \ket{I_z = \frac{1}{2},- \frac{1}{2}}$. Note that the negligible overlap with the remaining basis states $\ket{\uparrow \Uparrow}$ and $\ket{\downarrow \Uparrow}$ are not shown. All calculations were based on diagonalizing the spin Hamiltonian for the ground state.}}
\caption{The singular values of the full sensitivity. \textcolor{red}{Red color} signifies for $h$, i.e. plume observation while the \textcolor{blue}{blue color} is for $\rmd z$, interface observations. Full line is using all steps while the dashed line is using steps $1,5,9,12$. %Left plot is for the original model, while right is for %the matched model. }
\caption{The singular values of the full sensitivity. \textcolor{red}{Red color} signifies for plume observation while \textcolor{blue}{blue color} is gravity, $h$.}
\caption{Comperaing different optimal design criterieas including timestep $1$,$5$,$9$ and $10$. The optimal points shown in \textcolor{red}{red} is for step $\n$}
\caption{Optimal design for the original model with standard with the original assumption of the errors for gravity and plume data. The \textcolor{red}{red} squares show the optimal design usind D-optimal criteria with $20$ possible observation. The upper row show the response singular value corresponding to plume and the lower row for the gravity. The columns represent the second, third and forth singular value.}
\caption{Three traps create three rings of magnetic nanoparticles. The rings interact with one another (see \textcolor{blue}{Visualization 3}). [From Masajada et al., Opt. Lett. \textbf{38}, 3910 (2013).]}
\caption{Three traps create three rings of magnetic nanoparticles. The rings interact with one another (see \textcolor{blue}{Visualization 3}). [From Masajada et al., Opt. Lett. \textbf{38}, 3910 (2013).]}
\caption{Experiments with simulated graphs with $|V|=25$ vertices. Each field in the table summarizes the per-graph wall-time over a set of 1000 graphs as well as the per-graph number of recursive calls, except for the brute-force method. The columns represent pivot selection strategies: (i) random, (ii) random multi-parent vertex (mpv), (iii) minimum bound, (iv) maximum degree, and (v) bottleneck. The rows represent successive additions of practical modules for speedups: (\protect\drawcircle{white}) basic approach from Algorithm \ref{algo:dag}, (\protect\drawcircle{myred}) pruning, (\protect\drawcircle{myred}\;\protect\drawcircle{myyellow}) pruning and hashing, (\protect\drawcircle{myred}\;\protect\drawcircle{myyellow}\;\protect\drawcircle{myblue}) pruning, hashing, and graph reversal.}
\caption{\label{fig:spatial_convergence} Each of the subplots shows the reflectivity $R$ vs number of grid points $N_x$ for different $N_y$ (\opensquare : $ N_y = 2^{5}$, \opencircle : $N_y = 2^{6}$, \opentriangle : $N_y = 2^7$, $\star$ : $N_y = 2^8$). \textcolor{black}{Spatial convergence in the $x$ direction is seen along the $x$ axis of the panels. Convergence along the $y$ direction is seen following the different symbols in each panel at fixed $N_x$}. Each row corresponds to a different value of the cut-off length $\Delta$. The first column shows the behaviour of $R$ in logarithmic scale, whereas the second shows a zoom into the convergence regime in linear scale. The amplitude of the corrugation is $A/L=0.1$.}
\caption{Panel (a) shows the atomic population for superpositions of two, $|0\rangle+ e^{i\pi/2}|1\rangle$ (dotted line), and seven, $\sum_{n=-3}^{3}e^{-in\pi/2}|n\rangle$ (dashed line) plane waves. Panel (b) shows atomic population for initial states composed of superpositions of nonconsecutive momentum states: $e^{-i\pi}|-2\rangle+e^{-i\pi/2}|-1\rangle+e^{i\pi/2}|1\rangle+e^{i\pi}|2\rangle$ (dashed line), $e^{-i\pi}|-2\rangle+e^{i\pi}|2\rangle$ (dotted line), and $e^{-i\pi/2}|-1\rangle+e^{i\pi/2}|1\rangle$ (dash-dotted line). For both panels, the solid lines represent the spatial distribution of the standing wave intensity. \red{From \cite{ni2016}}.}
\caption{Panel (a) shows theoretical data for the atomic spatial distribution FWHM as a function of the number of consecutive plane waves in the atomic state. \red{Panel (b) shows the effective force of the standing wave as a function of the momentum state range. The circles represent the case where the momentum states are consecutive, diamonds the situation where one state is missing within the range.} Panel (c) shows the effective force of the standing wave as the function of the offset phase $\gamma$ with a initial state containing seven consecutive momentum states $\sum_{n=-3}^{3}e^{-in\pi/2}|n\rangle$. }
\caption{\red{Schematic showing the geometry of the experiment and the sequence of Bragg and kick pulses used to study quantum ratchets.}}
\caption{Panel (a) illustrates the experimental results of the variation of the normalized dispersion versus number of kick with $\beta=0.5$ and $T=T_{1/2}$ for ratchets with initial states $\sum_{n=-3}^{3}|n\rangle$ (squares), $\sum_{n=-1}^{2}|n\rangle$ (triangles), $\sum_{n=-1}^{1}|n\rangle$ (diamonds), and $\sum_{n=0}^{1}|n\rangle$ (circles). Panel (b) illustrates the experimental results of the variation of the mean momentum versus the number of plane waves with $t=5$, $\beta=0.5$, and $T=T_{1/2}$ for ratchets. Momenta are plotted in units of two-photon recoils. Error bars are small \red{compared to the plotted point size} and are not shown here. \red{Panel (a) from \cite{ni2016}.}}
\caption{Experimental results showing variation of the mean momentum versus kick number with $\beta=0.5$ and $T=T_{1/2}$ for ratchets with initial states $\sum_{n=-2}^{2}|n\rangle$ (asterisks), $\sum_{n=-1}^{1}|n\rangle$ (diamonds), and $\sum_{n=0}^{1}|n\rangle$ (circles). Momenta are plotted in units of two-photon recoils. Error bars are small \red{compared to the plotted point size} and not shown.}
\caption{Experimental momentum distributions after kicking the initial state $|0\rangle+e^{-i\gamma}|1\rangle$ by the short pulses of an off-resonant optical standing wave. The distributions are displayed as a function of (a) offset from resonance pulse period (10 kicks, $\phi_{d}=2.6$, and $\gamma=-\pi/2$) and (b) kick number ($|\varepsilon|=0.18$, $\phi_{d}=1.8$, and $\gamma=-\pi/2$). Image of each column was captured in a separate time-of-flight experiment. In both panels, weighting of the distributions are towards positive momenta at small values of each control variable followed by a tendency towards negative mean momenta around the middle of the horizontal axis range. In (b) the mean momenta tends asymptotically to 0 due to decoherence effects from spontaneous emission after about 15 kicks. \red{From \cite{shrestha2012controlling}.}}
\caption{Scaled mean momentum $\langle p\rangle/(\phi_{d}t\sin\gamma)$ as a function of the dimensionless scaling variable $z=\sqrt{\phi_{d}|\varepsilon|}t$ for $l=0$. $z$ was varied by scanning over kick number $t$ for different combinations of $\phi_{d}$ and $|\varepsilon|$. The solid line is the function $S(z)/z$ given by Eq. (10). \red{Adapted from \cite{shrestha2012controlling}.}}
\caption{Scaled mean momentum $\langle p\rangle/(\phi_{d}t\sin\gamma)$ as a function of the scaling variable $z=\sqrt{\phi_{d}|\varepsilon|}t$ for $l=1$. $z$ was varied by scanning over $|\varepsilon|$ with $t=10$ and $\phi_{d}=1.8$ (circles), and $\phi_{d}$ with $|\varepsilon|=0.18$ and $t=8$ (triangles), and $t$ with $|\varepsilon|=0.18$ and $\phi_{d}=1.8$ (asterisks). The solid line is the function $S(z)/z$ given by Eq. (10). This demonstrates that no matter how $z$ is obtained the scaled mean momentum is approximately universal. \red{Adapted from \cite{shrestha2012controlling}.}}
\caption{ Spin-selective electron transmission through a metallic $N=6$ CNT wrapped with a right- or a left-handed molecule, $ \varphi=\pm \pi/6$. The spin polarization $P = \left(T_{\uparrow}-T_{\downarrow}\right) / \left(T_{\uparrow}+T_{\downarrow}\right)$ is computed from the transmission probabilities $T_{s}$ for electrons of spin $s$. It is substantial at energies where the partial gap opens (shown on the sides). The two cases, a right- and a left-handed molecule, exhibit spin polarization with equal magnitude but opposite sign. } \label{figure4} \end{figure} We now turn to analyze the effects of a DNA molecule. The SOC terms acquire additional contributions that are non-zero only along lines defined by $y=(x+Nam)\tan{\varphi}$ where $m \in \mathbb{Z}$, and the wrapping angle $\varphi\in[-\pi/2,\pi/2]$ is positive (negative) for right (left) handed molecules. As a result, the SOC terms in Eq.~\ref{eq:SOCmom} are no longer diagonal in the momentum along the $y$-direction. For example, taking the wrapping angle to be $\varphi=\pi/6$, the additional contributions to the SOC are \begin{align} \label{eq:SOCmomMol} &\delta{h}_{\text{so}_1}\propto se^{i\frac{k_ya}{2\sqrt{3}}}\sin\left[\frac{\pi \ell}{N}\right]\delta_{k_y+Q(\ell-\ell'),k_y'} \delta_{s,s'}\hspace{0.5mm},\\\nonumber &\delta{h}_{\text{so}_2}\propto e^{i\frac{k_ya}{2\sqrt{3}}}\cos\hspace{-1mm}\left[\frac{\pi(2\ell+s)}{2N}\right] \delta_{k_y+Q(\ell+s-\ell'),k_y'} \delta_{-s,s'}\hspace{0.5mm}, \end{align} where $Q(x)=2\pi x/\sqrt{3}Na$. The broken translation symmetry suggest using a reduced Brillouin zone, with bands labeled by $n$ in addition to $\ell$, $s$ and the reduced momentum $k_y^{\text{red}}$ where $k_y=k_y^{\text{red}}+Q(n)$. Bands that cross in the reduced Brillouin zone couple when satisfying the delta-function involving $Q$ in Eq.~\ref{eq:SOCmomMol}. Thus, as before, the states $(\ell,\uparrow)$ and $(\ell+1-n,\downarrow)$ mix but not $(\ell,\downarrow)$ and $(\ell+1-n,\uparrow)$. The $2\times2$ Hamiltonian in the vicinity of these energies is $\mathcal{H}=\Psi_{\ell,\ell',\eta}^{\dagger}(q)h_{\ell,\ell'}^{\eta}(q)\Psi_{\ell,\ell',\eta}(q)\delta_{k_y-Q(n),k_y'}\delta_{\ell',\ell+s-n}$, where $h_{\ell,\ell'}^{\eta}(q)$ has the same structure as in Eq.~\ref{eq:Linearized1}. The main differences between this Hamiltonian and the case without a molecule (discussed above) is in the definition of the states $\Psi_{\ell,\ell',\eta}^T=[\psi_{\ell,k_0+q,\uparrow}^{\eta},\psi_{\ell',k_0+q-Q(\ell+s-\ell'),\downarrow}^{\eta}]$ and $G_{\ell,\ell'}(q\rightarrow0)\neq 0$. \begin{figure} \includegraphics[width=\linewidth]{Figure_5.pdf} \caption{ Spin-dependent transmission through clean (a) and weakly disordered (b) semiconducting $N=11$ CNTs of length $154a/\sqrt{3}$ wrapped with DNA at $\varphi=\pi/6 $. The transmission is significant only above the semiconducting gap $\approx0.17t$. In both cases, transmission of $s=~\downarrow$ electrons is strongly suppressed compared to $s=~\uparrow$ at energies within the partial gap. In the clean case (a) the spin polarization diminishes at below $\approx0.18t$. In contrast, in the weakly disordered case (b) significant spin polarization persists to the lowest energies above the band gap.} \label{figure5} \end{figure} The spectrum of the full Hamiltonian, Eqs.~\ref{eq:Hopping H}-\ref{eq:Sx}, in the presence of the molecule (encoded in the position-dependent SOC) is presented in Fig.~\ref{figure3}. There we show the energy spectra for metallic ($N=3$) and semiconducting ($N=5$) zig-zag CNTs wrapped with a DNA molecule at $\varphi=\pi/6$. The couplings are taken to be $\Gamma_1=3\cdot10^{-4}t/N$ and $\Gamma_2=3\cdot10^{-2}t/N$ on all sites unaffected by the molecule (corresponding to $\Delta_{\text{ASOC}}\approx5\cdot{10}^{-3}t$). On sites underneath the molecule, we took these parameters to be larger by a factor of 10. The relevant parameters for experimental systems are discussed below. The opening of helicity-dependent gaps of size $0.03t$ (metallic) and $0.01t$ (semiconducting) is clearly visible in Fig.~\ref{figure3}---for comparison, see Fig.~\ref{figure2} for the spectra in the absence of DNA molecules. Within the energies of the partial gap, only states of one helicity remain. We note that no helicity-dependent gaps open near the neutrality point of the metallic CNT. \section{Spin dependent transmission} To study the spin-filtering properties of CNTs wrapped with DNA, we numerically calculate the transmission probability per spin, $T_{s}$, as a function of energy. For this purpose, we assume the molecule wraps only a segment of the CNT, and let the uncovered parts acts as electrodes. In Fig.~\ref{figure4} we show the spin-polarization, $P=(T_{\uparrow}-T_{\downarrow})/(T_{\uparrow}+T_{\downarrow})$, for electron transport through a metallic CNT ($N=6$) as a function of energy. A clear window with spin polarization of more than $90\%$ occurs at the energies where helicity dependent gaps open. For the transport calculation, we took large SOC parameters, $\Gamma_1=10^{-3}t$ and $\Gamma_2=10^{-1}t$ on sites directly below the molecule and $10$ times smaller elsewhere. This choice of parameters allowed us to perform our simulation on relatively short molecules. This is because the spin polarization increases exponentially with length, $P\sim P_{\text{max}} - e^{-r/\xi}$, where $\xi$ is inversely proportional to the partial gap (for the full length dependence see Appendix~\ref{appendixC}). For metallic CNTs, the first partial gap opens at a relatively high energy. In the experiment,~\cite{Pramanik2015,Pramanik2017} in contrast, semiconducting CNTs have been measured and the largest spin polarization was observed at the lowest voltage where current flows. The spectrum of a representative semiconducting CNT shows a helicity-dependent gap opening not far from the bottom of the conduction band, see Fig.~\ref{figure3}. Consequently, a clean semiconducting CNT does not exhibit significant spin filter at the lowest energies within our model. However, the band below the partial gap is relatively narrow and is thus expected to easily localize in the presence of disorder. To test this, we calculated the transmission of such a CNT with a realistic circumference ($N=11$) and weak chemical potential disorder $\delta \mu \in[-0.05t,0.05t]$. The spin-dependent transmissions in the clean and weakly disordered cases are shown in Fig.~\ref{figure5}. The presented data is obtained by averaging $T_s$ over 25 realizations, and only small sample variations were observed. We find that, unlike the clean case, the disordered one exhibits significant spin-polarization at the lowest energies, similar to the experimental findings. Our discussion so far has focused on the effect of the molecule within a low-energy single-orbital model. Allowing for hybridization of $\pi$ and $\sigma$ orbitals induced by the molecule potential gives rise to a Rashba-type SOC.~\cite{Guinea2006,Diniz2012} The effect of such a SOC, on spin transport of metallic CNTs wrapped with DNA molecules has been analyzed in Ref.~\onlinecite{Diniz2012}. There, spin filtering has been found near the neutrality point. As we explicitly show in appendix~\ref{appendixB}, replacing the curvature induced SOC with the Rashba type gives qualitatively the same result as in our earlier discussion. This is not surprising since the only difference between the Rashba SOC and Eqs.~\ref{eq:Sy}-\ref{eq:Sx} is that the former does not vanish for hopping parallel to the cylinder axis ($y'$-direction). These additional contributions, however, do not intertwine the spin and angular momentum. In other words, although they include spin flips their effect is closer to $h_{\text{so}_1}$ than to $h_{\text{so}_2}$. Consequently, the Rashba SOC does not open any new helicity-dependent gaps, and does not give rise to further spin filtering on top of the one discussed above. \begin{figure} \includegraphics[width=\linewidth]{Figure_6.pdf} \caption{(a) Twisted CNTs can be viewed as rolled-up sheets of distorted graphene. Specifically, we tilt the vector $\boldsymbol{\delta}_1$ by an angle $\varphi\approx \pi/5$, while $\boldsymbol{\delta}_{2,3}$ are unchanged. The corresponding spectrum for a metallic $N=6$ CNT with $\Gamma_1=2\cdot10^{-1}t$ and $\Gamma_2=2\cdot10^{-3}t$ on all sites is plotted in (b) with a zoom-in on the helicity-dependent gap in the inset. } \label{figure6} \end{figure} \section{Connection to experiment} The above theoretical analysis suggests that the spin-selective transport measured in Refs.~\onlinecite{Pramanik2015,Pramanik2017} results from the partial gaps that open due to the combined effects of the curvature-induced SOC and of the DNA molecule. To estimate the size of these helicity-dependent gaps within our model, knowledge of the effective SOC parameters underneath the molecule is required. As shown in appendix~\ref{appendixA}, the effective SOC is not solely determined by the direct change in $\Gamma_{1,2}(\mathbf{r})$; it also incorporates the electrostatic potential generated by the molecule, as well as the modified hopping amplitudes. Precise values of the parameters can only be determined via ab-initio calculations, but a rough estimate can be obtained as follows. The potential on the carbon sites due to the molecule was found~\cite{Yarotski2009} to be around $0.1eV$. In appendix~\ref{appendixA} we show that the size of the partial gap induced by a helical potential of strength $0.5t$ is equivalent to the one due to SOC with $\Gamma_2=3\cdot10^{-1}t/N$ underneath the molecule. For graphene, $t\approx2.7eV$, which yields gaps of size $1meV-10meV$ in CNTs with radius of a few nanometers. This is consistent with the characteristic energy scale extracted from the observed temperature dependence of spin-polarization in transport, which is substantial up to $40K$. The voltage-dependence of the spin-selective transport is more complicated, and its analysis likely requires including electron-electron interactions. \section{Twist-induced spin selectivity} Our analysis of CNTs wrapped with DNA molecules suggests that any spiral perturbation will lead to similar modifications of the spectrum and give rise to spin-dependent transport. An alternative way would be twisting the CNT. For a zig-zag CNT this can be encoded by rotating the $\boldsymbol{\delta}_1$ bond by an angle $\varphi$ (see Fig.~\ref{figure6}), i.e., the lattice coordinates become $(x,y)\rightarrow (x+y\sin\varphi,y)$ while the basis vector $\boldsymbol{\delta}_2$ remains unaltered. Such a twist modifies the spin-independent hopping $t_1$ as well as the SOC. To focus on spin-related phenomena, we neglect the change in $t_1$ and analyze the SOC given by Eqs.~\ref{eq:Sy} and~\ref{eq:Sx}. The twist renders the angle $\theta_r=x+y\sin\varphi$ in $\mathcal{H}_{\text{so}_2}$ spatially non-uniform (in both directions), and results in non-zero SOC terms on all bonds. In momentum space $\mathcal{H}_{\text{so}_{1},\text{so}_{2}}={h}_{\text{so}_{1},\text{so}_{2}}^{s,s'}(\ell,\ell';k_y,k_y') A_{\ell,k_y,s}^{\dagger} B_{\ell',k_y',s'}+\text{h.c.}$, with \begin{align} \label{eq:SOCmomTwist1} {h}_{\text{so}_1} \propto &~s (\hat{\delta}_i\cdot\hat{x}) e^{-i\mathbf{k}\cdot\boldsymbol{\delta}_i}\delta_{k_y,k_y'} \delta_{s,s'}\left(\delta_{\ell,\ell'}+\delta_{2N-\ell,\ell'}\right)\hspace{0.5mm}, \end{align} and \begin{align} \label{eq:SOCmomTwist2} {h}_{\text{so}_2} \propto &~(\hat{\delta}_i\cdot\hat{x})^2(\hat{\delta}_i\cdot\hat{y}) e^{-i\mathbf{k}\cdot\boldsymbol{\delta}_i-i\pi s(\hat{\delta}_i\cdot\hat{x})/N} \delta_{-s,s'}\\\nonumber &\times \delta_{k_y+\tilde{Q}(\ell+s-\ell'),k_y'} \left(\delta_{\ell+s,\ell'}+\delta_{2N-\ell+s,\ell'}\right)\hspace{0.5mm}, \end{align} where $\tilde{Q}(x)=2\pi x\sin\varphi/Na$. The outcome of such SOC terms can be analyzed analogously to Eq.~\ref{eq:SOCmomMol}. Specifically, the $2\times2$ Hamiltonian near these crossing points is $\mathcal{H}=\Psi_{\ell,\ell',\eta}^{\dagger}h_{\ell,\ell'}^{\eta}\Psi_{\ell,\ell',\eta}\delta_{k_y-\tilde{Q}(n),k_y'}(\delta_{\ell',\ell+s}+\delta_{\ell',2N-\ell+s})$. Here, $h_{\ell,\ell'}^{\eta}(q)$ is given by Eq.~\ref{eq:Linearized1}, and $\Psi_{\ell,\ell',\eta}^T=[\psi_{\ell,k_0+q,\uparrow}^{\eta},\psi_{\ell',k_0+q-\tilde{Q}(\ell+s-\ell'),\downarrow}^{\eta}]$. For small twist angles, $\varphi\ll1$, the crossing points shift by $\delta k_0\propto2\pi(\ell+s-\ell')\varphi/Na$ compared to their position in a bare CNT. There, we saw that $ G_{\ell,\ell'}(q)$ vanished at the crossings. This is no longer the case once these points are shifted, and $G_{\ell,\ell'}(q)\propto\Gamma_2\varphi$. Consequently, partial gaps open as shown in Fig.~\ref{figure6}. There, a helicity-dependent gap of size $0.03t$ arises for metallic CNT with $N=6$, $\Gamma_1=2\cdot10^{-3}t$, $\Gamma_2=2\cdot10^{-1}t$ and $\varphi=\sin^{-1}(1/\sqrt{3})\approx\pi/5$. Twisted CNTs with angles up to $\pi/18$ are routinely created in experiments~\cite{Hall2007}, and it would be interesting to look for spin-dependent transport in these systems. For CNTs with $R\approx1nm$ and $\Delta_{\text{ASOC}}\approx10 meV$ such a twist would result in partial gaps of order $100\mu eV$. \section{Conclusion} We have shown that helicity-dependent gaps can be created in CNTs via helix-shaped potentials or through creating a (mechanical) twist. Within these partial gaps, all bulk states of left-moving electrons carry a fixed spin-projection which is reversed for right-moving electrons. As a result, transport through such CNTs is highly spin-dependent. %We emphasize that this effect does not arise in chiral CNTs without an additional potential or twist \david{Maybe too strong, since there is a very weak one, as you discussed in symmetry part}. For the case of DNA-wrapped CNTs, a recent experiment has indeed found that the resistance strongly depends of the spin of injected electrons. This paves a clear path towards CNT-based spin valves and filters without magnetic field.~\cite{Michaeli2017} Our results also suggest that CNTs are a natural platform to study various exciting phenomena that have been predicted to arise in one-dimensional helical systems. For example, by coupling a DNA-wrapped (or twisted) CNT to a quantum dot, one could probe the Kondo effect in a helical liquid.~\cite{Maciejko2009,Altshuler2015} Another very interesting setup would be Josephson junctions, formed by coupling CNTs to superconducting leads. Here, unconventional Andreev bound states are expected to occur which can strongly modify the transport properties compared to a conventional junction. \textit{\ Acknowledgements:} We thank Binghai Yan and Philip Kim for helpful discussions. This work is supported by the Israel Science Foundation Grant No.~1889/16 and the Minerva Foundation (KM). \bibliographystyle{apsrev4-1} \bibliography{ciss} \appendix \section{Alternative models}\label{appendixA} \begin{figure}[b] \includegraphics[width=\linewidth]{Figure_A4.pdf} \caption{ Energy spectra of a metallic $N=3$ CNT for different models of implementing a DNA molecule wrapped at angle $ \varphi= \pi /6$. The effect of the DNA is incorporated by adding to the bare Hamiltonian (Eqs.~\ref{eq:Hopping H}-\ref{eq:Sx}, with $\Gamma_1=10^{-4}t$ and $\Gamma_1=10^{-2}t$) different spatially-modulated terms on sites directly below the molecule. In (a) we added a potential with $U=0.5t$ underneath the molecule and $U=0$ otherwise. In (b) we modified the atomic SOC entering Eqs.~\ref{eq:Sy} and~\ref{eq:Sx} by taking $\Gamma_{1,2}\rightarrow10\Gamma_{1,2}$ underneath the molecule. In (c) the spin-independent hopping amplitudes are taken to be $t\rightarrow1.5t$ underneath the molecule. Panel (d) combines the modifications described in (b) and (c). In all four cases, helicity-dependent gaps of similar magnitude open at equal momenta and energies. } \label{figureA1} \end{figure} In the main text, we incorporated the molecule into the CNT Hamiltonian of Eqs.~(\ref{eq:Hopping H}-\ref{eq:Sx}) by adding a spatial modulation to both spin-independent and spin-dependent hopping amplitudes. In that model, the dominant spin-dependent effects could be isolated by studying only the latter contribution, as we showed in the main text. An alternative approach is to introduce a spatially varying potential, $\delta\mathcal{H}=U\sum_{\mathbf{r},s}\left[A_{\mathbf{r},s}^{\dagger}A_{\mathbf{r},s}+B_{\mathbf{r},s}^{\dagger}B_{\mathbf{r},s}\right]$, where $U_{\mathbf{r}}\neq0$ is non-zero only for sites directly below the molecule. In Fig.~\ref{figureA1} we compare the electronic spectra of these different models for a metallic zig-zag CNT with $N=3$. We find that any of the implementations of the molecule, via a potential [panel (a)], through spin-dependent hopping [panel (b)] or through the spin-independent hopping [panel (c)] lead to similar gaps. The same holds for a combination of the latter two [panel (d)]. While the opening of helicity-dependent gaps is common to all of these, the details of the band structure may differ. In panels (b), the bands feature a two-fold degeneracy that is lifted in (a), (c) and (d). We note, however, that this non-universal property is \textit{spin independent} and consequently does not affect spin filtering. % \begin{figure}[b] % \includegraphics[width=1\linewidth]{Figures/Figure_A11.pdf} % \caption{Zoomed out version of the energy spectrum in Fig.~\ref{3} to cover the full energy range. The partial gap around $E\approx0.5t$ leaves behind only states with a well-defined helicity, while at higher energies, states of both helicities are present. Still, there are partial gaps where the number of modes with the two opposite helicities are imbalanced. As a result, we expect spin-polarization to arise at these energies.} % \label{figureA3} %\end{figure} \section{Comparison between spatially modulated and Rashba SOC}\label{appendixB} \begin{figure}[b] \includegraphics[width=\linewidth]{Figure_A3.pdf} \caption{ Comparison between the energy spectra of metallic $N=3$ CNTs wrapped by a DNA molecule ($ \varphi= \pi /6$) in the presence of two types of spatial modulated SOC: (a) Rashba and (b) curvature-induced. In (a) we perturb the bare graphene Hamiltonian, Eq.~\ref{eq:Hopping H}, with the Rashba term of Eq.~ \ref{eq:Rashba} (we set $\Lambda=10^{-2}t$). To facilitate comparison with the curvature-induced SOC we re-plot the corresponding spectrum of Fig.~\ref{figure3} in (b). As anticipated, both mechanisms give rise to similar spectra; in particular, both open helicity-dependent gaps at the same energies and momenta.} \label{figureA2} \end{figure} \begin{figure}[b] \includegraphics[width=1\linewidth]{Figure_A2.pdf} \caption{ Spin-polarization as a function of length for electrons transmitted through a semiconducting $N=5$ CNT wrapped by a right-handed DNA molecule ($ \varphi= \pi /6$). The spin-polarization saturates when the length exceeds $\approx 20 nm$, consistent with the estimate provided in the text. } \label{figureA4} \end{figure} The model analyzed in the main text included only the $\pi$ orbitals, and neglected hybridization with the $\sigma$ orbitals due to the molecular potential. Allowing for such a mixing gives rise to a Rashba SOC~\cite{Guinea2006,Diniz2012} \begin{align} \label{eq:Rashba} \mathcal{H}_{\text{Rashba}}&=\sum_{\mathbf{r},i,s} \Lambda(\boldsymbol{r})\left[s (\hat{\delta}_i\cdot\hat{x}) A_{\boldsymbol{r}+\boldsymbol{\delta}_i,s}^{\dagger} B_{\boldsymbol{r},s}\right.\\\nonumber&\left.+ (\hat{\delta}_i\cdot\hat{y}) e^{-is(\theta_{\boldsymbol{r}}+\theta_{\boldsymbol{r}+\boldsymbol{\delta}_i})/{2}} A_{\boldsymbol{r}+\boldsymbol{\delta}_i,s}^{\dagger}B_{\boldsymbol{r},-s}+\text{h.c}\right], \end{align} where $\Lambda(\boldsymbol{r})$ is non-zero only on sites directly below the molecule. The structure of this SOC resembles the curvature-induced one in Eqs.~\ref{eq:Sy} and~\ref{eq:Sx}; the main difference lies in the parameters $(\hat{\delta}_i\cdot\hat{x})$ and $(\hat{\delta}_i\cdot\hat{y})$, which are independent of the curvature. Unsurprisingly, adding Eq.~\ref{eq:Rashba} to the bare graphene Hamiltonian of Eq.~\ref{eq:Hopping H} leads to helicity-dependent gaps in the energy spectrum, as illustrated in Fig.~\ref{figureA2}. %\section{Full band structure and length-dependence of spin filtering}\label{appendixC} \section{Length-dependence of spin filtering}\label{appendixC} %The electronic spectra plotted in the main text (Fig~\ref{figure3}) focus on a narrow energy window. In Fig.~\ref{figureA3}, we provide a `zoomed-out' version of the band structure of the metallic $N=3$ CNT. The absence of helicity-dependent gap at zero energy is clearly visible, as well as the appearance of additional partial gaps at higher energies. The energy range over which spin polarization is significant is comparable to the magnitude of the partial gap, provided a sufficiently long segment of the CNT is wrapped with DNA.The distance where spin-polarization saturates is determined by $\xi=2v/\Delta$, where $\Delta$ is the size of the partial gap and $v$ is the velocity of the unperturbed (ungapped) modes at these energies. In Fig.~\ref{figureA4} we show the spin-polarization as a function of energy and length for a semiconducting $N=5$ CNT. From its energy spectrum (Figs.~\ref{figure2} and~\ref{figure3}), we extract the magnitude of the partial gap $\Delta\approx0.02t$ and the velocity $v\approx0.5 ta$. The resulting saturation length is $\xi\approx50a=20nm$, which is in agreement with our numerical result shown in Fig.~\ref{figureA4}. % Disorder change parmeters % translate from Delta_{ASOC} to Gamma_{1,2} and also \Lambda_{1,2} % 4-terminal Current % Twist sin^2 + experimental realization + Hamiltonian % folding in Fig. 2 % reference to the length depedent % fixing parameters everywhere \end{document} }
\caption{\label{fig:data-dragging-distance}(a) The dragging distance of oil droplets (\Circle ) and glass beads (\textifsymbol[ifgeo]{99}) for different freezing velocities. The lines are fits of equation \ref{eq:dragging distance water-ice} with \emph{r}= 10 and 22 \emph{\textgreek{m}m}, \emph{A\textgreek{x}}=0.59 and 0.17, \emph{\textgreek{d}}=1.9 and 0.2 \emph{nm}, \emph{n}=357 and 0.1 respectively. (b) (\textifsymbol[ifgeo]{80}) The thickness of the ice lenses formed during directional freezing of a colloidal solution of water and monodispersed glass beads. The y-axes are logarithmic in all graphs. In the insets the x-axes is the reciprocal of the main graph x-axes. A straight line in the inset corresponds to an agreement with our model $L\sim\exp(\frac{1}{v\cdot r})$ (eq. \ref{eq:dragging distance}). We see that the model (eq. \ref{eq:dragging distance}) describes the data well. The data is taken from \citet{dedovets2017freezing} for the oil droplets (\Circle ) and from \citet{saruya2013experimental} for ice lenses growth (\textifsymbol[ifgeo]{80}). The data for the glass beads (\textifsymbol[ifgeo]{99}) was measured using a standard directional freezing setup for this study \citep{Directional}. (c-d) The lamellar spacing as a function of the freezing velocity \emph{v} (c) and $\frac{1}{v}$ (d). The structures formed during freeze casting experiments for different freezing velocities of a solution of water with 0.8 \emph{\textgreek{m}m} diameter alumina particles at different volume concentrations in the of 5-30 \%. Different temperature of the cooling plate were used -10, -20, -30 and -50 C, which represent different thermal gradients at the interface. The data for the freeze casting structures (c-d) is taken from \citet{WASCHKIES2011}.}